Springer Series in Materials Science: Series Editors: Z.M. Wang C. Jagadish R. Hull R.M. Osgood J. Parisi

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 370

Springer Series in Materials Science

Series Editors:
Z.M. Wang
C. Jagadish
R. Hull
R.M. Osgood
J. Parisi

For further volumes:


http://www.springer.com/series/856
Junqiao Wu Jinbo Cao Wei-Qiang Han
l l

Anderson Janotti Ho-Cheol Kim


l

Editors

Functional Metal Oxide


Nanostructures
Editors
Junqiao Wu Jinbo Cao
Department of Materials Science Energy Storage & Conversion
and Engineering Materials Technologies
University of California GE Global Research
210 Hearst Memorial Mining Building 1 Research Circle, K1-3D1C
Berkeley, CA 94720, USA Niskayuna, NY 12309, USA
wuj@berkeley.edu caoj@ge.com

Wei-Qiang Han Anderson Janotti


Center for Functional Nanomaterials Materials Department
Brookhaven National Laboratory University of California
Upton, NY 11973, USA Santa Barbara, CA 93106-5050, USA
whan@bnl.gov ajanotti@mrl.ucsb.edu

Ho-Cheol Kim
IBM Research Division
Almaden Research Center
650 Harry Road
San Jose, CA 95120-6099, USA
hckim@almaden.ibm.com

ISSN 0933-033X
ISBN 978-1-4419-9930-6 e-ISBN 978-1-4419-9931-3
DOI 10.1007/978-1-4419-9931-3
Springer New York Dordrecht Heidelberg London

Library of Congress Control Number: 2011935737

# Springer Science+Business Media, LLC 2012


All rights reserved. This work may not be translated or copied in whole or in part without the written
permission of the publisher (Springer Science+Business Media, LLC, 233 Spring Street, New York,
NY 10013, USA), except for brief excerpts in connection with reviews or scholarly analysis. Use in
connection with any form of information storage and retrieval, electronic adaptation, computer software,
or by similar or dissimilar methodology now known or hereafter developed is forbidden.
The use in this publication of trade names, trademarks, service marks, and similar terms, even if they
are not identified as such, is not to be taken as an expression of opinion as to whether or not they are
subject to proprietary rights.

Printed on acid-free paper

Springer is part of Springer Science+Business Media (www.springer.com)


Preface

Metal oxides are known to possess unique functionalities that are absent or inferior
in other solid materials. Their nanostructures have emerged as an important class of
materials with a rich collection of properties and great potential for device applica-
tions. These include transparent electrodes, high-mobility transistors, gas sensors,
photovoltaics, photonic devices, energy harvesting and storage devices, and non-
volatile memories. The research interest in oxide nanostructures is reflected by the
exponential growth of publications in this field in recent years, as shown in Fig. 1.
The impact of these publications is broad, as manifested by their large number of
citations. In line with this growth, there have always been symposia dedicated to the
field of oxide nanostructures in all of the recent Spring and Fall Materials Research
Society meetings. From these meetings we have witnessed great excitement in this
field, as these symposia attracted a large group of attendees from around the world.

Fig. 1 Growth of the number of publications with topics on “oxide” and “nanostructures” and their
citations. Statistics from ISI Web of Knowledge

v
vi Preface

We therefore invited some of the most active researchers to contribute book


chapters on topics that are at the frontiers in this field.
This book is divided into three parts: Basic Properties, which includes Chaps.
1–7; Synthesis and Processing, which includes Chaps. 8–10; and Applications,
which includes Chaps. 11–14. Chapter 1 is focused on an important effect, strain,
in oxide nanostructures. Chapters 2–4 describe conductivity control in oxide semi-
conductors, effects of point defects, and electron transport in oxide nanostructures,
respectively. Chapter 5 reviews spectroscopic investigations and Chap. 6 discusses
electronic properties of oxide surfaces. This part ends with a case study in Chap. 7:
electronic and magnetic properties of strontium ruthenate ultra-thin layers. In the
Synthesis and Processing part, the widely used solution processing is detailed in
Chap. 8 with titania as an example. Chapter 9 presents a review of biologically
templated oxide nanostructure growth, while Chap. 10 details epitaxial stabilization
of low-dimensional oxide layers. The Applications section is comprised of reviews
of three major applications of metal oxide nanostructures: Catalysis (Chap. 11),
batteries (Chaps. 12 and 14), and memory (Chap. 12).
We are grateful to all the authors who have contributed chapters to this book. We
also thank the staff at Springer for their assistance and patience in editing this book.
Junqiao Wu would like to acknowledge financial support from the National Science
Foundation and the US Department of Energy.

Berkeley, CA Junqiao Wu
Niskayuna, NY Jinbo Cao
Upton, NY Wei-Qiang Han
Santa Barbara, CA Anderson Janotti
San Jose, CA Ho-Cheol Kim
Contents

Part I Basic Properties

1 New Opportunities on Phase Transitions of Correlated


Electron Nanostructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Jinbo Cao and Junqiao Wu
2 Controlling the Conductivity in Oxide Semiconductors. . . . . . . . . . . . . . 23
A. Janotti, J.B. Varley, J.L. Lyons, and C.G. Van de Walle
3 The Role of Defects in Functional Oxide Nanostructures . . . . . . . . . . . . 37
C. Sudakar, Shubra Singh, M.S. Ramachandra Rao, and G. Lawes
4 Emergent Metal–Insulator Transitions Associated
with Electronic Inhomogeneities in Low-Dimensional
Complex Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
An-Ping Li and Thomas Z. Ward
5 Optical Properties of Nanoscale Transition Metal Oxides . . . . . . . . . . . 87
Janice L. Musfeldt
6 Electronic Properties of Post-transition Metal Oxide
Semiconductor Surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
T.D. Veal, P.D.C. King, and C.F. McConville
7 In Search of a Truly Two-Dimensional Metallic Oxide . . . . . . . . . . . . . . 147
Priya Mahadevan and Kapil Gupta

Part II Synthesis and Processing

8 Solution Phase Approach to TiO2 Nanostructures . . . . . . . . . . . . . . . . . . . 157


John D. Bass and Ho-Cheol Kim
9 Oxide-Based Photonic Crystals from Biological Templates. . . . . . . . . . 175
Michael H. Bartl, Jeremy W. Galusha, and Matthew R. Jorgensen

vii
viii Contents

10 Low Dimensionality and Epitaxial Stabilization


in Metal-Supported Oxide Nanostructures:
MnxOy on Pd(100) MnxOy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
Cesare Franchini and Francesco Allegretti

Part III Applications

11 One-Dimensional Oxygen-Deficient Metal Oxides . . . . . . . . . . . . . . . . . . . . 241


Wei-Qiang Han
12 Oxide Nanostructures for Energy Storage . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
Yuan Yang, Jang Wook Choi, and Yi Cui
13 Metal Oxide Resistive Switching Memory. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
Shimeng Yu, Byoungil Lee, and H.‐S. Philip Wong
14 Nanostructured Metal Oxides for Li-Ion Batteries . . . . . . . . . . . . . . . . . . . 337
Juchen Guo and Chunsheng Wang

Index. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365
Contributors

Francesco Allegretti Institute of Physics, Surface and Interface Physics,


Karl-Franzens University Graz, Graz 8010, Austria
Michael H. Bartl Department of Chemistry, University of Utah,
Salt Lake City, UT 84112, USA
Department of Physics, University of Utah, Salt Lake City, UT 84112, USA
John D. Bass IBM Research Division, Almaden Research Center,
650 Harry Road, San Jose, CA 95120-6099, USA
Jinbo Cao Department of Materials Science and Engineering,
University of California, Berkeley, CA, USA
Materials Sciences Division, Lawrence Berkeley National Laboratory,
Berkeley, CA, USA
Jang Wook Choi Department of Materials Science and Engineering,
Stanford University, Stanford, CA, USA
Yi Cui Department of Materials Science and Engineering,
Stanford University, Stanford, CA, USA
Cesare Franchini Center for Computational Materials Science,
Universität Wien, Wien 1090, Austria
Jeremy W. Galusha Department of Chemistry, University of Utah,
Salt Lake City, UT 84112, USA
Department of Physics, University of Utah, Salt Lake City, UT 84112, USA
Juchen Guo Department of Chemical and Biomolecular Engineering,
University of Maryland, College Park, MD 20742, USA
Kapil Gupta S.N. Bose National Centre for Basic Sciences,
Salt Lake, Kolkata, India

ix
x Contributors

Wei-Qiang Han Center for Functional Nanomaterials, Brookhaven


National Laboratory, Upton, NY 11973, USA
A. Janotti Materials Department, University of California,
Santa Barbara, CA, USA
Matthew R. Jorgensen Department of Chemistry, University
of Utah, Salt Lake City, UT 84112, USA
Department of Physics, University of Utah, Salt Lake City, UT 84112, USA
Ho-Cheol Kim IBM Research Division, Almaden Research Center,
650 Harry Road, San Jose, CA 95120-6099, USA
P.D.C. King Department of Physics, University of Warwick,
Coventry, CV4 7AL, UK
G. Lawes Department of Physics and Astronomy, Wayne State University,
Detroit, MI 48201, USA
Byoungil Lee Department of Electrical Engineering, Center
for Integrated Systems, Stanford University, Stanford, CA, USA
An-Ping Li Center for Nanophase Materials Sciences,
Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA
J.L. Lyons Materials Department, University of California,
Santa Barbara, CA, USA
Priya Mahadevan S.N. Bose National Centre for Basic Sciences,
Salt Lake, Kolkata, India
C.F. McConville Department of Physics, University of Warwick,
Coventry, CV4 7AL, UK
Janice L. Musfeldt University of Tennessee, Knoxville, TN, USA
M.S. Ramachandra Rao Department of Physics, Indian Institute
of Technology, Chennai, Tamil Nadu 600036, India
Nano Functional Materials Technology Centre, Indian
Institute of Technology, Chennai, Tamil Nadu 600036, India
Shubra Singh Department of Physics, Indian Institute of Technology,
Chennai, Tamil Nadu 600036, India
Nano Functional Materials Technology Centre, Indian
Institute of Technology, Chennai, Tamil Nadu 600036, India
C. Sudakar Department of Physics and Astronomy, Wayne State University,
Detroit, MI 48201, USA
Department of Physics, Indian Institute of Technology,
Chennai, Tamil Nadu 600036, India
Contributors xi

J.B. Varley Materials Department, University of California,


Santa Barbara, CA, USA
Department of Physics, University of California,
Santa Barbara, CA, USA
T.D. Veal Department of Physics, University of Warwick, Coventry,
CV4 7AL, UK
C.G. Van de Walle Materials Department, University of California,
Santa Barbara, CA, USA
Chunsheng Wang Department of Chemical and Biomolecular Engineering,
University of Maryland, College Park, MD 20742, USA
Thomas Z. Ward Materials Science and Technology Division,
Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA
H.-S. Philip Wong Department of Electrical Engineering, Center
for Integrated Systems, Stanford University, Stanford, CA, USA
Junqiao Wu Department of Materials Science and Engineering,
University of California, Berkeley, CA, USA
Materials Sciences Division, Lawrence Berkeley National Laboratory,
Berkeley, CA, USA
Yuan Yang Department of Materials Science and Engineering,
Stanford University, Stanford, CA, USA
Shimeng Yu Department of Electrical Engineering, Center
for Integrated Systems, Stanford University, Stanford, CA, USA
Part I
Basic Properties
Chapter 1
New Opportunities on Phase Transitions
of Correlated Electron Nanostructures

Jinbo Cao and Junqiao Wu

1.1 Introduction

Correlated electron materials (CEMs) exhibit rich and fascinating properties


including high-TC superconductivity, colossal magnetoresistance, and nonlinear
optical behavior [1, 2]. Most of these remarkable properties originate from the
interplay between spin, lattice, charge, and orbital degrees of freedom of the material
[1, 3]. These competing factors typically result in the coexistence of near degenerate
states and cause a spatial phase inhomogeneity or multiple domain structures at the
micro- and nanometer scale [4]. Although the spatial electronic phase separation is
believed to be associated with many exotic properties of CEMs, investigation of the
fundamental properties of these materials has been hampered due to the multiple
domain structures [1, 2, 5]. For example, transport and optical measurements on
devices which are much larger than the intrinsic electronic phase domains only
provide an averaged response of the inhomogeneous ensembles to external
parameters. Nanoscale materials, on the other hand, can be smaller or comparable
to the characteristic domain size of CEMs. When the electronic phases are spatially
confined, the conduction path can be better defined by precluding percolative
behavior that occurs in thin film or bulk samples. It is therefore possible to

J. Cao (*)
Department of Materials Science and Engineering, University of California, Berkeley, CA, USA
Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA, USA
Current contact address: Energy Storage & Conversion Materials Technologies,
GE Global Research, Niskayuna, NY, USA
e-mail: caoj@ge.com
J. Wu
Department of Materials Science and Engineering, University of California, Berkeley, CA, USA
Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA, USA

J. Wu et al. (eds.), Functional Metal Oxide Nanostructures, Springer Series 3


in Materials Science 149, DOI 10.1007/978-1-4419-9931-3_1,
# Springer Science+Business Media, LLC 2012
4 J. Cao and J. Wu

probe the properties of different electronic phases and investigate the fundamental
physical properties of CEMs in nanoscale specimens.
The fact that multiple phase coexistence has been observed in many different
CEMs such as manganites and cuprates near a phase transition raises questions
related to the origin of the phase inhomogeneity and its role in many interesting
properties of CEMs [6–10]. Despite decades of investigation, the question of
whether the phase separation is intrinsic or caused by external stimuli (extrinsic)
still remains open [7, 11–14]. For example, in colossal magnetoresistive
manganites that undergo a phase transition from ferromagnetic metal to antifer-
romagnetic insulator, these two phases typically coexist at length scales ranging
from nano- to micrometer, displaying a spatial phase inhomogeneity [7, 8, 15].
Existing theories explain the origin of phase inhomogeneity by either an intrinsic
mechanism arising from inherent properties of such correlated electron systems
[11, 12] or extrinsic mechanisms based on the effects of chemical disorder or local
strain distribution [13]. In the model of elastically mediated phase coexistence,
the structural aspect will be the primary reason that causes the multiphase
coexistence. External stimuli such as strain can be used to sensitively manipulate
patterns of metallic and insulating regions [13]. The recent findings that aniso-
tropic electronic domains of La5/8  xPrxCa3/8MnO3 (x ¼ 0.3) (LPCMO) thin film
can be induced by epitaxially locking it to an orthorhombic NdGaO3 substrate
suggest that the origin of phase coexistence can be strongly influenced by elastic
energy rather than local chemical inhomogeneities [16].
Continuously tuning of lattice strain in CEM nanostructures, on the other hand,
would be more desirable to uncover the origin of the phase inhomogeneity.
If phase inhomogeneity is absent in strain-free, single-crystal specimens,
but can be introduced and modulated by external strain, it would be concluded
that strain is responsible for the phase inhomogeneity in CEMs. Compared to thin
films, CEM nanostructures are dislocation-free and can be subjected to coherent
and continuously tunable external strain. CEM phase transitions and domain
dynamics can then be explored through in situ microscopic experiments varying
strain and temperature independently. Such an approach would enable, for the
first time, probe of CEMs at the single domain level under continuous tuning of
their lattice degrees of freedom.
In addition to elucidating the origin of the phase inhomogeneity, the strain
tuning could also be employed to modify the properties of CEMs. In contrast to
conventional materials, where elastic deformation causes minor variations in
material properties, lattice strain has profound influence on the electrical, optical,
and magnetic properties of CEMs through coupling between the charge, spin, and
orbital degrees of freedom of electrons. Control and engineering of strain have
become an important strategy for achieving novel functionalities and probing
exotic properties of CEMs. However, bulk inorganic materials can only sustain
extremely low nonhydrostatic strain (typically <0.1%) before plastic deformation
or fracture occurs. Recent advances in the synthesis of CEM nanostructures offer
new opportunities of studying the strain effect on CEMs. As these materials are
typically single crystals and dislocation free, they can sustain an extraordinary
1 New Opportunities on Phase Transitions of Correlated Electron Nanostructures 5

amount of uniaxial strain without fracturing and the strain is also continuously
tunable by varying the external stress.
This chapter focuses on recent reports of vanadium dioxide (VO2), a
representative CEM that displays interesting phase transitions at convenient
temperatures. Research of nanostructured VO2 has led to new discoveries and helps
resolve many outstanding questions that cannot be addressed in bulk or thin film
specimens. With the contribution of nanoscale or micrometer scale VO2 samples, we
specifically focus on the following topics (1) origin of phase inhomogeneity in VO2
[5, 17, 18]; (2) domain organization and manipulation [5, 19, 20]; (3) driving
mechanism of the phase transition [21–25]; (4) superelasticity in VO2 [26]; (5) new
phase stabilization under stress [27–29]; and (6) thermoelectric effects of the domain
walls [30]. After discussing the application of VO2 nanostructures on the
above questions, we conclude with a short discussion of how these novel probing
techniques and methods could be extended to other CEMs and resolve the compli-
cated questions there.

1.2 Electrical and Structural Transitions in VO2

In strain-free state,VO2 undergoes a first-order metal–insulator phase transition


(MIT) at TC0 ¼ 341 K [31–33] with a change in conductivity by several orders
of magnitude. The MIT is accompanied by the structural phase transition
from the high-temperature, tetragonal, metallic phase (rutile structure, R) to the
low-temperature, monoclinic, insulating phase (M1). Upon cooling through the
MIT, the vanadium ions dimerized and these pairs tilt with respect to the R-phase
c-axis (cR ) [31–33]. It is known that the transition can be profoundly affected by
uniaxial strain as well as doping. In addition to the aforementioned low tempera-
ture monoclinic structure M1 phase, another insulating monoclinic structure
M2 phase can also be induced by doping with Cr or uniaxial compression
perpendicular to cR [31, 32, 34]. In M2, only half of the vanadium atoms dimerize
while the other half form zigzag chains [31, 32]; therefore, it can be viewed as an
intermediate structure between M1 and R.
The vanadium dimerization leads to unit cell doubling in the monoclinic M1
and M2 structures. During the MIT, the structural transition from M1 to R
effectively shrinks the specimen along cR direction by e0  1% [31]. Along the
tetragonal aR and bR axes, on the other hand, the lattice expands by 0.6 and 0.1%,
respectively, causing a volume shrinkage of 0.3% [31]. The transition from M1 to
M2 expands it along cR direction by e0  0:3% [32]. Note that the lattice constant
is temperature dependent and the exact lattice parameters of these three phases at
the same temperature do not exist in literatures. The above numbers are
approximated from the discontinuous change from the phase diagram of Cr-doped
VO2 [31, 32]. A triclinic phase (T) might also derive from the M1 structure,
but only with a continuous change in lattice constant and angles [31, 32].
6 J. Cao and J. Wu

1.3 Experimental Methods

Single-crystalline VO2 beams (nanometer and micrometer size beams) were


synthesized using a vapor transport method [35]. The size distribution, lattice
structure, and crystal orientation of these beams were characterized by different
techniques such as X-ray diffraction, scanning electron microscopy (SEM),
transmission electron microscopy, and selected area electron diffraction. These
VO2 beams grow along the tetragonal cR -axis with {110} planes as the bounding
side faces. A percentage of VO2 beams were grown one end anchored on the SiO2
substrate during the synthesis, and naturally formed cantilevers for in situ optical
and SEM bending experiments.
During the electrical measurement, devices were typically fabricated on
as-grown, bottom-clamped VO2 beams via sputtering depositing adhesion layer
(~20 nm Ti or V) and a Au layer (~400 nm) onto the photolithography defined
patterns. The uncovered part of SiO2 was later etched using buffered
oxide etchant to liberate the VO2 beam into an end–end-clamping configuration.
To apply compressive stress to VO2 beams, single-crystal VO2 beams were
transferred to a polycarbonate or Kapton substrate and bonded by silver epoxy.
Addition uniaxial compressive stress was achieved by three-point concave bend-
ing bendable substrate along the length direction.
Laue diffraction was performed at the mXRD end station (Beamline 12.3.2) in
the advanced light source (ALS) at Lawrence Berkeley National Laboratory.
Both analytical model and phase field simulations were employed to model the
domain nucleation and stabilization. In both cases, the total energy arises from
thermodynamic bulk energy, interfacial energy, and strain energy. The equilibrium
domain structures were determined by numerical minimization of the total energy
under different conditions.

1.4 Results and Discussions

1.4.1 Phase Inhomogeneity and Domain Organization

Until the advent of VO2 single-crystal beams, measurements on crystal bulk and
polycrystalline thin films have suffered many problems near the MIT. Due to the
changes of the lattice parameters associated with the transition, bulk crystals tend to
crack across the MIT and degrade upon repeated cycling [36]. On the other hand,
polycrystalline thin films often display a broadened transition associated with phase
inhomogeneity [17, 37, 38]. Scanning near-field infrared microscopy was recently
employed to image the metallic and insulating phases in VO2 thin films and it was
found that these electronic phases are separated in a percolative manner [17]. As
discussed below, the phase inhomogeneity might originate from the inevitable
nonuniform local stress in thin films and therefore prevents detailed investigations
1 New Opportunities on Phase Transitions of Correlated Electron Nanostructures 7

Fig. 1.1 (a) Four-probe resistance of single-crystal VO2 beams as a function of temperature. The
free-standing VO2 beam shows single domain behavior, whereas the clamped beam shows multiple
domains during the transition [5]. (b) Optical images of the multiple-domain devices showing the
coexistence of metallic (dark) and insulating (bright) domains at intermediate temperatures [5]. The
scale bar is 5 mm. (c) Optical image of clamped VO2 at 338 K showing periodic metallic and
insulating domains. The width of the VO2 beam is around 1.5 mm. (d) Schematic diagram showing
the periodic domain pattern of a VO2 beam coherently strained on an SiO2 substrate [19]. Blue and
red correspond to tensile and compressive strain, respectively. “M” denotes metallic phase

of VO2 within the stress and temperature space. Single-crystal VO2 nanometer and
micrometer size beams will avoid the crackling of the specimens and eliminate the
phase separation and domain formation that are widely observed in bulk and thin
films, thus providing opportunities to investigate the intrinsic properties of VO2 at
the single domain level.
Depending on whether a stress is imposed on the sample or not, VO2 beams
show rather distinct optical and electronic properties. By adjusting the synthesis
conditions, VO2 beams can be grown with either a weakly coupled beam–substrate
interface where the stress can be easily released, or a strongly coupled, clamped
interface that pins the beam to the substrate. In the latter case thermal stress is
imposed on the beams after cooling from the growth temperature to room tempera-
ture. Both types of beams were incorporated into fourprobe devices using lithogra-
phy for transport measurement. As shown in Fig. 1.1a, devices made from
unstrained sample display a sharp drop of resistance at TC0 ¼ 341 K (single-domain
device), accompanied by an abrupt change in optical reflection, with dark reflection
corresponding to metallic phase and bright reflection to insulating phase [5].
Different behavior was observed for the stressed beams bottom-clamped on the
substrate surface. The electrical resistance of the clamped VO2 beams decreases
gradually over a wide phase transition temperature range, showing an effective
second-order phase transition. High-magnification optical imaging of the clamped
8 J. Cao and J. Wu

beams in Fig. 1.1b revealed multiple metallic and insulating domains appearing
during the transition, where dark domains nucleated in the bright phase and grew
with increasing temperature, finally merging into a single metallic phase [5]. The
broadening of transition therefore is a direct consequence of the nucleation and
growing/shrinking of metallic/insulating domains as a function of temperature in
stressed VO2 beams [5]. The electronic phase separation in VO2 beams is reminis-
cent of the phase inhomogeneity observed in other CEM thin films, although the
domain patterns in VO2 beams are more regular due to the lateral confinement and
more coherent strain imposed on them. That the multiple phase coexistence can be
observed in stressed VO2 beam but not in free-standing specimens [5] suggest that
the lattice strain be responsible to the phase inhomogeneity observed in polycrys-
talline VO2 films [17], and shed light on the origins of phase inhomogeneity in other
CEMs [13]. The broadening of phase transition in VO2 thin films is therefore a
direct consequence of the phase inhomogeneity across the transition, which arises
from the inevitable nonuniform local stress of the films. The above simultaneous
optical and electronic experiments also established a direct correlation between
optical contrast and the electronic phases in VO2 beams, where bright and dark
reflection indicates the insulating and metallic phase, respectively.
A fully coherently strained VO2 beam bottom-clamped on the substrate exhibits
periodic metal–insulator domains in high-resolution optical microscopy within the
transition range (Fig. 1.1c). The multiple domain coexistence can be understood
through analysis of energy minimization [19]. Such a domain pattern forms sponta-
neously as a result of competition between strain energy in the elastically mismatched
VO2/substrate system and domain wall energy in the VO2. The period of the pattern
is determined by the balance between the strainenergy minimization that favors
small, alternating metallic–insulating domains and domain-wall energy minimiza-
tion that opposes them [19]. The spatial periodicity can be described quantitatively
using a model originally developed for the strained ferroelectric systems [39]. The
total energy in unit nanobeam–substrate interface area is given by [19, 39]
P
1
EðlÞ ¼ ple3 1e2ð2jþ1Þpt=l
ð2jþ1Þ3
þ gtl þ ð fM2þfI Þt . Here l is the spatial period of the domain
j¼0
pattern, e is the volume density of the elastic misfit energy, g is the domain-wall
energy per unit domain-wall area, t is the nanobeam thickness, and fM and fI are the
free energy densities of the metallic and insulating phases, respectively.
The equilibrium domain period for different beams can therefore be determined
by numerical minimization of EðlÞ. Figure 1.1d displays schematic diagram of the
periodic domain patterns of a VO2 nanobeam strained on an SiO2 substrate [19].

1.4.2 Domain Dynamics and Manipulation

The dynamics of domains in response to external stimuli is also attractive as it


provides an effective route to control and manipulate the domains at small scales.
1 New Opportunities on Phase Transitions of Correlated Electron Nanostructures 9

a b
298 K

333 K

343 K

353 K

363 K
373 K

383 K

c d
1.2 298 K
no external strain
1.0
0.8 compressive strain

0.6
η

0.4 st
1 up 343 K
0.2 st
1 down
nd
2 up
0.0 nd
2 down

−0.2
−1.4 −1.6 −1.8 −2.0 −2.2
Strain (%)

Fig. 1.2 Domain dynamics and manipulation with strain in VO2 [5]. (a) Optical images of an
array of triangular metallic domains nucleated and costabilized by tensile and compressive strain
along a bent microbeam. (b) Phase-field modeling of domain formation in a bent VO2 beam. From
top to bottom: first, initial state of random phase distribution; second, equilibrium phase distribu-
tion at the natural MIT transition TC0 ; third, equilibrium strain (exx ) distribution at TC0 : yellow and
dark green denote the maximum tensile and maximum compressive strain, respectively; forth,
equilibrium strain energy density distribution: yellow denotes the highest strain energy density,
dark green denotes the lowest. (c) Uniaxial compression reversibly induces a metal–insulator
transition at room temperature. Here  is the fraction of metal phase along the beam. (d) Strain
engineering domains in a flexible VO2 microbeam. Scale bars in (a), (c), and (d) 10 mm

A current-driven phase oscillation and domain-wall propagation have been


observed in tungsten-doped VO2 nanobeams [20]. The domain oscillation occurs
through the axial drift of a single metallic–insulating domain wall driven by a
combined effect of Joule heating and Peltier cooling.
Through actively bending single VO2 beams, it has been shown that strain can
produce ordered arrays of metallic and insulating phases along the length [5].
The nonclamped beams on the substrate were bent by pushing part of the beam
with a microprobe. A large compressive strain will result near the inner edge of
the high-curvature regions of the bent beam and a tensile strain will occur near the
outer edge. Figure 1.2a shows the development of an array of triangular domains
along a bent VO2 beam imaged at different temperatures [5]. The bent beam
10 J. Cao and J. Wu

was in insulating phase at room temperature. With increasing temperatures,


submicron, periodic triangular metallic domains started to nucleate at the inner
edge of the bent region where the strain was the most compressive. These domains
continued to grow and expand with increasing temperature, while the triangular
geometry and periodic arrangement were maintained. At natural phase transition
temperature T  TC0 ¼ 341 K, the straight, strain-free part of the beam switched
abruptly to the metallic phase as expected, while the bent part of the beam showed
a nearly 50–50% coexistence of metallic and insulating phases. As temperature
was further increased, the metallic phase expanded toward the outer edge and
finally completely eliminated the insulating phase.
The domain nucleation and organization in a bent beam can be successfully
captured through two-dimensional phase-field modeling as shown in Fig. 1.2b [5].
The total energy F(f) is equal to the sum of bulk thermodynamic energy, interfacial
(domain wall) energy, and strain energy:
Z    
b2 2 1 
FðfÞ ¼ f ðfÞ þ jrfj þ Cijkl eij  eij ekl  ekl dA:
T T
2 2

The parameter f denotes the phase. f(f) describes the relative thermodynamic
energy of the two phases and is temperature dependent. The second term reflects
the interfacial energy. The last term is the elastic energy where C is the elastic
modulus tensor, e is the strain, and eT is the lattice mismatch between the two
phases. In the initial state, the phase distribution is random. At natural transition
temperature TC0 ¼ 341 K, the equilibrium phase distribution exhibits a periodic,
triangular domain pattern, agreeing well with experimental observations.
This pattern nearly completely relieves the strain energy in the bent beam, with
some remnant strain at the triangular tips (Fig. 1.2b). The period varies for
different interfacial energy density and elastic constants. Specifically, the period
is determined by competing effects of strain energy relaxation and interfacial
energy minimization: smaller period results in more effective strain energy relief,
but at the cost of introducing more interfacial area.
Strain was also used to lower the temperature of the transition from its bulk value
of 341 K to room temperature, which paves the way for future potential applications
[5]. Compressive stress was reversibly applied along the length of a VO2 beam
clamped onto a polycarbonate substrate and the fraction of metal phase Z was
monitored as a function of applied stress. As shown in Fig. 1.2c, VO2 beam remained
entirely insulating ( ¼ 0) until stressed to a total strain of approximately 1.9%,
then entered a strain regime where periodic metallic and insulating domains
coexisted, ultimately reaching a full metallic state ( ¼ 1) at total strain of approxi-
mately 2.1%. As the resistance of the VO2 beam changes by several orders of
magnitude across the MIT, this strain-controlled, room-temperature phase transition
can be used as a “strain-Mott” transistor. Enabled by the sensitivity of the electronic
phases to local strains, external stress was used to manipulate and engineer these
functional domains. As displayed in Fig. 1.2d [5], an array of metallic–insulating
1 New Opportunities on Phase Transitions of Correlated Electron Nanostructures 11

triangular domains is created and eliminated by simply modulating the local strain
state of a bent VO2 beam. Initially the beam was strain free, and therefore in pure
insulating phase at room temperature and pure metallic state at 343 K. When the
beam was locally bent at 343 K, an array of triangular insulating domains was
created in the high curvature region in response to the local tensile strain.
These domains were highly mobile and could be driven to different location along
the beam by slight modifications of the bending geometry. The ability to engineer
phase inhomogeneity and phase transition with strain in VO2 opens opportunities for
designing and controlling functional domains for device and sensor applications.
As distinctly different physical and chemical properties are associated with
the metallic and insulating phases, interfacing strain-engineered VO2 with other
molecular, nano-, or polymeric materials may provide new assembly strategies to
achieve collective and externally tunable properties. Similar approaches can be
applied in other CEMs and may lead to new revolutions of strain engineering
solid-state materials to achieve collective functionalities.

1.4.3 Investigation of Phase Transition at the Single


Domain Level

By investigating the electrical and optical properties of VO2 at the single domain
level, it is possible to discover new aspects of its phase transition in unprece-
dented detail. For example, it has been a topic of debate for decades whether the
MIT is fundamentally driven by electron–electron correlation, therefore is a Mott
transition, or by electron–lattice interaction, therefore a Peierls transition
[17, 22–24, 40–42]. In the Mott transition picture, the insulating phase is a Mott
insulator where the bandgap opens because of Coulomb blockade between
strongly localized d electrons at the vanadium sites. In the Peierls transition
picture, the bandgap exists because of lattice potential modulation by the vana-
dium dimerization. There are many experimental evidences that support both
mechanisms [17, 22–24, 40–42]. The controversy is largely due to the poorly
understood near-threshold behavior in thin films or bulk crystals, where phase
separation and domain formation prevent direct measurements of intrinsic
electronic and optical properties. To circumvent this issue, the electrical
properties of VO2 beams were investigated along the phase boundary in the
stress–temperature phase space at the single domain level. Such an electrical
measurement precisely along the phase boundary is not possible in thin films
or bulk because of percolative conduction and inhomogeneous strain in
those systems.
According to Gibbs’ phase rule, the number of degrees of freedom (F) in a pure
component system is intimately linked with the number of components (C) and the
number of phases (P) through equation F ¼ C  P þ 2 ¼ 3  P. Here C ¼ 1 for
single component system. Under a single phase (P ¼ 1) condition, two variables
12 J. Cao and J. Wu

a b
40

Resistivity of insulator (Ω cm)


Metal
Internal stress

20
r1 r1c
1
2
10
3 8
Insulator 6

Temperature 30 40 50 60 70 80 90 100110120
Temperature (°C)
c d

Fig. 1.3 (a) Schematic view of the path for a fixed length VO2 beam moving in the uniaxial
stress–temperature phase diagram [43]. (b) Resistivity of the insulating phase of VO2 as a function
of temperature. A constant resistivity (at 60–100 C) is observed when the system moves along the
phase boundary [21]. (c) Isobaric (A), isothermal (B), and combined sequential processes (C) in
the uniaxial stress–temperature phase diagram of VO2 showing phase transition from the
insulating (I) phase to metallic (M) phase. Green dots show schematically the position where
the transition threshold is defined. (d) Measured resistivity of VO2 beams at 300 K, at the
threshold, and in M-phase, respectively. The error is mostly from uncertainties in determining
the effective height and length of the microbeam. Also shown is the apparent threshold resistivity
from a VO2 thin film. The horizontal line shows the constancy of rth in I phase

(F ¼ 2) such as temperature and stress can be controlled to any selected pair of


values. However, if the system enters the two phase region (P ¼ 2), the temperature
and stress cannot be varied independently any more. When VO2 beam is clamped
onto a substrate at its two ends with fixed length, the system will be forced to move
along the metal/insulator phase boundary in the stress–temperature phase space,
as shown in Fig. 1.3a. A constant resistivity of the insulating phase was observed
when the system moves along the phase boundary Fig. 1.3b [21, 43].
More detailed experiments to test such a constant threshold resistivity over a wider
range of space including both tensile and compressive states were also performed
[25]. As shown in Fig. 1.3c, in ambient conditions VO2 is in insulating state.
1 New Opportunities on Phase Transitions of Correlated Electron Nanostructures 13

The metallic state can be reached either following route A at constant stress (akin to
the isobaric process for a liquid–vapor transition), or following route B at constant
temperature (isothermal process). Moreover, routes A and B can be combined to form
sequential processes following routes C. The devices allow investigation of the MIT
in the two-dimensional temperature–stress phase space by independently varying
these two external parameters. Despite the different transition route, the threshold
resistivity (rth ) of the insulating phase right before the system switches to metallic
phase and stays constant over a wide range. As shown in Fig. 1.3d, the systems have
different transition temperature and room-temperature resistivities under different
strain states. However, the values of rth all fall onto a constant line.
Such a constant rth has deep implications to the physics of the MIT. In a Mott
picture, the transition occurs when the density-dependent screening reaches a
critical strength. As suggested by Mott theory, the density is the maximum
semiconducting carriers coming from the insulating side, which is proportional
to the threshold resistivity rth. The observation that a constant critical free carrier
density has to be reached to trigger the insulator to metal transition indicates that
the transition is fundamentally driven by electron–electron interactions. On the
contrary, a phonon-driven mechanism (Peierls transition) would not be expected
to be sensitive to the carrier density in the insulator.

1.4.4 Superelasticity in Phase Transition

Superelasticity (or pseudoelasticity) has been widely observed in shape memory


alloys, where an applied stress can cause a reversible phase transformation between
the austenitic and martensitic phases of a crystal [44]. The superelasticity originates
from the reversible creation and motion of domain boundaries during the phase
transformation rather than from bond stretching or the introduction of defects in the
crystal lattice. CEMs are typically brittle and can only be stretched or compressed for
an extremely small percentage. In CEMs with phase transitions, control of the domain
wall motion may provide an effective way to realize superelasticity. This effect has
been recently demonstrated in the first-order MIT of VO2 [26]. An atomic force
microscopy (AFM) tip was used to bend a VO2 nanocantilever and the force–dis-
placement ( f–w) curves were recorded, as shown in Fig. 1.4a, b. At small deflections,
the f–w curve is linear as expected [26]. The slope of this linear part can be used to
determine the Young’s modulus of VO2. At large deflections (but still within the
reversible, elastic regime), VO2 beam shows superelasticity, as evidenced by the
appearance of kinks and nonlinearity. These kinks are reproducible upon repeated
bending of the VO2 nanobeams and occur at lower displacements at temperatures that
are closer to the natural transition temperature (TC0 ). At the same degree of deflection,
no such kinks were observed when bending nanobeams with similar size but made of
materials without phase transition, such as ZnTe nanowires. The critical stress from
the first kink on the f–w curve is calculated and plotted on the stress–temperature phase
14 J. Cao and J. Wu

Fig. 1.4 Superelastic metal–insulator phase transition in VO2 beams [26]. (a) Schematic view of
bending a VO2 nanocantilever using an AFM tip. (b) Force–displacement curves of a VO2
cantilever measured at various temperatures. The curves are vertically offset for clarity. Arrows
show the position of the first kink on each curve. (c) Calculated stress at the root of the VO2 beam
at the measured first kink of the force–displacement curves plotted in the stress–temperature phase
space. Solid squares and empty triangles represent the stress in the loading and unloading
processes, respectively. The dashed line is the phase boundary calculated from the Clapeyron
equation using a latent heat of 1,020 Cal/mol. (d) Optical images of side bending a wide VO2
beam, showing the nucleation of new domains with increasing stress

diagram in Fig. 1.4c. The metal–insulator phase boundary is calculated from the
Clapeyron equation using a latent heat of the MIT of 1,020 Cal/mol. The measured
critical stress points are distributed along the phase boundary line, consistent with the
Clapeyron equation. Figure 1.4d shows optical images of sidebending a wider VO2
beam, which suggests that the nonlinearity and kinks in the f–w curve are related to
the creation and motion of new domains. Phase-field modeling was also used
to simulate the f–w curve and obtained qualitatively similar results, as shown
in Fig. 1.5e, f [26].
The superelastic behavior observed during the MIT in VO2 provides a new
route to investigate the solid–solid phase transitions in general. The dynamics of
force–displacement curve during the transition can be easily understood
by comparing it with the condensation–evaporation process between a liquid
and its vapor. When the vapor is compressed isothermally below the critical
temperature, the pressure first increases following the ideal gas law, then reaches
a plateau where liquid droplets nucleate out of the vapor forming a liquid–vapor
1 New Opportunities on Phase Transitions of Correlated Electron Nanostructures 15

Fig. 1.5 (a) Simulated force–displacement curves for a beam demonstrating the slope change that
occurs at the onset of new domain formation. (b) Domain distribution obtained by two-dimen-
sional phase field simulation incorporating local strain relaxation for a beam of length 7.5 mm and
height 0.75 mm. The applied terminal load force is 20 mN, for temperature below and above the
natural transition temperature, respectively. (c) Residual strain energy distributions corresponding
to the parameters in part (b) illustrating how new domains relieve strain energy

coexisting system [45]. In this coexisting state the compressibility of the system
diverges, because the decrease in total volume is accounted for by the conversion of
more vapor into much denser liquid, rather than to increase pressure as in the pure ideal
gas. The pressure–volume curve in this system is equivalent to the f–w curve in VO2
nanobeam. At temperatures lower than the natural transition temperature, the bottom
edge at the root of the nanobeam will be compressively stressed with increasing
bending. At a critical compressive stress, new metallic domains start to nucleate out
of the original insulating phase, and the bottom portion of the nanobeam root enters a
metal–insulator phase coexisting state. As the nano-beam is further bent beyond the
critical stress, the metallic domains start to grow in response to the increasing uniaxial
compression and a vanishing Young’s modulus is expected from the metal–insulator
coexisting part (bottom portion) of the nanobeam. The top portion of the nanobeam is
under tension and therefore remains in the original insulating phase. The overall
Young’s modulus measured is the effective Young’s modulus of the “composite”
beam and therefore a lower but nonzero slope of the f–w curves is observed in this
region. At temperatures higher than the natural transition temperature, similar scenario
exists except that now new insulating domains nucleate out of the metallic phase in the
top portion of the nanobeam as a result of the maximum tensile stress there [26].

1.4.5 New Phase Stabilization with Strain

The role of the structural transition from the monoclinic to rutile phases during the
MIT in VO2 has been well documented in literature. This monoclinic phase is
typically referred to as M1 phase. It is known that strain can complicate the phase
transition and induce another monoclinic structure of VO2, the M2 phase [34].
16 J. Cao and J. Wu

The structure of M1 phase is characterized by a dimerization of vanadium atoms and


a tilt of these pairs with respect to the rutile cR -axis. In M2 structure, however, only
half of the vanadium atoms dimerize while the other half form zigzag chains. M2
structure has been widely observed in Cr-doped samples [31, 32, 46]. However, due
to the difficulty in accessing the undoped M2 phase, it has not been established how
conductive it is and what role it plays in the MIT of VO2 [31]. In fact, the free energy
of M2 in undoped VO2 is believed to be very close to that of M1 around TC0 , making it
difficult to stabilize a pure M2 phase [34]. Recent experiments show more evidence
that under certain strain state M2 can act as a transitional structure for the phase
transition from M1 to the R phase [18, 27–29, 47]. Extensive measurements using
infrared scattering-scanning near-field optical microscopy, Raman spectroscopy, and
electrical transport have been performed to probe these phase transitions [18, 27, 29].
The complicated phase diagram covering all three phases (M1, M2, and R) has been
mapped out using VO2 beams taking advantage of their single crystallinity and
superior elastic flexibility [29].
During the vapor-transport synthesis [35], single-crystal VO2 beams grow on
molten SiO2 surface and could be clamped in different strain states varying from
beam to beam, depending on local conditions during the cooling process.
High-resolution SEM imaging of these two types of microbeams revealed new
information, as shown in Fig. 1.6a, b. The free-standing beams always have
smooth, featureless surface that does not change as temperature increases.
The bottom-clamped beams, however, often show periodic stripes well aligned
in parallel to the cR direction at a period of ~150 nm. At room temperature these
stripes distribute over the entire length of the beam. With increasing temperature,
new domains with featureless surface start to nucleate out of the striped phase and
eventually eliminate the stripes at high temperatures. AFM shows that a weak
surface corrugation is responsible for the striped contrast in the SEM images.
Figure 1.6c shows the morphology of the beam surface obtained at 333 K using
AFM. The surface corrugation has an amplitude of 0.7 nm, corresponding to a
corrugation angle of ~0.6 . mXRD was employed to elucidate the crystal structure
of the submicron-sized domains along the VO2 microbeams. For free-standing
beams with no surface corrugation, Laue pattern of mXRD shows M1 structure at
T<TC0 and R structure at T>TC0 as expected. For the bottom-clamped beams that
show striped phase, however, a splitting of the mXRD spots was observed in the
Laue pattern, as shown in Fig. 1.6d. The VO2 spots were indexed to the M2
structure with two polysynthetic twins identified and labeled as M2a and M2b as
shown in Fig. 1.6d. This configuration resembles the structures observed in,
for example, polysynthetic twins in ferroelectric Rochelle salts [48, 49] and
twinning superlattices recently found in doped InP nanowires [50].
The appearance of M2 phase during the transition was also demonstrated
by Raman spectroscopy [18, 27]. The three structural phases (M1, M2, and R) of
VO2 can be differentiated by Raman spectra through the distinct phonon modes at
608 cm1 (M1 only), 645 cm1 (M2 only), or featureless spectrum (metallic, R).
With increasing temperature, it is found that VO2 beam initially in M1 phase
1 New Opportunities on Phase Transitions of Correlated Electron Nanostructures 17

Fig. 1.6 (a) Top-viewed high-resolution SEM image of a free-standing VO2 beam at 298 K
showing featureless surface. No change was observed at elevated temperatures. (b) SEM images of
a bottom-clamped VO2 beam at 298, 323, 333, and 368 K (left to right), showing diminishing of
the striped phase at elevated temperatures. (c) AFM image of the clamped VO2 beam at 333 K,
showing periodic surface corrugation of the striped phase. (d) Typical room-temperature mXRD
pattern of a bottom-clamed VO2 beam indexed as a twinned monoclinic M2 phase (M2a/M2b).
The brightest spots come from the Si substrate, while the remaining ones showing splitting are
from the VO2 beam. The scale bar is 1 mm in (a and b)

will convert to M2 phase before the final conversion to full metallic phase as
M1 ! M1 + R ! M2 + R ! R. While those crystals initially in the M2 phase
will transform to R phase directly without passing the M1 phase as M2 ! M2 +
R ! R [18]. By simultaneously performing transport experiments and Raman
spectroscopy on the same VO2 specimen, it is possible to deduce the electrical
properties of different phases and associate the structural domain formations with
the MIT. The activation energy is reported to be 0.09 eV for M1 phase and 0.43 eV
for M2 phase [27]. By direct injection of charge into clamped devices, a metallic
monoclinic phase was also produced [27]. This demonstrates that electrical and
structural phase transitions in VO2 can be decoupled, and a “true” Mott
metal–insulator transition can be revealed.
The complex stress–temperature phase diagram was determined through a
combination of in situ optical microscope and SEM [29]. Enabled by the superior
mechanical properties of single-crystal beams, the uniaxial stress–temperature phase
diagram of VO2 was expanded more than an order of magnitude wider than that
18 J. Cao and J. Wu

was previously achieved. Electrical resistance across the insulating M1 and M2


and metallic R phases was measured over the extensive phase diagram, and the M2
phase was found to be more resistive than the M1 phase by a factor of 3.
The approach can be applied to other strongly CEMs to explore remote phase
space and reveal and engineer new properties.

1.4.6 Thermoelectric Across the Metal–Insulator


Domain Walls

The metal–insulator domain walls in VO2 beams offer a platform to investigate the
thermoelectric effect of Schottky junctions [30]. High-performance thermoelectric
materials are currently one of the focuses in materials research for energy conversion
technologies [51–54]. A good thermoelectric material should have a relatively high
thermopower (Seebeck coefficient) [51, 55]. Interfacing different materials has
been proposed as a strategy to enhance the Seebeck effect from that of the constituent
materials alone. Effect of single or a few Schottky junctions on the Seebeck coeffi-
cient has not been experimentally tested, mainly due to the lack of materials suitable
for accurate determination of the small change in the Seebeck coefficient.
Accompanied with the formation of the multiple metal–insulator domains in VO2
microbeams, one or a few Schottky junctions can be reversibly created and
eliminated in the plane perpendicular to the current and heat flow direction.
This offers a material platform where the thermoelectric effect can be measured
from the same specimen with or without the Schottky junction, so that an accurate
extraction of the net junction effect becomes possible [30].
As schematically shown in Fig. 1.7a, the VO2 beam is in pure insulating phase
at low temperatures (<323 K) and pure metallic phase at high temperatures
(>373 K). At intermediate temperatures (323–373 K), both phases coexist. The total
resistance (Rtotal ) of the middle segment of the VO2 beam was measured as a
function of T in a four-probe geometry. Rtotal in the pure insulating phase (namely,
T<323 K) is RI and was fitted using the standard equation for semiconductors,
RI ðTÞ ¼ R0I expðEa =kB T Þ. This fitted RI ðTÞ was extrapolated to higher tem-
peratures, from which the fraction of the insulating phase in the middle segment
was calculated using xðTÞ ¼ Rtotal ðTÞ=RI ðTÞ.
In the pure insulating or metallic phase regimes, the Seebeck coefficient SðTÞ was
measured directly and agreed well with the values in bulk VO2 [30, 56]. In the phase
coexisting regime, the Seebeck coefficient can be either measured directly from the
devices or predicted by a linear combination of contributions from the insulating and
metallic domains. If one neglects the contribution from the metal–insulator domain
walls, the total Seebeck voltage is expected to be a sum of contributions from both
domains following the equation Stotal ðTÞ ¼ xðTÞSI ðTÞ þ ½1  xðTÞSM , where SI ðTÞ
is extrapolated from the pure insulating phase regime and SM ðTÞ is constant over
the temperature range. The measured and expected Stotal ðTÞ were plotted in Fig. 1.7c.
1 New Opportunities on Phase Transitions of Correlated Electron Nanostructures 19

Fig. 1.7 (a) Schematics of multiple M–I domains forming along the device at intermediate
temperatures, showing the creation and elimination of one or a few Schottky junctions on the
same specimen. (b) Four-probe resistance of a VO2 beam taken right after the Seebeck voltage
measurement at each global temperature. Solid line is a fit of the resistance in pure I phase with
standard equation and extrapolated to 393 K. (c) Seebeck coefficient of a VO2 beam measured as a
function of temperature. Solid blue line is the Seebeck coefficient expected from the measured
resistance in (b)

It can be seen that in the multiple metal–insulator domain regime, the measured
Stotal ðTÞ is significantly lower than the expected value, differing by up to a factor of 2.
The exact reason is still under investigation but is believed to be related to the
correlated electron nature of the system [30].

1.5 Conclusions

In this chapter, we have presented a comprehensive review of recent reports on VO2


nanostructures. VO2 is a prototypical CEM that exhibits coupled electrical and
structural phase transitions. Probing the transition and domain physics in vast phase
spaces allows one to reveal many fundamental mechanisms underlying the transition.
Despite decades of investigations, there are still many unresolved questions in VO2
that are largely associated with the electronic phase separation and multiple domain
structures near the phase transition. The unique size and extremely flexible
mechanical properties of VO2 beams (nanometer – or micrometer-sized beams)
offer a platform to investigate many fundamental properties of VO2. It is
demonstrated that lattice strain is responsible for the phase inhomogeneity in VO2
and can be employed to control and manipulate metal–insulator domains.
The fundamental driving mechanism of the phase transition in VO2 is more
20 J. Cao and J. Wu

electron–electron correlation related than electron–phonon interactions. The superior


mechanical properties of VO2 beams can be used to probe the superelasticity of the
solid-state phase transition and expand the uniaxial stress–temperature phase
diagram. Stress was employed to stabilize M2 phases of VO2 that is otherwise
thermodynamically inaccessible. In the end, we demonstrate that the unique ordered
domain patterns in VO2 can be used to investigate the thermoelectric properties of
Schottky junctions.
Research on nanostructured VO2 has revealed many interesting new aspects
that cannot be achieved in its bulk or thin film counterparts. The probing methods
and techniques in VO2 beams can be extended to other CEMs upon the successful
synthesis of desired nanostructures. Systematic investigations, especially in situ
experiments, are key components to resolve many complicated questions such as
the origin of phase inhomogeneity in manganites, flux quanta in high-TC
superconductors, domain wall conductivity in ferroelectrics, and many other
unique properties in CEMs. CEM nanostructures can sustain an extraordinary
amount of uniaxial strain as well as be subjected to continuously tunable external
stress, making them suitable for a variety of in situ experiments. The methods
discussed in this chapter offer new insights into and opportunities to the study of
materials system in condensed matter physics community.

Acknowledgment We greatly acknowledge the financial support of National Science Foundation


under Grant No. EEC-0425914.

References

1. Cox, P.A.: Transition Metal Oxides: An Introduction to Their Electronic Structure and
Properties. Oxford University Press, Oxford (1992)
2. Balamurugan, B., et al.: Size-dependent conductivity-type inversion in Cu2O nanoparticles.
Phys. Rev. B 69(16), 165419 (2004)
3. Spaldin, N.A., Fiebig, M.: The renaissance of magnetoelectric multiferroics. Science 309
(5733), 391–392 (2005)
4. Dagotto, E.: Complexity in strongly correlated electronic systems. Science 309(5732),
257–262 (2005)
5. Cao, J., et al.: Strain engineering and one-dimensional organization of metal-insulator domains
in single-crystal VO2 beams. Nat. Nanotechnol. 4, 732–737 (2009)
6. Mathur, N., Littlewood, P.: Mesoscopic texture in manganites. Phys. Today 56(1), 25–30
(2003)
7. Dagotto, E.: Nanoscale Phase Separation and Colossal Magnetoresistance. Springer,
Berlin (2002)
8. Dagotto, E.: Open questions in CMR manganites, relevance of clustered states and analogies
with other compounds including the cuprates. New J. Phys. 7, 67 (2005)
9. Shenoy, V.B., Sarma, D.D., Rao, C.N.R.: Electronic phase separation in correlated oxides: The
phenomenon, its present status and future prospects. Chemphyschem 7(10), 2053–2059 (2006)
10. Pan, S.H., et al.: Microscopic electronic inhomogeneity in the high-Tc superconductor
Bi2Sr2CaCu2O8+x. Nature 413(6853), 282–285 (2001)
1 New Opportunities on Phase Transitions of Correlated Electron Nanostructures 21

11. Moreo, A., Yunoki, S., Dagotto, E.: Phase separation scenario for manganese oxides and
related materials. Science 283(5410), 2034–2040 (1999)
12. Burgy, J., Moreo, A., Dagotto, E.: Relevance of cooperative lattice effects and stress fields in
phase-separation theories for CMR manganites. Phys. Rev. Lett. 92(9), 097202 (2004)
13. Ahn, K.H., Lookman, T., Bishop, A.R.: Strain-induced metal-insulator phase coexistence in
perovskite manganites. Nature 428(6981), 401–404 (2004)
14. Milward, G.C., Calderon, M.J., Littlewood, P.B.: Electronically soft phases in manganites.
Nature 433(7026), 607–610 (2005)
15. Dagotto, E., Hotta, T., Moreo, A.: Colossal magnetoresistant materials: The key role of phase
separation. Phys. Rep. 344(1–3), 1–153 (2001)
16. Ward, T.Z., et al.: Elastically driven anisotropic percolation in electronic phase-separated
manganites. Nat. Phys. 5, 885–888 (2009)
17. Qazilbash, M.M., et al.: Mott transition in VO2 revealed by infrared spectroscopy and
nano-imaging. Science 318(5857), 1750–1753 (2007)
18. Jones, A., et al.: Nano-optical investigations of the metal-insulator phase behavior of
individual VO2 microcrystals. Nano Lett. 10(5), 1574–1581 (2010)
19. Wu, J., et al.: Strain-induced self organization of metal-insulator domains in single-crystalline
VO2 nanobeams. Nano Lett. 6(10), 2313–2317 (2006)
20. Gu, Q., et al.: Current-driven phase oscillation and domain-wall propagation in WxV1-xO2
nanobeams. Nano Lett. 7(2), 363–366 (2007)
21. Wei, J., et al.: New aspects of the metal-insulator transition in single-domain vanadium dioxide
nanobeams. Nat. Nanotechnol. 4, 420–424 (2009)
22. Wentzcovitch, R.M., Schulz, W.W., Allen, P.B.: VO2: Peierls or Mott-Hubbard? A view from
band theory. Phys. Rev. Lett. 72(21), 3389–3392 (1994)
23. Rice, T.M., Launois, H., Pouget, J.P.: Comment on “VO2: Peierls or Mott-Hubbard? A view
from band theory”. Phys. Rev. Lett. 73(22), 3042 (1994)
24. Cavalleri, A., et al.: Evidence for a structurally-driven insulator-to-metal transition in VO2:
A view from the ultrafast timescale. Phys. Rev. B 70(16), 161102 (2004)
25. Cao, J., et al.: Constant threshold resistivity in the metal-insulator transition of VO2: Phys.
Rev. B, Rapid Commun. 82, 241101 (2010)
26. Fan, W., et al.: Superelastic metal-insulator phase transition in single-crystal VO2 nanobeams.
Phys. Rev. B Condens. Matter Mater. Phys. 80, 241105(R) (2009)
27. Zhang, S., Chou, J.Y., Lauhon, L.J.: Direct correlation of structural domain formation with the
metal insulator transition in a VO2 nanobeam. Nano Lett. 9(12), 4527–4532 (2009)
28. Sohn, J.I., et al.: Surface-stress-induced Mott transition and nature of associated spatial phase
transition in single crystalline VO2 nanowires. Nano Lett. 9(10), 3392–3397 (2009)
29. Cao, J., et al.: Mapping and exploration of extensive stress-temperature phase diagram of
vanadium dioxide. Nano Lett. 10(7), 2667–2673 (2009)
30. Cao, J., et al.: Thermoelectric effect across the metal  insulator domain walls in VO2
microbeams. Nano Lett. 9, 4001–4006 (2009)
31. Eyert, V.: The metal-insulator transitions of VO2: A band theoretical approach. Ann. Phys. 11
(9), 650–704 (2002)
32. Marezio, M., et al.: Structural aspects of the metal-insulator transitions in Cr-doped VO2. Phys.
Rev. B 5(7), 2541–2551 (1972)
33. Rakotoniaina, J.C., et al.: The thermochromic vanadium dioxide: I. Role of stresses and
substitution on switching properties. J. Solid State Chem. 103(1), 81–94 (1993)
34. Pouget, J.P., et al.: Electron localization induced by uniaxial stress in pure VO2. Phys. Rev.
Lett. 35(13), 873 (1975)
35. Guiton, B.S., et al.: Single-crystalline vanadium dioxide nanowires with rectangular cross
sections. J. Am. Chem. Soc. 127(2), 498–499 (2005)
36. Maurer, D., Leue, A.: Investigation of transition metal oxides by ultrasonic microscopy. Mater.
Sci. Eng. A 370(1–2), 440–443 (2004)
22 J. Cao and J. Wu

37. Kim, H.-T., et al.: Mechanism and observation of Mott transition in VO2-based two- and
three-terminal devices. New J. Phys. 6, 52 (2004)
38. Kim, H.-T., et al.: Monoclinic and correlated metal phase in VO2 as evidence of the Mott
transition: Coherent phonon analysis. Phys. Rev. Lett. 97(26), 266401 (2006)
39. Roytburd, A.L.: Thermodynamics of polydomain heterostructures. II. Effect of microstresses.
J. Appl. Phys. 83(1), 239–245 (1998)
40. Biermann, S., et al.: Dynamical singlets and correlation-assisted Peierls transition in VO2.
Phys. Rev. Lett. 94(2), 026404 (2005)
41. Qazilbash, M.M., et al.: Correlated metallic state of vanadium dioxide. Phys. Rev. B Condens.
Matt. Mater. Phys. 74(20), 205118 (2006)
42. Hilton, D.J., et al.: Enhanced photosusceptibility near TC for the light-induced insulator-to-
metal phase transition in vanadium dioxide. Phys. Rev. Lett. 99(22), 226401 (2007)
43. Natelson, D.: Correlated electron systems: Better than average. Nat. Nanotechnol. 4(7),
406–407 (2009)
44. Otsuka, K., Wayman, C.: Shape Memory Materials. Cambridge University Press, New York
(1999)
45. McQuarrie, D.A., Simon, J.D.: Physcial Chemistry: A Molecular Approach. University
Science, Sausalito (1997)
46. Pouget, J.P., et al.: Dimerization of a linear Heisenberg chain in the insulating phases of
V1xCrxO2. Phys. Rev. B 10(5), 1801 (1974)
47. Booth, J.M., Casey, P.S.: Production of VO2 M1 and M2 nanoparticles and composites and the
influence of the substrate on the structural phase transition. ACS Appl. Mater. Interfaces 1(9),
1899–1905 (2009)
48. Mitsui, T., Furuichi, J.: Domain structure of rochelle salt and KH2PO4. Phys. Rev. 90(2), 193
(1953)
49. Klassen-Neklyudova, M.V.: Mechanical Twinning of Crystals. Consultants Bureau, New York
(1964)
50. Algra, R.E., et al.: Twinning superlattices in indium phosphide nanowires. Nature 456(7220),
369–372 (2008)
51. Majumdar, A.: Materials science. Thermoelectricity in semiconductor nanostructures. Science
303, 777 (2004)
52. Boukai, A.I., et al.: Silicon nanowires as efficient thermoelectric materials. Nature 451, 168
(2008)
53. Hochbaum, A.I., et al.: Rough silicon nanowires as high performance thermoelectric materials.
Nature 451, 163 (2008)
54. Heremans, J.P., et al.: Enhancement of thermoelectric efficiency in PbTe by distortion of the
electronic density of states. Science 321, 554 (2008)
55. Mahan, G.D., Sofo, J.O.: The best thermoelectric. Proc. Natl Acad. Sci. USA 93, 7436 (1996)
56. Berglund, C.N., Guggenheim, H.J.: Electronic properties of VO2 near the semiconductor-
metal transition. Phys. Rev. 185(3), 1022 (1969)
Chapter 2
Controlling the Conductivity in Oxide
Semiconductors

A. Janotti, J.B. Varley, J.L. Lyons, and C.G. Van de Walle

2.1 Introduction

Oxide semiconductors occur in a variety of crystal structures and exhibit diverse


electronic and optical properties. Controlling the electrical conductivity in oxide
thin films and nanostructures is an important step toward their application in
electronics and optoelectronics. Here we discuss the results of first-principles studies
of the effects of native point defects and impurities on the electronic properties of
semiconducting oxides such as ZnO, SnO2, and TiO2. We address the possible
causes of the often observed unintentional n-type conductivity in these oxides
and the prospects of achieving p-type conductivity. In the case of ZnO and SnO2,
it is found that the unintentional conductivity is not due to oxygen vacancies or
cation interstitials, but rather to the incorporation of donor impurities, with hydrogen
being a likely candidate. Although the calculations were aimed at understanding
the behavior of defects and impurities in bulk single crystals, the main results and
conclusions are expected to be valid for thin films and nanostructures.
ZnO, SnO2, and TiO2 are wide-band-gap materials and can be made highly
conductive [1–8]. Yet the control of conductivity poses serious challenges and
constitutes an important step toward the development of electronic devices based
on oxide thin films or nanostructures. It has long been thought that intrinsic defects
such as oxygen vacancies are responsible for the observed n-type conductivity,
largely based on experiments in which the conductivity is measured as a function of
the oxygen content in the annealing environment [9–15]. Still the identification

A. Janotti (*) • J.L. Lyons • C.G. Van de Walle


Materials Department, University of California, Santa Barbara, CA, USA
e-mail: janotti@engineering.ucsb.edu
J.B. Varley
Materials Department, University of California, Santa Barbara, CA, USA
Department of Physics, University of California, Santa Barbara, CA, USA

J. Wu et al. (eds.), Functional Metal Oxide Nanostructures, Springer Series 23


in Materials Science 149, DOI 10.1007/978-1-4419-9931-3_2,
# Springer Science+Business Media, LLC 2012
24 A. Janotti et al.

of oxygen vacancies has remained quite elusive, and recent efforts to enhance the
performance of these oxides have highlighted the fact that the causes and
mechanisms of conduction are poorly understood. First-principles calculations for
ZnO and SnO2 have casted severe doubts on the hypothesis that oxygen vacancies
are sources of conductivity. The results indicate that oxygen vacancies are deep
rather than shallow donors and cannot cause conductivity [16–19].
Recent studies indicate that unintentional impurities are most likely the source
of the observed unintentional conductivity in oxides [4, 5, 19–26]. Most growth
techniques introduce impurities through the sources or as contaminants; even in
ultrahigh vacuum, impurities such as hydrogen are present at high enough levels
to incorporate in sizable concentrations in materials in which their solubility is
high [22]. Hydrogen is indeed a particularly insidious impurity in this respect,
since it is notoriously difficult to detect experimentally.
Based on first-principles calculations it has been suggested that interstitial hydro-
gen is a plausible cause of unintentional doping in ZnO [20–22], a proposal now
confirmed by numerous experimental studies [27–31]. Later, it has been proposed
that two forms of hydrogen can act as electrically active impurities: interstitial
hydrogen, which prefers to attach to an oxygen host atom and diffuses easily, and
substitutional hydrogen on an oxygen site, which is more stable and can alternatively
be regarded as a complex consisting of hydrogen and an oxygen vacancy [23]. Both
these species were predicted to act as shallow donors in ZnO [23] and SnO2 [19, 24].
These predictions have been recently confirmed by experiments [32, 33].
In Sect. 2.2, the formalism for calculating defect formation energies and the
computational approach are described; Sect. 2.3 addresses the electronic properties
of native defects and impurities in ZnO, SnO2 and TiO2, and Sect. 2.4 concludes the
paper.

2.2 Formalism and Computational Approach

The formation energy is a key quantity in the description of the electronic


structure and stability of point defects and impurities in solids. Defects that
occur in low concentrations have a small or negligible impact on conductivity;
only those whose concentration exceeds a threshold will have observable effects.
The concentration is determined by the formation energy through the expression

c ¼ Nsites expðEf =kTÞ; (2.1)

where Ef is the formation energy, Nsites is the number of sites on which the defect
can be incorporated, k is the Boltzmann constant, and T is the temperature.
To illustrate the definition of the formation energy for a point defect [34–36],
we take the specific example of an oxygen vacancy in a 2+ charge state in ZnO:

Ef ðVO2þ Þ ¼ Etot ðVO2þ Þ  Etot ðZnOÞ þ mO þ 2EF ; (2.2)


2 Controlling the Conductivity in Oxide Semiconductors 25

where Etot ðVOq Þ is the total energy of the supercell containing the defect,
and Etot(ZnO) is the total energy of the ZnO perfect crystal in the same supercell.
The Fermi energy EF is the energy of the reservoir in the solid, with which
electrons are exchanged. The oxygen atom that is removed is placed in a reservoir
with energy mO, i.e., the oxygen chemical potential. We note that mO is a variable
in the formalism, corresponding to notion that ZnO can be grown under conditions
that vary from the oxygen-rich to the oxygen-poor limit. The oxygen chemical
potential mO is subject to an upper bound equal to the energy per atom of an O2
molecule. The sum of mO and mZn corresponds to the energy of ZnO, which is
essentially the stability condition of ZnO. An upper bound on mZn, set by the
energy per atom of bulk Zn, therefore leads to a lower bound on mO. Therefore,
the range over which the chemical potentials can vary is given by the enthalpy of
formation of ZnO (exp.: 3.60 eV [37]).
Defects are usually electrically active, occurring in charge states other than
neutral. For each position of the Fermi level, one particular charge state has the
lowest energy. The Fermi-level positions at which the lowest energy charge state
changes are called transition levels. Once the formation energies are known,
the transition levels immediately follow by taking energy differences:

eðq=q0 Þ ¼ ½Ef ðDq ; EF ¼ 0Þ  Ef ðDq 0 ; EF ¼ 0Þ=ðq0  qÞ; (2.3)

where Ef(Dq; EF ¼ 0) is the formation energy of the defect D in the charge state q
when the Fermi level is at the valence-band maximum (EF ¼ 0). When atomic
relaxations are fully included in the calculation of the formation energies for both
charge states, a thermodynamic transition level is obtained. The experimental
significance of this level is that for Fermi-level positions below e(q/q0 ) charge
state q is stable, while for Fermi-level positions above e(q/q0 ), charge state q0
is stable. The transition levels should not be confused with the single-particle
Kohn-Sham states that result from band-structure calculations for a single charge
state. We also note that in optical experiments (luminescence or absorption)
the final state may not be completely relaxed, leading to different values for
optical levels [35].
Formation energies, such as in (2.2), can be explicitly calculated based on
density functional theory (DFT) [38] calculations. DFT calculations have tradition-
ally used the local density approximation (LDA) [39] or generalized gradient
approximation (GGA) [40, 41]. The use of DFT-LDA/GGA implies that the band
gap is not properly described, and states within the band gap will therefore be
affected as well. If these states are occupied with electrons, the formation energy of
the defect will also reflect these errors. Several approaches have been developed
to overcome these problems: (1) the LDA + U approach [42], which was used to
correct the semicore d states in ZnO, thus providing a partial correction to the band
gap; in conjunction with LDA, the LDA + U results were used to correct defect
formation energies and transition levels in ZnO [16–18]. (2) The screened hybrid
functional of Heyd, Scuseria, and Ernzerhof (HSE) [43], which is based on the
26 A. Janotti et al.

inclusion of a small fraction of nonlocal exchange in the Hamiltonian within a


sphere of a certain radius, thus describes metals and insulators on the same footing;
this is important since it allows for the computation of formation energies that take
metals as limiting phases in the evaluation of chemical potentials [26, 44]. (3) The
Green’s-function-based quasiparticle GW method [45, 46], which combined with
LDA allowed for precise calculations of defect formation energies and transition
levels, as illustrated with the case of the self-interstitial in Si [47]. The calculations
discussed in this chapter are based on the DFT within the LDA/LDA + U approach,
or the screened hybrid functional (HSE). These calculations made use of projected
augmented wave potentials [48, 49] to separate valence from core electrons, as
implemented in the VASP code [50, 51].

2.3 Results and Discussion

2.3.1 ZnO

Zinc oxide has a direct band gap of 3.4 eV, an exciton binding energy of 60 meV,
and is available as large single crystals [1–5]. As such, ZnO has been considered
a promising material for light emitting diodes and laser diodes that operate in the
blue-UV spectral region, and for high-power, high-frequency, or thin-film transistors.
It has also been demonstrated that ZnO can be made in nanostructures with a variety
of morphologies such as wires, helices, belts, and springs which can be useful in gas
sensors, transducers, and actuators at the nanoscale [52]. However, the development
of ZnO for these various applications has been hindered by a lack of understanding
and difficulties in controlling the electrical conductivity [1–5]. ZnO in bulk and
thin-film forms is almost always n-type [2, 53], the cause of which has been hotly
debated. In addition, many reports on p-type ZnO have appeared in the literature
[54–60], but reliability and reproducibility are still questionable.
The observed n-type conductivity in “undoped” ZnO has long been attributed
to the presence of native point defects such as oxygen vacancies or zinc
interstitials [2, 9–11]. However, the identification of such defects in as-grown
(as opposed to irradiated) material has been rather vague, and the evidence of their
relation to the observed conductivity has always been indirect, e.g., based on the
variation of conductivity with O2 partial pressure in the annealing environment.
On the contrary, first-principles calculations indicate that neither O vacancies nor
Zn interstitials can explain the observed n-type conductivity in ZnO [16–18].
Recent experiments on high-quality bulk single crystals indeed support these
results [61, 62].
The calculated formation energy as a function of the Fermi-level position for
native donor defects in ZnO, under O-poor conditions, is shown in Fig. 2.1. These
calculations were based on a combination of LDA and LDA + U as described in
[17, 18]. The results indicate that oxygen vacancy is a deep donor with a transition
2 Controlling the Conductivity in Oxide Semiconductors 27

Fig. 2.1 Formation energy as


a function of Fermi-level
position for donor centers in
ZnO: substitutional hydrogen
HO, interstitial hydrogen Hi,
oxygen vacancy VO, zinc
interstitial Zni, and zinc
antisite ZnO. These were
calculated according to the
LDA/LDA + U method as
described in [18]. Only the
results for oxygen-poor
conditions are shown. The
zero of Fermi level
corresponds to the valence-
band maximum. The slope
of the line segments indicates
the charge state. The kink
in the formation-energy curve
of VO at EF ¼ 2.4 eV
indicates the e(+2/0)
transition level

level e(2+/0) about 1 eV below the conduction band. Therefore, VO is stable in


the neutral charge state in n-type ZnO and, thus, cannot explain the observed
n-type conductivity. The ionization energy of 1 eV indicates that VO will not be
ionized even at temperatures well above room temperature. The zinc interstitial is
a shallow donor, but it is not thermally stable. It has high formation energy in
n-type ZnO and migrates with an energy barrier of only 0.6 eV [18], i.e., Zn
interstitials are mobile even below room temperature. Zinc antisites (ZnO) are also
shallow donors, stable in the 2+ charge state for Fermi-level positions near the
conduction band. The large off-site displacement of the Zn atom indicates that
Zn2þ 0 2þ
O is actually a complex of VO and Zni . The high formation energy in n-type

ZnO indicates that ZnO is unlikely to play a role in the observed unintentional
conductivity in as-grown or annealed materials, unless Zn2þ O is created by
nonequilibrium processes such as irradiation. Similar results and conclusions
based on more sophisticated and computationally demanding hybrid functional
methods have been published more recently [63].
Having established that native defects cannot explain the observed unintentional
n-type conductivity in ZnO, one has to consider the role of impurities that are
most likely to be present in different growth environments and act as donors.
One such impurity is hydrogen. First-principles calculations have shown that
interstitial hydrogen behaves as a shallow donor in ZnO [20, 21]. This is some-
what counterintuitive because hydrogen typically acts as a passivating agent in
semiconductors, reducing the electrical conductivity rather than being a source of
doping. The theoretical prediction was quickly confirmed in numerous experiments
(summarized in [4]).
28 A. Janotti et al.

Nevertheless, interstitial hydrogen is highly mobile [64, 65] and thus can be
removed from ZnO by annealing at relatively modest temperatures (150 C).
There are clear experimental indications, however, that hydrogen also exists as a
more thermally stable donor that persists upon annealing [29, 66] at temperatures up
to 500 C. Based on first-principles calculations it has been proposed that this
additional hydrogen-related donor species consists of a hydrogen atom occupying a
substitutional oxygen site [23]. The substitutional hydrogen species is more stable
than interstitial hydrogen in ZnO, with a diffusion barrier consistent with the
observed reduction in hydrogen activity above 500 C [29, 66].
The calculated formation energies of interstitial hydrogen Hi and substitutional
hydrogen HO are also shown in Fig. 2.1. Interstitial hydrogen strongly bonds to
oxygen, by breaking a Zn–O bond. It can occupy different positions in the ZnO
lattice: bond-center and antibonding sites next to oxygen, parallel to the c-axis or
forming an angle of about 112 with the c-axis. In all these configurations,
interstitial hydrogen results in effective-mass shallow donor levels, and has
similar formation energies, within less than 0.2 eV. We find that the bond-center
configuration with an O–H distance of 1.05 Å gives the lowest formation energy.
Substitutional hydrogen HO, on the other hand, bonds equally to all four
nearest-neighbor zinc atoms in a multicenter bond configuration and also results
in an effective-mass shallow donor. Its formation energy is only  0.1 eV higher
than that of interstitial hydrogen in oxygen-poor conditions. The electronic
structure and bonding properties of HO were discussed in [23]. The formation
energy, and hence the solubility of substitutional hydrogen, is consistent with
observed concentrations; furthermore, because it replaces oxygen, hydrogen can
also explain the observed dependence of unintentional n-type conductivity on the
oxygen partial pressure in the growth or annealing environments [23].
It has been recently pointed out that other impurities such as Al, Ga, and Si
are also present in as-grown ZnO single crystals and act as donors [4]. Al and
Ga substitute for Zn and act as shallow donors [67, 68], but are not observed
in high enough concentrations to explain the unintentional n-type conductivity in
bulk single crystals. On the other hand, Si is a double donor when substituting
on Zn site [26] and has been found in concentrations that are compatible
with free electron concentrations in ZnO single crystals grown by different
techniques [25].
Note that controlling the n-type conductivity is a necessary step toward
achieving the so-desired p-type conductivity in ZnO. Despite many reports in
the literature, reliable p-type doping of ZnO remains difficult. Low solubility of
p-type dopants and the compensation by abundant donor defects and impurities
are often raised as the main issues. Known acceptor impurities include group-I
elements Li, Na, K, Cu, and Ag, and group-V elements N, P, and As. However,
many of these form deep acceptors and do not result in p-type conduction at room
temperature. Even N, which has been regarded as the most promising p-type
dopant in ZnO, has recently been shown to act as a deep acceptor with ionization
energy over 1 eV. Calculated results for N-related absorption energy are in good
agreement with experimental observations [26].
2 Controlling the Conductivity in Oxide Semiconductors 29

As an alternative route, it has been proposed that p-type conductivity can


be achieved by the selective incorporation of interstitial fluorine impurities in
ZnO [69]. The F interstitial would complete its octet by extracting an electron
from the valence-band maximum. The resulting hole would be bounded to the
F impurity in an effective-mass state, as a typical shallow acceptor. It is anticipated
that technical difficulties may arise in attempting to selectively introduce interstitial
F as majority defects.

2.3.2 SnO2

Tin dioxide crystallizes in the rutile structure and has a direct band gap of
3.6 eV [12]. The ease of making it n-type, its highly dispersive conduction
band (small effective mass), and the large energy difference between the
conduction-band minimum and the next higher conduction band at G contribute
to SnO2 supporting high carrier concentrations while still maintaining a high
degree of optical transparency [12]. The fabrication of SnO2 nanowires and
nanobelts has been reported, and gas sensors based on these nanostructures have
been demonstrated [70, 71]. These applications crucially depend on the transport
of electrons and the control of conductivity. As in ZnO, an unintentional n-type
conductivity in SnO2 is likely caused by impurities.
SnO2 can be doped n-type by adding impurities such as Sb or F, which incorpo-
rate on Sn and O sites, respectively [12]. In addition, it has been widely believed
that oxygen vacancies are also a source of n-type conductivity. In analogy to ZnO,
the evidence for oxygen vacancies has been based on measurements of conductivity
as a function of the oxygen partial pressure in annealing experiments: increasing
the oxygen partial pressure leads to lower conductivities [12, 72–75]. However,
the attribution of conductivity to oxygen vacancies is not supported by recent
first-principles calculations [19, 24].
In Fig. 2.2 we show the calculated formation energies of donor native point
defects in SnO2. These results were obtained from a combination of LDA and
LDA + U calculations as described in [19]. As shown in Fig. 2.2, oxygen vacancy
is a deep donor, occurring in the neutral charge state if the Fermi level
is positioned near the conduction-band minimum. It has also been concluded
that Sn interstitials and Sn antisites are unlikely sources of conductivity due to
their high formation energies. Therefore, the unintentional n-type conductivity is
probably caused by the presence of impurities. As shown in Fig. 2.2, it is found
that hydrogen in either the interstitial form or substituting for oxygen has also
been predicted to act as a shallow donor in SnO2 [19, 24].
Reports on p-type SnO2 have been scarce. It has been proposed that Ga or In
substituting on Sn site would result in shallow acceptors. We anticipate that
30 A. Janotti et al.

Fig. 2.2 Formation energy as


a function of Fermi-level
position for donor-type
centers in SnO2: oxygen
vacancy VO, tin interstitial
Sni, tin antisite SnO, hydrogen
interstitial Hi, and
substitutional hydrogen HO.
These results were obtained
using the LDA/LDA + U
approach [19]. The zero of
Fermi level corresponds to the
valence-band maximum. For
Fermi-level positions near the
conduction band VO is stable
in the neutral charge state
whereas Sni and SnO are
stable in the 4+ charge state

difficulties in making SnO2 p-type may arise due to the formation of small hole
polarons. The valence band of SnO2 is quite flat, and instabilities of the O lattice
atoms near the impurities may favor the formation of bound hole polarons.

2.3.3 TiO2

Titania is most stable in the rutile crystal structure and has a band gap of 3.1 eV
[37]. The upper part of the valence band is composed of O 2p states, and the lower
part of the conduction band of Ti 3d states [44]. TiO2 can be made n-type by the
incorporation of shallow donor impurities (e.g., Nb, F, and H) and by annealing in
reducing environments [8, 14]. Because its conductivity varies with O2 partial
pressure, it is often argued that oxygen vacancies and/or titanium interstitials are
sources of conductivity in TiO2 [8, 14].
In Fig. 2.3 we show the calculated formation energies for oxygen vacancies in
TiO2, based on HSE hybrid functional calculations [44]. It was concluded that
oxygen vacancies are shallow donors, with VOþ and VO0 higher in energy than VO2þ
for any value of the Fermi level within the band gap [44]. The formation energy of
VO2þ in the extreme oxygen-poor limit is relatively low even when the Fermi level
is positioned near the conduction-band minimum. This might lead to the conclu-
sion that oxygen vacancies are the cause of conductivity in vacuum-annealed
TiO2. However, care should be taken, since the extreme oxygen-poor limit
corresponds to oxygen partial pressures that are not experimentally accessible.
We also need to keep in mind that impurities that act as shallow donors, such as
hydrogen, also likely contribute to the observed conductivity [76].
2 Controlling the Conductivity in Oxide Semiconductors 31

Fig. 2.3 Formation energy as


a function of Fermi-level
position for the oxygen
vacancy VO in TiO2. Only the
results for the O-poor limit
are shown. These were
obtained using the HSE
hybrid functional. The zero of
Fermi level corresponds to the
valence-band maximum.
VO2þ is lower in energy than
VOþ and VO0 even for the Fermi
level positioned at the
conduction-band minimum

2.4 Concluding Remarks

We have discussed the causes of conductivity in ZnO, SnO2, and TiO2, which
are representatives of a large family of wide-band-gap oxide semiconductors.
First-principles computational studies indicate that the conventional wisdom of
assigning the cause of unintentional conductivity in these materials to the presence
of native point defects, such as oxygen vacancies, must be reviewed. In particular it
is shown that oxygen vacancies in ZnO and SnO2 cannot explain the observed
conductivity because they are deep donors. Therefore, the widespread view that
oxygen vacancies are correlated with conductivity or are easily created during
growth or annealing in oxygen-poor, reducing environments must be seriously
reconsidered. In addition, first-principles calculations reveal that other native point
defects, e.g., cation interstitials, are also unlikely sources of conductivity since they
are not thermally stable.
Instead, it is argued that unintentional n-type conductivity in oxide semiconductors
is more likely caused by the unintentional incorporation of impurities, hydrogen being
a prime candidate. Hydrogen is an ubiquitous impurity, being a fast diffuser as an
interstitial impurity and also capable of assuming substitutional positions. In both
forms, interstitial and substitutional, hydrogen has been predicted to act as shallow
donor in these materials. These findings are quite unexpected and have important
technological implications, and have been confirmed in the case of ZnO. More
experiments need to be performed in the case of SnO2 and TiO2.
32 A. Janotti et al.

We also would like to emphasize that hydrogen is by no means the only


shallow donor impurity that can be unintentionally incorporated; many other
impurities may act as shallow donors, such as those on the right (n-type) side
(in the Periodic Table of Elements) of the element being substituted, although it is
quite unlikely that they are present in all growth or annealing environments.
Yet hydrogen is present in almost all growth environments, generally hard to
detect experimentally, and is not usually considered as a potential donor.
Therefore, special attention should be devoted to its potential effects on the
electronic properties of oxide semiconductors.

Acknowledgments This work was supported by the NSF MRSEC Program under award No.
DMR05-20415 and by Saint-Gobain Research. Collaborations with M. D. McCluskey,
M. Scheffler, A. K. Singh, N. Umezawa, P. Rinke, G. Kresse are gratefully acknowledged.

References

1. Jagadish, C., Pearton, S.J. (eds.): Zinc Oxide Bulk, Thin Films, and Nanostructures. Elsevier,
New York (2006)
2. Look, D.C.: Recent advances in ZnO materials and devices. Mater. Sci. Eng. B 80, 383 (2001)
€ Alivov, Y.I., Liu, C., Teke, A., Reshchikov, M.A., Dogan, S., Avrutin, V., Cho,
3. Ozg€ur, U.,
S.-J., Morkoç, H.: A comprehensive review of ZnO materials and devices. J. Appl. Phys. 98,
041301 (2005)
4. Janotti, A., Van de Walle, C.G.: Fundamentals of zinc oxide as a semiconductor. Rep. Prog.
Phys. 72, 126501 (2009)
5. McCluskey, M.D., Jokela, S.J.: Defects in ZnO. J. Appl. Phys. 106, 071101 (2009)
6. Gordon, R.G.: Criteria for choosing transparent conductors. Mater. Res. Soc. Bull. 25, 52
(2000)
7. Hosono, H.: Recent progress in transparent oxide semiconductors: Materials and device
application. Thin Solid Films 515, 6000 (2007)
8. Linsebigler, A.L., Lu, G., Yates Jr., J.T.: Photocatalysis on TiO2 Surfaces: Principles,
Mechanisms, and Selected Results. Chem. Rev. 95, 735 (1995)
9. Kr€oger, F.A.: The Chemistry of Imperfect Crystals. North Holland Publishing, Amsterdam
(1974)
10. Look, D.C., Hemsky, J.W., Sizelove, J.R.: Residual native shallow donor in ZnO. Phys. Rev.
Lett. 82, 2552–2555 (1999)
11. Tomlins, G.W., Routbort, J.L., Mason, T.O.: Zinc self-diffusion, electrical properties, and
defect structure of undoped, single crystal zinc oxide. J. Appl. Phys. Rev. 87, 117–123 (2000)
12. Dawar, A.L., Jain, A.K., Jagadish, C.: Semiconducting Transparent Thin Films. Institute of
Physics, London (1995)
13. Nowotny, J., Radecka, M., Rekas, M., Sugihara, S., Vance, E.R., Weppner, W.: Electronic and
ionic conductivity of TiO2 single crystal within the n-p transition range. Ceram. Int. 24, 571
(1998)
14. Diebold, U.: The surface science of titanium dioxide. Surf. Sci. Rep. 48, 53 (2003)
15. Nowotny, M.K., Bak, T., Nowotny, J.: Electrical properties of single crystal TiO2. I. Electrical
conductivity. J. Phys. Chem. B 110, 16270 (2006)
16. Janotti, A., Van de Walle, C.G.: Oxygen vacancies in ZnO. Appl. Phys. Lett. 87, 122102
(2005)
17. Janotti, A., Van de Walle, C.G.: New insights into the role of native point defects in ZnO.
J. Cryst. Growth 287, 58 (2006)
2 Controlling the Conductivity in Oxide Semiconductors 33

18. Janotti, A., Van de Walle, C.G.: Temperature dependence of Raman scattering in ZnO. Phys.
Rev. B 75, 165202 (2007)
19. Singh, A.K., Janotti, A., Scheffler, M., Van de Walle, C.G.: Sources of electrical conductivity
in SnO2. Phys. Rev. Lett. 101, 055502 (2008)
20. Van de Walle, C.G.: Hydrogen as a cause of doping in zinc oxide. Phys. Rev. Lett. 85, 1012
(2000)
21. Van de Walle, C.G., Neugebauer, J.: Universal alignment of hydrogen levels in
semiconductors, insulators and solutions. Nature 423, 626 (2003)
22. Van de Walle, C.G.: Hydrogen as a shallow center in semiconductors and oxides. Phys. Status
Solidi B 235, 89 (2003)
23. Janotti, A., Van de Walle, C.G.: Hydrogen multicentre bonds. Nat. Mater. 6, 44 (2007)
24. Varley, J.B., Janotti, A., Singh, A.K., Van de Walle, C.G.: Hydrogen interactions with acceptor
impurities in SnO2: First-principles calculations. Phys. Rev. B 79, 245206 (2009)
25. McCluskey, M.D., Jokela, S.J.: Sources of n-type conductivity in ZnO. Phys. B 401–402, 355
(2007)
26. Lyons, J.L., Janotti, A., Van de Walle, C.G.: Role of Si and Ge as impurities in ZnO. Phys.
Rev. B 80, 205113 (2009)
27. Cox, S.F.J., Davis, E.A., Cottrell, S.P., King, P.J.C., Lord, J.S., Gil, J.M., Alberto, H.V., Vilão,
R.C., Piroto Duarte, J., Ayres de Campos, N., Weidinger, A., Lichti, R.L., Irvine, S.J.C.:
Experimental confirmation of the predicted shallow donor hydrogen state in zinc oxide. Phys.
Rev. Lett. 86, 2601 (2001)
28. Lavrov, E.V., B€ orrnert, F., Weber, J., Van de Walle, C.G., Helbig, R.: Hydrogen-related
defects in ZnO studied by infrared absorption spectroscopy. Phys. Rev. B 66, 165205 (2002)
29. Jokela, S.J., McCluskey, M.D.: Structure and stability of O-H donors in ZnO from
high-pressure and infrared spectroscopy. Phys. Rev. B 72, 113201 (2005)
30. Lavrov, E.V., B€orrnert, F., Weber, J.: Dominant hydrogen-oxygen complex in hydrothermally
grown ZnO. Phys. Rev. B 71, 035205 (2005)
31. Alvin Shi, G., Stavola, M., Pearton, S.J., Thieme, M., Lavrov, E.V., Weber, J.: Hydrogen local
modes and shallow donors in ZnO. Phys. Rev. B 72, 195211 (2005)
32. Lavrov, E.V., Herklotz, F., Weber, J.: Identification of two hydrogen donors in ZnO. Phys.
Rev. B 79, 165210 (2009)
33. Hlaing Oo, W.M., Tabatabaei, S., McCluskey, M.D., Varley, J.B., Janotti, A., Van de Walle,
C.G.: Calibrating dipolar interaction in an atomic condensate. Phys. Rev. B 82, 193201 (2010)
34. Van de Walle, C.G., Laks, D.B., Neumark, G.F., Pantelides, S.T.: First-principles calculations
of solubilities and doping limits: Li, Na, and N in ZnSe. Phys. Rev. B 47, 9425 (1993)
35. Van de Walle, C.G., Neugebauer, J.: First-principle calculations for defects and impurities:
Applications to III-nitrides. J. Appl. Phys. 95, 3851 (2004)
36. Van de Walle, C.G., Lyons, J.L., Janotti, A.: Controlling the conductivity of InN. Phys. Status
Solidi A 207, 1024 (2010)
37. Dean, J.A.: Lange’s Handbook of Chemistry, 14th edn. McGraw-Hill, New York (1992)
38. Hohenberg, P., Kohn, W.: Inhomogeneous electron gas. Phys. Rev. 136, B864 (1964)
39. Kohn, W., Sham, L.J.: Self-consistent equations including exchange and correlation effects.
Phys. Rev. 140(4A), A1133–A1138 (1965)
40. Perdew, J.P., Wang, Y.: Liquid-drop model for crystalline metals: Vacancy-formation,
cohesive, and face-dependent surface energies. Phys. Rev. Lett. 66, 508 (1991)
41. Perdew, J.P., Burke, K., Ernzerhof, M.: Generalized gradient approximation made simple.
Phys. Rev. Lett. 77, 3865 (1996)
42. Anisimov, V.I., Aryasetiawan, F., Lichtenstein, A.I.: First-principles calculations of the
electronic structure and spectra of strongly correlated systems: the LDA+U method. J. Phys.
Condens. Matter 9, 767 (1997)
43. Heyd, J., Scuseria, G.E., Ernzerhof, M.: Hybrid functionals based on a screened Coulomb
potential. J. Chem. Phys. 118, 8207 (2003)
34 A. Janotti et al.

44. Janotti, A., Varley, J.B., Rinke, P., Umezawa, N., Kresse, G., Van de Walle, C.G.: Hybrid
functional studies of the oxygen vacancy in TiO2. Phys. Rev. B 81, 085212 (2010)
45. Hedin, L.: New method for calculating the one-particle Green’s function with application to
the electron-gas problem. Phys. Rev. 139, A796–A823 (1965)
46. Godby, R.W., Schl€ uter, M., Sham, L.J.: Accurate exchange-correlation potential for silicon
and its discontinuity on addition of an electron. Phys. Rev. Lett. 56, 2415–2418 (1986)
47. Rinke, P., Janotti, A., Scheffler, M., Van de Walle, C.G.: Defect formation energies without the
band-gap problem: combining density-functional theory and the GW approach for the silicon
self-interstitial. Phys. Rev. Lett. 102, 026402 (2009)
48. Bl€ochl, P.E.: Projector augmented-wave method. Phys. Rev. B 50, 17953 (1994)
49. Kresse, G., Joubert, D.: From ultrasoft psuedopotentials to the projector augmented-wave
method. Phys. Rev. B 59, 1758 (1999)
50. Kresse, G., Furthm€ uller, J.: Efficient iterative schemes for ab initio total-energy calculations
using a plane-wave basis set. Phys. Rev. B 54, 11169 (1996)
51. Kresse, G., Furthm€ uller, J.: Efficiency of ab-initio total energy calculations for metals and
semiconductors using a plane-wave basis set. Comput. Mat. Sci. 6, 15 (1996)
52. Wang, Z.L.: Zinc oxide nanostructures: growth, properties and applications. J. Phys. Condens.
Matter 16, R829–R858 (2004)
53. Look, D.C., Reynolds, D.C., Sizelove, J.R., Jones, R.L., Litton, C.W., Cantwell, G., Harsch,
W.C.: Electrical properties of bulk ZnO. Solid State Commun. 105, 399 (1998)
54. Look, D.C., Reynolds, D.C., Litton, C.W., Jones, R.L., Eason, D.B., Cantwell, G.:
Characterization of homoepitaxial p-type ZnO grown by molecular beam epitaxy. Appl.
Phys. Lett. 81, 1830 (2002)
55. Minegishi, K., Koiwai, Y., Kikuchi, Y., Yano, K., Kasuga, M., Shimizu, A.: Growth of p-type
zinc oxide films by chemical vapor deposition. Jpn. J. Appl. Phys. 2 36, L1453 (1997)
56. Ye, Z.-Z., Lu, J.-G., Chen, H.-H., Zhang, Y.-Z., Wang, L., Zhao, B.-H., Huang, J.-Y.:
Preparation and characterization of p-type ZnO films by DC reactive magnetron sputtering.
J. Cryst. Growth 253, 258 (2003)
57. Kim, K.K., Kim, H.S., Hwang, D.K., Lim, J.H., Park, S.J.: Realization of p-type ZnO thin films
via phosphorous doping and thermal activation of the dopant. Appl. Phys. Lett. 83, 63 (2003)
58. Ryu, Y.R., Lee, T.S., White, H.W.: Properties of arsenic-doped p-type ZnO grown by hybrid
beam deposition. Appl. Phys. Lett. 83, 87 (2003)
59. Xiu, F.X., Yang, Z., Mandalapu, L.J., Zhao, D.T., Liu, J.L.: High-mobility Sb-doped p-type
ZnO by molecular beam epitaxy. Appl. Phys. Lett. 87, 152101 (2005)
60. Tsukazaki, A., Ohtomo, A., Onuma, T., Ohtani, M., Makino, T., Sumiya, M., Ohtani,
K., Chichibu, S.F., Fuke, S., Segawa, Y., Ohno, H., Koinuma, H., Kawasaki, M.: Repeated
temperature modulation epitaxy for p-type doping and light-emitting diode based on ZnO. Nat.
Mater. 4, 42 (2005)
61. Vlasenko, L.S., Watkins, G.D.: Optical detection of electron paramagnetic resonance in
room-temperature electron-irradiated ZnO. Phys. Rev. B 71, 125210 (2005)
62. Wang, X.J., Vlasenko, L.S., Pearton, S.J., Chen, W.M., Buyanova, I.A.: Oxygen and zinc
vacancies in As-grown ZnO single crystals. J. Phys. D Appl. Phys. 42, 175411 (2009)
63. Oba, F., Togo, A., Tanaka, I., Paier, J., Kresse, G.: Defects energetics in ZnO: A hybrid
Hartee-Fock density functional study. Phys. Rev. B 77, 245202 (2008)
64. Thomas, D.G., Lander, J.J.: Hydrogen as a donor in zinc oxide. J. Chem. Phys. 25, 1136 (1956)
65. Wardle, M.G., Goss, J.P., Briddon, P.R.: First-principle study of the diffusion of hydrogen in
ZnO. Phys. Rev. Lett. 96, 205504 (2006)
66. Shi, G.A., Stavola, M., Pearton, S.J., Thieme, M., Lavrov, E.V., Weber, J.: Hydrogen local
modes and shallow donors in ZnO. Phys. Rev. B 72, 195211 (2005)
67. Myong, S.Y., Baik, S.J., Lee, C.H., Cho, W.Y., Lim, K.S.: Extremely transparent and
conductive ZnO: Al thin films prepared by photo-assisted metalorganic chemical vapor
deposition (photo-MOCVD) using AlCl3(6H2O) as new doping material. Jpn. J. Appl. Phys.
2 36, L1078 (1997)
2 Controlling the Conductivity in Oxide Semiconductors 35

68. Ko, H.J., Chen, Y.F., Hong, S.K., Wenisch, H., Yao, T., Look, D.C.: Ga-doped ZnO films
grown on GaN templates by plasma-assisted molecular-beam epitaxy. Appl. Phys. Lett. 77,
3761 (2000)
69. Janotti, A., Snow, E., Van de Walle, C.G.: A pathway to p-type wide-band-gap
semiconductors. Appl. Phys. Lett. 95, 172109 (2009)
70. Kolmakov, A., Klenov, D.O., Lilach, Y., Stemmer, S., Moskovits, M.: Enhanced gas sensing
by individual SnO2 nanowires and nanobelts functionalized with Pd catalyst particles. Nano
Lett. 5, 667 (2005)
71. Baik, J., Zielke, M., Kim, M.H., Turner, K.L., Wodtke, A.M., Moskovits, M.:
Tin-oxide-nanowire-based electronic nose using heterogeneous catalysis as a functionalization
strategy. ACS Nano 4, 3117 (2010)
72. Jarzebski, Z.M., Morton, J.P.: Physical properties of SnO2 materials. J. Electrochem. Soc. 123,
299C (1976)
73. Fonstad, C.G., Rediker, R.H.: Electrical properties of high-quality stannic oxide crystals.
J. Appl. Phys. 42, 2911 (1971)
74. Samson, S., Fonstad, C.G.: Defect structure and electronic donor levels in stannic oxide
crystals. J. Appl. Phys. 44, 4618 (1973)
75. Nagasawa, M., Shionoya, S.: Properties of oxidized SnO2 single crystals. Jpn. J. Appl. Phys.
10, 727 (1971)
76. Ohlsen, W.D., Johnson, O.W.: “Vacuum reduction” of rutile. J. App. Phys. 44, 1927 (1973)
Chapter 3
The Role of Defects in Functional Oxide
Nanostructures

C. Sudakar, Shubra Singh, M.S. Ramachandra Rao, and G. Lawes

3.1 Introduction

The burgeoning interest in nanoscale metal oxides arises from the recognition that
the material properties of these systems depend strongly on morphology, allowing
the development of new or enhanced characteristics in geometrically restricted
samples [1–3]. Finite size effects can produce significant changes in a number
of intrinsic properties in systems having reduced length scales, including the elec-
tronic band gap [4], the magnetic coercivity [5], and elastic modulus [6], to name
only a few of the characteristics that are strongly sensitive to sample geometry.
Simultaneously, the large surface to volume ratio in nanomaterials, realized
most dramatically in nanoparticles, can also substantially affect the electronic [7],
magnetic [8], optical [9], and elastic properties [10] of these systems. Because of
their relatively larger surface to volume ratio, the defect concentration in metal oxide
nanostructures is generally higher than that found in bulk systems. These defects can
have a profound effect on the physical properties of nanomaterials, so it is crucially
important that they be fully considered when characterizing metal oxide
nanostructures. There are a number of thorough and accessible reviews on defects
in particular metal oxide systems, including ZnO [11–14], TiO2 [15], and CuO [16]

C. Sudakar
Department of Physics and Astronomy, Wayne State University, Detroit, MI 48201, USA
Department of Physics, Indian Institute of Technology, Chennai, Tamil Nadu 600036, India
S. Singh • M.S.R. Rao
Department of Physics, Indian Institute of Technology, Chennai, Tamil Nadu 600036, India
Nano Functional Materials Technology Centre, Indian Institute of Technology,
Chennai, Tamil Nadu 600036, India
G. Lawes (*)
Department of Physics and Astronomy, Wayne State University, Detroit, MI 48201, USA
e-mail: av4599@wayne.edu

J. Wu et al. (eds.), Functional Metal Oxide Nanostructures, Springer Series 37


in Materials Science 149, DOI 10.1007/978-1-4419-9931-3_3,
# Springer Science+Business Media, LLC 2012
38 C. Sudakar et al.

to name a few, along with more comprehensive reports [17, 18]. Rather than
attempting to provide a general overview of how defects modify the physical
properties of oxides, this particular report is more narrowly focused on briefly
presenting the entirely new properties and characteristics that can emerge in metal
oxide nanosystems due to the presence of defects.
The chapter is structured as follows. We begin with a very short review of the
basic types of defects in metal oxides. Rather than considering the multitude of
possible defect structures, we sharply limit our discussion to point defects. We will
specifically focus on oxygen defect vacancies (VO), metal ion vacancies (VM),
and metal interstitials (MI) since these are generally the most important and widely
studied intrinsic point defects in metal oxides [17, 18]. The bulk of the review will
center on a discussion of the novel electrical, optical, and magnetic properties that
can arise in defect-rich metal oxide systems. We will focus uniquely on the new
physical behavior that emerges due to the presence of defects and do not consider in
any depth the rather more widely studied problem of understanding the role of
defects in perturbing the existing properties of oxides. We conclude with a short
discussion of how these defect-induced properties can be used to integrate new
functionalities into metal oxide nanostructures.

3.2 Defects in Metal Oxide Nanostructures

We broadly limit the scope of our discussion to point defects. As the emphasis of this
review is to consider defects in metal oxide nanostructures, we further restrict
ourselves to discussing only intrinsic defects, and only very briefly touch on dopant
ions as point defects. Within these constraints, a large number point defects can
be considered: oxygen vacancies (VO), metal vacancies (VM), oxygen interstitials
(OI), metal interstitials (MI), and antisite defects (MO or OM). In a large number of
cases, the most stable and/or physically important defects are VO, VM, and MI [17,
18], so we focus primarily on these specific examples. The interactions among
different types of defects can play an important role in determining the physical
properties of metal oxides. For example, because Sn is multivalent, SnI point defects
can readily form in SnO2, which, in turn, supports the formation of VO defects
leading to n-type conductivity in defect-rich SnO2 films [19]. Rather remarkably,
the presence of these defects alone in metal oxide nanomaterials is sufficient to
produce new physical properties that are not simply perturbations of the intrinsic
characteristics of the defect-free parent oxide. Heuristically, these emergent
properties can, in general, be understood to arise from interactions among the
point defects leading to collective behavior. Certain metal oxide nanostructures
are typically able to support a relatively high concentration of VO defects, with
oxygen nonstoichiometry reaching several percent near the surfaces of nanoscale
systems [20]. Because of this relatively large defect concentration, defect–defect
correlations can affect the response of the system [20–22].
3 The Role of Defects in Functional Oxide Nanostructures 39

Fig. 3.1 Schematic


illustration of possible
structural defect in metal
oxides (adapted from [23]) Dislocation

Anion vacancy

Substitutional Anion interstitial


dopant
Dopant-vacancy
pair defect
Cation interstitial
Interstitial dopant
Cation vacancy

Ordered anion
vacancy phase

3.2.1 Defect Structures in Metal Oxide Nanostructures

The concentration and distribution of defects determine a number of properties of


crystalline solids. In high quality crystals the concentration of defects can be
extremely small, leading to considerable experimental challenges in accurately
determining this defect concentration [23]. Crystalline solids contain a number of
different types of structural defects. Vacany defects develop due to the absence of
atoms in some lattice sites while interstitials arise from extra atoms occupying the
space between the atoms in the lattice [18, 23]. Vacancies and interstitial atoms
are point defects as these imperfections are limited to one unit cell and lead to
deviations from the crystalline order only in the immediate vicinity of the defect.
In addition to point defects, line and plane defects are very often found in real
crystal systems [18, 23]. Line defects are dislocations that are characterized by
displacements in the crystal structure along specific directions [18, 23]. Examples
of plane defects comprise stacking faults, grain boundaries, and internal and
external surfaces. A schematic diagram of possible defects in crystalline metal
oxide nanostructures is shown in Fig. 3.1. These defects can profoundly modify
the physical properties of materials, including electrical, optical, and magnetic
response as we will discuss in the following.
Many metal oxides, including ZnO, TiO2, SnO2, and In2O3, exhibit marked
deviations from stoichiometry under specific annealing conditions including thermal
annealing in vacuum [20, 24, 25] and under a finite metal vapor pressure [26, 27].
A relatively small degree of off-stoichiometry can be supported by the inclusion of
point defects, including VO and VM [23], and annealing can also promote the
formation of interstitials and antisite defects. At small concentrations (0.1–1 at.%)
these point defects are typically assumed to be randomly distributed throughout
40 C. Sudakar et al.

Fig. 3.2 Defects in In2O3 (a–c) and TiO2 nanoparticles (d–e) and micron-sized particles (f–g).
High resolution transmission electron micrographs of typical surface regions for (a) as-deposited
and (b) vacuum-annealed In2O3 samples. (c) A magnified view of a section of figure
(b) with arrows showing typical of several distortions in the crystal lattice. The square region
of HRTEM in (c) corresponds to a unit cell of In2O3 shown in the ball and stick model projected
along (1 0 0) plane. The insets (i and ii) in (c) are the simulated HRTEM images with (i) a oxygen
vacancy and (ii) a oxygen vacancy with two adjacent In atoms clustering models.
Bright field TEM images show surface regions of TiO2 nanoparticles for air-annealed (d)
and vacuum-annealed (e) samples. The TEM (f) and HRTEM (g) images of vacuum-annealed
sputter-deposited films show the interior of the crystallites with large number of parallel twin
running along the (0 1 1)

the lattice [23]. At higher defect concentrations, a number of different types of defects
can develop, including multiple charge state defects and pairs or complexes of defects
[17, 18, 23]. However, as we will discuss, many of the properties of defect-rich metal
oxides can be understood by considering segregated (though possibly interacting)
point defects, so we center our discussion to this class of structures.

3.2.2 Imaging Defects in Metal Oxide Nanostructures

At relatively small concentrations, the distribution of point defects in oxides is


determined solely by entropy considerations and consists of randomly distributed
defects [23]. At higher defect concentrations, enthalpy begins to affect the
distribution, which leads to the formation of new structures including defect
clusters, superlattice ordering and extended defects, shear planes, and discrete
intermediate phases [23]. Defect-rich TiO2 (see Fig. 3.2) is an example of an
oxide in which extended defects, including planar defects, are formed by the
accumulation and elimination of point defects, such as oxygen vacancies in
vacuum-annealed samples, along specific crystallographic planes [28].
These vacancies are eliminated by the formation of shear planes in the crystal,
which in turn produces a fault in the cation sublattice [28]. Because of this
3 The Role of Defects in Functional Oxide Nanostructures 41

interplay between point defects and extended defect structures, in some metal
oxide systems it can be difficult to completely separate the two, as we illustrate in
the case of TiO2 and In2O3 in Fig. 3.2.
We first consider ideal, isolated point defects in ZnO as a representative metal
oxide system. In wurtzite ZnO the possible point defects are oxygen and zinc
vacancies (VO, VZn), interstitials (Oi, Zni), and antisite defects (OZn and ZnO).
VO and Zni were most generally considered to be the defects responsible for
modifying the electric and magnetic properties of the system [29–34]. However
recent work [11] suggests that these defects exhibit high formation energies under
equilibrium conditions. Zn interstitials (Zni) are shallow donors and fast diffusers
with a low migration barrier, 0.57 eV, and are therefore not stable at room
temperature [11, 35]. VZn, which has a low formation energy, is a deep acceptor,
so is able to act as a compensating center in n-type ZnO, and may be relevant for
the green luminescence observed in ZnO [11, 35]. Oi has a high energy and acts as
a deep acceptor at the octahedral site Oi1(oct) in n-type ZnO [11, 36].
The antisite defects (ZnO and OZn) have very high formation energies and are
unlikely under equilibrium conditions [11, 35, 36].
The defect structure in metal oxide nanostructures can be imaged using high
resolution transmission electron microscopy (HRTEM). Comparing real-space
images of air-annealed and vacuum-annealed In2O3 thin films clearly demonstrates
the effects of point defects on the nanostructure [20]. Figure 3.2a shows an HRTEM
image of an air-annealed In2O3 nanoparticle, showing a well-ordered lattice with no
obvious defects. Vacuum annealing this sample introduces oxygen vacancies, as well
as other point defects, as shown in Figs. 3.2b, c. The agglomeration of point defects
leads to a 2–3-nm thick surface-disordered layer, as shown in Fig. 3.2b. Additional
point defects, both VO and InI, can be seen in the bulk of the sample. These additional
point defects are highlighted by arrows in Fig. 3.2c. The insets of Fig. 3.2c show
simulated HRTEM images for an oxygen vacancy defect (i) and an oxygen vacancy
with and adjacent cluster of two In (ii).
Similar defect-induced structures can be observed in nanoscale TiO2
(Fig. 3.2d–g) [25]. Air-annealed TiO2 thin films consisting particles ranging
from 300 to 500 nm show good crystalline order with few defects, as illustrated
in Fig. 3.2d. Conversely, vacuum-annealed films show numerous crystallographic
twin boundaries, with individual grains often containing several parallel twins
(Fig. 3.2f). Additionally, these particles exhibit a highly disordered surface phase
of few nanometers thick [25]. These are common [28] microstructural features in
nonstoichiometric TiO2 and are intimately related to the formation of shear
structures discussed above. The twinning produced in the rutile subcell structure
is parallel to (0 1 1), which is the common twinning plane for TiO2 [28, 37, 38].
This twinning is not observed in TiO2 thin films formed from nanoparticles [25].
However, a substantial nonstoichiometric disordered phase develops at the
surface of the nanoparticles (Fig. 3.2e), which suggests that these planar defects
may be more readily diffuse to the surface under thermal annealing in films
comprising smaller particles.
42 C. Sudakar et al.

3.2.3 Stability of Intrinsic Point Defects


in Metal Oxide Nanostructures

In order for these point defects to have any meaningful effect on the properties of the
metal oxide nanostructures in the context of device applications, they should be stable
under ambient conditions. The determination of whether these defects are stable
depends strongly on the details of the specific metal oxide being considered. In ZnO
for example, the VO defects are believed to become stable in the presence of transition
metal dopants such as Co [24]. In this particular study, oxygen defects were introduced
in thin film ZnO samples by annealing at high temperatures (600 C) and low pressures
(~106 Torr). Raman spectral modes related to –Zn–O–Co– local disordered
vibrations in the stoichiometric (or defect-poor) Co:ZnO films disappear after vacuum
annealing as the oxygen vacant sites are localized near Co site (–Zn–VO–Co–). In a
number of systems, however, oxygen vacancy defects may not be stable. Studies on
oxygen-deficient TiO2 found that the concentration of VO defects decreases rapidly
under ambient conditions [25], although these defects can apparently be stabilized by
transition metal doping [32, 39, 40]. Conversely, in In2O3 nanostructured films,
the oxygen vacancy defects are stable for a timescale of years under the same
conditions [41]. Because of this sensitive dependence of the persistence of point
defects on the specific compound being considered, it is important to properly
characterize the stability of these defects in a particular metal oxide when determining
the effects such defects may have on the physical properties of the material.

3.3 Electrical Response

Metal oxides exhibit a range of electrical transport properties, from metallic to


insulating to superconducting [37, 42, 43]. The introduction of point defects
generically affects all types of electrical transport, through mechanisms ranging
from increased scattering in metallic systems to the introduction of additional
charge carriers in insulators. We are particularly interested in exploring systems
in which the inclusion of point defects qualitatively changes the electrical
transport properties. We therefore limit our discussion to considering the onset
of metallic or quasimetallic conductivity induced by point defects in systems
where the undoped metal oxide is insulating or semiconducting.

3.3.1 Point Defects and Charge Carriers

In general terms, point defects in metal oxide nanostructures act like charge
centers [44], which can lead to very high electrical conductivities. Experimentally,
a number of metal oxide systems that are insulating when prepared as perfectly
3 The Role of Defects in Functional Oxide Nanostructures 43

stoichiometric samples develop good electrical conductivity with the introduction


of point defects [14, 45]. In the simplest models, this increase in conductivity requires
shallow donors near the conduction band [46], or acceptors near the valence band
[47]. Many, though by no means all, defect-rich oxide materials are found to exhibit
n-type conductivity, pointing to an abundance of excess electrons associated with the
defects. VO sites, which normally act as electron donors, are a possible point defect
in all metal oxide systems and it is often believed that the conducting properties in
these materials arise from oxygen vacancy defects [48]. While we see that oxygen
vacancy defects do play a crucial role in mediating electrical conductivity in many
metal oxide materials, other types of point defects can also have a significant effect on
transport in defect-rich samples.
ZnO represents one of the most intensely investigated metal oxide system
[49–51] in the past decade. While oxygen vacancy defects, possibly together
with Zn interstitials, had been widely considered to be the origin of the n-type
conductivity in this system [12, 48], recent studies suggest that other point defects
may be more relevant for determining the electrical transport properties [11].
These investigations find that VO sites are deep donors, falling approximately
1 eV below the conduction band, and are thus unlikely to introduce any significant
characteristic of n-type charge carriers. The concentration of Zn interstitials is
found to have a high formation energy in n-type materials, making ZnI defects
unlikely as the source for increased conductivity in defect-rich ZnO [11].
Upon considering all native point defects in ZnO, the authors conclude that
none of these is likely to produce the observed n-type conductivity and instead
propose that the charge carriers arise from the accidental inclusion of substitu-
tional hydrogen, HO, which can act as a shallow donor [52]. Along a similar line,
experimental studies on the conductivity of ZnO films prepared by pulsed laser
deposition provide evidence that nitrogen inclusions may play an important role
in the development of n-type conductivity in ZnO [46], with other work pointing
to the importance of hydrogen donors [53].
Indium oxide is another widely studied electronic material, but there still remain
a number of unanswered questions concerning the fundamental transport properties
in this system [54]. In2O3 can exhibit a high degree of nonstoichiometry and shows
extremely good n-type dopability [55–58]. It has been suggested that In2O3 is
an anion-deficient n-type conductor, but that the small oxygen vacancy defect
population, corresponding to approximately 1% of the anions, limits the electron
concentration [59]. Experimentally, it is found that oxygen-deficient In2O3 is highly
compensated, with the ratio of n-type free charge carriers to oxygen vacancy defect
sites being approximately 1:5, rather than the 2:1 one would expect that each
oxygen vacancy contributes two electrons [20, 59]. Recent density functional
theory calculations on defect-rich In2O3 find that oxygen vacancies, rather than
In interstitials, are the likely sources of n-type conductivity [60]. Furthermore,
both indium vacancies and oxygen interstitials are identified as possible charge
compensation sites.
The dramatic effects of point defects on the electrical transport properties of
transition metal oxides are clearly demonstrated by the remarkable change in
44 C. Sudakar et al.

Fig. 3.3 Temperature-


dependent resistivity for an
air-annealed defect-poor
In2O3 thin film (open
symbols), and the same film
after vacuum annealing
(defect rich, closed symbols)

conductivity of In2O3 thin films upon vacuum annealing, illustrated in Fig. 3.3.
As-prepared In2O3 thin films, which were crystallized by annealing in air and are
presumed to be close to stoichiometry, are highly resistive and show insulating
behavior below room temperature. On vacuum annealing, which introduces oxygen
vacancies and may also produce other types of point defects, the films develop n- type
conductivity with a carrier concentration on the order of n ¼ 1020 cm3. Concomitant
with this increase in carrier concentration, the room-temperature resistivity of the
films drops by three to four orders of magnitude and the samples exhibit metallic
conductivity to low temperatures, with a small upturn in resistivity below ~80 K.
Rather remarkably, defect-rich In2O3 films remain optically transparent, despite
the high conductivity. The optical band gap is found to increase from approximately
3.3–3.6 eV on vacuum annealing, which can be attributed to the Burstein Moss shift,
but there is practically no change in the optical transmission, which remains above
80% for visible wavelengths [41]. Similar conducting and optically transparent
features are observed in SnO2 samples, where it is argued that the high oxygen
vacancy defect concentration required for producing conductivity is stabilized by the
presence of multivalent Sn interstitials [19]. These same density functional studies
suggest that the donor electrons are not heavily compensated due to the paucity of
acceptor defects (VSn and OI). Furthermore, it is found that these donors do not have
direct optical transitions in visible wavelengths, so do not directly affect the optical
transparency. Since In has a fixed formal valence of +3, the same mechanism is not
likely to apply for In2O3, but the result on SnO2 highlights the importance of
interstitials in stabilizing oxygen vacancy defects.

3.3.2 Defects and p-Type Conductivity

A number of defect-rich metal oxide systems exhibit p-type rather than n-type
conductivity [37, 42, 47]. First-principles calculations on Cu2O find that the lowest
energy defects are Cu vacancy point defects, VCu, and a point defect complex
consisting of a Cu interstitial, CuI, located between two VCu defects [61]. The VCu
defects are found to produce delocalized holes near the top of the valence band,
3 The Role of Defects in Functional Oxide Nanostructures 45

leading to p-type conductivity. Measurements on intentionally undoped Cu2O thin


films find a p-type carrier concentration on the order of 1015 cm3, resulting in a
resistivity of approximately 150 O cm [47]. p-Type conductivity can also develop
in the Mott insulator NiO. Careful measurements have established that VNi sites are
the dominant point defects for determining the electrical properties of NiO samples,
rather than OI sites [62]. It is estimated that the VNi defect concentration can reach
1016 cm3 in nanostructured samples, leading to six to eight orders of magnitude
increase in conductivity over undoped NiO single crystal samples [63].

3.3.3 Defects and Conduction Mechanisms

In addition to understanding the origin of charge carriers in defect-rich metal oxide


nanostructures, it is also important to consider the mechanisms for electrical conduc-
tion. Depending on the details of the electronic structure, a number of different effects
can be relevant for electronic transport. Careful investigations on the low temperature
resistivity and Hall resistance of oxygen-deficient TiO2, having oxygen vacancy
defect concentrations in the range from 4  1018 cm3 to 5  1019 cm3, find that
the low temperature transport is consistent with hopping conductivity for high and
low VO concentrations (n[VO]) [64]. However, the transport falls in the inter-
mediate range between hopping and metallic conduction for 8  1018 cm3 < n
[VO] < 2  1019 cm3, where the donor separation is estimated to be five times the
effective Bohr radius of the donor electron [64]. TiO2 does not develop metallic
behavior because increasing the defect concentration leads to the development of
planar defects [64]. Hopping conduction has also been established as the origin of
electrical transport in defect rich NiO films [63, 65]. Frequency-dependent resistivity
studies find that the transport can be well modeled by the correlated barrier hopping
model, which supposes that holes hop from Ni3+ sites to Ni2+ sites with a barrier
height that depends on the separation between defects. A fit to the data yields a
separation of 1.9 eV between the ground state of the defect and the valence band [65].
Band conduction can also be observed in defect-rich metal oxides. ZnO films
prepared in an oxygen-deficient environment were found to be highly conducting,
with a band-like mechanism for conduction having an activation energy of ~1 meV
[46]. More stoichiometric samples were found to exhibit Arrhenius conductivity,
associated with the thermionic emission of band electrons from grain boundaries at
higher temperatures and thermally assisted hopping at lower temperatures [46].
There have also been a number of theoretical studies on the impurity band structure
in defect-rich ZnO, as there are proposals that the ferromagnetism in this system
(discussed in more detail in Sect. 3.5) may arise from spin split impurity bands [66].
Ab initio density functional calculations suggest that VZn point defects should
produce metallic behavior in defect-rich ZnO, while OI defects give rise to a
semiconducting electronic structure [67].
Surface effects are also expected to affect the electrical transport properties
in metal oxides, which is particularly relevant for nanostructured materials having a
46 C. Sudakar et al.

Fig. 3.4 Optical absorption


spectra for defect-rich (upper
curve) and defect-poor (lower
curve) In2O3 thin films

high surface area to volume ratio. Theoretical studies on SnO2 find no evidence for
defect-induced states in the gap [68], while subsurface oxygen defect vacancies in
TiO2 can lead to states falling 0.7 eV below the conduction band edge [69]. A surface
conduction layer, presumably arising from defect states, is found in ZnO; the conduc-
tivity of this layer is reduced on exposure to oxygen [12]. This surface conduction
layer provides an additional channel for electronic transport. In2O3 films and
nanostructures can develop a chemical depletion layer near the surface, corresponding
to a higher oxygen content at the interface [70]. This system also has a high density of
electronic surface states, which produces relatively large band bending.

3.3.4 Plasmon Response in Defect-Rich Oxide Nanostructures

One of the more striking realizations of the collective response of an electron gas is
the phenomenon of plasma oscillations. These plasmons are excited at the plasma
frequency op, given by o2p ¼ 4pne2 =meff ; with n the carrier density and meff
the effective mass. This plasma frequency is typically large for most metals,
with ћop ~ 11 eV for bulk plasmons in Al [71], and depends strongly on the charge
carrier concentration n. Because plasmons reflect collective behavior of the charge
carriers, they represent emergent response in insulating metal oxides driven by
point defects, which is completely absent in the parent system. To illustrate the
clear development of this electronic collective behavior in defect-rich metal oxide
nanostructures, we consider the optical response of as-prepared (defect-poor)
and vacuum-annealed (VO defect-rich) In2O3 thin films, as plotted in Fig. 3.4.
The as-prepared sample is insulating and, as expected for transparent materials,
has negligible absorption at energies well below the band gap. Conversely,
the vacuum-annealed In2O3 sample has a high concentration of VO defects leading
to an n-type charge carrier concentration of roughly 1020 cm3, as measured by the
Hall effect [41]. This high concentration of charge carriers in the defect-rich sample
leads to qualitatively different behavior in the low energy optical properties.
3 The Role of Defects in Functional Oxide Nanostructures 47

We observe a clear plasmon peak in the absorption falling at 0.53 eV. While the
magnitude of the absorbance associated with this peak falls well below the band
gap absorbance, this plasmon resonance represents the emergence of a distinct
electronic response that is absent in the defect-poor parent metal oxide structure.

3.4 Optical Response

As discussed in the previous section, charge carriers in oxides arise from a number
of different sources including interstitial metal ion impurities, substitutional doping
ions, and oxygen vacancies. Oxygen vacancies present in the lattice can act as
divalent electron donors, with these defects normally acting as shallow donors.
Within the metal oxide systems, point defect ionization occurs in a similar manner
to that found in doped semiconductors. Although the scattering of charge carriers in
defect-rich oxide systems arises primarily from ionized impurity scattering, the
majority of the intrinsic optical phenomena in these materials arise from ionized
defects occupying energy states lying in the band gap, at least for wide band gap
oxides. When such systems are obtained in the low-dimensional form the optical
properties of these materials are often modified due to the increase in surface energy
and surface defect states, which can lead to a number of interesting possibilities for
applications.
The incorporation of metal oxide nanostructures into electrooptical devices
relies on their ability to efficiently emit or absorb light; these application prospects
are heavily influenced by the energy band structure and lattice dynamics of the
system [72–74]. This change in optical response in defect-rich oxides is reflected in
the altered band to band transitions and absorption energies [75]. Moreover, oxide
nanostructures have lower threshold lasing energies due to quantum effects that
increase the density of states near band edges [76, 77]. For a number of varied
electrooptical applications it is necessary for all charge carriers, both electrons and
holes, to be confined [78]. One-dimensional wide band gap nanostructures are the
best candidates for this class of applications due to their remarkable physical and
chemical properties. However, it is also crucial to understand how these optical
properties may be affected by the almost unavoidable incorporation of point defects
in these nanostructured materials.

3.4.1 Photoluminescence from Point Defects


in Oxide Nanostructures

Nanocrystalline zinc oxide (nano-ZnO) is a wide band gap semiconductor that is


particularly promising for a number of optoelectronic properties, including ultravi-
olet (UV) light emitting diodes, UV laser diodes, and UV photodetectors, because
48 C. Sudakar et al.

of its very high excitonic binding energy (60 meV) [79] compared to GaN (25 meV)
and relative ease of band gap engineering [80]. Nanostructured ZnO possesses a
remarkable photoluminescence (PL) spectrum. The optical response of ZnO
changes significantly on the introduction of point defects, rather by doping or by
the incorporation of intrinsic defects. The PL spectrum of nano-ZnO consists
mainly of two emission peaks, one in the UV, falling near 385 nm, which is ascribed
to near-band-edge emission [81], with the other peak located in the visible region,
occurring in the green around 500 nm [82–85]. The origin of green luminescence
band is still not well understood; this has been attributed to the presence of a variety
of different impurities and defects present in the ZnO lattice. ZnO exhibits lumi-
nescence defect centers such as oxygen vacancies (located at 50 and 190 meV
below the conduction band edge), zinc interstitials (located at 2.5 eV below the
conduction band edge) [86], and various other native defects [82–85]. However it is
interesting to note that the intensity of the green emission can be controlled in a
systematic manner by oxidation and reduction [87]. Nanostructures such as
nanoislands of ZnO show PL emission whose origin can be explained on the
basis of zinc vacancy (VZn) complex defects [88]. The intensity of PL emission
from samples containing such islands is much smaller than that from normal thin
films due to a smaller area being covered by the islands. It has also been found that
upon biomolecule attachment, nano-ZnO powders exhibit further induced changes
in peak intensities and/or peak shifts [89]. Photoluminescence of nano-ZnO parti-
cle/SiO2 aerogel assemblies has shown very strong PL band at 500 nm whose
luminescence intensities are 10–50 times higher than those of nanostructured bulk
ZnO. The quantum efficiency is found to lie between 0.2 and 1%. This enhancement
is attributed to the increase of the singly ionized oxygen vacancies in nano-ZnO
particles, which are located in nanopores of the SiO2 aerogel [90].
Transition metal (TM) ion dopants, such as Ni and V substituting for Zn, suppress
the UV emission peak, indicating that the TM doping increases nonradiative recom-
bination processes in this material [91]. These nonradiative transitions arise when
free electrons recombine through a process involving a TM ion impurity level instead
of populating donor acceptor pairs [92, 93]. The suppression of the UV PL peak can
also be partially attributed to energy transfer processes from intrinsic donor–acceptor
pairs to neighboring TM ions [92, 93].
The conduction band in wurtzite ZnO is constructed mainly from s-type states,
while the valence band is formed from p-type states, which is split into three bands
due to the influence of crystal-field and spin–orbit interactions [94]. The related
free-exciton transitions (FX) from the conduction band to these three valence bands
or vice versa are usually denoted by A (also referred to as the heavy hole),
B (also referred to as the light hole), and C (also referred to as crystal-field split
band). Our previous studies have suggested that the PL spectrum of Ni-doped ZnO
nanoneedles at 10 K is dominated by neutral donor-bound exciton emissions [95].
We also observe a free A-exciton transition in Ni-doped ZnO nanoneedles grown
in an Ar atmosphere at FXA ¼ 3.375 eV (Fig. 3.5) at 10 K. The neutral shallow
3 The Role of Defects in Functional Oxide Nanostructures 49

Fig. 3.5 Logarithmic plot of


low temperature PL spectrum
of Ni: ZnO grown in Ar
atmosphere at 10 K

donor-bound exciton dominates because of the presence of donors due to uninten-


tional (or doped) impurities and/or shallow donor-like defects. The free A-exciton
bound to a neutral donor is positioned at 3.36 eV (D0XA). The energy separation
between the FXA and D0XA peak gives us the binding energy of the related donor-like
defect which is of the order of 15 meV.
We have observed that pure ZnO nanorods calcined at 500 C exhibit higher
defect emission combined with lower excitonic emission as compared to ZnO
nanorods treated at 600 C. This is attributed to a sharp increase in the volumetric
surface defect concentration with increase in surface area. As the calcination
temperature is reduced from 600 C to 500 C, the surface area to volume ratio for
the ZnO nanorods increases by approximately three orders of magnitude due to the
small size of these ZnO rods. The dramatic changes in the emission spectra
associated with this increase in defect concentration are illustrated in Fig. 3.6(a),
with structural changes shown in Fig. 3.6(b); a more complete discussion included
in [96]. The higher surface defect concentration in the samples calcined at low
temperatures (samples with lower dimensions) results in a sharp drop in the band-
edge intensity near 380 nm and the growth of a very broad peak centered near
475 nm. In this spectrum, the green emission in the range of 450–500 nm is believed
to originate from a transition between the electron in the conduction band and a
deep level. This hypothesis is consistent with the luminescence mechanism pro-
posed by Dijken et al. involving an electron in a conduction band and a deeply
trapped hole [96].
Besides ZnO a number of other wide band gap semiconductors, including IIIB
and IVB group oxides like In2O3 and SnO2 nanostructures ,are also actively
considered as candidates for fabricating electronic and optoelectronic nanodevices
[33, 97]. It is known that bulk In2O3 (Eg ¼ 3.6 eV) does not emit light at room
temperature [98–100]. However, In2O3 nanoparticles show PL signals at 430, 480,
520, and 637 nm. The origin of most of these peaks from In2O3 films have been
attributed to oxygen vacancies [88, 101, 102].
50 C. Sudakar et al.

Fig. 3.6 (a) PL spectra of ZnO nanorods calcined at different temperatures. (b) SEM images of
ZnO nanorods synthesized at different temperatures (a more complete discussion is included in
[96])

3.4.2 Raman Studies on Oxide Nanostructures

Raman spectroscopy provides a powerful tool to probe the structural characteristics


of oxide nanostructures. The local symmetry in oxide nanoparticles, specifically
ZnO, can be different from that of bulk samples, although the macroscopic crystal
structure is identical for both samples [103]. This is illustrated from the Raman
spectra comparing bulk and nanostructured samples (Fig. 3.7). These bulk and
nanostructured ZnO samples all have identical crystal structure [104], but markedly
different Raman characteristics. Undoped bulk ZnO shows clear Raman peaks
at 663 cm1 [A1(LO) + E2(low)], 538 cm1 (2LA mode), 437 cm1 (attributed to a
high frequency nonpolar optical phonon E2 mode of ZnO), 407 cm1 [E1(TO) mode],
and 381 cm1 [A1(TO) mode] [105, 106]. The E2 (high) mode at 437 cm1 is the
strongest mode in the wurtzite crystal structure and any broadening or weakening of
this peak indicates the presence of defects in the host lattice. This particular mode,
along with the less intense mode at 579 cm1 and the A1(TO) mode at 381 cm1, is
strongly suppressed in the nanostructured ZnO brushes and droplets as compared to
bulk samples, which is attributed to defect-induced changes in the local symmetry of
3 The Role of Defects in Functional Oxide Nanostructures 51

Fig. 3.7 Comparison of


E2 phonon shift in Raman
spectra of the as-deposited
nanostructured and the bulk
ZnO samples. The peaks at
579 and 381 cm1 occurring
in bulk ZnO powder are
suppressed for the brushes as
well as droplets. Inset shows
shift in the peak occurring at
437 cm1

these samples due to surface defects. In some cases Raman spectra show the presence
of ZnO optical phonon mode, which is red shifted when compared to bulk ZnO.
These are attributed to optical phonon confinement effects [107] or the presence of
intrinsic defects on the nanoparticles [108]. However in the as-grown ZnO nanorods
with much bigger size than Bohr exciton radii (~2.34 nm), phonon confinement effect
cannot be expected to be the main reason of the shift [109].
Other shifts in the Raman response for ZnO can also be observed in bulk samples
with the introduction of substitutional point defects. In the V-, Ni-, Ti-, and Fe-
doped ZnO bulk samples, the Raman peak frequencies are uniformly red shifted to
lower frequencies [91]. Such shifts in Raman frequency are believed to depend on
residual stress, disorder, and crystal defects present in the samples [110, 111]. The
defect-induced disorder disrupts long-range ordering in the ZnO lattice, which
weakens the electric field associated with a mode [111]. Furthermore, the inclusion
of point defects in the ZnO lattice can lead to the presence of additional Raman
modes, which are referred to as anomalous modes. Two possible mechanisms have
been proposed to account for these anomalous modes: disorder-activated scattering
and local lattice vibration [106]. Low-frequency Raman modes have been identified
for Fe-doped (19 and 39 cm1) and Mn-doped (22 and 46 cm1) ZnO nanoparticles
having mean crystallite sizes of ~10 and 43 nm, respectively [112]. The position of
these modes has been connected to the dimension of particles and dopant concen-
tration [112], highlighting the importance of considering the density defects when
interpreting or tuning the optical properties of metal oxide nanostructures.
Metal oxide-based nanostructures also offer opportunities for promoting new
approaches in Raman spectroscopy. Surface enhanced Raman scattering (SERS) is
exhibited when a nanoscale dielectric core is surrounded by a metal shell (often called
a nanoshell) [113]. This effect provides a huge increase in the intensity of the Raman
scattering signal, leading to a considerable enhancement of Raman spectroscopy as a
tool for designing biological or chemical sensors [114]. This enhancement in the
Raman signal is attributed to a local electromagnetic field enhancement at the metal
52 C. Sudakar et al.

surface or rough metal structures due to the surface plasmon polaritons [114, 115].
This effect may also be promoted by a chemical enhancement arising from an
electronic resonance transfer between surface absorbed molecules and the metal
surface [116–118]. Among metal oxide nanostructures SERS has been observed for
Au-coated ZnO nanorods having a biomodal size distribution with diameters of 150
and 400 nm prepared on a Si (1 0 0) substrate [119]. These structures show large
Raman enhancement factor (EF) values of the order of 106. This enhancement factor
is defined as

ISERS =Nads
EF ¼ ;
Ibulk =Nbulk

where ISERS is the intensity of the vibrational mode in the SERS spectrum, Ibulk is
the intensity of the same mode in the Raman spectrum, and Nads and Nbulk represent
the number of the corresponding analytic molecules effectively excited by the laser
beam [120]. Highly surface enhanced Raman spectra have also been obtained using
indium tin oxide coated gold nanotriangles as well as gold nanoparticle-
immobilized indium tin oxide [121]. Experimental reports on ZnO crystalline
samples covered with Ag nanoparticles suggest that the resonant Raman scattering
process is assisted by metal-induced gap states at the Ag/GaN and Ag/ZnO
interfaces. This study provides a view on electron-mediated enhanced Raman
scattering SERS of lattice vibrations in oxide semiconductors [122]. The presence
of defect sites such as metal pinholes can diminish the enhancement [123].

3.4.3 MagnetoOptical Properties of Oxide Nanostructures

As will be discussed in more detail in Sect. 3.5, defects in metal oxide nanostructures
can also have strong effects on the magnetic properties of these systems. These
induced spin structures can, in turn, affect the optical response of the nanostructures.
As an example, the formation of antiphase boundary defects in metal oxides can give
rise to large internal strains [124]. To determine the coercive field of a given sample,
longitudinal magnetooptic Kerr effect (MOKE) magnetometry is measured using a
light source. The optical and magnetooptical properties of oxides, such as ZnMnO,
in the Faraday configuration give us an estimate of the exchange constant [125].
In this context we emphasize that MOKE effect is very sensitive to the strain,
stoichiometry, and film thickness. A large mismatch between the lattice constants
of the thin film and substrate can lead to a large residual strain due to the formation of
antiphase boundary defects [126, 127]. In turn, these antiphase boundary defects
may subsequently reduce the net magnetization by changing the exchange interac-
tion across an antiphase boundary. The magnetooptical properties of metal oxide
nanostructures are therefore sensitive to the defect structure in the samples. These
structural defect-induced changes in the magnetic properties can be investigated by a
number of different techniques, including MOKE measurements [128].
3 The Role of Defects in Functional Oxide Nanostructures 53

Magnetooptical probes provide a powerful tool for identifying the origins of


different bound exciton transitions [129]. Exciton bound to ionized impurities can
often be identified by the nonlinear splitting of their transitions in an applied magnetic
field [130, 131]. Such splitting has been observed in ZnO with a magnetic field
applied along the c-axis. This approach can be valuable in characterizing defects in
oxide nanostructures [130].

3.5 Magnetic Response

Metal oxides exhibit a very wide range of magnetic properties, ranging from
ferrimagnetism with relatively large saturation magnetizations in Fe3O4 [132],
to antiferromagnetic order in NiO [133] and CoO [134], to simple diamagnetism
in ZnO [135] and TiO2 [136], to more complex spin structures in Mn3O4 [137].
Expanding this rich set of possible magnetic characteristics, it is well known that
the magnetic properties of nanostructure materials are often very different from
what is observed in bulk systems [138]. Given the diverse nature of metal oxides
and specific nanostructures, there is a bewildering array of magnetic properties
that are manifested in metal oxide nanostructures, before even considering
modifications arising from defects. Rather than attempting to completely summa-
rize the effects of point defects on the magnetic properties of all categories of metal
oxide nanostructures, we instead consider only specific examples of systems in
which these defects can induce weak ferromagnetic behavior. We will first briefly
discuss some results concerning the development of superparamagnetism and weak
ferromagnetic moments in antiferromagnetic metal oxide nanoparticles before
visiting the emergence of ferromagnetic order in diamagnetic semiconducting
metal oxides induced by point defects.

3.5.1 Magnetism in Metal Oxide Nanoparticles

Measurements on the weak ferromagnetism in nanoscale antiferromagnetic metal


oxide systems can be challenging, because of the possibility of accidentally
incorporating ferromagnetic secondary phases, such as metallic Co inclusions in
CoO nanoparticles or thin films [139]. Despite these difficulties, there is a growing
realization that antiferromagnetic metal oxide nanostructures often exhibit ferro-
magnetic properties that cannot be necessarily ascribed to impurity phases [138].
The presence of surfaces (or interfaces) in antiferromagnetic materials can typically
lead to the presence of uncompensated spins arising from the incomplete cancel-
ation of the sublattice magnetizations in the antiferromagnetic spin structure [140].
As the surface to volume ratio is exceeding large in nanostructures, the fraction of
such uncompensated spins can be a considerable fraction of the total. In addition to
such “native” uncompensated moments, produced solely by geometrical restrictions,
54 C. Sudakar et al.

the inclusion of point defects can also yield uncompensated spins, which can
also exhibit paramagnetic or weak ferromagnetic behavior [22]. More complicated
magnetic effects, including the onset of multisublattice antiferromagnetic order in
nanoparticles [141] or modifications of the electronic orbitals at the metal oxide
surface [142], have also been proposed, although we omit any discussion of these
properties in the following.
The magnetic properties of CoO nanoparticles are widely studied [143–146],
in part because Co/CoO core/shell nanoparticles represent a model system for the
investigation of exchange bias coupling [145]. It has been found that the uncompen-
sated moments present in CoO can be roughly divided into two classes: those
moments that are strongly coupled to the antiferromagnetic lattice and those that
are not [146]. At least a portion of spins falling in the latter category have been
attributed to point defects, and these are believed to produce a paramagnetic or
superparamagnetic response at low temperatures. Similar effects have been observed
in NiO nanoparticles, which have been shown to exhibit a superparamagnetic
response that increases with decreasing particle size [147]. Careful measurements
on small NiO nanoparticles find that the effective moment arising from these
uncompensated spins exceeds 2,000 mB [148], which is considerably larger than
what would be expected simply from uncompensated surface spins.
More generally, it has been suggested that low temperature superparamagnetic
behavior is a general characteristic of antiferromagnetic metal oxide nanoparticles
[138], including MnO and NiO [149]. Measurements on both NiO and MnO
nanoparticles find evidence for superparamagnetic behavior, with saturation
magnetizations for the ferromagnetic component on the order of a few emu/g
depending on particle size [149]. These samples show hysteretic behavior at low
temperatures, which vanishes at higher temperatures, characteristic of super-
paramagnetism. These investigations also find that the superparamagnetic blocking
temperature increases with increasing particle size for the NiO nanoparticles but,
surprisingly, decreases with increasing size for the MnO nanoparticles [149].
This difference in the size dependence of the magnetic properties at least hints at
the possibility that the origins for superparamagnetism in the two samples may be
different. Because weak ferromagnetism in antiferromagnetic systems can arise
both from discontinuities in the magnetic structure at the surface and from point
defects, it is challenging to unambiguously assign the observed superparamagnetic
moments to one mechanism or the other. Nevertheless, it is clear that structural
defects cansignificantly modify the magnetic response in antiferromagnetic metal
oxide nanostructures.

3.5.2 Ferromagnetism in Defect-Rich Semiconducting


Metal Oxides

A more dramatic example of how point defects can affect the magnetic properties
ofmetal oxides can be found in the observation of ferromagnetism in defect-rich,
intrinsically diamagnetic semiconducting oxides [21, 39, 138, 150–152]. The original
3 The Role of Defects in Functional Oxide Nanostructures 55

studies on this class of materials highlighted the development of ferromagnetism


in metal oxide films doped with magnetic transition metal ions, in particular,
Co substituted into TiO2 [153] and Mn substituted into ZnO [154]. Measurements
on this class of materials found considerable sample-to-sample variation in the
magnetic properties [155, 156], leading to suggestions that the magnetic properties
were produced by precipitates of a secondary ferromagnetic phase [157]. Further-
more, measurements on nearly stoichiometric Co- and Mn-doped ZnO samples
found no evidence for ferromagnetism and identified only weak antiferromagnetic
nearest-neighbor coupling between the dopant ions [158, 159]. Subsequent
experiments on Co-doped ZnO found evidence for the crucial role played by
oxygen vacancy defects in the development of ferromagnetic order in this class of
materials, with air-annealed films (low oxygen vacancy defect concentration)
having negligible magnetizations while vacuum-annealed films (high oxygen
vacancy defect concentration) exhibiting distinct ferromagnetism [24].
Despite the recognition that point defects play an important role in the develop-
ment of ferromagnetic order in transition metal-doped semiconducting oxides, it is
difficult to disentangle the effects of defects from the contributions arising from the
magnetic dopant ions [160]. However, over the past several years it has become
apparent that ferromagnetism can develop in diamagnetic metal oxides, which is
believed to be driven solely by the presence of point defects [161]. Signatures of
ferromagnetic order were observed in undoped HfO2 [151] and subsequently in a
range of other metal oxides, including TiO [39, 40], In2O3 [21], ZnO [162], and
CeO2 [163, 164], among many others [138, 152]. Because most of these systems do
not have thermodynamically stable magnetic compositions, it is unlikely that the
ferromagnetism arises from the precipitation of ferromagnetic impurity phases. It is
found that the magnetic properties of these systems depend strongly on the nature of
the point defects present [165], leading to suggestions of defect-mediated ferro-
magnetism in metal oxide nanostructures [166].
It is known that point defects in diamagnetic metal oxides can introduce
local moments [22]. The details of this local formation depend sensitively on the
compound. For example, oxygen vacancies [167] and zinc interstitials have been
predicted to be nonmagnetic in wurtzite ZnO, although there are reports of Zn
interstitials enhancing the magnetic properties in doped ZnO [168], while
oxygen interstitials and zinc vacancies are expected to show sizeable moments,
ranging from roughly 0.2 mB [167] to 2 mB [67]. However, both oxygen and
cerium vacancies are expected to contribute to the magnetic moment in defect-
rich CeO2 [164]. However, in the complete absence of interactions, defect-induced
moments would be expected to result in paramagnetic, rather than ferromagnetic
behavior, so the emergence of ferromagnetism in these metal oxide materials is
rather surprising.
The strong connection between defects and ferromagnetism in metal oxide
nanostructures is demonstrated by studies on TiO2 thin films [25]. The as-prepared
TiO2 thin films are relatively defect free, as shown in the HRTEM images in
Fig. 3.2d, and have a very small magnetization (Fig. 3.8a). This small moment can
be attributed to residual oxygen-defect vacancies that remain after air annealing [25].
56 C. Sudakar et al.

Fig. 3.8 Room temperature magnetization curves for an air-annealed defect-poor TiO2 film
(a) and for a vacuum-annealed defect-rich TiO2 film (b)

Conversely, vacuum-annealed TiO2 films, presumably having a much higher


concentration of oxygen vacancies, exhibit a much higher concentration of defects,
leading to an amorphous structure at the surface (Fig. 3.2e). Introducing oxygen
vacancy defects leads to a considerable enhancement in the magnetization.
This increase depends on film thickness, pointing to an intimate connection among
microstructure, point defects, and the emergence of ferromagnetism, and reaches
40 emu/cm3 for films having a thickness of 25 nm (Fig. 3.8b). Comparing the
size of the magnetic signal with an estimate of the total volume occupied by the
surface-disordered layer in the TiO2 films leads to the suggestion that the magnetism
in these samples may develop solely in the defect-rich regions of the sample [25].
Most significantly, the magnetization decreases systematically when the films are
exposed to air under ambient conditions [25], again highlighting both the role of
oxygen vacancy defects in developing magnetic order and the importance of properly
characterizing the stability of point defects when considering their effects on metal
oxide nanostructures.
Investigations on CeO2, Al2O3, ZnO, In2O3, and SnO2 [138, 169] nanoparticles,
among others, find evidence for weak ferromagnetism, having small saturation
magnetizations but clear hysteresis loops, albeit often with almost negligible
coercivities. The moments in these nanostructured samples are very small, on the
order of only 104 to 103 emu/g [138], but significantly larger than the completely
negligible magnetizations observed in diamagnetic bulk samples. Sintering the
samples at high temperatures in the presence of oxygen completely suppresses
the magnetization [138], leading to the conclusion that the ferromagnetism may be
intimately connected with the defect structure, specifically including oxygen
vacancy defects, at the surface of the nanoparticles. It is suggested that unpaired
electrons trapped on oxygen vacancies may be relevant for the development of
ferromagnetism and, furthermore, that such ferromagnetic order may be a general
characteristic of all metal oxide nanoparticles [161, 166].
3 The Role of Defects in Functional Oxide Nanostructures 57

Fig. 3.9 (a) Room temperature magnetization curve for an oxygen-deficient In2O3 thin film.
(b) Point contact Andreev reflection measurement on the same oxygen-deficient In2O3 thin film,
measured at T ¼ 2 K using a Nb tip. A fit to this curve yields an estimate spin polarization of
P ¼ 45%

3.5.3 Spin Polarization in Defect-Rich Metal Oxide


Nanostructures

There is considerable debate concerning the observations of ferromagnetism in


undoped metal oxide nanostructures, including the concern that these magnetic
features may arise from the accidental incorporation of ferromagnetic impurities
during sample preparation or handling [160, 170]. This uncertainty arises mainly
because the very small magnetizations observed in these measurements could,
in many cases, be produced by almost negligibly small amounts of contaminants,
which could easily be missed by even the most thorough sample characterization.
It is therefore desirable to probe the development of magnetic order in these
systems using some technique that is not sensitive to trace amounts of ferromag-
netic impurity phases. A number of different approaches to this problem have been
considered, including magnetotransport measurements [171] and magnetic dichro-
ism spectroscopy [172]. In the following, we discuss another approach based on
measurements of the charge carrier spin polarization at normal/superconducting
interface [173].
Thin films of undoped In2O3 exhibit a small but distinct ferromagnetic
signature with the inclusion of oxygen vacancy defects, introduced by vacuum
annealing [21]. Room temperature magnetic hysteresis loops, showing a satura-
tion moment of 0.3  1 emu/cm3 and a coercive field of 50–200 Oe, are shown in
Fig. 3.9a. Because these vacuum-annealed In2O3 films remain conducting to low
temperatures, as discussed in Sect. 3.3, it is possible to probe the spin polarization
of the charge carriers using point contact Andreev reflection (PCAR) [173].
The results of PCAR measurements made at T ¼ 2 K with a Nb tip are shown
in Fig. 3.9b. The zero voltage dip in conductance is characteristic of a finite
58 C. Sudakar et al.

spin polarization of the In2O3 charge carriers. A more careful analysis of the
conductance curve yields an estimated spin polarization of approximately 50%,
indicative of ferromagnetism in these defect-rich metal oxide nanostructures [21].
While these investigations do not unambiguously prove the existence of intrinsic,
carrier-mediated ferromagnetic order, they do firmly establish that the measured
magnetization is at least strongly coupled to the conduction electrons. Evidence
for a finite spin polarization in Co-doped ZnO films has also been inferred from
low temperature tunneling magnetoresistance measurements on Co/Al2O3/Co:
ZnO heterostructures [174].

3.5.4 Mechanisms for Magnetism in Metal Oxide Nanostructures

There are a number of proposals for the origin of weak ferromagnetism in defect-rich
semiconducting metal oxide nanostructures [161, 166]. As the stoichiometric parent
compounds are diamagnetic, the point defects must both provide the magnetic
moments and introduce interactions among these defect moments. While oxygen
vacancy defects are predicted to yield magnetic moments of approximately 2 mB [67],
this relatively small moment is insufficient to produce the high Curie temperatures,
typically well above room temperature [175], observed in many defect-rich metal
oxide nanostructures. It has been suggested that cation defects may offer much larger
magnetic moments, leading to correspondingly larger Curie temperatures [21].
For example, density functional calculations on CeO2 find a moment of 2 mB per
oxygen vacancy, associated with the Ce 4f electrons, with a much larger moment of
4 mB attached to Ce vacancies arising from O 2p orbitals [176]. Similar computations
on SnO2 find that Sn vacancies are magnetic, carrying a moment of approximately
4 mB, while oxygen vacancy defects are nonmagnetic [177].
The presence of defect-induced magnetic moments alone is insufficient to
ferromagnetism in these semiconducting metal oxide nanostructures; room temper-
ature ferromagnetism requires relatively large interactions among these moments.
Local density approximation calculations on SnO2 find evidence for an oscillating
exchange interaction between moments associated with Sn vacancies (VSn), with
strong ferromagnetic coupling arising for an average separation of 0.55 nm [177].
This exchange coupling can be modeled approximately by a Ruderman–Kittel–K-
asuya–Yosida (RKKY)-type interaction, with a kF of 0.12 nm1, suggesting the
importance of charge carriers in mediating the ferromagnetism. The possible role of
defect-induced conduction electrons in promoting ferromagnetic order in defect-
rich metal oxides has also been discussed for In2O3 thin films [178, 179]. It has been
proposed that the n-type carriers from oxygen vacancy defects are highly, though
not completely, compensated by indium vacancy defects. The residual n-type
charge carriers may have a relatively high density of states at the Fermi level, due
to the quasilocalized nature of the donors, promoting high temperature ferromag-
netic order [21].
3 The Role of Defects in Functional Oxide Nanostructures 59

It has recently been suggested that the ferromagnetism in metal oxide


nanostructures can be produced by Stoner-type band splitting rather than exchange
coupling between local moments [180]. This is motivated, in part, by the argument
that, for reasonable materials parameters, RKKY-mediated ferromagnetic order in
these systems should develop only below 20 K. In this model, the defects produce a
density of states NS(E), with a peak lying close to the Fermi level. The introduction
of a local charge reservoir can shift the Fermi level to align with a peak in NS(E),
which can satisfy the Stoner criterion leading to a spin-split impurity band. In this
proposal, the charge reservoir is introduced by mixed valence transition metal
dopants [66], although other mechanisms are also suggested [22, 138, 180]. In
principle, metal ion vacancies, which can act as electron acceptors, could provide
the charge reservoir. This model is particularly relevant for nanostructured metal
oxide systems, as it is suggested that the ferromagnetism would arise only in defect-
rich areas or surfaces, where the defect density of states would be large [138].
While there remains considerable uncertainty surrounding the mechanisms
giving rise to ferromagnetism in metal oxide nanostructures, experimentally it is
clear that weak ferromagnetism is very generally observed in both antiferromag-
netic and diamagnetic systems. Furthermore, a number of studies confirm that this
magnetization is correlated with the density of point defects, whether oxygen
vacancies, metal ion vacancies, or others. Although the saturation magnetization
associated with this ferromagnetism is generally small, and the coercive fields are
also small or vanishing at room temperature for superparamagnetic systems, such
ferromagnetic order represents an emergent property driven solely by point defects,
which is relevant for understanding how these defects modify the behavior of metal
oxide nanostructures.

3.6 Defect Engineering in Metal Oxide Nanostructures

When considering defects in metal oxide nanostructures, it is crucial to recognize that


all of the negative connotations of the word “defect” are strictly associated only with
imperfections in the crystal lattice of the system and should not prejudice the
interpretation of the induced changes in the physical properties. While lattice defects
can have a large effect on the properties of oxide materials, these changes can be
beneficial, detrimental, or neutral depending on the specific application being consid-
ered. An enhanced conductivity produced by charge hopping between defect sites is
certainly a drawback for a metal oxide insulating layer in a MOSFET [181, 182],
but can be crucial for developing optically transparent thin film electrodes [41].
The weak ferromagnetism arising from surface defects in nanoparticles may offer a
route to developing spintronic devices [183], but could also produce spurious mag-
netic signals in nanoscale sensors. It is, however, essential to consider the potential
effects of defect-induced properties when investigating metal oxide nanostructures
and to recognize the opportunities presented by defects in introducing new
functionalities into the system.
60 C. Sudakar et al.

Controlling the defect chemistry in metal oxide nanostructures offers a route to


tuning existing material properties or incorporating new characteristics. This can be
an attractive approach for modifying materials, since this does not involve the
addition of new elements, which reduces the potential for the formation of spurious
secondary phases. The defect structure in nanomaterials can normally be tuned by
varying the conditions during sample preparation, or by post preparation
techniques. The possibility of reversibly controlling the defect structure, such as
vacuum annealing to introduce oxygen vacancy defects then annealing in an
oxygen-rich environment to remove these defects, leads to a tunability that is
completely absent when controlling the material properties by doping. However,
it can be exquisitely difficult to experimentally parameterize defect-induced
properties because of problems associated with quantifying the concentration of
native defects. It is much more straightforward to measure a small percentage of Co
doped into ZnO [32] than it is to determine the degree of oxygen nonstoichiometry
for slightly oxygen deficient samples [59]. While simulations provide a great deal
of insight into the relationship between the concentration of native defects and their
effects on the physical properties of oxides, the ability to tailor the defect structure
in metal oxides for specific applications will require the development of improved
tools to more precisely parameterize the point defects in real systems.
One of the central themes of this short review is that native point defects in
oxides can lead to the emergence of completely new materials properties that are
absent in the stoichiometric parent compound. Since nanostructured materials can
typically develop much higher defect concentrations than bulk systems, researchers
working with oxide nanomaterials should remain cognizant of this effect. One
further complication associated with the defect chemistry of metal oxides arises
from the fact that in some, but by no means all, cases the native point defects can be
unstable. In some materials, oxygen vacancy defects are removed as materials are
held under ambient conditions, while in other systems, metal ion interstitials can
readily diffuse even at room temperature [12]. This can lead to a very complicated
time dependence for the induced physical properties. For example, weak ferromag-
netism in oxygen-deficient TiO2 vanishes after only a few hours for samples stored
under ambient conditions [25], while the magnetic signal in In2O3 films persists for
years [21]. This dynamical evolution in physical properties is likely to be detrimen-
tal for many applications, so it is important to fully investigate the stability of
defect-rich metal oxide nanostructures, along with their response, before they can
be considered for incorporation into devices.

3.7 Conclusions

We have presented a brief and somewhat idiosyncratic overview of defects in metal


oxide nanostructures and how these defects modify the physical properties of these
systems. We have focused primarily on native point defects in binary oxides, mainly
considering only defects and interstitials. The specific nature of the most relevant
3 The Role of Defects in Functional Oxide Nanostructures 61

point defects, whether VO, MI, or other, varies considerably among different
materials, and, in many cases, remains a topic of lively debate. Beyond perturbing
the existing materials properties, such as introducing an impurity paramagnetic
response or increasing the conductivity, the modifications offered by these point
defects can lead to the emergence of entirely new physical behavior. The presence of
point defects can produce metallic conductivity, at least over some range of
temperatures, optical responses characteristic of collective behavior, and weak
ferromagnetism. While the specific mechanisms producing these features vary
considerably from system to system, the broad nature of the response is somewhat
universal, particularly considering the magnetic characteristics. The ability to not
only modify but build new physical properties into metal oxide nanostructures
through defect chemistry greatly expands the functionality of these materials and
is expected to play a crucial role in the next generation of oxide devices.

Acknowledgments We have greatly benefitted from many conversations with A. Dixit,


P. Kharel, B. Nadgorny, R. Naik, V.M. Naik, R. Panguluri, R. Seshadri, R. Suryanarayanan, and
J. Thakur. We acknowledge support from the National Science Foundation through DMR-
0644823, from the Jane and Frank Warchol Foundation, and from the Institute for Manufacturing
Research at Wayne State University.

References

1. Rodrı́guez, J.A., Fernández-Garcı́a, M. (eds.): Synthesis, Properties and Applications of


Oxide Nanomaterials. Wiley, Hoboken, NJ (2007)
2. Cao, G. (ed.): Nanostructures & Nanomaterials: Synthesis Properties & Applications. Impe-
rial College Press, London (2004)
3. Poole Jr., C.P., Owens, F.J.: Introduction to Nanotechnology. Wiley, Hoboken, NJ (2003)
4. Li, M., Li, J.C.: Size effects on the band-gap of semiconductor compounds. Mater. Lett. 60
(20), 2526–2529 (2006)
5. Liou, S.H., et al.: Enhancement of coercivity in nanometer-size CoPt crystallites. J. Appl.
Phys. 85(8), 4334–4336 (1999)
6. Weertman, J.R.: Mechanical behaviour of nanocrystalline metals. In: Koch, C.C. (ed.)
Nanostructured Materials, Processing, Properties and Applications. Noyes, Norwich, NY
(2002)
7. Balamurugan, B., et al.: Size-dependent conductivity-type inversion in Cu2O nanoparticles.
Phys. Rev. B 69, 165419 (2004)
8. Hono, K., Ohnuma, M.: Microstructures and properties of nanocrystalline and nanogranular
magnetic materials. In: Nalwa, H.S. (ed.) Magnetic Nanostructures, p. 300. American Scien-
tific Publishers, Stevenson Ranch, CA (2002)
9. Djurisić, A.B., Leung, Y.H.: Optical properties of ZnO nanostructures. Small 2(8–9),
944–961 (2006)
10. Han, B.Q., Laverni, E.J., Mohamed, F.A.: Mechanical properties of nanostructured materials.
Rev. Adv. Mater. Sci. 9, 1–16 (2005)
11. Janotti, A., Van de Walle, C.G.: Native point defects in ZnO. Phys. Rev. B 76, 165202 (2007)
12. McCluskey, M.D., Jokela, S.J.: Defects in ZnO. J. Appl. Phys. 106(7), 071101 (2009)
13. Ischenko, V., et al.: Zinc oxide nanoparticles with defects. Adv. Funct. Mater. 15, 1945–1954
(2005)
62 C. Sudakar et al.

14. Janotti, A., Van de Walle, C.G.: Fundamentals of zinc oxide as a semiconductor. Rep. Prog.
Phys. 72, 126501 (2009)
15. Diebold, U.: The surface science of titanium dioxide. Surf. Sci. Rep. 48(5–8), 53–229 (2003)
16. Carel, C., Mouallem-Bahout, M., Gaude, J.: Re-examination of the non-stoichiometry and
defect structure of copper(II) oxide or tenorite, Cu1zO or CuO1e – a short review. Solid
State Ionics 117(1–2), 47–55 (1999)
17. Smyth, D.M.: The Defect Chemistry of Metal Oxides. Oxford University Press, Oxford
(2000)
18. Tilley, R.J.D.: Defects in Solids. Wiley, Hoboken, NJ (2008)
19. Kilic, C., Zunger, A.: Origins of coexistence of conductivity and transparency in SnO2. Phys.
Rev. Lett. 88, 095501 (2002)
20. Sudakar, C., et al.: Coexistence of anion and cation vacancy defects in vacuum-annealed
In2O3 thin films. Scripta Materialia 62(2), 63–66 (2010)
21. Panguluri, R.P., et al.: Ferromagnetism and spin-polarized charge carriers in In2O3 thin films.
Phys. Rev. B 79, 165208 (2009)
22. Stoneham, M.: The strange magnetism of oxides and carbons. J. Phys.: Condens. Matter 22,
074211–074218 (2010)
23. Kofstad, P., Norby, T.: Defects and Transport in Crystalline Solids. University of Oslo, Oslo
(2007)
24. Sudakar, C., et al.: Raman spectroscopic studies of oxygen defects in Co-doped ZnO films
exhibiting room-temperature ferromagnetism. J. Phys.: Condens. Matter 19(2), 026212
(2007)
25. Sudakar, C., et al.: Room temperature ferromagnetism in vacuum-annealed TiO2 thin films.
J. Magn. Magn. Mater. 320(5), L31–L36 (2008)
26. Halliburton, L.E., et al.: Production of native donors in ZnO by annealing at high temperature
in Zn vapor. Appl. Phys. Lett. 87(17), 172108 (2005)
27. Weber, M.H., et al.: Defect engineering of ZnO. Appl. Surf. Sci. 255(1), 68–70 (2008)
28. Reece, M., Morrell, R.: Electron-microscope study of nonstoichiometric titania. J. Mater. Sci.
26(20), 5566–5574 (1991)
29. Erhart, P., Albe, K., Klein, A.: First-principles study of intrinsic point defects in ZnO: Role of
band structure, volume relaxation, and finite-size effects. Phys. Rev. B 73, 205203 (2006)
30. Janotti, A., Van de Walle, C.G.: Oxygen vacancies in ZnO. Appl. Phys. Lett. 87, 122102
(2005)
31. Kohan, A.F., et al.: First-principles study of native point defects in ZnO. Phys. Rev. B 61(22),
15019–15027 (2000)
32. Ali, B., et al.: Interplay of dopant, defects and electronic structure in driving ferromagnetism
in Co-doped oxides: TiO2, CeO2 and ZnO. J. Phys.: Condens. Matter 21(45), 456005 (2009)
33. Kim, D., et al.: The origin of oxygen vacancy induced ferromagnetism in undoped TiO2.
J. Phys.: Condens. Matter 21(19), 195405 (2009)
34. Khalid, M., et al.: Defect-induced magnetic order in pure ZnO films. Phys. Rev. B 80(3),
035331 (2009)
35. von Wenckstern, H., et al.: Anionic and cationic substitution in ZnO. Prog. Solid State Chem.
37(2–3), 153–172 (2009)
36. Zhang, S.B., Wei, S.H., Zunger, A.: Intrinsic n-type versus p-type doping asymmetry and the
defect physics of ZnO. Phys. Rev. B 63, 075205 (2001)
37. Cox, P.A.: Transition Metal Oxides. Clarendon, Oxford (1992)
38. Catlow, C.R.A., James, R.: Disorder in TiO2x. Proc. R. Soc. Lond. A. 384, 157–173 (1982)
39. Yoon, S.D., et al.: Oxygen-defect-induced magnetism to 880 K in semiconducting anatase
TiO2-delta films. J. Phys.: Condens. Matter 18(27), L355–L361 (2006)
40. Yoon, S.D., et al.: Magnetic semiconducting anatase TiO2-[delta] grown on (1 0 0) LaAlO3
having magnetic order up to 880 K. J. Magn. Magn. Mater. 309(2), 171–175 (2007)
41. Dixit, A., et al.: Undoped vacuum annealed In2O3 thin films as a transparent conducting
oxide. Appl. Phys. Lett. 95(19), 192105 (2009)
3 The Role of Defects in Functional Oxide Nanostructures 63

42. Rao, C.N.R., Raveau, B.: Transition Metal Oxides. VCH, New York (1995)
43. Norton, D.P.: Synthesis and properties of epitaxial electronic oxide thin-film materials.
Mater. Sci. Eng. R Rep. 43(5–6), 139–247 (2004)
44. Pacchioni, G.: Ab initio theory of point defects in oxide materials: Structure, properties,
chemical reactivity. Solid State Sci. 2, 161–179 (2000)
45. Domaradzki, J., et al.: Transparent oxide semiconductors based on TiO2 doped with V, Co
and Pd elements. J. Non-Cryst. Solids 352, 2324–2327 (2006)
46. Heluani, S.P., et al.: Electrical conductivity mechanisms in zinc oxide thin films deposited by
pulsed laser deposition using different growth environments. Thin Solid Films 515,
2379–2386 (2006)
47. Ishizuka, S., et al.: Control of hole carrier density of polycrystalline Cu2O thin films by Si
doping. Appl. Phys. Lett. 80, 950–952 (2002)
48. Tomlins, G.W., Routbort, J.L., Mason, T.O.: Zinc self-diffusion, electrical properties, and
defect structure of undoped, single crystal zinc oxide. J. Appl. Phys. 87, 117–123 (2000)
49. Ozgur, U., et al.: A comprehensive review of ZnO materials and devices. J. Appl. Phys. 98(4),
041301 (2005)
50. Morkoc, H., Ozgur, U.: Zinc Oxide. Wiley-VCH, Weinheim (2009)
51. Coleman, V.A., Jagdish, C.: Zinc Oxide Bulk, Thin Films and Nanostructures. Elsevier,
Oxford (2006)
52. Selim, F.A., et al.: Nature of native defects in ZnO. Phys. Rev. Lett. 99, 085502 (2007)
53. van de Walle, C.G.: Defect analysis and engineering in ZnO. In: 21st International Confer-
ence on Defects in Semiconductors, ICDS-21, Elsevier, Netherlands, 16–20 July 2001.
54. Tomita, T., et al.: The origin of n-type conductivity in undoped In2O3. Appl. Phys. Lett.
87(5), 051911 (3p) (2005)
55. Hamberg, I., Granqvist, C.G.: Evaporated Sn-doped In2O3 films: Basic optical properties and
applications to energy-efficient windows. J. Appl. Phys. 60, R123 (1986)
56. Hartnagel, H.L., Dawar, A.K.L., Jagadish, C.: Semiconducting Transparent Thin Films.
Institute of Physics Publishing, Bristol (1995)
57. de Wit, J.H.W.: The high temperature behavior of In2O3. J. Solid State Chem. 13, 192 (1975)
58. Frank, G., Kostlin, G.: Electrical properties and defect model of tin-doped indium oxide
layers. Appl. Phys. A 27, 197–206 (1982)
59. Bellingham, J.R., Mackenzie, A.P., Phillips, W.A.: Precise measurements of oxygen content:
Oxygen vacancies in transparent conducting indium oxide films. Appl. Phys. Lett.
58, 2506–2508 (1991)
60. Agoston, P., et al.: Geometry, electronic structure and thermodynamic stability of intrinsic
point defects in indium oxide. J. Phys.: Condens. Matter 21, 455801–455812 (2009)
61. Soon, A., et al.: Native defect-induced multifarious magnetism in nonstoichiometric cuprous
oxide: first-principles study of bulk and surface properties of Cu2dO. Phys. Rev. B 7, 035205
(2009)
62. Jang, W.-L., et al.: Point defects in sputtered NiO films. Appl. Phys. Lett. 94, 062103 (2009)
63. Biju, V., Khadar, M.A.: AC conductivity of nanostructured nickel oxide. J. Mater. Sci.
36, 5779–5787 (2001)
64. Hasiguti, R.R., Yagi, E.: Electrical conductivity below 3 K of slightly reduced oxygen-
deficient rutile TiO2x. Phys. Rev. B 49, 7251–7256 (1994)
65. Lunkenheimer, P., et al.: Correlated barrier hopping in NiO films. Phys. Rev. B 44,
5927–5930 (1991)
66. Coey, J.M.D., et al.: Charge-transfer ferromagnetism in oxide nanoparticles. J. Phys. D 41,
134012 (2008)
67. Xu, Z., et al.: Ferromagnetism in pure wurtzite zinc oxide. J. Appl. Phys. 105, 507–508
(2009)
68. Henrich, V.E.: The surfaces of metal oxides. Rep. Prog. Phys. 48, 1481–1541 (1985)
69. Munnix, S., Schmeits, M.: Electronic structure of oxygen vacancies on TiO2(110) and
SnO2(110) surfaces. In: 10th International Vacuum Congress (IVC-10), 6th International
64 C. Sudakar et al.

Conference on Solid Surfaces (ICSS-6) and 33 rd National Symposium of the American


Vacuum Society, USA, 27–31 Oct 1986 (1987)
70. Harvey, S.P., et al.: Surface versus bulk electronic/defect structures of transparent conducting
oxides: I. Indium oxide and ITO. J. Phys. D: Appl. Phys 39, 3959–3968 (2006)
71. Niemann, D., et al.: Plasmon excitation by multiply charged Neq+ ions interacting with an Al
surface. Phys. Rev. Lett. 80, 3328–3331 (1998)
72. Demangeot, F., et al.: Experimental study of LO phonons and excitons in ZnO nanoparticles
produced by room-temperature organometallic synthesis. Appl. Phys. Lett. 88, 71921 (2006)
73. Pai-Chun, C., et al.: Finite size effect in ZnO nanowires. Appl. Phys. Lett. 90, 113101 (2007)
74. Wang, J.B., et al.: Raman scattering and high temperature ferromagnetism of Mn-doped ZnO
nanoparticles. Appl. Phys. Lett. 88, 252502 (2006)
75. Lima, S.A.M., et al.: Luminescent properties and lattice defects correlation on zinc oxide. Int.
J. Inorg. Mater. 3(7), 749–754 (2001)
76. Kong, Y.C., et al.: Ultraviolet-emitting ZnO nanowires synthesized by a physical vapor
deposition approach. Appl. Phys. Lett. 78, 407–409 (2001)
77. Lin, H.Y., et al.: Laser action in Tb(OH)3/SiO2 photonic crystals. Opt. Express 16,
16697–16703 (2008)
78. Guichard, A.R., et al.: Tunable light emission from quantum-confined excitons in TiSi2-
catalyzed silicon nanowires. Nano Lett. 6, 2140–2144 (2006)
79. Hummer, K.: Interband magnetoreflection of ZnO. Phys. Status Solidi B 56, 249–260 (1973)
80. Makino, T., et al.: Band gap engineering based on MgxZn1xO and CdyZn1yO ternary alloy
films. Appl. Phys. Lett. 78, 1237–1239 (2001)
81. Jin, B.J., Im, S., Lee, S.Y.: Violet and UV luminescence emitted from ZnO thin films grown
on sapphire by pulsed laser deposition. Thin Solid Films 366, 107–110 (2000)
82. Bagnall, D.M., et al.: Room temperature excitonic stimulated emission from zinc oxide
epilayers grown by plasma-assisted MBE. In: Eighth International Conference on II-VI
Compounds, Elsevier, Netherlands, 25–29 Aug 1997 (1998)
83. Bylander, E.G.: Surface effects on the low-energy cathodoluminescence of zinc oxide.
J. Appl. Phys. 49, 1188–1195 (1978)
84. Garces, N.Y., et al.: Role of copper in the green luminescence from ZnO crystals. Appl. Phys.
Lett. 81, 622–624 (2002)
85. Prosanov, I.Y., Politov, A.A.: Intrinsic defects and luminescence of zinc oxide. Inorg. Mater.
31, 663 (1995)
86. Egehaaf, J.H., Oelkrug, D.: Luminescence and nonradiative deactivation of excited states
involving oxygen defect centers in polycrystalline ZnO. J. Cryst. Growth 161, 190–194
(1996)
87. Vanheusden, K., et al.: Mechanisms behind green photoluminescence in ZnO phosphor
powders. J. Appl. Phys. 79, 7983–7990 (1996)
88. Zhou, H., Cai, W., Zhang, L.: Photoluminescence of indium-oxide nanoparticles dispersed
within pores of mesoporous silica. Appl. Phys. Lett. 75, 495–497 (1999)
89. Soares, J.W., et al.: Novel photoluminescence properties of surface-modified nanocrystalline
zinc oxide: Toward a reactive scaffold. Langmuir 24, 371–374 (2008)
90. Mo, C.M., et al.: Enhancement effect of photoluminescence in assemblies of nano-ZnO
particles/silica aerogels. J. Appl. Phys. 83, 4389–4391 (1998)
91. Singh, S., Rao, M.S.R.: Optical and electrical resistivity studies of isovalent and aliovalent
3d transition metal ion doped ZnO. Phys. Rev. B 80, 045210–045220 (2009)
92. Godlewski, M.: ODMR study of the nature of iron-red photoluminescence in ZnS crystals.
Solid State Commun. 47, 811–813 (1983)
93. Goldewski, M., Skowronski, M.: Effective deactivation of the ZnS visible photolumi-
nescence by iron impurities. Phys. Rev. B 32, 4007–4013 (1985)
94. Mang, A., Reimann, K., Rubenacke, S.: Band gaps, crystal-field splitting, spin-orbit coupling,
and exciton binding energies in ZnO under hydrostatic pressure. Solid State Commun.
94, 251–254 (1995)
3 The Role of Defects in Functional Oxide Nanostructures 65

95. Singh, S., et al.: Investigation of low-temperature excitonic and defect emission from
Ni-doped ZnO nanoneedles and V-doped ZnO nanostructured film. New J. Phys.
12, 023007–023018 (2010)
96. van Dijken, A., et al.: The kinetics of the radiative and nonradiative processes in
nanocrystalline ZnO particles upon photoexcitation. J. Phys. Chem. B 104(8), 1715–1723
(2000)
97. Ginley, D.S., Bright, C.: Transparent conducting oxides. MRS Bulletin 25, 15–21 (2000)
98. Dai, L., et al.: Fabrication and characterization of In2O3 nanowires. Appl. Phys. A A75,
687–689 (2002)
99. Ng, Y.H., et al.: Fabrication of hollow carbon nanospheres encapsulating platinum
nanoparticles using a photocatalytic reaction. Adv. Mater. 19, 597–601 (2007)
100. Soulantica, K., et al.: Synthesis of indium and indium oxide nanoparticles from indium
cyclopentadienyl precursor and their application for gas sensing. Adv. Funct. Mater.
13, 553–557 (2003)
101. Lee, M.S., et al.: Characterization of the oxidized indium thin films with thermal oxidation.
Thin Solid Films 279, 1–3 (1996)
102. Wu, P., et al.: Synthesis and photoluminescence property of indium oxide nanowires. Appl.
Surf. Sci. 255, 3201–3204 (2008)
103. Singh, S., et al.: Formation of ZnO nanobrushes in direct atmosphere using a carbon catalyst
and a Zn metal source. Nano 3, 361–365 (2008)
104. Yang, J.H., et al.: Low temperature hydrothermal growth and optical properties of ZnO
nanorods. Cryst. Res. Technol. 44, 87–91 (2009)
105. Ashkenov, N., et al.: Infrared dielectric functions and phonon modes of high-quality ZnO
films. J. Appl. Phys. 93, 126–133 (2003)
106. Yang, Y.H., et al.: Radial ZnO nanowire nucleation on amorphous carbons. Appl. Phys. Lett.
87, 183109 (2005)
107. Yang, R.D., et al.: Photoluminescence and micro-Raman scattering in ZnO nanoparticles:
The influence of acetate adsorption. Chem. Phys. Lett. 411, 150–154 (2005)
108. Rajalakshmi, M., et al.: Optical phonon confinement in zinc oxide nanoparticles. J. Appl.
Phys. 87, 2445–2448 (2000)
109. Alim, K.A., Fonoberov, V.A., Balandin, A.A.: Origin of the optical phonon frequency shifts
in ZnO quantum dots. Appl. Phys. Lett. 86, 053103 (2005)
110. Bundesmann, C., et al.: Raman scattering in ZnO thin films doped with Fe, Sb, Al, Ga, and Li.
Appl. Phys. Lett. 83, 1974–1976 (2003)
111. Manjon, F.J., et al.: Silent Raman modes in zinc oxide and related nitrides. J. Appl. Phys.
97, 053516 (2005)
112. Kostic, R., et al.: Low-frequency Raman scattering from ZnO(Fe) nanoparticles. Acta
Physica Polonica A 116, 65–67 (2009)
113. Zhang, P., Guo, Y.: Surface-enhanced Raman scattering inside metal nanoshells. J. Am.
Chem. Soc. 131(11), 3808–3809 (2009)
114. Hirsch, L.R., et al.: A whole blood immunoassay using gold nanoshells. Anal. Chem.
75, 2377–2381 (2003)
115. Jackson, J.B., et al.: Controlling the surface enhanced Raman effect via the nanoshell
geometry. Appl. Phys. Lett. 82, 257–259 (2003)
116. Albrecht, M.G., Creighton, J.A.: Anomalously intense Raman spectra of pyridine at a silver
electrode. J. Am. Chem. Soc. 99, 5215 (1977)
117. Chao, S., et al.: A double substrate sandwich structure for fiber surface enhanced Raman
scattering detection. Appl. Phys. Lett. 92, 103107 (2008)
118. Yi, Z., et al.: Liquid core photonic crystal fiber sensor based on surface enhanced Raman
scattering. Appl. Phys. Lett. 90, 193504 (2007)
119. Gu, Y., et al.: Quantum confinement in ZnO nanorods. Appl. Phys. Lett. 85, 3833 (2004)
66 C. Sudakar et al.

120. Sakano, T., et al.: Surface enhanced Raman scattering properties using Au-coated ZnO
nanorods grown by two-step, off-axis pulsed laser deposition. J. Phys. D: Appl. Phys.
41, 235304 (2008)
121. Pettinger, B.: Light scattering by adsorbates at Ag particles: Quantum-mechanical approach
for energy transfer induced interfacial optical processes involving surface plasmons,
multipoles, and electron-hole pairs. J. Chem. Phys. 85, 7442–7451 (1986)
122. Liu, C.Y., et al.: Anomalously enhanced Raman scattering from longitudinal optical phonons
on Ag-nanoparticle-covered GaN and ZnO. Appl. Phys. Lett. 96, 033109 (2010)
123. Cullum, B., et al.: Characterization of multilayer-enhanced surface-enhanced Raman scatter-
ing (SERS) substrates and their potential for SERS nanoimaging. NanoBioTechnology 3(1),
1–11 (2007)
124. Habermeier, H.U.: Strategies towards controlling strain-induced mesoscopic phase separa-
tion in manganite thin films. J. Phys.: Condens. Matter 20(43), 434228 (2008)
125. Przezdziecka, E.: Optical and magnetooptical properties of the p-type ZnMnO. E-MRS Fall
Meeting Symposium F (2006)
126. Arora, S.K., et al.: Antiphase boundaries induced exchange coupling in epitaxial Fe3O4 thin
films. J. Magn. Magn. Mater. 286, 463–467 (2005)
127. Eerenstein, W., et al.: Origin of the increased resistivity in epitaxial Fe3O4 films. Phys. Rev. B
66, 201101 (2002)
128. Siang Huei, L., Jian Ping, W., Teck Seng., L.: Study of in-depth defects using magneto-
optical Kerr effect by measuring the magnetic hardness coefficient in magnetic thin films.
In: INTERMAG 2000 Digest of Technical Papers. 2000 IEEE International Magnetics
Conference, IEEE, USA, 9–13 April 2000
129. Thomas, D.G., Hopfield, J.J.: Optical properties of bound exciton complexes in cadmium
sulfide. Phys. Rev. 128(5), 2135 (1962)
130. Loose, P., Rosenzweig, M., W€ ohlecke, M.: Zeeman effect of bound exciton complexes in
ZnO. Phys. Status Solidi (b) 75(1), 137–144 (1976)
131. Reynolds, D.C., Litton, C.W., Collins, T.C.: Zeeman effects in the edge emission and
absorption of ZnO. Phys. Rev. 140(5(A)), A1726 (1965)
132. Margulies, D.T., et al.: Origin of the anomalous magnetic behavior in single crystal Fe3O4
films. Phys. Rev. Lett. 79, 5162–5165 (1997)
133. Barbier, A., et al.: Surface and bulk spin ordering of antiferromagnetic materials: NiO(111).
Phys. Rev. Lett. 93, 257208–2572081 (2004)
134. Tang, Y.J., et al.: Finite size effects on the moment and ordering temperature in antiferro-
magnetic CoO layers. Phys. Rev. B 67, 054408 (2003)
135. Kapilashrami, M., et al.: Transition from ferromagnetism to diamagnetism in undoped ZnO
thin films. Appl. Phys. Lett. 95, 033104 (2009)
136. Khaibullin, R.I., et al.: Formation of anisotropic ferromagnetic response in rutile (TiO2)
implanted with cobalt ions. Nucl. Instr. Meth. B 257, 369–373 (2007)
137. Dwight, K., Menyuk, N.: Magnetic properties of Mn3O4 and the canted spin problem. Phys.
Rev. 119(5), 1470 (1960)
138. Sundaresan, A., et al.: Ferromagnetism as a universal feature of nanoparticles of the other-
wise nonmagnetic oxides. Phys. Rev. B 74, 161306 (2006)
139. Riveiro, J.M., et al.: CoO1d layers in a reactively sputtered exchange-bias system. New
J. Phys. 10, 083028 (2008)
140. Mishra, S.R., et al.: Anomalous magnetic properties of mechanically milled cobalt oxide
nanoparticles. J. Nanosci. Nanotechnol. 5, 2076–2081 (2005)
141. Jagodic, M., et al.: Surface-spin magnetism of antiferromagnetic NiO in nanoparticle and
bulk morphology. J. Phys.: Condens. Matter 21, 215302 (2009)
142. Morales, M.A., et al.: Surface anisotropy and magnetic freezing of MnO nanoparticles. Phys.
Rev. B75, 134423 (2007)
143. Gruyters, M.: Spin-glass-like behavior in CoO nanoparticles and the origin of exchange bias
in layered CoO/ferromagnet structures. Phys. Rev. Lett. 95, 077204 (2005)
3 The Role of Defects in Functional Oxide Nanostructures 67

144. Tomou, A., et al.: Weak ferromagnetism and exchange biasing in cobalt oxide nanoparticle
systems. J. Appl. Phys. 99, 123915 (2006)
145. Tracy, J.B., Bawendi, M.G.: Defects in CoO in oxidized cobalt nanoparticles dominate
exchange biasing and exhibit anomalous magnetic properties. Phys. Rev. B 74, 184434
(2006)
146. Tracy, J.B., et al.: Exchange biasing and magnetic properties of partially and fully oxidized
colloidal cobalt nanoparticles. Phys. Rev. B 72, 064404 (2005)
147. Richardson, J.T., et al.: Origin of superparamagnetism in nickel oxide. J. Appl. Phys. 70,
6977–6982 (1991)
148. Makhlouf, S.A., et al.: Magnetic anomalies in NiO nanoparticles. In: 41st Annual Conference
on Magnetism and Magnetic Materials, AIP, USA, 12–15 Nov 1996 (1997)
149. Ghosh, M., et al.: MnO and NiO nanoparticles: Synthesis and magnetic properties. J. Mater.
Chem. 16, 106–111 (2006)
150. Coey, J.M.D.: High-temperature ferromagnetism in dilute magnetic oxides. J. Appl. Phys. 97
(10), 10D313 (2005)
151. Coey, J.M.D., et al.: Magnetism in hafnium dioxide. Phys. Rev. B 72, 024450 (2005)
152. Hong, N.H., et al.: Room-temperature ferromagnetism observed in undoped semiconducting
and insulating oxide thin films. Phys. Rev. B 73, 132404 (2006)
153. Matsumoto, Y., et al.: Room-temperature ferromagnetism in transparent transition metal-
doped titanium dioxide. Science 291(5505), 854–856 (2001)
154. Sharma, P., et al.: Ferromagnetism above room temperature in bulk and transparent thin films
of Mn-doped ZnO. Nat. Mater. 2(10), 673–677 (2003)
155. Prellier, W., Fouchet, A., Mercey, B.: Oxide-diluted magnetic semiconductors: A review of
the experimental status. J. Phys.: Condens. Matter 15(37), R1583–R1601 (2003)
156. Liu, C., Yun, F., Morkoc, H.: Ferromagnetism of ZnO and GaN: A review. J. Mater. Sci.:
Mater. Electron. 16(9), 555–597 (2005)
157. Park, J.H., et al.: Co-metal clustering as the origin of ferromagnetism in Co-doped ZnO thin
films. Appl. Phys. Lett. 84(8), 1338–1340 (2004)
158. Rao, C.N.R., Deepak, F.L.: Absence of ferromagnetism in Mn- and Co-doped ZnO. J. Mater.
Chem. 15(5), 573–578 (2005)
159. Lawes, G., et al.: Absence of ferromagnetism in Co and Mn substituted polycrystalline ZnO.
Phys. Rev. B 71, 045201 (2005)
160. Abraham, D.W., Frank, M.M., Guha, S.: Absence of magnetism in hafnium oxide films.
Appl. Phys. Lett. 87(25), 252502 (2005)
161. Coey, J.M.D.: Dilute magnetic oxides. Curr. Opin. Solid State Mater. Sci. 10(2), 83–92
(2006)
162. Hong, N.H., Sakai, J., Brize, V.: Observation of ferromagnetism at room temperature in ZnO
thin films. J. Phys.: Condens. Matter 19(3), 036219 (2007)
163. Liu, Y.L., et al.: Size dependent ferromagnetism in cerium oxide (CeO2) nanostructures
independent of oxygen vacancies. J. Phys.: Condens. Matter 20(16), 165201 (2008)
164. Han, X., Lee, J., Yoo, H.-I.: Oxygen-vacancy-induced ferromagnetism in CeO2 from first
principles. Phys. Rev. B 79, 100403 (2009)
165. Song, Y.Q., et al.: Direct evidence of oxygen vacancy mediated ferromagnetism of Co doped
CeO2 thin films on Al2O3(0001) substrates. J. Phys.: Condens. Matter 20(33), 255510 (2008)
166. Coey, J.M.D.: d0 Ferromagnetism. Solid State Sci. 7(6), 660–667 (2005)
167. Kim, D., Yang, J.-H., Hong, J.: Ferromagnetism induced by Zn vacancy defect and lattice
distortion in ZnO. J. Appl. Phys. 106(1), 013908 (2009)
168. Lau, S.P., et al.: Zn-interstitial-enhanced ferromagnetism in Cu-doped ZnO films. J. Magn.
Magn. Mater. 315, 107–110 (2007)
169. Kharel, P., et al.: Room temperature ferromagnetism in Cr-doped In2O3 on high vacuum
annealing of thin films and bulk samples. J. Appl. Phys. 101(9), 09H117 (2007)
170. Potzger, K., Zhou, S.: Non-DMS related ferromagnetism in transition metal doped zinc oxide.
Phys. Status Solidi (b) 246(6), 1147–1167 (2009)
68 C. Sudakar et al.

171. Park, Y.D., et al.: Anomalous Hall effect in wide bandgap diluted magnetic semiconductors
co-doped with non-magnetic impurities. Proc. Electrochem. Soc. 6, 300–311 (2004)
172. Ney, A., et al.: Absence of intrinsic ferromagnetic interactions of isolated and paired Co
dopant atoms in Zn1xCoxO with high structural perfection. Phys. Rev. Lett. 100, 157201
(2008)
173. Soulen Jr., R.J., et al.: Measuring the spin polarization of a metal with a superconducting
point contact. Science 282(5386), 85–88 (1998)
174. Qingyu, X., et al.: Spin manipulation in co-doped ZnO. Phys. Rev. Lett. 101, 076601 (2008)
175. Ueda, K., Tabata, H., Kawai, T.: Magnetic and electric properties of transition-metal-doped
ZnO films. Appl. Phys. Lett. 79, 988–990 (2001)
176. Fernandes, V., et al.: Dilute-defect magnetism: Origin of magnetism in nanocrystalline CeO2.
Phys. Rev. B 80, 035202 (2009)
177. Rahman, G., Garcia-Suarez, V.M., Soon Cheol, H.: Vacancy-induced magnetism in SnO2:
A density functional study. Phys. Rev. B 78, 184404 (2008)
178. Philip, J., et al.: Carrier-controlled ferromagnetism in transparent oxide semiconductors. Nat.
Mater. 5, 298–304 (2006)
179. Raebiger, H., Lany, S., Zunger, A.: Control of ferromagnetism via electron doping in In2O3:
Cr. Phys. Rev. Lett. 101, 4 (2008)
180. Coey, J.M.D., Venkatesan, M., Fitzgerald, C.B.: Donor impurity band exchange in dilute
ferromagnetic oxides. Nat. Mater. 4(2), 173–179 (2005)
181. Hutin, L., et al.: GeOI pMOSFETs scaled down to 30-nm gate length with record off-state
current. IEEE Electron Device Lett. 31(3), 234–236 (2010)
182. Qiang, F., Wagner, T.: Interaction of nanostructured metal overlayers with oxide surfaces.
Surf. Sci. Rep. 62, 431–498 (2007)
183. Wolf, S.A., et al.: Spintronics: A spin-based electronics vision for the future. Science 294,
1488–1495 (2001)
Chapter 4
Emergent Metal–Insulator Transitions
Associated with Electronic Inhomogeneities
in Low-Dimensional Complex Oxides

An-Ping Li and Thomas Z. Ward

4.1 Introduction

It is becoming increasingly clear that the exotic properties displayed by correlated


electronic materials such as CMR in manganites, high-Tc superconductivity in
cuprates, and non-Fermi liquid behavior in heavy-fermion compounds are inti-
mately related to the coexistence of competing nearly degenerate states which
couple with active degrees of freedom – charge, lattice, orbital, and spin states
[1, 2]. The striking phenomena associated with these materials are due in large part
to spatial electronic inhomogeneities or nanoscale phase separation [3]. In many of
these hard materials, the functionality is a result of the soft electronic components
that lead to self-organization [4]. These materials require a new approach to their
study. With so many layers of electronic correlation driving the observable self-
organized properties, it is vital to find methods which will allow for single
ingredients to be tuned independently while at the same time offering the ability
to recognize small local shifts that drive the larger macroscopic properties. By using
low-dimensional materials and multilength scale transport probing techniques, it is
possible to isolate individual contributions from specific interactions through their
inherent length scales. In this chapter, we will discuss several novel approaches on
complex oxides that are helping to untangle these interactions.
While much work has been done on colossal electroresistance, CMR, high-Tc
superconductivity, and the MIT in complex materials, there is yet no known
theoretical model that is capable of fully explaining any one of these behaviors

A.-P. Li (*)
Center for Nanophase Materials Sciences, Oak Ridge National Laboratory,
Oak Ridge, TN 37831, USA
e-mail: apli@ornl.gov
T.Z. Ward
Materials Science and Technology Division, Oak Ridge National Laboratory,
Oak Ridge, TN 37831, USA

J. Wu et al. (eds.), Functional Metal Oxide Nanostructures, Springer Series 69


in Materials Science 149, DOI 10.1007/978-1-4419-9931-3_4,
# Springer Science+Business Media, LLC 2012
70 A.-P. Li and T.Z. Ward

Fig. 4.1 (a) Diagram of effects of confinement on transport in a phase-separated material.


(b) Electronic correlation lengths offer new length scales beyond statistical and quantum limits
to exploit

let alone a unifying understanding capable of explaining the effects of complexity


on emergent behavior as a whole. The one thing that many of the materials
exhibiting these properties share is electronic phase separation (PS) [5]. For this
reason, a better understanding of electronic phase separation will have far reaching
implications across a wide range of materials. Many systems with correlated
electrons, such as manganites, ruthenates, cobaltates, and cuprates, have been
shown to contain inherent electronic inhomogeneity near a phase transition. The
fact that electronic phase separation is present in so many materials of such varying
character raises many questions as to its origin and its role in correlated electronic
systems.
A particularly interesting approach is to use spatial confinement techniques on
complex oxides. In transport measurements on unconfined systems where device
size is larger than the inherent electronic phase domains, current bypasses regions
of high resistance in favor of regions with lower resistance, because the probing
electrons will follow the path of least resistance. By confining complex materials
exhibiting emergent phenomena to length scales smaller than the electronic phase
domains that reside within them, it is possible to simultaneously probe multiple
resistive regions (Fig. 4.1a). This method allows for a much more complete view of
the phases residing in a material and gives vital information on phase formation,
movement, and fluctuation. Since these phase-separated regions also possess varied
properties, this technique promises to lead to unexpected functionalities for future
device applications.
Similarly, spatial confinement and local probes should make it possible to
probe the delicate interplay between spin, charge, orbital, and lattice interactions
that drive the mechanisms governing emergent physical phenomena in correlated
systems. Each layer of interaction that contributes to the observable self-organized
behaviors is thought to carry with it a unique correlation length scale [6–8]. By
confining these materials and locating characterization probes across a spectrum of
lengths, it should be possible to create systems in which individual contributions from
specific interactions can be manipulated through their correlation lengths (Fig. 4.1b).
4 Emergent Metal–Insulator Transitions... 71

These studies will be invaluable in unraveling these systems since varied correlation
groupings can be experimentally produced through doping, strain, and degree
of confinement.

4.2 Experimental Approach

4.2.1 Fabrication of Spatially Confined Oxide Nanostructures

While there are several methods to directly write confined structure using scanning
tunneling microscopes (STMs) [9], atomic force microscopes [10], and self-assembly
during growth [11, 12], it is often desirable to start with a macroscopic sample and
etch to a desired geometry. This allows for measurements to be taken at macroscales
and repeated through successive confinements. There are several methods at our
disposal to spatially confine thin films. These methods include focused ion beam
(FIB) milling, atomic force microscopy (AFM) lithography, electron beam lithogra-
phy (e-beam), direct laser writing, and optical lithography.
The FIB and AFM methods mechanically remove atoms and do not require a
mask or photoresist layer to be put down [10, 13]. AFM lithography uses direct
probe tip contact to mechanically remove material. This technique can shape down
to 10 nm; however, it is very slow, and there is a danger of disrupting the crystalline
structure of the film. FIB milling is based on a system very similar to scanning
electron microscopy (SEM), but, instead of using focused electrons, it uses a beam
of focused ions with a high enough energy to etch the film from the substrate.
Though this method is relatively slow it has a resolution of tens of nanometer
(Fig. 4.2) but carries with it the danger of ion contamination [14].
E-beam and laser writing are similar to traditional wide beam photoresistive
lithography in their etching procedure but do not require a mask during resist
development which increases the flexibility to quickly create novel geometries

Fig. 4.2 SEM images of (a) 1 mm  5 mm  50 nm wire created using the FIB method, (b) and
(c) show a comparison of edge roughness on 1 mm-wide wires using FIB milling and wet-etch
photolithography, respectively
72 A.-P. Li and T.Z. Ward

Fig. 4.3 Diagram of etching procedure steps and optical image of two confined etches from single
thin film having constriction wires of 10 mm  50 mm  50 nm and 1 mm  5 mm  50 nm with
gold contact pads for four-contact transport

without the need to fabricate new masks while creating structures of 10–100 nm.
However the exposure times can be considerably longer due to the need to finely
scan the active beam [15, 16].
The most widely used and fastest method to create confined structures is
photolithography (Fig. 4.3). In this method, a film is grown on a lattice-matched
or slightly mismatched substrate. A layer of photoresist is applied to the film. This
layer is reactive to UV light. A metallic mask having the desired geometry is placed
on top of the photoresist layer and exposed to a UV light source. This creates a
patterned photoresist layer on the film. The structure is then placed in an etching
solution. The etchant removes the exposed film while leaving the film under the
photoresist layer. The remaining photoresist layer is then removed using a solvent
and leaves the confined structure without damage. This method is only capable of
creating structures larger than about 500 nm and limits the geometry to the
predefined parameters of the mask used. To create new geometries, a new mask
must be made.
The methods discussed each have strengths and weaknesses that must be
matched to the material system under consideration. For the study of very long-
range interactions in a material such as PS manganites, photolithography is a very
attractive and useful method; however, as correlation lengths and organization
levels of interest shrink to nanometer scales, FIB milling and e-beam lithography
become the only methods capable of allowing for consecutive scaling transport
measurements from a single sample.
4 Emergent Metal–Insulator Transitions... 73

4.2.2 Cryogenic Four-Probe STM

Measuring electron transport behaviors in nanostructured materials presents a


significant challenge to experimental physicists. Conventional transport
electrodes and probes are very invasive; namely, they change what we are trying
to measure. For a large conductor, the probes only represent a minor perturbation.
But for a small conductor, especially at the nanoscale, the probes can very well be
the dominant source of scattering. Therefore, using weakly coupled scanning
tunneling probes to detect transport phenomena around individual scatterers
appears to be the best approach [17]. The advantage of using the four-point
method is that it separates the current supplying electrodes from the voltage-
probing electrodes and thus can eliminate the parasitic resistance introduced by
the probes. Furthermore, for observing quantum behavior of electrons, it is often
necessary for transport measurements to be carried out in the nanometer scale at
cryogenic temperatures so that electronic excitation processes can be controlled
and studied in such nanosystems.
Recently, several groups have developed four-probe STMs that can perform
four-point electrical measurements with probe spacing down to the microscopic
scale [18–23]. With only a few exceptions [18, 22], these systems provide a
cryogenic environment only for the sample stage, while the STM probes stay
at an unregulated higher temperature. Transport experiments in these systems
would involve hot electrons injected from probes into samples at a much lower
temperature. Such a temperature gradient between the probe and the sample
severely impacts the stability and thus the reliability of the experiments. More-
over, injected hot electrons will take a long time to reach thermal equilibrium with
the sample (lattice temperature) usually through an electron–phonon coupling
that is very weak at low temperatures, imposing an additional level of complexity
to the interpretation of the experiments. The temperature management in these
multiple-probe STM systems is a serious challenge.
The cryogenic four-probe STM represented in Fig. 4.4 directly addresses
the demand of nanotransport research, which allows for in situ sample preparation
and treatment, nanoscale precision of sample and probe positioning, together with
the four-point transport probing potential. The cryogenic multiprobe STM system
has integrated multiple functional modules in a single system to provide
capabilities of material synthesis, high resolution microscopy, chemical analysis,
cryogenic temperature control both for sample and probes, atomic resolution
imaging and spectroscopy, four independently controllable STM probes,
and local electrical transport probing. In the cryogenic four-probe STM, both
sample and probes are kept at the same low temperature. Each of the four probes
works independently as a cryogenic STM [25], and together they provide the
capability of four-point electrical transport measurement for nanostructured
materials [18, 24, 26].
74 A.-P. Li and T.Z. Ward

Fig. 4.4 (a) Photo of ORNL quadraprobe STM system: (A) a four-probe STM, (B) an SEM
column, (C) a liquid helium Dewar, (D) an electron energy analyzer for SAM, (E) a probe load-
lock chamber, (F) an MBE chamber, (G) a sample load-lock chamber, and (H) three air legs for
vibration suppression. (b) Photograph of four STM scanners and a sample holder on the cryostat
stage. (c) SEM image showing four STM probes located on a single CNT. Adapted from [18, 24]

4.3 Results and Discussions

4.3.1 Percolative Mott Transition in Sr3(Ru1  xMnx)2O7

The layered Sr3Ru2O7 belongs to the Ruddlesden–Popper series with a stacking of


two layers of cornersharing RuO6 octahedra separated by cation oxide rock-salt
layers [27]. The lamellar structure of the ruthenate makes the crystal amenable to
creating a fresh ab plane by cleaving insitu in an ultrahigh vacuum (UHV) system.
This offers a controlled way to investigate electronic phases using multiple surface
analytical techniques. In general, 4d TMOs (e.g., ruthenates) have weaker correla-
tion effects than 3d TMOs (e.g., manganites, cuprates, and cobaltates), as the radial
extension of 4d wave functions is significantly larger than that of transition-metal
(TM)-3d orbitals [28]. On the other hand, the extended wave functions of the 4d TM
enable an interesting competition between local and itinerant physics. Doping
the 4d host with a dilute 3d TM is thus extremely effective in tuning valence,
spin, and orbital characteristics and, in turn, the macroscopic physical properties as
manifested in Sr3(Ru1  xMnx)2O7 [28, 29]. In the undoped compound of Sr3Ru2O7,
a fieldinduced quantum phase transition was discovered [30, 31]. With the substi-
tution of Ru by a few percent of Mn, an MIT and the emergence of a Mott
antiferromagnetic (AF) ground state were revealed [28, 29]. The MIT is believed
to be driven by the onset of AF correlations with a clear phase boundary between
the paramagnetic (PM) metal and AF insulator regions at Tc ~ 80 K for 10% Mn
substitution [28, 29].
4 Emergent Metal–Insulator Transitions... 75

To study the Mott MIT in ruthenates, we substituted the Ru site with 20%
Mn and grew single crystals of Sr3(Ru0.8Mn0.2)2O7 using the floating zone tech-
nique [32]. Direct visualization of phase separation was carried out on cleaved
single-crystal surfaces in a cryogenic four-probe STM system [18]. An UHV
(pressure <1.0  1010 Torr) field emission SEM, scanning tunneling spectros-
copy (STS), and four-probe STM were used to probe electronic properties on
surfaces in situ.
Figure 4.5 shows the striking domain structures in a Sr3(Ru0.8Mn0.2)2O7 surface
at low temperatures revealed by SEM operated in the secondary electron emission
mode. At 81 K, micrometer-scale dark bubble- and stripe-like (Fig. 4.5a, b)
domains are clearly seen in an otherwise bright matrix of the cleaved surface.
However there is no domain structure at room temperature (Fig. 4.5c) or for an
undoped sample.
Contrast in the field-emission SEM image reflects the difference in the second-
ary electron (SE) yield which, in turn, is determined by topography, chemical
composition, and conductance of the sample [34, 35]. In particular, the differences
in conductance between the neighboring areas affect the work function, stopping
power, and electron mean free path, all intimately associated with the SE emission
process. In fact, there is a growing interest in utilizing the conductance contrast of
SEM to map electron density distributions, particularly in semiconductors [36].
Across the MIT point of Sr3(Ru1  xMnx)2O7, topographical and chemical contrasts
are not expected to show appreciable change. However, the formation of an energy
gap near the MIT point modifies the electron band structures relative to the
corresponding metallic phase. As the energy distribution of the SE is narrow and
peaked at very low energy, generally in the range of 2–5 eV [34, 35], even a small
shift in band structure can generate a remarkable change in SE yield and a clear
contrast at the boundary of the insulator and the metal phases.
The temperature (T)-dependent evolutions of stripe domains have been
displayed in Fig. 4.6. When warming up the sample from 50 K, the stripe domains
start to melt gradually with T and disappear completely at 150 K. Upon cooling
down to 50 K again, the stripes reappear. Compared to the domain pattern obtained
in the first thermal cycle, the new pattern shows a few new remarkable regions –
a new stripe developed, a stripe disappeared in the second thermal cycle, and some
stripes changed locations. Based on theoretical simulations, the chemical disorder
introduced by the impurity ions of different radii can form largescale phase
separations [37, 38]. The inhomogeneity of a chemical origin would occur at
more or less the same location although the shape may be different in different
thermal cycles. On the other hand, pure electronic interactions can also induce
inhomogeneity of charge density distribution; however, this inhomogeneity would
have totally random distribution at the nanometer scale [3, 38]. Thus, neither
electronic interactions nor chemical inhomogeneity can be used to fully explain
the above observations.
To investigate the origins of the bright/dark domain contrast in SEM images, we
have performed STM/STS measurements on cleaved surfaces. STS spectra were
taken on individual surface domains, which probes the integrated density of
76 A.-P. Li and T.Z. Ward

Fig. 4.5 Electronic phase


separations displayed in
the cleaved surfaces of
Sr3(Ru1  xMnx)2O7.
(a) Bubble- and (b) Stripe-
shaped domain structures
revealed by SEM for
x ¼ 0.20 at 81 K. (c) No
domain structures observed
at 300 K for x ¼ 0.20
sample or in an undoped
sample (x ¼ 0). Adapted
from [33]

electronic states near the Fermi energy, EF. The derivative of I(V) data (Fig. 4.7a)
reveals an energy gap near EF in the dark domains, in contrast to a finite density of
states in the bright regions. Therefore, the SEM contrast observed here reflects the
electronic state difference between dark and bright regions that correspond
4 Emergent Metal–Insulator Transitions... 77

Fig. 4.6 T-dependent phase percolations in a Sr3(Ru0.8Mn0.2)2O7. (a–e) Domain images


measured at various T for the same sample location. (f) Superimposed view of images (a) and
(e), showing in blue and purple, respectively. Adapted from [33]

to insulating and metallic phases, respectively. The Mott transition nature is


reflected in the abrupt opening of the energy gap measured by STS and shown in
Fig. 4.7b. At the microscopic scale, the onset of the MIT (namely Tc) varies from
location to location, consistent with the observation of the continuous transition
behavior in domain percolation and the bulk MIT.
Furthermore, the conductance difference between the bright and the dark regions
has been verified by microscopic transport measurements with the four STM
probes. Figure 4.8 shows the transport R(T) curves measured along the dark stripes,
while keeping the interprobe spacing constant (as indicated in the inset of Fig. 4.8).
The local MIT occurs near 135 K, below which R increases with decreasing T and
reaches more than five orders of magnitude higher resistance at 10 K than that for its
metallic state. Between 81 and 135 K, the single stripe turns into multiple discon-
nected stripes with smaller sizes, which vanish completely when reaching 135 K.
The percolative growth of the domains with T suggests that the phases are not
pinned. It has been theoretically argued that lattice distortions and long-range
strains can play an important role in TMOs [1, 39]. In the presence of strain,
induced either externally (stress) or internally (chemical substitution), the
electron–lattice interactions can lead to local energetically favorable configurations
and electronic PS over micrometer scales [1]. If strain interactions are the principal
cause of PS, then it should be possible to control the elastic energy landscape with
strain engineering methods [40].
Figure 4.9a–c shows the dynamical evolution of phase domains under uniaxial
compressive stress applied in the ab plane with an in situ piezoelectric sample
holder. The insulating domain appears brighter than the metallic matrix, inverted
from the contrast in Figs. 4.5 and 4.6. Such a contrast inversion is due to electron
78 A.-P. Li and T.Z. Ward

Fig. 4.7 STS reveals the a


conductance difference
between bright and dark 60
200

dl/dV (pA/V)
domains. (a) Derivative
conductance (dI/dV) curves
measured in bright and dark 100

Tunneling current I (pA)


bubble domains. (b) The 30
energy gap width as a 0
function of T for a −0.6 −0.3 0.0 0.3 0.6
semiconducting (dark) V (V)
domain measured by STS. 0
Adapted from [33]
dark domain
bright domain
−30

−0.6 −0.3 0.0 0.3 0.6


Sample voltage V (V)

b
Gap width (eV)

0.5

0.0

0 100 200 300


T (K)

Fig. 4.8 Local resistance


1010
change with temperature
measured in stripe domains. V
Inset: SEM image showing 109
the transport measurement
method with four-probe STM. 108
Resistance R (W)

Adapted from [33]


107 I
106

105 Local resistance

104

103
0 50 100 150
Temperature T(K)
4 Emergent Metal–Insulator Transitions... 79

Fig. 4.9 Strain-induced domain evolutions in the cleaved surface of a Sr3(Ru0.8Mn0.2)2O7 at 81 K.


(a) Domain image before stress. (b) After applying an uniaxial compressive stress in ab plane for
10 min and (c) 14 h. (d) Evolution of the insulating domain area with time under stress. Adapted
from [33]

charging in the insulating domains, not uncommon in electron microscopy. For a


metal, the energy barrier (work function) for secondary electron emission is closely
related to its Fermi energy. However, for an insulator, charge carriers can be
trapped in the surface area, which can severely shift surface bands and thus change
the work function and even lead to contrast inversion. Such a contrast inversion
corroborates the coexistence of metal and insulator phases near the Mott transition.
The domain area has expanded significantly after stress, growing two dimen-
sionally in the ab plane with a slight overgrowth along the stress direction
(Fig. 4.9b). The stripe shapes of domains displayed in Fig. 4.5 may thus reflect
the local distributions of residual strain in the cleaved surface. The magnitude of
strain, deduced from the displacement of the piezoelectric holder, is extremely
small (105); however, the domain area has more than doubled as shown in
Fig. 4.9c. The enormous response of the phase domains to the strain field in
Sr3(Ru0.8Mn0.2)2O7 can stem from the interplay between the spatially confined
Mn 3d orbitals and the extended, yet anisotropic, Ru 4d–O 2p electronic backbone
of the Ru–O host. Undoped Sr3Ru2O7 is a tetragonally distorted system with RuO6
octahedra rotated about the c-axis [41]. A changeover from Fermi liquid metal to
80 A.-P. Li and T.Z. Ward

AF has been reported below 70 K under large hydrostatic pressure (~1 GPa) in
Sr3Ru2O7 [42]. Owing to a smaller ionic radius, the Mn substitution for Ru creates
an octahedral compression [29], which alters the relative energies of d orbitals and
enhances AF with a possibly orbitally ordered state at low T [28]. A minute amount
of external strain superimposed onto the distortions induced by Mn substitution has
thus resulted in a gigantic phase transition response in Sr3(Ru0.8Mn0.2)2O7.
Strain-induced dynamical phase evolution exhibits a long time constant. As
shown in Fig. 4.9d, the domain expands slowly over a period of hours for the
crystal kept at constant T in a strain field, showing a nearly logarithmic time-
dependent relaxation effect. This strain-field relaxation is reminiscent of diffuse
phase transitions reported in Cr-doped Nd1/2Ca1/2MnO3 [43] and PbMg1/3Nb2/3O3
[44], dubbed as relaxor ferromagnet and relaxor ferroelectric, respectively.
The slow relaxation in these systems was explained using the random field model
based on the stability consideration of the ordered state under a quenched random
field conjugate to an order parameter with a continuous symmetry [45]. In Cr-doped
Nd1/2Ca1/2MnO3, an orbital deficiency on Cr impurity sites provides the quenched
random field that destroys the orbital ordering and produces the FM clusters [43].
In PbMg1/3Nb2/3O3 nanoscale ferroelectric domains embedded in a paraelectric
matrix are sources of quenched random electric fields [43, 44]. There is a strong
parallel between these relaxor ferroics and Sr3(Ru0.8Mn0.2)2O7. Here, the Mn
impurities introduce a local strain field, and the quenched disorder leads to PS
near the Mott transition. In this scenario, the conjugate order parameter is the orbital
orientation, which is consistent with the previous suggestions that the AF low-T
phase has a long-range orbital order [28, 29].
In fact, electronic inhomogeneities are believed to be a common feature in
correlated TMOs and nanoscale inhomogeneities have widely been observed in
manganites [46], ruthenates [47], cuprates [48], and VO2 [49]. However, microme-
ter scale PS has only been shown in real space in manganites that were tuned by
heating and magnetic field [50, 51]. To the best of our knowledge, we observe, for
the first time, the real space PS and the phase percolation in a ruthenate 
Sr3(Ru0.8Mn0.2)2O7. Furthermore, a rare glimpse of the electronic nature of the
individual phase domains has now been provided in the vicinity of the MIT that
enabled establishment of the correlation of global functionality with the micro-
scopic phase percolations. The classic T-induced MIT is demonstrated in real space
to be tunable via a strain field, an extraordinary manifestation of the strong
correlations of electron and lattice degrees of freedom.

4.3.2 Confinement Effects and Tunable Emergent Behavior


in La5/8xPrxCa3/8MnO3

As discussed above, electronic phase separation is thought to play an important role


in CMR, the MIT and high-TC superconductivity [5]. By spatially confining com-
plex materials to length scales smaller than the electronic phase domains that reside
4 Emergent Metal–Insulator Transitions... 81

Fig. 4.10 Comparison of electronically phase-separated (PS) material to corresponding resistor


networks. PS film behaves as a parallel resistor network. Confining the same film creates a serial
resistor network where single domains can dominate the transport signal. Adapted from [54]

within them it is possible to access previously hidden electronic phases using


simple transport measurements [52, 53]. Conceptually, a PS unconfined film can
be thought of as a large parallel resistor network composed of two types of resistors
with different values (Fig. 4.10). When confined to the same scale as the electronic
phases in the material, the system can be treated as a serial resistor network.
A single domain’s transport signal in the parallel system is unseen as other transport
channels are open to current flow. However, in the serial network, a single domain’s
signal can have a huge effect on the observed signal.
Using this technique, behaviors that were previously hidden to transport
measurements have been observed. One such discovery showed that multiple
MITs in certain manganite materials can reside in a single crystal [54]. This
behavior was attributed to balanced strain contributions arising from lattice and
chemical pressures. Of particular importance was that a desired property in a
complex system could be selectively tuned (Fig. 4.11) through doping and substrate
selection. While the reemergent behavior was present in the film geometry, only
through confinement could it be observed. These findings further hinted at the key
role that strain plays in the creation of electronic phase separation in manganites.
Indeed, these findings had an immediate impact by pointing out that slight strain
differences could influence macroscale phase separation. By tuning the energetic
landscape through anisotropic strain fields, it was shown that electronic phase
domains could be forced to preferentially seed with an elongated aspect ratio
[55]. This was accomplished by epitaxially locking a pseudocubic perovskite single
crystal thin film of La5/8  xPrxCa3/8MnO3 (x ¼ 0.3) to an orthorhombic NdGaO3
substrate. This created a strain environment in which the two in-plane axes were
held under differing strains. Simultaneous temperature-dependent resistivity
measurements along the two perpendicular in-plane axes exhibited large
differences in the MIT temperatures and extraordinarily high anisotropic resistive
behaviors on macroscales. The mechanism for these differences was attributed to
metallic domain formation along the strained axis; in effect, the metallic domain
82 A.-P. Li and T.Z. Ward

Fig. 4.11 Effects of strain tuning on reemergent metal–insulator transitions in La5/8  xPrxCa3/8
MnO3 (LPCMO) 10 mm wide wires under increasing magnetic field (black for lowest field, blue for
highest field). Red spheres in perovskite crystal structure denote a-site. Red boxes denote strain
configurations where reemergent transition is observable. Adapted from [54]

was confined by the strain field along one direction which opened transport
channels much more readily. This lead to in-plane resistivity differences of
20,000% and a shift in the MIT temperature of 25 K. These findings confirmed
what was hinted at from the early confinement work on strain balancing by showing
that PS domain formation at the macroscale could be guided through strain. Strain-
controlled domain formation promises new tunable device applications while
answering fundamental questions on the role of electronic phase separation in
manganites.
The control of electronic phase separation will be vital to its application in future
emergent electronic devices, but to actually study the dynamics at play within these
systems, it is again useful to turn to spatial confinement techniques to isolate single
PS domains. The dynamics of a first-order electronic phase transition in a complex
material is not well understood. However this understanding is crucial if we are to
find a comprehensive model for the emergent phenomena of electronic phase
separation.
By reducing a model LPCMO system to the scale of its inherent electronic
charge-ordered insulating and ferromagnetic metal phase domains, it is possible to
directly observe single electronic phase domain fluctuations by balancing at the
MIT temperature (TMIT) where single domains can dominate the transport signal.
4 Emergent Metal–Insulator Transitions... 83

Fig. 4.12 (a) Wire shows abrupt drop in resistivity at the metal–insulator transition while the film
shows a smooth transition (inset). (b) Resistivity of wire when held at the transition temperature
shows clear jumps associated with single electronic domain fluctuations; this behavior is not
observed in the film (inset). Taken from [56]

Using this method, several key pieces to the puzzle of how the MIT manifests at a
tipping point were found in manganites [56, 57]. While previous experimental work
on bulk and thin film samples gave no indication that the MIT was anything other
than a smooth process across a wide temperature window, confinement studies
showed that this transition actually comprised many individual domains indepen-
dently flipping phase across a narrow temperature window (Fig. 4.12). This finding
was accompanied by the first observation of domain fluctuation lifetimes and the
realization that the domains themselves are correlated with other domains within
the material. While answering several important questions about the PS domain
dynamics, this work also offered proof that spatial confinement can glean new
information from well-studied systems and suggests that this technique will work
on any system where length scale-dominated behaviors play a key role in
influencing electronic properties.

4.4 Conclusion

In this chapter, we have reviewed our recent reports on the emergent MITs
associated with electronic phase separation and phase domain percolations in
functional metal oxides including a Mott transition in a single-crystalline
ruthenate surface and emergent conduction behaviors in manganite nanowires.
Utilizing a newly developed low temperature four-probe STM system, we imaged
the microscopic phase separation using a scanning electron microscope while
simultaneously interrogating the electronic properties of each domain using the
scanning tunneling probes in the spectroscopic or transport modes. We observed
dramatic changes in the phase percolations in response to thermal cycling and
mechanical stress. A quantitative correlation between the global Mott transition
and the microscopic phase percolation has been revealed, which clearly indicates
84 A.-P. Li and T.Z. Ward

that the functionality of the MIT originates from the phase competition at the
microscopic level. Strikingly, we discovered that the phase separation was not
associated with inhomogeneous doping and that the MIT can be controlled by an
external uniaxial stress. Moreover, by reducing material systems with inherent
correlation lengths across a spectrum of lengths, we found it possible to use robust
transport probing methods to gather information on dynamics and degrees of
electronic interaction. In effect, these techniques allow the experimenter to
increase the resolution of electronic transport measurements by creating a system
in which multiple regions of varying character can dominate a transport signal.
The methods discussed in this chapter can be of immense value to the study of any
material system having coexisting intrinsic length scales from nanometers
to micrometers.

Acknowledgments This research was conducted at the Center for Nanophase Materials Sciences,
which is sponsored at Oak Ridge National Laboratory by the Division of Scientific User Facilities
(APL) and the Division of Materials Sciences and Engineering (TZW), Office of Basic Energy
Sciences, U.S. Department of Energy.

References

1. Ahn, K.H., Lookman, T., Bishop, A.R.: Strain-induced metal-insulator phase coexistence in
perovskite manganites. Nature 428, 401–404 (2004)
2. Sarma, D.D., et al.: Direct observation of large electronic domains with memory effect in
doped manganites. Phys. Rev. Lett. 93, 097202 (2004)
3. Dagotto, E., Hotta, T., Moreo, A.: Colossal magnetoresistant materials: the key role of phase
separation. Phys. Rep. 344, 1–3 (2001)
4. Moreo, A., Mayr, M., Feiguin, A., Yunoki, S., Dagotto, E.: Giant cluster coexistence in doped
manganites and other compounds. Phys. Rev. Lett. 84, 5568–5571 (2000)
5. Shenoy, V.B., Sarma, D.D., Rao, C.N.R.: Electronic phase separation in correlated oxides: the
phenomenon, its present status and future prospects. Chemphyschem 7, 2053–2059 (2006)
6. Du, C.H., et al.: Critical fluctuations and quenched disordered two-dimensional charge stripes
in La5/3Sr1/3NiO4. Phys. Rev. Lett. 84, 3911–3914 (2000)
7. Yamada, K., et al.: Doping dependence of the spatially modulated dynamical spin correlations
and the superconducting-transition temperature in La2xSrxCuO4. Phys. Rev. B 57,
6165–6172 (1998)
8. Zimmermann, M.V., et al.: Interplay between charge, orbital, and magnetic order in
Pr1xCaxMnO3. Phys. Rev. Lett. 83, 4872–4875 (1999)
9. Tseng, A.A., Notargiacomo, A., Chen, T.P.: Nanofabrication by scanning probe microscope
lithography: a review. J. Vac. Sci. Technol. B 23, 877–894 (2005)
10. Hamada, M., Eguchi, T., Akiyama, K., Hasegawa, Y.: Nanoscale lithography with frequency-
modulation atomic force microscopy. Rev. Sci. Instrum. 79, 123706 (2008)
11. Sarkar, T., Ghosh, B., Raychaudhuri, A.K., Chatterji, T.: Crystal structure and physical
properties of half-doped manganite nanocrystals of less than 100-nm size. Phys. Rev. B 77,
235112 (2008)
12. Shankar, K., Raychaudhuri, A.K.: Low-temperature polymer precursor-based synthesis of
nanocrystalline particles of lanthanum calcium manganese oxide (La0.67Ca0.33MnO3) with
enhanced ferromagnetic transition temperature. J. Mater. Res. 21, 27–33 (2006)
4 Emergent Metal–Insulator Transitions... 85

13. Petit, D., Faulkner, C.C., Johnstone, S., Wood, D., Cowburn, R.P.: Nanometer scale patterning
using focused ion beam milling. Rev. Sci. Instrum. 76, 026105 (2005)
14. Pallecchi, I., et al.: Investigation of FIB irradiation damage in La0.7Sr0.3MnO3 thin films.
J. Magn. Magn. Mater. 320, 1945–1951 (2008)
15. Grigorescu, A.E., Hagen, C.W.: Resists for sub-20-nm electron beam lithography with a focus
on HSQ: state of the art. Nanotechnology 20, 292001 (2009)
16. Rhee, H.-G., Kim, D.-I., Lee, Y.-W.: Realization and performance evaluation of high speed
autofocusing for direct laser lithography. Rev. Sci. Instrum. 80, 073103 (2009)
17. Sun, Y., et al.: From tunneling to point contact: correlation between forces and current. Phys.
Rev. B 71, 193407 (2005)
18. Kim, T.-H., et al.: A cryogenic Quadraprobe scanning tunneling microscope system with
fabrication capability for nanotransport research. Rev. Sci. Instrum. 78, 123701 (2007)
19. Grube, H., Harrison, B.C., Jia, J., Boland, J.J.: Stability, resolution, and tip-tip imaging by a
dual-probe scanning tunneling microscope. Rev. Sci. Instrum. 72, 4388–4392 (2001)
20. Guise, O., et al.: Development and performance of the nanoworkbench: a four tip STM for
conductivity measurements down to submicrometer scales. Rev. Sci. Instrum. 76, 045107
(2005)
21. Hansen, T.M., et al.: Resolution enhancement of scanning four-point-probe measurements on
two-dimensional systems. Rev. Sci. Instrum. 74, 3701–3708 (2003)
22. Hobara, R., et al.: Variable-temperature independently driven four-tip scanning tunneling
microscope. Rev. Sci. Instrum. 78, 053705 (2007)
23. Tsukamoto, S., Siu, B., Nakagiri, N.: Twin-probe scanning tunneling microscope. Rev. Sci.
Instrum. 62, 1767–1771 (1991)
24. Kim, T.-H., Wendelken, J.F., Li, A.P., Du, G.H., Li, W.Z.: Probing electrical transport in
individual carbon nanotubes and junctions. Nanotechnology 19, 485201 (2008)
25. Zeng, C.G., Kent, P.R.C., Kim, T.-H., Li, A.P., Weitering, H.H.: Charge-order fluctuations in
one-dimensional silicides. Nat. Mater. 7, 539–542 (2008)
26. Kim, T.-H., et al.: Probing microscopic variations of superconductivity on the surface of Ba
(Fe1xCox)2As2 single crystals. Phys. Rev. B 80, 214518 (2009)
27. Ruddlesden, S.N., Popper, P.: The compound Sr3Ti2O7 and its structure. Acta Cryst. 11, 54–55
(1958)
28. Hossain, M.A., et al.: Crystal-field level inversion in lightly Mn-doped Sr3Ru2O7. Phys. Rev.
Lett. 101, 016404 (2008)
29. Mathieu, R., et al.: Impurity-induced transition to a Mott insulator in Sr3Ru2O7. Phys. Rev.
B 72, 092404 (2005)
30. Grigera, S.A., et al.: Magnetic field-tuned quantum criticality in the metallic ruthenate
Sr3Ru2O7. Science 294, 329–332 (2001)
31. Perry, R.S., et al.: Metamagnetism and critical fluctuations in high quality single crystals of the
bilayer ruthenate Sr3Ru2O7. Phys. Rev. Lett. 86, 2661–2664 (2001)
32. Stone, M.B., et al.: Temperature-dependent bilayer ferromagnetism in Sr3Ru2O7. Phys. Rev. B
73, 174426 (2006)
33. Kim, T.-H., et al.: Imaging and manipulation of the competing electronic phases near the Mott
metal-insulator transition. Proc. Natl Acad. Sci. USA 107, 5272–5275 (2010)
34. Joseph Goldstein, D.E.N., Echlin, P., Lyman, C.E., Joy, D.C., Lifshin, E., Sawyer, L.C.,
Michael, J.R.: Scanning Electron Microscopy and X-Ray Microanalysis. Kluwer, New York
(2003)
35. Reimer, L.: Scanning Electron Microscopy: Physics of Image Formation and Microanalysis.
Springer, Berlin (1985)
36. Castell, M.R., Perovic, D.D., Lafontaine, H.: Electronic contribution to secondary electron
compositional contrast in the scanning electron microscope. Ultramicroscopy 69, 279–287
(1997)
37. Dagotto, E.: Complexity in strongly correlated electronic systems. Science 309, 257–262
(2005)
86 A.-P. Li and T.Z. Ward

38. Moreo, A., Yunoki, S., Dagotto, E.: Solid state physics – phase separation scenario for
manganese oxides and related materials. Science 283, 2034–2040 (1999)
39. Millis, A.J.: Lattice effects in magnetoresistive manganese perovskites. Nature 392, 147–150
(1998)
40. Ikeda, S.-I., Maeno, Y., Nakatsuji, S., Kosaka, M., Uwatoko, Y.: Ground state in Sr3Ru2O7:
Fermi liquid close to a ferromagnetic instability. Phys. Rev. B 62, R6089–R6092 (2000)
41. Shaked, H., Jorgensen, J.D., Chmaissem, O., Ikeda, S., Maeno, Y.: Neutron diffraction study of
the structural distortions in Sr3Ru2O7. J. Solid State Chem. 154, 361–367 (2000)
42. Sushko, Y.V., et al.: Hydrostatic pressure effects on the magnetic susceptibility of ruthenium
oxide Sr3Ru2O7: evidence for pressure-enhanced antiferromagnetic instability. Solid State
Commun. 130, 341–346 (2004)
43. Kimura, T., Tomioka, Y., Kumai, R., Okimoto, Y., Tokura, Y.: Diffuse phase transition and
phase separation in Cr-doped Nd1/2Ca1/2MnO3: a relaxor ferromagnet. Phys. Rev. Lett. 83,
3940–3943 (1999)
44. Westphal, V., Kleemann, W., Glinchuk, M.D.: Diffuse phase-transitions and random-field-
induced domain states of the relaxor ferroelectric PbMg1/3Nb2/3O3. Phys. Rev. Lett. 68,
847–850 (1992)
45. Imry, Y., Ma, S.-K.: Random-field instability of the ordered state of continuous symmetry.
Phys. Rev. Lett. 35, 1399–1401 (1975)
46. Ma, J.X., Gillaspie, D.T., Plummer, E.W., Shen, J.: Visualization of localized holes in
manganite thin films with atomic resolution. Phys. Rev. Lett. 95, 237210 (2005)
47. Zhang, J.D., et al.: Dopant-induced nanoscale electronic inhomogeneities in Ca2xSrxRuO4.
Phys. Rev. Lett. 96, 066401 (2006)
48. Hanaguri, T., et al.: A ‘checkerboard’ electronic crystal state in lightly hole-doped
Ca2xNaxCuO2Cl2. Nature 430, 1001–1005 (2004)
49. Qazilbash, M.M., et al.: Mott transition in VO2 revealed by infrared spectroscopy and nano-
imaging. Science 318, 1750–1753 (2007)
50. Uehara, M., Mori, S., Chen, C.H., Cheong, S.W.: Percolative phase separation underlies
colossal magnetoresistance in mixed-valent manganites. Nature 399, 560–563 (1999)
51. Zhang, L., Israel, C., Biswas, A., Greene, R.L., De Lozanne, A.: Direct observation of
percolation in a manganite thin film. Science 298, 805–807 (2002)
52. Wu, T., Mitchell, J.F.: Creation and annihilation of conducting filaments in mesoscopic
manganite structures. Phys. Rev. B 74, 214423 (2006)
53. Zhai, H.-Y., et al.: Giant discrete steps in metal-insulator transition in perovskite manganite
wires. Phys. Rev. Lett. 97, 167201 (2006)
54. Ward, T.Z., et al.: Reemergent metal-insulator transitions in manganites exposed with spatial
confinement. Phys. Rev. Lett. 100, 247204 (2008)
55. Ward, T.Z., et al.: Elastically driven anisotropic percolation in electronic phase-separated
manganites. Nat. Phys. 5, 885–888 (2009)
56. Ward, T.Z., et al.: Time-resolved electronic phase transitions in manganites. Phys. Rev. Lett.
102, 087201 (2009)
57. Ward, T.Z., et al.: Dynamics of a first order electronic phase transition in manganites. Phys.
Rev. B 83, 125125 (2011)
Chapter 5
Optical Properties of Nanoscale Transition
Metal Oxides

Janice L. Musfeldt

5.1 Physical, Chemical, and Size–Shape Tunability


in Transition Metal Oxides

The interplay between charge, structure, and magnetism is the origin of rich
physics in complex oxides. Because these interactions are so strong, oxides are
“on the knife’s edge,” straddling several competing regions of phase space [1–8].
These competing interactions give rise to very rich H-T-P phase diagrams, often
with exotic electronic and magnetic ground states that are the result of delicately
balanced charge-spin-lattice coupling. Signatures of this coupling include, for
instance, inhomogeneous spin and charge texture and coupled excitations like
electromagnons, magnon sidebands, and electron–phonon interactions [9–14]. An
important consequence of the aforementioned phase proximity is the physical
property tunability. Small external perturbations can change important energy and
length scales, driving the system into a new state with very different properties.
Here, the Goodenough–Kanamori–Anderson rules are often used to understand
the influence of charge, structure, and orbital overlap on magnetism [15]. Occa-
sionally, one discovers the opportunity to drive completely new functionality by
tuning across phase boundaries. For instance, temperature can induce
metal–insulator, charge and magnetic ordering, or superconducting transitions
[16–24]. Likewise, high magnetic fields, pressure, and electric fields can be used
to drive superconducting, density wave, ferromagnetic, and other exotic magnetic
transitions [25–38].
Chemical substitution is another powerful method for tuning the properties of a
material. The copper oxide superconductors are a classic example [21, 23]. In these

J.L. Musfeldt (*)


University of Tennessee, Knoxville, TN, USA
e-mail: musfeldt@ion.chem.utk.edu

J. Wu et al. (eds.), Functional Metal Oxide Nanostructures, Springer Series 87


in Materials Science 149, DOI 10.1007/978-1-4419-9931-3_5,
# Springer Science+Business Media, LLC 2012
88 J.L. Musfeldt

doped antiferromagnets, oxygen stoichiometry controls the carrier concentration.1


Chemical substitution also gives rise to exotic phase diagrams in the manganites
[3, 9]. Here, rare earth doping drives the system through a series of ground states
whose properties depend upon the charge (and spin) of the Mn centers and coupling
to the lattice. Chemical substitution can also be used to manipulate the properties of
ferroelectric perovskites such as Pb(Zr1  xTix)O3 [39–41]. According to recent
theoretical predictions, a magnetic ion placed on the perovskite “A site” may activate
ferroelectrically induced ferromagnetism [42, 43]. The ability of chemical substitu-
tion to manipulate properties is, of course, not limited to oxides or extended solids.
For instance, molecular hydrogen-bonded systems [44–52] and molecular magnets
[53–61] employ subtle chemical substitutions to control architecture, bonding, and
reactivity, often with the use of electron donating or withdrawing substituents as
structure-directing agents. Photomagnetic metal-cyanide compounds [62–71] also
display rich chemical tunability.
The discovery that complex solids can form nanoscale objects provides an
exciting opportunity to investigate bulk vs. nanoscale chemistry using molecular-
level strain as the tuning parameter [72–77]. At very small length scales, surface
states and strain, defect states, local structure, local composition, and finite size
effects dominate the properties. Some of the beautiful, flexible, and functional
nanomorphologies include tubes, wires, octahedra, particles, tetrahedra, and
spheres [78–113]. These systems demonstrate the promise of molecular-level
control of size and shape, and hold out the potential for control of confinement
and crossover related effects. Flagship materials include organics such as carbon
nanotubes and graphene, and inorganics such as gold, CdSe, ZnO, PtFe, and MoS2
[76, 78, 81, 89, 92, 107, 108, 114–116]. Here, quantum confinement influences
transport, magnetic, optical, and mechanical properties [78, 104, 105, 117–122].
Interestingly, inorganic nanomaterials have many of the same scientific challenges
as their bulk counterparts with regard to superconductivity and magnetism, mixing
of spin-lattice-charge-orbital degrees of freedom, unusual phase transitions, micro
and nanoscopic “texture,” and strain [123–129]. As a consequence of intrinsic
physical and chemical interactions, competition between different states in nano-
scale solids remains an area with challenging scientific opportunities [130].
The goal of this review is to illustrate how optical spectroscopy can reveal
fundamental aspects of nanoscale transition metal oxides. Just as physical and
chemical tuning are at the heart of our ability to design, understand, and control
advanced materials in their bulk form, size–shape tuning is of emerging importance
in the field of functional oxide nanomaterials. This is because it (1) expands the
usable structure–property phase space and (2) it can drive emergent phenomena.
And in the same way that optical spectroscopy reveals the dynamics of bulk
materials, it is poised to do so in nanoscale compounds. I selected several physical
systems that illustrate these opportunities, placing an emphasis on chemically

1
In addition to controlling carrier concentration, oxygen vacancies are also related to domain wall
pinning, leakage currents, degradation, resistances switching behavior in thin film samples.
5 Optical Properties of Nanoscale Transition Metal Oxides 89

simple materials where both bulk and nanoscale properties have been investigated.
In making these selections, I carefully considered the definition of a nanomaterial
and elected to take a broad view encompassing nanoscale texture in bulk materials,
thin films oxides, and more traditional nanoscale objects such as particles and tubes,
as well as chemical aspects of binding in porous materials. In addition to advancing
the fundamental understanding of novel materials, basic research on the optical
properties of nanomaterials can lead to practical applications [65, 131–155] partic-
ularly in energy-related technologies.

5.2 Optical Spectroscopy as a Probe of Complex Oxides

Optical spectroscopy is a sensitive microscopic probe of various physical phenom-


ena in complex solids [156, 157]. A common target of an infrared investigation is
the dipole-allowed vibrational modes, typically characterized as intramolecular and
intermolecular features for a molecular material, or collective phonons and lattice
modes for a more extended solid [158–160]. Vibrational spectroscopy is an impor-
tant complement to neutron scattering for the investigation of complex materials.
Molecular group theory, correlation group, and double group methods are
employed to assess symmetry properties of a material [161, 162]. There is a natural
connection between vibrational spectra and molecular dynamics theory because
such calculations provide resonance frequencies, displacement patterns, and sym-
metry [106, 163–166]. From a combined analysis of vibrational mode trends and
displacement patterns, it is possible to assess local lattice distortions, the local
chemical environment, and charge ordering patterns [160, 167–170]. Rattling, soft,
and collective modes are other types of low-energy excitations and can contribute
to ionic conductivity, negative thermal expansion, thermoelectric response, density
wave and ferroelectric transitions, and superconductivity [137, 171–173].
Electronic effects can also be investigated by optical techniques. These include the
mobile carrier response, localization, confinement effects, charge ordering/dispropor-
tionation, dimensional behavior (such as van Hove singularities and Luttinger liquid
behavior), charge transfer, color properties and crystal field splitting, effective mass,
oscillator strength, and the nature of the optical gap, just to name a few [106, 163,
170, 174–195]. Many of these effects occur within 200 meV (~1,600 cm1) of the
Fermi surface, making them sensitive to physical and chemical modifications. For
certain features, electronic structure calculations (and partial densities of states
information) offer insight into the microscopic origins of observed electronic
excitations, gaps, and impurity levels; calculated geometry and magnetic properties
are also of interest [100, 101, 103, 106, 196–209].
A variety of interaction mechanisms lift traditional spectroscopic selection rules.
For instance, vibronic coupling connects many well-known degrees of freedom
with appropriate symmetry in molecular solids [18, 210], activates crystal field
excitations [195], and gives nanoscale texture in superconducting cuprates [211].
Microscopic strain, local structural modifications, intermolecular interactions and
90 J.L. Musfeldt

Table 5.1 Characteristic energy scales probed by optical spectroscopy compared with typical
temperature, magnetic field, pressure, and electric field scales
Low energy Medium energy High energy
30 cm1 (1 THz) 100–3,000 cm1 0.5–20 eV
(12 meV–0.4 eV)
10 K 100–1,000 K 104–105 K
1T 50 T 1,000 T
1 kbar 25 kbar 1 Mbar
1 mV/cm 100 mV/cm–1 V/cm 1,000 kV/cm
Magnetic excitations, lattice Vibrational modes, low energy Localized electronic, charge
modes, free carrier processes charge transfer excitations transfer excitations

bonding, and the charge environment each affects the vibrational response via local
symmetry considerations [26, 120, 161, 212, 213]. Low-energy magnetic
excitations can also be spectroscopically probed due to the lifting of traditional
selection rules [214–217]. For instance, spin-orbit coupling can activate the singlet-
triplet gap [216, 218–220], and spin-phonon coupling has been of contemporary
interest in molecular magnets [221, 222] and frustrated systems [223].
Clearly, optical spectroscopy is a flexible and microscopic probe of the disper-
sive and lossy response of a material [156, 157]. Considering the developing
evidence for nanoscale analogs of most functional oxides, optical spectroscopy of
these systems is a rich and wide-open area of research, and it uniquely provides
microscopic information that complements various bulk properties studies where
different length and time scales are important. Combining optical spectroscopy
with physical and chemical tuning techniques makes it even more powerful [224].
Characteristic energy scales for temperature, magnetic field, pressure, and electric
field effects are summarized in Table 5.1. It is clearly of great interest to evaluate
how nanoscale strain compares (and competes) with characteristic temperature,
magnetic field, pressure, and electric field ranges.

5.3 Quantitative Models

Optical spectroscopy is a technique that is blessed with an excellent fundamental


connection to theory. This is because the matrix element of an allowed excitation is,
at its heart, simply a measurement of the joint density of states with a particular
moment operator. As a result, traditional electronic structure theory, lattice dynam-
ics calculations, and group theory are excellent tools with which to understand the
electronic excitations, vibrational properties, symmetry, and selection rules. These
approaches work well in polar oxide thin films and carbon nanotubes where
periodic boundary conditions, crystal symmetries, and helical symmetries can be
invoked. That said, one is often presented with issues (e.g., a size-dependent band
5 Optical Properties of Nanoscale Transition Metal Oxides 91

Fig. 5.1 Sliding scale of confinement for a spherical electron-hole system that depends upon the
radius of the quantum well (R) and the exciton Bohr radius (aB ) [226]. When R=aB  1,
confinement is weak and the character of the exciton remains quasiparticle-like. The exciton
loses its quasiparticle character in the strong confinement regime where R=aB  1

gap, the need to account for inhomogeneities, or the need to compare the wave-
length of light with the length scale of a nanostructure) that do not regularly arise
with a single crystalline sample. Cases where local structure and symmetry are
important (e.g., at an interface or at the uncompensated edge of a nanoparticle) are
also less amenable to standard approaches, although they still represent a reason-
able place to start an analysis. Sometimes, quantitative methods appropriate to bulk
structures completely fail to describe the relevant science in nanomaterials. In this
case, other models must be sought to quantify finite length scale effects. Several
prominent approaches are discussed below.

5.3.1 Confinement Models

One example of a framework that was developed to capture quantum size effects is
the confinement model. These models take on a variety of forms depending upon
their intended application, as described in Yoffe’s excellent review article [225].
All involve foundational assumptions regarding particle size and the characteristic
size of the exciton. Kayanuma [226] defines three different regimes that depend
upon the R=aB ratio (Fig. 5.1). Here, R is the radius of the quantum well (or particle)
and aB is the excitation Bohr radius. This is a sliding scale that depends essentially
upon the chemical nature of the material (which determines the exciton Bohr
radius) and the characteristic size of the nanoscale object (which depends upon
synthetic techniques). From the point of view of dramatically different or emergent
properties, the strong confinement regime is the most interesting.
Brus presents a useful analytical expression for the first excited electronic state
of a nanoparticle [227] that is relevant to the strong confinement regime in Fig. 5.1.
This model is based on the effective mass approximation and was extended by
Kayanuma [226] and others [228–231]. Here,
 
p2 h2 1 1 1:8e2
Egap;nano ¼ Egap;bulk þ þ   0:248ERyd : (5.1)
2R2 me mh ER
92 J.L. Musfeldt

5.0
APPROACH OF CLUSTER LOWEST EXCITED ELECTRIC
4.5 STATE TO THE BULK BAND GAP

4.0 CLUSTER

3.5 ZnO
BAND GAP
3.0
ENERGY (eV)

2.5 cds

2.0

1.5 GaAs

1.0

0.5
InSb

0
30 50 100 200 300 500
°)
DIAMETER ( A

Fig. 5.2 Calculated energy of the cluster lowest excited electronic state in relation to the bulk
band gap [227]

The second particle-in-a-box-like term in this expression describes quantum locali-


zation effects and acts to increase the energy of the first excited state. It has the
practical effect of blue-shifting electronic excitations from their position in the bulk
material. The third term describes Coulomb interactions and acts to shift Egap,nano to
lower energy as R1 . Because of the competition between these terms, the apparent
band gap will always increase for small R but will show variations at larger R
depending on the importance of various terms. The fourth term describes binding
energy effects and acts to decrease the apparent gap when ERyd is large. Egap,nano is
predicted to asymptotically approach Egap,bulk with increasing particle size
(Fig. 5.2) [227], an effect that has been observed in a wide variety of materials
[225]. TiO2 and ZnO nanomaterials provide a classic illustration of spectral blue
shifts due to quantum confinement and are discussed in later sections of this review
[232, 233]. CdO and Bi2O3 nanoparticles display similar trends [234, 235].
Although these and other materials display spectral blue shifts due to confinement,
it is worth noting that there are several emerging cases where opposite effects are
observed. Another interesting prediction of these models is a confinement-induced
redistribution of oscillator strength [226, 231]. This modification of the optical
properties derives from the breakdown of translational invariance. In the exciton
confinement regime (large R=aB ), oscillator strength grows as R3, whereas
5 Optical Properties of Nanoscale Transition Metal Oxides 93

oscillator strength asymptotically approaches an enhanced value in the individual


particle confinement regime [226].
Confinement effects also manifest themselves in phase transition behavior. This
is because phase transitions are intrinsically long-range, cooperative phenomena
that are governed by an appropriate correlation length [236]. Interesting things
happen when sample size is tuned through the length scale of the important
correlation. For instance, in triglycine sulfate, the ferroelectric polarization and
Curie temperature are reduced with decreasing film thickness as DT ¼ Tc(1) 
Tc(D) ~ 1/D where D is the film thickness [237, 238]. The superconducting transi-
tion temperature depends on thickness in a-MoGe thin films as well [239]. In model
ferroelectrics like PbTiO3 and BaTiO3, X-ray diffraction and Raman scattering
were combined to correlate the reduced ferroelectric transition temperature vs.
nanoparticle size with soft mode trends and local structure distortions [240–242].
In PbTiO3, the E(1TO) soft mode frequency decreases with particle size as ~1/
(D – Dcrit), where D is particle diameter and Dcrit is the critical particle diameter
[240, 242]. Even in the absence of microscopic models for these trends, there are
clearly many opportunities to understand phonon confinement and its relation to
phase transitions in functional oxide nanomaterials, particularly in systems where
the particle size distribution is under good control.

5.3.2 Descriptions of Inhomogeneous Media

The spectroscopic response of inhomogeneous media is an area that benefits from


the well-developed Maxwell-Garnett and Bruggeman Effective Medium Approxi-
mation models [243–245]. These models are traditionally applied to mixtures and
compounds, as described by Carr et al. in their excellent 1985 review article [246].
They are now finding application in nanoscale materials. The Maxwell-Garnett and
Bruggeman Effective Medium Approximation models are both based upon the
assumption that particle (or domain) size is much larger than the wavelength of
light used in the spectroscopic analysis. This approach allows us to consider
inhomogeneous media to be homogeneous on the length scale of the light and to
possess a well-defined effective dielectric response function. The models differ
mainly in the way in which the medium surrounding the grain under consideration
is treated whether it is surrounded by one of the mixture constituents or by a
medium characterized by averaged properties [246]. The Maxwell-Garnett model
is most appropriate for dilute systems, where the gains of the minority constituent
are well separated [243, 244]. We can write the effective dielectric response,
typically labeled EMGT as

3f ðEa  Eb Þ
EMGT ¼ Eb þ Eb ; (5.2)
ð1  f ÞðEa  Eb Þ þ 3Eb
94 J.L. Musfeldt

where Ea is the dielectric function of the grain in question, Eb is the dielectric


function of the surrounding medium, and f is the filling factor. Two issues that
immediately arise are (1) the symmetric nature of this expression and (2) the
continuous variation of the effective dielectric response of the mixture, EMGT. The
first issue can be finessed with a focus on well-separated minority grains in a
surrounding matrix. The second issue is more physical and relates to the lack of a
predicted percolation threshold, a limitation that can be traced back to the assump-
tion that embedded grains are not in contact. That said, the Maxwell-Garnett model
has been widely employed to understand stained glass, metallic particles like Al,
Au, and Ag, granular superconductors like Sn and Pb, and carbon nanotubes (with
an appropriate depolarization factor) [243, 244, 247–250].
The effective medium approximation provides an alternate framework for the
description of inhomogeneous media [245]. In a two-component system with
spherical grains, Ea is the dielectric response of the first component (present with
volume fraction f ), and Eb is the dielectric response of the second component
(present with volume fraction (1 – f )). In this model, each individual grain is
considered to be surrounded by a homogeneous host that possesses the average
dielectric properties of the medium. Within the effective medium theory, the
effective response of a two-component system with spherical grains is given as

Ea  EEMA Eb  EEMA
f þ ð1  f Þ : (5.3)
Ea þ 2EEMA Eb þ 2EEMA

Equation (5.3) can be extended to induce a depolarization factor that depends on the
shape of a grain or component. This approach differs from the Maxwell-Garnett
theory in that (1) it is not restricted to a particular range of concentration and (2) it
predicts a metal–insulator transition at a critical volume fraction of 1/3. Overall, the
effective medium approximation gives a superior description of the optical
properties for most systems. It has been used to understand the optical properties
of small Ag grains, a-MoGe thin film superconductors in magnetic field, and
the temperature-driven insulator–metal transition in VO2 [123, 251–253]. In the
a-MoGe superconductor, the effective medium approximation provides a good fit to
changes in transmittance ratios with magnetic field, from which the authors find that
the gap shrinks due to field-assisted pair breaking [252]. The superconducting
fraction also decreases linearly with magnetic field.
The optical properties of nanocomposites can be complicated by many other
factors including particle–particle interactions and scattering effects [254–256].
These effects are important in a variety of dense media such as coatings, paints,
and device applications. For larger particles (particle size  l), the effective
medium approximation breaks down and both absorption and scattering must be
taken into account [257]. Both dependent and independent scattering models can be
used to describe these effects, depending upon particle size, loading, and agglom-
eration effects.
5 Optical Properties of Nanoscale Transition Metal Oxides 95

5.3.3 Inhomogeneous Media and Surface Plasmons

A surface plasmon is a natural, collective oscillation of the electron gas inside of a


metallic nanomaterial that propagates parallel to the surface [258]. It is different
from the bulk plasma frequency for a material, op [156]. The basic problem is
solved by considering a metallic sphere with an incident electromagnetic wave with
electric field E0eiot. This incident wave activates the collective excitation. Writing
the internal electric field as a function of the Drude dielectric response of the
metallic sphere, the host dielectric, and the applied electric field, one extracts a
surface plasmon resonance frequency of

op
o ¼ pffiffiffiffiffiffiffiffiffiffiffiffi : (5.4)
E0 þ 2

Within the effective medium approximation, the surface plasmon peak develops
naturally above the percolation threshold. This electric dipole-allowed excitation
displays a characteristic peak in the absorption spectrum that often determines
the color properties of a material. It dominates the optical properties of a tradi-
tional system like Ag nanoparticles embedded in an insulating matrix [259]. It is
easy to understand why the plasmon resonance frequency is sensitive to quantum
size effects. As indicated in (5.3), the resonance frequency depends on the
dielectric function of the nanomaterial (which determines op) and the dielectric
properties of the host. The former is very sensitive to size effects, making it an
incisive probe of confinement in nanomaterials. To the extent that shape affects
the complex dielectric properties of a material, it will also influence the plasmon
resonance frequency. Studies of shape and aspect-ratio effects in silver
nanoparticles indicate important changes to both absorption profiles and scatter-
ing cross sections that emanate from the way electric field and charge are
concentrated on areas of high curvature or corners (Fig. 5.3) [256, 260–263].
Nanoshell structures offer electric field enhancements (and “hot volume” effects)
for similar reasons [264].

5.3.4 Charge and Bonding Models

Quantitative methods to evaluate charge and bonding in nanomaterials have also


attracted attention [161, 195]. Here, Born’s original formalism was extended in a
powerful way to include powdered samples – an approach that is very useful for
nanomaterials. Born effective charge (ZB ) can be calculated mode by mode from a
96 J.L. Musfeldt

Fig. 5.3 E-field enhancement contours external to the Ag trigonal prism, for a plane that is
perpendicular to the trigonal axis and that passes midway through the prism. The light is chosen to
have k along the trigonal axis and E along the abscissa. Left: 770 nm. Right: 460 nm. Side
length ¼ 100 nm and thickness ¼ 16 nm [260]

knowledge of oscillator strength (Sj) and the transverse optic phonon frequency
(oTO,j) as

X Nc2 X ðZB Þ2k


4p2 c2 o2TO; j Sj ¼ : (5.5)
j
E0 V k mk

With appropriate density and orientation corrections, this rendering is well


suited to the analysis of nanomaterials and was recently employed to evaluate
the Born effective charge for both 2H-MoS2 and analogous nanoparticles
(Fig. 5.4) [161]. In MoS2, the intralayer Born effective charge of the nanoparticles
is decreased significantly compared to the layered bulk. Of course, Born
effective charge describes the static and dynamic polarizations [342]. Local charge
(Z*) and polarizability (a) provide a way to separate these effects. They can be
extracted as

Z
ZB ¼ (5.6)
1  n Na
V

and

Na
Eð1Þ ¼ 1 þ V
: (5.7)
1  n Na
V

In the MoS2 nanoparticles, the significant decrease in the intralayer Born effective
charge was attributed to structural strain and the resulting change in polarizability
5 Optical Properties of Nanoscale Transition Metal Oxides 97

Fig. 5.4 (a) Crystal structure [265] and displacement patterns of infrared active E1u and A2u
vibrational modes [161, 266] in 2H-MoS2. These modes are sensitive to charge and bonding in the
intralayer and interlayer directions, respectively. (b) Close-up view of the 300 K reflectance
spectra of bulk and nanoscale MoS2 along with oscillator fits [161]. The inset shows a high-
resolution TEM image of the nanoparticles. (c) Born effective charge, local charge, and polariz-
ability of 2H- and IF-MoS2 in the two principle directions and a schematic view of the intralayer
electron clouds where spheres indicate a generalized orbital [161]

in the nanoparticles [161, 343, 344]. Thus far, this method has been applied only to
bulk and nanoscale MoS2, but it can be extended to oxides, where it is anticipated
that confinement will also modify ionicity.

5.4 Charge–Structure–Function Relationships


in Model Nanoscale Materials

The remainder of this chapter focuses on charge–structure–function relations in


nanomaterials using a set of case studies selected for chemical simplicity and
variety. Simple model systems are most amenable to quantitative approaches,
although there is still a lot to learn from the theoretical perspective about confine-
ment, local strain, charge and bonding, excitons, the role of defects, and how the
spectroscopic signatures of these features vary with size.
Another advantage of model systems is that they generally form the chemical
basis for a much wider class of materials. As a result, studies of finite length
scale effects in chemically simple model compounds have wide applicability.
Several physical systems are of interest including (1) the Mott–Hubbard compound
98 J.L. Musfeldt

VO2, (2) La1/2Sr3/2MnO4 in high magnetic field, (3) pristine and chemically
substituted vanadium oxide nanoscrolls, (4) quantum size effects in ZnO and
TiO2, (5) polar oxide thin films and nanoparticles based upon BiFeO3, and (6) H2
binding in metal-organic frameworks (MOFs). Discussion will focus on the
consequences of nanometer-sized physical length scales in these materials on
functionality.

5.4.1 Mott Transition in VO2 Revealed by Infrared Spectroscopy

Vanadium dioxide (VO2) is a chemically simple transition metal oxide that is at the
heart of the correlated electron vs. band structure/Peierls instability debate [267, 268].
At issue is whether electron correlations or electron–phonon interactions dominate
physics, particularly very near an insulator–metal transition. The general facts have
been clear for some time. VO2 is a monoclinic insulator at low temperature. It
displays an insulator-to-metal transition at ~342.5 K, above which VO2 is tetragonal
and metallic. The infrared properties of VO2 thin films in the low and high tempera-
ture phases are consistent with this picture, with a 0.5 eV gap in the insulating range
and a broad Drude-like feature in the metallic regime [123, 253, 269].
The optical response through the insulator–metal transition is of special
interest. Here, the gap fills gradually (borrowing strength from various high
energy electronic excitations [269]), and there is an isosbestic point in the optical
conductivity (Fig. 5.5a) that is seen in other Mott systems, which in chemical
kinetics is indicative of a simple, straightforward mechanism. Combining these
far-field infrared results with near-field infrared spectroscopy offers a way for-
ward [123, 253]. The near-field results, obtained with scanning near-field infrared
microscopy in tapping mode, provide detailed information on scattering
amplitudes of the insulating and metallic components and their relative spatial
positioning. This visualization (and the subsequent analysis) is made possible
because of the excellent dielectric contrast between the insulating and metallic
phases at the probe frequency (at 930 cm1) and the 10–20 nm linear system
resolution. Images of the near-field scattering amplitude clearly show the devel-
opment of metallic domain structure with increasing temperature in the transition
regime (Fig. 5.5b) [123, 253]. At the onset of the transition, these metallic islands
are embedded in an insulating matrix. They increase in size and begin to connect
with increasing temperature. The observed coexistence of metallic and insulating
regimes in the transition regime demonstrates the percolative nature of the
insulator-to-metal transition. As the metallic islands grow, their optical properties
evolve into those of the high temperature rutile phase of VO2. The transition is
complete by 360 K, as evidenced by the lack of contrast in the scanning images.
The metallic puddles in the transition regime are small (1 mm2) below the
percolation threshold, and they have several unusual properties, as described
below. Combining the far- and near-field infrared data with a modified effective
5 Optical Properties of Nanoscale Transition Metal Oxides 99

Fig. 5.5 (a) Optical conductivity of VO2 as a function of frequency for various temperatures.
(b) Images of near-field scanning amplitude over a 4  mm area showing the insulating and
metallic domain structure and how it develops through the transition regime. (c) Optical conduc-
tivity of the metallic domains extracted from a modified effective medium analysis. (d, e) The
relaxation rate and effective mass vs. temperature [123]

medium approach, the authors extracted the optical conductivity of the metallic
puddles [123, 253]. The latter shows a very narrow Drude-like feature along with
a pinned electronic band centered at 0.25 eV, quite different from the spectrum of
the fully metallic high temperature state. The metallic puddles also display a
divergent effective carrier mass and a pseudogap-like structure in the relaxation
rate (Fig. 5.5c). Both signal the importance of electron correlation. The diverging
effective mass in particular is an unambiguous signature of the Mott transition
[270]. This indicates that while lattice effects may play a role in the temperature-
driven insulator-to-metal transition, VO2 is primarily a correlated electron mate-
rial. This experimental approach will be very useful in addressing questions
regarding charge dynamics in other inhomogeneous correlated oxides like
cuprates and manganites.
100 J.L. Musfeldt

Fig. 5.6 Polarizing microscope images in the ab plane of La1/2Sr3/2MnO4 at 180 K in magnetic
fields of (a) 0 T, (b) 20 T, (c) 26 T, and (d) 27 T. (e) Averaged intensity of a part of the polarizing
microscope images [marked by red squares in (a–d)] as a function of applied field. The data at each
temperature are normalized so as to reproduce the intensity at each temperature in zero field [271]

5.4.2 Visualizing Charge and Orbitally Ordered Domains


in La1/2Sr3/2MnO4 at High Magnetic Fields

Interest in micro- and nanoscale texture in high magnetic fields is also driving new
instrumentation. An important example is a recently developed polarizing micro-
scope equipped with a variable temperature cryostat, pulsed magnet (to 35 T), and
high speed camera [271]. This system captures domain structure images and their
changes with magnetic field, essentially representing a fusion of magneto-optical
techniques and spatially resolved imaging. This setup was recently used to visualize
and elucidate domain structure changes through the field-induced collapse of the
charge and orbitally ordered state in the mixed valent manganite La1/2Sr3/2MnO4
(Fig. 5.6) [271]. In these images, the charge ordered state is birefringent and
appears as a bright area under crossed polarizers. Local intensity is thus propor-
tional to the order parameter of the transition from the charge and orbitally ordered
5 Optical Properties of Nanoscale Transition Metal Oxides 101

state to the quenched state, although the functional form is not yet clear. The dark
image in Fig. 5.6d at full field demonstrates that the charge and orbital ordering is
disordered, at least over the length scales probed in these measurements (4.8 mm/
pixel). Clearly, development and widespread use of this instrumentation will enable
major contributions to our understanding of magneto-optical effects in complex
transition metal oxides, especially as the image resolution improves. The magneto-
structural transitions recently reviewed in [272] may provide especially interesting
opportunities.

5.4.3 Discovery of Bound Carrier Excitation in Metal-Exchanged


Vanadium Oxide Nanoscrolls and Size Dependence
of the Equatorial Stretching Modes

The discovery that low-dimensional inorganic solids can curve or fold into nanoscale
objects provides an exciting opportunity to investigate bulk vs. nanoscale chemistry
using molecular-level strain as the tuning parameter [72, 75, 76]. Among the transi-
tion metal oxides, vanadates show particularly rich chemistry due to the tunable
oxidation state and flexible coordination environment, which ranges from octahedral
to square pyramidal to tetrahedral with increasing vanadium oxidation state [273].
They also display open framework structures, making them prospective materials for
ion intercalation, exchange, and storage [110, 111, 274, 275]. One nanoscale system
of interest is mixed-valent (amine)yVOx with x ~ 2.4. These compounds are actually
nanoscrolls, consisting of vanadium oxide layers between which organic molecules
are intercalated [110, 111, 276–278]. The latter controls sheet distance – essentially
the scroll winding. The optical response of mixed-valent vanadate nanoscrolls and
their metal-exchanged derivatives was investigated to understand the charge dynam-
ics in these compounds (Fig. 5.7). Unexpectedly, a bound carrier excitation was
observed in the metal-exchanged nanoscrolls, rather than the predicted free carrier
response [139]. This excitation is localized near 400 cm1 (0.05 eV) at all
temperatures and yields an extrapolated dc conductivity of ~1 O1 cm1, indicating
that the Mn2+-exchanged nanoscrolls are weakly metallic in their bulk form [279].
Detailed analysis suggests that this excitation is localized due to a combination of
inhomogeneous charge disproportionation, electron–electron correlation, and chemi-
cal disorder, motivating an alternate band filling picture that accounts for these effects
(Fig. 5.7c) [279]. Resistivity and thermopower measurements find a similar energy
required to propagate charge by polaronic mechanisms (0.09 eV) [280]. The higher
energy electronic structure of the (amine)yVOx materials consists of an 0.56 eV
charge gap at 300 K that does not depend on scroll size, superimposed V4+ d ! d
and V4+ ! V5+ transitions centered at ~1 eV, and a series of charge transfer
excitations above 3.2 eV [104, 279]. The feature near 5 eV depends on scroll size,
suggesting that the excitation is polarized preferentially in the equatorial direction.
102 J.L. Musfeldt

Fig. 5.7 (a) Structural information on (C12H25NH3)-VOx scrolls. (b) 300 K optical conductivity
of pristine VOx scrolls and the Mn-exchanged compound in the far-infrared regime. The dotted
line guides the eye, highlighting the additional bound carrier contribution in the substituted scrolls.
(c) Schematic representation of possible electronic structure changes emanating from the ion
exchange process in the scrolled vanadates, showing the effects of additional carriers,
electron–electron interactions, and chemical disorder. (d) Expanded view of the 300 K optical
conductivity of pristine VOx scrolls and the Mn-exchanged compound [279]

The vibrational properties of the (amine)yVOx scrolls also display a sheet


distance dependence. In the pristine (amine)yVOx materials, the equatorial V-O-V
stretching modes are particularly sensitive to sheet curvature, blueshifting, and
broadening with increasing sheet distance [104]. The blueshift is a typical quantum
size effect, and the broadening and peak asymmetry at small scroll sizes emanate
from k 6¼ 0 mode contributions to the response. Axially directed modes are much
less sensitive to the microscopic manifestations of strain. The 113 cm1 mode,
assigned as screw-like motion of the VOx tubes in analogy to the radial breathing
mode in carbon nanotubes, is a unique feature of these nanoscrolls [104]. Analysis
of vibrational structures in the metal-substituted scrolls indicated that ion exchange
modifies local charge and symmetry. The 575 cm1 V-O-V equatorial stretching
mode is especially sensitive to ion substitution [279].
5 Optical Properties of Nanoscale Transition Metal Oxides 103

a D0X 10 K
ZnO thin film

PL Intensity [a.u.] EX

b D0X 10 K

As-grown
ZnO QDs

3.10 3.35 3.60


Photon Energy [eV]

Fig. 5.8 Photoluminescence spectra of the ZnO thin film and quantum dots grown on the SiO2/Si
(111) substrate at 550
C measured at 10 K. (a) ZnO thin film grown for 60 min and (b) ZnO
quantum dots grown for 90 s. The free exciton peak at 3.377 eV is labeled “EX.” It corresponds to
the 10 K bulk band gap [232]

5.4.4 Classic Test Cases: Quantum Size Effects in ZnO and TiO2

The photophysical response of the II-VI compound semiconductor ZnO provides a


classic example of confinement effects [231, 281, 282]. Since the Bohr radius of
ZnO is 2 nm, small nanoscale objects are the most interesting (Fig. 5.1) [231].
Here, the direct gap is determined by the exciton [281, 282]. It blueshifts from a
room temperature value of ~400 nm in the bulk to 340 and 325 nm in 9.3 and 6.1 nm
particles, respectively – a clear signature of quantum size effects [281]. The exciton
binding energy is ~60 meV, making ZnO useful for short wavelength optical device
applications. Light emission in nanoscale ZnO has been extensively investigated
[232, 283–287].2 Figure 5.8 displays a comparison of the photoluminescence for
ZnO in thin film and nanoparticle form [232]. Importantly, light emission is observed
at significantly higher energies in the nanoparticles. This blueshift is attributed to
the quantum confinement effect. Kim et al. estimate the band gap enhancement in
nanoscale ZnO using the second term of (5.1) along with appropriate electron and
hole masses [232]. Other authors include the Coulomb contribution [284] (third
term in (5.1)) and the exciton binding contribution [283] (the fourth term in (5.1)).
Even these modified expressions predict exciton shifts that are somewhat larger

2
The general results are clear, with some features sensitive to the growth conditions (and the
defect states associated with the surface).
104 J.L. Musfeldt

than observed in experiment [232, 283, 284], a discrepancy that may be related to
the actual barrier potential in ZnO as well as assumptions about the importance of
Coulomb interactions in the model. The additional width of the nanoparticle
spectrum in Fig. 5.8 is attributed to size distribution effects [232, 283, 284]. In
the future, ZnO nanostructures with novel shapes, engineered defects, and differ-
ent chemical doping [288–294] may allow quantitative testing of models related
to size dependence of shallow and deep trapped states [227].
Nanoscale TiO2 also provides a classic illustration of quantum size effects [233,
295–297]. Rutile and anatase forms of TiO2 have indirect band gaps at 3.0 and
3.2 eV, respectively, followed by direct gaps at 3.3 and 3.4 eV, respectively
[298–300]. The recent synthetic breakthrough by Satoh et al. involving the use of
dendrimers to control nanoparticle size and distribution enabled very important
studies of confinement and growth in both the rutile and anatase forms [233].
Particle sizes below 2 nm are again most interesting for confinement studies, and
loss of translational invariance at small sizes seems to convert an indirect transition
to a direct one. The spectral blueshift with decreasing size is clearly observed for
both crystalline forms of TiO2 (Fig. 5.9), and the authors fit their data with a
modified version of the Brus equation [227, 228] (5.1) that better accounts for finite
well depth effects [233]. Clearly, fundamental understanding of band gap trends
will impact photocatalytic and solar cell applications [301, 302]. Analysis of small
cluster trends combined with modeling studies is revealing distinctions in the early
stage growth mechanism of rutile vs. anatase morphologies as well [233].

5.4.5 Optical Properties of Polar Oxide Thin Films


and Nanoparticles

Iron-based compounds and mixed metal oxides are promising materials with which
to harvest solar energy [303]. One system that has attracted attention in this regard
is BiFeO3, currently the only single phase room temperature multiferroic. Important
predictions from first-principles electronic structure calculations include strong
transition metal-oxygen hybridization, a stereochemically active Bi lone pair that
mixes Bi and O states, and insulating behavior in calculations that employ strong
exchange correlation [304–307]. The optical properties of BiFeO3 and related
systems are of great current interest for understanding the nature and size of the
band gap as well as chemical bonding and hybridization [195, 308]. The preparation
of high-quality thin films with precisely controlled oxygen stoichiometry, chemical
substitution, and epitaxial strain (to stabilize nonequilibrium phases) allows us to
compare predictions from first-principles electronic structure calculations with
optical property measurements. Recent reports of photoconductivity [308], a pho-
tovoltaic effect [309], and a 0.8–0.9 V open circuit voltage in a working ferroelec-
tric solar device [310] illustrate the potential of polar oxides as active photovoltaic
materials [308–310]. The light harvesting efficiency in BiFeO3 thin films is,
5 Optical Properties of Nanoscale Transition Metal Oxides 105

a
1.4
3.0
1.2 2.5
2.0

(αhν)0.5
1.0
1.5
Absorbance

0.8 1.0
0.5
0.6
0
3.0 3.5 4.0 4.5 5.0
0.4
Energy (eV)

0.2

0
250 300 350 400 450
Wavelength (nm)
b
1.4
3.0
1.2 2.5
2.0
(αhν)0.5

1.0
1.5
Absorbance

0.8
1.0

0.6 0.5
0
0.4 3.0 3.5 4.0 4.5 5.0
Energy (eV)
0.2

0
250 300 350 400 450
Wavelength (nm)

Fig. 5.9 Band gap measurements for 6TiO2 (green), 14TiO2 (red), and 30TiO2 (blue). Optical
wave guide spectra for hydrolyzed (rutile) TiO2 (a) and thermolyzed (anatase) TiO2. (b) The insets
show Tauc plots for the hydrolyzed and thermolyzed TiO2, respectively. From the relation
(ahn)0.5 ~ (hn Eg), we can obtain Eg from the Tauc plots for the indirect band gap. Taking into
account the lowest direct band gap, we obtained the regression lines in the regions 4.00–4.26,
3.60–3.94, and 3.47–3.78 eV for the hydrolyzed 6TiO2, 14TiO2, and 30TiO2 and 3.88–4.10,
3.67–3.92, and 3.44–3.77 eV for the thermolyzed 6TiO2, 14TiO2, and 30TiO2, respectively
(estimated error in energies  0.02 eV, coefficients of determination  95%) [233]

however, still limited by a variety of factors including carrier lifetimes, recombina-


tion rates, and matching the optical band gap and overall electronic structure with
the solar spectrum. Optical spectroscopy can support the development of new thin
film polar oxides for solar devices by tackling the band gap problem. It is, of course,
worth remembering that oxides are attractive because they are cheap, abundant, and
106 J.L. Musfeldt

stable, and their properties can be tuned by chemical substitution. Among the
oxides, ferroelectrics present an alternative pathway to charge separation of photo-
excited carriers potentially eliminating the need for a p-n junction. The latest
BiFeO3-based all-oxide heterostructures are characterized by open-circuit voltages
greater than 0.8–0.9 eV and quantum efficiencies up to 10% when illuminated with
light of energies above the band gap [310]. This efficiency is at least an order of
magnitude larger than the maximum efficiency under sunlight (AM 1.5) thus far
reported for ferroelectric thin film devices [310]. To extend the impact of this
discovery and possibly provide a new approach to solar energy conversion, it is
desirable to tune the band gap and electronic excitations to coincide with the
550–700 nm peak in the solar spectrum and to take advantage of the long near
infrared tail. This approach to light harvesting is complementary to many others
such as catalytically driven H2O-splitting reactions, dye sensitized cells, and molec-
ular electronic-based assemblies [311]. The discovery that BiFeO3 also forms
nanoparticles [312–314] provides additional avenues for exploration.
The optical properties of rhombohedral BiFeO3 (Fig. 5.10a) are well
documented [195, 308, 319, 320]. The 300 K direct charge gap is at 2.67 eV.3
It is preceded by a small shoulder centered at ~2.5 eV, which yields an absorption
onset near 2.2 eV. Peaks at 3.2 and 4.5 eV are dipole-allowed charge transfer
excitations [195, 308, 319, 320]. Our efforts to tune the band gap have focused on
(1) tetragonal BiFeO3 stabilized via a compressive strain from the substrate
(Fig. 5.10b), (2) a series of BiFeO3 nanoparticles (Fig. 5.10c), and (3) a series of
alloys with Mn substitution, Bi(Fe,Mn)O3, including the end member BiMnO3
(Fig. 5.10d) [315–317]. Taken together, these systems offer an excellent test
of band gap tuning methodologies, previewing how chemical substitution and
nanoscale strain can impact the match with the solar spectrum, and by corollary,
energy conversion efficiencies of a device. The optical properties of tetragonal
and nanoscale BiFeO3 allow us to evaluate the potential of strain as a tuning
parameter. The absorption spectrum of the tetragonal film is overall blueshifted
compared with that of the rhombohedral material [315]. It shows an absorption
onset near 2.25 eV, a direct 3.1 eV bandgap, and charge transfer excitations that are
~0.4 eV higher than those of the rhombohedral counterpart (Fig. 5.10b). In the
nanoparticles, the band gap decreases from 2.7 to ~2.3 eV, and the well-known 3.2
and 4.5 eV charge transfer excitations split into multiplets (Fig. 5.10c) [316].
Again, the band gap is defined by the charge transfer edge. The spectral redshift
with decreasing particle size is different from normal confinement-induced
blueshifts [227] (5.1) but in line with recent results on BiFeO3 micro-cubes [321],
indicative of important Coulomb interactions in BiFeO3. These results can be
understood in terms of structure, strain, and symmetry breaking. Focusing on the
onset to optical absorption and the direct gap analysis via an (aE)2 vs. E plot in

3
The theory of energy gap determination in solids is, of course, well established. The absorption
coefficient, a(E), consists of contributions from both the direct and the indirect band gap
transitions, the expression for which can be found in any standard optics text [157].
5 Optical Properties of Nanoscale Transition Metal Oxides 107

Fig. 5.10 (a) 300 K optical absorption of a rhombohedral BiFeO3 thin film compared with the
solar spectrum [195, 308]. (b) Response of a tetragonal phase BiFeO3 film compared with that of
the rhombohedral material [315]. Here, strain blueshifts the gap and the electronic excitations. (c)
Optical properties of 31 nm nanoparticles showing a redshifted gap compared with the
rhombohedral film [316]. (d) 300 K optical absorption spectrum of BiMnO3, the end member in
a series of alloy films, compared with BiFeO3 [317, 318]. The Mn-substituted material displays a
lower band gap

Fig. 5.10d, we find that Mn substitution yields an overall redshift of the


oscillator strength and a 1.1 eV charge gap in the BiMnO3 end member [317,
318]. This gap emanates from strongly hybridized O 2p + Mn 3d ! Mn 3d charge
transfer excitations. Taken together, these data demonstrate the potential of band
gap tuning to improve polar oxide-based ferroelectric solar cell efficiency. Given
the recent report of a piezoelectric response in strained BiFeO3 [322], control of
the strain-driven morphotropic phase boundary may also give rise to electro-
optical effects.
108 J.L. Musfeldt

5.4.6 Spectroscopic Determination of H2 Binding Sites


and Energies in Metal Organic Framework Materials

Local charge and structure govern the properties of many complex materials
including superconductors, thermoelectrics, metamaterials, and multiferroics
[171, 279, 323–326]. Charge and structure are also intimately connected to func-
tionality in porous materials like metal organic framework materials compounds
that have nanometer-sized pores, channels, and active sites [77, 327]. Here, absorp-
tion, binding, and chemical reactivity are the functionalities of interest, and they
have direct application to hydrogen storage, carbon/nitrogen sequestration, and
catalysis [328–334].4 Infrared vibrational spectroscopy is well suited to analyze
the chemical and physical aspects of small molecules absorbed on high surface area
materials like metal organic framework materials because, as a microscopic tech-
nique, it can provide information on the relative energies of different binding sites
in these materials, especially when combined with first-principles calculations.
Pore size, shape, and local charge seem to determine much of the hydrogen
binding chemistry in metal organic framework materials [335]. Corners in the
framework structure or exposed metal coordination sites within the pore structure
concentrate charge, change the local acid/base character, and present convenient
attachment sites. Monte Carlo calculations predict optimum pore sizes to be ~7 and
10 Å at 300 and 77 K, respectively [336]. From the bulk properties point of view,
H2 has weak surface interactions at 300 K, so it is hard to get good high temperature
performance in metal organic framework materials. This is because thermal energy
exceeds binding energy. Reasonable bulk hydrogen uptake data can, however, be
obtained at 77 K [337, 338]. To advance hydrogen storage applications, it is clearly
desirable to measure, tune, and control the H2 binding energy.
MOF-5 is one system for which the uptake and binding site problem is
reasonably well understood [337, 339]. The structure of MOF-5 consists of a
cubic lattice of Zn4O tetrahedra connected by 1,4-benzenedicarboxylate groups
[340]. The zinc centers are fully coordinated and unexposed. In their work,
Bordiga et al. exploit the fact that interaction of the surface with a hydrogen
molecule breaks local symmetry, perturbs the charge density, and activates the
symmetric stretching mode of H2. Interaction with the surface also induces a
frequency shift that can be related to the local charge, structure, and accessibility
of the binding site. Because the frequency shift increases with increasing interac-
tion energy, different absorbing sites yield distinct infrared features [339].
Figure 5.11 displays a closeup view of the infrared absorption response of
MOF-5 as a function of H2 uptake [339]. With increasing uptake, sharp new
bands appear in the 4,110–4,150 cm1 range, direct evidence for at least three

4
Work on hydrogen storage is, of course, directed toward meeting Department of Energy
hydrogen storage system targets.
5 Optical Properties of Nanoscale Transition Metal Oxides 109

Fig. 5.11 Dependence of the equilibrium constant for absorption, on 1/T for H2 adsorbed on
MOF-5 sites labeled 1, 2, and 3 in the inset. Adsorption enthalpies (DH) calculated from slope lines
are 7.4 0.2 kJ/mol for sites 1 and 2 and 3.5 0.1 kJ/mol for site 3. Inset: Infrared spectra of H2
adsorbed on MOF-5 as a function of decreasing temperature (upper curve 30 K and 0.019 bar
equilibrium pressure; lower curve 110 K and 0.035 bar equilibrium pressure) [339]

distinct binding sites [339]. Interaction energies can be determined from an


equilibrium analysis with the van’t Hoff equation [339]. Here,

yðTÞ
Kads ¼ ; (5.8)
½1  yðTÞ PH2

where y(T) is the fraction of sites covered by H2 (determined from the relative
infrared intensity), PH2 is the loading pressure, and Kads is the equilibrium constant
for the absorption process. The slope of ln Kabs vs. 1/T gives the absorption energy.
These dependencies are plotted in Fig. 5.11 for the active sites in MOF-5 [339]. The
strongest H2 binding (3.5 0.1 kJ/mol) occurs at the corners of the pores, at the
carboxylate O2 centers behind which lie the Zn2+ ions [339].
Metal organic framework materials with exposed metal sites are providing a way
forward with improved bulk H2 uptake results and significantly larger H2. . . metal
binding energies [338, 340, 341]. For instance, a site-specific absorption energy was
recently determined for CPO-27-Ni, a metal organic framework materials system
with accessible Ni2+ sites. Vibrational spectroscopy reveals an interaction energy of
13.5 kJ/mol [340]. Clearly, exposed metal sites are poised to play a very important
role in the design of new metal organic framework materials. Charge variation on
110 J.L. Musfeldt

the unsaturated metal site may provide both control of the H2 binding energy and
mechanistic insight. At the same time, structures that contain a higher density of
absorption sites may offer additional opportunities to boost H2 uptake [340].

5.5 Summary and Outlook

Finite length scale effects are a major new theme for transition metal oxides. They
present compelling opportunities for discovery-class science and new areas of
phase space within which to search for exciting physical properties. Optical spec-
troscopy is well positioned to capitalize on these opportunities. This is because it is
a versatile and sensitive microscopic probe that can easily be combined with
various physical tuning techniques. Certainly, one goal of this review is to advance
the idea that size–shape effects are on par with temperature, magnetic field,
pressure, and electric field in their potential to tune complex materials. At the
same time, optical spectroscopy connects easily with existing models and can be
used to build upon and test new theoretical approaches. This review takes the case
study approach, highlighting recent findings in several prototypical transition metal
oxides including (1) the Mott–Hubbard compound VO2, (2) La1/2Sr3/2MnO4 in high
magnetic field, (3) pristine and chemically substituted vanadium oxide nanoscrolls,
(4) quantum size effects in ZnO and TiO2, (5) polar oxide thin films and nanoparticles
based upon BiFeO3, and (6) H2 binding in metal organic framework materials. In
these examples, efforts were made to bring simple chemical systems together with
appropriate models of insulator-to-metal transitions, confinement, local strain, charge
and bonding, excitons, and the role of defects for high level physical understanding.
Discussion focuses on the consequences of quantum size effects on properties and
functionality. There are clearly many opportunities for spectroscopists to advance the
field of nanoscience. Investigations that emphasize systematic structure–property
correlations or combine optical spectroscopy with physical tuning techniques and
mechanistic models will be particularly valuable.

Acknowledgments JLM thanks the Materials Science Division, Basic Energy Sciences, U.S.
Department of Energy and the Joint Directed Research and Development Program at the Univer-
sity of Tennessee for support of this work.

References

1. Rao, C.N.R., Raveau, B.: Transition Metal Oxides. VCH, New York (1995)
2. Hwu, S.-J.: Structurally confined transition-metal oxide layers, chains and oligomers in
molecular and extended magnetic solids. Chem. Mater. 10(10), 2846–2859 (1998)
3. Tokura, Y. (ed.): Colossal Magnetoresistive Oxides. Gordon and Breach, New York (2000)
4. Julien, C., Pereira-Ramos, J.P., Momchilov, A. (eds.): New Trends in Intercalation
Compounds for Energy Storage, NATO Science Series II: Mathematics, Physics, and Chem-
istry, vol. 61. Kluwer, New York (2002)
5 Optical Properties of Nanoscale Transition Metal Oxides 111

5. Gr€uner, G.: The dynamics of charge density waves. Rev. Mod. Phys. 60, 1129–1181 (1988)
6. Greenblatt, M.: Monophosphate tungsten bronzes: a new family of low-dimensional, charge-
density-wave oxides. Acc. Chem. Res. 29, 219–228 (1996)
7. Kahn, O.: Molecular Magnetism. VCH, New York (1993)
8. Canadell, E.: Dimensionality and Fermi surface of low-dimensional metals. Chem. Mater. 10,
2770–2786 (1998)
9. Dagotto, E.: Nanoscale Phase Separation and Colossal Magneto-Resistance. Springer Series
on Solid State Sciences. Springer, Berlin (2003)
10. Cuk, T., Lu, H.H., Zhou, X.J., Shen, Z.-X., Devereaux, T.P., Nagosa, N.: A review of
electron-phonon coupling seen in the high-Tc superconductors by angle-resolved photoemis-
sion studies. Phys. Status Solidi B 242, 11–29 (2005)
11. Pimenov, A., Rudolf, R., Mayr, F., Loidl, A., Mukhin, A.A., Balbashov, A.M.: Coupling of
phonons and electromagnons in GdMnO3. Phys. Rev. B 74, 100403(R) (2006)
12. Sushkov, A.B., Valdés-Aguilar, R., Park, S., Cheong, S.-W., Drew, H.D.: Electromagnons in
multiferroic YMn2O5 and TbMn2O5. Phys. Rev. Lett. 98, 027202 (2007)
13. White, R.M., Yen, W.M.: On the discovery of magnon sidebands in insulating
antiferromagnets. Low Temp. Phys. 31, 777–779 (2005)
14. Born, M., Huang, K.: Dynamical Theory of Crystal Lattices. Oxford University Press,
London (1954)
15. Goodenough, J.: Magnetism and the Chemical Bond. Wiley, New York (1963)
16. Egami, T., Billinge, S.J.L.: Underneath the Bragg Peaks: Structural Analysis of Complex
Materials. Pergamon Materials Series, vol. 7. Pergamon, Elsevier, Oxford (2003)
17. Kamihara, Y., Watanabe, T., Hirano, M., Hosono, H.: Iron-based layered superconductor La
[O1xFx]FeAs (x ¼ 0.05–0.12) with Tc ¼ 26 K. J. Am. Chem. Soc. 130, 3296–3297 (2008)
18. Dressel, M., Drichko, N.: Optical properties of two-dimensional organic conductors:
signatures of charge ordering and correlation effects. Chem. Rev. 104, 5689–5716 (2004)
19. Zhou, H.D., Lumata, L.L., Kuhns, P.L., Reyes, A.P., Choi, E.S., Dalal, N.S., Lu, J., Jo, Y.J.,
Balicas, L., Brooks, J.S., Wiebe, C.R.: Ba3NbFe3Si2O14: A new multiferroic with a 2D
triangular Fe3+ motif. Chem. Mater. 21(1), 156–159 (2009)
20. Sasmal, K., Lev, B., Lorenz, B., Guloy, A.M., Chen, F., Xue, Y.-Y., Chu, C.-W.:
Superconducting Fe-based compounds (A1xSrx)Fe2As2 with A ¼ K and Cs with transition
temperatures up to 37 K. Phys. Rev. Lett. 101, 107007 (2008)
21. Orenstein, J., Millis, A.J.: Advances in the physics of high-temperature superconductivity.
Science 288, 468–474 (2000)
22. Yoo, C.S., Maddox, B., Klepeis, J.-H.P., Iota, V., Evans, W., McMahan, A., Hu, M.Y., Chow, P.,
Somayazulu, M., Husermann, D., Scalettar, R.T., Pickett, W.E.: First-order isostructural Mott
transition in highly compressed MnO. Phys. Rev. Lett. 94, 115502 (2005)
23. Basov, D.N., Timusk, T.: Electrodynamics of high-Tc superconductors. Rev. Mod. Phys. 77,
721 (2005)
24. Hur, N., Park, S., Sharma, P.A., Ahn, J.S., Guha, S., Cheong, S.-W.: Electronic polarization
reversal and memory in a multiferroic materials induced by magnetic fields. Nature 429,
392–395 (2004)
25. Brooks, J.S.: Magnetic field dependent and induced ground states in organic conductors. Rep.
Prog. Phys. 71, 126501 (2008)
26. Snow, C.S., Karpus, J.F., Cooper, S.L., Kidd, T.E., Chiang, T.-C.: Quantum melting of the
charge density wave state in 1T-TiSe2. Phys. Rev. Lett. 91, 136402 (2003)
27. Goddard, P.A., Singleton, J., Sengupta, P., McDonald, R.D., Lancaster, T., Blundell, S.J.,
Pratt, F.L., Cox, S., Harrison, N., Mason, J.L., Southerland, H.I., Schlueter, J.A.: Experimen-
tally determining the exchange parameters of quasi-two-dimensional Heisenberg magnets.
New J. Phys. 10, 083025 (2008)
28. dela Cruz, C.R., Lorenz, B., Sun, Y.Y., Wang, Y., Park, S., Cheong, S.-W., Gospodinov, M.
M., Chu, C.W.: Pressure-induced enhancement of ferroelectricity in multiferroic RMn2O5
(R¼Tb,Dy,Ho). Phys. Rev. B 76, 174106 (2007)
112 J.L. Musfeldt

29. Graf, D., Choi, E.S., Brooks, J.S., Almeida, M.: Pressure-induced quantum limit in a Q1D
system in high magnetic fields. J. Low Temp. Phys. 142, 179–184 (2006)
30. Uji, S., Terashima, T., Nishimura, M., Takahide, Y., Konoike, T., Enomoto, K., Cui, H.,
Kobayashi, H., Kobayashi, A., Tanaka, H., Tokumoto, M., Choi, E.-S., Tokumoto, T., Graf,
D., Brooks, J.S.: Vortex dynamics and the Fulde-Ferrell-Larkin-Ovchinnikov state in a
magnetic-field-induced organic superconductor. Phys. Rev. Lett. 97, 157001 (2006)
31. Gupta, R., Kim, M., Barath, H., Cooper, S.L., Cao, G.: Field- and pressure-induced phases in
Sr4Ru3O10: A spectroscopic investigation. Phys. Rev. Lett. 96, 067004 (2006)
32. Kim, K.H., Harrison, N., Jaime, M., Boebinger, G.S., Mydosh, J.A.: Magnetic field-induced
quantum critical point and competing order parameters in URu2Si2. Phys. Rev. Lett. 91(25),
256401 (2003)
33. Zapf, V.S., Zocco, D., Hansen, B.D., Jaime, M., Harrison, N., Batista, C.D., Kenzelmann, M.,
Niedermayer, C., Lacerda, A., Paduan-Filho, A.: Bose-Einstein condensation of S ¼ 1 nickel
spin degrees of freedom in NiCl2-4SC(NH2)2. Phys. Rev. Lett. 96, 077204 (2006)
34. dela Cruz, C.R., Lorenz, B., Sun, Y.Y., Chu, C.W., Park, S., Cheong, S.-W.: Magnetoelastic
effects and the magnetic phase diagram of multiferroic DyMn2O5. Phys. Rev. B 74, 180402
(2006)
35. Sologubenko, A.V., Berggold, K., Lorenz, T., Rosch, A., Shimshoni, E., Phillips, M.D.,
Turnbull, M.M.: Magnetothermal transport in the spin-1/2 chains of copper pyrazine
dinitrate. Phys. Rev. Lett. 98, 107201 (2007)
36. Kunes, J., Lukoyanov, A.V., Anisimov, V.I., Scalettar, R.T., Pickett, W.E.: Collapse of
magnetic moment drives the Mott transition in MnO. Nat. Mater. 7, 198–202 (2008)
37. Itkis, M.E., Brill, J.W.: Electromodulated infrared transmission in blue bronze. Phys. Rev.
Lett. 72, 2049 (1994)
38. Ahn, C.H., Triscone, J.-M., Mannhart, J.: Electric field effect in correlated oxide systems.
Nature 424, 1015–1018 (2003)
39. Jaffe, B., Cook, W.R., Jaffe, H.: Piezoelectric Ceramics. Academic, London (1971)
40. Cooper, V.R., Grinberg, I., Martin, N.R., Rappe, A.M.: Local structure of PZT. In: Cohen, R.
E. (ed.) Fundamental Physics of Ferroelectrics, pp. 26–35. AIP, Melville, NY (2002)
41. Egami, T.: Local structure and dynamics of ferroelectric solids. In: Dalal, N.S., Bussmann-
Holder, A. (eds.) Structure and Bonding, Vol. 124, Ferro-and Antiferroelectricity. Springer,
Berlin (2007)
42. Ederer, C., Fennie, C.J.: Electric-field switchable magnetization via the Dzyaloshinskii-
Moriya interaction: FeTiO3 versus BiFeO3. J. Phys. Condens. Matter 20, 434219 (2008)
43. Fennie, C.J.: Ferroelectrically induced weak ferromagnetism by design. Phys. Rev. Lett. 100,
167203 (2008)
44. Holman, K.T., Pivovar, A.M., Ward, M.D.: Engineering crystal symmetry and polar order in
molecular host frameworks. Science 294, 1907–1911 (2001)
45. Ferrer, J.R., Lahti, P.M., George, C., Oliete, P., Julier, M., Palacio, F.: Role of hydrogen
bonds in benzimidazole-based organic magnetic materials: crystal scaffolding or exchange
linkers. Chem. Mater. 13, 2447–2454 (2001)
46. Murata, H., Aboaku, S., Lahti, P.M.: Molecular recognition in a heteromolecular radical pair
system with complementary multipoint hydrogen-bonding. Chem. Comm. 29, 3441–3443
(2008)
47. Hayward, M.A., Cussen, E.J., Claridge, J.B., Bieringer, M., Rosseinsky, M.J., Kiely, C.J.,
Blundell, S.J., Marshall, I.M., Pratt, F.L.: The hydride anion in an extended transition metal
oxide array: LaSrCoO3H0.7. Science 295, 1882–1884 (2002)
48. Ward, M.D.: Directing assembly of molecular crystals. MRS Bull. 30, 705–712 (2005)
49. Orendacova, A., Kajnakova, M., Cernak, J., Park, J.-H., Cizmar, E., Orendac, M., Vlcek, A.,
Kravchyna, O.V., Anders, A.G., Feher, A., Meisel, M.W.: Hydrogen bond mediated magne-
tism in [CuII(en)2(H2O)][CuII(en)2Ni2CuI2(CN)10]·2H2O. Chem. Phys. 309, 115–125 (2005)
50. Baddeley, C., Yan, Z., King, G., Woodward, P.M., Badjic, J.D.: Structure-function studies of
modular aromatics that form molecular organogels. J. Org. Chem. 72(19), 7270–7278 (2007)
5 Optical Properties of Nanoscale Transition Metal Oxides 113

51. Manson, J.L., Conner, M.M., Schlueter, J.A., McConnell, A.C., Southerland, H.I., Malfant, I.,
Lancaster, T., Blundell, S.J., Brooks, M.L., Pratt, F.L., Singleton, J., McDonald, R.D., Lee,
C., Whangbo, M.-H.: Experimental and theoretical characterization of the magnetic
properties of CuF2(H2O)2 (pyz) (pyz ¼ pyrazine): A two-dimensional quantum magnet
arising from supersuperexchange interactions through hydrogen bonded paths. Chem.
Mater. 20(24), 7408–7416 (2008)
52. Murata, H., Miyazaki, Y., Inaba, A., Paduan-Filho, A., Bindilatti, V., Fernandes Oliveira Jr.,
N., Delen, Z., Lahti, P.M.: 2-(4,5,6,7-Tetrafluorobenzimidazol-2-yl)-4,4,5,5-tetramethyl-4,5-
dihydro-1H-imidazole-3-oxide-1-oxyl, A hydrogen-bonded organic quasi-1D ferromagnet.
J. Am. Chem. Soc. 130, 186–194 (2008)
53. Manriquez, J.M., Yee, G.T., McLean, R.S., Epstein, A.J., Miller, J.S.: A room temperature
molecular/organic-based magnet. Science 252, 1415–1417 (1991)
54. Kaul, B.B., Durfee, W.S., Yee, G.T.: Dialkyldicyanofumarate diesters: tunable building
blocks for molecular-based ferromagnets. J. Am. Chem. Soc. 121, 6862–6866 (1999)
55. Berlinguette, C.P., Baughn, D., Canada-Vilalta, C., Galán-Mascarós, J.R., Dunbar, K.R.:
A trigonal-bipyramidal cyanide cluster with single molecular behavior: synthesis, structure,
magnetic properties of [MnII(tmpen)2]3[MnIII(CN)6]2. Angew. Chem. Int. Ed. 42, 1523–1526
(2003)
56. Miller, J.S., Epstein, A.J.: Designer magnets. Chem. Eng. News 73(40), 30 (1995)
57. Hill, S., Anderson, N., Wilson, A., Takahashi, S., Petukhov, K., Chakov, N.E., Murugesu, M.,
del North, J.M., Barco, E., Kent, A.D., Dalal, N.S., Christou, G.: A comparison between high-
symmetry Mn12 single-molecule magnets in different ligand/solvent environments. Polyhe-
dron 24, 2284–2292 (2005)
58. Miller, J.S.: Organometallic and organic based magnets: New chemistry and new materials
for the new millennium. Inorg. Chem. 39, 4392–4408 (2000)
59. Pokhodnya, K.I., Epstein, A.J., Miller, J.S.: Thin film V(TCNE)x magnets. Adv. Mater. 12,
410–413 (2000)
60. Pejaković, D.A., Manson, J.L., Miller, J.S., Epstein, A.J.: Photoinduced magnetism, dynam-
ics, cluster glass behavior of a molecule-based magnet. Phys. Rev. Lett. 85, 1994–1997
(2000)
61. Gı̂rtu, M.A., Wynn, C.M., Zhang, J., Miller, J.S., Epstein, A.J.: Magnetic properties and
critical behavior of Fe(tetracyanoethylene)2 · x(CH2Cl2): a high-Tc molecule-based magnet.
Phys. Rev. B 61, 492–500 (2000)
62. Sato, O., Iyoda, T., Fujishima, A.: Photoinduced magnetization of a cobalt-iron cyanide.
Science 272, 704–705 (1996)
63. Miller, J.S.: Three-dimensional network-structured cyanide-based magnets. MRS Bull. 25,
60–64 (2000)
64. Ohta, H., Nomura, K., Hiramatsu, H., Ueda, K., Kamiya, T., Hirano, M., Hosono, H.: Frontier
of transparent oxide semiconductors. Solid State Electron. 47, 2261–2267 (2003)
65. Kaye, S.S., Long, J.R.: Hydrogen storage in the dehydrated pressian blue analogues Mn3[Co
(CN)6]2(M ¼ Mn, Fe, Co, Ni, Cu, Zn). J. Am. Chem. Soc. 127, 6506–6507 (2005)
66. Culp, J.T., Park, J.H., Frye, F., Huh, Y.D., Meisel, M.W., Talham, D.R.: Magnetism of metal
cyanide networks assembled at interfaces. Coord. Chem. Rev. 249, 2642–2648 (2005)
67. Berlinguette, C.P., Dragulescu-Andrasi, A., Sieber, A., Galán-Mascarós, J.R., G€ udel, H.,
Achim, C., Dunbar, K.R.: A charge-transfer-induced spin transition in the discrete cyanide-
bridged complex [Co(tmphen)2]3[Fe(CN)6]2. J. Am. Chem. Soc. 126, 6222–6223 (2004)
68. Holmes, S.M., Girolami, G.S.: Sol-gel synthesis of KVII[CrIII(CN)6] · 2H2O: a crystalline
molecule-based magnet with a magnetic ordering temperature above 100
C. J. Am. Chem.
Soc. 121, 5593–5594 (1999)
69. Escax, V., Bleuzen, A., Itie, J.P., Munsch, P., Varret, F., Verdaguer, M.: Nature of the long-
range structural changes induced by the molecular photoexcitation and by the relaxation in
the Prussian blue analogs Rb1.8Co4[Fe(CN)6]3.3 · 13H2O:a synchrotron diffraction study.
J. Phys. Chem. B. 107, 4763–4767 (2003)
114 J.L. Musfeldt

70. Li, D., Clérac, R., Roubeau, O., Harté, E., Mathoniére, C., Le Bris, R., Holmes, S.M.:
Magnetic and optical bistability driven by thermally and photo-induced intramolecular
electron transfer in a molecular cobalt-iron Prussian blue analogue. J. Am. Chem. Soc. 130,
242–259 (2007)
71. Berlinguette, C.P., Draguleschu-Andrasi, A., Sieber, A., G€ udel, H.-U., Achim, C., Dunbar, K.
R.: A charge-transfer-induced spin transition in a discrete complex: the role of extrinsic
factors in stabilizing three electronic isomeric forms of a cyanide-bridged Co/Fe cluster.
J. Am. Chem. Soc. 127, 6766–6779 (2005)
72. Rao, C.N.R., Nath, M.: Inorganic nanotubes. Dalton Trans. 1, 1–24 (2003)
73. Bonnell, D.A.: Materials in nanotechnology: new structures, new properties, new complexity.
J. Vac. Sci. Technol. A 21, S194–S206 (2003)
74. Tenne, R., Rao, C.N.R.: Inorganic nanotubes. Philos. Trans. R. Soc. A 362, 2099–2125
(2004)
75. Halford, B.: Inorganic menagerie. Chem. Eng. News 83, 30–33 (2005)
76. Tenne, R.: Inorganic nanotubes and fullerene-like nanoparticles. Nat. Nanotechnol. 1,
103–111 (2006)
77. Long, J.W., Rolison, D.R.: Architectural design, interior decoration, three-dimensional
plumbing en route to multifunctional nanoarchitectures. Acc. Chem. Res. 40, 854–862 (2007)
78. Nobile, C., Kudera, S., Fiore, A., Carbone, L., Chilla, G., Kipp, T., Heitmann, D., Cingolani,
R., Manna, L., Krahne, R.: Confinement effects on the optical phonons in spherical, rod-,
tetrapod-shaped nanocrystals detected by Raman spectroscopy. Phys. Stat. Sol. (a) 204, 483
(2007)
79. Moses, M.J., Fettinger, J.C., Wichhorn, B.W.: Interpenetrating As20 fullerene and Ni12
iscosahedra in the onion-skin [As@Ni12@As20]. Science 300, 778–780 (2003)
80. Wu, Y., Messer, B., Yang, P.: Superconducting MgB2 nanowires. Adv. Mater. 13, 1487–1489
(2001)
81. Wang, Z.L., Kong, X.Y., Ding, Y., Gao, P., Hughes, W.L., Yang, R., Zhang, Y.: Semicon-
ducting and piezoelectric oxide nanostructures induced by polar surfaces. Adv. Funct. Mater.
14, 943–956 (2004)
82. Liu, C., Rondinone, A.J., Zhang, Z.J.: Sol-gel synthesis of free-standing ferroelectric lead
zirconate titanate nanoparticles. J. Am. Chem. Soc. 123, 4344–4345 (2001)
83. Kovtyukhova, N.I., Mallouk, T.E., Mayer, T.S.: Templated surface sol-gel synthesis of SiO2
nanotubes and SiO2-insulated metal nanowires. Adv. Mater. 15, 780–785 (2003)
84. Tian, M., Wang, J., Han, T., Kumar, N., Kobayashi, Y., Liu, Y., Mallouk, T.E., Chan, M.W.
H.: Superconductivity in granular Bi nanowires fabricated by electrochemical deposition at
ambient pressure. Nano Lett. 6, 2773–2780 (2006)
85. Rusakova, I., Ould-Ely, T., Hofmann, C., Prieto-Centurión, D., Levin, C.S., Halas, N.J.,
L€uttge, A., Whitmire, K.H.: Nanoparticle shape conservation in the conversion of MnO
nanocrosses into Mn3O4. Chem. Mater. 19, 1369–1375 (2007)
86. Ould-Ely, T., Prieto-Centurion, D., Kumar, A., Guo, W., Knowles, W.V., Asokan, S., Wong,
M.S., Rusakova, I., L€ uttge, A., Whitmire, K.H.: Manganese(II) oxide nanohexapods: Insight
into controlling the form of nanocrystals. Chem. Mater. 18, 1821–1829 (2006)
87. Tian, M.L., Wang, J.G., Kurtz, J.S., Liu, Y., Chan, M.H.W., Mayer, T.S., Mallouk, T.E.:
Dissipation in quasi-one-dimensional superconducting single crystal Sn nanowires. Phys.
Rev. B 71, 104521 (2005)
88. O’Dwyer, C., Navas, D., Lavayen, V., Benavente, E., Santa Ana, M.A., Gonzales, G.,
Newcomb, S.B., Sotomayor Torres, C.M.: Nanourchin: the formation and structure of high-
density spherical clusters of vanadium oxide nanotubes. Chem. Mater. 18, 3016–3022 (2006)
89. Danier, M.-C., Astruc, D.: Gold nanoparticles: Assembly, supramolecular chemistry, quan-
tum size related properties, applications toward biology, catalysis, and nanotechnology.
Chem. Rev. 104, 293–346 (2004)
5 Optical Properties of Nanoscale Transition Metal Oxides 115

90. Mai, L.W., Lao, C.S., Hu, B., Qi, Y.Y., Chen, W., Gu, E.D., Wang, Z.L.: Structure and
electrical transport for single-crystal NH4 V3 O8 nanobelts. J. Phys. Chem. B 110,
18138–18141 (2006)
91. Szlufarska, I., Nakano, A., Vashishta, P.: A crossover in the mechanical response of
nanocrystalline ceramics. Science 309, 911–914 (2005)
92. Wu, B., Heidelberg, A., Boland, J.J.: Mechanical properties of ultrahigh-strength gold
nanowires. Nat. Mater. 4, 525–529 (2003)
93. Tenne, R., Margulis, L., Genut, M., Hodes, G.: Polyhedral and cylindrical structures of
tungsten disulphide. Nature 360, 444–446 (1992)
94. Margulis, L., Salitra, G., Tenne, R.: Nested fullerene-like structures. Nature 365, 113–114
(1993)
95. Parilla, P.A., Dillon, A.C., Jones, K.M., Riker, G., Schulz, D.L., Ginley, D.S., Heben, M.J.:
The first true inorganic fullerenes? Nature 397, 114 (1999)
96. Yao, B.D., Chan, Y.F., Zhang, X.Y., Zhang, W.F., Yang, Z.T., Wang, N.: Formation
mechanism of TiO2 nanotubes. Appl. Phys. Lett. 82, 281–283 (2003)
97. Remskar, M., Mrzel, A., Skraba, Z., Jesih, A., Ceh, M., Demsar, J., Stadelmann, P., Lévy, F.,
Mihailovic, D.: Self-assembly of subnanometer-diameter single-wall MoS2 nanotubes. Sci-
ence 292, 479–481 (2001)
98. Mickelson, W., Aloni, S., Han, W.-Q., Cumings, J., Zettl, A.: Packing C60 in boron nitride
nanotubes. Science 300, 467–469 (2003)
99. Levy, P., Leyva, A.G., Troiani, H.E., Sánchez, R.D.: Nanotubes of rare earth manganese
oxide. Appl. Phys. Lett. 83, 5247–5249 (2003)
100. Cabria, I., Mintmire, J.W.: Stability and electronic structure of phosphorus nanotubes.
Europhys. Lett. 65, 82–88 (2004)
101. Seifert, G., Enyashin, A.N.: Structure, stability, electronic properties of TiO2 nanostructures.
Phys. Stat. Sol. 242, 1361–1370 (2005)
102. Enyanshin, A.N., Seifert, G., Ivanovskii, A.L.: Calculation of the electronic and thermal
properties of C/BN nanotubular heterostructures. Inorg. Mater. 41, 595–603 (2005)
103. Saha-Dasgupta, T., Valenti, R., Capraro, F., Gros, C.: Na2V3O7, a frustrated nanotubular
system with spin-1/2 diamond rings. Phys. Rev. Lett. 95, 107201 (2005)
104. Cao, J., Choi, J., Musfeldt, J.L., Lutta, S., Whittingham, M.S.: Effect of sheet distance on the
optical properties of vanadate nanotubes. Chem. Mater. 16, 731–736 (2004)
105. Cao, J., Choi, J., Musfeldt, J.L., Lutta, S., Whittingham, M.S.: Vibrational response of
vanadium oxide nanotubes: Exploring sheet distance effects and variable temperature
properties. In: Sahu, S.N., Choudhury, R.K., Jena, P. (eds.) Nano-Scale Materials: From
Science to Technology. Nova Science, New York (2006)
106. Baruah, T., Zope, R.R., Richardson, S.L., Pederson, M.R.: Electronic structure and rebonding
in the onionlike As@Ni12@As20. Phys. Rev. B 68, 241404 (2003)
107. Chen, S.H., Wang, Z.L., Ballato, J., Foulger, S.H., Carroll, D.L.: Monopod, bipod, tripod and
tetrapod gold nanocrystals. J. Am. Chem. Soc. 125, 6186–16187 (2003)
108. Kong, X.Y., Wang, Z.L.: Spontaneous polarization-induced nanohelixes, nanosprings,
nanorings of piezoelectric nanobelts. Nano Lett. 3, 1625–1631 (2003)
109. Dames, C., Poudel, B., Wang, W.Z., Huang, J.Y., Ren, Z.F., Sun, Y., Oh, J.I., Opeil, C.,
Naughton, M.J., Chen, G.: Low-dimensional phonon specific heat of titanium dioxide
nanotubes. Appl. Phys. Lett. 87, 031901 (2005)
110. Muhr, J.-J., Krumeich, K., Schonholzer, U.P., Bieri, F., Neiderberger, M., Gauckler, L.J.,
Nesper, R.: Vanadium oxide nanotubes: a new flexible vanadate nanophase. Adv. Mater. 12,
231–234 (2000)
111. Worle, M., Krumeich, F., Bieri, F., Muhr, H.-J., Nesper, R.: Flexible V7O16 layers as the
common structural element of vanadium oxide nanotubes and a new crystalline vanadate. Z.
Anorg. Allg. Chem. 628, 2778–2784 (2002)
112. Friedman, J.R., Sarachik, M.P., Tejada, J., Ziolo, R.: Macroscopic measurement of resonant
magnetization tunneling in high-spin molecules. Phys. Rev. Lett. 76, 3830–3833 (1996)
116 J.L. Musfeldt

113. Tasiopoulos, A.J., Vinslava, A., Wernsdorfer, W., Abboud, K.A., Christou, G.: Giant single-
molecule magnets: A Mn84 torus and its supramolecular nanotubes. Angew. Chem. Int. Ed.
43, 2117–2121 (2004)
114. Iijima, S.: Helical microtubules of graphitic carbon. Nature 354, 56–58 (1991)
115. Bethune, D.S.: Cobalt-catalysed growth of carbon nanotubes with single-atomic-layer walls.
Nature 363, 605–607 (1993)
116. Geim, A.K., Novoselov, K.S.: The rise of grapheme. Nat. Mater. 6, 183–191 (2007)
117. Hone, J., Batlogg, B., Benes, Z., Johnson, A.T., Fischer, J.E.: Quantized phonon spectrum of
single-wall carbon nanotubes. Science 289, 1730–1733 (2000)
118. Sun, C.Q., Pan, L.K., Li, C.M., Li, S.: Size-induced acoustic hardening and optic softening of
phonons in InP, CeO2, SnO2, CdS, Ag, Si nanostructures. Phys. Rev. B 72, 134301 (2005)
119. Bersani, D., Lottici, P.P., Ding, X.-Z.: Phonon confinement effects in the Raman scattering by
TiO2 nanocrystals. Appl. Phys. Lett. 72, 73–75 (1998)
120. Luttrell, R.D., Brown, S., Cao, J., Musfeldt, J.L., Rostenveig, R., Tenne, R.: Understanding
the dynamics of bulk vs. nanoscale WS2: Local strain and charging effects. Phys. Rev. B 73,
035410 (2006)
121. Brown, S., Musfeldt, J.L., Mihut, I., Betts, J.B., Migliori, A., Zak, A., Tenne, R.: Bulk vs.
nanoscale WS2: Finite size effects and solid state lubrication. Nano Lett. 7, 2365–2369 (2007)
122. Smith, M.B., Page, K., Siegrist, T., Redmond, P.L., Walter, E.C., Seshadri, R., Brus, L.E.,
Steigerwald, M.L.: Crystal structure and the paraelectric-to-ferroelectric phase transition of
nanoscale BaTiO3. J. Am. Chem. Soc. 130, 6955–6963 (2008)
123. Qazilbash, M.M., Brehm, M., Chae, B.-G., Ho, P.-C., Andreev, G.O., Kim, B.-J., Yun, S.J.,
Balatsky, A.V., Maple, M.B., Keilmann, F., Kim, H.-T., Basov, D.N.: Mott transition in VO2
revealed by infrared spectroscopy and nano-imaging. Science 318, 1750–1753 (2007)
124. Filippetti, A., Hill, N.A.: Magnetic stress as a driving force of structural distortions: the case
of CrN. Phys. Rev. Lett. 85, 5166–5169 (2000)
125. Sun, Z., Chuang, Y.-D., Fedorov, A.V., Douglas, J.F., Reznik, D., Weber, F., Aliouane, N.,
Argyriou, D.N., Zheng, H., Mitchell, J.F., Kimura, T., Tokura, Y., Revcolevschi, A., Dessau,
D.S.: Quasiparticle-like peaks, kinks, electron-phonon coupling at the (p,0) regions in the
CMR oxide La22xSr1+2xMn2O7. Phys. Rev. Lett. 97, 056401 (2006)
126. Mazin, I.I., Johannes, M.D.: A critical assessment of the superconducting pairing symmetry
in NaxCoO2 · yH2O. Nat. Phys. 1, 91–93 (2005)
127. Monteverde, M., Nunez-Regueiro, M., Rodado, N., Regan, K.A., Hayward, M.A., He, T.,
Lourerir, S.M., Cava, R.J.: Pressure dependence of the superconducting transition tempera-
ture of magnesium diboride. Science 292, 75–77 (2001)
128. Homes, C.C., Vogt, T., Shapiro, S.M., Wakimoto, S., Ramirez, A.P.: Optical response of high
dielectric constant perovskite-related oxide. Science 293, 673–676 (2001)
129. Lake, B., Tennant, D.A., Frost, C.D., Nagler, S.E.: Quantum criticality and universal scaling
of a quantum antiferromagnets. Nat. Mater. 4, 329–334 (2005)
130. Feynman, R.P.: There’s plenty of room at the bottom. http://www.zyvex.com/nanotech/
feynman.html (1959).
131. Aviram, A.: Molecules for memory, logic, amplification. J. Am. Chem. Soc. 110, 5687–5692
(1988)
132. Wang, Y.Z., Gebler, D.D., Blatchford, J.W., Jessen, S.W., Wang, H.L., Epstein, A.J.: AC
light emitting device based on conjugated polymers. Appl. Phys. Lett. 68, 894–896 (1996)
133. Lui, C., Pan, H., Fox, M.A., Bard, A.J.: High-density nanosecond charge trapping in thin
films of the photoconductor ZnODEP. Science 261, 897–899 (1993)
134. Balasubramanian, K., Burghard, M., Kern, K., Scolari, M., Mews, M.: Photocurrent imaging
of charge transport barriers in carbon nanotube devices. Nano Lett. 5, 507–510 (2005)
135. Leuenberger, M.N., Loss, D.: Quantum computing in molecular magnets. Nature 410,
789–793 (2001)
136. Kagan, C.R., Mitzi, D.B., Dimitrakopoulos, C.D.: Organic-inorganic hybrid materials as
semiconducting channels in thin-film field-effect transistors. Science 286, 945–947 (1999)
5 Optical Properties of Nanoscale Transition Metal Oxides 117

137. Sales, B.C., Mandrus, D., Williams, R.K.: Filled skutterudites antimonides: a new class of
thermoelectric materials. Science 272, 1325–1328 (1996)
138. Rueckes, T., Kim, K., Joselvich, E., Tseng, G.Y., Cheung, C.L., Lieber, C.M.: Carbon
nanotube-based non-volatile random access memory for molecular computing. Science
289, 94–97 (2000)
139. Krusin-Elbaum, L., Newns, D.M., Zeng, H., Derycke, V., Sun, J.Z., Sandstrom, R.: Room-
temperature ferromagnetic nanotubes controlled by electron or hole doping. Nature 431,
672–676 (2004)
140. Rothschild, A., Cohen, S.R., Tenne, R.: WS2 nanotubes as tips in scanning probe microscopy.
Appl. Phys. Lett. 75, 4025–4027 (1999)
141. Maeda, K., Eguchi, M., Youngblood, W.J., Mallouk, T.E.: Niobium oxide nanoscrolls as
building blocks for dye-sensitized hydrogen production from water under visible light
irradiation. Chem. Mater. 20, 6770–6778 (2008)
142. Rapoport, L., Fleischer, N., Tenne, R.: Fullerene-like WS2 nanoparticles: superior lubricants
for harsh conditions. Adv. Mater. 15, 651–655 (2003)
143. Nemanic, V., Zumer, M., Zajec, B., Pahor, J., Remskar, M., Mrzel, A., Panjan, P., Mihailovic,
D.: Field-emission properties of molybdenum disulfide nanotubes. Appl. Phys. Lett. 82,
4573–4575 (2003)
144. Hill, S., Edwards, R.S., Aliaga-Alcalde, N., Christou, G.: Quantum coherence in an
exchange-coupled dimer of single molecule magnets. Science 302, 1015–1018 (2003)
145. Wang, J., Stucky, G.D.: Mesostructured composite materials for multibit-per-site optical data
storage. Adv. Funct. Mater. 14, 409–415 (2004)
146. Li, S., Yu, Z., Yen, S.-F., Tang, W.C., Burke, P.J.: Carbon nanotube transistor operation at
2.6 GHz. Nano Lett. 4, 753–756 (2004)
147. Jacoby, M.: Single-nanotube photodetector. Chem. Eng. News 81(27), 5 (2003)
148. Moore, J.G., Lochner, E.J., Ramsey, C., Dalal, N.S., Stiegman, A.E.: Transparent, superpar-
amagnetic KCo[Fe(CN)6]-silica nanocomposites with tunable photomagnetism. Angew.
Chem. Int. Ed. 42, 2741–2743 (2003)
149. Novoselov, K.S., Jiang, Z., Zhang, Y., Morozov, S.V., Stormer, H.L., Zeitler, U., Maan, J.C.,
Boebinger, G.S., Kim, P., Geim, A.K.: Room-temperature quantum Hall effect in grapheme.
Science 315, 1379 (2007)
150. Zŭtić, I., Fabian, J., Das Sarma, S.: Spintronics: fundamentals and applications. Rev. Mod.
Phys. 76, 323–410 (2004)
151. Cage, B., Russek, S.E., Shoemaker, R., Barker, A.J., Stoldt, C., Ramachandaran, V., Dalal, N.
S.: The utility of the single-molecule magnet Fe8 as a magnetic resonance imaging contrast
agent over a broad range of concentration. Polyhedron 26(12), 2413–2419 (2007)
152. Cao, G., Durairaj, V., Chikara, S., DeLong, L.E., Schlottmann, P.: Observation of strong spin
valve effect in bulk Ca3(Ru1-xCrx)2O7. Phys. Rev. Lett. 100, 016604 (2008). 159902
153. Poudel, B., Hao, Q., Ma, Y., Lan, Y., Minnich, A., Yu, B., Yan, X., Wang, D., Muto, A.,
Vashaee, D., Chen, X., Liu, J., Dresselhaus, M.S., Chen, G., Ren, Z.: High thermoelectric
performance of nanostructured bismuth antimoney telluride bulk alloys. Science 320,
634–638 (2008)
154. Bibes, M., Barthélémy, A.: Multiferroics: Towards a magnetoelectric memory. Nat. Mater. 7,
425–426 (2008)
155. Chapline, M.G., Wang, S.X.: Room-temperature spin filtering in a CoFe2O4/MgAl2O4/Fe3O4
magnetic tunnel barrier. Phys. Rev. B 74, 014418 (2006)
156. Wooten, F.: Optical Properties of Solids. Academic, New York (1972)
157. Pankove, J.I.: Optical Processes in Semiconductors. Dover, New York (1971)
158. Love, S.P., Worl, L.A., Donohoe, R.J., Huckett, S.C., Swanson, B.I.: Origin of the fine
structure in the vibrational spectrum of [Pt(C2H8N2)2] [Pt(C2H8N2)2] (ClO4)4: vibrational
localization in a quasi-1D system. Phys. Rev. B 46, 813–816 (1992)
118 J.L. Musfeldt

159. Jones, B.R., Varughese, P.A., Pigos, J.M., Landee, C.P., Turnbull, M.M., Olejniczak, I., Carr,
G.L.: Lattice dynamics of the 1D S ¼ 1/2 Heisenberg antiferromagnet CuPzN. Chem. Mater.
13, 2127–2134 (2001)
160. Brown, S., Cao, J., Musfeldt, J.L., Conner, M.M., McConnell, A.C., Southerland, H.I.,
Manson, J.L., Schlueter, J.A., Phillips, M.D., Turnbull, M.M., Landee, C.P.: Hydrogen
bonding and multiphonon structure in copper pyrazine coordination polymers. Inorg.
Chem. 46, 8577–8583 (2007)
161. Sun, Q.-C., Xu, X.S., Vergara, L.I., Rosentsveig, R., Musfeldt, J.L.: Dynamical charge and
structural strain in inorganic fullerne-like MoS2 nanoparticles. Phys. Rev. B 79, 205205
(2009)
162. MnO, Q.C. Sun, X.S. Xu, S.N. Baker, A.D. Christianson, and J.L. Musfeldt, Chem. Mater. 23,
2956 (2011)
163. Magneto-elastic coupling in bulk and nanoscale MnO, Q.C. Sun, S.N. Baker, A.D.
Christianson, and J.L. Musfeldt, Phys. Rev. B 84, 014301 (2011)
162. Xu, X.S., Brinzari, T.V., McGill, S., Zhou, H.D., Weibe, C.R., Musfeldt, J.L.: Absence of
spin liquid behavior: Magneto-optical study of Nd3Ga5SiO14. Phys. Rev. Lett. 103, 267402
(2009)
163. Forró, L., Mihály, L.: Electronic properties of doped fullerenes. Rep. Prog. Phys. 64, 649–699
(2001)
164. Kortus, J., Pederson, M.R.: Magnetic and vibrational properties of the uniaxial Fe13O8
cluster. Phys. Rev. B 62, 5755–5759 (2000)
165. Zhu, Z.-T., Musfeldt, J.L., Kamarás, K., Adams, G.B., Page, J.B., Kashevarova, L.S.,
Rakhmanina, A.V., Davydov, V.A.: Far-infrared vibrational properties of linear C60
polymers: A comparison between neutral and charged materials. Phys. Rev. B 67, 045409
(2003)
166. Fateley, W.G., Dollish, F.R., McDevitt, N.T., Bentley, F.F.: Infrared and Raman Selection
Rules for Molecular and Lattice Vibrations: The Correlation Method. Wiley, New York
(1972)
167. Barath, H., Kim, M., Cooper, S.L., Abbamonte, P., Fradkin, E., Mahns, I., R€ ubhausen, M.,
Aliouane, N., Argyriou, D.N.: Domain fluctuations near the field-induced incommensurate-
commensurate phase transition of TbMnO3. Phys. Rev. B 78, 134407 (2008)
168. Gasparov, L.V., Brown, K.G., Wint, A.C., Tanner, D.B., Berger, H., Margaritondo, G., Gaal,
R., Forro, L.: Phonon anomaly at the charge ordering transition in 1T-TaS2. Phys. Rev. B 66,
094301 (2002)
169. Musfeldt, J.L., Brown, S., Mazuumdar, S., Clay, R.T., Mas-Torent, M., Rovira, C., Dias, J.C.,
Henriques, R.T., Almeida, M.: Infrared investigation of the charge ordering pattern in the
organic spin ladder candidate (DTTTF)2Cu(mnt)2. Solid State Sci. 10, 1740–1744 (2008)
170. Choi, J., Musfeldt, J.L., He, J., Jin, R., Thompson, J.R., Mandrus, D., Lin, X.N., Bondarenko,
V.A., Brill, J.W.: Probing localization effects in Li0.9Mo6O17: An optical properties investi-
gation. Phys. Rev. B 69, 085120 (2004)
171. Barath, H., Kim, M., Karpus, J.F., Cooper, S.L., Abbamonte, P., Fradkin, E., Morosan, E.,
Cava, R.J.: Quantum and classical mode softening near the charge density wave-supercon-
ductor transition of CuxTiSe2. Phys. Rev. Lett. 100, 106402 (2008)
172. Choi, J., Musfeldt, J.L., Whangbo, M.H., Galy, J., Millet, P.: Optical investigation of
Na2V3O7 nanotubes. Chem. Mater. 14, 924–930 (2002)
173. Ernst, G., Broholm, C., Kowach, G.R., Ramirez, A.P.: Phonon density of states and negative
thermal expansion in ZrW2O8. Nature 396, 147–149 (1998)
174. Frey, G.L., Elani, S., Homyonfer, M., Feldman, Y., Tenne, R.: Optical absorption spectra of
inorganic fullerene-like MS2 (M ¼ Mo, W). Phys. Rev. B 57, 6666–6671 (1998)
175. Bommeli, F., Degiogi, L., Wachter, R., Legez, Ö., Janossy, A., Oszlanyi, G., Chauvet, O.,
Forro, L.: Metallic conductivity and metal-insulator transition in (AC60)n (A ¼ K, Rb, Cs)
linear polymer fullerides. Phys. Rev. B 51, 14794–14797 (1995)
5 Optical Properties of Nanoscale Transition Metal Oxides 119

176. Hicks, L.D., Dresselhaus, M.S.: Effect of quantum-well structures on the thermodynamic
figure of merit. Phys. Rev. B 47, 12727–12731 (1993)
177. Hasegawa, T., Akutagawa, T., Nakamura, T., Mochida, T., Kondo, R., Kagoshima, S., Iwasa,
Y.: Neutral-ionic phase transition of (BEDT-TTF)(ClMeTCNQ) under pressure. Phys. Rev.
B 64, 085106 (2001)
178. Schrama, J.M., Rzepniewski, E., Edwards, R.S., Singleton, J., Ardavan, A., Kurmoo, M.,
Day, P.: Millimeter-wave magneto-optical determination of the anisotropy of the
superconducting order parameter in the molecular superconductor k-(BEDTTTF)2Cu
(NCS)2. Phys. Rev. Lett. 83, 3041–3044 (1999)
179. Nuttall, C.J., Hayashi, Y., Yamazaki, K., Mitani, T., Iwasa, Y.: Dipole dynamics in the
endohedral metallofullerene La@C82. Adv. Mater. 14, 293–296 (2002)
180. Islam, M.F., Milkie, D.E., Kane, C.L., Yodh, A.G., Kikkawa, J.M.: Direct measurement of
the polarized optical absorption cross section of single-wall carbon nanotubes. Phys. Rev.
Lett. 93, 037404 (2004)
181. Tsvetkov, A.A., Mena, F.P., van Loosdrecht, P.H.M., van der Marel, D., Ren, Y., Nugroho,
A.A., Menovsky, A.A., Elfimov, I.S., Sawatzky, G.A.: Structural, electronic, and magneto-
optical properties of YVO3. Phys. Rev. B 69, 075110 (2004)
182. Turner, G.M., Beard, M.C., Schmuttenmaer, C.A.: Carrier localization and cooling in dye-
sensitized nanocrystalline titanium dioxide. J. Phys. Chem. B 106, 11716 (2002)
183. Basov, D.N., Homes, C.C., Singley, E.J., Strongin, M., Timusk, T., Blumberg, F., van der
Marel, D.: Unconventional energetics of the pseudogap state and superconducting state in the
high Tc cuprates. Phys. Rev. B 63, 134514 (2001)
184. Degiorgi, L.: The electrodynamic response of heavy-electron compounds. Rev. Mod. Phys.
71, 687–734 (1999)
185. Dora, B., Vanyolos, A., Maki, K., Virosztek, A.: Gapped optical excitations from gapless
phases: imperfect nesting in unconventional density waves. Phys. Rev. B 71, 245101 (2005)
186. Jones, B.R., Olejniczak, I., Dong, J., Pigos, J.M., Zhu, Z., Garlach, A.D., Musfeldt, J.L.,
Schlueter, J.A., Koo, H.-J., Whangbo, M.-H., Nixon, P.G., Winter, R.W., Gard, G.L.: Optical
properties of b00 -(ET)2SF5RSO3 (R ¼ CH2CF2, CHFCF2, CHF): changing electronic
properties via chemical tuning of the counterion. Chem. Mater. 12, 2490–2495 (2000)
187. Bockrath, M., Cobden, D.H., Lu, J., Rinzler, A.G., Smalley, R.E., Balents, L., McEuen, P.L.:
Luttinger-liquid behaviour in carbon nanotubes. Nature 397, 598–601 (1999)
188. Rauf, H., Pichler, T., Knupfer, M., Fink, J., Kataura, H.: Transition from a Tomonaga-
Luttinger liquid to a Fermi liquid in potassium-intercalated bundles of single-wall carbon
nanotubes. Phys. Rev. Lett. 93, 096805 (2004)
189. Ishii, H., Kataura, H., Shiozawa, H., Yoshioka, H., Otsubo, H., Takayama, Y., Miyahara, T.,
Suzuki, S., Achiba, Y., Nakatake, M., Narimura, T., Higashiguchi, M., Shimada, K.,
Namatame, H., Taniguchi, M.: Direct observation of Tomonaga-Luttinger-liquid state in
carbon nanotubes at low temperatures. Nature 426, 540–544 (2003)
190. Zhu, Z.-T., Musfeldt, J.L., Teweldemedhin, Z.S., Greenblatt, M.: Anisotropic ab-plane
optical response of the charge-density-wave superconductor P4W14O50. Phys. Rev. B 65,
214519 (2002)
191. Shin, S., Suga, S., Taniguchi, M., Fujisawa, M., Kanzaki, H., Fujimori, A., Daimon, H., Ueda,
Y., Kosuge, K., Kachi, S.: Vacuum-ultraviolet reflectance and photoemission study of the
metal-insulator phase transitions in VO2, V6O13, and V2O3. Phys. Rev. B 41, 4993–5009
(1990)
192. Benckiser, E., R€uckamp, R., M€ uller, T., Taetz, T., Moller, A., Nugroho, A., Palstra, T.T.M.,
Uhrig, G.S., Gr€ uninge, M.: Collective orbital excitations in orbitally ordered YVO3 and
HoVO3. New J. Phys. 10, 053027 (2008)
193. Dawlaty, J.M., Shivaraman, S., Strait, J., George, P., Chandrashekhar, M., Raman, F.,
Spencer, M.G., Veksler, D., Chen, Y.: Measurement of the optical absorption of epitaxial
graphene from terahertz to visible. Appl. Phys. Lett. 93, 131905 (2008)
120 J.L. Musfeldt

194. Xu, X.S., Angst, M., Brinzari, T.V., Hermann, R.P., Musfeldt, J.L., Christianson, A.D.,
Mandrus, D., Sales, B.C., McGill, S., Kim, J.-W., Islam, Z.: Charge order, dynamics,
magnetostructural transition in multiferroic LuFe2O4. Phys. Rev. Lett. 101, 227602 (2008)
195. Xu, X.S., Sun, Q.-C.,Rosentsveig, R., Musfeldt, J.L.: Musfeldt, Evaluation of Born and local
effective charges in unoriented materials from vibrational spectra, Phys. Rev. B 80, 014303
(2009)
196. Hart, G.L.W., Pickett, W.E., Kurmaev, E.Z., Hartmann, D., Neumann, M., Moewes, A.,
Ederer, D.L., Endoh, R., Taniguchi, K., Nagata, S.: Electronic structure of Cu1-xNixRh2S4 and
CuRh2Se4: band structure calculations, x-ray photoemission, and fluorescence measurements.
Phys. Rev. B 61, 4230–4237 (2000)
197. Reddy, B.V., Khanna, S.N., Jena, P.: Structure and magnetic ordering in Cr8 and Cr13
clusters. Phys. Rev. B 60, 15597–15600 (1999)
198. Islam, Z., Detlefs, C., Song, C., Goldman, A.I., Antropoc, C., Harmon, B.N., Bud’ko, S.L.,
Wiener, T., Canfield, P.C., Wermeille, D., Finkelstein, K.D.: Effects of band filling on
magnetic structures: the case of RNi2Ge2. Phys. Rev. Lett. 83, 2817–2820 (1999)
199. Pederson, M.R., Khanna, S.N.: Magnetic anisotropy barrier for spin tunneling in Mn12O12
molecules. Phys. Rev. B 60, 9566–9572 (1999)
200. Zeng, Z., Guenzburger, D., Ellis, D.E.: Electronic structure, spin couplings, hyperfine
properties of nanoscale molecular magnets. Phys. Rev. B 59, 6927–6937 (1999)
201. Haule, K., Oudovenko, V., Savrasov, S.Y., Kotliar, G.: The a ! g transition in Ce:
A theoretical view from optical spectroscopy. Phys. Rev. Lett. 94, 036401 (2005)
202. Boukhvalov, D.W., Lichtenstein, A.I., Dobrovitski, V.V., Katsnelson, M.I., Harmon, B.N.,
Mazurenko, V.V., Anisimov, V.I.: Electronic structure and exchange interactions in V15
magnetic molecules: LDA + U results. Phys. Rev. B 70, 054417 (2004)
203. Xiang, H.J., Whangbo, M.-H.: Charge order and the origin of giant magnetocapacitance in
LuFe2O4. Phys. Rev. Lett. 98, 246403 (2007)
204. Stoltzfus, M.W., Woodward, P.M., Seshadri, R., Klepeis, J.-H., Bursten, B.: Structure and
bonding in SnWO4, PbWO4, and BiVO4: Lone pairs vs. inert pairs. Inorg. Chem. 46(10),
3839–3850 (2007)
205. Marianetti, C.A., Kotliar, G., Ceder, G.: A first-order Mott transition in LixCO2. Nat. Mater.
3, 627–631 (2004)
206. Barbour, A., Luttrell, R.D., Choi, J., Musfeldt, J.L., Zipse, D., Dalal, N.S., Boukhvalov, D.S.,
Dobrovitski, V.V., Katsnelson, M.I., Lichtenstein, A.I., Harmon, B.N., K€ ogerler, P.: Under-
standing the gap in polyoxovanadate molecule-based magnets. Phys. Rev. B 74, 014411
(2006)
207. Oudovenko, V.X., Palsson, G., Savrasoc, S.Y., Haule, J., Kotliar, G.: Calculations of optical
properties in strongly correlated materials. Phys. Rev. B 70, 125112 (2005)
208. Hoinkis, M., Sing, M., Schafer, J., Klemm, M., Horn, S., Benthien, H., Jeckelmann, E., Saha-
Dasgupta, T., Pisani, L., Valenti, R., Claessen, R.: Electronic structure of the S ¼ 1/2 quan-
tum magnet TiOCl. Phys. Rev. B 72, 125127 (2005)
209. Kasinathan, D., Kunes, J., Koepernik, K., Diaconu, C.V., Martin, R.L., Prodan, I.D., Scuseria,
G.E., Spaldin, N., Petit, L., Schulthess, T.C., Pickett, W.E.: Mott transition of MnO under
pressure: A comparison of correlated band theories. Phys. Rev. B 74, 195110 (2006)
210. Wesolowski, R., Haraldsen, J.T., Cao, J., Musfeldt, J.L., Olejniczak, I., Wang, Y.J.,
Schlueter, J.A.: Understanding the totally symmetric intramolecular vibrations in type k-
phase organic superconductors. Phys. Rev. B 71, 214514 (2005)
211. Egami, T.: Electron-phonon coupling in high-Tc superconductors. In: M€ uller, K.A.,
Bussmann-Holder, A. (eds.) Superconductivity in Complex Systems Structure and Bonding,
vol. 114, pp. 267–286. Springer, Berlin (2005)
212. Long, V.C., Musfeldt, J.L., Kamrás, K., Adams, G.B., Page, J.B., Iwasa, Y., Mayo, W.E.: Far-
infrared vibrational properties of high-pressure high-temperature C60 polymers and the C60
dimer. Phys. Rev. B 61, 13191–13201 (2000)
5 Optical Properties of Nanoscale Transition Metal Oxides 121

213. Cronin, S.B., Swan, A.K., Unl€€ u, M.S., Goldberg, B.B., Dresselhaus, M.S., Tinkham, M.:
Measuring the uniaxial strain of individual single-wall carbon nanotubes: resonance Raman
spectra of atomic-force-microscope modified single-wall nanotubes. Phys. Rev. Lett. 93,
167401 (2004)
214. FitzGerald, S.A., Sievers, A.J.: Far-infrared properties of C60 and C70 compacts. J. Chem.
Phys. 101, 7283–7289 (1994)
215. Taylbayev, D., Mihály, L., Zhou, J.: Antiferromagnetic resonance in LaMnO3 at low temper-
ature. Phys. Rev. Lett. 93, 017202 (2004)
216. Cao, J., Haraldsen, J.T., Brown, S., Musfeldt, J.L., Thompson, J.R., Zvyagin, S., Krzytek, J.,
Whangbo, M.-H., Nagler, S.E., Torardi, C.C.: Understanding low-energy magnetic
excitations and hydrogen bonding in VOHPO4 · H2O. Phys. Rev. B 72, 214421 (2005)
217. Valdés-Aguilar, R., Sushkov, A.B., Zhang, C.L., Choi, Y., Cheong, J.S.-W., Drew, H.D.:
Colossal magnon-phonon coupling in multiferroic Eu0.75T0.25MnO3. Phys. Rev. B 76,
060404(R) (2007)
218. Room, T., H€uvonen, D., Nagel, U., Wang, Y.-J., Kremer, R.K.: Low-energy excitations and
dynamic Dzaloshinskii-Moriya interaction in a -NaV2O5 by far-infrared spectroscopy. Phys.
Rev. B 69, 144410 (2004)
219. Room, T., Nagel, U., Lippmaa, E., Kageyama, H., Onizuka, K., Ueda, Y.: Far-infrared study
of the two-dimensional dimer spin system SrCu(BO3)2. Phys. Rev. B 61, 14342 (2004)
220. Room, T., H€uvonen, D., Nagel, U., Hwang, J., Timusk, T., Kageyama, H.: Far-infrared
spectroscopy of spin excitations and Dzyaloshinskii-Moriya interactions in a Shastry-
Sutherland compound SrCu(BO3)2. Phys. Rev. B 70, 144417 (2004)
221. Sushkov, A.B., Musfeldt, J.L., Wang, Y.J., Achey, R.M., Dalal, N.S.: Spin-vibrational
coupling in the far-infrared spectrum of Mn12-acetate. Phys. Rev. B 66, 144430 (2002)
222. Pederson, M.R., Bernstein, N., Kortus, J.: Fourth-order magnetic anisotropy and tunnel
splittings in Mn12 from spin-orbit-vibron interactions. Phys. Rev. Lett. 89, 097202 (2002)
223. Sushkov, A.B., Tchernyshyov, O., Ratcliff II, W., Cheong, S.W., Drew, H.D.: Probing spin
correlations with phonons in the strongly frustrated magnet ZnCr2 O4. Phys. Rev. Lett. 94,
137202 (2005)
224. Basov, D.N., Averitt, R.D., van der Marel, D., Dressel, M., Haule, K.: Electrodynamics of
correlated electron materials. Rev. Mod. Phys.
225. Yoffe, A.D.: Low-dimensional systems: quantum size effects and electronic properties of
semiconductor microcrystallites (zero-dimensional systems) and some quasi-two-dimensional
systems. Adv. Phys. 42, 173–266 (1993)
226. Kayanuma, Y.: Quantum-size effects of interacting electrons and holes in semiconductor
microcrystals with spherical shape. Phys. Rev. B 38, 9797–9805 (1988)
227. Brus, L.: Electronic wave functions in semiconductor clusters: experiment and theory.
J. Phys. Chem. 90, 2555–2560 (1986)
228. Nosaka, Y.: Finite depth spherical well model for excited states of ultrasmall semiconductor
particles: an application. J. Phys. Chem. 95, 5054–5058 (1991)
229. Franceschetti, A., Zunger, A.: Direct pseudopotential calculation of exciton Coulomb and
exchange energies in semiconductor quantum dots. Phys. Rev. Lett. 78, 915–918 (1997)
230. Uozumi, T., Kayanuma, Y.: Excited states of an electron-hole pair in spherical quantum dots
and their optical properties. Phys. Rev. B 65, 165318 (2002)
231. Senger, R.T., Bajaj, K.K.: Optical properties of confined polaronic excitons in spherical ionic
quantum dots. Phys. Rev. B 68, 045313 (2003)
232. Kim, S.-W., Fujita, S., Fujita, S.: Self-organized ZnO quantum dots on SiO2/Si substrates by
metalorganic chemical vapor deposition. Appl. Phys. Lett. 81, 5026–5038 (2002)
233. Satoh, N., Nakashima, T., Kamikura, K., Tamamoto, K.: Quantum size effect in TiO2
nanoparticles prepared by finely controlled metal assembly on dendrimer templates. Nat.
Nanotechnol. 3, 106–111 (2008)
234. Dong, W., Zhu, C.: Optical properties of surface-modified CdO nanoparticles. Opt. Mater. 22,
227–233 (2003)
122 J.L. Musfeldt

235. Dong, W., Zhu, C.: Optical properties of surface-modified Bi2O3 nanoparticles. J. Phys.
Chem Sol. 64, 265–271 (2003)
236. Landau, L.D., Lifshitz, E.M.: Statistical Physics Part 1, Vol. 5, Course of Theoretical Physics,
3rd edn. Pergamon, London (1994)
237. Batra, I.P., Wurfel, P., Silverman, B.D.: Phase transition, stability, depolarization Field in
ferroelectric thin films. Phys. Rev. B 8, 3257 (1973)
238. Batra, I.P., Wurfel, P., Silverman, B.D.: New type of first-order phase transition in ferroelec-
tric thin films. Phys. Rev. Lett. 30, 384 (1973)
239. Tashiro, H., Graybeal, J.M., Tanner, D.B., Nicol, E.J., Carbotte, J.P., Carr, G.L.: Unusual
thickness dependence of the superconducting transition of a-MoGe thin films. Phys. Rev. B
78, 014509 (2008)
240. Ishikawa, K., Yoshikawa, J., Okada, N.: Size effect on the ferroelectric phase transition in
PbTiO3 ultrafine particles. Phys. Rev. B 37, 5852 (1988)
241. Uchino, K., Sadanaga, E., Hirose, T.: Dependence of the crystal structure on particle size in
barium titanate. J. Am. Ceram. Soc. 72, 1555–1558 (1989)
242. Zhong, W.L., Jiang, B., Zhang, P.L., Ma, J.M., Cheng, H.M., Yang, Z.H., Li, L.X.: Phase
transition in PbTiO3 ultrafine particles of different sizes. J. Phys. Condens. Matter 5,
2619–2624 (1993)
243. Maxwell-Garnett, J.C.: Colours in metal glasses and in metallic films. Philos. Trans. R. Soc.
Lond. A 203, 385–420 (1904)
244. Maxwell-Garnett, J.C.: Colours in metal glasses and in metallic films. Philos. Trans. R. Soc.
Lond. 205, 237–288 (1906)
245. Bruggeman, D.A.G.: Berechnung verschiedener physikalischer Konstanten von heterogenen
substanzen. Ann. Phys. (Leipzig) 24, 636–679 (1935)
246. Carr, G.L., Perkowitz, S., Tanner, D.B.: Far-infrared properties of inhomogeneous media.
In: Button, K. (ed.) Infrared and Millimeter Waves, 13th edn. Academic, New York (1985)
247. Kim, Y.H., Tanner, D.B.: Far-infrared absorption by aluminum small particles. Phys. Rev. B
39, 3585 (1989)
248. Carr, G.L., Henry, R.L., Russell, N.E., Garland, J.C., Tanner, D.B.: Anomalous far-infrared
absorption in random small-particle composites. Phys. Rev. B 24, 777 (1981)
249. Carr, G.L., Garland, J.C., Tanner, D.B.: Anomalous infrared absorption in granular
superconductors. Phys. Rev. Lett. 50, 1607 (1983)
250. Akima, N., Iwasa, Y., Brown, S., Barbour, A.M., Cao, J., Musfeldt, J.L., Matsui, J., Toyota,
N., Shiraishi, M., Shimoda, H., Zhou, O.: Strong anisotropy in the far infrared absorption
spectra of stretch-aligned single-walled carbon nanotubes. Adv. Mater. 18, 1166–1169 (2006)
251. Cummings, K.D., Garland, J.C., Tanner, D.B.: Optical properties of a small-particle compos-
ite. Phys. Rev. B 30, 4170 (1984)
252. Carr, G.L.: Unpublished results.
253. Qazilbash, M.M., Brehm, M., Andreev, G.O., Frenzel, A., Ho, P.-C., Chae, B.-G., Kim, B.J.,
Yun, S.J., Kim, H.-T., Valatsky, A.V., Shpyrko, O.G., Maple, M.B., Keilmann, F., Basov, D.
N.: Infrared spectroscopy and nano-imaging of the insulator-to-metal transition in vanadium
dioxide. Phys. Rev. B 79, 075107 (2009)
254. Mie, G.: Beitr¨age zur optik tr€
uber medien speziell kolloidaler goldlosungen (contributions to
the optics of diffuse media, especially colloidal metal solutions). Ann. Phys. 25, 377–445
(1908)
255. Bohren, C.F., Huffman, D.R.: Absorption and Scattering of Light by Small Particles. Wiley,
New York (1983)
256. Nome, R.A., Guffey, M.J., Scherer, N.F., Gray, S.K.: Plasmonic interactions and optical
forces between Au bipyramidal nanoparticle dimers. J. Phys. Chem. A 113(16), 4408–4415
(2009)
257. Niklasson, G.A.: Modeling the optical properties of nanoparticles. SPIE Newsroom (2006).
doi:10.1117/2.1200603.0182
258. Maier, S.A.: Plasmonics: Fundamentals and Applications. Springer, New York (2007)
5 Optical Properties of Nanoscale Transition Metal Oxides 123

259. Stroud, D.: Percolation effects and sum rules in the optical properties of composities. Phys.
Rev. B 19, 1783 (1979)
260. Kelley, E.L., Coronado, E., Zhao, L.L., Schatz, G.C.: The optical properties of metal
nanoparticles: The influence of size, shape, dielectric environment. J. Phys. Chem. B 107,
668–677 (2003)
261. Sosa, I.O., Noguez, C., Barrera, R.G.: Optical properties of metal nanoparticles with arbitrary
shapes. J. Phys. Chem. B 107, 6269–6275 (2003)
262. Zhang, Q., Ge, J., Pham, T., Goebl, J., Hu, Y., Lu, Z., Yin, Y.: Reconstruction of silver
nanoplates by uv irradiation: tailored optical properties and enhanced stability. Angew.
Chem. Int. Ed. 121, 3568–3571 (2009)
263. Lassiter, J.B., Knight, M.W., Mirin, N.A., Halas, N.J.: Reshaping the plasmonic properties of
an individual nanoparticle. Nano Lett. 9, 4326–4332 (2009)
264. Bardhan, R., Chen, W., Perez-Torres, C., Bartels, M., Huschka, R.M., Zhao, L., Morosan, E.,
Pautler, R., Joshi, A., Halas, N.J.: Nanoshells engineered for targeted, simultaneous enhance-
ment of magnetic and optical imaging and photothermal therapeutic response. Adv. Func.
Mater. 19(24), 3901–3909 (2009)
265. Schonfeld, B., Huang, J.J., Moss, S.C.: Anisotropic mean-square displacements (MSD) in
single-crystals of 2H and 3R-MoS2. Acta Cryst. B 39, 404–407 (1983)
266. Verble, J.L., Wieting, T.J.: Lattice mode degeneracy in MoS2 and other layer compounds.
Phys. Rev. Lett. 25, 362 (1970)
267. Mott, N.F.: Metal-Insulator Transitions. Taylor & Francis, London (1990)
268. Imada, M., Fujimori, A., Tokura, Y.: Metal-insulator transitions. Rev. Mod. Phys. 70,
1039–1263 (1998)
269. Qazilbash, M.M., Schafgans, A.A., Burch, J.S., Yun, S.J., Chae, B.G., Kim, B.J., Kim, H.T.,
Basov, D.N.: Electrodynamics of the vanadium oxides VO2 and V2O3. Phys. Rev. B 77,
115121 (2008)
270. Brinkman, W.F., Rice, T.M.: Application of Gutzwiller’s variational method to the metal-
insulator transition. Phys. Rev. B 2, 4302 (1970)
271. Katakura, I., Tokunaga, M., Matsuo, A., Kawaguchi, K., Kindo, J., Hitomi, M., Akahoshi, D.,
Kuwahara, H.: Development of high-speed polarizing imaging system for operation in high
pulsed magnetic field. Rev. Sci. Inst. 81, 043701 (2010)
272. MRS Bull. 34(11) (2009)
273. Zavalij, P.Y., Whittingham, M.S.: Structural chemistry of vanadium oxides with open
frameworks. Acta Cryst. B 55, 627–663 (1999)
274. O’Dwyer, C., Navas, D., Lavayen, V., Benavente, E., Santa Ana, M.A., González, G.,
Newcomb, S.B., Sotomayor Torres, C.M.: Nano-Urchin: The formation and structure of
high-density spherical clusters of vanadium oxide nanotubes. Chem. Mater. 18, 3016–3022
(2006)
275. Wang, Y., Cao, G.: Synthesis and enhanced intercalation properties of nanostructured
vanadium oxides. Chem. Mater. 18, 2787–2804 (2006)
276. Krumeich, F., Muhr, H.-J., Niederberger, M., Bieri, F., Schnyder, B., Nesper, R.: Morphology
and topochemical reactions of novel vanadium oxide nanotubes. J. Am. Chem. Soc. 121,
8324–8331 (1999)
277. Niederberger, M., Muhr, H.-J., Krumeich, F., Bieri, F., G€ unther, D., Nesper, R.: Low-cost
synthesis of vanadium oxide nanotubes via two novel non-alkoxide routes. Chem. Mater. 12,
1995–2000 (2000)
278. Reinoso, J.M., Muhr, H.-J., Krumeich, F., Bieri, F., Nesper, R.: Controlled uptake and release
of metal cations by vanadium oxide nanotubes. Helv. Chim. Acta 83, 1724–1733 (2000)
279. Cao, J., Musfeldt, J.L., Mazumdar, S., Chernova, N., Whittingham, M.S.: Pinned low energy
excitation in metal exchanged vanadium oxide nanoscrolls. Nano Lett. 7, 2351–2355 (2007)
280. Sipos, B., Duchamp, M., Magrez, A., Forró, L., Barisić, N., Kis, A., Seo, J.W., Bieri, F.,
Krumeich, F., Nesper, R., Patzke, G.R.: Mechanical and electronic properties of vanadium
oxide nanotubes. J. App. Phys. 105, 074317 (2009)
124 J.L. Musfeldt

281. Mahamuni, S., Borgohain, K., Bendre, B.S., Leppert, B.J., Risbud, S.H.: Spectroscopic and
structural characterization of electrochemically grown ZnO quantum dots. J. Appl. Phys. 85,
2861–2865 (1999)
282. Guo, L., Yang, S., Yang, C., Yu, P., Wang, J., Ge, W., Wong, P.K.W.: Highly monodisperse
polymer-capped ZnO nanoparticles: Preparation and optical properties. Appl. Phys. Lett. 76,
2901–2903 (2000)
283. Kim, K.-K., Koguchi, N., Ok, Y.-W., Seong, T.Y., Park, S.-J.: Fabrication of ZnO quantum
dots embedded in an amorphous oxide layer. Appl. Phys. Lett. 84, 3810–3812 (2004)
284. Du, X.-W., Fu, Y.-S., Sun, J., Han, X., Liu, J.: Complete uv emission of ZnO nanoparticles in
a PMMA matrix. Semicond. Sci. Technol. 21, 1202–1206 (2006)
285. Oftadeh, M., Salavati-Niasari, M., Davar, F.: Synthesis of ZnO nanoparticles and their optical
properties. Int. J. Nanopart. 2, 307–313 (2009)
286. Zheng, M.J., Zhang, L.D., Li, G.H., Shen, W.Z.: Fabrication and optical properties of large-
scale uniform zinc oxide nanowire arrays by one-step electrochemical deposition technique.
Chem. Phys. Lett. 363, 123–128 (2002)
287. Stichtenoth, D., Ronning, C., Niermann, T., Wischmeier, L., Voss, T., Chien, C.-J., Chang,
P.-C., Lu, J.G.: Optical size effects in ultrathin ZnO nanowires. Nanotechnology 18, 435701
(2007)
288. Pal, E., Dekany, I.: Structural, optical, photoelectric properties of indium-doped zinc oxide
nanoparticles prepared in dimethyl sulphoxide. Coll. Surf. A 318, 141–150 (2008)
289. Hilli, S.M., Willander, M.: Optical properties of zinc oxide nanoparticles embedded in
dielectric medium for uv region: numerical simulations. J. Nanopart. Res. 8, 79–97 (2006)
290. Kahn, M.L., Monge, M., Colliére, V., Senocq, F., Maisonnat, A., Chaudrei, B.: Size and
shape-control of crystalline zinc oxide nanoparticles: a new organometallic synthetic method.
Adv. Funct. Mater. 15, 458–461 (2005)
291. Oo, W.M.H., McCluskey, M.D., Lalonde, A.D., Norton, M.G.: Infrared spectroscopy of ZnO
nanoparticles containing CO2 impurities. Appl. Phys. Lett. 86, 073111 (2005)
292. Ischenko, V., Polarz, S., Grote, D., Stavarache, V., Fink, K., Driess, M.: Zinc oxide
nanoparticles with defects. Adv. Funct. Mater. 15, 1945–1954 (2005)
293. Zhao, X., Chen, Z., Luo, Y., Wang, L.: Giant enhanced infrared and orange emissions of ZnO
nanoparticles induced by rich oxygen atmosphere. Solid State Commun. 147, 447–451
(2008)
294. Carrasco, J., Illas, F., Bromley, S.T.: Ultralow-density nanocage-based metal-oxide
polymorphs. Phys. Rev. Lett. 99, 235502 (2007)
295. Lassaletta, G., Fernandez, A., Espinos, J.P., González-Elipe, A.R.: Spectroscopic characteri-
zation of quantum-sized TiO2 supported on silica: influence of size and TiO2-SiO2 interface
composition. J. Phys. Chem. 99, 1484–1490 (1995)
296. Monticone, S., Tufeu, R., Janaev, A.V., Scolan, E., Sanchez, C.: Quantum size effect in TiO2
nanoparticles: does it exist? Appl. Surf. Sci. 162–163, 565–570 (2000)
297. Lin, Y., Wu, G.S., Yuan, X.Y., Zie, T., Zhang, L.D.: Fabrication and optical properties
of TiO2 nanowire arrays made by sol-gel electrophoresis deposition into anodic alumina
membrane. J. Phys. Condens. Matter 15, 2917–2922 (2003)
298. Frova, A., Boddy, P.J., Chen, Y.S.: Electromodulation of the optical constants of rutile in the
uv. Phys. Rev. 157, 700–708 (1967)
299. Jiménez González, A.E., Gelover Santiago, S.: Structural and optoelectronic characterization
of TiO2 films prepared using the sol-gel technique. Semicond. Sci. Techol. 22, 709–716
(2007)
300. Gao, Y., Masuda, Y., Peng, Z.-F., Yonezawa, T., Koumoto, K.: Room temperature deposition
of a TiO2 thin film from aqueous peroxotitanate solution. J. Mater. Chem. 13, 608–613 (2003)
301. Lee, H.-S., Woo, C.-S., Youn, B.-K., Kim, S.-Y., Oh, S.-T., Sung, Y.-E., Lee, H.-I.: Bandgap
modulation of TiO2 and its effect on the activity in photocatallytic oxidation of 2-isopropyl-6-
methyl-4-pyrimidinol. Top. Catal. 35, 255–260 (2005)
5 Optical Properties of Nanoscale Transition Metal Oxides 125

302. Oregan, B., Gratzel, M.: A low-cost, high-efficiency solar-cell based on dye-sensitized
colloidal TiO2 films. Nature 353, 737–740 (1991)
303. Mor, G.P., Prakasam, H.E., Varghese, O.K., Shankar, K., Grimes, C.A.: Vertically oriented
Ti-Fe-O nanotube array films: toward a useful material architecture for solar spectrum water
photoelectrolysis. Nano. Lett. 7, 2356 (2007)
304. Hill, N.A., Rabe, K.M.: First-principles investigation of ferromagnetism and ferroelectricity
in bismuth manganite. Phys. Rev. B 59, 8759–8769 (1999)
305. Seshadri, R., Hill, N.A.: Visualizing the role of Bi 6s “lone pairs” in the off-center distortion
in ferromagnetic BiMnO3. Chem. Mater. 13, 2892–2899 (2001)
306. Baettig, P., Seshadri, R., Spaldin, N.A.: Anti-polarity in ideal BiMnO3. J. Am. Chem. Soc.
129, 9854–9855 (2007)
307. Baettig, P., Seshadri, R., Spaldin, N.A.: Private communication.
308. Basu, S.R., Martin, L.W., Chu, Y.H., Gajek, M., Ramesh, R., Rai, R.C., Xu, X., Musfeldt, J.
L.: Photoconductivity in BiFeO3 thin films. App. Phys. Lett. 92, 091905 (2008)
309. Choi, T., Lee, S., Choi, Y.J., Kiryukhin, V., Cheong, S.-W.: Switchable ferroelectric diode
and photovoltaic effect in BiFeO3. Science 324, 63 (2009)
310. Yang, S.Y., Martin, L.W., Byrnes, S.J., Conry, T.E., Basu, S.R., Paran, D., Reichertz, L.,
Ihlefeld, J., Adamo, C., Melville, A., Chu, Y.-H., Yang, C.-H., Musfeldt, J.L., Schlom, D.G.,
Ager, J.W., Ramesh, R.: Photovoltaic effects in BiFeO3. Appl. Phys. Lett. 95, 062909 (2009)
311. Arunachalam, V.S., Fleischer, E.L.: The global energy landscape and materials innovation.
MRS Bull. 33, 264–276 (2008)
312. Park, T.-J., Mao, Y., Wong, S.S.: Synthesis and characterization of multiferroic BiFeO3
nanotubes. Chem. Commun. 23, 2708–2709 (2004)
313. Park, T.-J., Papaefthymiou, G.C., Viescas, A.J., Moodenbaugh, A.R., Wong, S.S.: Size-
dependent magnetic properties of single-crystalline mulitferroic BiFeO3 nanoparticles.
Nano Lett. 7, 766–772 (2007)
314. Selbach, S.M., Tybell, T., Einarsrud, A.-A., Grande, T.: Size-dependent properties of
multiferroic BiFeO3 nanoparticles. Chem. Mater. 19, 6478–6484 (2007)
315. Chen, P., Podraza, N.J., Xu, X.S., Melville, A., Vlahos, E., Ramesh, R., Gopalan, V., Schlom,
D.G., Musfeldt, J.L.: Optical properties of tetragonal BiFeO3 thin film. Appl. Phys. Lett. 96,
131907 (2010)
316. Chen, P., Xu, X.S., Santulli, A.C., Koenigsmann, C., Wong, S.S., Musfeldt, J.L.: Size-
dependent infrared phonon modes and ferroelectric phase transition in BiFeO3 nanoparticles.
Nano Lett. 10, 4526–4532 (2010)
317. Xu, X.S., Ihlefeld, J.F., Lee, J.H., Vlahos, E., Ramesh, R., Gopalan, V., Schlom, D.G.,
Musfeldt, J.L.: Tunable band gap in Bi(Fe1-xMnx)O3 alloy films. Appl. Phys. Lett. 96,
192901 (2010)
318. Lee, J.H., Ke, X., Misra, R., Ihlefeld, J.F., Xu, X.S., Mei, Z.G., Meeg, T., Roecherath, M.,
Schubert, J., Liu, Z.K., Musfeldt, J.L., Schiffer, P., Schlom, D.G.: Absorption-controlled
growth of BiMnO3 thin films by molecular-beam epitaxy. Appl. Phys. Lett. 96, 262905 (2010)
319. Ihlefeld, J.F., Podraza, N.J., Liu, Z.K., Rai, R.C., Xu, X., Heeg, T., Chen, Y.B., Li, J., Collins,
R.W., Musfeldt, J.L., Pan, X.Q., Schubert, J., Ramesh, R., Schlom, D.G.: Optical band gap of
BiFeO3 grown by molecular-beam epitaxy. Appl. Phys. Lett. 92, 142908 (2008)
320. Kumar, A., Rai, R.C., Podraza, N.J., Denev, S., Ramirez, M., Chu, Y.-H., Ihlefeld, J.F., Heeg, T.,
Schubert, J., Schlom, D.G., Orenstein, J., Ramesh, R., Collins, R.W., Musfeldt, J.L., Gopalan, V.:
Linear and nonlinear optical constants of BiFeO3. Appl. Phys. Lett. 92, 121915 (2008)
321. Li, S., Lin, Y.-H., Zhang, B., Wang, Y., Nan, C.: Controlled fabrication of BiFeO3 uniform
microscrystals and their magnetic and photocatalytic behaviors. J. Phys. Chem. C 114, 2903
(2010)
322. Reches, J., Rossell, M.D., Zhang, J.X., Hatt, A.J., He, Q., Yang, C.-H., Kumar, A., Wang, C.
H., Melviite, A., Adamo, C., Sheng, G., Chu, Y.-H., Ihlefeld, J.F., Erni, R., Ederer, C.,
Gopalan, V., Chen, L.Q., Schlom, D.G., Spaldin, N.A., Martin, L.W., Ramesh, R.:
A strain-driven morphotropic phase boundary in BiFeO3. Science 326, 977–980 (2009)
126 J.L. Musfeldt

323. Homes, C.C., Dordevic, S.V., Gu, G.D., Li, Q., Valla, T., Tranquada, J.M.: Charge order,
metallic behavior, superconductivity in La2-xBaxCuO4 with x ¼ 1/8. Phys. Rev. Lett. 96,
257002 (2006)
324. Poudeu, P.F.P., D’Angelo, J., Downey, A.D., Short, J.L., Hogan, T.P., Kanatzidis, M.G.:
High thermoelectric figure of merit and nanostructuring in bulk p-type Na1-xPbmSbyTem+2.
Angew. Chem. Int. Ed. 45, 3835–3839 (2006)
325. Hochbaum, A.I., Chen, R., Delgado, R.D., Liang, W.J., Garnett, E.C., Najarian, M.,
Majumdar, A., Yang, P.D.: Enhanced thermoelectric performance of rough silicon
nanowires. Nature 451, 163–167 (2008)
326. Boukai, A.I., Bunimovich, Y., Tahir-Kheli, J., Yu, J.-K., Goddard III, W.A., Heath, J.R.:
Silicon nanowires as efficient thermoelectric materials. Nature 451, 168–171 (2008)
327. Long, J.R., Yaghi, O.M.: The pervasive chemistry of metal-organic frameworks. Chem. Soc.
Rev. 38, 1213–1214 (2009)
328. Bonino, F., Chavan, S., Vitillo, J.G., Groppo, E., Agostini, G., Lamberti, C., Dietzel, P.D.C.,
Prestipino, C., Bordiga, S.: Local structure of CPO-27-Ni metallorganic framework upon
dehydration and coordination of NO. Chem. Mater. 20, 2957–4968 (2008)
329. Britt, D., Furukawa, H., Wang, B., Glover, T.G., Yaghi, O.M.: Highly efficient separation of
carbon dioxide by a metal-organic framework replete with open metal sites. Proc. Natl. Acad.
Sci. USA 106, 20637–20640 (2009)
330. Chavan, S., Vitillo, J.G., Groppo, E., Bonino, F., Lamberti, C., Dietzel, P.D.C., Bordiga, S.:
CO absorption on CPO-27-Ni coordination polymer: spectroscopic feature and interaction
energy. J. Phys. Chem. C 113, 3292–3299 (2009)
331. Czaja, A.U., Trukhan, N., M€ uller, U.: Industrial applications of metal-organic frameworks.
Chem. Soc. Rev. 38, 1284–1293 (2009)
332. Han, S.S., Furukawa, H., Yaghi, O.M., Goddard, W.A.: Covalent organic frameworks as
exceptional hydrogen storage materials. J. Am. Chem. Soc. 130, 11580–11581 (2008)
333. Zecchina, A., Groppo, E., Bordiga, S.: Selective catalysis and nanoscience: an inseparable
pair. Chemistry 13, 2440–2460 (2007)
334. Wang, B., Coté, A.P., Furukawa, H., O’Keeffe, M., Yaghi, O.M.: Colossal cages in zeolitic
imidazolate frameworks as selective carbon dioxide reservoirs. Nature 453, 207–211 (2008)
335. Zecchina, G., Spoto, S.: Bordiga, Probing the acid sites in confined spaces of microporous
materials by vibrational spectroscopy. Phys. Chem. Chem. Phys. 7, 1627–1642 (2005)
336. Rzepka, M., Kamp, P., de la Casa-Lillo, M.A.: Physisorption of hydrogen on microporous
carbon and carbon nanotubes. J. Phys. Chem. B 102, 10894–10898 (1998)
337. Britt, D., Tranchemontagne, D.J., Yaghi, O.M.: Metal-organic frameworks with high capac-
ity and selectivity for harmful gases. Proc. Natl. Acad. Sci. USA 105, 11623–11627 (2008)
338. Dinca, M., Dailly, A., Liu, Y., Brown, C.M., Neumann, D.A., Long, J.R.: Hydrogen storage
in a microporous metal-organic framework with exposed Mn2+ coordination sites. J. Am.
Chem. Soc. 128, 16876–16883 (2006)
339. Bordiga, S., Bitillo, J.G., Ricchiardi, G., Regli, L., Cocina, D., Zecchina, A., Arstad, B.,
Bjorgen, M., Harizocic, J., Lillerud, K.P.: Interaction of hydrogen with MOF-5. J. Phys.
Chem. B 109, 18237–18242 (2005)
340. Vitillo, J.G., Regli, L., Chavan, S., Ricchiardi, G., Spoto, G., Dietzel, P.D.C., Bordiga, S.,
Zecchina, A.: Role of exposed metal sites in hydrogen storage in MOFs. J. Am. Chem. Soc.
130, 8386–8396 (2008)
341. Roswell, J.L.C., Yaghi, O.M.: Strategies for hydrogen storage in metal-organic frameworks.
Angew. Chem. Int. Ed. 44, 4670–4679 (2005)
342. Xu, X.S., Sun, Q.-C., Rosentsveig, R., Musfeldt, J.L.: Evaluation of Born and local effective
changes in unoriented materials from vibrational spectra. Phys. Rev. B 80, 014303 (2009)
343. Sun, Q.C., Xu, X.S., Baker, S.N., Christianson, A.D., Musfeldt, J.J.: Experimental determi-
nation of ionicity in MnO. Chem. Mater. 23, 2956 (2011)
344. Sun, Q.C., Baker, S.N., Christianson, A.D., Musfeldt, J.L. Magneto-elastic coupling in bulk
and nanoscale MnO. Phys. Rev. B. 84, 014301 (2011)
Chapter 6
Electronic Properties of Post-transition Metal
Oxide Semiconductor Surfaces

T.D. Veal, P.D.C. King, and C.F. McConville

6.1 Introduction

The post-transition metal oxides (PTMOs) have traditionally been classified as


transparent conductors. The normally contradictory properties of transparency to
visible light and electrical conductivity have made materials such as Sn-doped
In2O3 and Al-doped ZnO suitable for use as transparent electrodes in devices such
as solar cells and flat panel displays. However, the PTMOs are increasingly being
considered as semiconductors for potential use as the active layer in a wide range of
transparent (opto)electronic devices.
The electronic properties of the surfaces of semiconductors are important for
many device applications. Due to different atomic coordination and broken trans-
lational symmetry at the surface, the electronic properties can differ markedly in
this region compared to in the bulk of a semiconductor. The presence of charged
surface states can induce a macroscopic redistribution of free carriers close to the
surface, in order to screen the electric fields associated with such a surface charge,
and therefore minimize energy. Such surface space-charge regions are associated
with a bending of the electronic bands relative to the Fermi level, over distances
defined by the electronic screening length of the material. Typical length scales for
such a band bending in semiconductors are on the order of 5–500 nm. Therefore,
while such surface electronic properties are important for many device applications,
they become even more important in nanostructured semiconductors, where the
surface space-charge region can account for a large proportion of the volume of the
material. Indeed, transport properties of semiconductor nanowires are known to
depend on the nature of the surface space charge of the particular material. For
example, in InN nanowires the conductivity increases as the diameter is reduced
due to the presence of surface electron accumulation, while in GaN nanowires

T.D. Veal (*) • P.D.C. King • C.F. McConville


Department of Physics, University of Warwick, Coventry, CV4 7AL, UK
e-mail: Timothy.Veal@warwick.ac.uk

J. Wu et al. (eds.), Functional Metal Oxide Nanostructures, Springer Series 127


in Materials Science 149, DOI 10.1007/978-1-4419-9931-3_6,
# Springer Science+Business Media, LLC 2012
128 T.D. Veal et al.

Fig. 6.1 The post-transition


metal elements and oxygen
highlighted on a portion of the
periodic table. The shading
corresponds to the
electronegativity of the
elements. The atomic number
(top left), atomic radius in pm
(top right) and the
electronegativity in Pauling
units (bottom) are also shown
for each element

reducing the diameter decreases the conductivity as a result of the surface electron
depletion layer [1, 2]. This follows the surface space-charge behaviour of thin films
of InN and GaN [3, 4]. The surface space-charge region is widely recognized as
important to the wide variety of applications of metal oxide nanostructures, but
frequently the only type of intrinsic space-charge region considered for n-type
semiconductors is an electron depletion layer [5]. However, as will be discussed
below, this is generally an incorrect assumption for PTMOs. The electronic
properties of metal oxide nanostructures provide the motivation for this chapter
on the electronic properties of the surfaces of PTMO thin films. Until recently,
many of the fundamental surface properties of these materials were unknown or
misunderstood. A knowledge and understanding of the fundamental surface space-
charge properties of this class of materials will be of interest to those researching
both the application of these materials in transparent (opto)electronics and the
transport properties of PTMO nanostructures.
The post-transition metallic elements and oxygen are highlighted in Fig. 6.1 on a
portion of the periodic table that depicts the electronegativity and atomic radius of
the elements. This chapter is limited to the surface electronic properties of the
period 4 (ZnO and Ga2O3) and period 5 (CdO, In2O3, and SnO2) PTMO
semiconductors. The period 6 PTMOs (HgO, Tl2O3, PbO, PbO2, and Bi2O3) have
not yet had their surface electronic properties investigated. In Sect. 6.2, the surface
space-charge properties will be presented for each of the PTMOs of periods 4 and 5.
This will be followed in Sect. 6.3 by an explanation of the overriding cause of these
and related properties in terms of the nature of the bulk band structure of the
materials and the location of the band edges with respect to the charge neutrality
level (CNL).
6 Electronic Properties of Post-transition Metal Oxide Semiconductor Surfaces 129

6.2 Surface Space-Charge Properties

The surface space-charge properties of an n-type semiconductor can be determined,


to a large extent, by measuring the surface and bulk Fermi levels, with the
difference between them being accounted for by the band bending. In practice,
the surface Fermi level with respect to the valence-band maximum (VBM) can
be determined using X-ray photoemission spectroscopy (XPS) and knowledge of
the band gap of the semiconductor enables the Fermi level with respect to the
conduction-band minimum (CBM) to be obtained.1 The bulk Fermi level can be
estimated by measuring either the free-electron density using the Hall effect, the
conduction electron plasma frequency using infrared reflectivity, or, for a degener-
ately doped material, by using optical absorption spectroscopy. The determined
surface and bulk Fermi levels can then be used as boundary conditions to solve the
Poisson equation within the modified Thomas–Fermi approximation [6]. The solu-
tion of the Poisson equation gives the band bending and charge profiles as a
function of depth below the semiconductor surface and also the surface state
density. Here, the surface space-charge properties of all the period 4 and 5
PTMOs investigated using these methods are presented within the context of the
previous understanding of their surface electronic properties.

6.2.1 ZnO

The first member of the period 4 PTMOs, ZnO, has the wurtzite structure and a
room temperature band gap of 3.35 eV. The electronic properties of its surfaces
have been studied for over 40 years [7]. Some of the earliest work showed that it is
possible to modify the surface space charge from electron depletion to strong
electron accumulation by varying the surface treatment. Exposure of the ZnO
surface to oxygen induces an electron depletion layer, whereas atomic hydrogen
treatment produces surface electron accumulation [7]. However, the “intrinsic”
state of the space charge at ZnO surfaces is still debated [8, 9] and the role of
surface conductivity in ZnO nanowires is far from being resolved [10].
Figure 6.2 shows valence-band XPS of bulk ZnO crystals grown by both the
hydrothermal (HT) and pressurized-melt (PM) techniques. Spectra were collected
from both the Zn- and O-polar faces of the same c-axis crystals grown by each method
and also from non-polar crystals (m-plane and a-plane). The room temperature bulk
electron densities were found, from multiple field Hall effect measurements, to be
2–3  1014 cm3 for the HT ZnO, 5  1016 cm3 for the a-plane PM ZnO, and

1
While this picture may be complicated somewhat by many-body effects within the surface space-
charge region, as will be discussed later in this chapter, it still provides a correct qualitative and
indeed semi-quantitative picture of the surface space-charge characteristics of a given material.
130 T.D. Veal et al.

Intensity (arb.units)
HT ZnO
m-plane face

ζ = 3.60 eV
Intensity (cps)

9 8 7 6 5 4 3 2 1 0 −1
Binding Energy (eV)
HT ZnO
Zn-polar face ζ = 3.71 eV
O-polar face ζ = 3.52 eV

9 8 7 6 5 4 3 2 1 0 −1
Binding Energy (eV)

b
Intensity (cps)

PM ZnO
a-plane face
Intensity (cps)

ζ = 3.54 eV

9 8 7 6 5 4 3 2 1 0 −1
Binding Energy (eV)
PM ZnO ζ = 3.61 eV
Zn-polar face
O-polar face ζ = 3.50 eV

9 8 7 6 5 4 3 2 1 0 −1
Binding Energy (eV)

Fig. 6.2 Valence-band XPS spectra from the Zn-polar and O-polar faces of (a) the same bulk
c-axis HT ZnO wafer and (b) the same bulk c-axis PM ZnO wafer, showing the extraction of x, the
valence-band maximum to Fermi level separation, for each face. The inset of (a) shows the
spectrum from the m-plane face of a separate HT ZnO wafer. The inset of (b) shows the spectrum
from the a-plane face of a separate PM ZnO wafer. Figure reproduced with permission from [9].
Copyright (2010) by the American Physical Society

1.5  1017 cm3 for the c-plane PM ZnO [9]. The separation, z, between the VBM
and the Fermi level at the surface is determined by extrapolating a linear fit to the
lower binding energy edge of the valence band photoemission to a line fitted through
the background data. The Fermi level is calibrated to be the zero of the binding energy
scale by using the Fermi edge of a silver reference sample.
6 Electronic Properties of Post-transition Metal Oxide Semiconductor Surfaces 131

From the XPS data, the surface Fermi level was found to be located between
3.50 and 3.71 eV above the VBM, with the highest values being for the Zn-polar
surface, the lowest values for the O-polar surface, and the non-polar surfaces in
between [9]. Since the measured bulk carrier densities correspond to bulk Fermi
levels ranging from 3.10 to 3.27 eV above the VBM, the surface Fermi level is
always higher than the bulk Fermi level, indicating the presence of surface electron
accumulation for all samples investigated. A greater difference between the surface
Fermi levels of the Zn- and O-polar surfaces was found for the HT ZnO than for the
much higher bulk carrier density PM ZnO. The greater electron accumulation at the
Zn-polar face, particularly for the HT ZnO, is attributed to the spontaneous polari-
zation [11] associated with the wurtzite structure. The effect is reduced for the PM
ZnO due to partial screening by the higher bulk electron density of these samples.
The surface band bending and carrier concentration profiles for the various ZnO
samples were determined by numerical solution of Poisson’s equation within the
modified Thomas–Fermi approximation [6] and are shown in Fig. 6.3.
From the measured amount of band bending, solving the Poisson equation
enabled the surface sheet densities to be determined as 7  1012 cm2
(3  1012 cm2) for the Zn-polar (O-polar) face of the HT ZnO and
5  1012 cm2 (2.5  1012 cm2) for the Zn-polar (O-polar) face of the PM
ZnO. These surface electrons have been found to have a significant influence on
the electrical properties of high resistivity HT ZnO at temperatures below 200 K
and are a major cause of the anomalously low maximum electron mobility
measured using single magnetic field Hall effect measurements. For PM ZnO, the
surface layer has little effect on its electrical properties above 50 K due to the higher
concentration of uncompensated shallow donors in this material [9]. Surface elec-
tron accumulation appears to be the natural state of both polar and non-polar ZnO
surfaces and can significantly influence electrical measurements made on heavily
compensated ZnO material. Consequently it must be taken into account when
investigating the electrical properties of potentially p-type material. Additionally,
it can be seen from Fig. 6.3 that, for the HT ZnO, the free-electron density 100 nm
away from the surface is still significantly higher than the bulk value, illustrating
how important surface properties can be for nanoscale structures.

6.2.2 Ga2O3

The second member of the period 4 PTMOs, Ga2O3, exhibits five different
polymorphs [12], with the most stable being b-Ga2O3 with a monoclinic structure
and a room temperature band gap in the range of 4.4–4.9 eV [13, 14]. In spite of the
surface electronic properties of Ga2O3 not receiving much attention, thin films of
this material have been successfully used as gas sensors. Rather than exhibiting
sensitivity to anion-like chemisorption, such as O2, they have been found to be good
sensors of reducing gases, such as H2 and CO, when chemisorption of donor-like
species occurs [15]. This behaviour was explained in terms of an electron
132 T.D. Veal et al.

Fig. 6.3 Poisson-MTFA calculations of the band banding and the carrier concentration profiles in
the electron accumulation layer at the Zn-polar (a) and (c) and O-polar faces (b) and (d) of HT
ZnO; and at the Zn-polar face (e) and (g) and O-polar face (f) and (h) of PM ZnO. Figure
reproduced with permission from [9]. Copyright (2010) by the American Physical Society
6 Electronic Properties of Post-transition Metal Oxide Semiconductor Surfaces 133

accumulation layer model of the Ga2O3 surface [16]. The presence of an electron
accumulation is also suggested by recent photoemission data, where the surface
Fermi level can be seen to be approximately 4.5 eV above VBM [17]. Such a
surface Fermi level, very close to or above the CBM, is above typical Ga2O3 bulk
Fermi levels, indicating downward band bending and surface electron accumula-
tion. Further investigations of Ga2O3 are, however, required to verify the nature of
its surface electronic properties.

6.2.3 CdO

The first member of the period 5 PTMOs, CdO, has the rock salt structure and a
room temperature direct band gap of 2.2 eV [18, 19], with two indirect band gaps of
less than 1.1 eV [19–21]; the VBM is not at the G-point due to repulsion between
the Cd 4d and O 2p states being symmetry forbidden at G for the rock salt structure
[22]. Until recently, CdO had long been assumed to exhibit surface electron
depletion [23]. However, several recent studies have contradicted this, finding
evidence of surface electron accumulation [24–26].
Specifically, a combination of valence-band XPS, Hall effect, and optical absorp-
tion and reflectivity has been used to determine the surface and bulk Fermi levels of
CdO(001) films grown by metal organic vapour phase epitaxy on sapphire substrates.
The valence-band XPS measurements, shown in Fig. 6.4a, give the L-point (indirect)
VBM to surface Fermi level separation as x ¼ 1.29  0.05 eV. Taking the separa-
tion of the G-point and L-point maxima of the valence band as 1.2 eV from
calculations [24], the surface Fermi level is found to be 2.49 eV above the G-point
VBM. Meanwhile, from Hall effect measurements (n ¼ 1.6  1019 cm3, m ¼ 106
cm2 V1 s1) and the optical absorption and IR reflectivity data shown in Fig. 6.4b, c,
the G-point VBM to bulk Fermi level separation was determined as  ¼ 2.23
 0.05 eV. Therefore, the Fermi level lies higher relative to the band extrema at
the surface than in the bulk, implying a downward bending of the bands at the surface
of 0.26 eV. The calculated band bending is shown in Fig. 6.4d along with the large
accumulation of electrons in the near surface region shown in Fig. 6.4e.
The surface electron accumulation has additionally been found to be quantized.
Angle-resolved photoemission spectroscopy (ARPES) has revealed that the con-
duction electrons at the surface occupy two-dimensional subbands created by the
confining potential well associated with the downward band bending [24, 26]. The
photoemission maps for a CdO surface-quantized two-dimensional electron gas
(Q2DEG) are shown in Fig. 6.5. Similar Q2DEG states are expected at the surface
of PTMO nanowires. However, if the nanowire size is sufficiently small, the
wavefunction tails from the Q2DEG states of one surface may overlap with those
from the surface on the opposite side of the nanowire. In this case, these states could
interact, potentially giving rise to novel physics between two spatially separated,
but nonetheless coupled, Q2DEGs.
134 T.D. Veal et al.

Fig. 6.4 (a) Valence-band XPS, (b) squared optical-absorption coefficient, and (c) measured
(points) and simulated (line) IR reflectivity spectra for an undoped CdO sample following
annealing at 600 C in UHV for 2 h. (d) Band bending [CBM, indirect (L-point) and direct
(G-point) VBMs] and (e) carrier concentration as a function of depth below the surface in the
electron accumulation layer. Figure reproduced with permission from [25]. Copyright (2009) by
the American Physical Society

Intriguingly, such detailed spectroscopic measurements of the electronic states


within a surface electron accumulation layer reveal a discrepancy with the simple
band-bending picture discussed so far. While the valence-band bending determined
by XPS (Fig. 6.4a), as well as core-level and other valence-band photoemission
measurements, (Fig. 6.5e) indicates a downward band bending of only 0.26 eV, a
deep enough quantum well to give the measured conduction-band subband as low
as 0.5 eV below the Fermi level requires the conduction band to bend downward by
1.1 eV. This discrepancy between the amount of valence- and conduction-band
bending can be resolved by considering many-body effects within the electron
accumulation layer [26]. Instead of the conduction and valence bands undergoing
the same amount of band bending as in the conventional one-electron picture of
6 Electronic Properties of Post-transition Metal Oxide Semiconductor Surfaces 135

Fig. 6.5 (a) Quantized conduction-band subbands of a surface Q2DEG. (b) E vs. k|| and
(c) constant energy cuts through the subband dispersions measured by ARPES, shown here for a
CdO(001) surface recorded with a photon energy of hn ¼ 30 eV at room temperature. The
constant energy cuts are integrated over 3 meV about (c1) the Fermi level and (c2) 0.15, (c3)
0.30, and (c4) 0.45 eV below the Fermi level, respectively. (d) Schematic of downward bending of
the conduction band and (inset) corresponding increase in free-carrier density within the semicon-
ductor surface electron accumulation layer (Q2DEG). (e) Angle-integrated PES measurements of
the valence bands and Cd 4d core levels in CdO, recorded at a photon energy of hn ¼ 120 eV.
Figure reproduced with permission from [26]. Copyright (2010) by the American Physical Society

surface space charge, band gap renormalization occurs in the accumulation layer
where the electron density is the highest, shrinking the band gap close to the surface,
as illustrated in Fig. 6.6. This band gap shrinkage has a large influence on the surface
sheet density; within the one-electron picture, the valence-band bending implies a
sheet density of 1  1013 cm2, while the actual sheet density, determined from the
Luttinger area, N2D ¼ kF2 =2p, with the Fermi wave vector, kF, taken directly from the
ARPES data, is 4.4  1013 cm2. Such many-body effects probably increase the
surface sheet electron densities of other materials discussed elsewhere in this chapter
compared with the reported values based on valence-band photoemission data.

6.2.4 In2O3

The second member of the period 5 PTMOs, In2O3, has the body-centred-cubic
(bcc) bixbyite structure or the less stable rhombohedral (rh) structure. The direct
band gap of cubic In2O3 was long thought to be 3.7 eV, based on the onset of optical
absorption, with an indirect gap of about 2.6 eV [27, 28]. However, recent theoreti-
cal studies have found no evidence for any significant difference between the direct
and indirect band gaps [29, 30]. Experimental studies using photoemission, X-ray
emission and absorption, and optical absorption are consistent with a band gap in
136 T.D. Veal et al.

Fig. 6.6 One-electron (solid


line) and schematic
renormalized (dashed line)
band bending at the surface of
CdO, and corresponding
calculated one-electron
(solid) and measured
renormalized (dashed)
subband energies. Figure
reproduced with permission
from [26]. Copyright (2010)
by the American Physical
Society

the range 2.6–2.95 eV [29, 31]. The theoretical calculations indicate that the weak
nature of optical absorption around this energy can be attributed to transitions
between the highest valence-band states and states at the CBM being dipole
forbidden or having only minimal dipole intensity [29, 30].
This revision of the band gap has brought about a re-evaluation of the surface space-
charge properties of In2O3. Before 2008, photoemission spectroscopy had been used to
locate the surface Fermi level at about 3 eV above the VBM [32, 33]. With the accepted
band gap at that time being 3.7 eV, this, along with a bulk Fermi level close to the
CBM, implied the presence of upward band bending and a surface electron depletion
layer. However, with the recently established band gap of 2.9 eV (3.0 eV) for bcc (rh)
In2O3, this and more recent photoemission data, such as that shown in Fig. 6.7 for both
bcc- and rh-In2O3, indicate that the surface Fermi level is above the bulk Fermi level,
corresponding to downward band bending and electron accumulation [31, 34].
For In2O3, with the shallow dispersion of its topmost valence bands [30], the
VBM position with respect to the Fermi level cannot be accurately determined by
extrapolating the leading edge of the valence-band photoemission; this method
leads to an underestimation of the VBM to surface Fermi level separation due to the
effects of instrumental and lifetime broadening. Instead, as shown in Fig. 6.7, the
experimental data are compared with the broadened calculated valence-band den-
sity of states (VB-DOS). A 3.4 eV (3.5 eV) shift is required to align the calculated
VB-DOS with the XPS data from undoped bcc (rh) In2O3, indicating that the
surface Fermi level is 0.5 eV above the CBM for both bcc and rh In2O3. Along
with the bulk Fermi levels for the different samples, determined from Hall effect
and infrared reflectivity measurements, the downward band bending was found to
be in the range 0.4–0.5 eV [31]. When heavily Sn-doped, the bulk Fermi level is
much closer to the surface Fermi level, resulting in approximately flat bands. The
band bending and carrier concentration profiles as a function of depth below
the surface, calculated by solving the Poisson equation, are shown in Fig. 6.8
[31]. The surface sheet density can be estimated from these calculations to be
1.3  1013 cm2 for all of the undoped In2O3 samples.
6 Electronic Properties of Post-transition Metal Oxide Semiconductor Surfaces 137

Fig. 6.7 Background-


subtracted valence-band
photoemission spectra from
(a) undoped and Sn-doped
bcc-In2O3(001) and undoped
(111) (grown by plasma-
assisted molecular-beam
epitaxy and metal organic
vapour phase epitaxy) and
(b) undoped rh-In2O3(0001).
The conduction-band
emission is magnified 25
times. The density functional
theory calculated valence-
band density of states is
shaded in grey and is also
shown after lifetime and
instrumental broadening.
Figure is adapted from [31]

Fig. 6.8 Poisson-MTFA


calculations of band bending
(a) and (c) and carrier
concentration profiles (b) and
(d) in the electron
accumulation layer at the
surface of bcc-In2O3 (a) and
(b) and rh-In2O3 (c) and (d).
Figure is adapted from [31]
138 T.D. Veal et al.

The presence of surface electron accumulation is further corroborated by a


comparison of Al Ka XPS (hn ¼ 1486.6 eV) and synchrotron radiation high energy
XPS (hn ¼ 6,000 eV) [35]. As can be seen in Fig. 6.7, an additional peak is present in
the photoemission spectra just below the Fermi level. This corresponds to photoemis-
sion from the occupied conduction-band states which have predominantly In 5s
character. After correcting for matrix element effects, Zhang et al. found a larger
contribution from filled conduction-band states in the more surface sensitive
1486.6 eV XPS than for the 6,000 eV XPS [35]. This indicates that there is an
increased concentration of conduction electrons close to the surface than deeper in the
bulk, confirming the picture of a surface electron accumulation layer in In2O3.

6.2.5 SnO2

The third and final member of the period 5 PTMOs, SnO2, has the rutile structure
and a band gap of 3.59 eV (3.50 eV at room temperature) [36, 37]. The surface
properties of SnO2 have been comprehensively reviewed quite recently [38].
The surface of SnO2 can exhibit surface electron depletion or accumulation,
depending on the stoichiometry of the surface and the bulk Fermi level. Similar
to ZnO, adsorption of anion species leads to electron depletion, while cation
adsorption leads to accumulation, making SnO2 a good material for gas sensing
via corresponding changes in conductivity. Indeed, nanoscale structures of SnO2
are particularly suited to this task [39], where the larger surface to bulk ratio
provides a relatively higher density of surface adsorption sites compared to thin
films, enabling the adsorbed species to have an even greater influence on the total
measured conductivity. The presence of an electron accumulation layer at the
surface of SnO2 under certain conditions is long established [40], but whether it
is the “intrinsic state” of the surface is not yet well established.
Valence-band XPS from an undoped bulk SnO2(100) crystal after cleaning by 2 h
of vacuum annealing at 400 C is shown in Fig. 6.9. The VBM to surface Fermi level
separation is 3.70 eV. With a room temperature band gap of 3.50 eV, this corresponds
to the surface Fermi level being 0.20 eV above the CBM. This is well above the non-
degenerate bulk Fermi level of undoped SnO2, indicating the presence of downward
band bending and electron accumulation. Before annealing a VBM to surface Fermi
level separation of 3.60 eV was observed, again indicating electron accumulation but
to a lesser extent than after the surface was reduced by annealing.
These results, parallels with the other PTMOs, some recent studies [37, 41], and
the discussion in the next section, indicate that SnO2 surfaces do have an inherent
tendency to exhibit electron accumulation in contrast to the surface depletion layer
of the majority of n-type semiconductors. However, the presence of anion species
when used as a gas detector or exposure to oxygen plasma treatment can result in
surface electron depletion [38, 41].
6 Electronic Properties of Post-transition Metal Oxide Semiconductor Surfaces 139

Fig. 6.9 Background-


subtracted valence-band
photoemission spectrum from
bulk-grown SnO2(100)

6.3 Bulk Band Structure Origin of Electron


Accumulation Propensity

From the preceding sections, it is clear that the PTMOs exhibit a tendency towards
surface electron accumulation, rather than the charge depletion common to almost
all conventional semiconductors. This raises the question whether some universal
characteristic of the PTMOs drives this apparently unconventional surface behaviour.
In addition to their propensity for strongly n-type surfaces, these materials also have a
proclivity towards n-type bulk properties, with many defect centres that are tradition-
ally considered to counteract the prevailing conductivity, including hydrogen
impurities [42–45] and several native defects, exhibiting an extreme propensity for
donor behaviour in PTMOs. Any model which explains the surface properties of
these materials must also be consistent with these striking bulk properties.
In fact, both the surface and bulk characteristics of PTMOs can naturally be
explained from gross features of their bulk band structure. A typical PTMO bulk
band structure, for In2O3, is shown in Fig. 6.10. While multiple and quite complex
valence bands are observed, the conduction band is characterized by a single low-
lying band at the zone centre. Throughout the rest of the Brillouin zone, the
conduction band lies much higher. The character of the so-called deep, that is
localized, defect, is determined from the band structure across the Brillouin zone
(that is, they have a very extended k-space nature). Indeed, for broken symmetry
defect states such as surface states, native defects, and hydrogen impurities, it is
possible to define an energy level at the mid-point of the average band gap across the
Brillouin zone which determines the charge transition from positively to negatively
charged defect centres [46–49]. If this so-called CNL lies within the fundamental
band gap, it is generally favourable for compensating defect centres and surface
charge depletion, as shown in Fig. 6.11. However, for a bulk band structure of the
form depicted in Fig. 6.10, the low-lying CBM can actually occur below the mid-gap
140 T.D. Veal et al.

Fig. 6.10 Calculated band structure of In2O3. Figure adapted from [30]

Fig. 6.11 Schematic formation energies for dominant native defects and interstitial hydrogen
impurities and schematic band bending in conventional semiconductors (left) and PTMOs (right).
The influence of the CNL, within the band gap in conventional semiconductors but above the CBM
in PTMOs, can clearly be seen. Figure adapted from [25]
6 Electronic Properties of Post-transition Metal Oxide Semiconductor Surfaces 141

Fig. 6.12 Band line-up of a number of PTMOs and other semiconductors, relative to the charge
neutrality level. The PTMOs and InN are highlighted and compared with the conventional
semiconductors, Si and GaAs. Ga2O3 is omitted due to the lack of quantitative information
about the location of its band edges with respect to the CNL. The positions of the bands relative
to the CNL are derived from the measurements and calculations in [25, 31, 50–53, 56]

position averaged across the entire Brillouin zone. In this case, even for bulk Fermi
levels above the bottom of this CBM, it remains favourable for donor-like surface
states to donate their electrons into the conduction band, forming a surface electron
accumulation. Similarly, native defects and hydrogen impurities can both gain energy
by donating free-carriers into the conduction band (see Fig. 6.11).
Consequently, the general propensity for donor-type surface states as well as donor
bulk defects and hydrogen in PTMOs can all be understood to result from the bottom of
the conduction band lying below the CNL in these materials. Indeed, from a number of
direct measurements [25, 34], relative valence-band alignments from valence-band
offset measurements [50–52], and calculations [53, 54], the CNL lying above the
bottom of the conduction band can indeed be seen as a universal property of the
PTMOs, as shown in Fig. 6.12. In all of these cases, the low-lying CBM gives rise to
this band alignment: such features of the bulk band structure are apparent in electronic
structure calculations for many PTMOs [22, 37, 55]. This itself can be understood as a
consequence of the large size and electronegativity mismatch between the constituent
post-transition metal element and oxygen [56] (see Fig. 6.1). A similar situation occurs
in the semiconductor InN, with the CBM also located below the CNL [56], as shown in
Fig. 6.12. Indeed, the properties of InN exhibit many similarities with the PTMOs
[56–58]. While we stress that this is not a microscopic model, such similarities, and
their differences from conventional semiconductors such as Si and GaAs, are well
explained by the CNL concept (see Fig. 6.12).
142 T.D. Veal et al.

6.4 Conclusion

Evidence of electron accumulation at the surfaces of the PTMOs ZnO, Ga2O3, CdO,
In2O3, and SnO2 has been presented. For the most ionic and cation–anion
mismatched compounds, CdO and In2O3, electron accumulation is universally
observed (in the absence of extremely high doping levels). The other PTMOs
have a strong tendency to exhibit surface electron accumulation, but can also
have an electron depletion layer under the appropriate surface stoichiometry
conditions or when certain anions are adsorbed.
This proclivity towards surface electron accumulation shown by the PTMOs is
due to their bulk band structure, whereby the G-point band extrema are low with
respect to the CNL, enabling donor surface states to be unoccupied and therefore
positively charged for typical bulk Fermi levels. These surface states donate their
electrons into the conduction band, resulting in surface electron accumulation and a
corresponding downward band bending. This band structure property of the PTMOs
is also responsible for the favourability of donor native defects, high dopability, and
donor hydrogen in already n-type material.
The surface electronic properties of the PTMOs determine the Ohmic or
Schottky behaviour of contacts, are important for their use as gas sensors, and
contribute to their conductivity. The electronic properties of thin films and bulk
crystals of the PTMO surfaces also provide information vital for the interpretation
of conductivity measurements of PTMO nanostructures, which are often dominated
by surface effects.

Acknowledgments The following people are gratefully acknowledged for fruitful collaborations
on determining the surface electronic properties of PTMOs: P. H. Jefferson, J. Zúniga-Peréz, V.
Muñoz-Sanjosé, Ch. Y. Wang, V. Cimalla, O. Ambacher, A. Bourlange, D. J. Payne, K. H. L.
Zhang, R. G. Egdell, M. W. Allen, S. M. Durbin, N. Peng, W. M. Linhart, G. R. Bell, I. Maskery,
L. F. J. Piper, K. E. Smith, E. D. L. Rienks, M. Fuglsang Jensen, Ph. Hofmann, F. Fuchs, A.
Schleife, J. Furthm€uller, and F. Bechstedt. The Engineering and Physical Sciences Research
Council, UK, is acknowledged for funding a Career Acceleration Fellowship for TDV (Grant
no. EP/G004447/1).

References

1. Calarco, R., Marso, M.: GaN and InN nanowires grown by MBE: A comparison. Appl. Phys.
A 87, 499–503 (2007)
2. Richter, T., L€uth, H., Sch€apers, Th, Meijers, R., Jeganathan, K., Estévez Hernández, S.,
Calarco, R., Marso, M.: Electrical transport properties of single undoped and n-type doped
InN nanowires. Nanotechnology 20, 405206 (1–6) (2009)
3. Mahboob, I., Veal, T.D., McConville, C.F., Lu, H., Schaff, W.J.: Intrinsic electron accumula-
tion at clean InN surfaces. Phys. Rev. Lett. 92, 036804 (1–4) (2004)
4. King, P.D.C., Veal, T.D., Lu, H., Jefferson, P.H., Hatfield, S.A., Schaff, W.J., McConville,
C.F.: Surface electronic properties of n- and p-type InGaN alloys. Phys. Stat. Sol. (b) 245,
881–883 (2008)
6 Electronic Properties of Post-transition Metal Oxide Semiconductor Surfaces 143

5. Comini, E., Baratto, C., Faglia, G., Ferroni, M., Vomiero, A., Sberveglieri, G.: Quasi-one
dimensional metal oxide semiconductors: Preparation, characterization and application as
chemical sensors. Prog. Mater. Sci. 54, 1–67 (2009)
6. King, P.D.C., Veal, T.D., McConville, C.F.: Nonparabolic coupled Poisson-Schr€ odinger
solutions for quantized electron accumulation layers: Band bending, charge profile, and
subbands at InN surfaces. Phys. Rev. B 77, 125305 (1–7) (2008)
7. Heiland, G., Kunstman, P.: Polar surfaces of ZnO crystals. Surf. Sci. 13, 72–84 (1969)
8. Schmidt, O., Kiesel, P., Ehrentraut, D., Fukuda, T., Johnson, N.M.: Electrical characterization
of ZnO, including analysis of surface conductivity. Appl. Phys. A 88, 71–75 (2007)
9. Allen, M.W., Swartz, C.H., Myers, T.H., Veal, T.D., McConville, C.F., Durbin, S.M.: Bulk
transport measurements in ZnO: The effect of surface electron layers. Phys. Rev. B 81, 075211
(1–6) (2010)
10. Chiu, S.-P., Lin, Y.-H., Lin, J.-J.: Electrical conduction mechanisms in natively doped ZnO
nanowires. Nanotechnology 20, 015203 (1–8) (2009)
11. Posternak, M., Baldereschi, A., Catellani, A., Resta, R.: Ab initio study of the spontaneous
polarization of pyroelectric BeO. Phys. Rev. Lett. 64, 1777–1780 (1990)
12. Roy, R., Hill, V.G., Osborn, E.F.: Polymorphism of Ga2O3 and the system Ga2O3-H2O. J. Am.
Chem. Soc. 74, 719–722 (1952)
13. Passlack, M., Schubert, E.F., Hobson, W.S., Hong, M., Moriya, N., Chu, S.N.G.: Ga2O3films
for electronic and optoelectronic applications. J. Appl. Phys. 77, 686–693 (1995)
14. Orita, M., Ohta, H., Hirano, M., Hosono, H.: Deep-ultraviolet transparent conductive
b- Ga2O3thin films. Appl. Phys. Lett. 77, 4166–4168 (2000)
15. Fleischer, M., Meixner, H.: Sensing reducing gases at high temperatures using long-term stable
Ga2O3thin films. Sens. Actuators B: Chem 7, 257–261 (1992)
16. Geistlinger, H.: Accumulation layer model for Ga2O3 thin-film gas sensors based on the
Volkenstein theory of catalysis. Sens. Actuators B: Chem 18–19, 125–131 (1994)
17. Lovejoy, T.C., Yitamben, E.N., Shamir, N., Morales, J., Villora, E.G., Shimamura, K., Zheng,
S., Ohuchi, F.S., Olmstead, M.A.: Surface morphology and electronic structure of bulk single
crystal b-Ga2O3(100). Appl. Phys. Lett. 94, 081906 (1–3) (2009)
18. Jefferson, P.H., Hatfield, S.A., Veal, T.D., King, P.D.C., McConville, C.F., Zúñiga-Pérez, J.,
Muñoz-Sanjosé, V.: Bandgap and effective mass of epitaxial cadmium oxide. Appl. Phys. Lett.
92, 022101 (1–3) (2008)
19. Benko, F.A., Koffyberg, F.P.: Quantum efficiency and optical transitions of CdO photoanodes.
Solid State Commun. 57, 901–903 (1986)
20. Koffyberg, F.P.: Thermoreflectance spectra of CdO: Band gaps and band-population effects.
Phys. Rev. B 13, 4470–4476 (1976)
21. King, P.D.C., Veal, T.D., Schleife, A., Zúñiga-Pérez, J., Martel, B., Jefferson, P.H., Fuchs, F.,
Muñoz-Sanjosé, V., Bechstedt, F., McConville, C.F.: Valence-band electronic structure of
CdO, ZnO, and MgO from x-ray photoemission spectroscopy and quasi-particle-corrected
density-functional theory calculations. Phys. Rev. B 79, 205205 (1–6) (2009)
22. Schleife, A., Fuchs, F., Furthm€ uller, J., Bechstedt, F.: First-principles study of ground- and
excited-state properties of MgO, ZnO, and CdO polymorphs. Phys. Rev. B 73, 245212 (2006)
23. Makuta, I.D., Poznyak, S.K., Kulak, A.I.: Hole diffusion transport and photocurrent generation
in the degenerate n-CdO/electrolyte junction. Solid State Commun. 76, 65–68 (1990)
24. Piper, L.F.J., Colakerol, L., King, P.D.C., Schleife, A., Zúñiga-Pérez, J., Glans, P.-A.,
Learmonth, T., Federov, A., Veal, T.D., Fuchs, F., Muñoz-Sanjosé, V., Bechstedt, F.,
McConville, C.F., Smith, K.E.: Observation of quantized subband states and evidence for
surface electron accumulation in CdO from angle-resolved photoemission spectroscopy. Phys.
Rev. B 78, 165127 (1–5) (2008)
25. King, P.D.C., Veal, T.D., Jefferson, P.H., Zúñiga-Pérez, J., Muñoz-Sanjosé, V., McConville,
C.F.: Unification of the electrical behavior of defects, impurities, and surface states in
semiconductors: Virtual gap states in CdO. Phys. Rev. B 79, 035203 (1–6) (2009)
144 T.D. Veal et al.

26. King, P.D.C., Veal, T.D., McConville, C.F., Zúñiga-Pérez, J., Muñoz-Sanjosé, V., Hopkinson,
M., Rienks, E.D.L., Fuglsang, J.M., Hofmann, Ph: Surface band-gap narrowing in quantized
electron accumulation layers. Phys. Rev. Lett. 104, 256803 (1–4) (2010)
27. Weiher, R.L., Ley, R.P.: Optical properties of indium oxide. J. Appl. Phys. 37, 299–302 (1966)
28. Matino, F., Persano, L., Arima, V., Pisignano, D., Blyth, R.I.R., Cingolani, R., Rinaldi, R.:
Electronic structure of indium-tin-oxide films fabricated by reactive electron-beam deposition.
Phys. Rev. B 72, 085437 (1–6) (2005)
29. Walsh, A., Da Silva, J.L.F., Wei, S.-H., K€ orber, C., Klein, A., Piper, L.F.J., DeMasi, A., Smith,
K.E., Panaccione, G., Torelli, P., Payne, D.J., Bourlange, A., Egdell, R.G.: Nature of the band
gap of In2O3 revealed by first-principles calculations and x-ray spectroscopy. Phys. Rev. Lett.
100, 167402 (1–4) (2008)
30. Fuchs, F., Bechstedt, F.: Indium-oxide polymorphs from first principles: Quasiparticle elec-
tronic states. Phys. Rev. B 77, 155107 (1–10) (2008)
31. King, P.D.C., Veal, T.D., Fuchs, F., Wang, C.Y., Payne, D.J., Bourlange, A., Zhang, H., Bell,
G.R., Cimalla, V., Ambacher, O., Egdell, R.G., Bechstedt, F., McConville, C.F.: Band gap,
electronic structure, and surface electron accumulation of cubic and rhombohedral In2O3.
Phys. Rev. B 79, 205211 (1–10) (2009)
32. Christou, V., Etchells, M., Renault, O., Dobson, P.J., Salata, O.V., Beamson, G., Egdell, R.G.:
High resolution x-ray photoemission study of plasma oxidation of indium–tin–oxide thin film
surfaces. J. Appl. Phys. 88, 5180–5187 (2000)
33. Gassenbauer, Y., Schafranek, R., Klein, A., Zafeiratos, S., H€avecker, M., Knop-Gericke, A.,
Schl€ogl, R.: Surface states, surface potentials, and segregation at surfaces of tin-doped In2O3.
Phys. Rev. B 73, 245312 (1–11) (2006)
34. King, P.D.C., Veal, T.D., Payne, D.J., Bourlange, A., Egdell, R.G., McConville, C.F.: Surface
electron accumulation and the charge neutrality level in In2O3. Phys. Rev. Lett. 101, 116808
(1–4) (2008)
35. Zhang, K.H.L., Payne, D.J., Palgrave, R.G., Lazarov, V.K., Chen, W., Wee, A.T.S.,
McConville, C.F., King, P.D.C., Veal, T.D., Panaccione, G., Lacovig, P., Egdell, R.G.: Surface
structure and electronic properties of In2O3(111) single-crystal thin films grown on
Y-stabilized ZrO2(111). Chem. Mater. 21, 4353–4355 (2009)
36. Fr€ohlich, D., Kenklies, R., Helbig, R.: Band-gap assignment in SnO2 by two-photon spectros-
copy. Phys. Rev. Lett. 41, 1750–1751 (1978)
37. Schleife, A., Varley, J.B., Fuchs, F., R€ odl, C., Bechstedt, F., Rinke, P., Janotti, A., Van de
Walle, C.G.: Tin dioxide from first principles: Quasiparticle electronic states and optical
properties. Phys. Rev. B 83, 035116 (2011)
38. Batzill, M., Diebold, U.: The surface and materials science of tin oxide. Prog. Surf. Sci. 79,
47–154 (2005)
39. Comini, E.: Metal oxide nano-crystals for gas sensing. Anal. Chim. Acta 568, 28–40 (2006)
40. De Frésart, E., Darville, J., Gilles, J.M.: Surface properties of tin dioxide single crystals. Surf.
Sci. 126, 518–522 (1983)
41. Nagata, T., Bierwagen, O., White, M.E., Tsai, M.-Y., Speck, J.S.: Study of the Au Schottky
contact formation on oxygen plasma treated n-type SnO2 (101) thin films. J. Appl. Phys. 107,
033707 (1–7) (2010)
42. Van de Walle, C.G., Neugebauer, J.: Universal alignment of hydrogen levels in
semiconductors, insulators and solutions. Nature 423, 626–628 (2003)
43. Cox, S.F.J., Davis, E.A., Cotrell, S.P., King, P.J.C., Lord, J.S., Gil, J.M., Alberto, H.V., Vilão,
R.C., Piroto Duarte, J., Ayres de Campos, N., Weidinger, A., Lichti, R.L., Irvine, S.J.C.:
Experimental confirmation of the predicted shallow donor hydrogen state in zinc oxide. Phys.
Rev. Lett. 86, 2601–2604 (2001)
44. King, P.D.C., McKenzie, I., Veal, T.D.: Observation of shallow-donor muonium in Ga2O3:
Evidence for hydrogen-induced conductivity. Appl. Phys. Lett. 96, 062110 (2010)
45. King, P.D.C., Lichti, R.L., Celebi, Y.G., Gil, J.M., Vilão, R.C., Alberto, H.V., Piroto Duarte,
J., Payne, D.J., Egdell, R.G., McKenzie, I., McConville, C.F., Cox, S.F.J., Veal, T.D.: Shallow
6 Electronic Properties of Post-transition Metal Oxide Semiconductor Surfaces 145

donor state of hydrogen in In2O3 and SnO2: Implications for conductivity in transparent
conducting oxides. Phys. Rev. B 80, 081201(R) (1–4) (2009)
46. Heine, V.: Theory of surface states. Phys. Rev. 138, A1689–A1696 (1965)
47. Inkson, J.C.: Deep impurities in semiconductors. I. Evanescent states and complex band
structure. J. Phys. C: Solid State Phys. 13, 369–381 (1980)
48. Tersoff, J.: Schottky barriers and semiconductor band structures. Phys. Rev. B 32, 6968–6971
(1985)
49. M€onch, W.: Electronic Properties of Semiconductor Interfaces. Springer, Berlin (2004)
50. Veal, T.D., King, P.D.C., Hatfield, S.A., Bailey, L.R., McConville, C.F., Martel, B., Moreno,
J.C., Frayssinet, E., Semond, F., Zúñiga-Pérez, J.: Valence band offset of the ZnO/AlN
heterojunction determined by x-ray photoemission spectroscopy. Appl. Phys. Lett. 93,
202108 (1–3) (2008)
51. King, P.D.C., Veal, T.D., Jefferson, P.H., McConville, C.F., Wang, T., Parbrook, P.J., Lu, H.,
Schaff, W.J.: Valence band offset of InN/AlN heterojunctions measured by x-ray photoelec-
tron spectroscopy. Appl. Phys. Lett. 90, 132105 (1–3) (2007)
52. King, P.D.C., Veal, T.D., Kendrick, C.E., Bailey, L.R., Durbin, S.M., McConville, C.F.:
InN/GaN valence band offset: High-resolution x-ray photoemission spectroscopy
measurements. Phys. Rev. B 78, 033308 (1–3) (2008)
53. Falabretti, B., Robertson, J.: Electronic structures and doping of SnO2, CuAlO2, and CuInO2.
J. Appl. Phys. 102, 123703 (1–5) (2007)
54. Schleife, A., Fuchs, F., R€ odl, C., Furthm€uller, J., Bechstedt, F.: Branch-point energies and
band discontinuities of III-nitrides and III-/II-oxides from quasiparticle band-structure
calculations. Appl. Phys. Lett. 94, 012104 (1–3) (2009)
55. He, H., Orlando, R., Blanco, M.A., Pandey, R., Amzallag, E., Baraille, I., Rérat, M.: First-
principles study of the structural, electronic, and optical properties of Ga2O3 in its monoclinic
and hexagonal phases. Phys. Rev. B 74, 195123 (1–8) (2006)
56. King, P.D.C., Veal, T.D., Jefferson, P.H., Hatfield, S.A., Piper, L.F.J., McConville, C.F.,
Fuchs, F., Furthm€ uller, J., Bechstedt, F., Lu, H., Schaff, W.J.: Determination of the branch-
point energy of InN: Chemical trends in common-cation and common-anion semiconductors.
Phys. Rev. B 77, 045316 (1–6) (2008)
57. Veal, T.D., McConville, C.F., Schaff, W.J. (eds.): Indium Nitride and Related Alloys. CRC
Press, Boca Raton, Fl (2009)
58. King, P.D.C., Veal, T.D., McConville, C.F.: Unintentional conductivity of indium nitride:
transport modelling and microscopic origins. J. Phys.: Condens. Matter 21, 174201 (1–7)
(2009)
Chapter 7
In Search of a Truly Two-Dimensional
Metallic Oxide

Priya Mahadevan and Kapil Gupta

7.1 Introduction

The existence of a metallic state in two dimensions has been an area that has
attracted considerable attention, especially since the metal–insulator transition
discovered in 1994 by Kravchenko and Pudalov [1]. The tremendous attention to
this experimental observation was motivated by theoretical understanding devel-
oped in the late 1970s and early 1980s [2] which asserted on firm grounds that one
could not have a metallic state in two dimensions. The slightest amount of disorder
would drive the system to an Anderson Insulator. So, the origin of the metallic state
discovered was a puzzle. Early suggestions were that this exotic metallic phase
arose out of electron–electron interaction effects. However, these suggestions had
to be reviewed again. This was because the ideal limit within Fermi liquid theory in
the absence of disorder was a metal. However, in the noninteracting limit, the
system was an insulator. Hence, switching on interactions adiabatically, one could
not go from one limit to the other. These ideas are at the heart of Fermi liquid
theory. So, an alternate mechanism was required to explain the insulator–metal
transition.
A possible reason [3] that was suggested was that the observed metallic behavior
arose from a crossover between a weak and a strong localized phase. The basic
physical picture is similar to what is invoked in semiconductors. The main source of
the disorder in semiconductors is the random distribution of ionized impurity
centers. At high carrier densities, the charged impurities are effectively screened.
However, as the carrier density is decreased without changing the ionized impurity
concentration, one reaches a point where the screening by the charge carriers is
inefficient and this creates inhomogeneities in the carrier distribution and finally

P. Mahadevan (*) • K. Gupta


S.N. Bose National Centre for Basic Sciences, Salt Lake, Kolkata, India
e-mail: priya@bose.res.in; priya.mahadevan@gmail.com

J. Wu et al. (eds.), Functional Metal Oxide Nanostructures, Springer Series 147


in Materials Science 149, DOI 10.1007/978-1-4419-9931-3_7,
# Springer Science+Business Media, LLC 2012
148 P. Mahadevan and K. Gupta

localization. So, a viewpoint that has emerged is that the metal–insulator transition
observed here is a classical percolation transition.
All the earlier work on two-dimensional metallicity was centered around semi-
conductor heterostructures. This was because the growth techniques available at
that time were optimal only for semiconductors. However, recent advances in
growth technology have brought the quality of oxide films closer to that obtained
for semiconductor films in the past few years. For instance, oxide heterostructures
of LaAlO3 grown on SrTiO3 were found to give rise to a two-dimensional electron
gas with very large mobilities for oxides [4, 5]. Recently, very large mobilities
reaching 30,000 V/cm2 were obtained [6] for SrTiO3 in the best made samples to
date. The metallic region in each of these cases is several nanometers thick. The
strong directionality associated with the bonding in oxides gives us an additional
advantage here. A proper choice of the metal atom could give us a truly two-
dimensional oxide film. This is the limit we begin our search for a two-dimensional
metal.
For the purpose of our study, we chose a system which is a metal. There are very
few examples of transition metal oxides which are metals in the absence of doping.
SrRuO3 is one such example [7]. An additional advantage of SrRuO3 is that it is a
4d transition metal atom. Correlation effects are expected to be weaker in a 4d oxide
and this is indeed what we find from a first-principles estimate of U [8]. Bulk films
of SrRuO3 have been found [9] to be well described by Fermi liquid theory, and so it
would be interesting to examine what would happen in truly two-dimensional films
of the metallic oxide.
In our calculations we find that the ultrathin limit is unfortunately insulating and
remains so up to 4 monolayers of SrRuO3 grown on SrTiO3. Examining the partial
density of states, we find it is just the top monolayer that contributes to the valence
and conduction bands. Hence we have a two-dimensional insulator. The two-
dimensional insulating state is interesting as it represents the rare occurrence of a
high spin state for Ru4+. The microscopic considerations responsible for this are
traced to the two dimensionality and the consequent increased localization. This
makes the t42g system very strongly Jahn–Teller active and results in a lattice
distortion that drives the Ru environment strongly square planar and consequently
the system insulating.

7.2 Methodology

SrRuO3 is found to crystallize in an orthorhombic structure with four formula units


per unit cell [10]. The Ru–O–Ru angles in the a–b plane are found to be around
168 . The substrates that the films have been grown on, SrTiO3, is found to occur in
a cubic unit cell. The GGA optimized lattice constant of ferromagnetic cubic
SrRuO3 is 3.99 Å, while that of cubic SrTiO3 is 3.94 Å. Therefore, SrRuO3
grown on SrTiO3 experiences a small in-plane compressive strain. To simulate
the ultrathin films of SrRuO3 grown on SrTiO3, we consider a 13-layer symmetric
7 In Search of a Truly Two-Dimensional Metallic Oxide 149

Fig. 7.1 Ball and stick model


showing the GdFeO3
distortion of the RuO6
octahedra in bulk SrRuO3.
The Ru atoms are at the center
of the octahedra and the O
atoms are at the corners

slab containing a central SrO layer, and alternating TiO2/SrO layers grown in the
(001) direction. The SrRuO3 films were grown on this backbone in our calculations.
In every case we have included 15 Å of vacuum and we have checked that this is
adequate and the results are converged with respect to vacuum. The in-plane lattice
constant of the TiO2 layers as well as the RuO2 layers is set equal to the equilibrium
lattice constant of SrTiO3. As the Ru–O–Ru angles in bulk SrRuO3 deviate from
180 , we have allowed for a GdFeO3 distortion of the RuO6 octahedra (Fig. 7.1) and
optimized the structure to find the minimum energy solution.
The electronic structure of bulk and of thin films of SrRuO3 were investigated
using the plane-wave pseudopotential implementation of density functional theory
and projector-augmented wave potentials in the VASP code [11–14]. The General-
ised gradient approximation (GGA) for the exchange-correlation functional was
used. Electron-correlation effects at the Ru sites were included through the GGA + U
method within the Dudarev et al. formalism [15] where U represents the onsite
Coulomb interaction strength and J represents the intra-atomic exchange. The value
of U ¼ 2.5 eV on the Ru 4d states was determined independently by a recently
proposed scheme based on the random-phase approximation [16] for bulk SrRuO3.
We use a value of J of 0.4 eV. Henceforth in the text the finite U results will
correspond to this value of U and J. Although for the surface, where due to reduced
screening we expect the U to be larger, we expect the qualitative aspects of the results
to remain the same. A Monkhorst-Pack k-point mesh of 6  6  6 was used for
integrations over the Brillouin zone of the bulk, while a mesh of 4  4  2 was used
for the slab calculations. A denser mesh of 8  8  4 was used for the density of
states calculations for the slabs. A cutoff energy of 250 eV was used for the plane
wave basis set.
150 P. Mahadevan and K. Gupta

7.3 Results and Discussion

Bulk SrRuO3 is a ferromagnetic metal and our calculations reproduce the experimen-
tally observed ground state. The basic electronic structure may be understood starting at
the molecular orbital limit. The six Oxygen atoms surrounding each Ru atom lift the
fivefold degeneracy of the Ru d levels into triply degenerate levels with t2g symmetry
and doubly degenerate levels with eg symmetry. These levels interact with those on the
oxygen atoms with the same symmetry forming bonding and antibonding levels. The
bonding levels have predominantly O p character while the antibonding levels are more
localized on the Ru atoms. Intra-atomic exchange interactions split the levels further.
Most ruthenates have a smaller exchange splitting than the crystal field splitting and
therefore favor low spin ground states. Ru has a formal valency of 4 in SrRuO3. This
leads to a configuration of t32g" t12g# . Although the above description was for the ionic
limit, even allowing for band formation the symmetry labelings remain intact. The
partial occupancy of the t2g minority spin band results in the favored ground state to be
ferromagnetic.
While correlation effects are important in compounds containing 3d transition
metal atoms, their effects are much weaker in 4d transition metal compounds. There
are conflicting reports of the importance of correlation effects in SrRuO3. To settle
this controversy we used a recently developed scheme to determine U from first
principles. This gave us a U of 2.5 eV. Taking this into account in GGA + U
calculations did not change the gross features of the density of states discussed
earlier.
Having understood the electronic structure of bulk SrRuO3 we proceeded by
examining the electronic structure of one atomic layer of RuO2 grown on the
SrTiO3 substrate. Before discussing the results from the calculations, we first
speculate on what to expect. Strong crystal field anisotropies at the surface are
expected to result in a level ordering of doubly degenerate dyz and dxz levels
followed by dxy levels. Thus if the system underwent a transition to a nonmagnetic
state the four electrons would go into the dyz ; dxz levels. This would explain the
origin of the insulating ground state. With this naive explanation in mind, we
examined various solutions. The antiferromagnetic state is found to be more stable
than the ferromagnetic state by 49 meV per RuO2 unit. Thus magnetism survives at
this ultrathin film limit. Most importantly, the transition to the antiferromagnetic
state is accompanied by a change in the electronic state from metal to insulator.
Thus the basic assumption that magnetism would be destroyed at the ultrathin limit
is wrong. The loss of coordination makes the Ru atom more correlated and
so magnetism survives. It is however interesting to observe that two-dimensional
magnetism survives in this system. This is of academic interest as the Mermin
Wagner theorem disallows any magnetic order in dimensions less than three. The
present system provides a playground for testing these ideas. An additional advan-
tage over other magnetic systems studied so far is that the substrate is a nonmag-
netic insulator and so does not easily get polarized due to the presence of the
magnetic overlayer.
7 In Search of a Truly Two-Dimensional Metallic Oxide 151

The origin of the insulating state is a puzzle. If Ru is magnetic as our calculations


reveal, then crystal field effects would lead to degeneracy lifting. This would result
in the dyz , dxz orbitals. This is indeed what we find in the structure generated by
merely truncating the SrRuO3 layer in the third direction. However, the U that we
have is insufficient to open up a gap. Allowing for relaxations, one finds something
strange. The Ru in the top layer was coordinated to five oxygen atoms. However it
loses the other oxygen in the z-direction and goes towards a fourfold coordination.
This structural distortion has a dramatic effect on the ordering of levels. Ru which
usually favors a low spin state is found to favor a high spin state in the present case.
The distortion is able to bring down one of the eg levels ðd3z2 r2 Þ below the minority
spin doubly degenerate dyz and dxz levels. The singly degenerate state that we have
as our highest occupied level drives the system antiferromagnetic as well as
insulating. So, even in a 4d transition metal oxide monolayer which one would
believe is moderately correlated, one finds that a metallic phase cannot be sustained
in two dimensions. The loss of coordination in the third direction makes the system
more localized and this drives strong Jahn–Teller effects which one normally does
not expect in a wide band 4d oxide. So, if a two-dimensional 4d oxide sustains such
strong Jahn–Teller effects destroying any possibility of a metallic state, no metallic
3d oxide can be expected to remain metallic when just one monolayer is grown on a
substrate. This is indeed consistent with recent experiments that have examined
monolayers of SrVO3 [17] and LaNiO3 [18] on various substrates and find them to
be insulating. Interface metallicity [19] has been observed, but the metallic layer is
not confined to just two dimensions.
Although we started off in search of a two-dimensional metal, we find that the
ground state of this bulk ferromagnetic metallic system is an antiferromagnetic
insulator. The layer-projected partial density of states is plotted in Fig. 7.2. One
finds that the TiO2 layer has a large band gap of almost 2 eV with some low intensity
states in the band gap region. The second SrO layer and the top RuO2 layer contribute
in this band gap region. On closer examination, we find that it is only the top layer
that contributes to the valence band and conduction band of this small gap semicon-
ductor. So, we have a two-dimensional insulator, and it would be interesting to see if
this two-dimensional insulator will show any insulator–metal transitions as a function
of temperature.
Adding a layer of SrO on top of the RuO2 layer (3 monolayer case) still keeps
the system insulating and antiferromagnetic. The strong out of plane distortions
that were found earlier are still there but reduced. The out of plane Ru–O bond
length for the single monolayer case was 2.12 Å. This is now reduced to
2.03–2.06 Å. Interestingly, the Ru–O–Ru angle in the present case reaches the
values found in bulk SrRuO3. The two monolayer case, however, had the angles at
near ideal values of 179 . Both these effects of bond-length elongation and
changes in the Ru–O–Ru angle contribute to band narrowing and drive the system
insulating.
The four monolayer limit shows some surprises. The favored magnetic state we
found was antiferromagnetic so far. At this thickness we find that the energy of the
antiferromagnetic solution is degenerate with the solution where the surface is
152 P. Mahadevan and K. Gupta

Fig. 7.2 The up-spin (red line) and down-spin (black line) partial density of states projected on
the (a) RuO2, (b) SrO, and (c) TiO2 layers of SrRuO3 grown on SrTiO3 calculated for the
antiferromagnetic state. The zero of energy corresponds to highest energy occupied state

antiferromagnetic while the bulk is ferromagnetic. The former solution is insulating


while the latter solution is metallic. Hence both solutions coexist with disorder pinning
a particular solution in its vicinity. So, experimental probes examining the local
electronic structure should find both solutions. Some experimental photoemission
results [20] exist as a function of thickness. They find an insulator–metal transition
to take place at four monolayers. There are also indirect indications from experiments
that ferromagnetism is lost below 5–6 monolayers [20]. Direct probes of the nature of
magnetic ordering are difficult. Exchange bias effects have been observed [21] at the
below the four monolayer limit at which the system becomes insulating, suggesting the
presence of coexisting ferromagnetic and antiferromagnetic regions. This is consistent
with our results that find an antiferromagnetic surface and ferromagnetic bulk.
We have examined the electronic and magnetic ground state favored by two-
dimensional films of a 4d oxide which is both ferromagnetic and metallic in the
bulk. One does not associate strong correlation effects with 4d oxides. However,
interesting differences are observed at the ultrathin limit. SrRuO3 is found to become
an antiferromagnetic insulator, which evolves into a ferromagnetic metal with thick-
ness. The thickness-dependent insulator–metal transition takes place at four mono-
layers. The ultrathin limit represents the rare realization of a high spin state for Ru.
7 In Search of a Truly Two-Dimensional Metallic Oxide 153

The strong changes in the properties at this limit occur because the system becomes
strongly localized and consequently strongly Jahn–Teller active. As a result, even a
seemingly uncorrelated 4d oxide does not favor a metallic ground state.

Acknowledgments PM thanks the Department of Science and Technology for financial support
and acknowledges an earlier collaboration with A. Janotti, F. Aryasetiawan, and T. Sasaki on a part
of the work.

References

1. Kravchenko, S.V., Kravchenko, G.V., Furneaux, J.E., Pudalov, V.M., D’Iorio, M.: Possible
metal-insulator transition at B¼0 in two dimensions. Phys. Rev. B 50, 8039 (1994)
2. Abrahams, E., Anderson, P.W., Licciardello, D.C., Ramakrishnan, T.V.: Scaling theory of
localization: Absence of quantum diffusion in two dimensions. Phys. Rev. Lett. 42, 673 (1979)
3. Das Sarma, S., Lilly, M.P., Hwang, E.H., Pfeiffer, L.N., West, K.W., Reno, J.L.: Two-
dimensional metal-insulator transition as a percolation transition in a high-mobility electron
system. Phys. Rev. Lett. 94, 136401 (2005)
4. Ohtomo, A., Muller, D.A., Grazul, J.L., Hwang, H.Y.: Artificial charge-modulationin atomic-
scale perovskite titanate superlattices. Nature 419, 378 (2002)
5. Ohtomo, A., Hwang, H.Y.: A high-mobility electron gas at the LaAlO3/SrTiO3
heterointerface. Nature 427, 423 (2004)
6. Son, J., Moetakef, P., Jalan, B., Bierwagen, O., Wright, N.J., Engel-Herbert, R., Stemmer, S.:
Epitaxial SrTiO3 films with electron mobilities exceeding 30,000 cm2v-1s-1. Nat. Mater. 9, 482
(2010)
7. Longo, J.M., Raccah, P.M., Goodenough, J.B.: Magntetic properties of SrRuO3 and CaRuO3.
J. Appl. Phys. 39, 1327 (1968)
8. Mahadevan, P., Aryasetiawan, F., Janotti, A., Sasaki, T.: Evolution of the electronic structure
of a ferromagnetic metal: Case of SrRuO3. Phys. Rev. B 80, 035106 (2009)
9. Kostic, P., Okada, Y., Collins, N.C., Schlesinger, Z., Reiner, J.W., Klein, L., Kapitulnik, A.,
Geballe, T.H., Beasley, M.R.: Non-Fermi-liquid behavior of SrRuO3: Evidence from infrared
conductivity. Phys. Rev. Lett. 81, 2498 (1998)
10. Nakatsugawa, H., Iguchi, E., Oohara, Y.: Electronic structures and magnetic properties in
Sr1xLaxRuO3(0.0  x  0.5). J. Phys.: Condens. Matter 14, 415 (2002)
11. Kresse, G., Furthm€ uller, J.: Efficient iterative schemes for ab initio total-energy calculations
using a plane-wave basis set. Phys. Rev. B. 54, 11169 (1996)
12. Kresse, G., Furthm€ uller, J.: Efficiency of ab initio total-energy calculations for metals and
semiconductors using a plane-wave basis set. Comput. Mat. Sci. 6, 15 (1996)
13. Blochl, P.E.: Projector augmented-wave method. Phys. Rev. B 50, 17953 (1994)
14. Kresse, G., Joubert, D.: From ultrasoft pseudopotentials to the projector augmented-wave
method. Phys. Rev. B 59, 1758 (1999)
15. Dudarev, S.L., Botton, G.A., Savrasov, S.Y., Humphreys, C.J., Sutton, A.P.: Electron-energy-
loss spectra and the structural stability of nickel oxide: An LSDA + U Study. Phys. Rev. B 57,
1505 (1998)
16. Aryasetiawan, F., Imada, M., Georges, A., Kotliar, G., Biermann, S., Lichtenstein, A.I.:
Frequency-dependent local interactions and low-energy effective models from electronic
structure calculations. Phys. Rev. B 70, 195104 (2004)
17. Yoshimatsu, K., Okabe, T., Kumigashira, H., Okamoto, S., Aizaki, S., Fujimor, A., Oshima,
M.: Dimensional-crossover-driven metal-insulator transition in SrVO3 ultrathin films. Phys.
Rev. Lett. 104, 147601 (2010)
154 P. Mahadevan and K. Gupta

18. Scherwitzl, R., Zubko, P., Lichtensteiger, C., Triscone, J.M.: Electric-field tuning of the metal-
insulator transition in ultrathin films of LaNiO3. Appl. Phys. Lett. 95, 222114 (2009)
19. Basletic, M., Maurice, J.L., Carretero, C., Harranz, G., Copie, O., Bibes, M., Jacquet, E.,
Bouzehouane, K., Fusil, S., Barthelemy, A.: Mapping the spatial distribution of charge carriers
in LaAlO3/SrTiO3 heterostructures. Nat. Mat. 7, 621 (2008)
20. Toyota, D., Ohkubo, I., Kumigashira, H., Oshima, M., Lipmaa, M., Takizawa, M., Fujimori,
A., Ono, K., Kawasaki, M., Koinuma, H.: Thickness-dependent electronic structure of ultra-
thin SrRuO3 films studied by in situ photoemission spectroscopy. Appl. Phys. Lett. 87, 162508
(2005)
21. Xia, J., Siemons, W., Koster, G., Beasley, M.R., Kapitulnik, A.: Critical thickness for itinerant
ferromagnetism in ultrathin films of SrRuO3. Phys. Rev B 79, 140407 (2009)
Part II
Synthesis and Processing
Chapter 8
Solution Phase Approach to TiO2
Nanostructures

John D. Bass and Ho-Cheol Kim

8.1 Introduction

Titanium dioxide (titania, TiO2) features prominently in a number of commercialized


and next-generation high-tech domains and is, by consequence, the most well-studied
metal oxide [1]. Industrially, TiO2 finds a prominent role mainly (~95%) in pigments
(over 5 million metric tons/year), with catalysts, ceramics, coated fabrics and textiles,
floor coverings, printing ink, and roofing granules constituting other major uses on a
consumption basis [2]. In terms of more high-tech domains, TiO2 is a material of
interest in photonics [3], membranes [4], biological supports [5–7], sensing [8, 9],
electrochromics [10], and environmental applications [11] including photoelectrolysis
of water [12] and catalytic and photocatalytic applications [13–16]. For example, in
photovoltaics (PV), TiO2 has emerged as the material of choice in dye-sensitized solar
cells (DSCs), a low cost PV technology that has recently reached commercialization
[17, 18]. TiO2 has also been employed in other PV systems, pairing as the n-type
semiconductor with extremely thin CdTe absorber layers [19], nanocomposite CuInS2
devices [20, 21], and PbS quantum dots [22].
In many of these emerging high-tech domains, the ability to control the structure of
TiO2 on the nanometer scales is the driving force behind technology development. In
the case of both the DSCs and the CuInS2 system, the nanostructuration of the TiO2
acceptor layer is vital in achieving any reasonable efficiency [17, 18, 21]. The high
surface area allows intimate contact between the thin absorbing material or monolayer
of charge transfer dye and the TiO2. This reduces or eliminates the need for carrier
diffusion to the interface while maintaining a sufficient quantity of absorber material
to achieve high light absorption. The nanostructuration can increase the available
surface area by 1,000 times or more for efficient absorption of sunlight [18].

J.D. Bass • H.-C. Kim (*)


IBM Research Division, Almaden Research Center, 650 Harry Road, San Jose,
CA 95120-6099, USA
e-mail: hckim@almaden.ibm.com

J. Wu et al. (eds.), Functional Metal Oxide Nanostructures, Springer Series 157


in Materials Science 149, DOI 10.1007/978-1-4419-9931-3_8,
# Springer Science+Business Media, LLC 2012
158 J.D. Bass and H.-C. Kim

Fig. 8.1 Schematic representation of a nanostructured heterojunction photovoltaic device that


decouples d, the path length relevant to minority carrier transport from h, the path length relevant
to light absorption

Indeed, nanostructuring the heterojunction between n-type and p-type materials


in photovoltaic devices, Fig. 8.1, has the potential to completely redefine the
relevant landscape of materials while markedly improving efficiency [23].
Light absorption in a photovoltaic semiconductor generates electron–hole pairs.
The minority charge carrier (e.g., electrons for a p-type absorber), either bound as an
exciton to the majority carrier or on its own, must diffuse to the p-n junction to allow
charge separation in the device. This junction is an interface or interfacial region
between a p-type material and an n-type material. In cases where the distance to the
junction is longer than the diffusion length afforded by the carrier lifetime, many of the
carriers recombine, limiting the output current. An alternative to bringing the carriers
to the junction is, in effect, to bring the junction to the carriers by nanostructuring the
junction. Nanostructures projecting into the absorbing layer can reduce the diffusion
distance while maintaining an adequate optical thickness for light absorption. This
allows the use of absorber materials with inherently lower diffusion lengths and/or the
use of lower quality (i.e., cheaper, more accessible, and easier to process) materials,
which also tend to suffer from higher recombination rates.
In this embodiment, 1D structures such as rods, wires, pillars, nanotubes, and
nanofibers have unique advantages, such as improved charge collection [24–26]
and lower tortuosity. In other applications, including photonics, catalysis, and
sensing, controlling 3D structures on one or more length scales is desirable.
These structures can be designed to guide light, maximize the available surface
area, or present a particular crystal polymorph or crystal facet known to optimize
the reactivity of the surface [3, 27–31]. Quantum size effects in TiO2 crystallites
can be achieved, though these observations are observed at small sizes, below
10 nm due to the small Bohr radius [32–34]. Consequently, novel physical and
chemical properties arising from such quantum confinement effects need to be
examined at this small size scale [35].
8 Solution Phase Approach to TiO2 Nanostructures 159

8.2 Approaches

A variety of methods have been developed to synthesize nanostructured TiO2.


Nonsolution phase approaches using gas phase precursors have been used to grow
a variety of shapes including rods, fibers, and ribbons on substrates at elevated
temperature [36, 37]. The most common methods, however, rely on solution phase
deposition, which is the main focus of this chapter. Solution phase synthesis has a
number of advantages in terms of process control, ease of handling, and flexibility.
In the simplest manifestation, nanoporous coatings can be achieved via the deposi-
tion colloidal TiO2 nanoparticles. This sort of deposition is compatible with com-
mon coating methods such as doctor blading, screen printing, and spray coating
[21]. The porosity can be controlled by varying the particle size, the addition of
additives, and the drying conditions. Using a narrow size dispersion of
nanoparticles, self-organized particle films can be obtained [38]. These materials
show very little mesoporosity as they are tightly packed together.

8.2.1 Porous Architectures Through Templated Self-Assembly

The templating of nanostructured TiO2 by self-assembly using structure-directing


materials such as surfactants [39] and block-copolymers [40–43] is a rich field that
has been intensively investigated in the recent years. Much is now known about the
assembly [44] and crystallization [45, 46] processes that yield polycrystalline TiO2
with highly ordered mesoporous structure. Films can be applied via a variety of
coating techniques including dip coating, spray coating, meniscus coating, and spin
coating. Templates used are either preformed nanosized templates or templates that
self-assemble during the film forming process such as in evaporation-induced self-
assembly (EISA).
Apart from the initial assembly of the organic and inorganic phases into
nanostructured domains, thermal posttreatment to condense the network, remove
the organic phase, and induce crystallization has emerged as a critical step in
determining the final characteristics of nanostructured TiO2 materials [47–49].
These thermally driven processes, covering dehydration, condensation, densifica-
tion, decomposition/pyrolysis, crystallization, and sintering, have recently received
in-depth attention using in situ techniques including SAXS/WAXS investigations
[46, 50] and ellipsometry [45]. With in situ ellipsometry, for example, the effect of
the precursor solution, substrate, composition of the calcination atmosphere, and
confinement can be elucidated in terms of the formation of the TiO2 matrix
(Fig. 8.2). Moreover, characterization of chemical processes occurring within the
pores, such as the kinetics of pyrolysis of the template, is also accessible.
Recent efforts have been put forth to synthesize vertically oriented structures
with open accessibility and direct conduction pathways to the surface. Structural
transformation of spherical domains upon heating to grid-like [44] and pillar-like
160 J.D. Bass and H.-C. Kim

Dense Film Nanostructured Film

h0 = 82 nm h0 = 262 nm

1.75 6
2.4
8 1.7
2.3
4
1.65

Index of Refraction (700 nm)


Index of Refraction (700 nm)

Derivative dn/dT (x1000)


2.2

Derivative dn/dT (x1000)


6
1.6 2
2.1
4 1.55
2 0
1.5
1.9 2

cr
−2

sin a I
1.45

ys

te liza
t
ri ti
1.8

ng on
1.4
0 −4
1.7
cr
sin a I

1.35
ys

te liza
t
ri ti

−2 pyrolysis
ng on

1.6 1.3 −6
0 200 400 600 0 200 400 600
Temperature (°C) Temperature (°C)

Fig. 8.2 In situ ellipsometry shows the thermal evolution by way of the change in index of
refraction of a nanostructured TiO2 thin film prepared via templated self-assembly as compared to
a dense TiO2 film. Adapted from [44]

structures [51], as well as tilted cylindrical arrays [52] has been achieved using the
Pluronic™ family of ethylene oxide and propylene oxide block copolymers.
The phase behavior in these systems is fairly predictive and is adjusted by changing
the volume ratio between the copolymer and inorganic components [53]. These
yield pores in the size range of 5–12 nm depending on the starting material and the
degree of sintering [45]. In a similar size range (10 nm), a bicontinuous double
gyroid TiO2 structure was formed in a multistep process involving self-assembly of
a block copolymer, selective removal of one block, and electrofilling with a TiO2
precursor [54]. Unlike in silica systems [55], the control of cylindrical pore
orientation normal to the supported film surface and with variable pore diameters
ranging up to 20 nm remains an open challenge for TiO2.

8.2.2 1D Structures from Anodization

A common and well-investigated approach to 1D TiO2 structures is through the


oxidation of titanium (Ti) foil. Nanotubes and nanowires have been prepared on Ti
supports seeded with TiO2 nanoparticles via hydrothermal synthesis under basic
conditions [26, 56]. For the nanotubes, outer tube diameters are around 12 nm,
while the inner diameter is ~4 nm. Tubes can be grown up to 10 mm and seem to
8 Solution Phase Approach to TiO2 Nanostructures 161

form from folded sheets. Their diffraction structure is suggested to be H2Ti3O7.


Nanowires were calcined at 500 C to transform them to the anatase phase and used
in DSCs [26].
In a technique that finds parallel in the anodization of Al, TiO2 nanotubes can be
formed by the oxidation of titanium in the presence of fluoride-based electrolytes
under anodic conditions [57]. Polycrystalline transparent TiO2 nanotube arrays
prepared in this manner have been investigated in DSCs [58]. Improved charge-
collection efficiencies over nanoparticle systems were reported [25]. Using electro-
chemical methods, these structures can be prepared directly on transparent
conducting oxides with lengths up to 33 mm long, reaching DSC efficiencies of
nearly 7% [59].

8.2.3 Imprinting and Molding

The development and application of top-down nanoimprint lithography approaches


has evolved to support the generation of submicron oxide features from sol-gel
precursors [60, 61]. Whitesides and coworkers demonstrated submicron patterning
of titanium silicates using poly(dimethyl siloxane) (PDMS) soft lithography [61].
Extension of this technique using perfluoropolyether (PFPEs) elastomers that
provide improved filling and release characteristics has been used in the DeSimone
group to pattern TiO2 and other oxides and mixed metal oxides [60]. Sub-200 nm
features and aspect ratios of up to 2.5 can be achieved [60].
The use of sacrificial templates such as water soluble templates for transfer
molding (TM) is another approach that can yield submicron patterned features
[62–64]. In this approach, water soluble poly(vinyl alcohol) (PVA) templates are
used for pattern transfer as shown schematically in Fig. 8.3. Daughter templates are
prepared en masse from hard template masters such as large area lithographically
patterned silicon. As illustrated in Fig. 8.3, this approach is amenable to creating
nanostructured TiO2 without expensive imprinting tools [65]. Here a solution
containing photosensitive TiO2 precursor is spun onto a prepatterned PVA tem-
plate. The photosensitive TiO2 precursor is an oligomeric titanate (OT) prepared
from an acetylacetone-chelated titanium alkoxide [66, 67]. The coated template is
bonded to a substrate and exposed to long wavelength UV radiation to induce
partial condensation of the precursor. Exposure to warm water dissolves the PVA
leaving a partially condensed amorphous network of nanostructured TiO2. Crystal-
lization to the anatase phase and removal of residual organics and accomplished by
thermal treatment to 450 C.
Figure 8.4 shows SEM micrographs of nanoscopic TiO2 posts prepared by TM
method using a PVA template. The PVA template was removed by water and TiO2
was calcined at 450 C. As shown in the micrograph, the TM technique provides
well-defined nanostructures over large area without defects. The inset shows higher
magnification of the TiO2 posts that have dimensions of 65 nm in diameter, 90 nm
in height, and approximately 200 nm in center-to-center distance.
162 J.D. Bass and H.-C. Kim

Fig. 8.3 Microtransfer molding using a water soluble PVA template applied to the nanostruc-
turation of TiO2

Fig. 8.4 Defect-free TiO2 nanoposts prepared using PVA templates

High aspect ratio nanoposts are desirable partly due to increase in surface area,
greater penetration into the surrounding media, and for optical (e.g., antireflection)
and photonic applications. To prepare high aspect ratio TiO2 posts by TM, it is
necessary to prepare high aspect ratio masters for generating PVA templates.
8 Solution Phase Approach to TiO2 Nanostructures 163

Fig. 8.5 (a) Silicon master and (b) hexagonally arranged TiO2 posts prepared through the PVA
template method. MT followed by calcination allows for the production of TiO2 posts over large
areas

A high aspect ratio silicon master can be prepared using conventional lithography
and plasma etching of silicon using oxide as an etch mask, Fig. 8.5a. Figure 8.5b
shows a top view and a cross-sectional SEM micrograph of high aspect ratio TiO2
posts derived from this type of silicon master. With this technique, TiO2 posts of
approximately 360 nm in height and 70 nm in diameter (aspect ratio ~5) can be
achieved over large areas [65].
High aspect ratio structures through MT can also be achieved through successive
stacking of templated layers. This is shown in the most trivial case with a two-layer
cell structure, Fig. 8.6a, b, that can be made in either a closed-cell or open-cell
morphology. The difference between the two structures is the concentration of the
titania precursor solution. Open-cell structures are prepared using low concentra-
tion solutions that, when spun on, just fill the 250 nm mesh PVA template. Closed-
cell structures are prepared at higher precursor concentrations where the amount of
fill material exceeds the template volume. In this situation, overfill forms a contin-
uous sheet between the templated structure and the previous layer. Stacking can be
continued to create quite thick films, Fig. 8.6c, and different templates can
be combined to form novel structures, Fig. 8.6d. These structures are possible
because under normal conditions the PVA near the templated area is crosslinked
by the titania precursor solution, becoming resistant to dissolution. This protects the
templated features against infill by successive layers. The residual PVA is removed
later by calcination.
TiO2 posts normal to the surface can also be generated by sol filling in high
aspect ratio templates. The preparation of well-defined high aspect ratio templates,
however, becomes nontrivial as feature sizes approach tens of nanometers and
below. One promising method to generate such templates is using oxygen plasma
etching of a polymeric transfer layer using a block copolymer pattern mask, Fig. 8.7
[66]. Reproducible templates with diameters from 8 to 25 nm and layer thickness of
hundreds of nanometers can be reliably achieved on a variety of surfaces. This
pattern layer/transfer layer motif is common in photolithography, using photopat-
ternable materials that contain high etch contrast materials such as silicon. Here, the
164 J.D. Bass and H.-C. Kim

Fig. 8.6 SEM micrographs of two-layer (a) closed-cell and (b) open-cell structures prepared from
the same template but using different concentrations of fill solution. Stacking can be used to create
successively thicker films such as (c), a five-layer structure. Mixed templates (d) can also be used
to create interesting stacked structures

block-copolymer pattern layer provides a self-assembly-based organization beyond


the current resolution of traditional photolithography. To the block copolymer
mixture is added a silicon containing organosilicate [OS, or polymethylsiloxane
(PMS)] resin that segregates into one of the domains, providing high etch contrast.
Controlled oxygen plasma etching followed by removal of the residual
organosilane yields the high aspect ratio polymeric template shown as an inset in
Fig. 8.7.
The high aspect ratio polymeric template can be used for molding TiO2 or other
materials. Partially chelated TiO2 precursor is spun onto the template and allowed
to fill in the pores by capillary action at elevated temperature (190 C) shown in
Fig. 8.8a. Partial lift off of this layer demonstrates that the TiO2 precursor infiltrates
down to the bottom substrate, as a clear pattern of TiO2 bumps can be seen
(Fig. 8.8b). Calcination to 450 C of the TiO2 infiltrated template removes the
polymeric template leaving behind TiO2 posts connected with a thin porous TiO2
top layer as shown in Fig. 8.8c. The interconnected thin top layer can be removed by
using low-voltage, broad beam ion milling prior to calcination (Fig. 8.8d).
8 Solution Phase Approach to TiO2 Nanostructures 165

Fig. 8.7 Preparation of high aspect ratio polymer templates. (Inset) Cross-sectional SEM micro-
graph of polymer template with 15 nm diameter holes

Fig. 8.8 (a) Cross-sectional SEM micrograph of nanoporous template after infilteration of TiO2
precursor, (b) cross-sectional SEM micrograph of sample after partial lift off, (c) cross-sectional
SEM micrograph of TiO2 nanoposts after thermal treatment at 450 C, (d) plane view SEM
micrograph of TiO2 nanoposts calcined after gentle ion milling of the sample shown in (a). Adapted
from [64]
166 J.D. Bass and H.-C. Kim

Fig. 8.9 Schematic


illustration of the
electrochemical deposition of
TiO2 into nanoporous
anodized alumina template.
Adapted from [71]

8.2.4 Templated Electrochemical Synthesis

One clear advantage of the high aspect ratio polymeric template described in
Sect. 8.2.3 is that it can be easily created on a variety of substrates, including
conductive substrates. High aspect ratio polymeric templates on conducting
substrates can be combined with electrochemical deposition to create titania posts
by electrochemical filling. Figure 8.9 shows a schematic illustration of electro-
chemical deposition of TiO2 using an anodized alumina template. Cathodic TiO2
deposition from acidic solutions has previously been demonstrated [68–70]. This
process is attractive in that low cost, easy to handle titanium oxysulfate precursors
can be used instead of alkoxide or titanium chloride precursors. A similar approach
has been used to deposit TiO2 in self-assembled scaffolds prepared after selective
removal of one block of a block-copolymer assembly [54].
Figure 8.10 shows SEM micrographs of TiO2 nanoposts generated by electrode-
position method using a high aspect ratio polymeric template. This approach
provides remarkably well-defined TiO2 nanoposts of approximately 20 nm in
diameter over large area of the substrate.
8 Solution Phase Approach to TiO2 Nanostructures 167

Fig. 8.10 TiO2 posts on a gold-coated silicon wafer achieved by electrochemical filling high
aspect ratio polymeric template

8.2.5 Single Crystalline 1D Structures by Solution Phase


Hydrothermal Growth

Controllable solution phase growth of single crystal 1D metal oxide structures on


surfaces has been achieved for a variety of metal oxides, most notably ZnO
[72–77]. Extremely long ZnO structures, up to 25 mm, with aspect ratios of 125
can be grown [74] and a high degree of vertical orientation with respect to the
substrate can be achieved [73].
Single crystalline 1D structures have potentially a number of advantages over their
polycrystalline counterparts. In the domain of photovoltaics, it is known that the
mobility of photogenerated electrons in polycrystalline, porous TiO2 networks is
several orders of magnitude slower than in single crystalline materials [78, 79]. Trap-
limited diffusion [79], tortuosity [80], and local field effects [78] all can serve to limit
electron mobility in polycrystalline, porous TiO2 networks. In DSCs electron trans-
port through the oxide layer occurs on the order of milliseconds [74, 81]. Work on
DSCs fabricated from single-crystal ZnO nanowires showed that the faster electron
transport afforded by the single-crystalline nanowires led to improved charge-
collection efficiency [82, 83].
Not to be overlooked, the well-defined surfaces of single-crystalline materials
can allow researchers in applied fields to draw on the significant body knowledge
garnered from fundamental surface science research. The TiO2 surface, especially
the rutile and anatase polymorphs, is the most thoroughly investigated of the metal
oxides [1]. This is because in comparison with ZnO and other metal oxide
materials, TiO2 has emerged as a material of choice for a variety of applications,
largely due to the fact that it is inexpensive and stable – biologically, chemically,
and photochemically [6, 84].
168 J.D. Bass and H.-C. Kim

Fig. 8.11 SEM images of rutile nanorods grown hydrothermally on TiO2-coated FTO substrate at
(a) low and (b) high precursor concentration

Solution phase synthesis of single-crystalline TiO2 structures on surfaces is less


well developed than that of ZnO, though much progress has been made recently.
Hydrothermal synthesis and characterization of rutile nanorods on glass substrates
starting from aqueous solutions of TiCl3 has been reported [85–87]. Typical
conditions used temperatures in the 160–200 C range and reaction times of several
hours and yielded rods of lengths ~500 nm. Temperatures down to 80 C could be
used, though the reaction times were much longer, up to 168 h [86]. Aspect ratios
were in the range of 10–20.
Rutile nanorods from alkoxide and TiCl4 precursors can be prepared directly on
transparent conducting oxides such as fluorine-doped tin oxide (FTO) using two
recently developed hydrothermal procedures [88–90]. In one, a biphasic solution of
toluene and hydrochloric acid with both TiCl4 and tetrabutyl titanate was allowed to
react at 180 C for between 30 min and 22 h. Depending on the time, this yielded
rutile rods from 2 to 5 mm with diameters averaging from 10 to 35 nm. This
procedure could also be used to grow rods on FTO coated with a TiO2 prelayer
grown from TiCl4 solution and was shown to be amenable to doping [88, 89]. DSC
efficiency of 5.02% was achieved with Ta doping.
A second methodology based on TiCl4 or various alkoxide precursors and a
hydrochloric acid solution was also shown to yield single-crystalline rutile
nanorods [90]. Rods of length up to 4 mm were grown; diameters were slightly
larger than that of rods grown in the biphasic system. Longer rods detached from
the surface giving freestanding TiO2 nanorod films. Rods were only observed to
form on FTO-coated substrates. This was believed to result from an epitaxial
relation between the FTO substrate and rutile TiO2.
In our hands, this second technique could also be adapted to FTO substrates
covered by TiO2 thin films, as shown in Fig. 8.11. The TiO2 prelayer was prepared
by spin coating a solution of the titania precursor Tyzor BTP partially chelated with
acetylacetone. Growth conditions were best between 130 and 150 C over a period
of 15 h. The density of the rods could be controlled through the amount of Ti(OBu)4
precursor used. As evident in the SEM images, the vertical orientation in these
8 Solution Phase Approach to TiO2 Nanostructures 169

systems results mainly from impingement of the growing nanorods. With a high
density of rods, those growing with a more vertical orientation are less likely to
experience arrested growth from running into a neighbor.

8.3 Conclusion

Solution-based approaches to TiO2 nanostructures represent a very attractive route


to achieving well-defined, controllable structures at low cost. As discussed above, a
variety of different strategies including self-assembly, anodization, transfer mold-
ing, electrodeposition, and hydrothermal growth are being actively pursued, each
showing its own promises and challenges. For each approach much work remains,
especially on device integration and realizing precise structural requirements for
targeted applications. Certainly excitement and surprises will continue in this rich
research field.

References

1. Diebold, U.: The surface science of titanium dioxide. Surf Sci Rep 48(5–8), 53–229 (2003)
2. Gambogi, J.: Titanium and titanium dioxide. U.S. Geological Survey, Mineral Commodity
Summaries. http://minerals.usgs.gov (2009)
3. Subramania, G., et al.: Log-pile TiO2 photonic crystal for light control at near-UV and visible
wavelengths. Adv. Mater. 22(4), 487–491 (2009)
4. Guliants, V.V., Carreon, M.A., Lin, Y.S.: Ordered mesoporous and macroporous inorganic
films and membranes. J. Membr. Sci. 235(1–2), 53–72 (2004)
5. Bass, J.D., et al.: Nanostructuration of titania films prepared by self-assembly to affect cell
adhesion. J. Biomed. Mater. Res. A 93A, 96 (2010)
6. Bass, J.D., et al.: Stability of mesoporous oxide and mixed metal oxide materials under
biologically relevant conditions. Chem. Mater. 19, 4349–4356 (2007)
7. Yan, X.X., et al.: Highly ordered mesoporous bioactive glasses with superior in vitro bone-
forming bioactivities. Angew. Chem. Int. Ed. 43(44), 5980–5984 (2004)
8. Nicole, L., et al.: Advanced selective optical sensors based on periodically organized
mesoporous hybrid silica thin films. Chem. Commun. 20, 2312–2313 (2004)
9. Wirnsberger, G., Scott, B.J., Stucky, G.D.: pH Sensing with mesoporous thin films. Chem.
Commun. 1, 119–120 (2001)
10. Ohsuku, T., Hirai, T.: An electrochromic display based on titanium dioxide. Electrochim. Acta
27(9), 1263–1266 (1982)
11. Bosc, F., et al.: Mesoporous TiO2-based photocatalysts for UV and visible light gas-phase
toluene degradation. Thin Solid Films 495(1–2), 272–279 (2006)
12. Fujishima, A., Honda, K.: Electrochemical photolysis of water at a semiconductor electrode.
Nature 238(5358), 37–38 (1972)
13. Fujishima, A., Rao, T.N., Tryk, D.A.: Titanium dioxide photocatalysis. J. Photochem.
Photobiol. C 1(1), 1–21 (2000)
14. Martinez-Ferrero, E., et al.: Nanostructured titanium oxynitride porous thin films as efficient
visible-active photocatalysts. Adv. Funct. Mater. 17(16), 3348–3354 (2007)
170 J.D. Bass and H.-C. Kim

15. Sakatani, Y., et al.: Optimised photocatalytic activity of grid-like mesoporous TiO2 films:
effect of crystallinity, pore size distribution, and pore accessibility. J. Mater. Chem. 16(1),
77–82 (2006)
16. Hoffmann, M.R., et al.: Environmental applications of semiconductor photocatalysis. Chem.
Rev. 95(1), 69–96 (1995)
17. O’Regan, B., Gratzel, M.: A low-cost, high-efficiency solar cell based on dye-sensitized
colloidal TiO2 films. Nature 353(6346), 737–740 (1991)
18. Gr€atzel, M.: Conversion of sunlight to electric power by nanocrystalline dye-sensitized solar
cells. J. Photochem. Photobiol. A 164(1–3), 3–14 (2004)
19. Ernst, K., et al.: Contacts to a solar cell with extremely thin CdTe absorber. Thin Solid Films
387(1–2), 26–28 (2001)
20. Nanu, M., Schoonman, J., Goossens, A.: Inorganic nanocomposites of n- and p-type
semiconductors: a new type of three-dimensional solar cell. Adv. Mater. 16(5), 453–456
(2004)
21. Nanu, M., Schoonman, J., Goossens, A.: Nanocomposite three-dimensional solar cells
obtained by chemical spray deposition. Nano Lett. 5(9), 1716–1719 (2005)
22. Pattantyus-Abraham, A.G., et al.: Depleted-heterojunction colloidal quantum dot solar cells.
ACS Nano 4(6), 3374–3380 (2010)
23. Fan, Z., et al.: Three-dimensional nanopillar-array photovoltaics on low-cost and flexible
substrates. Nat. Mater. 8(8), 648–653 (2009)
24. Jennings, J.R., et al.: Dye-sensitized solar cells based on oriented TiO2 nanotube arrays:
transport, trapping, and transfer of electrons. J. Am. Chem. Soc. 130(40), 13364–13372 (2008)
25. Zhu, K., et al.: Enhanced charge-collection efficiencies and light scattering in dye-sensitized
solar cells using oriented TiO2 nanotubes arrays. Nano Lett. 7(1), 69–74 (2006)
26. Enache-Pommer, E., Boercker, J.E., Aydil, E.S.: Electron transport and recombination in
polycrystalline TiO2 nanowire dye-sensitized solar cells. Appl. Phys. Lett. 91(12), 123116-3
(2007)
27. Yang, H.G., et al.: Anatase TiO2 single crystals with a large percentage of reactive facets.
Nature 453(7195), 638–641 (2008)
28. Wang, H., Wu, Y., Xu, B.-Q.: Preparation and characterization of nanosized anatase TiO2
cuboids for photocatalysis. App. Catal. B 59(3–4), 139–146 (2005)
29. Dambournet, D., Belharouak, I., Amine, K.: Tailored preparation methods of TiO2 anatase,
rutile, brookite: mechanism of formation and electrochemical properties. Chem. Mater. 22(3),
1173–1179 (2010)
30. Wu, B., et al.: Nonaqueous production of nanostructured anatase with high-energy facets.
J. Am. Chem. Soc. 130(51), 17563–17567 (2008)
31. Barnard, A.S., Curtiss, L.A.: Prediction of TiO2 nanoparticle phase and shape transitions
controlled by surface chemistry. Nano Lett. 5(7), 1261–1266 (2005)
32. Satoh, N., et al.: Quantum size effect in TiO2 nanoparticles prepared by finely controlled metal
assembly on dendrimer templates. Nat. Nanotechnol. 3(2), 106–111 (2008)
33. Luca, V.: Comparison of size-dependent structural and electronic properties of anatase and
rutile nanoparticles. J. Phys. Chem. C 113(16), 6367–6380 (2009)
34. Monticone, S., et al.: Quantum size effect in TiO2 nanoparticles: does it exist? App. Surf. Sci.
162–163, 565–570 (2000)
35. Liu, C., Yang, S.: Synthesis of angstrom-scale anatase titania atomic wires. ACS Nano 3(4),
1025–1031 (2009)
36. Chi, B., Jin, T.: Synthesis of titania nanostructure films via TiCl4 evaporation – deposition
route. Cryst. Growth Des. 7(4), 815–819 (2007)
37. Wu, J.-M., et al.: Thermal evaporation growth and the luminescence property of TiO2
nanowires. J. Cryst. Growth 281(2–4), 384–390 (2005)
38. Burnside, S.D., et al.: Self-organization of TiO2 nanoparticles in thin films. Chem. Mater.
10(9), 2419–2425 (1998)
8 Solution Phase Approach to TiO2 Nanostructures 171

39. Antonelli, D.M., Ying, J.Y.: Synthesis of hexagonally packed mesoporous TiO2 by a modified
sol-gel method. Angew. Chem. Int. Ed. Engl. 34(18), 2014–2017 (1995)
40. Grosso, D., et al.: Highly organized mesoporous titania thin films showing mono-oriented 2D
hexagonal channels. Adv. Mater. 13(14), 1085–1090 (2001)
41. Smarsly, B., et al.: Highly crystalline cubic mesoporous TiO2 with 10-nm pore diameter made
with a new block copolymer template. Chem. Mater. 16(15), 2948–2952 (2004)
42. Yang, P., et al.: Generalized syntheses of large-pore mesoporous metal oxides with semicrys-
talline frameworks. Nature 396(6707), 152–155 (1998)
43. Yang, P., et al.: Block copolymer templating syntheses of mesoporous metal oxides with large
ordering lengths and semicrystalline framework. Chem. Mater. 11(10), 2813–2826 (1999)
44. Crepaldi, E.L., et al.: Controlled formation of highly organized mesoporous titania thin films:
from mesostructured hybrids to mesoporous nanoanatase TiO2. J. Am. Chem. Soc. 125(32),
9770–9786 (2003)
45. Bass, J.D., et al.: Pyrolysis, crystallization, and sintering of mesostructured titania thin films
assessed by in situ thermal ellipsometry. J. Am. Chem. Soc. 130(25), 7882–7897 (2008)
46. Choi, S.Y., et al.: Evolution of nanocrystallinity in periodic mesoporous anatase thin films.
Small 1(2), 226–232 (2005)
47. Grosso, D., et al.: Preparation, treatment and characterisation of nanocrystalline mesoporous
ordered layers. J. Sol-Gel Sci. Technol. 40(2–3), 141–154 (2006)
48. Grosso, D., et al.: Fundamentals of mesostructuring through evaporation-induced self-assem-
bly. Adv. Funct. Mater. 14(4), 309–322 (2004)
49. Sanchez, C., et al.: Design, synthesis and properties of inorganic and hybrid thin films having
periodically organized nanoporosity. Chem. Mater. 20(3), 682–737 (2008)
50. Grosso, D., et al.: Highly porous TiO2 anatase optical thin films with cubic mesostructure
stabilized at 700 degrees C. Chem. Mater. 15(24), 4562–4570 (2003)
51. Wu, C.-W., et al.: Formation of highly ordered mesoporous titania films consisting of crystal-
line nanopillars with inverse mesospace by structural transformation. J. Am. Chem. Soc. 128
(14), 4544–4545 (2006)
52. Koganti, V.R., et al.: Generalized coating route to silica and titania films with orthogonally
tilted cylindrical nanopore arrays. Nano Lett. 6(11), 2567–2570 (2006)
53. Alberius, P.C.A., et al.: General predictive syntheses of cubic, hexagonal, and lamellar silica
and titania mesostructured thin films. Chem. Mater. 14(8), 3284–3294 (2002)
54. Crossland, E.J.W., et al.: A bicontinuous double gyroid hybrid solar cell. Nano Lett. 9(8),
2807–2812 (2008)
55. Freer, E.M., et al.: Oriented mesoporous organosilicate thin films. Nano Lett. 5(10),
2014–2018 (2005)
56. Tian, Z.R., et al.: Large oriented arrays and continuous films of TiO2-based nanotubes. J. Am.
Chem. Soc. 125(41), 12384–12385 (2003)
57. Mor, G.K., et al.: A review on highly ordered, vertically oriented TiO2 nanotube arrays:
fabrication, material properties, and solar energy applications. Sol. Energy Mater. Sol. Cells
90(14), 2011–2075 (2006)
58. Mor, G.K., et al.: Use of highly-ordered TiO2 nanotube arrays in dye-sensitized solar cells.
Nano Lett. 6(2), 215–218 (2005)
59. Varghese, O.K., Paulose, M., Grimes, C.A.: Long vertically aligned titania nanotubes on
transparent conducting oxide for highly efficient solar cells. Nat. Nanotechnol. 4(9),
592–597 (2009)
60. Hampton, M.J., et al.: The patterning of sub-500 nm inorganic oxide structures. Adv. Mater.
20(14), 2667–2673 (2008)
61. Marzolin, C., et al.: Fabrication of glass microstructures by micro-molding of sol-gel
precursors. Adv. Mater. 10(8), 571–574 (1998)
62. Schaper, C.D., Alan, M.: Polyvinyl alcohol templates for low cost, high resolution, complex
printing. J. Vac. Sci. Technol. B 22(6), 3323–3326 (2004)
172 J.D. Bass and H.-C. Kim

63. Schaper, C.D.: Planarizing surface topography by polymer adhesion to water-soluble


templates with replicated null pattern. Langmuir 20(1), 227–231 (2003)
64. Schaper, C.D.: Patterned transfer of metallic thin film nanostructures by water-soluble polymer
templates. Nano Lett. 3(9), 1305–1309 (2003)
65. Bass, J.D., Schaper, C.D., Rettner, C.T., Arellano, N., Alharbi, F.H., Miller, R.D., Kim, H.-C.:
Transfer molding of nanoscale oxides using water-soluble templates. ACS Nano 5(5),
4065–4072 (2011)
66. Park, O.-H., et al.: High aspect-ratio cylindrical nanopore arrays and their use for templating
titania nanoposts. Adv. Mater. 20(4), 738–742 (2008)
67. Park, O.-H., et al.: Formation and photopatterning of nanoporous titania thin films. Appl. Phys.
Lett. 90(23), 233102–233103 (2007)
68. Karuppuchamy, S., et al.: Cathodic electrodeposition of TiO2 thin films for dye-sensitized
photoelectrochemical applications. Chem. Lett. 30(1), 78–79 (2001)
69. Karuppuchamy, S., et al.: Cathodic electrodeposition of oxide semiconductor thin films and
their application to dye-sensitized solar cells. Solid State Ion. 151(1–4), 19–27 (2002)
70. Natarajan, C., Nogami, G.: Cathodic electrodeposition of nanocrystalline titanium dioxide thin
films. J. Electrochem. Soc. 143(5), 1547–1550 (1996)
71. Miao, Z., et al.: Electrochemically induced sol-gel preparation of single-crystalline TiO2
nanowires. Nano Lett. 2(7), 717–720 (2002)
72. Greene, L.E., et al.: Low-temperature wafer-scale production of ZnO nanowire arrays. Angew.
Chem. Int. Ed. 42(26), 3031–3034 (2003)
73. Ito, D., Jespersen, M.L., Hutchison, J.E.: Selective growth of vertical ZnO nanowire arrays
using chemically anchored gold nanoparticles. ACS Nano 2(10), 2001–2006 (2008)
74. Law, M., et al.: Nanowire dye-sensitized solar cells. Nat. Mater. 4(6), 455–459 (2005)
75. Tian, Z.R., et al.: Biomimetic arrays of oriented helical ZnO nanorods and columns. J. Am.
Chem. Soc. 124(44), 12954–12955 (2002)
76. Vayssieres, L.: Growth of arrayed nanorods and nanowires of ZnO from aqueous solutions.
Adv. Mater. 15(5), 464–466 (2003)
77. Vayssieres, L., et al.: Three-dimensional array of highly oriented crystalline ZnO microtubes.
Chem. Mater. 13(12), 4395–4398 (2001)
78. Hendry, E., et al.: Local field effects on electron transport in nanostructured TiO2 revealed by
terahertz spectroscopy. Nano Lett. 6(4), 755–759 (2006)
79. Kopidakis, N., et al.: Ambipolar diffusion of photocarriers in electrolyte-filled, nanoporous
TiO2. J. Phys. Chem. B 104(16), 3930–3936 (2000)
80. Benkstein, K.D., et al.: Influence of the percolation network geometry on electron transport in
dye-sensitized titanium dioxide solar cells. J. Phys. Chem. B 107(31), 7759–7767 (2003)
81. Kopidakis, N., et al.: Transport-limited recombination of photocarriers in dye-sensitized
nanocrystalline TiO2 solar cells. J. Phys. Chem. B 107(41), 11307–11315 (2003)
82. Galoppini, E., et al.: Fast electron transport in metal organic vapor deposition grown dye-
sensitized ZnO nanorod solar cells. J. Phys. Chem. B 110(33), 16159–16161 (2006)
83. Martinson, A.B.F., McGarrah, J.E., Parpia, M.O.K., Hupp, J.T.: Dynamics of charge transport
and recombination in ZnO nanorod array dye-sensitized solar cells. Phys. Chem. Chem. Phys.
8, 4655–4659 (2006)
84. Hoffmann, M.R., et al.: Environmental applications of semiconductor photocatalysis. Chem.
Rev. 95(1), 69–96 (2002)
85. Hosono, E., et al.: Growth of submicrometer-scale rectangular parallelepiped rutile TiO2 films
in aqueous TiCl3 solutions under hydrothermal conditions. J. Am. Chem. Soc. 126(25),
7790–7791 (2004)
86. Kakiuchi, K., et al.: {1 1 1}-faceting of low-temperature processed rutile TiO2 rods. J. Cryst.
Growth 293(2), 541–545 (2006)
87. Li, Y., et al.: Hydrothermal synthesis and characterization of TiO2 nanorod arrays on glass
substrates. Mater. Res. Bul. 44(6), 1232–1237 (2009)
8 Solution Phase Approach to TiO2 Nanostructures 173

88. Feng, X., et al.: Tantalum-doped titanium dioxide nanowire arrays for dye-sensitized solar
cells with high open-circuit voltage. Angew. Chem. Int. Ed. 48(43), 8095–8098 (2009)
89. Feng, X., et al.: Vertically aligned single crystal TiO2 nanowire arrays grown directly on
transparent conducting oxide coated glass: synthesis details and applications. Nano Lett. 8(11),
3781–3786 (2008)
90. Liu, B., Aydil, E.S.: Growth of oriented single-crystalline rutile TiO2 nanorods on transparent
conducting substrates for dye-sensitized solar cells. J. Am. Chem. Soc. 131(11), 3985–3990
(2009)
Chapter 9
Oxide-Based Photonic Crystals
from Biological Templates

Michael H. Bartl, Jeremy W. Galusha, and Matthew R. Jorgensen

9.1 Introduction

The ability to organize crystalline and amorphous metal oxide compounds into
periodically ordered three-dimensional structures has led to a range of novel functional
materials. Today, periodically ordered metal oxide frameworks with periodicities
spanning several orders of magnitude from the microscale (several Ångstroms) to
the mesoscale (2–50 nm) and to the macroscale are available [1–6]. Regardless of the
framework feature size, it is the combination of inherent metal oxide properties with
three-dimensional structure organization that gives periodically ordered composites
their unique functionalities and has led to a broad field of new applications. While
micro- and mesostructured metal oxides organized by small molecules [7, 8] and
supramolecular self-assembly chemistry [5, 9, 10], respectively, are useful in sorption,
separation, and catalysis [2, 3, 8], as low-k dielectrics [11], and for various optical,
optoelectronic, and energy applications [10, 12–14], macrostructured metal oxides
with periodic feature sizes from a few hundred nanometers to several micrometers [6,
15–17] are prime candidates for photonic crystals [18, 19].
Photonic crystals, originally proposed by John [20] and Yablonovich [21] in 1987,
are an emerging type of optical materials with the potential to manipulate light in
revolutionary new ways. The defining characteristic of photonic crystals is a periodic
variation of the dielectric constant (or refractive index) with periodicities on the order
of the light wavelength of interest. Due to this periodic variation of the dielectric
constant, the behavior of light in photonic crystals is governed by band structure
concepts – similar to how electrons are affected in crystalline atomic crystals [19].
This unique, nonclassical behavior has the potential to lead to fundamentally new

M.H. Bartl (*) • J.W. Galusha • M.R. Jorgensen


Department of Chemistry, University of Utah, Salt Lake City, UT 84112, USA
Department of Physics, University of Utah, Salt Lake City, UT 84112, USA
e-mail: bartl@chem.utah.edu

J. Wu et al. (eds.), Functional Metal Oxide Nanostructures, Springer Series 175


in Materials Science 149, DOI 10.1007/978-1-4419-9931-3_9,
# Springer Science+Business Media, LLC 2012
176 M.H. Bartl et al.

optical principles based on light localization [20, 22, 23], making photonic band
structure materials particularly interesting for next-generation all-optical information
processing and advanced energy technologies [24].
Interestingly, millions of years before we “invented” photonic band structure
materials, nature had already made use of these optical concepts for creating
spectacular structural colors in many insects, birds, and marine animals. In particu-
lar, the strikingly colorful world of insects is in large part the result of optical
interference produced by the interaction of light with precisely ordered, periodic
biopolymeric structures, incorporated into their wings and exoskeletons [25–33].
These biological systems have evolved to create astonishingly complex photonic
architectures – structures that are still far out of our synthetic reach [34–36]. While
biological photonic crystals present exciting structural alternatives to our current
limited photonic engineering capabilities, unfortunately, they do not possess
the high photo and heat stability required for most advanced applications. In
addition, biological photonic structures are composed of electrically insulating,
low-dielectric biopolymers, which strongly limits their device integration. These
problems can be overcome, by converting biopolymeric photonic structures into
inorganic, high-dielectric compounds by biotemplating methods [35–43].
In this chapter we will describe and compare different approaches for creating
three-dimensional photonic crystals. Particular focus will be on oxide-based
photonic crystals with complex lattice structures created using replication routes
from various biological photonic templates.
We will cover three main topics, beginning with a brief introduction of photonic
crystals, emphasizing the nonclassical optical properties of these band structure
materials. We will review the most important top-down and bottom-up fabrication
methods for engineering three-dimensional photonic crystals operating in the
infrared and visible regions of the electromagnetic spectrum and also discuss the
current limitations of these approaches.
Next, we will introduce biological systems that employ photonic band structure
concepts to create a wide variety of structural colors, spanning the entire visible
range. A new structure evaluation method based on sequential focused ion
beam milling and scanning electron microscopy imaging will be discussed. This
method enables a high-resolution three-dimensional reconstruction of photonic
architectures and gives unprecedented insights into the structure-optical properties
relationship of complex photonic crystal lattices.
Then, we will discuss the potential of biological photonic crystal architectures as
biotemplates for conversion into oxide-based high-dielectric replicas. We will briefly
introduce and compare different biotemplating methods, including atomic layer depo-
sition, conformal-evaporated-film-by-rotation, and sol–gel chemistry. The structural
and optical properties of oxide-based photonic crystals obtained from these three
methods will be addressed. Special focus will be directed toward sol–gel biotemplating
methods that have recently made significant progress toward the fabrication of a
photonic crystal with a complete band gap at visible frequencies. We will finish our
chapter with some conclusions and a brief outlook on the potential and promise of
oxide-based three-dimensional photonic crystals for novel optical phenomena at
visible frequencies.
9 Oxide-Based Photonic Crystals from Biological Templates 177

9.2 Engineered Photonic Crystals

9.2.1 Characteristics of Photonic Band Structure Materials

In the following section we will give a brief introduction into the characteristics and
properties of photonic crystals. It is, however, beyond the scope of this chapter to
provide a detailed discussion of the physics of photonic crystals and the foundation
of photonic band structure properties and we refer the interested reader to the
excellent literature available; see, for example [18, 19, 23, 24] and references
therein.
Photonic band structure materials, also known as photonic crystals, are
periodically ordered dielectric composite structures with periodicities [18, 19].
The fascination with these materials stems from their ability to strongly affect the
propagation of light in nonclassical ways, making them interesting candidates for
next-generation optoelectronic applications and all-optical information processing
[24]. The nonclassical properties of photonic crystals are the direct result of their
periodic variation of the dielectric constant. Light with wavelengths comparable to
the variation periodicity of a given photonic crystal is strongly affected by Bragg
diffraction events – similar to how electrons are affected in atomic crystals. As a
consequence, many concepts of solid state physics can be applied to describe the
properties of photonic crystals, leading to direction- and frequency-dependent
photonic band structure diagrams with allowed optical bands and photonic band
gaps (Fig. 9.1).
While photonic crystal can be designed with periodicities in one, two, or three
dimensions, of particular interest are three-dimensional photonic crystals, since
they can have a complete (or omnidirectional) photonic band gap – a frequency

0.7

0.6

0.5 Diamond
FREQUENCY

0.4

0.3

0.2

0.1

0.0
X U L Γ X W K
Wavevector

Fig. 9.1 Calculated photonic band structure diagram for a photonic crystal consisting of dielectric
spheres with a refractive index of 3.6 arranged in a diamond structure. Adapted from [50]
178 M.H. Bartl et al.

Fig. 9.2 Diamond structure models. (a) Dielectric spheres arranged in the diamond lattice.
(b) Dielectric rods connecting nearest-neighbor sites in the diamond lattice. Adapted from [52]

range for which propagation of light is strictly forbidden [19–21]. Photonic band
gaps are the basis for several new optical concepts such as control of spontaneous
emission in bulk materials, slow light-enhanced photocatalysis, low-threshold light
amplification, and quantum information processing [22–24, 44–48].
Motivated by the range of new and exciting optical phenomena predicted
for photonic band gap materials, numerous efforts have been undertaken to fabricate
photonic crystals with band gaps in the microwave, infrared, and visible regions of
the electromagnetic spectrum. In general, the band structure properties of a photonic
crystal are determined by its lattice morphology and the refractive index contrast of
its dielectric building blocks. Along with possessing a high transparency in the
wavelength range of interest, the two different (high- and low-) dielectric materials
building up the photonic crystal lattice should also have strongly differing refractive
indices; for example, air and silicon (for photonic crystals operating in the infrared
region) or air and titanium dioxide (for photonic crystals operating in the visible). In
addition to having a high refractive index contrast, it is also of great importance that
these dielectric building blocks are arranged in an optimal lattice structure. While
every periodically ordered structure results in a direction-dependent energy disper-
sion of photonic states, only a few selected lattices also give rise to the formation of
complete photonic band gaps [45, 49–55]. Among them the diamond crystal structure
is the clear “champion” [52]. Moreover, the diamond crystal structure has been
shown to be rather insensitive to deviations from the ideal lattice configuration,
resulting in a variety of diamond-based (or diamond-like) morphologies that allow
a complete band gap to open for refractive index ratios as low as 2.1 (Fig. 9.2)
[50, 52, 53].
To date, tremendous progress in photonic structure engineering has been made
in the microwave and infrared regimes. Using top-down microfabrication and
bottom-up colloidal self-assembly techniques, various three-dimensional photonic
crystal structures have been synthesized with complete band gaps at infrared
frequencies [56–64]. In contrast to the successes in the infrared regime, complete
photonic band gaps at visible frequencies have proven elusive due mainly to
difficulties in creating efficient three-dimensional lattices with feature sizes in
9 Oxide-Based Photonic Crystals from Biological Templates 179

the hundred-nanometer range. In the following section we will briefly review


fabrication strategies for photonic crystals operating in the infrared and visible
regimes. We will discuss the advantages and limitations of different top-down and
bottom-up fabrication techniques focusing on attainable feature sizes, crystal
lattices, and high-dielectric compounds.

9.2.2 Photonic Crystals Operating in the Infrared Region

The fabrication of photonic band gap crystals operating at infrared frequencies has
benefited tremendously from powerful microprocessing techniques that have been
optimized in the semiconductor industry during the last 50 years. These techniques
can generally be classified into direct and indirect methods. In the former, a desired
photonic crystal structure is formed directly out of a high-dielectric semiconductor
compound. For example, Lin and coworkers used a comprehensive multilevel
stacking process consisting of a repeated deposition, lithographic patterning, and
etching to successfully fabricate dielectric woodpile (a diamond-based lattice struc-
ture) photonic crystals with band gaps in the infrared regime [58]. Subsequently,
Noda et al. developed a wafer-fusion-based method to create woodpile structures
made out of GaAs with a complete band gap at near infrared wavelengths [59],
whereas Johnson, Joannopoulos and coworkers designed and fabricated a nine-layer
photonic crystal with a wide (up to 25% gap-to-mid-gap ratio) band gap out of
silicon by sequential layer-by-layer scanning-electron-beam lithography [61].
A common disadvantage of these direct methods is that fabrication of high-quality
three-dimensional photonic crystals is very time consuming, expensive, and is gener-
ally limited to only a few layers. Indirect methods, on the other hand, use a template
structure created out of inexpensive polymers. This structure serves as a sacrificial
mold for templating high-index compounds such as silicon or germanium. Success-
fully applied methods to create such polymeric photonic crystal template structures,
including the highly efficient diamond-based woodpile lattices, are multibeam
holography, multiphoton lithography, and direct laser writing methods [64–66].
An interesting alternative to these light-patterning routes is the direct ink writing
method originally developed by Lewis and coworkers [67, 68]. In this technique, a
cylindrical filament approximately 1 mm in diameter is formed by the deposition of
a fluidic polyelectrolyte/water ink into an alcohol-rich reservoir. Braun, Lewis and
coworkers demonstrated that this filament can then be patterned in a layer-by-layer
sequence to build a woodpile structure with photonic crystal feature sizes in the
near infrared region [57, 69].
All of these polymeric templates can be converted into high-dielectric photonic
crystals made out of silicon or germanium. Since the polymeric templates would
not withstand the high deposition temperatures required for typical semiconductor
deposition techniques such as chemical vapor deposition, they are first protected by
a metal oxide (silica or alumina) coating formed by atomic layer deposition. For
example, Ozin and coworkers showed that depending on the amount of metal oxide
180 M.H. Bartl et al.

Fig. 9.3 Scanning electron microscopy images of an inverse germanium woodpile structure
fabricated by direct ink writing and a combination of atomic layer deposition and chemical
vapor deposition. Adapted from [57]

deposition (complete backfilling or deposition of a thin coating) it is possible to


create high-dielectric photonic crystals in the form of a positive replica or an
inverse of the original template structure [64, 70]. While a silicon double inversion
procedure produced a woodpile photonic crystal with a complete (up to 9% wide)
band gap in the infrared region [70], Hermatschweiler et al. showed that silicon
inverse woodpile photonic crystals with a more than 14% wide complete band gap
centered at a wavelength of around 2.5 mm can be fabricated by a silicon single
inversion method [64]. Braun, Lewis and coworkers used a similar – although
independently developed – technique to convert direct ink-writing-created wood-
pile templates into germanium photonic crystals with wide (up to 25%) complete
band gaps centered at a wavelength of around 6 mm (Fig. 9.3) [57].
An interesting – fast, simple, and low-cost – alternative to these rather labor-
intensive routes is colloidal self-assembly [71–73]. In this bottom-up photonic crystal
fabrication technique, monodisperse nano and microspheres are deposited onto
planar substrates by self- or directed assembly in close-packed face-centered cubic
or hexagonally close-packed colloidal crystals (also called artificial opals, since these
colloidal crystals closely resemble the microstructure of natural opal gemstones).
9 Oxide-Based Photonic Crystals from Biological Templates 181

Similar to the indirect methods described above, these colloidal crystals are then used
as templates and are infiltrated with an infrared-transparent high-dielectric compo-
nent. After selective removal of the opal template an inverse opal photonic crystal
(a close-packed face-centered-cubic lattice of air spheres in a high dielectric material)
is obtained [45]. While inverse opal photonic crystals are less effective (i.e., less
efficient in affecting and controlling the propagation of light) than diamond-based
lattices, it was shown that the formation of a complete photonic band gap is possible
provided the high-dielectric material has a refractive index of 3 or higher versus air as
the low-dielectric component [49]. Using polycrystalline silicon as the high-dielectric
component (with a refractive index of 3.2–3.4), John and coworkers [56] and Norris
and coworkers [62] successfully fabricated inverse opal photonic crystals with a
complete band gap in the near infrared region.

9.2.3 Photonic Crystals Operating at Visible Frequencies

Compared to the enormous progress achieved in fabricating photonic crystals


operating in the infrared region, photonic structure engineering in the visible region
is far less advanced – due mainly to the difficulties in shaping visible light-transpar-
ent, high-dielectric materials into efficient morphologies with periodicities at visible
wavelengths. Unlike infrared photonic crystals with complete band gaps of up to
20–30% gap-to-mid-gap ratios [57–63], enabled by infrared-transparent materials
with refractive indices of 3.2 and higher, the lack of visible light-transparent
dielectrics with comparable refractive indices embosses an enormous challenge for
achieving complete band gaps at visible wavelengths. The best compounds for
photonic crystals in the visible region are cadmium chalcogenides and oxide
semiconductors such as zinc oxide and titanium dioxide (titania) with refractive
indices in the range of 2.0–2.6. Calculations show that the lowered refractive index
ratio as compared to infrared compounds not only reduces the width of potential
complete band gaps to below 10%, but also limits photonic crystal morphologies with
potentially complete band gaps at visible frequencies to pyrochlore and diamond-
based crystal lattices [50–52].
Unfortunately, successful synthesis of such lattices with feature sizes at visible
length scales is extremely challenging. On the one hand, typical microfabrication
methods used to create diamond-based photonic crystal structures in the infrared
region rely on lithographic/holographic or direct-writing methods and are therefore
very difficult to successfully implement at the small length scales required for
three-dimensional structures with band gaps in the visible region. Subramania
and coworkers showed that this obstacle could be overcome by electron beam
direct writing of a titania woodpile structure with lattice parameters in the visible
region [74, 75]. However, creating large-scale structures with this technique has
proven to be extremely challenging, limiting the fabricated photonic crystals to nine
alternating layers. On the other hand, bottom-up supramolecular self-assembly of
polymeric building blocks into diamond-based lattices is limited to feature sizes
182 M.H. Bartl et al.

Fig. 9.4 Top (a–c) and side view (d) scanning electron microscopy images of planar open-surface
titania inverse opal photonic crystals. Images are taken along the [111] direction (a, b) and the
[100] direction (c). Adapted from [15]

less than 150 nm due to the increasingly slowed assembly kinetics of the required
ultra-high molecular weight monomers, restricting this technique to the fabrication
of photonic crystals operating in the ultraviolet regime [76].
A fabrication technique that readily bridges the feature-size gap between
microfabrication and supramolecular assembly is the colloidal self-assembly of
submicrometer spheres [71–73]. Indeed, colloidal crystals can be formed with
periodicities spanning the entire visible range and can be converted into inverse
opals by infiltration with visible light-transparent dielectric compounds [6, 15, 17,
77–79]. This approach has been used to create inverse opal photonic crystals
composed of high-dielectric metal oxide compounds using various infiltration
techniques. For example, successful procedures were reported by Stein and
coworkers [6, 80] and Vos and coworkers [17, 81], who used alkoxide condensation
in air to create titania, zirconia, and alumina inverse opals and Pine and coworkers
[77, 79], who showed that titania and silica inverse opals can also be prepared by
infiltration of opals with ultrafine colloidal particles of the oxides. Recently, Bartl
and coworkers developed a titania sol–gel precursor and combined it with a liftoff/
turnover infiltration/processing technique to fabricate planar titania inverse opals
with an open surface and defined thickness (Fig. 9.4) [15].
9 Oxide-Based Photonic Crystals from Biological Templates 183

Another powerful method to infiltrate polymeric templates is low-temperature


atomic layer deposition [82]. For example, King et al. used this method to fabricate
titanium dioxide inverse structures from various self-assembled and holographi-
cally prepared templates with highly controlled filling fractions and excellent
quality [83–85].
Despite this intense research in developing high-quality inverse opals, these
materials fail to produce complete photonic band gaps in the visible region. The
reason lies in the crystal structure of opals, which lacks the desired diamond-based
symmetry. As mentioned above, spherical building blocks have a natural tendency
to form densely packed lattices. Such opal-based crystals, however, require refrac-
tive index ratios greater than 3 to form complete photonic band gaps [49]. Unfortu-
nately, this is far beyond the reach of visible light-transparent compounds and
consequently, even the best inverse opal photonic crystals have only directional
(incomplete) band gaps at visible frequencies.
To conclude, despite promising proposals for directed assembly routes [51, 54]
and proof-of-principle electron beam direct-writing fabrication of titania woodpile
structures [74, 75] and sphere-by-sphere assembly using a micromanipulator [86],
large-scale synthetic photonic crystals with complete band gaps at visible
frequencies are yet to be achieved.

9.3 Natural Photonic Crystals

9.3.1 Structural Colors in Biology

In contrast to our current limitations in photonic structure engineering at visible length


scales, biological systems have developed a wealth of photonic structures composed of
biopolymeric components such as chitin and keratin. These structures efficiently
interact with visible light for the production of vibrant colors in wings, feathers,
hairs, or as part of insect exoskeleton [25–35, 87–89]. Biological structural colors
have been developed for a large variety of reasons ranging from camouflaging to
frightening predators and attracting mates. This requires an array of optical effects:
From strongly iridescent to near angle-independent coloration, from shimmering
bright to matte hues, from brilliantly sparkling to pastel colors and various
combinations thereof. Moreover, biological photonic structures were optimized to
operate under various illumination conditions; while many species create their optical
effects in bright sunlight, others operate under highly scattering conditions produced
by water droplets and wet leaves in rain forests. Others have to function under dim
illumination at the forest floor or within dense vegetation. These different applications,
optical effects, and environmental conditions have led to a diversity of photonic
structures in biology that is virtually limitless.
For example, simple one-dimensional multilayer structures are found in many
species of butterflies and beetles, which employ alternating biopolymeric layers
184 M.H. Bartl et al.

Fig. 9.5 (a) Photograph of the weevil Lamprocyphus augustus. (b) Optical micrograph of
individual exoskeleton scales of L. augustus under white light illumination. (c) Cross-sectional
scanning electron microscopy image of a single scale. (d) Detailed cross-sectional scanning
electron microscopy image of a region of a scale. Adapted from [34]

with slightly different refractive indices [90–92]. More sophisticated examples are
found in marine animals such sea urchins [93] and the sea mouse [94], whose
skeleton and hairs, respectively, have an internal two-dimensional periodic lattice.
Two-dimensional photonic structures are also found in the feathers of several
species of birds [87, 88, 95]. In addition to these simpler one- and two-dimensional
structures, many species of butterflies (Lepidoptera) and beetles (Coleoptera)
obtain their striking coloration from various three-dimensional architectures with
lattice constants spanning the entire visible regime (see, for example, Fig. 9.5).
Examples of such cuticular (biopolymeric) photonic crystal architectures range
from quasiperiodic lattices to chiral, honeycomb, and non-close-packed ball-stick
structures as well as various cubic lattice symmetries [28, 29, 31, 34, 35, 89,
96–100].
It is interesting that most of the photonic structures found in butterflies and beetles
display symmetry-breaking strategies to produce tailored optical effects. The role of
symmetry in photonic crystals is twofold, consisting of a delicate interplay between
the overall symmetry of the system given by the crystal lattice structure and the
building blocks occupying the individual lattice sites. While the symmetry of the
9 Oxide-Based Photonic Crystals from Biological Templates 185

overall structure needs to be such that the resulting Brillouin zone is as spherical as
possible, the symmetry of the building blocks need not be high. In fact, lowering the
symmetry of the building blocks (for example, from spherical to cylindrical or even
lower) can strongly enhance the photonic properties in these materials [101]. Using
designed building block shapes provides a means to finetune photonic properties and
engineer photonic crystal structures with specific optical qualities.

9.3.2 Structure Evaluation Methods

The concept of symmetry breaking/lowering is a widely used strategy in biological


photonic crystals, making it difficult to evaluate their exact three-dimensional lattice
architecture [34, 89, 91, 102–104]. Typical top view and cross-sectional scanning
electron microscopy imaging is a powerful technique to investigate structural features
at the relevant tens to hundreds of nanometer length scale. However, such randomly
taken two-dimensional views lack sufficient morphological details to enable a com-
plete understanding of the three-dimensional lattice architecture. This difficulty is
particularly exacerbated for biological photonic structures in which the lattice build-
ing blocks often deviate strongly from simple spherical shapes. The situation is
further complicated by the fact that photonic structures in most biological systems
are polycrystalline, consisting of several micrometer-sized arrays of differently ori-
ented single-crystalline domains [29, 34]. These varying hierarchical orderings and
myriad structural fine features make it exceedingly difficult to find a single method for
solving the lattice geometries of all the different biological photonic architectures.
Consequently, a range of structure investigation techniques have been developed and
applied to biological structures. In general, these techniques can be divided into
scattering methods (various scatterometry and spectrophotometry techniques) and
structural methods (electron microscopy imaging-based procedures). As an example,
we will briefly discuss a novel electron microscopy-based structure evaluation
technique in the following section and for structural methods we refer the interested
reader to an excellent review by Vukusic and Stavenga and references therein [104].
To solve hierarchically ordered three-dimensional photonic structures hidden
within exoskeleton scales of colored beetles, Galusha et al. developed a high-
resolution structure evaluation method that combines scanning electron microscopy
with focused ion beam milling and subsequent volume rendering/computer
modeling [29, 34, 89]. This technique allows a three-dimensional reconstruction
of biological photonic structures with unprecedented resolution (~30 nm). In short,
serial sectioning was performed by consecutively milling away ~30 nm sections of
the structure using a focused ion beam. After each milling step, the freshly exposed
cross-sectional two-dimensional view of the structure was imaged by scanning
electron microscopy. This resulted in a three-dimensional data set consisting of a
series or “stack” of two-dimensional images, each with a “thickness” of ~30 nm –
significantly less than the lattice constants of the periodic structures
(~300–500 nm). This stack of images was smoothed and aligned using a recursive
186 M.H. Bartl et al.

Fig. 9.6 (a) Scanning electron microscopy image of a scale’s top surface exposed by focused ion
beam milling. (b–d) Scanning electron microscopy images of individual single-crystalline
domains indicated in (a) with corresponding calculated dielectric functions (black-framed insets).
Adapted from [34]

algorithm in ImageJ [105, 106]. The preprocessed stack of consecutive scanning


electron microscopy images was then imported into the SCIRun visualization
package for volume rendering [107]. High-resolution structural information was
obtained by comparing and analyzing a large number of oblique “cuts” through the
virtual reconstruction of the three-dimensional biological photonic structure. For
example, Fig. 9.6 shows as comparison of the calculated/reconstructed structures of
different crystal faces with the crystal faces of the actual three-dimensional
photonic crystal lattice of the weevil Lamprocyphus augustus (details are given in
the following section).

9.3.3 Examples of Biological Photonic Structures

In this section we will introduce and discuss specific biopolymeric photonic crystal
structures found in the colored wings and exoskeleton scales of butterflies and
beetles. From the seemingly infinite pool of biological photonic architectures, the
ones described below were chosen because of their unique – and in most cases
synthetically out-of-reach – lattice structures. These particular biopolymeric
lattices have therefore been the templates of choice in bioinspired photonic crystal
fabrication, as will be discussed later in this chapter.
9 Oxide-Based Photonic Crystals from Biological Templates 187

Fig. 9.7 Scanning electron microscopy view of a fractured photonic structure of a Morpho scale
showing a side view of three ridges with their lamellae (l) and associated microribs (mr). Also
shown are parts of several crossribs (cr; here fractured) that join the ridges as well as the bottom
layer (bl) of the scale and several pillars (p) that connect the bottom scale layer with the photonic
structure. Adapted from [111]

Among the large group of structural color effects in butterfly wing scales, those
found in the Nymphalidae, Lycaenidae, and Papilionidae families are particularly
interesting due to their diversity in structure/coloration and underlying hierarchical
architectures, consisting of ridges, lamellae, and ribs [31, 33, 103, 108–110]. While
these main structural features are found in most colored wing scales, it is their
spatial arrangement, relative size, and orientation that results in the wide variation
of optical effects produced by butterflies. The famous tropical Morpho butterflies,
for example, obtain their distinct blue coloration from an elaborate ridge/lamellae/
microrib structure acting as a combination of multilayer reflector and diffraction
grating [29, 33, 108]. In detail, each ridge is built up of alternating stacked lamellar
features with a precise periodic spacing. These lamellar features are connected by
microribs and act as a multilayer reflector for a particular band of frequencies. In
addition, the spacing between ridges and their relative orientation can be designed
to function as a diffraction grating. As an example, Fig. 9.7 shows different electron
microscopy views of the photonic structure of the butterfly Morpho sulkowskyi.
Potyrailo et al. investigated this butterfly and discovered an interesting phenome-
non; namely, the hierarchical architecture of the scales of M. sulkowskyi also
displays a highly selective optical response to various alcoholic vapors and thus
can be used as a photonic gas sensor [111].
In addition to various ridge-based reflector and grating structures, the wing
scales of the Lycaenidae and Papilionidae butterfly families often possess photonic
188 M.H. Bartl et al.

Fig. 9.8 Transmission electron microscopy images of the photonic wing scale structure from the
butterflies P. sesostris (a) and T. imperialis (b). Adapted from [109]

crystal architectures in which the biopolymeric matrix underneath the ridges can be
sculpted into complicated three-dimensional periodic lattices. For example,
Michielsen and Stavenga found that the Green Hairstreak butterfly (Callophrys
rubi) obtains its distinctive wing coloration from scales with internal gyroid-
structured photonic crystals [100]. Gyroid-based lattices are very interesting
photonic structures since they are known to have efficient/robust band structure
properties. Other interesting examples of combined ridge-photonic crystal scale
structures are found, for example, in the wings of Parides sesostris and Teinopalpus
imperialis in which the ridges and the three-dimensional photonic lattice are
connected by a matrix composed of a series of vertical columns organized in so-called
honeycomb arrays (Fig. 9.8) [109]. Additionally, in most cases, the three-dimensional
photonic crystal domains are composites of differently oriented subdomains with the
same lattice structure [29]. This structural feature is often also found in another type of
structural color in biology, namely, in exoskeleton scales of colored beetles [34].
Bartl and coworkers studied the origin of structural colors in various colored
beetles of the longhorn (Cerambycidae) and weevil (Curculionidae) families
[34, 35, 89, 112]. This group discovered several interesting biopolymeric structures
with three-dimensional lattice geometries and multidomain arrangements. These
structures include highly ordered non-close-packed sphere lattices and quasiperi-
odic lattices of spheres found in the longhorn beetles Glenea celia and Anoplophora
elegans, respectively, as well as elaborate interpenetrating three-dimensional
architectures such as the highly desired diamond-based structures, found in
several weevils (L. augustus, Eupholus schoenherri, Eudiagogus pulcher, and
Pachyrrhynchus moniliferus).
In particular, the discovered diamond-based architectures are of great impor-
tance since they resemble some of the most efficient photonic crystal lattice
structures. Galusha et al. performed a detailed investigation of the Brazilian weevil
L. augustus and discovered that its near-angle-independent greenish hue is the
result of exoskeleton scales with an interior diamond-based photonic crystal
9 Oxide-Based Photonic Crystals from Biological Templates 189

Fig. 9.9 (a–c) Scanning electron microscopy images of the photonic structure of the weevils
E. schoenherri, P. moniliferus, and E. pulcher, respectively. (d) Calculated dielectric function,
showing three orthogonal planes of the reconstructed diamond-based crystal structure with a 50%
volume fraction (air: dark; dielectric: light). Adapted from [35]

structure (see also Fig. 9.5) [34]. Remarkably, a high-resolution reconstruction


revealed that the beetle’s lattice structure is near-identical to a theoretically derived
and property-optimized photonic crystal structure by Johnson and Joannopoluos –
the only major difference being a higher dielectric filling fraction in the beetle
structures [113]. In more detail, the diamond-based photonic structure of
L. augustus consists of an ABC stacking of layers of hexagonally ordered air
cylinders in a surrounding cuticular matrix (Fig. 9.9). The air cylinders have an
average radius and height of 0.20 and 0.77, respectively, in units of the lattice
constant, which was found to be 450 nm (with an inherent 10–20% variation of
structural dimensions within single scales and between scales taken from different
parts of the beetles). The filling fraction of the cuticular matrix was found to be
between 50 and 60%. Moreover, the reconstruction over larger areas revealed each
scale is a hierarchically organized photonic structure composed of an array of
diamond-based single-crystalline domains selectively oriented with their [100],
[110], or [210] crystal axes normal or slightly off-normal to the scale top surface.
A comparison with photonic band structure calculations and multi directional
optical reflectance spectroscopy investigations showed that it is this multi domain
orientation of the diamond-based crystal lattice that gives L. augustus its angle-
independent green coloration.
Subsequently, the same group also discovered diamond-based photonic crystal
structures in other weevils. In fact, they found that pixilated arrays of diamond-based
architectures seem to be the dominant photonic lattices among colored weevils.
190 M.H. Bartl et al.

For example, the weevils E. schoenherri, P. moniliferus, and E. pulcher all possess
diamond-based photonic crystal lattices (Fig. 9.9) [35, 89]. While these crystal
structures are isomorphous with the structure originally found in L. augustus, they
display differences in both lattice constants and filling fraction. Such variations in
lattice constants as well as filling fraction strongly influence the resulting photonic
properties. While biology uses these strategies to create different structural colors and
fine-tune optical effects, they also provide enormous possibilities for photonic
materials research. For the first time, some of the most sought-after photonic crystal
structures operating at visible frequencies are available – lattice symmetries that are
not yet available via synthetic structure engineering methods. In the following
sections we will discuss the possibilities of using these biological structures as unique
templates to create novel oxide-based optical materials.

9.4 Biotemplated Photonic Crystals

9.4.1 General Considerations

The large variation in crystal structure symmetry, lattice constant, and dielectric
filling fraction of biological photonic structures dramatically extends the currently
available “synthetic” photonic crystal structures fabricated through micropro-
cessing or bottom-up assembly techniques. Unfortunately, the direct use of
biological photonic crystals for advanced optical/optoelectronic applications is
strongly limited since most of these biological structures are composed of
biopolymeric compounds that lack several properties critical to emerging optical
applications, such as a high refractive index and structural robustness, as well as heat
and photostability, necessitating conversion into more stable inorganic materials.
From a materials viewpoint, biological structures are therefore similar to syn-
thetic polymeric photonic crystal templates created by direct ink writing, laser
writing, holography, or colloidal self-assembly discussed previously. In these
examples, the fabricated polymeric structures have to be converted into inorganic
positive or inverse structures by infiltration with transparent high-dielectric
compounds. Consequently, similar infiltration methods can be used to convert
biological photonic frameworks into high-dielectric replicas, such as atomic layer
deposition, low-temperature evaporation techniques, and sol–gel chemistry routes
[36]. Compared to synthetic template structures, however, there are several new
considerations that have to be taken into account for successfully converting
biological structures into oxide-based high-dielectric replicas. These considerations
include accessibility of the photonic structure for infiltration, avoidance of lattice
shrinkage during processing, and formation of a dense high-dielectric framework
while preserving structural integrity. In the following, we will discuss these general
factors in more detail, before introducing several specific examples of successfully
applied biotemplating methods in section 9.4.2.
9 Oxide-Based Photonic Crystals from Biological Templates 191

In contrast to the free-standing, open, and easily accessible lattice framework


of synthetic photonic crystal templates such as polymeric woodpile and opal
structures, biological photonic structures are integrated into larger body parts
(feathers, wings, hair, exoskeleton). Consequently, biological photonic structures
are often buried, hidden, or embedded within a structureless matrix. This is especially
true for photonic structures of most butterflies and beetles, which are contained in
wings and exoskeleton scales and are thus partially or completely surrounded by an
impermeable biopolymeric shell (see also Fig. 9.5) [28, 35, 108, 110]. In addition,
most biopolymeric structures are covered by a hydrophobic, wax-like film that can
cause problems during infiltration. Successful infiltration/replication of biological
photonic crystals thus often requires preprocessing steps such as treatment with
organic solvents or acids to remove the wax-like layer and cutting or microtoming
to provide access to the encapsulated lattice structures for high-dielectric filling
compounds. The template-infiltration process is generally followed by densification
and/or crystallization of the inorganic replica framework. This step is of great
importance since achieving a dense, highly crystalline photonic crystal framework
with a high refractive index is of the utmost importance in photonic structure
engineering via templating of both synthetic and biological polymeric structures.

9.4.2 Biotemplating Techniques

9.4.2.1 Deposition and Evaporation Methods

Low-temperature atomic layer deposition is an excellent method for creating oxide-


based inorganic replica of biological photonic structures [40, 41, 82, 114, 115].
It combines a noncorrosive reaction environment and mild pH conditions with rela-
tively low deposition temperatures of around 100–200 C. Furthermore, since the
infiltrated oxide compound is formed by a layer-by-layer atomic deposition process,
the degree of infiltration can easily be tuned by controlling deposition cycles. The
precursors used are in general gaseous compounds and therefore readily infiltrate even
complex three-dimensional frameworks as long as the internal structure is fully acces-
sible. Since the majority of photonic structures found in the wings of butterflies are
open frameworks and require no or little preinfiltration cutting, they are the templates
of choice for most successful atomic layer deposition-based bioreplication attempts.
For example, Wang and coworkers created aluminum oxide (alumina, Al2O3)
replicas of the photonic structure of wing scales from the butterfly M. peleides by
atomic layer deposition [40]. Using a low-temperature atomic layer deposition
process at 100 C and trimethyl aluminum (Al(CH3)3) and water as precursor
sources, the biopolymeric photonic structure of M. peleides was coated with a
structurally stable layer of amorphous alumina. The thickness of the alumina
coating was gradually increased by about 10 nm steps until a final thickness of
40 nm was achieved. Characteristic of photonic crystal structures, the infiltration
process can be monitored and precisely controlled by analyzing the color of the
192 M.H. Bartl et al.

Fig. 9.10 (a) An optical microscope image of the alumina coated Morpho peleides butterfly wing
scales, of which the color changed from original blue to pink. (b) Scanning electron microscopy
image of alumina replicas of butterfly wing scales. (c) Energy dispersive X-ray spectrum of
alumina replicas shown in (b). (d) Higher magnification scanning electron microscopy image of
alumina replicas of butterfly wing scales and two broken rib tips (e). Adapted from [40]

reflected light. As shown in Fig. 9.10a, a red shift (from blue to pink) of the reflected
light was observed due to a change in periodicity and effective refractive index of
the composite as the thickness of the coating increased.
When a desired layer thickness was obtained, the biopolymeric/alumina com-
posite was heated to 800 C in air. Under these conditions, the butterfly template
completely decomposes and the amorphous alumina coating is transformed into a
polycrystalline framework. The resulting structure is a shell-like copy of the
original M. peleides photonic lattice made out of transparent, polycrystalline
alumina. Scanning electron microscopy images are given in Fig. 9.10 and show
that even nanoscale structural features of the original biotemplate were perfectly
9 Oxide-Based Photonic Crystals from Biological Templates 193

preserved by this process. Furthermore, optical reflectance spectroscopy studies


revealed alumina replica features very similar to that of the original butterfly scales
in terms of reflection peak wavelength position and shape.
Gaillot et al. developed a similar atomic layer deposition route to fabricate
biotemplated organic–inorganic composite photonic crystal structures [41]. In
their approach, the photonic scales covering the wings of the green swallowtail
butterfly P. blumei were used as biological templates. Two different templating
methods via titania low-temperature atomic layer deposition were employed. While
in the first method entire scales were surrounded by a uniformly thin (50 nm) layer
of amorphous titania, in the second method both the scale exterior and the interior
photonic structure were covered with a titania layer. The intrascale deposition of
titania was enabled by diffusion and subsequent deposition of the gaseous titania
precursors through surface cracks created by razor blades or sharp tips. Gaillot et al.
also conducted detailed structural and optical studies on both organic–inorganic
replica types [41]. Experimental results were compared to theoretical modeling and
band structure calculations and provided valuable insights regarding the ability to
tune the optical properties of these photonic crystals through slight variations of the
deposited high-dielectric compound. In addition, analyses of the properties of these
different types of oxide-replica butterfly wings indirectly provided new insights
into their structural complexity; with various intersecting nanochannels and
connected chambers in addition to the photonic crystal structure.
Another interesting and elegant biotemplating method is the so-called confor-
mal-evaporated-film-by-rotation technique developed by Lakhtakia and coworkers
[38, 39, 116]. This method, which evolved from the oblique angle deposition
technique, combines thermal evaporation with simultaneous substrate tilting and
rotation in a low-pressure chamber. Since the reaction chamber pressure is in the
micro-Torr regime, the evaporation/deposition process can be performed at low
temperatures and under a noncorrosive environment, providing excellent conditions
for replicating sensitive biological structures. Additionally, the high-speed rotation
of the biological templates during the evaporation process facilitates the formation
of a homogeneous and dense inorganic film, even on highly curved and structured
surfaces.
The first successful application of the conformal-evaporated-film-by-rotation
technique on biological samples was the high-fidelity replication of various body
parts of a fruit fly, including the eyes, head, and wings [38]. For example, the
replicated eye structures displayed the same long-range features (micrometer-scale)
and fine features (nanometer-scale) as found in the original fly eye. The same group
later extended this biotemplating technique to replicate photonic structures found in
the wings of butterflies [39, 116]. For example, Fig. 9.11 shows a high-resolution
scanning electron microscopy image of replicated micrometer and nanometer
features of the photonic framework found in the wing scales of the butterfly Battus
philenor. While these initial replicas were all composed of chalcogenide glasses with
a nominal composition Ge28Sb12Se60, the conformal-evaporated-film-by-rotation
technique can also be used to create oxide-based compounds [117].
194 M.H. Bartl et al.

Fig. 9.11 Scanning electron microscopy image of a replica of the scales of B. philenor created by
the conformal-evaporated-film-by-rotation technique. Adapted from [39]

9.4.2.2 Sol–Gel Chemistry Methods

Sol–gel chemistry is an attractive alternative to deposition and evaporation methods


for its simplicity enabled by flexible processing parameters [118]. However, pro-
ducing oxide-based bioreplicas of similar quality to those obtained from evapora-
tion and atomic layer deposition in terms of uniformity and degree of template
infiltration, compositional control, and structural integrity requires precise tuning
and optimization of processing parameters along their many degrees of freedom.
Among these, the most important are the precursor components and solvents, the
sol composition and concentration, and the postinfiltration conditions such as sol
drying time, humidity during the gelation process, and temperature and time of the
heat treatment used to induce complete solidification and (in some cases) crystalli-
zation of the oxide-based replica framework.
Similar to evaporation and deposition approaches, the majority of sol–gel
biotemplation attempts have focused on replicating the photonic structures found in
the wing scales of various butterflies due to their open framework. For example,
Zhang and coworkers created zinc oxide and titanium dioxide wing scale replicas
from butterflies Papilio paris and Thaumantis diores [42, 43]. A simple sol–gel route
was developed to infiltrate the biological templates with ethanol solutions of the
respective metal salts followed by heat treatment at 500 C in air to induce crystalliza-
tion of the oxide framework and remove the biopolymeric structure. Despite structural
shrinkage during the heat-based framework densification–crystallization treatment,
9 Oxide-Based Photonic Crystals from Biological Templates 195

the replicated materials displayed optical band structure features in the blue-green
region. Zhang and coworkers showed that these interesting optical properties can be
used to create new materials for applications in solar cells and light emission. For
example, titanium dioxide photonic architectures templated from butterfly wings can
be used as photoanodes with enhanced light harvesting efficiencies under visible light
illumination [42]. On the other hand, the patchy blue and black colored wing scales of
P. paris replicated into zinc oxide have interesting roomtemperature cathodolumi-
nescence properties [43]. While replicas from both the black and the blue patches of
the scales showed sharp near-band-edge cathodoluminescence in the UV region, the
emission from the former also displayed strong green emission, which the authors
attributed to the internal sponge-like “beehive photonic nanostructure” found in the
black patch area of the scales.
Zhu et al. reported an interesting bioreplication approach of butterfly wings by
combining sol–gel templating with sonochemistry [119]. In their method, the internal
biopolymeric nanostructure of blue-colored wings from Morpho butterflies was
impregnated with an ethanol–water–based precursor solution followed by high-
intensity ultrasonication for several hours at room temperature. Using this approach
the authors were able to replicate the photonic wing structure into a variety of oxide
ceramics including titanium dioxide, tin dioxide and silicon dioxide. After sonication,
the composites were heat-treated to remove the biopolymeric template and, in the
case of titanium dioxide and tin dioxide, induce the formation of a polycrystalline
framework. All of the replicated samples possessed highly preserved photonic
nanostructures stemming from the original butterfly wings and displayed optical
reflectance features in the visible region. In addition, it was found that the crystalline
tin dioxide replica can be used as a gas sensor, displaying high sensitivity for ethanol
vapor with a fast response time (8 s) and a short recovery time (15 s).
In addition to replicating the open-scale structures of butterfly wings, sol–gel
chemistry can also be used to template photonic lattices enclosed in the
biopolymeric shells typically found in beetles [35, 37, 112]. Interestingly, small
openings of the scale, which naturally form during scale removal from the insect’s
exoskeleton or can be created by cutting with a razor blade, are sufficient to
infiltrate the entire internal photonic crystal framework. The key here lies in the
liquid nature of the sol–gel precursor, which allows infiltration of the internal
photonic lattice framework through small openings by capillary forces. Galusha
et al. used hybrid organic/inorganic silica sol–gel chemistry combined with acid-
based template removal approach to successfully create glass-like replicas of
different photonic crystal structures from a variety of colored beetles [35]. The
rationale for using a hybrid (SBA-type) silica sol–gel precursor (originally devel-
oped by Stucky and coworkers [120]) as the infiltration material lies in the superior
templating properties of this organic/inorganic compound. The hybrid nature of this
organic/inorganic precursor results in enhanced structural stability upon sol drying/
solidification and template removal as compared to pure silica. While the latter
shows significant crack formation and structural long-range order degradation in
the replicated structure, the use of organic/inorganic silica sol not only preserves to
196 M.H. Bartl et al.

Fig. 9.12 (a) Scanning


electron microscopy image of
a sol–gel-derived hybrid
silica inverse structure
templated from a scale of G.
celia. (b) Magnified look of
the top surface of the inverse
structure. Adapted from [35]

a large extent the macroscopic morphology of the original scale template, but also
the structural fine features of the photonic structure (compare also Fig. 9.12).
Another crucial factor in converting biopolymeric templates into oxide-based
replicas is template removal after the infiltration/solidification event. A common
method is heat treatment under oxidative conditions at temperatures between 300
and 500 C. This process is effective in complete removal of the organic template;
however, it can also cause significant structural damage and shrinkage with a
concomitant reduction of the lattice periodicity of up to 30%. Due to the pro-
nounced structure–property relationship in photonic crystals, shrinkage is
accompanied by blue shifting of the optical features of the replicated photonic
structures, which is problematic for applications that require band gaps at particular
frequency ranges in the visible region. In synthetic structure templating, this
shrinkage can be circumvented simply by starting with templates with larger lattice
constants (for example, to fabricate a titania inverse opal crystal with a lattice
constant of 500 nm, a polymeric opal template with a lattice constant of 700 nm is
used). Unfortunately, this strategy cannot be applied for biological photonic
9 Oxide-Based Photonic Crystals from Biological Templates 197

templates, since they have fixed lattice constants and are designed and optimized to
create defined structural colors by interaction with visible light. Lattice shrinkage in
the range of 30% of these structures during the infiltration/replication process
would therefore lead to photonic crystals with band gap features shifted out of the
desired visible range into the deep blue or ultraviolet range. To eliminate this
obstacle, Galusha et al. developed a low-temperature acid-etching technique for
biopolymeric template removal [37]. Treating the infiltrated composite with a
mixture of concentrated nitric and perchloric acid led to the complete removal of
the biological template while greatly reducing shrinkage and cracking of the silica-
based replica framework. Furthermore, the acid treatment was also beneficial for
low-temperature silica sol–gel processing, accelerating condensation and densifi-
cation due to its catalytic effect on these processes. The combined beneficial effects
of this acid treatment resulted in robust replicas with well-preserved lattice features
and minimal structural shrinkage in the range of 5%.
Apart from using oxide-based replicas of biological photonic structures for
enhancing optical properties in “conventional” applications such as sensors,
emitters, and photocatalytic cells, some of these structures are also attractive for
fabricating optical materials with entirely new – nonclassical – optical properties.
In the final section of this chapter, we will describe the sol–gel fabrication and
properties of a biotemplated titania structure that was the first example of a photonic
crystal with a calculated complete band gap at visible frequencies – a type of optical
material that is still not achievable by any other engineering technique.

9.4.3 Biotemplated Band Gap Crystals

The discovery by Bartl and coworkers that several beetles obtain their striking
coloration from light interacting with diamond-based photonic crystal lattices built
into their exoskeleton [34, 89] (see also Fig. 9.9) opened the door to exciting new
possibilities in photonic band gap research. Modeling and photonic band structure
calculations showed the diamond-based lattice found in the weevil L. augustus
possesses a complete photonic band gap in the green region of the electromagnetic
spectrum when fabricated out of polycrystalline titania [37]. However, these
calculations also revealed that the formation of a complete band gap for the
replicated structure occurs only within a narrow window of dielectric and
lattice properties. The most important findings of these theoretical studies are:
(1) The inverse of the diamond-based photonic structure produced by a single-
templation process lacks a complete band gap, regardless of the volume fraction
and refractive index of the high-dielectric component. (2) For the original beetle
structure with a high-dielectric volume fraction larger than 50%, a complete
photonic band gap would require a refractive index of at least 3.4 for the high-
dielectric component – far beyond the reach of visible light-transparent compounds.
(3) Reducing the high-dielectric volume fraction significantly lowers the refractive
index required for opening of a complete photonic band gap and can be as low
198 M.H. Bartl et al.

Fig. 9.13 Illustration of the double-imprint sol–gel biotemplating route for converting photonic
scales of the weevil L. augustus into high-dielectric titania replica. (a) Weevil L. augustus and its
green colored photonic scales (inset). (b–d) Scanning electron microscopy images of cross-
sectional views of (b) the original biopolymeric photonic structure, (c) the inverse structure
made of hybrid silica and (d) the titania replica templated from the intermediary hybrid silica
structure. Scale bars are 200 mm (a) and 1 mm (b–d). Adapted from [37]

as 2.1 for a volume fraction between 30 and 40%. From these findings it can
be concluded that only structures based on the original (not the inverse!) dia-
mond-based photonic lattice will possess a complete band gap in the visible region
when fabricated out of a material with a refractive index larger than 2.1 and
a volume fraction below 40%.
In response to these predicted requirements, Galusha et al. developed a sol–gel
chemistry-based double imprint biotemplation method (Fig. 9.13) [37]. The strat-
egy was to first create an inverse of the original beetle diamond-structured crystal,
which then could be used as a sacrificial new template to fabricate a high-dielectric
copy (“inverse-of-the-inverse”) of the original structure with a reduced volume
fraction. Hybrid organic/inorganic silica was used as the sacrificial template mate-
rial, which was infiltrated with a titania sol–gel precursor using a dip-coating
approach followed by heating of the silica/titania composite at 500 C for 2 h.
This induced titania nanocrystallization and resulted in the formation of a high-
dielectric framework with a measured refractive index of 2.3  0.1. In addition,
the authors were able to tune the volume fraction of infiltrated titania into the
required 30–40% range by successively repeating the infiltration–solidification–
crystallization cycle. Finally, the intermediary silica-based template was removed
9 Oxide-Based Photonic Crystals from Biological Templates 199

Fig. 9.14 (a) Scanning electron microscopy cross-sectional view of the biotemplated diamond-
based titania photonic crystal lattice. (b) Tilted view of the same titania structure. (c) Corr-
esponding calculated band structure diagram. The complete photonic band gap between the second
and third band is indicated by a gray rectangle. Adapted from [37]

using hydrofluoric acid etching, producing a polycrystalline anatase titania framework


with a diamond-based lattice structure and lattice periodicities at visible
wavelengths.
The obtained titania photonic crystal was investigated by a range of structural and
optical characterization techniques [37]. Structural studies were performed
by focused ion beam milling and scanning electron microscopy imaging, revealing
an excellent preservation of the diamond-based photonic lattice after the double-
templating procedure (Fig. 9.13). Examples of the final structures are also given
in Fig. 9.14 and a detailed analysis showed the original diamond-based framework – a
lattice of ABC stacked layers of hexagonally ordered air cylinders in a surrounding
high-dielectric matrix – was excellently preserved. Moreover, even after two sol–gel
200 M.H. Bartl et al.

templation steps, it was found that the overall shrinkage was kept below 15%, giving
a final lattice constant of 366  24 nm and volume fractions between 30 and 40%.
Both values are within the required range for opening a complete band gap at visible
wavelengths. The corresponding calculated band structure diagram is shown in
Fig. 9.14 and reveals a 5% wide gap-to-mid-gap ratio.
The same authors also showed that the calculated complete band gap agrees
excellently with the optical properties of the diamond-structured titania replica
determined by multidirectional reflectance microspectroscopy measurements [37].
These measurements were performed in two different ways. In the first method,
reflectance spectra were recorded normal or slightly off-normal to particular crystal
axes of the diamond-based lattice. Taking advantage of the pixilated multidomain
organization of the photonic lattice within each scale [34], the optical reflectance
properties along various crystal axes ([100], [110], and [210] directions) could
be evaluated. Analysis of the obtained spectra revealed significant overlap of the
directional reflectance peaks from the high-index titania replica, in contrast to
the well-separated reflectance features of the original low-index biopolymeric
structure. Furthermore, a comparison with the calculated band structure shows an
excellent agreement with the positions and widths of the [100], [110], and [210]
directional gaps, which are narrow and well separated for the beetle photonic
structure, but are wide and strongly overlapping for the titania replica photonic crystal.
In addition, to cover an even larger range of directions, a series of angle-dependent
reflectance spectra covering a 30 range were collected. The obtained series of
intensity-normalized spectra displayed no significant dependence of the reflectance
peak position on the recording angle. This agrees with the band structure calculations
of the first complete photonic band gap at visible frequencies – obtained by sol–gel
replicating a biological diamond-based photonic crystal structure into crystalline
titanium dioxide.

9.5 Conclusions

In biological systems, structurally complex architectures with feature sizes covering


several length scales are engineered under rather simple environmental conditions
and with limited resources – strategies still largely unmatched by our synthetic
abilities. Prime examples of such elaborate architectures are found in the amazingly
colorful world of insects, birds, and fishes in which myriad of hues and optical effects
are in large part the result of structural colors. In contrast to pigmented colors,
structural colors arise from a delicate interplay of light with periodically organized
dielectric lattices with feature sizes of a few hundreds of nanometers. For example, it
is the periodic variation of biopolymeric compounds embedded into wings and
exoskeletons that lends many butterflies and beetles their iridescent appearance.
While the pure beauty of natural iridescence in the form of opal gem stones, jewel
beetles, and feather ornaments has fascinated and inspired mankind for thousands of
years, architectural colors have recently gained tremendous interest with the
9 Oxide-Based Photonic Crystals from Biological Templates 201

introduction of band structure concepts to electromagnetism. Such materials, termed


photonic crystals, offer revolutionary new ways to use and manipulate light and are
the cornerstones for future optical devices based on light localization and control of
spontaneous emission.
In this chapter, we briefly reviewed the basic concepts of photonic crystals and
gave examples of successful synthetic photonic crystal fabrication based on top-
down microprocessing techniques as well as bottom-up writing and assembly-based
routes. While these techniques are very well suited to produce photonic crystals
with complete band gaps in the infrared regime, they have severe limitations for
engineering robust photonic crystal lattices operating in the visible region. In
contrast to our current inability to shape dielectric compounds into effective
photonic architectures with feature sizes at visible wavelengths, biological systems
have developed a wealth of elaborate structures optimized to efficiently interact
with visible light. We discussed several examples of colored butterflies and beetles,
demonstrating that they have evolved to create highly efficient and complex
photonic structures embedded into their wings and exoskeletons. The fact that
some of the most sought-after photonic structures, including diamond-based lattice
geometries, occur – to date, exclusively – in biological systems not only reflects the
ingenuity of structural engineering in nature, but also opens new avenues in optical
materials design and fabrication.
One of these highly promising new areas uses biological templates to produce
novel photonic materials. In biotemplating the best of two worlds – biology and
materials science – are combined in a synergistic approach by merging unique
structural engineering in biology with state-of-the-art materials synthesis and
processing. We reviewed several successful biotemplating strategies using three-
dimensional photonic lattices from butterflies and beetles as unique molds for
replication into oxide-based inorganic compounds such as titania, silica, and alu-
mina. While typical deposition and evaporation based techniques are efficient in
infiltrating and replicating open-surface photonic structures such as those found in
the wings of butterflies, sol–gel chemistry routes are particularly well suited to
replicate lattices surrounded by biopolymeric shells.
All of these biotemplating methods have been shown to create oxide-based
replica with high fidelity, preserving even nanoscale fine features of the original
biological photonic structures. In addition, the combination of interesting
biophotonic lattice geometries with the intrinsic properties of the oxide-based
replica compounds creates new functionalities and interesting applications as
sensors, photoanodes, and light emitters. Finally, the recently reported step toward
a complete photonic band gap at visible frequencies – the first of its kind – by titania
sol–gel replication of a diamond-based photonic structure found in a weevil,
underlines the enormous potential of biotemplating-based photonic engineering
and paves the road for novel optical materials and phenomena based on light
localization at visible frequencies.
202 M.H. Bartl et al.

References

1. Bartl, M.H., Boettcher, S.W., Frindell, K.L., Stucky, G.D.: 3-D molecular assembly of
function in titania-based composite material systems. Acc. Chem. Res. 38, 263 (2005)
2. Corma, A.: From microporous to mesoporous molecular sieve materials and their use in
catalysis. Chem. Rev. 97, 2373 (1997)
3. Davis, M.E.: Ordered porous materials for emerging applications. Nature 417, 813 (2002)
4. Soler-illia, G.J.D., Sanchez, C., Lebeau, B., Patarin, J.: Chemical strategies to design textured
materials: From microporous and mesoporous oxides to nanonetworks and hierarchical
structures. Chem. Rev. 102, 4093 (2002)
5. Yang, P.D., Zhao, D.Y., Margolese, D.I., Chmelka, B.F., Stucky, G.D.: Generalized
syntheses of large-pore mesoporous metal oxides with semicrystalline frameworks. Nature
396, 152 (1998)
6. Holland, B.T., Blanford, C.F., Stein, A.: Synthesis of macroporous minerals with highly
ordered three-dimensional arrays of spheroidal voids. Science 281, 538 (1998)
7. Davis, M.E., Lobo, R.F.: Zeolite and Molecular-Sieve Synthesis. Chem. Mater. 4, 756 (1992)
8. Stucky, G.D., Huo, Q., Firouzi, A., Chmelka, B.F., Schacht, S., Voigt-Martin, I.G., Schuth,
F.: Progress in zeolite and microporous materials, studies in surface science and catalysis.
In: Chon, H., Ihm, S.-K., Uh, Y.S. (eds.) Directed synthesis of organic/inorganic composite
structures, vol. 105. Elsevier, Amsterdam (1996)
9. Schuth, F., Schmidt, W.: Microporous and mesoporous materials. Adv. Mater. 14, 629 (2002)
10. Sanchez, C., Boissiere, C., Grosso, D., Laberty, C., Nicole, L.: Design, synthesis, and
properties of inorganic and hybrid thin films having periodically organized nanoporosity.
Chem. Mater. 20, 682 (2008)
11. Wirnsberger, G., Yang, P.D., Scott, B.J., Chmelka, B.F., Stucky, G.D.: Mesostructured
materials for optical applications: from low-k dielectrics to sensors and lasers. Spectrochim.
Acta A 57, 2049 (2001)
12. Bartl, M.H., Stucky, G.D.: Mesostructured thin film oxides. In: Ramanathan, S. (ed.) Thin
Film Metal-Oxides, pp. 255–279. Springer, New York (2010)
13. Scott, B.J., Wirnsberger, G., Stucky, G.D.: Mesoporous and mesostructured materials for
optical applications. Chem. Mater. 13, 3140 (2001)
14. Marlow, F.: Optical materials based on nanoscaled guest/host composites. Mol. Cryst. Liquid
Cryst. 341, 289 (2000)
15. Galusha, J.W., Tsung, C.K., Stucky, G.D., Bartl, M.H.: Optimizing sol-gel infiltration and
processing methods for the fabrication of high-quality planar titania inverse opals. Chem.
Mater. 20, 4925 (2008)
16. Imhof, A., Pine, D.J.: Ordered macroporous materials by emulsion templating. Natrue 389,
948 (1997)
17. Wijnhoven, J., Vos, W.L.: Preparation of photonic crystals made of air spheres in titania.
Science 281, 802 (1998)
18. Joannopoulos, J.D., Meade, R.D., Winn, J.N.: Photonic Crystals: Molding the Flow of Light.
Princeton Press, Princeton (1995)
19. Joannopoulos, J.D., Villeneuve, P.R., Fan, S.H.: Photonic crystals: Putting a new twist on
light. Nature 386, 143 (1997)
20. John, S.: Strong localization of photons in certain disordered dielectric superlattices. Phys.
Rev. Lett. 58, 2486 (1987)
21. Yablonovitch, E.: Inhibited spontaneous emission in solid-state physics and electronics.
Phys. Rev. Lett. 58, 2059 (1987)
22. John, S.: Localization of light. Phys. Today 44, 32 (2008)
23. Woldeyohannes, M., John, S.: Coherent control of spontaneous emission near a photonic
band edge. J. Opt. B 5, R43 (2003)
24. Soukoulis, C.M. (ed.): Photonic Crystals and Light Localization. Kluwer, Dordrecht (2001)
25. Mason, C.W.: Structural colors in insects I. J. Phys. Chem. 30, 383 (1926)
9 Oxide-Based Photonic Crystals from Biological Templates 203

26. Mason, C.W.: Structural colors in insects II. J. Phys. Chem. 31, 321 (1927)
27. Doucet, S.M., Meadows, M.G.: Iridescence: a functional perspective. J. R. Soc. Interface 6,
S115 (2009)
28. Seago, A.E., Brady, P., Vigneron, J.P., Schultz, T.D.: Gold bugs and beyond: a review of
iridescence and structural colour mechanisms in beetles (Coleoptera). J. R. Soc. Interface 6,
S165 (2009)
29. Vukusic, P., Sambles, J.R.: Photonic structures in biology. Nature 424, 852 (2003)
30. Fan, T.-X., Chow, S.-K., Zhang, D.: Biomorphic mineralization: From biology to materials.
Prog. Mater. Sci. 54, 542 (2009)
31. Biro, L.P., Kertesz, K., Vertesy, Z., Mark, G.I., Balint, Z., Lousse, V., Vigneron, J.P.: Living
photonic crystals: Butterfly scales - nanostructure and optical properties. Mater. Sci. Eng. C
27, 941 (2007)
32. Parker, A.R.: Conservative photonic crystals imply indirect transcription from genotype to
phenotype. Recent Res. Dev. Entomol. 5, 59 (2006)
33. Srinivasarao, M.: Nano-optics in the biological world: Beetles, butterflies, birds, and moths.
Chem. Rev. 99, 1935 (1999)
34. Galusha, J.W., Richey, L.R., Gardner, J.S., Cha, J.N., Bartl, M.H.: Discovery of a diamond-
based photonic crystal structure in beetle scales. Phys. Rev. E 77, 050904 (2008)
35. Galusha, J.W., Richey, L.R., Jorgensen, M.R., Gardner, J.S., Bartl, M.H.: Study of natural
photonic crystals in beetle scales and their conversion into inorganic structures via a sol-gel
bio-templating route. J. Mater. Chem. 20, 1277 (2010)
36. Jorgensen, M.R., Bartl, M.H.: Biotemplating routes to three-dimensional photonic crystals.
J. Mater. Chem. 21, 10583 (2011)
37. Galusha, J.W., Jorgensen, M.R., Bartl, M.H.: Diamond-structured titania photonic bandgap
crystals from biological templates. Adv. Mater. 22, 107 (2010)
38. Martin-Palma, R.J., Pantano, C.G., Lakhtakia, A.: Replication of fly eyes by the conformal-
evaporated-film-by-rotation. Nanotechnology 19, 5 (2008)
39. Martı́n-Palma, R.J., Pantano, C.G., Lakhtakia, A.: Biomimetization of butterfly wings by the
conformal-evaporated-film-by-rotation technique for photonics. Appl. Phys. Lett. 93, 083901
(2008)
40. Huang, J., Wang, X., Wang, Z.L.: Controlled replication of butterfly wings for achieving
tunable photonic properties. Nano. Lett. 6, 2325 (2006)
41. Gaillot, D.P., Deparis, O., Welch, V., Wagner, B.K., Vigneron, J.P., Summers, C.J.:
Composite organic-inorganic butterfly scales: Production of photonic structures with atomic
layer deposition. Phys. Rev. E 78, 031922 (2008)
42. Zhang, W., Zhang, D., Fan, T., Gu, J., Ding, J., Wang, H., Guo, Q., Ogawa, H.: Novel
photoanode structure templated from butterfly wing scales. Chem. Mater. 21, 33 (2008)
43. Zhang, W., Zhang, D., Fan, T., Ding, J., Gu, J., Guo, Q., Ogawa, H.: Biosynthesis of
cathodoluminescent zinc oxide replicas using butterfly (Papilio paris) wing scales as
templates. Mater. Sci. Eng. C 29, 92 (2009)
44. Lopez, C.: Materials aspects of photonic crystals. Adv. Mater. 15, 1679 (2003)
45. Tetreault, N., Miguez, H., Ozin, G.A.: Silicon inverse opal - a platform for photonic bandgap
research. Adv. Mater. 16, 1471 (2004)
46. Chen, J.I.L., von Freymann, G., Choi, S.Y., Kitaev, V., Ozin, G.A.: Amplified photochemis-
try with slow photons. Adv. Mater. 18, 1915 (2006)
47. Halaoui, L.I., Abrams, N.M., Mallouk, T.E.: Increasing the conversion efficiency of dye-
sensitized TiO2 photoelectrochemical cells by coupling to photonic crystals. J. Phys. Chem.
B 109, 6334 (2005)
48. Mihi, A., Miguez, H.: Origin of light-harvesting enhancement in colloidal-photonic-crystal-
based dye-sensitized solar cells. J. Phys. Chem. B 109, 15968 (2005)
49. Busch, K., John, S.: Photonic band gap formation in certain self-organizing systems. Phys.
Rev. E 58, 3896 (1998)
204 M.H. Bartl et al.

50. Ho, K.M., Chan, C.T., Soukoulis, C.M.: Existence of a photonic gap in periodic dielectric
structures. Phys. Rev. Lett. 65, 3152 (1990)
51. Hynninen, A.P., Thijssen, J.H.J., Vermolen, E.C.M., Dijkstra, M., Van Blaaderen, A.: Self-
assembly route for photonic crystals with a bandgap in the visible region. Nat. Mater. 6, 202
(2007)
52. Maldovan, M., Thomas, E.L.: Diamond-structured photonic crystals. Nat. Mater. 3, 593
(2004)
53. Moroz, A.: Metallo-dielectric diamond and zinc-blende photonic crystals. Phys. Rev. B 66,
115109 (2002)
54. Ngo, T.T., Liddell, C.M., Ghebrebrhan, M., Joannopoulos, J.D.: Tetrastack: Colloidal dia-
mond-inspired structure with omnidirectional photonic band gap for low refractive index
contrast. Appl. Phys. Lett. 88, 242920 (2006)
55. Yablonovitch, E., Gmitter, T.J., Leung, K.M.: Photonic band structure: The face-centered-
cubic case employing nonspherical atoms. Phys. Rev. Lett. 67, 2295 (1990)
56. Blanco, A., Chomski, E., Grabtchak, S., Ibisate, M., John, S., Leonard, S.W., Lopez, C.,
Meseguer, F., Miguez, H., Mondia, J.P., Ozin, G.A., Toader, O., van Driel, H.M.: Large-scale
synthesis of a silicon photonic crystal with a complete three-dimensional bandgap near 1.5
micrometres. Nature 405, 437 (2000)
57. Garcia-Santamaria, F., Xu, M.J., Lousse, V., Fan, S.H., Braun, P.V., Lewis, J.A.: A germa-
nium inverse woodpile structure with a large photonic band gap. Adv. Mater. 19, 1567 (2007)
58. Lin, S.Y., Fleming, J.G., Hetherington, D.L., Smith, B.K., Biswas, R., Ho, K.M., Sigalas, M.
M., Zubrzycki, W., Kurtz, S.R., Bur, J.: A three-dimensional photonic crystal operating at
infrared wavelengths. Nature 394, 251 (1998)
59. Noda, S., Tomoda, K., Yamamoto, N., Chutinan, A.: Full three-dimensional photonic
bandgap crystals at near-infrared wavelengths. Science 289, 604 (2000)
60. Maldovan, M., Thomas, E.L., Carter, C.W.: Layer-by-layer diamond-like woodpile structure
with a large photonic band gap. Appl. Phys. Lett. 84, 362 (2004)
61. Qi, M.H., Lidorikis, E., Rakich, P.T., Johnson, S.G., Joannopoulos, J.D., Ippen, E.P., Smith,
H.I.: A three-dimensional optical photonic crystal with designed point defects. Nature 429,
538 (2004)
62. Vlasov, Y.A., Bo, X.Z., Sturm, J.C., Norris, D.J.: On-chip natural assembly of silicon
photonic bandgap crystals. Nature 414, 289 (2001)
63. Wong, S., Deubel, M., Perez-Willard, F., John, S., Ozin, G.A., Wegener, M., von Freymann, G.:
Direct laser writing of three-dimensional photonic crystals with complete a photonic bandgap in
chalcogenide glasses. Adv. Mater. 18, 265 (2006)
64. Hermatschweiler, M., Ledermann, A., Ozin, G.A., Wegener, M.: vonFreymann, G.: Fabri-
cation of silicon inverse woodpile photonic crystals. Adv. Mater. 17, 2273 (2007)
65. Sun, H.-B., Matsuo, S., Misawa, H.: Three-dimensional photonic crystal structures achieved
with two-photon-absorption photopolymerization of resin. Appl. Phys. Lett. 74, 786 (1999)
66. Deubel, M., von Freymann, G., Wegener, M., Pereira, S., Busch, K., Soukoulis, C.M.: Direct
laser writing of three-dimensional photonic-crystal templates for telecommunications. Nat.
Mater. 3, 444 (2004)
67. Gratson, G.M., Xu, M., Lewis, J.A.: Microperiodic structures: Direct writing of three-
dimensional webs. Nature 428, 386 (2004)
68. Lewis, J.A.: Direct ink writing of 3D functional materials. Adv. Funct. Mater. 16, 2193
(2006)
69. Lewis, J.A., Braun, P.V.: Direct-write assembly of three-dimensional photonic crystals:
Conversion of polymer scaffolds to silicon hollow-woodpile structures. Adv. Mater. 18,
461 (2006)
70. Tétreault, N., von Freymann, G., Deubel, M., Hermatschweiler, M., Pérez-Willard, F., John, S.,
Wegener, M., Ozin, G.A.: New route to three-dimensional photonic bandgap materials: Silicon
double inversion of polymer templates. Adv. Mater. 18, 457 (2006)
9 Oxide-Based Photonic Crystals from Biological Templates 205

71. Jiang, P., Bertone, J.F., Hwang, K.S., Colvin, V.L.: Single-crystal colloidal multilayers of
controlled thickness. Chem. Mater. 11, 2132 (1999)
72. Norris, D.J., Arlinghaus, E.G., Meng, L., Heiny, R., Scriven, L.E.: Opaline photonic crystals:
How does self-assembly work? Adv. Mater. 16, 1393 (2004)
73. Shimmin, R.G., DiMauro, A.J., Braun, P.V.: Slow vertical deposition of colloidal crystals:
A Langmuir-Blodgett process? Langmuir 22, 6507 (2006)
74. Subramania, G., Lee, Y.J., Brener, I., Luk, T.S., Clem, P.G.: Nano-lithographically fabricated
titanium dioxide based visible frequency three dimensional gap photonic crystal. Opt.
Express 15, 13049 (2007)
75. Subramania, G., Lee, Y.-J., Fischer, A.J., Koleske, D.D.: Log-pile TiO2 photonic crystal for
light control at near-UV and visible wavelengths. Adv. Mater. 22, 487 (2010)
76. Urbas, A.M., Maldovan, M., DeRege, P., Thomas, E.L.: Bicontinuous cubic block copolymer
photonic crystals. Adv. Mater. 14, 1850 (2002)
77. Manoharan, V.N., Imhof, A., Thorne, J.D., Pine, D.J.: Photonic crystals from emulsion
templates. Adv. Mater. 13, 447 (2001)
78. Subramania, G., Constant, K., Biswas, R., Sigalas, M.M., Ho, K.-M.: Inverse face-centered
cubic thin film photonic crystals. Adv. Mater. 13, 443 (2001)
79. Subramanian, G., Manoharan, V.N., Thorne, J.D., Pine, D.J.: Ordered macroporous materials
by colloidal assembly: A possible route to photonic bandgap materials. Adv. Mater. 11, 1261
(1999)
80. Holland, B.T., Blanford, C.F., Do, T., Stein, A.: Synthesis of highly ordered, three-dimen-
sional, macroporous structures of amorphous or crystalline inorganic oxides, phosphates, and
hybrid composites. Chem. Mater. 11, 795 (1999)
81. Wijnhoven, J., Bechger, L., Vos, W.L.: Fabrication and characterization of large
macroporous photonic crystals in titania. Chem, Mater. 13, 4486 (2001)
82. Kim, H., Lee, H.-B.-R., Maen, W.-J.: Applications of atomic layer deposition to
nanofabrication and emerging nanodevices. Thin Solid Films 517, 2563 (2009)
83. King, J.S., Gaillot, D.P., Graugnard, E., Summers, C.J.: Conformally back-filled, non-close-
packed inverse-opal photonic crystals. Adv. Mater. 18, 1063 (2006)
84. King, J.S., Graugnard, E., Roche, O.M., Sharp, D.N., Scrimgeour, J., Denning, R.G.,
Turberfield, A.J., Summers, C.J.: Infiltration and inversion of holographically defined poly-
mer photonic crystal templates by atomic layer deposition. Adv. Mater. 18, 1561 (2006)
85. King, J.S., Graugnard, E., Summers, C.J.: TiO2 inverse opals fabricated using low-
temperature atomic layer deposition. Adv. Mater. 17, 1010 (2005)
86. Garcia-Santamaria, F., Miyazaki, H.T., Urquia, A., Ibisate, M., Belmonte, M., Shinya, N.,
Meseguer, F., Lopez, C.: Nanorobotic manipulation of microspheres for on-chip diamond
architectures. Adv. Mater. 14, 1144 (2002)
87. Vigneron, J.P., Colomer, J.F., Rassart, M., Ingram, A.L., Lousse, V.: Structural origin of the
colored reflections from the black-billed magpie feathers. Phys. Rev. E 73, 021914 (1993)
88. Prum, R.O., Dufresne, E.R., Quinn, T., Waters, K.: Development of colour-producing beta-
keratin nanostructures in avian feather barbs. J. R. Soc. Interface 6, S253 (2009)
89. Galusha, J.W., Richey, L.R., Bartl, M.H.: High-resolution three-dimensional reconstruction
of photonic crystal structure found in beetle scales. IEEE LEOS Summer Conference, Adv.
Biophotonics 83 (2008)
90. Ghiradella, H., Aneshansley, D., Eisner, T., Silberglied, R.E., Hinton, H.E.: Ultraviolet
reflection of a male butterfly: Interference color caused by thin-layer elaboration of wing
scales. Science 178, 1214 (1972)
91. Parker, A.R.: The diversity and implications of animal structural colours. J. Exp. Biol. 201,
2343 (1998)
92. Vigneron, J.P., Colomer, J.F., Vigneron, N., Lousse, V.: Natural layer-by-layer photonic
structure in the squamae of Hoplia coerulea (Coleoptera). Phys. Rev. E 72, 904 (2005)
206 M.H. Bartl et al.

93. Ha, Y.H., Vaia, R.A., Lynn, W.F., Costantino, J.P., Shin, J., Smith, A.B., Matsudaira, P.T.,
Thomas, E.L.: Three-dimensional network photonic crystals via cyclic size reduction/
infiltration of sea urchin exoskeleton. Adv. Mater. 16, 1091 (2004)
94. Parker, A.R., McPhedran, R.C., McKenzie, D.R., Botten, L.C., Nicorovici, N.A.P.: Photonic
engineering – Aphrodite’s iridescence. Nature 409, 36 (2001)
95. Li, Y., Lu, Z., Yin, H., Yu, X., Liu, X., Zi, J.: Structural origin of the brown color of barbules
in male peacock tail feathers. Phys. Rev. E 72, 010902 (2000)
96. Kertesz, K., Molnar, G., Vertesy, Z., Koos, A.A., Horvath, Z.E., Mark, G.I., Tapaszto, L.,
Balint, Z., Tamaska, I., Deparis, O., Vigneron, J.P., Biro, L.P.: Photonic band gap materials in
butterfly scales: A possible source of "blueprints". Mater. Sci. Eng. B 149, 259 (2008)
97. Kinoshita, S., Yoshioka, S., Miyazaki, J.: Physics of structural colors. Rep. Prog. Phys. 71,
76401 (2008)
98. Prum, R.O., Quinn, T., Torres, R.H.: Anatomically diverse butterfly scales all produce
structural colours by coherent scattering. J. Exp. Biol. 209, 748 (2006)
99. Schultz, T.D., Rankin, M.A.: Developmental changes in the interference reflectors and
colorations of tiger beetles (Cicindela). J. Exp. Biol. 117, 111 (1985)
100. Michielsen, K., Stavenga, D.G.: Gyroid cuticular structures in butterfly wing scales:
biological photonic crystals. J. R. Soc. Interface 5, 85 (2008)
101. Galusha, J.W., Carter, K., Bartl, M.H.: 3-D photonic band structure engineering in self-
assembled photonic crystals. Mater. Res. Soc. Symp. Proc. 988, QQ05-08 (2006)
102. Shawkey, M.D., Saranathan, V., Palsdottir, H., Crum, J., Ellisman, M.H., Auer, M., Prum, R.O.:
Electron tomography, three-dimensional Fourier analysis and colour prediction of a three-
dimensional amorphous biophotonic nanostructure. J. R. Soc. Interface 6, S213 (2009)
103. Wilts, B.D., Leertouwer, H.L., Stavenga, D.G.: Imaging scatterometry and microspectro-
photometry of lycaenid butterfly wing scales with perforated multilayers. J. R. Soc. Interface
6, S185 (2009)
104. Vukusic, P., Stavenga, D.G.: Physical methods for investigating structural colours in
biological systems. J. R. Soc. Interface 6, S133 (2009)
105. Abramoff, M.D., Magelhaes, P.J., Ram, S.J.: Image Processing with ImageJ. Biophotonics
Int. 11, 36 (2004)
106. Thevenaz, P., Ruttimann, U.E., Unser, M.: A pyramid approach to subpixel registration based
on intensity. IEEE Trans. Image Process. 7, 27 (1998)
107. Weinstein, D.M., Parker, S.G., Simpson, J., Zimmerman, K., Jones, G.: In: Hansen, C.D.,
Johnson, C.R. (eds) The Visualization Handbook, pp. 615. Elsevier, Burlington (2005)
108. Ghiradella, H.T., Butler, M.W.: Many variations on a few themes: a broader look at
development of iridescent scales (and feathers). J. R. Soc. Interface 6, S243 (2009)
109. Poladian, L., Wickham, S., Lee, K., Large, M.C.J.: Iridescence from photonic crystals and its
suppression in butterfly scales. J. R. Soc. Interface 6, S233 (2009)
110. Biro, L.P.: Photonic nanoarchitectures of biologic origin in butterflies and beetles. Mater. Sci.
Eng. B (2009). doi:10.1016/j.mseb.2009.10.027
111. Potyrailo, R.A., Ghiradella, H., Vertiatchikh, A., Dovidenko, K., Cournoyer, J.R., Olson, E.:
Morpho butterfly wing scales demonstrate highly selective vapour response. Nat. Photon 1,
123 (2007)
112. Galusha, J.W., Jorgensen, M.R., Richey, L.R., Gardner, J.S., Bartl, M.H.: Oxide-based
photonic crystals from biological templates. Proc. SPIE 7401, 74010G (2009)
113. Johnson, S.G., Joannopoulos, J.D.: Three-dimensionally periodic dielectric layered structure
with omnidirectional photonic band gap. Appl. Phys. Lett. 77, 3490 (2000)
114. Huang, J., Wang, X., Wang, Z.L.: Bio-inspired fabrication of antireflection nanostructures by
replicating fly eyes. Nanotechnol. 19, 025602 (2008)
115. Knez, M., Kadri, A., Wege, C., G€ osele, U., Jeske, H., Nielsch, K.: Atomic layer deposition on
biological macromolecules: Metal oxide coating of tobacco mosaic virus and ferritin. Nano
Lett. 6, 1172 (2006)
9 Oxide-Based Photonic Crystals from Biological Templates 207

116. Lakhtakia, A., Martin-Palma, R.J., Motyka, M.A., Pantano, C.G.: Fabrication of free-standing
replicas of fragile, laminar, chitinous biotemplates. Bioinsp. Biomim. 4, 034001 (2009)
117. Lakhtakia, A., Martin-Palma, R.J., Pantano, C.G.: Towards replication of the exoskeleton of
Lamprocyphus augustus for photonic applications. Proc. SPIE 7401, 74010H (2009)
118. Brinker, D.J., Scherrer, G.W.: Sol Gel Science, The Physics and Chemistry of Sol Gel
Processing. Academic, San Diego (1990)
119. Zhu, S., Zhang, D., Chen, Z., Gu, J., Li, W., Jiang, H., Zhou, G.: A simple and effective
approach towards biomimetic replication of photonic structures from butterfly wings.
Nanotechnol. 20, 315303 (2009)
120. Zhao, D., Yang, P., Melosh, N., Feng, J., Chmelka, B.F., Stucky, G.D.: Continuous
mesoporous silica films with highly ordered large pore structures. Adv. Mater. 10, 1380
(1998)
Chapter 10
Low Dimensionality and Epitaxial Stabilization
in Metal-Supported Oxide Nanostructures:
MnxOy on Pd(100) MnxOy

Cesare Franchini and Francesco Allegretti

We present a survey of the growth and structure of manganese oxide nanolayers on


a Pd(100) substrate, investigated in two different thickness regimes through a
plethora of surface science techniques [scanning tunneling microscopy (STM),
atomic force microscopy (AFM), low energy electron diffraction (LEED), SPA-
LEED, X-ray photoemission spectroscopy (XPS), X-ray absorption spectroscopy
(XAS), and high-resolution electron energy loss spectroscopy (HREELS)] and
state-of-the-art theoretical tools [density functional theory (DFT) and hybrid DFT
approaches]. In the high thickness regime, above 10–15 monolayers, depending on
the preparation conditions, different films with specific growth direction and stoi-
chiometry are formed. At low and intermediate pressure (1  107 and
2–5  107 mbar, respectively), the oxide structures are already bulk-like in
terms of their in-plane lattice constant and can be assigned to MnO(111) and
MnO(100), respectively. At high pressures (5  106 mbar), (001)-oriented
Mn3O4 layers are obtained by oxidation of MnO(100). To explore the epitaxial
(geometric) relationships that favor the growth of the different oxide phases, we
have investigated the atomistic details of different oxide/oxide and oxide/metal
interfaces. In particular, we have addressed the issue of the stability of the
Mn3O4(001)/MnO(001) interface and determined the phase stability diagram of
MnxyOy/Pd(100) phases at a Mn coverage of about 1 ML. In the latter low thickness
regime, we have identified nine different two-dimensional (2D) phases, which are
novel in terms of their structural and electronic properties. These nanophases can be
classified according to similar building block units and described either as O–Mn–O
MnO(111)-like trilayers or in terms of metal-deficient MnO(100)-like monolayers,

C. Franchini (*)
Center for Computational Materials Science, Universit€at Wien, Wien 1090, Austria
e-mail: cesare.franchini@univie.ac.at
F. Allegretti
Institute of Physics, Surface and Interface Physics, Karl-Franzens University Graz,
Graz 8010, Austria

J. Wu et al. (eds.), Functional Metal Oxide Nanostructures, Springer Series 209


in Materials Science 149, DOI 10.1007/978-1-4419-9931-3_10,
# Springer Science+Business Media, LLC 2012
210 C. Franchini and F. Allegretti

and therefore we argue that they mediate the epitaxial growth of MnO thicker films
on Pd(100) by providing structurally graded interfaces. Moreover, the formation of
O or Mn vacancies drives the transition between 2D phases with similar structural
units but different lattice periodicity, indicating that ion vacancies, mixed valence
states, and substoichiometry lie at the basis of the architectural flexibility in the
monolayer regime. Interestingly, the latter concepts play a major role in the more
complex class of functional oxides such as the manganites, of which binary
manganese oxides are the simplest parent compounds.

10.1 Introduction

Since several decades, the study of transition metal oxides represents a field of
active and intense research. Thanks to the uniquely rich spectrum of structural,
electronic, and magnetic properties, which cover the range from metallic behavior
to magnetic insulators and encompass spectacular phenomena as superconductivity,
charge ordering, and colossal magnetoresistance, these compounds have attracted
the interest of physicists, solid state chemists, and materials scientists. Renewed
interest has been fuelled by the advances in the synthesis processes of oxide
materials, which now allow control over the structure and stoichiometry at the
level of a unit cell. This has opened up new far-reaching perspectives for both
fundamental studies and technological applications, in particular in the field of
functional and multifunctional oxides, many of which incorporate in their structure
one or more transition metal atom species. Functional oxide materials have unique
physical and chemical properties, which can be suitably controlled and modified by
means of external stimuli, such as changes in temperature and pressure, electric or
magnetic fields, and the adsorption of foreign atoms or molecules, thereby enabling
the development of new functionalities. As a result, a wide range of intriguing
phenomena can be observed, ranging from superconductivity to ferroelectricity,
piezoelectricity, and exotic magnetic behaviors. The ability to tailor and tune the
properties of these functional materials thus guarantees a high potential for
applications in micro- and nanoelectronics, spintronics, heterogeneous catalysis,
gas sensing, energy harvesting, etc.
In addition to these developments, novel properties can also arise by the scaling
down of the dimensions of oxide-based devices. At the nanoscale, in fact, effects
arising from the reduced dimensionality, geometric confinement, and proximity of
surfaces and interfaces as well as effects arising from the coupling of oxide to the
supporting material (so-called substrate-induced effects, such as electronic
hybridization, charge transfer, and elastic strain) may convey new physical and
chemical properties and therefore new functionalities to oxide systems. This poten-
tial is indeed reflected in the enormous impulse given to fundamental and applied
research in the fields of nanoscience and nanotechnology. Among low-dimensional
nanostructured oxide systems, a prominent role is played by transition metal oxide
ultrathin films supported on metal surfaces. Just to name a few technological
10 Low Dimensionality and Epitaxial Stabilization. . . 211

applications, they are typically employed as gate dielectrics and tunneling barrier
layers in conventional and novel electronic devices [8, 62], as protective layers in
corrosion prevention and inhibition [41], as gas sensor materials [33, 70], and as
support surfaces in the field of heterogeneous catalysis [21]. In fundamental
scientific studies, on the other hand, thin films of transition metal oxides grown
on metal single crystal surfaces constitute preferred model systems for the elucida-
tion of emerging phenomena at the atomistic scale. Not only the presence of the
metal substrate allows circumventing charging problems arising from the insulating
character of many oxides, but also it may lead – through the active participation in
the elastic and electronic coupling – to the stabilization of novel hybrid systems,
whose structural, electronic, and magnetic properties bear no correspondence to
those of bulk oxides [59].
With this general frame in mind, in this chapter we focus on the study of ultrathin
manganese oxide films epitaxially grown on Pd(100), providing a brief review of a
combined theoretical and experimental investigation performed by our research
groups at the University of Vienna and at the Karl-Franzens University of Graz. All
calculations presented in this chapter have been performed using the Vienna ab
initio Simulation Package (VASP) [45, 46] in the framework of DFT [44] and
hybrid DFT [39].
Due to the ability of manganese to assume different oxidation states, ranging
from +2 to +7 [38], Mn oxides in bulk compounds exhibit a number of
stoichiometries with a complex phase diagram [27]. Their architecture is dictated
by the ability of the Mn/O complexes to assemble by corner sharing, edge sharing,
or double corner sharing, such that more than 30 Mn oxide mineral phases occur in
nature [63]. As a consequence, Mn oxides display a richness of behaviors which
render them attractive systems in applications of heterogeneous catalysis [9] (for
example, as electrocatalysts in fuel cells [71, 76]) as well as in applications as
electrode materials in solid state batteries [7] and in the environmental waste
treatment [20, 74]. Moreover, Mn oxides are parent compounds of a particular
class of functional oxides, the manganites, the properties of which are determined
by the complex interplay among spin, orbital, and lattice degrees of freedom, which
results in outstanding phenomena [65] such as giant magnetoresistance [75],
metal–insulator transitions, and orbital orderings [40]. These spectacular properties
promise to find application in the fields of magnetic recording and novel spintronic
devices. As to the study of Mn oxides in reduced dimensions, an increasing effort
has been devoted in the last decade, which has revealed the richness of unusual
behaviors and an even higher degree of complexity relative to bulk Mn oxides. For
example, the ferromagnetic behavior of small MnO nanoparticles, anomalous with
respect to the observed antiferromagnetic ordering in the bulk phase, has been first
predicted by DFT [58] and then confirmed by subsequent experiments [49].
Recently, ferromagnetism of MnO and Mn3O4 nanowires has also been observed
[55], whereas unconventional exchange bias interaction has been reported for
oxide-coated Mn nanoparticles with a Mn3O4 shell [72], the latter effect being
attributed to an unusual spin alignment sequence at the interface.
212 C. Franchini and F. Allegretti

To understand the unconventional properties of low-dimensional Mn oxides,


fundamental studies on model systems are desirable, which require a detailed
knowledge of the oxide structures at the atomistic level in the effort to elucidate
the structure–properties relationship. So far, these studies on model systems have
mainly focused on the growth of MnO thin films on the Ag(001) surface. Due to the
relatively large in-plane lattice constant of MnO(001) (aMnO ¼ 3:14 Å), the tem-
plate has been chosen to ensure a reasonable overlayer–substrate matching
(aAg ¼ 2:89 Å), which is expected to favor better epitaxial growth. Indeed, despite
the still large mismatch (about 9%), films of good quality have been obtained
[54, 73]. Further investigations cast light on the evolution of the structural and
electronic properties with the oxide thickness due to the partial release of the
epitaxial strain [18, 56, 57]. However, no detailed studies of the oxide–Ag(001)
interface in terms of the Mn oxide phases formed at coverage below or about one
monolayer (ML) are available in the literature. Interface-stabilized MnxOy
monolayers have instead been reported on two different substrates, Rh(100) [60]
and Pt(111) [37], but the literature is still sparse and the unambiguous structural
assignment of the oxide phases has not been yet accomplished. In this context, we
have recently performed a joint experimental–theoretical investigation aimed to
extensively characterize the growth and structure of MnxOy ultrathin films on a Pd
(100) substrate. The film thickness range extended from a few angstroms,
corresponding to the interface-stabilized phases of the monolayer regime, up to
30–50 Å, where bulk-like behavior sets in. The detailed theory–experiment com-
parison for the Mn oxide systems, which will be presented in the next section,
proved a decisive factor for the unambiguous assignment of the oxide phases, and it
enabled to identify the interface-stabilized oxide structures that mediate the epitax-
ial growth. Moreover, it proved highly beneficial in that it allowed assessing the
reliability of DFT and post-DFT approaches applied to the Mn oxides. This is a very
important issue in the light of the well-known difficulties encountered by standard
DFT in dealing with strongly correlated materials [1].

10.2 Growth of MnxOy Layers on Pd(100)

The growth of Mn oxides on Pd(100) has been the subject of a series of recent
papers [3, 4, 15, 25, 26, 50], which explore in a rather systematic way the phase
stability diagram at different oxide thicknesses. In particular, it has been shown that
below 1 ML a complex surface phase diagram with a multitude of novel Mn oxide
structures develops, the oxide structures being characterized by specific structural
building blocks and vibrational properties [25, 26, 50]. In the high thickness regime,
upon deposition of 20–30 ML, MnO(100) films exhibiting good long-range order
can be grown epitaxially [3], despite the considerable lattice mismatch with the
substrate (14%). The MnO(100) structure can be preferentially converted either
into MnO(111) [3, 4] or into Mn3O4(001) [15] depending on the combination of
temperature and applied oxygen pressure. The evolution of the physical properties
10 Low Dimensionality and Epitaxial Stabilization. . . 213

of the manganese oxide layers on Pd(100) with respect to the changes in film
thickness provides a unique method to separate surface (2D) and bulk (3D) effects.
In fact, a decrease of film thicknesses from the high-coverage (multilayer) to the
low-coverage [(sub)monolayer] regime corresponds to an increase of the surface to
bulk ratio, and low-dimensional effects become more distinguishable. The funda-
mental characteristics of the MnxOy/Pd(100) system are discussed in detail in
Sects. 10.2.1 and 10.2.2 and can be summarized as follows:

Mnx Oy =Pdð100Þ

Low coverage regime, 0.75 monolayer (see Sect. 10.2.1). Ultrathin layers of
variable MnxOy stoichiometry, only 1–2 ML thick, can be formed at the
metal–MnO interface. At least nine different 2D MnxOy phases on Pd(100) are
found, which are novel in terms of the known Mn oxide bulk crystal structures and
which are stabilized by the metal–oxide interface and by the confinement in the
direction perpendicular to the surface. These low-dimensional interface-stabilized
phases may mediate the epitaxial growth of thicker layers by providing structurally
graded interfaces.
High coverage regime, 20–30 layers (see Sect. 10.2.2). MnO(100) with a bulk-
like in-plane lattice constant is stable under a wide range of pressure and tempera-
ture. Through the appropriate tuning of temperature and oxygen pressure the MnO
(100) films can be transformed into MnO(111) [annealing at elevated temperatures
or reactive evaporation at lower oxygen pressures (pO2 <1  107 mbar)] or
transformed into Mn3O4(001) surfaces [high-temperature oxidation at relatively
high oxygen pressure (pO2  5  106 mbar)].

10.2.1 Low Coverage Regime

The experimental phase stability diagram of Mn oxides on Pd(100) below 1 ML is


depicted schematically in Fig. 10.1, where the various Mn oxide nanolayer phases
are ordered as a function of the oxygen pressure pO2 and of the oxygen chemical
potential mO during the preparation procedure, and they are represented by their
corresponding STM profile.
The manifold Mn oxide phase diagram comprises nine different nanophases
which are characterized by specific windows in the parameter space of the thermo-
dynamic variables, temperature (in the total range 600–800 K) and oxygen pressure
(in the total range 5  106–5  108 mbar). We distinguish
1. The oxygen-rich regime (5  106 mbar > pO2 < 5  107 mbar):
(a) Two hexagonal phases (HEX-I and HEX-II), which are both obtained at high
O2 pressures.
214 C. Franchini and F. Allegretti

Fig. 10.1 Experimental schematic phase stability diagram of the interfacial Mn oxides, presented
as a function of the oxygen pressure pO2 and of the oxygen chemical potential mO . The nominal
coverage of Mn on Pd(100) is: 0.75 ML. From [50] with permission

2. The oxygen-intermediate regime (5  107 mbar > pO2 < 1  107 mbar):
(a) A c(4  2) structure and a stripe phase described as a uniaxially compressed
c(4  2), which are both stabilized at intermediate O2 pressures.
(b) Two structures which were called chevrons (CHEV-I and CHEV-II),
because of their STM appearance.
3. The oxygen-poor regime (1  107 mbar > pO2 < 5  108 mbar):
(a) Two reduced phases with complex structures, named waves and labyrinth.
(b) At the most reducing conditions, a third hexagonal phase (HEX-III), com-
mensurate with the Pd(100) substrate (2) along one of the two <011>
directions.
The extraordinary architectural flexibility of the interfacial Mn oxides on Pd
(100) and the electronic properties of the novel 2D phases can be rationalized
and understood through a synergic combination of experimental (STM, LEED,
HREELS, and XPS) and computational (DFT and hybrid functionals) techniques.
Indeed, in modern surface science, theoretical calculations have become an effi-
cient complementing tool to experimental observations. Theoretical models based
on educated guesses of possible structures can be tested and directly compared with
the experiments in order to clarify the structural aspects and provide an atomistic
interpretation of the measured properties. In the present case, the joint experi-
mental–theoretical analysis reveals that the two oxygen-rich phases (HEX-I
and HEX-II) can be described in terms of O–Mn–O trilayers with MnO(111)-
like structures (see Sect. 10.2.1.1), whereas the intermediate oxygen regime
[c(4  2) and chevrons] is based on a compressed MnO(100)-like monolayer
model (see Sect. 10.2.1.2). At low oxygen chemical potentials the waves and
labyrinth structures show in STM very complex unit cells whose link to the other
phases is less clear but seems still to be related to a MnO(100)-like wetting layer,
while the additional hexagonal phase HEX-III is of uncertain attribution (see
Sect. 10.2.1.3).
10 Low Dimensionality and Epitaxial Stabilization. . . 215

Fig. 10.2 HREELS phonon


spectra of the four Mn oxide
submonolayer phases at high
and intermediate oxygen
partial pressures. From the
top to the bottom: HEX-I,
HEX-II, c(4  2), and
CHEV-I. The statistical
uncertainty in the peak
position is 0.5 meV. For
every Mn oxide phase,
measurements performed on
samples freshly prepared in
different days agree within
1 meV. Adopted from [50]

The recurrent MnO(100)- and MnO(111)-like structural features are reflected in


the phonon-loss spectrum displayed in Fig. 10.2. As expected, the common building
blocks shared between the different oxide phases result in similar phonon losses.
All spectra exhibit a clear single peak structure. The phonon loss is centered at
around 70 meV for the two HEX-I and HEX-II phases and shifts down to
44–45 meV for the c(4  2) and CHEV-I phases. Interestingly, the phonon spectra
(not shown) of the CHEV-II, waves, and HEX-III phases, which are obtained by
further lowering the O chemical potential, are also characterized by a single peak at
43–45 meV. These findings provide a clear indication of two distinct regimes, a
MnO(111)-like regime comprising phases with a higher energy phonon loss
(70 meV) and an MnO(100)-like regime with a single phonon loss around 44 meV.

10.2.1.1 MnO(111)-Like Phases (Oxygen-Rich Regime)

In this pressure regime two Mn oxide phases, HEX-I and HEX-II, have been
detected, which are characterized by hexagonal or quasi-hexagonal (i.e., distorted)
symmetry linked to that of MnO(111), and which display a phonon spectrum with a
single loss peak centered at around 70 meV.

HEX-I

For the sake of clarity, it is instructive to recall the structure of MnO(111), which is
made out of alternating O and Mn hexagonal layers with ABC stacking sequence,
each layer having in-plane lattice constant aMnO ¼ 3:14 Å (see Fig. 10.3a).
Fig. 10.3 (a) Real and (b) reciprocal lattice for a (1  1) MnO(111)-like hexagonal structure on a Pd
(100) surface. The lattice parameter of the overlayer is assumed to be the measured bulk value
aMnO ¼ 3:14Å, whereas apd ¼ 2:75 Å. In (a) only one of the two symmetry domains is reported for
clarity. (c) SPA-LEED two-dimensional pattern measured at E ¼ 90 eV, for the undistorted HEX-I
phase. (d) LEED pattern of the distorted HEX-I phase, recorded at E ¼ 96 eV. (e) Real and (f)
reciprocal lattice associated with the distorted hexagonal MnO(111)-like phase. Only one symmetry
domain is shown in (e) for clarity, for which the lattice parameter in the direction of the distortion
([011]pd) is bMnOx ¼ 2:94 Å. In total, four symmetry domains contribute to the reciprocal lattice in (f):
two are obtained from (e) with aMnO at either +60 or 60 from bMnOx , and the other two are obtained
by rotating aMnO by 90 relative to the substrate mesh. (g) Hard sphere model simulating in the real
space the moiré pattern that originates from the interference of the quasi-hexagonal oxide lattice of
panel (e) with the square mesh of the Pd(100) substrate. Figure adapted from [25] 51 with permission
10 Low Dimensionality and Epitaxial Stabilization. . . 217

A perfect MnO(111) film would give rise to a hexagonal reciprocal space


pattern with lattice parameter 4p=3aMnO ¼ 2:30Å1. The latter value is very
close to the reciprocal space lattice parameter of Pd(100) (2p=apd ¼ 2:28Å1).
The resulting LEED pattern averaged over two hexagonal domains rotated by 90 ,
which account for the different symmetry of substrate and overlayer, is a circular
array of 12 extra spots superimposed on the (10) spots of the substrate, as
illustrated in Fig. 10.3b. Although such a LEED pattern has been indeed observed
in the experiments, as demonstrated in Fig. 10.3c, the structural details sensibly
depend on the preparation conditions. It appears that this undistorted hexagonal
structure is favored at slightly lower coverage (0.5–0.6 ML) or after repeated
oxidation cycles. In contrast, at 0.75 ML and after a single oxidation step at high
pressure the LEED pattern reported in Fig. 10.3d is reproducibly observed: it still
reflects a hexagonal-like lattice, but the characteristic elongation of the overlayer
spots indicates a clear distortion of the “perfect” (1  1) structure into a distorted
hexagonal structure. With respect to the perfect hexagonal lattice, the distorted
LEED pattern is distinctive of an incommensurate overlayer with a contracted
lattice vector bMnOx ¼ 2:94 Å along [011]pd (Fig. 10.3e). The simulated reciprocal
space pattern of the distorted structure, shown in Fig. 10.3f, reproduces very well
the experimental LEED pattern (Fig. 10.3d). Interestingly, the comparison
between the undistorted and distorted real space lattices of Fig. 10.3a, e suggests
that the better matching between overlayer and substrate might be the driving
force for the distortion: the initial mismatch of 14% along the [011]pd direction is
in fact reduced to ðbMnOx  apd Þ=apd ¼ 7% upon distortion. This better epitaxial
relationship is likely to be at the origin of the increased stability and reproducibil-
ity of the distorted HEX-I phase.
In passing, we note that the distorted model explains also the moiré pattern
observed experimentally in the STM images (Fig. 10.4c). The latter results from
the interference of the quasi-hexagonal Mn oxide overlayer with the square mesh of
the Pd(100) surface. Taking a ratio bMnOx =apd  16=15, a moiré pattern can be
generated with a geometrical model that displays modulations in the form of
broad lines inclined with respect to the [011]pd direction and with an average
periodicity of about 22 Å parallel to [011]pd (see Fig. 10.3g). In this model, only
one type of overlayer atoms is considered for simplicity, and a different color
gradation (white, gray, black) is used to emphasize the different lateral registry of
these overlayer atoms relative to the underlying matrix of substrate atoms. For on-
top/bridge location, the white/black color of the overlayer atoms reflects the different
height above the surface. The remarkable similarity between the modeled moiré
pattern (Fig. 10.3g) and the experimental STM image (Fig. 10.4c) suggests implicitly
that the HEX-I phase may consist of alternately stacked layers with quasi-hexagonal
symmetry containing only one single atom species, as it is realized in the MnO(111)
structure, which is therefore the natural toy-model for the computational analysis.
The HEX-I model used in the DFT analysis is constructed by cleaving the MnO
structure perpendicular to the [111] direction to form the smallest block containing
1 ML of Mn atoms in the O–Mn–O stacking, with a formal stoichiometry of MnO2.
A rigid sphere sketch of the MnO2 trilayer is given in Fig. 10.4a, along with the
218 C. Franchini and F. Allegretti

Fig. 10.4 (a–b) Top view of the geometrical models for undistorted (a) and distorted (b) HEX-I
MnO(111)-like trilayer phases: red spheres: O atoms; light gray spheres: Mn atoms. Dashed lines
delimit the 2D unit cell. (c–e) Experimental (c, d) and simulated (e) STM images of the HEX-I
phase; Experimental images: sample bias U ¼ +0.5 V (c), +0.6 V (d); tunneling current I ¼ 0.13
nA (c), 0.15 nA (d). The simulated STM image has been calculated considering tunneling into
empty states between 0 and +0.5 eV. Taken from [25] with permission

corresponding distorted trilayer (Fig. 10.4b) obtained by shrinking the lattice


constant b by 7%. Due to the free-standing (i.e., unsupported trilayers) computational
setup, which is unavoidable because of the incommensurate registry of the substrate/
overlayer system, the optimized 2D lattice constants are significantly smaller than
the measured ones. Energetic considerations on the computed relative stability of the
two HEX-I models reveal that although the perfect hexagonal structure is the most
favorable, only a low energy cost of 70 meV/cell is required to convert it into a
distorted hexagonal structure. Therefore, the interaction with the substrate, neglected
in the unsupported trilayer model, might easily provide the energy gain to reverse this
energy sequence through the better matching discussed above.
The soundness of the MnO2 trilayer model is confirmed by a direct comparison
of the HREELS and STM data with the computed counterparts. A comparison
between the measured and simulated atomically resolved STM images is shown in
Fig. 10.4d, e. Despite the low resolution of the experimental map the agreement is
acceptable, in particular in that the arrangement of large bright protrusions is fairly
reproduced; moreover, the theoretical data allow assigning the bright spots to the
topmost oxygen atoms. Further convincing support for the proposed model comes
from the comparison between the experimental (70.5 meV) and computed
(73 meV) distinct single dipole-active phonon mode, which is associated to the
vertical antiphase vibration of O and Mn layers [25]. Finally, it is also worth noting
that this trilayer model is well compatible with the O 1s core level photoemission
data reported in [50], which show a strong component at 529.1 eV attributable to the
surface O layer and a weaker line chemically shifted to higher binding energy
presumably related to the O layer at the interface with the Pd(100) substrate.
Indeed, a model consisting of a MnO(111)-like bilayer, with only one layer of
10 Low Dimensionality and Epitaxial Stabilization. . . 219

Fig. 10.5 (a) LEED pattern of the HEX-II phase (E ¼ 80 eV). (b) Top view of the Mn4O7-a HEX-
II model: red spheres: O atoms; light gray spheres: Mn atoms. Dashed lines delimit the 2D unit
cell, whereas triangles highlight the octopolar Mn–O3 pyramids and the O3 trimers of the a
termination. (c) Experimental and simulated STM images of the hexagonal a HEX-II phase;
sample bias U ¼ +0.75 V; tunneling current I ¼0.1 nA. The simulated STM image has been
calculated considering tunneling into empty states between 0 and +0.75 eV. Unit cells are
indicated by full lines. Figure adapted from [25] 51 with permission

chemically equivalent O atoms either at the interface or at the surface to yield a


formal MnO stoichiometry, is not only incompatible with the photoemission data
but is also shown to be much less stable than the corresponding trilayer model in a
broad window of the O chemical potential [25].

HEX-II

The O–Mn–O MnO(111) trilayer model is the crucial key for the structural under-
standing of the HEX-II phase, which is obtained from the HEX-I phase upon
reduction of the oxygen partial pressure. Unlike HEX-I, the STM appearance of
HEX-II is characterized by a well-ordered hexagonal array of triangularly shaped
maxima (Fig. 10.5c), and the lattice constant derived from the STM and LEED
measurements (Fig. 10.5a) is a  6.0 Å, corresponding to a compressed (2  2)
reconstruction of the MnO(111) termination. The DFT analysis reveals that the
so-called O-terminated a reconstruction displays the lowest surface energy in a
wide range of oxygen partial pressures [25]. The a structure is obtained by
removing one topmost oxygen atom per (2  2) unit cell from the HEX-I structure,
resulting in a O3  Mn4  O4 trilayer with the in-plane hexagonal lattice constant
of 5.95 Å, as shown in Fig. 10.5b.
Compelling evidence for the O-terminated a model comes from the comparison
between the simulated STM and HREELS data and the corresponding experimental
results. The hexagonal pattern of triangular bright protrusions in the experimental
STM picture is well reproduced by the simulated STM image (Fig. 10.5c), and the
triangular protrusions are therefore ascribed to the oxygen trimers outlined in the
sketch of Fig. 10.5b. As for the HREELS spectrum, in analogy with the precursor
220 C. Franchini and F. Allegretti

HEX-I phase, the HEX-II a phase displays a single phonon loss peak centered at
69.5 meV (Fig. 10.5b), characteristic of the MnO(111)-like structures. The simula-
tion reproduces well the position of the vibrational peak (within 1–2 eV) and
ascribes it to the concerted movement of the surface oxygen trimers against the
manganese atoms underneath.

10.2.1.2 MnO(100)-Like Phases (Intermediate Oxygen Regime)

At intermediate oxygen pressures, four different MnxOy phases have been observed.
At the high-pressure end of this regime two phases coexist (see Fig. 10.1): a quasi-
hexagonal stripe phase characterized by bands of parallel stripes separated by
5.5 Å (2  apd ) and the c(4  2) structure, commensurate with the Pd(100)
substrate as expressed by the transformation matrix M ¼ [2, 1/2, 1]. At the
low-pressure end, instead, the c(4  2) phase is often observed in coexistence
with the so-called chevron structures, chevron I and chevron II, corresponding to
the transformation matrices M ¼ [2, 1/3, 2] and [2, 1/5, 3]. The phase
boundaries between the two chevron structures and the c(4  2) structure are
smooth and continuous, suggesting that from the structural point of view these
phases are closely related. Indeed, the unit cells of the chevron structures can
be generated from the c(4  2) structure by simply extending the b2 unit cell
vectors to adjacent antiphase positions as illustrated in the geometric models of
Fig. 10.6b. In the following, the structural analogies of these phases are rationalized
in terms of MnO(100)-like monolayers.

cð4  2Þ  Mn3 O4 =Pdð100Þ

The LEED pattern of the c(4  2) structure, the central phase in this pressure
regime, is displayed in Fig. 10.6a and compared with the simulated reciprocal space
pattern of (c), which corresponds to the real space periodicity given in (b). The
structural model that explains the c(4  2) MnxOy phase is depicted in Fig. 10.7.
It is constructed by placing a compressed MnO(100) monolayer on top of Pd(100)
and by creating a regular rhombic array of Mn vacancies, in close analogy with the
Ni/Co oxide monolayer phases, c(4  2)-Ni3O4/Pd(100) and c(4  2)-Co3O4/Pd
(100), respectively, obtained upon reactive evaporation of nickel/cobalt on Pd(100)
[2, 5]. It is, however, to be noted that the lattice mismatch between MnO (14%) and
the substrate is much larger than that for NiO (7%) and CoO (9.5%). The lateral
compression required to fit a single layer of the bulk MnO(100) structure onto the
Pd(100) substrate and the consequent formation of the Mn vacancies dominate the
structural rearrangement and determine the occurrence of two distinct types of Mn
atoms with different environment, as illustrated in Fig. 10.7a: the Mn1 atoms are
localized between two nearest neighbor vacancies, while the Mn2 atoms form zig-
zag Mn–Mn chains along the [01 1] direction. Large relaxations take place in the
Mn3O4 layer, which is found to be well separated from the substrate by z  2.3 Å,
approximately 20% larger than the Pd bulk interlayer distance. The lateral
10 Low Dimensionality and Epitaxial Stabilization. . . 221

Fig. 10.6 (a) LEED patterns, (b) real lattice models, and (c) reciprocal lattice patterns of the
c(4  2) (left panel), chevron I (middle panel) and chevron II (right panel) structures. The
respective LEED energies in (a) are 116, 108, and 60 eV. From [50]

compression is accompanied by a significant rumpling of the monolayer, yielding a


corrugation of 0.23 Å, whereas the Mn vacancy breaks the local Mn–O bonds and
induces a planar outward relaxation dO of the oxygen atoms, which move closer to
Mn1, resulting in the waving short-ranged ordering highlighted in Fig. 10.7a.
The comparison with the experimental data, given in Fig. 10.7b, c, validates the
proposed c(4  2)-Mn3O4 structural model. The measured STM image (inset of b
in Fig. 10.7) is characterized by a rhombic arrangement of bright spots and dark
depressions connected by weak segments. These features are very well reproduced
by the simulated STM picture and can be interpreted in the following way: the black
depressions reflect the network of Mn vacancies, the single bright spots can be
assigned to the Mn1 species, whereas the light segments correspond to the zig-zag
Mn2–Mn2 chains embedded in the regular array of oxygen atoms, which are not
222 C. Franchini and F. Allegretti

Fig. 10.7 (a) Optimized structural model for the c(4  2)-Mn3O4/Pd(100) system. The Pd(100)
substrate is displayed with light gray large spheres, whereas Mn and O atoms are depicted with
small blue and pink spheres. Two distinct Mn species are distinguishable: Mn1, sandwiched
between two vacancies, and two Mn2 atoms forming the zig-zag arrangements highlighted by
the horizontal lines. Atoms surrounding the Mn vacancies experience considerable strain
(indicated by arrows). (b) Comparison between simulated and experimental (inset) STM images
for the c(4  2)-Mn3O4/Pd(100) phase. Experimental parameters: sample bias U ¼ +0.8 V,
tunneling current I ¼ 0.25 nA.(c) Comparison between the measured HREELS phonon value
(vertical bar) and PBE (dashed line) and HSE (full line) predicted dipole active modes. Adapted
from [25, 26]

seen in the STM image. Further support for the c(4  2)-Mn3O4 model is provided
by the comparison of the calculated phonon spectrum with the HREELS
measurements (Fig. 10.7c). The single peak structure centered at 43 meV
revealed by HREELS is very well reproduced by the hybrid functional calculations
(standard PBE, instead, underestimates the phonon energy by 7 meV). The
diagonalization of the dynamical matrix enables to assign this specific phonon
mode to the collective vertical vibration of the oxygen sublattice against the Mn
sublattice, a vibrational mode typical of the ideal MnO(100) surface [25, 37].

Chevron Structures

The chevron structures are adjacent to the c(4  2) phase in the phase stability
diagram and are obtained by lowering the O2 partial pressure. These phases, the
name of which originates from the “chevron-like” STM motifs, can coexist locally
with the c(4  2) phase, as shown in Fig. 10.8a. Due to the close link with the
10 Low Dimensionality and Epitaxial Stabilization. . . 223

Fig. 10.8 (a) STM image showing the characteristic appearance of the chevron structures and the
local coexistence of the c(4  2), chevron I and chevron II phases (140 Å  100 Å; U ¼ +1.0 V;
I ¼ 0.2 nA). (b) Top view of the geometrical model for the Pd(100) supported chevron I phase: red
spheres: O atoms; light gray spheres: Mn atoms; green small spheres: Pd atoms underneath. Dashed
lines delimit the 2D unit cell and circles denote the position of the vacancies. (c) Experimental and
simulated STM images of the chevron I phase; sample bias U ¼ +1.0 V; tunneling current I ¼ 0.13
nA. The simulated STM image has been calculated considering tunneling into empty states between
0 and +1.0 eV. Unit cells are indicated by full lines. Figure adapted from [25]

c(4  2) phase, presented in a pictorial fashion in Fig. 10.6b, the chevron structures
can be rationalized in terms of structural models inspired by the c(4  2) structural
features and by invoking the concept of vacancy propagation. As discussed above,
the periodicity of the c(4  2) phase is given by the Mn vacancies distributed in a
regular array with the rhombic unit cell. Therefore, within this picture and based on
Fig. 10.6b, the chevron I and II structures may be naturally obtained from the
c(4  2) by propagating a Mn vacancy to a neighboring lattice site. Accordingly,
the stoichiometry becomes Mn6 O7 and Mn10 O11 , respectively, i.e., closer to that of
MnO. Compared to the c(4  2) phase, the chevron structures exhibit a much
higher degree of long-range order, with large defect-free areas extending over
typical distances of 200 Å and interrupted by ordered domain boundaries, which
convey the characteristic “chevron-like” appearance (see Fig. 10.8a).
The structural model for the chevron I structure, obtained by removing 14% of Mn
atoms from the compressed MnO(100) monolayer, is depicted in Fig. 10.8b. The
geometrical optimization again yields an oxygen-terminated surface with a corruga-
tion of 0.33 Å, and in analogy with the cð4  2Þ - Mn3 O4 structure a moderate
displacement (0.13 Å) of Mn atoms toward the vacancy is found, indicated by the
arrows in Fig. 10.8b. Figure 10.8c shows a comparison between the experimental and
simulated STM images. Although the experimental resolution does not allow a
detailed atomic comparison between theory and experiment, the overall agreement
is satisfactory. The dark circular depressions in the images represent the position of
the Mn vacancies and determine the 2D unit cell of this phase, as outlined by the full
lines. The bright protrusions are clearly due to manganese atoms, suggesting that
electronic effects contribute predominantly to the STM topography. Two distinct
bright features are observed in the experimental picture, one large spot in the center of
the unit cell and weaker flecks at the edges. The simulated STM image together with
the optimized geometrical data allows the atomically resolved identification of these
two kinds of features. The manganese sublayer is itself slightly corrugated due to a
small upward shift (0.05 Å) of the Mn atoms lying closer to the vacancy and aligned
224 C. Franchini and F. Allegretti

in the [01 1] direction. These atoms, indicated by filled dots in Fig. 10.8b, are
responsible for the two large bright spots in the simulated STM image, which
merge together in the experimental image thereby giving rise to the wavy lines of
bright protrusions. The remaining spots (four per unit cell) arise from the lower
manganese atoms. In the experimental STM image their intensity partially mixes up
with the topmost Mn spots thus contributing to the large bright feature centered in the
cell and partially providing the weaker spots along the edges. Finally, the DFT-based
analysis establishes that the Mn6 O7 chevron I structure is characterized by a single
dipole-active phonon mode at about 40 meV, in good agreement with the HREELS
measurements (phonon loss at 44.5 meV). This peak reflects the vibrational finger-
print of the parent MnO(100) not only in terms of the phonon loss energy, very
similar to the cð4  2Þ - Mn3 O4 value (43.5 meV, see Fig. 10.2), but also in terms of
the atomic displacements producing this vibration, which are associated to the
vertical opposite movement of manganese and oxygen atoms.

10.2.1.3 The Reduced Phases (Oxygen-Poor Regime)

There is a smooth transition from the c(4  2) and chevron structures of the
intermediate oxygen pressure regime to the structures in the oxygen-poor regime
denoted as labyrinth and waves. These two phases are named according to their
STM pattern shown in Fig. 10.1. Although the complexity of these structures is a
major obstacle for computational machinery, some fundamental properties can be
deduced from a careful analysis of STM and LEED data (not shown) [50].
The labyrinth structure is described by the lattice vectors b1 ¼ 13.5  0.5 Å and
b2 ¼ 11.0  0.5 Å, y ¼ 85  2 (y: angle between lattice vectors b1 and b2), the
unit cell being rotated by a ¼ 21  2 with respect to the Pd [011] direction. The
unit cell of the waves structure is characterized by b1 ¼ 13.8  0.2 Å, b2 ¼ 34
 1 Å, y ¼ 90  2 , which corresponds approximately to a (5  12) superstruc-
ture with respect to the Pd(100) substrate. On the basis of the measured phonon loss
spectrum and of the occurrence of ordered phase boundaries with the c(4  2) and
chevron structures, these two reduced phases with complex appearance are tenta-
tively assigned to MnO(100)-like wetting layers.
Finally, at the most reducing conditions, a different structure is observed, the
HEX-III structure, shown on the left side of the phase stability diagram of Fig.p10.1.
ffiffiffi
The HEX-III structure ispdescribed
ffiffiffi ffi the transformation matrix M ¼ ½0; 2= 3; 1
pffiffiby
(or equivalently
pffiffiffi p Mffiffiffi ¼ ½ 3 ; 1= 3; 1) with respect to Pd(100), but corresponds
also to a ð 3  3ÞR30 superstructure with respect to a MnO(111) surface. The
last observation suggests that this phase might be related to a MnO(111)-like
model, possibly a bilayer with one-third of the surface atoms removed. On the
other hand, the matrix notation shows that the periodicity of the HEX-III phase can
be also obtained upon a unidirectional 14% compression of the c(4  2) structure.
As a result, a (7  2) coincidence lattice relative to the Pd(100) surface is observed.
This interpretation would be compatible with the very similar phonon-loss spectra
of the HEX-III and c(4  2) phases, and with certain features of the STM
10 Low Dimensionality and Epitaxial Stabilization. . . 225

appearance at particular bias voltages (indeed, the c(4  2) periodicity can itself
be regarded as due to a distorted hexagonal lattice of vacancies). This further
compression could hypothetically produce a more pronounced rumpling of the
corrugated monolayer, possibly resulting in a well-defined bilayer. The precise
assignment of the HEX-III requires detailed structural investigations, and to this
purpose DFT calculations are underway. It is, in any case, possible that the HEX-III
structure is an interfacial phase that mediates the growth of MnO(111)-oriented
films with hexagonal symmetry, which will be discussed in the next section.

10.2.2 High Coverage Regime

The reactive evaporation of 20–30 ML Mn on Pd(100) at moderate temperature


(620 K) and an oxygen pressure in the range of 2  107–5  107 mbar leads to
the stabilization of (100)-oriented MnO films. The formation of well-ordered MnO
(100) is evident from the spot profile analysis LEED (SPA-LEED) image presented
in Fig. 10.9a, which displays sharp reflexes arranged in a square pattern and
superimposed on a very low background. The surface lattice constant determined
from the separation of the LEED spots measures 3.14  0.03 Å, identical to the
(100) in-plane lattice parameter of bulk MnO. The frequency modulation atomic
force microscopy (FM-AFM) image (Fig. 10.9b) shows that the MnO(100) surface is
atomically flat and consists of terraces with lateral dimensions up to 500 Å [3]. The
MnO stoichiometry of the oxide film has been confirmed by the characteristic
fingerprint provided by core level photoemission spectroscopy (see [3] for details)
and by a comparison between the valence band photoemission spectrum and the
calculated [14] (PBE + Hubbard U method) density of states of MnO given in
Fig. 10.9c, which provides further evidence of the good quality of the MnO(100)
multilayers. The main spectral features, namely the triple peak structure with maxima
at 2.2, 3.5, and 4.9 eV, are reasonably well reproduced by the theory, although the
calculated peak with highest binding energy is upshifted by about 1 eV with respect
to the experimental one. Finally, in Fig. 10.9d, the HREELS phonon spectrum
exhibits a main phonon-loss peak at 65 meV and a weaker structure at 48 meV,
which can be connected with the bulk optically allowed phonons at 33.6–36.4 meV
[19] and 62 meV [36]. This finding is in line with the single peak structure at 71 meV
detected in MnO(100) crystals [48]. The comparison of the phonon spectrum of the
MnO(100) thick films with those of the MnO(100)-related monolayer structures
discussed previously [c(4  2)-Mn3O4 and chevrons-Mn6O7(I)/Mn10O11(II)] clearly
indicates that in the monolayer regime the dipole-active phonon mode is strongly
redshifted by about 20 meV. A similar thickness dependence of the main phonon
frequency has been also reported for the MnO(100)/Pt(111) system [37].
We have shown above that the number of oxidation states available to manga-
nese in the solid state is reflected in the complex stoichiometric and architectural
flexibility in the interface-dependent monolayer regime. Similarly, the manganese
oxide surface can support a range of manganese oxidation states and oxide phases
226 C. Franchini and F. Allegretti

Fig. 10.9 30 ML MnO(100) on Pd(100) [3, 14]. (a) SPA-LEED pattern, recorded with an electron
energy of 90 eV. (b) FM-AFM image of the MnO(100) surface (size 2,500  2,500 Å2, detuning
df ¼ 5 Hz, bias U ¼ 0.2 V). (c) Valence band spectra excited by a photon energy of 120 eV
(solid line) compared with the calculated density of states (dashed line). (d) HREELS spectrum of
the MnO(100) film. Figure adapted from [3]

because of the similar electronic character of the simplest Mn oxides (MnO,


Mn3O4, and Mn2O3), which ultimately differ by a diverse electronic population
and splitting of the partially filled d shell. Langell et al. have shown that by proper
adjustment of temperature and oxidizing/reducing conditions it is possible to select
the dominant oxide forms, MnO, Mn2O3, or Mn3O4, present at the MnO(100)
surface [48]. Within this context, it is not unexpected that under suitable prepara-
tion conditions (5  106–2  105 mbar O2 and elevated temperature) the Pd
(100)-supported MnO(100) structure can be converted into the Mn3O4(001) surface
[15]. It is nonetheless more surprising that either by annealing the MnO(100) films
at an elevated temperature (770 K) in vacuo or by reactive evaporation at a
relatively low pressure (<1  107 mbar) the transformation of the MnO(100)
structure into a MnO(111) surface occurs [3]. The discussion of these two phase
transitions will be addressed in the following two sections.
10 Low Dimensionality and Epitaxial Stabilization. . . 227

Fig. 10.10 Comparison between the initial 20-ML thick MnO(001) film on Pd(100) and the
Mn3 O4 phase prepared by oxidation of MnO(001). (a) Mn 3s core level spectra excited by a photon
energy of 180 eV. The corresponding exchange spin splitting is indicated. (b) Mn L2,3-edge X-ray
absorption spectroscopy (XAS) spectra recorded in total yield mode. (c) HREELS spectra. The
energies of the main energy losses are reported. (d) 2D SPA-LEED pattern of the 20 ML thick
MnO(001) film, and (e) of the Mn3 O4 ð001Þ phase. Both diffraction patterns were recorded with an
electron energy of 125 eV. (f) Corresponding 1D SPA-LEED profiles centered on the (00) reflex
and recorded along the [110] substrate direction at Ep ¼ 125 eV. Adapted from [15]

10.2.2.1 Formation of Mn3O4 on MnO(001)

Although, like in other spinel-like structured compounds, Mn3 O4 terminations are


formally considered to be unstable because of the uncompensated electrostatic poten-
tial perpendicular to the surface, it is reported that the cleavage of Mn3 O4 leads
distinctly to a (001)-oriented surface [17]. Indeed, Noguera has rationalized that this
formal instability is an artificial consequence of the oversimplistic ionic model and can
be removed by a number of natural mechanisms involving structural and electronic
reconstructions [61]. The epitaxial growth of Mn3 O4 ð001Þ has been achieved on MgO
(001) [34, 35] and circumstantial evidence indicates that Mn3 O4 -like surfaces can be
obtained by mild oxidation of MnO(001) single crystals [48]. In this section, we
discuss the formation of Mn3 O4 ð001Þ surfaces on Pd(100)-supported MnO(100)
[denoted, equivalently, as MnO(001)/Pd(100)] by high-temperature oxidation at
intermediate and high molecular oxygen pressures.
The Mn 3s core level spectra shown in Fig. 10.10a indicate that exposing for
20 min Pd(100)-supported MnO(001) layers (20 ML) to molecular oxygen at a
pressure of 2  105 mbar with the sample kept at 770 K induces a marked
228 C. Franchini and F. Allegretti

decrease of the 3s splitting down to 5.4 eV, a value that is in line with the measured
exchange splitting for bulk Mn3 O4 (5.3–5.5 eV) and with the expected change in
the electronic configuration of the d shell. Further experimental evidence for the
formation of Mn3O4-like films upon oxidation is provided by the Mn L2,3-edge
X-ray absorption data reported in Fig. 10.10b: as evident, the characteristic finger-
print of MnO (bottom spectrum) is completely lost upon oxidation (top spectrum),
and the intensity ratio of the L3 to L2 lines is reduced from 3.8 to 2.6. Since the
I(L3)/I(L2) ratio can be related to the occupancy of the 3d orbitals in the Mn ion
[47], the overall changes are again consistent with the (partial) presence of Mn ions
with 3d 4 configuration. In particular, the profile of the L3 peak is considerably
broadened and exhibits a triplet fine structure that is a distinctive signature of
Mn3 O4 [28]. The Mn3 O4 stoichiometry of the oxidized film is also confirmed by
the HREELS phonon spectrum (Fig. 10.10c). While the MnO(001) film is
characterized by a Fuchs–Kliewer phonon-loss peak at 65 meV and a much
weaker structure at 48 meV (the peak at 130 meV is a double loss of the
65 meV peak) as noted above, the oxidized film is dominated by a new phonon
loss at 83 meV, very close to the analogous vibrational peak at 650–654 cm1
(81 meV) that has been detected by Raman spectroscopy measurements from
mineral and synthetic hausmannite Mn3 O4 crystals [16, 42].
The issue of the surface symmetry and growth direction of the Mn3 O4 layers is
successfully addressed by SPA-LEED measurements, see Fig. 10.10d, e. Upon
oxidation, the sharp (1  1) square pattern characteristic of well-ordered MnO
(001) evolves into a weaker (2  2)-like pattern with broader spots, which
preserves the square symmetry of the unit cell but with a larger periodicity in
the real space. The line scan along the <110> high symmetry direction
(Fig. 10.10f) indicates that the spots appearing at roughly  2 positions are the
first-order spots of Mn3 O4 , which are located at 53  1% of the MnO(001)
Brillouin zone and in real space describe a square lattice with a unit cell parameter
of 5.9  0.1 Å. From the geometry of the bulk hausmannite spinel structure it is
inferred that the Mn3 O4 film is (001)-oriented, with the a and b sides of the unit
cell aligned along the <110> directions of the MnO(001) substrate. The data also
demonstrate that due to the 9% lattice mismatch with the 2 supercell of the
underlying MnO(001) the supported Mn3 O4 phase grows strained relative to the
bulk phase (5.76 Å).
The DFT-based study of this system complements the experimental information,
giving additional insight into the atomic-level characterization of the isolated Mn3 O4
surfaces and of the Mn3 O4 (001)/MnO(001) interface [15]. In line with the experi-
mental observations, the (001) orientation of Mn3 O4 is found to be the energetically
most favorable, with a surface energy more than 40 meV/Å2 lower than that of the
(110) and (100) surfaces. Among the two possible as-cleaved (001) terminations, the
mixed oxygen and manganese surface with stoichiometry Mn2O4 is more stable than
the manganese-terminated one by about 20 meV/Å2. The structural modifications of
the Mn2O4-terminated Mn3 O4 ð001Þ crystal induced by the presence of the surface
are localized within the topmost two layers: they can be described in terms of a
buckling (0.2 Å) of the outermost species and of the changes in the interlayer
10 Low Dimensionality and Epitaxial Stabilization. . . 229

Fig. 10.11 Model structure for the interface between Mn3 O4 ð001Þ and MnOð001Þ and calculated
relative stability (PBE). Black and red circles indicate Mn and O atoms, respectively. From [15]

Table 10.1 Calculated interlayer distances (dij) for the Mn3O4(001)/MnO(001) optimized struc-
ture, expressed either in Å (first row) or in percent with respect to the bulk interlayer spacing of the
strained Mn3O4 structure (second row). dI1 and dI2 indicate the distances between the two MnO
(001) layers and the first two Mn3O4(001) layers at the interface, respectively, whereas dI is the
interfacial distance between the MnO(001) and Mn3O4(001) contact planes
d12 d23 d34 d45 d56 d67 d78 dI2 dI dI1
0.74 1.26 0.95 0.95 1.02 1.00 1.10 1.02 2.19 2.27
29.5 +20.0 0.1 0.1 – 0.1 +0.1 0.0 0.0 –

distances with respect to the bulk geometry (in particular, there is a very pronounced
contraction of 26% of the first interlayer distance, partially compensated by an
appreciable expansion of 9% of the second interlayer distance) [15].
Based on the two distinct Mn3 O4 ð001Þ terminations and considering that the
MnO(001) surface possesses a single termination with planar unit containing an
equal number of manganese and oxygen atoms, two models of the interface emerge
naturally: (1) interface A, which is constructed by placing the Mn-terminated (Mn-t)
Mn3 O4 ð001Þ surface on the MnO(001) substrate and (2) interface B, which directly
joins the MnO surface layer with the Mn2 O4  t layer of Mn3 O4 ð001Þ. Different
junctions can be constructed by changing the registry of the interfacial Mn3 O4 ð001Þ
slice. The resulting models are sketched in Fig. 10.11: submodels A1 and B1
correspond to an atop site registry of the Mn3 O4 ð001Þ interfacial manganese atoms
on the MnO(001Þ oxygen atoms underneath, whereas submodels A2 and B2 refer to
the corresponding bridge site MnMn3 O4  OMnO registry. The calculated relative
stability of the explored models, listed in Fig. 10.11, indicates that interface B1 has
the lowest energy and represents by far the most favorable model.
The structural characteristics of the Mn3 O4 (001)/MnO(001) interface are
summarized in Table 10.1 (optimized geometry) and Fig. 10.12 (layer by-layer
decomposed charge density plot) [15]. The structural optimization was performed
230 C. Franchini and F. Allegretti

Fig. 10.12 Charge density plots for the Mn3 O4 (001)/MnO(001) interface: (a) layer-by-layer
decomposition of the charge density along the [001] direction. The dashed lines A, B, C, and D
(only A and D labels are given in the figure) are 2D projections of the (001) planes displayed in
(b) and (c) showing vertical cuts of the charge density along the [100] and [010] directions,
respectively. Dark gray (red) and light blue circles indicate the position of oxygen and manganese
atoms, respectively. The arrows highlight the most significant internal structural relaxations. The
yellow line marks the separation between the Mn3 O4 and MnO components, whereas the blue line
indicates the unrelaxed starting position of the topmost Mn3 O4 ð001Þ layer. From [15]

using a (2  2) unit cell relative to the MnO lattice constant and preserving the
bulk internal geometry in the three layers at the bottom of the slab. A large
vertical contraction of 20% of the Mn3 O4 region of the slab is obtained, which
compensates for the in-plane tensile strain. As a consequence of the latter strain,
the distance between adjacent surface oxygen rows increases by 0.7 Å, in contrast
with the contraction observed in the unstrained Mn3 O4 ð001Þ surface.
In terms of interlayer distances it is found that the relaxations are well localized
within the first three layers and do not affect the rest of the slab. By referring the
changes of the interlayer distances to the bulk interlayer spacing corresponding to
the bulk Mn3O4 strained structure (i.e., with a ¼ aMnO ), a significant compression
of the first interlayer distance (d12 ¼ 29.5%) is observed, which is partially
canceled out by the expansion of d23 (+20.0%). Deep down in the Mn3 O4 part of
the slab the situation remains practically unchanged, including the distance dI2
between I2 and (S-7). At the interface, the distance dI between the contact Mn3 O4
and MnO layers is found to be slightly larger than the MnO(001) interlayer distance.
The formation of the interface leads to an upward shift of the manganese atoms
placed in the MnO layer, thus inducing a relatively small buckling of 0.2 Å between
these atoms and the interfacial MnO layer oxygens. It is worth noting that the bare
10 Low Dimensionality and Epitaxial Stabilization. . . 231

MnO(001Þ surface exhibits a larger displacement of the topmost manganese atoms


and an almost doubled surface buckling [14]. The rest of the MnO slab remains
unaffected and reproduces the bulk-like behavior. In conclusion, the growth of
strained Mn3 O4 ð001Þ films on a MnO(001)-oriented substrate inhibits the structural
relaxations observed in the clean MnOð001Þ surface. Finally, because of the large
vertical compression experienced in the Mn3 O4 ð001Þ side of the slab there is a
general in-plane rearrangement of the atoms. As shown by the yellow arrows in
Fig. 10.12b, c, the displacements mainly involve the movement of oxygens along
the [010] direction, which leads to an important distortion of the octahedral
O–Mn–O chain.

10.2.2.2 Epitaxial Stabilization of MnO(111) Overlayers

This conclusive section is devoted to the transformation of MnO(100) layers into a


polar MnO(111) surface, which is observed upon annealing at a high temperature in
ultrahigh vacuum (UHV) [3, 4]. Keeping in mind the intrinsic thermodynamic
stability of the neutral MnO(100), this is an interesting and somewhat unexpected
result and therefore deserves an in-depth analysis. As already mentioned, the
problem with the stability of polar ionic crystal surfaces has been long recognized,
and recently its main lines have been discussed in a number of review articles
[13, 32, 61]. However, while the (111) surfaces of rock salt oxides such as MgO
[6, 30, 31, 43], NiO [10–12, 53, 77], FeO [29, 64, 67], and CoO [51, 52] have been
intensively investigated experimentally and theoretically, there is instead surpris-
ingly little work done on the polar MnO(111) surface. Using the grazing incidence
X-ray scattering technique, Renaud and Barbier have found that the MnO(111)
single crystal surface is nonstoichiometric, containing a mixture of MnO(111) and
Mn3O4(111) phases [66]. Rizzi et al. have grown epitaxial MnO(111) films on a Pt
(111) substrate and characterized their structure by X-ray photoelectron diffraction
(XPD) and LEED, suggesting that the nonreconstructed MnO(111) surface can be
stabilized by adsorption of OH groups [68, 69]. Finally, DFT-based thermodynam-
ics has found that a (2  2) octopolar reconstructed MnO(111) is the most stable
surface over a wide range of the oxygen chemical potential [24].
LEED patterns corresponding to manganese oxide films of various thickness
obtained by reactive evaporation of Mn at low oxygen pressures and high
temperatures are displayed in Fig. 10.13. The LEED reflexes now exhibit a pure
hexagonal arrangement due to the MnO(111) overlayer, and the presence of a faint
streakiness signals that faceting occurs to some extent. The phenomenon of the
formation of facets becomes more pronounced at higher oxide coverage and high
postannealing temperatures, as visible in MnO films of about 30 ML
(Fig. 10.13b–d). A detailed analysis [3] reveals that the facet rods are inclined
with respect to the (00) rod of the MnO(111) surface by an angle of 53 3 , which
is very close to the angle of 54.7 , expected between bulk (111) and (100) planes.
This clearly demonstrates that the facets are of (100) type. Geometrical arguments,
supported by FM-AFM profiles (not shown), indicate that the MnO(111) surface is
232 C. Franchini and F. Allegretti

Fig. 10.13 (a) Conventional LEED picture of 10 ML MnO(111) on Pd(100) evaporated at
670 K in p(O2) ¼ 1  107 mbar and annealed in UHV at 770 K. (b–c) SPA-LEED patterns of 30
ML MnO(111) evaporated at 670 K in p(O2) ¼ 1  107 mbar and annealed at 820 K, recorded
with electron energies of 54 and 70 eV, respectively. (d) SPA-LEED pattern of the surface in (b)
and (c) annealed at 920 K in UHV, recorded with an electron energy of 70 eV. Adapted from [3]

covered by triangular pyramids exposing neutral (100) facets with the lowest
surface energy. Therefore, the formation of the (100)-faceted pyramids on top of
the MnO(111) surface provides a channel for minimizing the total energy. Similar
(100)-faceted pyramids have been observed with AFM by Mocuta et al. on NiO
(111) films supported on an a-Al2O3(0001) surface [53].
However, the replacement of the MnO(100) surface by the polar MnO(111)
cannot be completely explained by the faceting argument alone. To understand this
result we need to take into account the strain energy at the metal–oxide interface,
which is determined by the lattice matching conditions. The MnO(100)–Pd(100)
interface is characterized by a large lattice mismatch of 14% along the [011] rows
(Fig. 10.14a), whereas for the MnO(111)–Pd(100) interface in one direction the
lattice mismatch is only 1%, which results in an almost perfect row matching
along the ½01 1 rows, as illustrated in Fig. 10.14b. This lowers the energy of the
MnO(111)–Pd(100) interface and presumably stabilizes the formation of the MnO
(111) overlayer. This finding is corroborated by the DFT results of Fig. 10.14c
showing the pressure-dependent variation of the energy of formation Ef for MnO
10 Low Dimensionality and Epitaxial Stabilization. . . 233

Fig. 10.14 Real-space model


of the (a) MnO(100)–Pd(100)
and (b) MnO(111)–Pd(100)
interfaces. Note the row-
matching condition along the
[011] rows at the MnO
(111)–Pd(100) interface.
DFT + U derived plot of the
formation energy of
freestanding 3 ML MnO(100)
and MnO(111) layers as a
function of the lattice
constant. From [3]

(100) and (111) thin slabs. Although the (100) termination remains the most stable
surface in the whole range of variation of the lattice constant a, the curves show that
there is a clear tendency of MnO(111) to gain stability with respect to MnO(100)
layers upon compression, i.e., when the MnO bulk lattice constant shrinks in order to
match that of Pd(100) the energy difference DEf ¼ Ef (111)  Ef (100) decreases by a
factor of 2.5 with respect to that calculated for a ¼ aMnO . This result shows that when
considering the stabilization mechanism of polar surfaces of oxide films, the
metal–oxide interface can play a crucial role. Indeed we have demonstrated above
that well-defined and well-ordered MnO(111)-like interfaces are formed at monolayer
coverage, which can provide structurally graded interfaces for the growth of thicker
MnO(111) polar films. Importantly, the epitaxial stabilization mechanism may also
allow the design of oxide surface orientations in thin films that are not stable in the
bulk form. These oxide phases can have interesting implications in the application
of ultrathin oxide films in diverse areas of the emerging nanotechnologies.

Acknowledgments We are very grateful to all coworkers mentioned in the references, in


particular to F. Netzer, S. Surnev, R. Podloucky, and G. Kresse for their scientific vision, unfailing
inspiration, and enthusiastic support, and to F. Li, G. Parteder and V. Bayer for their invaluable
assistance during different stages of this work. Financial support by the Austrian Science Funds
FWF, by the sixth Framework Programme of the European Community (GSOMEN and
ATHENA), and by the seventh Framework Programme of the European Community (ERC
Advanced Grant SEPON) is thankfully acknowledged.
234 C. Franchini and F. Allegretti

References

1. Anisimov, V.I. (ed.): Strong Coulomb Correlation in Electronic Structure Calculations:


Beyond the Local Density Approximation. Gordon and Breach Science Publishers, The
Netherlands (2000)
2. Agnoli, S., Sambi, M., Granozzi, G., Schoiswohl, J., Surnev, S., Netzer, F.P., Ferrero, M.,
Ferrari, A.M., Pisani, C.: Experimental and theoretical study of a surface stabilized monolayer
phase of nickel oxide on Pd(100). J. Phys. Chem. B 109, 17197 (2004)
3. Allegretti, F., Franchini, C., Bayer, V., Leitner, M., Parteder, G., Xu, B., Fleming, B., Ramsey,
M.G., Podloucky, R., Surnev, S., Netzer, F.P.: Epitaxial stabilization of MnO(111) overlayers
on a Pd(100) surface. Phys. Rev. B 75, 224120 (2007)
4. Allegretti, F., Leitner, M., Parteder, G., Xu, B., Fleming, B., Ramsey, M.G., Surnev, S., Netzer,
F.P.: The (100) ! (111) Transition in Epitaxial Manganese Oxide Nanolayers. In: Cat, D.T.,
Pucci, A., Wandelt, K. (eds.) Physics and Engineering of New Materials, pp. 163–170.
Springer, Heidelberg (2009)
5. Allegretti, F., Parteder, G., Gragnaniello, L., Surnev, S., Netzer, F.P., Barolo, A., Agnoli, S.,
Granozzi, G., Franchini, C., Podloucky, R.: Strained c(4  2) CoO(100)-like monolayer on Pd
(100): experiment and theory. Surf. Sci. 604, 529 (2010)
6. Arita, R., Tanida, Y., Entani, S., Kiguchi, M., Saiki, K., Aoki, H.: Polar surface engineering in
ultrathin MgO(111)/Ag(111): Possibility of a metal-insulator transition and magnetism. Phys.
Rev. B 69, 235423 (2004)
7. Armstrong, A.R., Bruce, P.G.: Synthesis of layered LiMnO2 as an electrode for rechargeable
lithium batteries. Nature 381, 499 (1996)
8. Bachmann, K.J.: The Materials Science of Microelectronics. Wiley-VCH, New York (1994)
9. Baldi, M., Finocchio, E., Milella, F., Busca, G.: Catalytic combustion of C3 hydrocarbons and
oxygenates over Mn3O4. Appl. Catal. B: Environ 16, 43 (1998)
10. Barbier, A., Mocuta, C., Kuhlenbeck, H., Peters, K.F., Richter, B., Renaud, G.: Atomic
structure of the polar NiO(111)- p(2  2) surface. Phys. Rev. Lett. 84, 2897 (2000)
11. Barbier, A., Mocuta, C., Renaud, G.: Structure, transformation, and reduction of the polar NiO
(111) surface. Phys. Rev. B 62, 16056 (2000)
12. Barbier, A., Mocuta, C., Neubeck, W., Mulazzi, M., Yakhou, F., Chesne, K., Sollier, A.,
Vettier, C., de Bergevin, F.: Surface and bulk spin ordering of antiferromagnetic materials:
NiO(111). Phys. Rev. Lett. 93, 257208 (2004)
13. Barbier, A., Stierle, A., Finocchi, F., Jupille, J.: Stability and stoichiometry of (polar) oxide
surfaces for varying oxygen chemical potential. J. Phys.: Condens. Matter 20, 184014 (2008)
14. Bayer, V., Franchini, C., Podloucky, R.: Ab-initio study of the structural, electronic, and
magnetic properties of MnO(100) and MnO(110). Phys. Rev. B 75, 035404 (2007)
15. Bayer, V., Podloucky, R., Franchini, C., Allegretti, F., Xu, B., Parteder, G., Ramsey, M.G.,
Surnev, S., Netzer, F.P.: Formation of Mn3O4(001) on MnO(001): surface and interface
structural stability. Phys. Rev. B 76, 165428 (2007)
16. Buciuman, F., Patcas, F., Craciun, R., Zahn, D.R.T.: Vibrational spectroscopy of bulk and
supported manganese oxides. Phys. Chem. Chem. Phys. 1, 185 (1999)
17. Caslavska, V., Roy, R.: Epitaxial growth of Mn3O4 single-crystal films. J. Appl. Phys. 41, 825
(1970)
18. Chassé, A., Langheinrich, Ch, M€ uller, F., H€
ufner, S.: Growth and structure of thin MnO films
on Ag(001) in dependence on film thickness. Surf. Sci. 602, 597 (2008)
19. Chung, E.M.L., Paul, D.M., Balakrischnan, G., Lees, M.R., Ivanov, A., Yethiray, M.: Role of
electronic correlations on the phonon modes of MnO and NiO. Phys. Rev. B 68, 140406(R)
(2003)
20. de Rudder, J., Van de Wiele, T., Dhooge, W., Comhaire, F., Verstraete, W.: Advanced water
treatment with manganese oxide for the removal of 17a-ethynylestradiol (EE2). Water Res.
38, 184 (2004)
10 Low Dimensionality and Epitaxial Stabilization. . . 235

21. Ertl, G., Kn€otzinger, H., Sch€


uth, F., Weitkamp, J. (eds.): Handbook of Heterogeneous Cataly-
sis, vol. 1–8, 2nd edn. Wiley-VCH, Weinheim (2008)
22. Franchini, C., Bayer, V., Podloucky, R., Paier, J., Kresse, G.: Density functional theory study
of MnO by an hybrid functional approach. Phys. Rev. B 72, 045132 (2005)
23. Franchini, C., Bayer, V., Podloucky, R., Parteder, G., Surnev, S., Netzer, F.P.: Density
functional study of the polar MnO(111) surface. Phys. Rev. B 73, 155402 (2006)
24. Franchini, C., Podloucky, R., Paier, J., Marsman, M., Kresse, G.: Ground-state properties of
multivalent manganese oxides: Density functional and hybrid density functional calculations.
Phys. Rev. B 75, 195128 (2007)
25. Franchini, C., Podloucky, R., Allegretti, F., Li, F., Parteder, G., Surnev, S., Netzer, F.P.:
Structural and vibrational properties of two-dimensional MnxOy layers on Pd(100):
Experiments and density functional theory calculations. Phys. Rev. B 79, 035420 (2009)
26. Franchini, C., Zabloudil, J., Podloucky, R., Allegretti, F., Li, F., Surnev, S., Netzer, F.P.:
Interplay between magnetic, electronic and vibrational effects in monolayer Mn3O4 grown on
Pd(100). J. Chem. Phys. 130, 124707 (2009)
27. Franke, P., Neusch€ utz, D. (ed.): Springer Materials – The Landolt-B€ ornstein Database, vol.
IV/19B4. doi: 10.1007/10757285_37
28. Gilbert, B., Frazer, B.H., Belz, A., Conrad, P.G., Nealson, K.H., Haskel, D., Lang, J.C., Srajer,
G., De Stasio, G.: Multiple scattering calculations of bonding and X-ray absorption spectros-
copy of manganese oxides. J. Phys. Chem. A 107, 2839 (2003)
29. Giordano, L., Pacchioni, G., Goniakowski, J., Nilius, N., Rienks, E.D.L., Freund, H.-J.:
Interplay between structural, magnetic, and electronic properties in a FeO/Pt(111) ultrathin
film. Phys. Rev. B 76, 075416 (2007)
30. Goniakowski, J., Noguera, C.: Microscopic mechanisms of stabilization of polar oxide
surfaces: Transition metals on the MgO(111) surface. Phys. Rev. B 66, 085417 (2002)
31. Goniakowski, J., Noguera, C., Giordano, L.: Using polarity for engineering oxide
nanostructures: Structural phase diagram in free and supported MgO(111) ultrathin films.
Phys. Rev. Lett. 93, 215702 (2004)
32. Goniakowski, J., Finocchi, F., Noguera, C.: Polarity of oxide surfaces and nanostructures.
Rep. Prog. Phys. 71, 016501 (2008)
33. Graf, M., Gurlo, A., Bârsan, N., Weimar, U., Hierlemann, A.: Microfabricated gas sensor
systems with sensitive nanocrystalline metal-oxide films. J. Nanoparticle Res. 8, 823 (2006)
34. Guo, L.W., Ko, H.J., Makino, H., Chen, Y.F., Inaba, K., Yao, T.: Epitaxial growth of Mn3O4
film on MgO(001) substrate by plasma-assisted molecular beam epitaxy (MBE). J. Cryst.
Growth 205, 531 (1999)
35. Guo, L.W., Peng, D.L., Makino, H., Inaba, K., Ko, H.J., Sumiyama, K., Yao, T.: Structural and
magnetic properties of Mn3O4 films grown on MgO(001) substrates by plasma-assisted MBE.
J. Magn. Magn. Mater. 213, 321 (1999)
36. Haywood, B.C.G., Collins, M.F.: Optical phonons in MnO. J. Phys. C Solid State Phys. 4, 1299
(1971)
37. Hagendorf, Ch, Sachert, S., Bochmann, B., Kostov, K., Widdra, W.: Growth, atomic structure,
and vibrational properties of MnO ultrathin films on Pt(111). Phys. Rev. B 77, 075406 (2008)
38. Henrich, V.E., Cox, P.A.: The Surface Science of Metal Oxides. Cambridge University Press,
Cambridge (1994)
39. Heyd, J., Scuseria, G.E., Ernzerhof, M.: Hybrid functionals based on a screened Coulomb
potential. J. Chem. Phys. 118, 8207 (2003)
40. Imada, M., Fujimori, A., Tokura, Y.: Metal-insulator transitions. Rev. Mod. Phys. 70, 1039
(1998)
41. Jones, D.A.: Principles and Prevention of Corrosion, 2nd edn. Prentice Hall, Upper Saddle
River, NJ (1996)
42. Julien, C.M., Massot, M., Poinsignon, C.: Lattice vibrations of manganese oxides: Part I.
Periodic structures. Spectrochim. Acta A 60, 689 (2004)
236 C. Franchini and F. Allegretti

43. Kiguchi, M., Entani, S., Saiki, K., Goto, T., Koma, A.: Atomic and electronic structure of an
unreconstructed polar MgO(111) thin film on Ag(111). Phys. Rev. 68, 115402 (2003)
44. Kohn, W.: Nobel lecture: Electronic structure of matter-wave functions and density
functionals. Rev. Mod. Phys. 71, 1253 (1999)
45. Kresse, G., Furthm€ uller, J.: Efficiency of ab-initio total energy calculations for metals and
semiconductors using a plane-wave basis set. Comput. Mater. Sci. 6, 15 (1996)
46. Kresse, G., Joubert, D.: From ultrasoft pseudopotentials to the projector augmented-wave
method. Phys. Rev. B 59, 1758 (1999)
47. Kurata, H., Colliex, C.: Electron-energy-loss core-edge structures in manganese oxides. Phys.
Rev. 48, 2102 (1993)
48. Langell, M.A., Hutchings, C.W., Carson, G.A., Nassir, M.H.: High resolution electron energy
loss spectroscopy of MnO(100) and oxidized MnO(100). J. Vac. Sci. Techol. A 14, 1656
(1996)
49. Lee, G.H., Huh, S.H., Jeong, J.W., Choi, B.J., Kim, S.H., Ri, H.-C.: Anomalous magnetic
properties of MnO nanoclusters. J. Am. Chem. Soc. 124, 12094 (2002)
50. Li, F., Parteder, G., Allegretti, F., Franchini, C., Podloucky, R., Surnev, S., Netzer, F.P.: Two-
dimensional manganese oxide nanolayers on Pd(100): Surface phase diagram. J. Phys.:
Condens. Matter 21, 134008 (2009)
51. Meyer, W., Hock, D., Biedermann, K., Gubo, M., M€ uller, S., Hammer, L., Heinz, K.: Coexis-
tence of rocksalt and wurtzite structure in nanosized CoO films. Phys. Rev. Lett. 101, 016103
(2008)
52. Meyer, W., Biedermann, K., Gubo, M., Hammer, L., Heinz, K.: Superstructure in the termina-
tion of CoO(111) surfaces: Low-energy electron diffraction and scanning tunneling micros-
copy. Phys. Rev. B 79, 121403(R) (2009)
53. Mocuta, C., Barbier, A., Renaud, G., Samson, Y., Noblet, M.: Structural characterization of
NiO films on Al2O3(0001). J. Magn. Magn. Mater. 211, 283 (2000)
54. M€uller, F., de Masi, R., Reinicke, D., Steiner, P., H€ ufner, S., St€
owe, K.: Epitaxial growth of
MnO/Ag(001) films. Surf. Sci. 520, 158 (2002)
55. Na, C.W., Han, D.S., Kim, D.S., Park, J., Jeon, Y.T., Lee, G., Jung, M.-H.: Ferromagnetism of
MnO and Mn3O4 nanowires. Appl. Phys. Lett. 87, 142504 (2005)
56. Nagel, M., Biswas, I., Peisert, H., Chassé, T.: Interface properties and electronic structure of
ultrathin manganese oxide films on Ag(0 0 1). Surf. Sci. 601, 4484 (2007)
57. Nagel, M., Biswas, I., Nagel, P., Pellegrin, E., Schuppler, S., Peisert, H., Chassé, T.: Ultrathin
transition-metal oxide films: Thickness dependence of the electronic structure and local
geometry in MnO. Phys. Rev. B 75, 195426 (2007)
58. Nayak, S.K., Jena, P.: Giant magnetic moments and magnetic bistability of stoichiometric
MnO clusters. Phys. Rev. Lett. 81, 2970 (1998)
59. Netzer, F.P., Allegretti, F., Surnev, S.: Low-dimensional oxide nanostructures on metals:
hybrid systems with novel properties. J. Vac. Sci. Technol. B 28, 1 (2010)
60. Nishimura, H., Tashiro, T., Fujitani, T., Nakamura, J.: Surface structure of MnO/Rh(100)
studied by scanning tunneling microscopy and low-energy electron diffraction. J. Vac. Sci.
Technol. A 18, 1460 (2000)
61. Noguera, C.: Polar oxide surfaces. J. Phys.: Condens. Matter 12, 367 (2000)
62. Ogale, S.B. (ed.): Thin Films and Heterostructures for Oxide Electronics. Springer, Boston
(2005)
63. Post, J.E.: Manganese oxide minerals: Crystal structures and economic and environmental
significance. Proc. Natl Acad. Sci. USA 96, 3447 (1999)
64. Ranke, W., Ritter, M., Weiss, W.: Crystal structures and growth mechanism for ultrathin films
of ionic compound materials: FeO(111) on Pt(111). Phys. Rev. B 60, 1527 (1999)
65. Rao, C.N.R., Raveau, B. (eds.): Colossal Magnetoresistance. Charge Ordering and Related
Properties of Manganese Oxides. World Scientific, Singapore (1998)
66. Renaud, G., Barbier, A.: The Chemical Physics of Solid surfaces, vol. 9, p. 256. Elsevier, New
York (2001)
10 Low Dimensionality and Epitaxial Stabilization. . . 237

67. Rienks, E.D.L., Nilius, N., Rust, H.-P., Freund, H.-J.: Surface potential of a polar oxide film:
FeO on Pt(111). Phys. Rev. B 71, 2414048 (2005)
68. Rizzi, G.A., Zanoni, R., Di Siro, S., Perriello, L., Granozzi, G.: Epitaxial growth of MnO
nanoparticles on Pt(111) by reactive deposition of Mn2 (CO)1O. Surf. Sci. 462, 187 (2000)
69. Rizzi, G.A., Petukhov, M., Sambi, M., Zanoni, R., Perriello, L., Granozzi, G.: An X-ray
photoelectron diffraction structural characterization of an epitaxial MnO ultrathin film on Pt
(111). Surf. Sci. 482–485, 1474 (2001)
70. Sahner, K., Tuller, H.L.: Novel deposition techniques for metal oxides: Prospect for gas
sensing. J. Electroceram. (2008). doi:10.1007/s10832-008-9554-7
71. Samant, P.V., Fernandes, J.B.: Nickel-modified manganese oxide as an active electrocatalyst
for oxidation of methanol in fuel cells. J. Power Sources 79, 114 (1999)
72. Si, P.Z., Li, D., Lee, J.W., Choi, C.J., Zhang, Z.D., Geng, D.Y., Br€ uck, E.: Unconventional
exchange bias in oxide-coated manganese nanoparticles. Appl. Phys. Lett. 87, 133122 (2005)
73. Soares, E.A., Paniago, R., de Carvalho, V.E., Lopes, E.L., Abreu, G.J.P., Pfannes, H.-D.:
Quantitative low-energy electron diffraction analysis of MnO(100) films grown on Ag(100).
Phys. Rev. B 73, 035419 (2005)
74. Tourney, J., Dowding, C., Worrall, F., McCann, C., Gray, N., Davenport, R., Johnson, K.: Mn
oxide as a contaminated-land remediation product. Mineral. Mag. 72, 513 (2008)
75. von Helmolt, R., Wecker, J., Holzapfel, B., Schultz, L., Samwer, K.: Giant negative magneto-
resistance in perovskitelike La2/3Ba1/3MnOx ferromagnetic films. Phys. Rev. Lett. 71, 2331
(1993)
76. Yang, J., Xu, J.J.: Nanoporous amorphous manganese oxide as electrocatalyst for oxygen
reduction in alkaline solutions. Electrochem. Commun. 5, 306 (2003)
77. Zhang, W.-B., Tang, B.-W.: Stability of the polar NiO(111) surface. J. Chem. Phys. 128,
124703 (2008)
Part III
Applications
Chapter 11
One-Dimensional Oxygen-Deficient
Metal Oxides

Wei-Qiang Han

11.1 Introduction

Metal oxides contain at least one metal cation and the oxygen anion (O2). They
usually are semiconductors and nonstoichiometric under typical experimental
processing conditions. Oxides can be classified into two main categories, as listed
in Table 11.1, viz. p-type and n-type, and into four subcategories, metal deficient,
metal excess, oxygen-deficient, and oxygen excess [1]. For example, cerium oxide
(CeO2) falls into the oxygen deficient category. For this oxide, O2 vacancies are
the cause of metal excess; the oxide will have the real formula CeO2  x to keep the
crystal neutral because two electrons are needed to be incorporated for each oxygen
ion that is removed. A good site for these electrons is the Ce4+ cation. If one
electron is associated with one Ce4+ cation, it will change from a Ce4+ ion to a Ce3+
ion. Interestingly, titanium oxide falls into both the metal-excess and oxygen-
deficient categories.
Nanostructured objects have attracted wide attention in recent years because of
their (1) new physics phenomena that affect physical properties (2) unusual quan-
tum effects and structural properties and (3) promising applications in optics,
electronics, thermoelectric storage, magnetic storage, and renewable energies.
One-dimensional (1D) systems are realized by creating nanostructures that are
quantum confined in one direction. When the dimensionality of the material is
lowered, the new variable of length scale becomes available for control of material
properties. Then, as the system’s size declines and approaches nanometer length
scales, it is possible to elicit dramatic differences in the density of electronic states,
opening new opportunities to alter the physical and chemical properties [2, 3]. 1D
nanostructures, including nanotubes (NTs) and nanowires (NWs), are used as

W.-Q. Han (*)


Center for Functional Nanomaterials, Brookhaven National Laboratory,
Upton, NY 11973, USA
e-mail: whan@bnl.gov

J. Wu et al. (eds.), Functional Metal Oxide Nanostructures, Springer Series 241


in Materials Science 149, DOI 10.1007/978-1-4419-9931-3_11,
# Springer Science+Business Media, LLC 2012
242 W.-Q. Han

Table 11.1 Semiconducting properties of binary metal oxides of nonstoichiometric


composition [1]
p-Type semiconductor n-Type semiconductor
Deficit of metal CoO, NiO, FeO, MnO, Cu2O –
Excess of oxygen UO2 –
Excess of metal – TiO2, ZnO, CdO
Deficit of oxygen – TiO2, CeO2, ZrO2, SnO2, Nb2O5, Ta2O5,
WO3, PuO2, Bi2O3, PbO2

elements for miniaturized electrical, nanofluidic, and optical nanodevices and have
played important roles in renewable energies [4, 5].
This chapter will introduce and detail the nonstoichiometric oxygen-deficient 1D
nanoceria (CeO2  x) and its applications in water–gas shifting (WGS) reaction.
It will also describe nonstoichiometric oxygen-deficient TiO2  x NWs with
nanocavities, substoichiometric Magnéli phases TinO2n  1 NWs and substoichiometric
Cr2O2.4 nanobelts (NBs) with modulation structures.

11.2 Oxygen-Deficient 1D Nano-CeO2  x and Its


Applications in the WGS Reaction

11.2.1 Crystal Structure of Cubic Ceria

Cerium with a 4f25d06s2 electronic configuration can exhibit both +3 and +4


oxidation states, and intermediate oxides whose composition is in the range
Ce2O3–CeO2 can be formed. The dioxide CeO2 crystallizes in the fluorite structure.
It has a face-centered cubic cell (fcc) with space group Fm3m (a ¼ 5.41134 Å,
JCPDS 43-1002). In this structure, each cerium cation is coordinated by eight
equivalent nearest-neighbor oxygen anions at the corner of a cube and each anion
is coordinated tetrahedrally by four cations. The structure, illustrated in Fig. 11.1, is
considered as a cubic close-packed array of cerium ions with oxygen ions
occupying all the tetrahedral holes.
Cerium oxides, ranging from Ce2O3 to CeO2, earlier were treated using the
classic point defect model of nonstoichiometry in which the oxygen-vacant sites
were considered to occur randomly in the lattice, in conformity with the law of
statistical thermodynamics. Later experiments indicated that nonstoichiometric
phases originating from the fluorite lattice were formed at low temperatures by
the removal of oxygen ions and ordering of the vacancies formed [6].
Reduced ceria results from the removal of O2 ions from the CeO2 lattice, so
generating a vacant anion site according to the following equation:

4Ce4þ þ O2 ! 4Ce4þ þ 2e =□ þ 0:5O2 ! 2Ce4þ þ 2Ce3þ þ □ þ 0:5O2


11 One-Dimensional Oxygen-Deficient Metal Oxides 243

Fig. 11.1 The crystal structure of CeO2 in the fluorite structure

where □ represents an empty position (anion-vacant site) originating from the


removal of O2 from the lattice, here represented as an oxygen tetrahedral site
(Ce4O). Electrostatic balance is maintained by the reduction of two cerium cations
from +4 and +3.

11.2.2 Background of the WGS Reaction

The WGS reaction, typically used to generate H2 through the reaction of a gas
mixture of CO and H2O (CO + H2O ! H2 + CO2, DH 298 ¼ 41.1 kJ/mol), is a
well-known catalytic process of industrial importance [7]. Based on thermodynamic
and kinetic considerations, a high conversion of CO is obtained with a two-bed
operation at low (180–250 C) and high temperatures (350–500 C). In continuously
operating industrial applications, the classical catalysts employed are Fe2O3–Cr2O3
for the first stage (high-temperature shift (HTS)), and Cu/ZnO/Al2O3 for subsequent
stages (low-temperature shift (LTS)), to obtain a good performance under steady-
state conditions.
For proton exchange membrane (PEM) fuel cells, the anode catalyst usually is
Pt/C, chosen as it is more sensitive to CO than those mentioned above, because the
PEM fuel cell operates at lower temperatures at which CO can deactive the Pt.
Usually, the CO in the fuel must be deeply reduced to <10 ppm. The WGS reaction
is a critical step in fuel processors for preliminary cleanup of CO, and the additional
244 W.-Q. Han

generation of hydrogen before the preferential oxidation of CO, or the methanation


step. WGS units sited downstream of the fuel reformer to further lower the CO
content, and improve the H2 yield. To obtain this equilibrium outlet concentration of
CO from the reformate fuel, the WGS catalyst must be active at low temperatures,
200–280 C, depending on the inlet concentrations of CO in it. The reaction is
moderately exothermic, with low temperatures resulting in low CO levels; however,
the kinetics of the reaction is favorable even at higher temperatures [8].
Employing Fe–Cr and Cu–ZnO catalysts in fuel processors poses a series of
disadvantages: the low activity of Fe–Cr as an HTS catalyst and its thermodynamic
limitations at high temperatures; the sensitivity of the Cu–ZnO catalyst to air or
temperature excursions; the lengthy preconditioning of such catalysts for intermit-
tent operation (pre-reduction/passivation); and the large reactor volume dictated by
the slow WGS kinetics of the Cu–ZnO catalyst at low temperatures. Therefore,
Fe–Cr and Cu–ZnO catalysts are unsuitable for automotive applications, where the
need for fast start-ups dictate the need for using a low volume of a nonpyrophoric
catalyst (fewer stages). Thus, it is critical to develop efficient safe catalysts for the
WGS reaction in the fuel cell process.
Ceria recently attracted great interest, particularly for reducing the emissions of
CO, NOx, and hydrocarbons from automobile exhaust, to abate soot formation in
diesel fuels, and to minimize CO content in fuel cells [9]. The key to these
applications is that CeO2  x easily produces oxygen vacancies in an oxygen-deficient
environment, shifting some Ce4+ to Ce3+ ions in the stable fluorite structure [10].
Oxygen vacancies also are crucial for binding catalytically active species to ceria.
Thus, oxygen vacancies in ceria are considered to play an essential role in catalytic
reactions [11]. Ceria-supported noble metal cocatalysts, such as Pt-, Au-, and Pd-
loaded ceria, exhibit very interesting properties for the WGS reaction with fuel cells
[12]. Under some conditions related to H2 production, the WGS reaction rates were
higher on noble-metals-loaded ceria than on commercial catalysts [13].
This section discusses the use of pure and Pd-loaded 1D nanoceria, a mixture of
NTs and NWs, as catalysts for the WGS reaction at lowtemperatures.

11.2.3 Synthesis of 1D Ceria

Various methods are used to prepare special ceria morphologies with enhanced
reducibility. Zhou et al. generated ceria nanoparticles (NPs) by adding an aqueous
ammonium hydroxide precipitant to a solution of cerium nitrate at room tempera-
ture and then introduced oxygen into the reactor to oxidize Ce3+ to Ce4+ [14]. Chen
et al. obtained ceria NWs by adding this same precipitant into cerium nitrate at
70 C, and subsequently allowed it to age at 0 C for 1 day [15]. Yu et al. prepared
ceria nanocrystals (NCs) in spherical, wire, and tadpole shapes from a
nonhydrolytic sol–gel reaction of cerium (III) nitrate and diphenyl in the presence
of surfactants [16]. Natile et al. synthesized ceria NPs by two different synthetic
11 One-Dimensional Oxygen-Deficient Metal Oxides 245

Fig. 11.2 Powder profile refinement of fresh 1D ceria

routes: precipitation from a basic solution (sizes around 8–15 nm) and microwave-
assisted heating hydrolysis (size around (3.3–4.0 nm) [17]. They found that the NPs
synthesized by the latter method were more reduced than those synthesized from
the former. Methanol oxidation is also favored on the ceria NPs prepared by the
latter method because of their high specific area and the presence of greater amount
of active sites of Ce3+ cations. Zr4+ and La3+ doped porous ceria NPs with a high
BET surface area of 160 m2/g exhibited a photovoltaic response, directly derived
from the NPs’ size; normal ceria does not show this response [18]. All these results
suggest that ceria with high surface area can increase the Ce3+ ratio that leads to
high reducibility.
The 1D nano-CeO2  x used for WGS reaction, described in the next section, was
synthesized in two successive stages: precipitation and aging. At the precipitation
stage, 1.5 g of cerium nitrate (Ce(NO3)3  6H2O) was added to 15 ml deionized
water and heated at 100 C. Once a large amount of vapor formed, 10 ml of 5%
ammonium hydroxide solution was added. A very fine yellowish precipitate formed
immediately and the mix started boiling. After 3 min, the solution was transferred
quickly to a 0 C refrigerator [19]. Figure 11.2 displays the powder profile refine-
ment of the as-produced material using GSAS/EXPGUI code.
A careful inspection reveals that there are two kinds of 1D nanostructures of
CeO2  x. One is a NW with consistent crosswise lattice, while the other is the NT
with weak contrast in the middle (Fig 11.3a). These characteristics can be seen
more clearly in high-resolution images of a CeO2  x NT and NW, respectively
(Fig. 11.3b, c). The transmission electron microscopy (TEM) image of the pure 1D
nano-CeO2  x shows polycrystalline ceria NWs and NTs (~80%), together with
246 W.-Q. Han

Fig. 11.3 Pure 1D ceria sample (a) typical morphology with three kinds of nanostructures: NPs,
NWs, and NTs; (b) a high-magnification TEM image of a ceria NT, with a wall of about 5.5 nm
thick; and, (c) a high-magnification TEM image of a ceria NR

ceria NPs (~20%) with a diameter similar to that of the 1D nano-CeO2  x


(Fig. 11.3a). Most 1D nanocerias are NWs whose diameters typically range from
6 to 25 nm and lengths up to a couple of microns. In Fig. 11.3b, the selected area
diffraction patterns (obtained by fast Fourier transform (FFT) techniques) in the
upper right corner correspond to cubic ceria. The direction of the incident
electronbeam is along <110>, i.e., the exposed crystal plane is (110). The axis of
the CeO2  x NT is along the <110> direction. Two kinds of lattice fringe directions
attributed to (111) and (200) are observed that, respectively, have an interplanar
spacing of 3.1 and 2.7 Å. For most NTs, the thickness of the wall is almost uniform
over the tube, though the thickness differs from tube to tube. Figure 11.3c shows
that the CeO2  x NW has the same crystalline features as the NT. For the 1D ceria,
11 One-Dimensional Oxygen-Deficient Metal Oxides 247

Fig. 11.4 EELS spectra


showing different M5 peak
intensity for CeO2  x NTs
with (a) d ¼ 14.6 nm,
(b) d ¼ 17.3 nm, and (c)
25.5 nm. The thicknesses
of the wall of the NTs are 5.5,
6.0, and 10.8 nm for (a), (b)
and (c), respectively. The
spectra are normalized for
the M4 peak

the preferred exposed crystal planes for both NWs and NTs are {110} and {100}.
Based on electron diffraction analyses and high-resolution imaging, the CeO2  x
NPs, NWs, and NTs were shown to have the same crystal structure, a cubic fluorite
structure, consistent with the X-ray measurements. The lattice parameters of the
CeO2  x NTs vary from 0.54 to 0.56 nm depending on their diameters. In general,
the lattice parameter increases with decreasing diameter of the NTs. Cerium nitrate
solution reacts with ammonium hydroxide to form Ce(OH)3 as an intermediate
product with a 1D nanostructure that is retained if the pH of the reaction is higher
than 8 [20]. Excess ammonium hydroxide was used in the present experiment so the
intermediate Ce3+ oxidized to Ce4+. Quickly cooling the samples to 0 C retained
the 1D nanostructure. The precipitates were dehydrated further, and recrystallized
during the aging time. Prolonging aging time leads to more 1D-like hollow struc-
ture, i.e., NTs.
The increase in the lattice parameter of the CeO2  x NTs implies that the
oxidation state of the CeO2  x NTs may differ from that of bulk CeO2. Electron
energy loss spectrometry (EELS) can analyze the chemical composition of TEM
specimens with a lateral resolution down to about 1 nm. The valency of the cerium
ions is determined from the relative intensity of the white lines (M4 and M5) of the
cerium in the EELS spectra. The NPs are almost completely reduced to CeO1.5
when the diameter < 3 nm. This reduced CeO2  x has a fluorite structure, the same
as that of bulk CeO2. Also, EELS spectra taken from the edge and center of the
NP indicated that for large NPs the valence reduction of cerium ions occurs mainly
at the surface, forming a Ce1.5 layer and leaving the core essentially as CeO2.
The fraction of Ce3+ ions in the NPs rapidly increased with declining size of the
NPs [21]. Figure 11.4 shows the M4 and M5 edges of the EELS spectra from three
NTs with diameter, d ¼ 14.6, 17.3, and 25.5 nm. It qualitatively illustrates that the
systematic change in the EELS spectra is correlated with the diameters of the NTs,
248 W.-Q. Han

that is the intensity of the M5 edge rises with the decrease in the diameter of the
NTs. To determine the relative amounts of cerium ions Ce3+ and Ce4+, the second
derivative method is used to measure the M5/M4 ratio, since it is insensitive to
variations in thickness. The M5/M4 ratio of these three NTs, d ¼ 14.6, 17.3, and
25.5 nm, respectively, are 1.27, 1.22, and 1.05; based on the M5/M4 ratio of 1.31 for
Ce3+ and 0.91 for Ce4+, the fraction of Ce3+ (Ce3+/[Ce3++Ce4+]) therefore is
estimated correspondingly as 0.90, 0.78, and 0.35. Compared with the CeO2  x
NPs of the same diameter, the fraction of Ce3+ in the CeO2  x NTs is significantly
larger. The main reason is that NTs have two surfaces: The outer surface and the
inner one. Actually, the total surface area depends on the thickness of the wall of the
NTs. If the cerium ions in the CeO2  x NTs follow the same distribution as that of
the CeO2  x NPs, that is, Ce3+ exists on the surface, while Ce4+ inside, the fraction
of Ce3+ would be determined mainly by the thickness of the wall. In fact, the
thicknesses of the wall of the NTs for Fig. 11.4a–c are about 5.5, 6.0, and 10.8 nm,
respectively. Oxygen vacancies in ceria NT combined with their inner and outer
surfaces could offer more functional, effective features and play an essential role in
applications, such as catalytic reactions. Techniques to make high-yield ceria NTs
with sustainable stability during the WGS reaction still is challenging and worthy of
further effort [19].

11.2.4 Testing 1D Ceria for the WGS Reaction

The in situ time-resolved X-ray diffraction (XRD) experiments were performed at


beam line X7B (l ¼ 0.922 Å) of the National Synchrotron Light Source (NSLS) at
Brookhaven National Laboratory. The in situ Ce LIII-edge X-ray absorption near
edge spectra (XANES) and the in situ Pd K-edge data were collected there, at beam
lines X19A and X18B, respectively [22]. The products from time-resolved XRD
and XAFS experiments were measured using a 0–100 amu quadruple mass spec-
trometer (QMS, Stanford Research Systems). The portion of the exit gas flow that
passed through a leak valve and into the QMS vacuum chamber provided the
relative pressure of the products [23].
Samples of 1–2 mg were loaded into a 1-mm sapphire capillary tube attached to
a flow system. The 1D nanoceria was exposed in pure H2 up to 400 C for activation
before the WGS reaction [24, 25]. A set up similar to that used for the WGS reaction
was employed for the temperature-programmed reduction and oxidation, for which
pure H2 and 5% O2 in He were used, respectively. The temperature ramp rate was
~2 C/min. The in situ time-resolved XRD patterns (Fig. 11.5a) showed the reten-
tion of the cubic fluorite structure, and peak widths that were nearly constant during
the reduction process, although there were significant changes in the lattice param-
eter from thermal expansion and the partial reduction of the cerium oxide, viz. from
5.43 Å at 25 C to 5.47 Å at 400 C. The reduction of pure 1D nanoceria in H2 started
at 150 C, a much lower temperature than those previously reported for bulk or 3D
ceria NPs (i.e., NPs with no preferred growth in any direction).
11 One-Dimensional Oxygen-Deficient Metal Oxides 249

Fig. 11.5 Pure 1D ceria sample. (a) A 3D plot of in situ time-resolved XRD patterns collected
during the hydrogen reduction process. (b) H2 and CO2 relative pressures during the WGS
reaction. (c) TEM image of the sample after the WGS reaction. (d) The lattice parameter of the
ceria determined from the in situ diffraction during WGS and H2 reduction conditions as a function
of temperature, which show relative cell expansion of H2 vs. WGS

Once the catalyst was cooled down to ambient temperature, the gas was switched
to 5% CO/He passing through a water bubbler, at ambient temperature, at a flow
rate of 10 ml/min. The H2O versus CO vapor ratio was ~0.35. After equilibration,
the WGS reaction was carried out isothermally at 200, 250, 300, and 350 C, with a
holding time of 4 h at each temperature. Figure 11.5b displays the relative pressure
of H2 and CO2 as a function of time. Catalytic activity increased with increasing
temperature, becoming significant at 300 C. This behavior is very different from
that of bulk and 3D ceria NPs, which exhibit a negligible catalytic activity for the
WGS reaction. A series of in situ XRD patterns collected during the WGS reaction
showed no obvious changes; however, further analysis (see below) revealed
alterations in the number of oxygen vacancies in the ceria structure, and in the
cell dimensions. It was proposed that ceria participates in the WGS reaction when
250 W.-Q. Han

the oxygen vacancies formed by CO reduction facilitate the breakdown of the H2O
to form H2 and O2 ions [26]. The number of oxygen vacancies can be determined
from the lattice parameters of the ceria under WGS conditions and H2 reduction
conditions because the cell expands when cerium is reduced from Ce+4 to Ce+3 [27].
However, the cell also displays thermal expansion; consequently, one can
only compare relative oxidation at a chosen temperature from lattice parameters.
Figure 11.5c is the TEM image of the 1D ceria (~60%) after the WGS reaction.
Compared to the prereaction sample, the 1D ceria still are crystalline and the
preferred exposed crystal planes are {110} and {100}, though the amount of 1D
ceria is slightly decreased. Figure 11.5d shows the lattice parameter of the ceria
determined from in situ diffraction during WGS and H2 reduction conditions as a
function of temperature. The cell expands more during H2 reduction conditions than
under WGS conditions at 350 C. This finding reveals that the presence of H2O
together with the CO partly reoxidizes the ceria, and is consistent with the hypoth-
esis that the reaction of H2O at the O vacancy sites produces adsorbed O2 and H2
gas [26].
To enhance catalytic activities at lowtemperatures, Pd (1% in weight) was
loaded on the 1D nanoceria [28]. It was deposited on the 1D nanoceria through
the drop wise addition of palladium nitrate (Pd(NO3)2  xH2O) (1 wt%) into an
aqueous suspension of 1D nanoceria. After stirring the solution for 1 day, it was
washed with ethanol and then with distilled water. Examination under TEM of the
Pd-loaded 1D ceria sample did not find any isolated Pd0 or Pd-ion NPs. EDS also
did not detect a Pd signal, signifying that there was no Pd-rich area. Figure 11.6a
displays the in situ XRD patterns of the Pd-loaded 1D nanoceria collected during
the reduction process with pure H2 up to 200 C. The cubic fluorite structure of ceria
was stable during the reduction process, similar to pure 1D nanoceria. At room
temperature, no other diffraction peaks were apparent, except those of 1D
nanoceria. Around 60 C, the diffraction peaks of Pd started to appear. However,
only Pd (111) was observed due to its low concentration as well as its broad
dispersion on the 1D nanoceria. Together, the XRD and TEM results lead to the
conclusion that Pd exists in the starting sample as a noncrystalline species.
In addition, the dimension of the ceria cell still was 5.449 Å (see Table 11.2)
when the sample was cooled under H2 flow to room temperature. This could reflect
the stabilization of the reduced structure of ceria in a flow of pure hydrogen or by
the storage of hydrogen in the ceria lattice [29].
Figure 11.6b displays the resulting H2 and CO2 relative pressures as a function of
WGS reaction time. Some WGS activity was observed at 200 C, and catalytic
activity increased at the other three temperature ranges. The corresponding in situ
time-resolved XRD patterns displayed structures of cubic ceria and metallic Pd
with small changes in the cells. Under identical experimental conditions, the WGS
reaction was repeated for the same sample in two additional passes; good catalytic
activity was observed during both passes. Figure 11.6c shows the H2 and CO2
relative pressures during the second pass of the WGS reaction using the Pd-loaded
1D nanoceria catalyst.
11 One-Dimensional Oxygen-Deficient Metal Oxides 251

Fig. 11.6 Pd-loaded 1D ceria sample. (a) A 3D plot of in situ time-resolved XRD patterns
collected during the reduction process. (b) H2 and CO2 relative pressures during the first pass of
the WGS reaction using the catalyst of Pd-loaded ceria. The catalyst was at first ramped to 200 C
in H2. (c) H2 and CO2 relative pressures during the second pass of the WGS reaction using the
catalyst of Pd-loaded ceria. (d) The lattice parameters of the ceria determined from the in situ
diffraction during WGS first pass and second pass

Table 11.2 Detailed analysis of the XRD patterns for Pd-loaded 1D nanoceria catalyst
Lattice parameter (Å)
Sample CeO2 CeO2 from 400 C peak Pd
Starting materials 5.423 – –
After reduced at RT 5.449 – –
Before 1st WF + GS 5.427 5.427 4.035
Before 2nd WGS 5.435 5.432 3.948
Before 3rd WGS 5.433 – 3.938

The XRD patterns of the Pd-loaded 1D nanoceria sample after the WGS runs in
CO/He/H2O showed Pd (111) had shifted to lower 2y angle after the first pass.
Table 11.2 gives details of the lattice parameters of the Pd and ceria. The lattice
252 W.-Q. Han

Fig. 11.7 Pd-loaded 1D ceria sample. (a) Pd K-edge XANES spectra during the reduction
process. (b) Ce LIII-edge XANES spectra during the reduction process. (c) Ce LIII-edges
XANES spectra during the WGS reaction

parameters of Pd decreased after each pass of the WGS reaction. However, after H2
reduction (RT, H2 flow), the lattice parameter of Pd at 4.035 was much larger than
´
that of the measured Pd lattice parameter of 3.889 Å (powder diffraction file 46-
´
1043). Since the lattice parameter of PdH0.76 is 4.02 Å (powder diffraction file 19-
0951), it is likely that the Pd is in the hydride form. The in situ time-resolved XRD
pattern of Pd-loaded 1D nanoceria during the pretreatment in pure hydrogen
displayed a shift of the diffraction peak of Pd (111) to a lower two-theta angle as
soon as the temperature was decreased to the ambient level. This suggests that PdHx
forms during cooling [30]. Figure 11.6d depicts the lattice parameters of ceria
determined from the in situ diffraction during the WGS first pass and second
pass; the cell parameters are found to be lower in the second one. This could be
related to the fall in activity after each pass, since the decline in the lattice
parameters of ceria indicates a lesser extent of reduction, and thus fewer oxygen
vacancies as in the succeeding of WGS reaction.
11 One-Dimensional Oxygen-Deficient Metal Oxides 253

In situ XANES measurements were used to study the oxidation state of the Pd
and Ce in 1D ceria during the catalytic process. Figure 11.7a displays the Pd K-edge
XANES spectra of the Pd-loaded 1D nanoceria sample. The absorption edge of the
starting material is at a higher energy position than that of the Pd foil, suggesting
that the Pd is ionic. Also, the much greater white-line intensity implies that Pd is
ionic in the starting material. After purging at room temperature in H2 for 3 h, the Pd
K-edge spectra of Pd-loaded 1D nanoceria resembled that of the Pd foil, indicating
that the Pd species was reduced, even at ambient temperatures, although the
spectrum still displayed a slightly higher white-line intensity than did the Pd foil;
this indicated that the Pd ions are partially reduced. Pd hydride might also be
formed since it has a similar XANES pattern to that of the Pd foil [31]. However,
Pd hydride displays a slightly larger white-line intensity in the Pd K-edge XANES
as well as some shift in peaks ii and iii to relatively lower energy. For a similar
cluster of Pd or Pd hydride, it was observed that (Er  Eb)R2 ¼ constant, where R is
the interatomic distance, Er is the energy of resonance, and Eb is the energy of a
bound state at threshold [32]. Thus, this shift essentially reflects an increase in the
Pd–Pd bond distance due to an increase in the lattice parameters during the
formation of the hydride, which the XRD pattern confirms. Also apparent in
Fig. 11.7a, at 300 C, is that the Pd K-edge spectrum of the catalyst is very similar
to that of the Pd foil, pointing to a complete reduction of Pd to its metallic form.
Figure 11.7b illustrates Ce LIII-edge XANES spectrum of the Pd-loaded 1D ceria
sample during the reduction process; this edge is shifted to a lower energy position
in a pure hydrogen atmosphere at ambient temperature, a finding consistent with the
results of the Pd K-edge XANES, where the sample of Pd-loaded 1D ceria is
partially reduced at ambient temperature. At 60 C, the Ce LIII-edge shifted to an
even lower position, indicating further reduction of the ceria. At 200 C, the shift of
the Ce LIII-edge does not shift much more. We also noted that there is a shoulder
peak at 5,727.4 eV in pristine Pd-ceria catalyst (highlighted with an arrow),
pointing to the existence of some Ce3+ ions. The intensity of this shoulder increased
gradually as the temperature rose, suggesting that the concentration of the Ce3+ ions
increased. In CO/H2O, the Ce LIII-edge spectrum of the Pd-loaded 1D nanoceria
was similar to that in H2 atmosphere at ambient temperature, indicating a slight
oxidation of the catalyst after it was fully reduced in H2 at 200 C (Fig. 11.7c).
During the WGS reaction, the Ce LIII-edge position shifted to a lower energy at
350 C, while its intensity increased at 5,727.4 eV, pointing to the reduction of ceria
in the Pd-ceria catalyst under the conditions of the WGS reaction [28].
A redox mechanism might be at the basis of the reaction [13]. Using s to
designate an adsorption site, the mechanism can be written as follows:

CO þ s ! COad

H2 O þ s ! Oad þ H2

COad þ sad ! CO2 þ 2 s


254 W.-Q. Han

Fig. 11.8 Changes in ceria lattice parameters during the temperature ramping of 1D nanoceria,
Pd-loaded 1D nanoceria and 3D ceria NPs

The catalyst is oxidized by H2O and reduced by CO. The adsorption sites for
COad and Oad need not be the same. For Pd/ceria catalysts, CO might adsorb onto
the Pd and then reduce the ceria near the Pd interface; subsequently, H2O reoxidizes
the reduced ceria [13].
To understand more comprehensively the interaction between the loaded Pd
clusters and the 1D nanoceria, the changes in the ceria lattice parameters of pure 1D
nanoceria, Pd-loaded 1D nanoceria and ceria 3D NPs during the hydrogen reduc-
tion process were compared (Fig. 11.8). For pure 1D ceria, the lattice parameters of
ceria barely changed at the beginning of the ramping process, but increased after the
onset temperature of 150 C. At 350 C, the lattice expanded by ~0.04 Å compared
with that at 150 C, as a consequence of the reduction of Ce4+ to Ce3+ ions, since the
radius of the Ce3+ (1.14 Å) is much larger than that of the Ce4+ (0.97 Å). This
temperature at which reduction occurred was much lower than that for ceria NPs
(490 C) [33]. During the hydrogen reduction process, a dramatic increase in the
ceria lattice parameter (0.025 Å) was observed at about 60 C for Pd-loaded 1D
nanoceria, and concurrently metallic Pd appeared. This finding suggested that the
interactions between the Pd clusters and the 1D nanoceria might facilitate the
reduction of Pd-loaded 1D nanoceria, which was consistent with reports that Au
(or Cu) can activate the surface of ceria and decrease its reduction temperature [34].
The Pd-loaded 1D nanoceria could be reduced at temperatures ~100 C lower
than pure 1D nanoceria, wherein large number of oxygen vacancies might be
produced to participate in the formation of the active sites for WGS reaction.
As observed from Fig. 11.8, pure 1D nanoceria was significantly reduced at
300 C, leading to the good catalytic activity observed in Fig. 11.5b, mainly due
11 One-Dimensional Oxygen-Deficient Metal Oxides 255

Fig. 11.9 TEM images of Pd-loaded 1D ceria reduced and pretreated in hydrogen at 200 C after
WGS reaction: (a) NWs and (b) nanoparticles; TEM images of Pd-loaded 1D ceria pretreated in
hydrogen at 400 C after WGS reaction: (c) NWs and (d) nanoparticles; (e) schematic image of the
evolution of (A) a NW breaking down into faceted NPs and (B) subsequently changing to
nonfaceted 3D NPs

to the 1D conformation of the ceria, whose preferred exposed crystal planes are
{110} and {100}. The {110}/{100}-dominated surface structures are more reactive
for CO oxidation than the {111}-dominated one.
Long-term stability is a concern with all WGS catalysts and particularly so with
ceria-supported precious metals. There is no consensus on what processes are most
important [13]. Zalc et al. proposed that deactivation occurs from ceria overreaction
[35]. Goguet et al. suggested that deactivation in the Pt/ceria catalysts is due to
formation of carbon during reaction at 400 C as a result of CO dissociation [36].
Wang et al. considered that the deactivation of Pt/ceria and Pd/ceria catalysts under
the WGS reaction originated from the loss of metal dispersions since the WGS rates
of a series of Pd/ceria catalysts increased linearly with surface area of the Pd [37].
To understand the deactivation mechanism, we studied the effect of pretreating the
catalysts with hydrogen by combining the activity measurements with subsequent
TEM measurements [28]. For Pd-loaded 1D nanoceria, the one pretreated in H2 at
200 C had much higher activity than the one treated in H2 at 400 C. Thus, the
256 W.-Q. Han

relative WGS catalytic activity at 250 C for samples of Pd-loaded (1%) 1D


nanoceria pretreated in H2 up to 400 and 200 C are, respectively, 1.83% and
2.96%. For both such samples, 25% of the 1D nanostructures remained after the
WGS reaction. Also, the preferred exposed crystal planes for these remaining 1D
nanocerias were {110} and {100} for both samples (Fig. 11.9a, c). In both samples,
a large number of 1D ceria nanostructures were broken down into NPs after a
couple of redox cycles. The TEM image, taken after WGS reaction, of the NPs of
the sample pretreated in hydrogen at 200 C often resemble polyhedrons with clear
crystalline facets, and the preferred exposed surfaces still are {100} and {110}
(Fig. 11.9b). However, many NPs of the sample pretreated in the hydrogen atmo-
sphere at 400 C after WGS reaction do not have clear crystalline facets, or a
preferred crystalline orientation (Fig. 11.8d). Besides {100} and {110}, there are
many other exposed surfaces, such as (111) that is an unfavorable surface for CO
oxidation [38]. This might explain why the 1D nanoceria sample pretreated in
hydrogen at 200 C has a better WGS activity. Another factor in the decrease in
surface area might be the aggregation of NPs after the reaction. In the present work,
a higher reducing temperature introduced more changes in crystallographic faces to
those that do not favor the activity. Figure 11.9e shows a schematic image of the
evolution of a NW breaking down into faceted NPs, and subsequently changing to
nonfaceted NPs, i.e., 3D NPs. The loss of the effective faceted surfaces of ceria
during the WGS process could be responsible for the loss of catalytic activity.

11.3 Substoichiometric Magnéli Phases 1D TinO2n  1

Stoichiometric titanium dioxide (TiO2) is one of the most widely studied transition
metal oxides because of its many potential photoelectrochemical applications [39].
Although many semiconductors were proposed as photoelectrodes, the general con-
clusion is that TiO2 is the one of the best due to its excellent chemical stability and
resistance to corrosion [40]. There are three forms of stoichiometric TiO2 crystals:
orthorhombic brookite, tetragonal anatase, and rutile [41]. Electronically, these three
TiO2 phases are wide band gap nonconducting materials. However, the wide band
gap and high electrical resistivity of stoichiometric TiO2 are two main factors
precluding the realization of the requirements for a highly efficient photoelectrode.
Hydrogen reduction of stoichiometric TiO2 generates nonstoichiometric TiO2  x
with significantly better conductivity [42]. Many researchers explored the relation
in nonstoichiometric TiO2  x between its electrochemical applications and its
microstructure, composition, defect formation, conduction mechanism, and optical
properties. In particular, studies focused on the substoichiometric titanium oxides,
TinO2n  1 (i.e., the Magnéli phases, where n is a number between 4 and 10 (i.e.,
1.90  x  1.75)) [43, 44]. The Magnéli phases comprise two-dimensional chains
of octahedral TiO2, with every nth layer missing oxygen atoms to accommodate
the loss in stoichiometry. The formed superstructures display recurrent shear
planes wherein the MO6 octahedra exhibit edge-sharing, rather than vertex sharing.
11 One-Dimensional Oxygen-Deficient Metal Oxides 257

Fig. 11.10 (a) XRD spectrum of the sample prepared by heating intermediate H2Ti3O7 at 850 C
under a hydrogen atmosphere; (b) an SEM image of the sample; (c) its low-magnification TEM
image; and (d) a high-magnification TEM image of a part of a NW

Consequently, there is added disorder in such superstructures compared to standard


idealized crystallographic phases. Magnéli phases are insoluble in acids, are electro-
chemically stable, and have a high electronic conductivity; accordingly, they served
as gas sensors for detecting hydrogen, oxygen, and carbon monoxide, in battery
electrodes, antireflection coatings in solar systems, and in photoelectrolysis [45].
In this section, a simple route is described to synthesize Magnéli phases 1D
TinO2n  1. Meanwhile, we measured their electrical transport and optical
properties.
The synthesis process entails two steps. First, the intermediate product, H2Ti3O7
NWs, is produced by treating anatase TiO2 particles with NaOH in an autoclave at
160–180 C for 2–3 days, and then subsequently washing the material with nitric
acid. XRD measurements revealed that the intermediate product is monoclinic
H2Ti3O7 [46, 47]. Scanning electron microscopy (SEM) and TEM images showed
that most products are straight NWs of 30–200 nm diameter, and up to 10 mm long.
In the second step, the intermediate product is heated under hydrogen for 1–4 h at
850–1,050 C. The final products from the reduction treatment were dark, rather
than being white, like the intermediate H2Ti3O7 [48].
Figure 11.10a shows an XRD spectrum of a sample of the final product. The
characteristic spectral peaks denote that the sample is triclinic Ti8O17 (JCPDF# 50-
´ ´ ´
0790, a ¼ 5.526 Å, b ¼ 7.133 Å, c ¼ 44.059 Å, a ¼ 66.54 , b ¼ 57.18 , g ¼ 108.51 ).
An SEM image and a low-magnification TEM image of the sample are shown in
Fig. 11.10b, c, respectively. Both show many NWs with diameters typically ranging
258 W.-Q. Han

from 30 to 200 nm, similar to the intermediate H2Ti3O7 NWs. Figure 11.10d is a high-
magnification TEM image of a part of a single-crystal NW. The labeled lattice
´
distance is 6.4 Å, corresponding to the (004) plane of the triclinic crystal. After
heating, the intermediate H2Ti3O7 is heated at 1,050 C in a hydrogen atmosphere,
the XRD spectrum reveals that the material is triclinic Ti4O7 (JCPDF# 50-0787,
´ ´ ´
a ¼ 5.5965 Å, b ¼ 7.11259 Å, c ¼ 12.4647 Å, a ¼ 95.05 , b ¼ 95.18 , g ¼ 108.77 ).
The TEM image shows that most products are quasi-one-dimensional fibers, with
diameters of about 1 mm, i.e., much larger than that of the intermediate H2Ti3O7 fibers.
Nanocavities embedded within anatase TiO2 NWs were generated by heating
intermediate H2Ti3O7 NWs in an oxidation atmosphere at 650 C. During the reac-
tion, monoclinic H2Ti3O7 transforms into tetragonal anatase TiO2 and H2O (H2Ti3O7
! 3TiO2 + H2O). Figure 11.11a shows numerous polyhedral nanocavities inside the
TiO2 NWs; Fig. 11.11b is a high-resolution TEM image viewed along the [100]
direction. The inset is the FFT from the whole area. The nanocavities have polyhedral
shape. The boundaries between the nanocavities and the NW body are sharp. The
boundaries planes are {01-1} {100}, and {001}, which all are low-index planes of the
anatase crystal and have the lowest surface formation energies (J/m2) of 0.44, 0.53,
and 0.90, respectively. Examining the near-edge fine structure of the Ti L-edge and O
K-edge (see Fig. 11.11c) revealed that the Ti/O ratio in the area of the nanocavity is
found to be 18% higher than in the “body” of the NW. For comparison, an average
spectrum, taken at lower magnification also is included. In addition, the Ti L3/L2 ratio
increases at the nanocavities, suggesting a decrease in the Ti valence, which might be
due to a decline in the O stoichiometry of the TiO2 around the nanocavity [49].
For forming the Magnéli phases, the H2Ti3O7 intermediate NWs were treated at
temperatures higher than 850 C under a reduction environment. The chemical
reactions for transforming intermediate H2Ti3O7 to the final Magnéli phases,
Ti8O15 and Ti4O7, are expressed as

8H2 Ti3 O7 þ 3H2 ! 3Ti8 O15 þ 11H2 O

4H2 Ti3 O7 þ 3H2 ! 3Ti4 O7 þ 7H2 O

During reduction processes at 850 C, monoclinic H2Ti3O7 NWs change into


triclinic Ti8O15 NWs. This latter phase is made up of two-dimensional chains of
TiO2 octahedra, wherein every eighth layer has missing oxygen atoms to accom-
modate for the loss in stoichiometry. At a higher temperature (1,050 C), monoclinic
H2Ti3O7 NWs are transformed into triclinic Ti4O7 fibers, viz. two-dimensional
chains of titania octahedra, wherein every fourth layer has oxygen atoms missing
to accommodate the loss in stoichiometry. The oxygen-deficient density of Ti4O7 is
double that of Ti8O15. This higher level of oxygen deficiency entails more changes
in the materials’ morphology and structure. Intermediate H2Ti3O7 NWs often are
closely packed, and during high-temperature exposures, these bundles may merge
to form large-diameter masses. Nevertheless, areas of the NWs with numerous
defects still break off, generating short thick fibers.
11 One-Dimensional Oxygen-Deficient Metal Oxides 259

Fig. 11.11 (a) Low-magnification TEM image of the TiO2 NWs, showing the high density of
nanocavities; (b) high-resolution image viewed along [100] direction, showing polyhedral
nanocavities in the NW; (c) core-loss EELS spectra from the center of the nanocavity and the
body of the NW

Figure 11.12a shows the UV–visible diffuse reflectance spectra of the starting
trititanate NWs, the TiO2 NWs with nanocavities, and the Ti8O15 NWs and the Ti4O7
fibers. Very different from the starting trititanate NWs and TiO2 NWs, whose absorp-
tion band covers only the UV region, the optical absorption bands of Ti8O15 and Ti4O7
cover the full range of visible light wavelengths and extend into the near-IR region. This
difference is attributed to the different ionization states of oxygen vacancies [42].
Figure 11.12b, c shows the I–V curves of the Ti8O17 and Ti4O7 samples,
respectively, measured at room temperature; clearly, both samples exhibit high
electrical conducting behavior at room temperature. The resistances of the Ti8O17
and Ti4O7 samples, respectively, are 9.74 and 0.965 O at room temperature;
their corresponding electrical conductivities at room temperature are 0.236 and
260 W.-Q. Han

Fig. 11.12 (a) UV–visible diffuse reflectance spectra of the starting H2Ti3O7 NWs, the products
of TiO2, Ti8O15, and Ti4O7. Room temperature I–V curves of (b) Ti8O15; (c) Ti4O7; and (d) anatase
TiO2

10.36 S/cm, respectively. The electrical conductivity of the Ti8O15 and Ti4O7
samples also were measured by dipping them into liquid nitrogen, starting from
room temperature. The values decreased by about four orders of magnitude to
2.4  105 and 4  103 S/cm, respectively, reflecting the lesser thermal excitation
of electrons at lower temperatures. The electrical resistivity of bulk Ti4O7 was
recorded previously. Ti4O7 shows two conductivity transitions, a semiconductor–
semiconductor one in the range of 130–140 K, and a semiconductor–metal one at
150 K [50]. For both transitions, there is steep increase in the electrical conductivity
with increasing temperature. In the metallic high-temperature phase, 3d electrons of
Ti ions are delocalized and contribute to the electrical conductivity. Below 150 K,
the 3d electrons are localized to form covalently bonded Ti3+–Ti3+ pairs. Since pair
formation involves lattice displacement, these pairs may be considered as two-
particle polarons, or bipolarons. While the ordered arrangement of the Ti3+ pairs is
established in the semiconducting low-temperature phase (T < 130 K), this
long-range order vanishes during the intermediate temperature phase (130 K < T
< 150 K) and then is described as a liquid state of the bipolarons [51].
11 One-Dimensional Oxygen-Deficient Metal Oxides 261

The electrical conductivity of TiO2  x depends strongly on the conditions of


treatment that engender different oxygen deficiencies, suggesting that such
deficiencies play a crucial part in electrical conductivity. TinO2n  1 can be consid-
ered as being made up of (n  1) TiO2 octahedra and one TiO octahedron. Rutile (or
anatase) TiO2 is a well-known wide-band gap semiconductor. In contrast to TiO2,
its crystal structure of TiO is based on a NaCl structure with ordered vacancies in
both the metal and the oxygen-sublattices; one-sixth of the titanium atoms and one-
sixth of the oxygen atoms are missing. The existence of the vacant sites within the
TiO structure is thought to permit sufficient contraction in the lattice such that 3d
orbitals on titanium overlap, thereby broadening the conduction band and allowing
electronic conduction [52]. The metallic behavior of TiO is responsible for the
metallic conductivity of TinO2n  1. Ti4O7 has a much higher TiO density than that
of other TinO2n  1, including Ti8O15, and thus, Ti4O7 is the most highly conductive
phase. This is consistent with the above measurement.

11.4 Substoichiometric Chromium Oxide


Nanobelts with Modulation Structures

The physical and chemical properties of transition metal oxides can be changed
dramatically because the transition metal elements can have different oxidation
states. The Cr atom has a 4s13d5 configuration. Chromium, with different valence
charges, forms different chromium oxides, such as CrO3, Cr2O5, CrO2, and Cr2O3.
Chromium dioxide (CrO2) is a half-metallic ferromagnet with almost 100% spin
polarization at the Fermi level such that electron conduction only occurs with one
orientation of electron spin; the other spin direction is insulating. CrO2 has a Curie
temperature of 397 K and thus offers promising applications in magnetic tunneling
and spin injection devices [53]. However, CrO2 is a metastable oxide that readily
decomposes into thermodynamically stable chromia (Cr2O3), which is an antifer-
romagnetic insulator with a Néel temperature of 307 K and is suitable as a tunnel
junction barrier both below and above this temperature [54]. For antiferromagnetic
nanomaterials, the surface spins increase as the size declines. The effect of uncom-
pensated spins on their surfaces is an important factor that may affect their magnetic
properties. Néel suggested that very fine particles of an antiferromagnetic material
should exhibit particular magnetic properties such as superparamagnetism and
weak ferromagnetism [55].
Size and shape are two important factors affecting the nanostructures’ properties
and performances. Chromia NPs and nanopores were prepared by several methods,
including a sol–gel process, gas condensation, microwave plasma, a sonochemical
reaction, laser-induced pyrolysis, hydrazine reduction followed by thermal treat-
ment, mechanochemical processing, urea-assisted homogeneous precipitation, and
a precipitation–gelation reaction [56]. In this section, I introduce a facile way to
prepare chromia NBs and NWs, and describe two types of superlattice structures
related to oxygen deficiency.
262 W.-Q. Han

Fig. 11.13 (a) SEM image showing the typical morphology of growth of NBs and NWs on the Cr
substrate; (b) XRD spectrum of rhombohedral Cr2O3 NBs. Cr (110) peak originated from Cr
substrate; (c) high-resolution image. The inset is the corresponding electron diffraction patterns
with beam parallel to the [001] direction. The electron diffraction patterns are indexed as
rhombohedral Cr2O3
The easy synthesis route for the chromia NBs and NWs is as follows: First, a
piece of chromium was inserted into the hot zone of a quartz tube horizontal
furnace. Argon was used as the protective gas during the rising temperature stage.
Once the temperature reached 800 C, an ethanol vapor is introduced while keeping
the temperature steady for 1 h. Next, the flow of ethanol vapor was stopped but
argon flow was maintained until the sample naturally cooled down [57].
Figure 11.13a is the SEM image taken from the as-produced sample; most of the
products are NBs, with some NWs. The lengths of the NBs are several micrometers
and their width ranges from tens nanometers to several hundred nanometers;
their thickness is about a few nanometers to tens of nanometers. In general, the
width and thickness of an individual NB are quite uniform. Interestingly, NBs with
large roots and a folded shape along the axis direction also were observed. An X-ray
spectrum (Fig. 11.13b) of the NBs indicates that peaks can be indexed as
rhombohedral Cr2O3 (JCPDF# 38-1479, space group R-3c (167), a ¼ 0.495876 nm
and c ¼ 1.35942 nm).
Figure 11.13c is the high-resolution image of a single-crystalline NB. The inset
is the corresponding electron diffraction pattern from the NB with the beam parallel
to the [001] direction. Tilting experiments confirm that the basic structure of the
NBs is rhombohedral with a ¼ 0.496 nm and c ¼ 1.359 nm. Many NBs are single
crystalline with the surface normal being the [001] direction.
11 One-Dimensional Oxygen-Deficient Metal Oxides 263

Fig. 11.14 High-resolution image (a) and electron diffraction pattern (b) from a NB with
modulation I. There is no modulation at the top left part as shown in the image and the top left
inset which is the Fourier transfer from the top-left area. The bottom-right part, however, exhibits
modulation I as shown in the image and the bottom-right inset which is the Fourier transfer from
the bottom-right area. The superlattices also are present in the diffraction (b); (c) high-resolution
image of modulation II with periodicity of 16.4 Å; (d) high-resolution image of modulation II with
a periodicity of 18.4 Å; (e) EELS spectra from NBs: A without modulation; B with modulation I; C
with modulation II; and D referenced micron-sized Cr2O3 particle

Figure 11.14a is a high-resolution image from another NB with the beam along
the [111] direction. The top left part in the image shows the basic lattice image of
Cr2O3, while the bottom right part exhibits additional modulation fringes as
indicated by the arrows. Moiré effects are ruled out since there are no slightly
mismatched or misorientated lattices, as shown in both the high-resolution image
(Fig. 11.14a) and the electron diffraction patterns (Fig. 11.14b). The diffraction
patterns (Fig. 11.14b) show the satellite spots, indicating the structure modulation
along the [110] direction. The modulation wave vector was determined to be
q1 ¼ 1/3(a* + b*), and the modulation periodicity in the real space to be 7.4 Å
(referred to as the modulation I). Besides this commensurate modulation along
the [110] direction, there is another incommensurate modulation along about the
[1-23]* (or [0-51]) direction (referred to as modulation II), as shown in Fig. 11.14c,
d. The modulation wave vector in Fig. 11.14c is determined as q2  0.133a*
 0.267b* + 0.401c*, and the periodicity in the real space to be 16.4 Å. The
modulation shown in Fig. 11.14d is similar to that in Fig. 11.14c but with a slightly
longer periodicity (18.4 Å) in real space. Most NBs have the second kind of
modulation structure, though the modulation periodicity among the different NBs
differs slightly.
264 W.-Q. Han

To determine the origin of the modulation, we performed EELS analysis for the
different NBs. Figure 11.14e(A–C) shows EELS spectra from the Cr2O3 NBs
without modulation, with the first kind of modulation, and second kind of modula-
tion, respectively. For comparison, a reference EELS spectrum was acquired
(Fig. 11.14e(D)) from a micron-sized Cr2O3 particle (Alfa-Aesar Co.). The spectra
were normalized to equalize their maximum intensities of Cr-L3. The quantitative
calculations show that the atomic ratio of oxygen to chromium for the NBs without
modulation (Fig. 11.14e(A)) is rO/Cr ¼ 1.47, close to that of the micron-sized Cr2O3
particles (~1.5). However, the rO/Cr of the NBs with the modulations usually is
markedly lower than that of micron-sized Cr2O3 particles. The rO/Cr of the NBs with
modulations I (Fig. 11.14e(B)) and II (Fig. 11.14e(C)) are 1.18 and 1.22, respec-
tively, indicating the oxygen deficiency in the nanobelts with modulation structure.
A series of EELS analysis shows that the composition of the NBs with the
modulation is Cr2O3  x with x ranging from 0.25 to 0.70. Interestingly, many
NBs with modulation have a composition close to Cr2O2.4. We also checked the
energy range related to C K-edge by EELS and do not find other elements in the
NBs. TEM studies show that about 80% NBs have the modulation structures.
Pop et al. reported an oxygen-deficient Cr2O2.4 powder that they produced [58].
They first synthesized an intermediate material Cr(OH)3gel using a sol–gel method:

2CrO3 þ 3C2 H5 OH ! 2Cr(OHÞ3 # þ3CH3 COH

Then, the gel was heated at 1,250 K in hydrogen atmosphere in order to stabilize
the oxygen vacancies. XRD study showed that the crystal structure of the
nonstoichiometric Cr2O3  x still is rhombohedral, the same as stoichiometric
Cr2O3, but with slight differences in the peaks’ relative intensity. The differences
arise from the oxygen deficiency in the nonstoichiometric compound due to a
modification of the structural factor |Fhkl|2, and a modification of the peak intensity.
The X-ray K-edge spectra revealed that the composition of the nonstoichiometric
Cr2O3  x is Cr2O2.4. Measurements of the temperature dependence of the reciprocal
susceptibility between 100 and 1,200 K disclosed an antiferromagnetic behavior,
with the Néel temperature at 318 K. For the nonstoichiometric sample, there is a
linear temperature dependence of the reciprocal magnetic susceptibility in
the paramagnetic region. The effective magnetic moment per chromium ion is
meff ¼ 4.86 mB, a big difference from 3.83 mB for a stoichiometric Cr2O3 [58].
Though the synthesis route, morphology, and microstructure described
here is different from those reported by Pop et al., many nanobelts with modulation
in the present work have a composition close to Cr2O2.4, The oxygen vacancies in the
nanobelts with modulation structures are ordered, and thus form the superstructures.
In the present work, the chemical reaction for the formation of chromia nanobelts can
be represented as [57]

2Cr þ ð1:5  0:5xÞO2 ! Cr2 O3x


11 One-Dimensional Oxygen-Deficient Metal Oxides 265

Oxygen here mainly comes from the decomposed ethanol at hightemperature.


Though the reactiontube was purged with 99.999% argon before heating, it still
may have contained some remnant oxygen. At high-temperature, Cr, like some
transition metals such as W [59], is very easily oxidized even at reduced atmo-
sphere. The presence of ethanol vapor ensures Cr can be oxidized, but under an
oxygen-deficient condition. To understand the role of ethanol, ethanol vapor is
replaced with water vapor or air, no NBs or NWs were formed, signifying that
ethanol plays an important role in their formation.
The Cr2O3  x NBs grow in the oxygen-deficient environment, so that most of
them display oxygen deficiency, as confirmed by the EELS measurements. The
ordered superstructures occur in the NBs that contained oxygen defects, or
vacancies, but not in NBs lacking oxygen defects. Hence, the modulations found
in the oxygen-deficient NBs reflect the ordering of the oxygen vacancies. This
finding is also consistent with the XRD measurements, since the X-ray scattering is
insensitive to oxygen. The oxygen deficiency chromia NBs with modulation
structures and folded shapes described here are expected to have distinct antiferro-
magnetic and catalytic properties.

11.5 Summary

1. 1D nanocerias (a mixture of NTs and NWs) were prepared by a mild hydrother-


mal reaction route, with cerium nitrate and ammonia hydroxide as reactants.
High precipitation temperature and prolonged aging times were keys for the
formation of tubular structure. The fraction of Ce3+/Ce4+ ions increased with
decreasing NT diameters. A high catalytic activity of pure and Pd-loaded 1D
nanoceria was observed in the WGS reaction at low temperature. Both ceria and
Pd played important roles during the WGS reaction process. While 1D nanoceria
was reduced easily at low temperature to produce oxygen vacancies, Pd can
activate the reduction of ceria. The special 1D feature could increase the
catalytic activity of nanoceria efficiently because the effective surface area is
extended by mitigating the problem of aggregation of NPs and whilst benefiting
from the double surfaces of NTs. Furthermore, the preferred exposed crystal
planes for 1D nanoceria are {100}/{110}, that is, the favored surfaces for CO
oxidation. The loss of the effective faceted surfaces of ceria during the WGS
process could be responsible for the loss of catalytic activity.
2. Ti8O15 NWs and Ti4O7 fibers were prepared by reducing H2Ti3O7 NWs at
different temperatures. Both samples show a high electrical conducting behavior
at room temperature and their absorption bands cover the full visible light region
and extend into the near IR region. Nanocavities embedded within anatase TiO2
NWs were generated by heating H2Ti3O7 NWs in an oxidation atmosphere at
650 C. The Ti L3/L2 ratio increases at the nanocavities, suggesting a decrease in
the Ti valence that might be due to the decline in O stoichiometry of the TiO2
around the nanocavity.
266 W.-Q. Han

3. Two types of unknown superlattice structures were found, related to oxygen


deficiency in chromium oxide NBs. The first, is a commensurate modulation
along the [1 1 0] direction, and the second is an incommensurate modulation
along about the [1 2 3]* (or [0 5 1]) direction. Cr2O2.4 NBs are produced by
simply heating a piece of chromium under an ethanol atmosphere.

Acknowledgments The author is grateful to Drs. L. J. Wu, W. Wen, J. C. Hanson, X.W. Teng
J. A. Rodriguez, and Y. M. Zhu in Brookhaven National Laboratory and Mr. Y. Zhang in SUNY at
Stony Brook for their cooperation and helpful discussions. This work is supported by the U. S.
DOE under contract DE-AC02-98CH10886 and Laboratory Directed Research and Development
Fund of Brookhaven National Laboratory.

References

1. Seebauer, E.G., Kratzer, M.C. (eds.): Charged Semiconductor Defects. Springer, London
(2009)
2. Han, W.Q., Fan, S.S., Li, Q.Q., Hu, Y.D.: Synthesis of gallium nitride nanorods through a
carbon nanotube-confined reaction. Science 277, 1287 (1997)
3. Dresselhaus, M.S., Chen, G., Tang, M.Y., Yang, R.G., Lee, H., Wang, D.Z., Ren, Z.F.,
Fleurial, J.P., Gogna, P.: New directions for low-dimensional thermoelectric materials. Adv.
Mater. 19, 1043 (2007)
4. Gur, I., Fromer, N.A., Geier, M.L., Alivisatos, A.P.: Air-stable all-inorganic nanocrystal solar
cells processed from solution. Science 310, 462 (2005)
5. Han, W.Q.: Anisotropic Hexagonal Boron Nitride Nanomaterials: Synthesis and Applications.
In: Kumar, C. (ed.) Mixed Metal Nanomaterials. Wiley, Weinheim (2009)
6. Trovarelli, A. (ed.): Catalysis by Ceria and Related Materials. Imperial College Press, London
(2002)
7. Twigg, M.V. (ed.): Catalyst Handbook, 2nd edn. Wolfe, London (1989)
8. Ghenciu, A.F.: Review of fuel processing catalysts for hydrogen production in PEM fuel cell
systems. Curr. Opin. Solid State Mater. Sci. 6, 389 (2002)
9. Shao, Z., Haile, S.M., Ahn, J., Ronney, P.D., Zhan, Z., Barnett, S.A.: A thermally self-
sustained micro solid-oxide fuel-cell stack with high power density. Nature 435, 795 (2005)
10. Tsunekawa, S., Ishikawa, K., Li, Z.Q., Kawazoe, Y., Kasuya, A.: Origin of anomalous lattice
expansion in oxide nanoparticles. Phys. Rev. Lett. 85, 3440 (2000)
11. Trovarelli, A.: Catal. Rev. Sci. Eng. 38, 439 (1996)
12. Fu, Q., Saltsburg, H., Flytzani-Stephanopoulos, M.: Active non-metallic Au and Pt species on
ceria-based water-gas shift catalysts. Science 301, 935 (2003)
13. Gorte, R.J., Zhao, S.: Studies of the water-gas-shift reaction with ceria-supported precious
metals. Catal. Today 104, 18 (2005)
14. Zhou, X.D., Huebner, W., Anderson, H.U.: Room-temperature homogeneous nucleation
synthesis and thermal stability of nanometer single crystal CeO2. Appl. Phys. Lett. 80, 3814
(2002)
15. Chen, H.I., Chang, H.Y.: Solid State Commun. 133, 593 (2005)
16. Yu, T., Joo, J., Park, J.Y.I., Hyeon, T.: Large-scale nonhydrolic sol-gel synthesis of uniform-
sized ceria nanocrystals with spherical, wire, and tadpole shapes. Angew. Chem. Int. Ed. 44,
7411 (2005)
17. Natile, M.M., Boccaletti, G., Glisenti, A.: Properties and reactivity of nanostructured CeO2
powders: comparison among two synthesis procedures. Chem. Mater. 17, 6272 (2005)
11 One-Dimensional Oxygen-Deficient Metal Oxides 267

18. Corma, A., Atienzar, P., Garcia, H., Chane-Chine, J.-Y.: Hierarchically mesostructured doped
CeO2 with potential for solar-cell use. Nat. Mater. 3, 394 (2004)
19. Han, W.Q., Wu, L.J., Zhu, Y.M.: Formation and oxidation state of CeO2-x nanotubes. J. Am.
Chem. Soc. 127, 12814 (2005)
20. Yamashita, M., Kameyama, K., Yabe, S., Yoshida, S., Fujishiro, Y., Kawai, T., Sato. T.:
J. Mater. Sci. 37, 683 (2002)
21. Wu, L., Wiesmann, H.J., Moodenbaugh, A.R., Klie, R.F., Zhu, Y., Welch, D.O., Suenaga, M.:
Oxidation state and lattice expansion of CeO2-x nanoparticles as a function of particle size.
Phys. Rev. B 69, 125415 (2004)
22. Wang, X., Rodriguez, J., Hanson, J., Gamarra, D., Martinez-Arias, A., Fernandez-Garcia, M.:
Unusual physical and chemical properties of Cu in Ce1-xCuxO2 oxides. J. Phys. Chem. B 109,
19595 (2005)
23. Rodriguez, J.A., Wang, W.Q., Hanson, J.C., Liu, G., Iglesia-Juez, A., Fernández-Garcia, M.:
The behavior of mixed-metal oxides: Structural and electronic properties of Ce1-xCaxO2 and
Ce1-xCaxO2-x. J. Chem. Phys. 119, 5659 (2003)
24. Han, W.Q., Wen, W., Ding, Y., Liu, Z.X., Maye, M.M., Lewis, L., Hanson, J., Gang, O.: Fe-
Doped trititanate nanotubes: formation, optical and magnetic properties, and catalytic
applications. J. Phys. Chem. C 111, 14339 (2007)
25. Wen, W., Liu, J., White, M.G., Marinkovic, N., Hanson, J.C., Rodriguez, J.A.: In situ time-
resolved characterization of novel Cu-MoO2 catalysts during the water-gas shift reaction.
Catal. Lett. 113, 1 (2007)
26. Rodriguez, J.A., Wang, X., Liu, P., Wen, W., et al.: Gold nanoparticles on ceria: importance of
O vacancies in the activation of gold. Top. Catal. 44, 73 (2007)
27. Sadi, F., Duprez, D., Gerard, F., Miloudi, A.: Hydrogen formation in the reaction of steam with
Rh/CeO2 catalysis: a tool for characterising reduced centres of ceria. J. Catal. 213, 226 (2003)
28. Han, W.Q., Wen, W., Hanson, J.C., Teng, X., Marinkovic, N., Rodriguez, J.A.: One-dimensional
ceria as catalyst for the low-temperature water-gas shift reaction. J. Phys. Chem. C 113, 21949
(2009)
29. Sohlberg, K., Pantilides, S.K., Pennycook, S.J.: Interactions of hydrogen with CeO2. J. Am.
Chem. Soc. 123, 6609 (2001)
30. Wang, Y., Sun, S.N., Chou, M.Y.: Total-energy study of hydrogen ordering in PdHx(0  x
 1). Phys. Rev. B 53, 1 (1996)
31. McCaulley, J.A.: In-situ x-ray absorption spectroscopy studies of hydride and carbide forma-
tion in supported palladium catalyst. J. Phys. Chem. 97, 10372 (1993)
32. Bianconi, A.: XANES Spectroscopy. In: Koningsberger, D.C., Prins, R. (eds.) X-ray Absorp-
tion: Principles, Applications, Techniques of EXAFS, SEXAFS and XANES, pp. 573–662.
Wiley, New York (1988)
33. Ranganathan, E.S., Bej, S.K., Thompson, L.T.: Methanol steam reforming over Pd/ZnO and
Pd/CeO2 catalyst. Appl. Catal. A 289, 153 (2005)
34. Fu, Q., Kudriavtseva, S., Saltsburg, H., Flytzani-Stephanopoulos, M.: Gold-ceria catalysts for
low-temperature water-gas shift reaction. Chem. Eng. J. 93, 41 (2003)
35. Zalc, J.M., Sokolovskii, V., Loffler, D.G.: Are noble metal based water-gas shift catalysts
practical for automotive fuel processing? J. Catal. 206, 169–171 (2002)
36. Gouet, A., Meunier, F., Breen, J.P., Burch, R., Petch, M.I., Faur-Ghenciu, A.: Study of the
origins of the deactivation of a Pt/CeO2 catalyst during reverse water gas shift (RWGS)
reaction. J. Catal. 226, 382 (2004)
37. Wang, X., Gorte, R.J., Wagner, J.P.: Deactivation mechanisms for Pd/Ceria during the
water-gas-shift reaction. J. Catal. 212, 225 (2002)
38. Mai, H.M., Sun, L.D., Zhang, Y.W., Si, R., Feng, W., Zhang, H.P., Liu, H.C., Yan, C.H.: Shape
selective synthesis and oxygen storage behavior of ceria nanopolyhedra, nanorods, and
nanocubes. J. Phys. Chem. B 109, 24380 (2005)
39. Honda, K., Fujishima, A.: Electrochemical photolysis of water at a semiconductor electrode.
Nature 238, 37 (1972)
268 W.-Q. Han

40. Gr€atzel, M.: Review article photoelectrochemical cells. Nature 414, 338 (2001)
41. Diebold, U.: The surface science of titanium dioxide. Surf. Sci. Rep. 48, 53 (2003)
42. Cronemeyer, D.C.: Electrical and optical properties of rutile single crystals. Phys. Rev. 87, 876
(1952)
43. Magnéli, A.: Non-stoichiometry and structural disorder in some families of inorganic
compounds. Pure Appl. Chem. 50, 1261 (1978)
44. Anderson, S., Collen, B., K€ uylenstierna, U., Magnéli, A.: Phase analysis studies on the
titanium-oxygen system. Acta Chem. Scand. 11, 1641 (1957)
45. Smith, J.R., Walsh, F.C., Clarke, R.L.: Electrodes based on Magneli phase titanium oxides: the
properties and applications of Ebonex® materials. J. Appl. Electrochem. 28, 1021 (1998)
46. Kasuga, T., Hiramatsu, M., Hoson, A., Sekino, T., Niihara, K.: Titania nanotubes prepared by
chemical processing. Adv. Mater. 11, 1307 (1999)
47. Chen, Q., Zhou, W., Du, G.H., Peng, L.M.: Titranate nanotubes made via a single alkali
treatment. Adv. Mater. 14, 1208 (2002)
48. Han, W.Q., Zhang, Y.: Magneli phases TinO2n-1 nanowires: formation, optical, and transport
properties. Appl. Phys. Lett. 92, 203117 (2008)
49. Han, W.Q., Wu, L., Klie, R.F., Zhu, Y.: Enhanced optical absorption induced by dense
nanocavities inside titania nanorods. Adv. Mater. 19, 2525 (2007)
50. Bartholemew, R.F., Frankl, D.F.: Electrical properties of some titanium oxides. Phys. Rev.
187, 828 (1969)
51. Lakkis, S., Schlenker, C.B., Chakraverty, B.K., Buder, R.: Metal-insulator transitions in Ti4O7
single crystals: crystal characterization, specific heat, and electron paramagnetic resonance.
Phys. Rev. B. 14, 1429 (1976)
52. Smart, L., Moore, E. (eds.): Solid State Chemistry. Chapman & Hall, New York (1992)
53. Kamper, K.P., Schmitt, W., Guntherodt, G., Gambino, R.J., Ruf, R.: CrO2 – A new half-
metallic ferromagnet? Phys. Rev. Lett. 59, 2788 (1987)
54. Kobylinski, T.P., Taylor, B.W.: The catalytic chemistry of nitric oxide: I. The effect of water
on the reduction of nitric oxide over supported chromium and iron oxides. J. Catal. 31, 450
(1973)
55. Coey, J.D.M., Berkowitz, A.E., Balcells, L., Putris, F.F., Barry, A.: Magnetoresistance of
chromium dioxide powder compacts. Phys. Rev. Lett. 80, 3815 (1998)
56. Kawabata, A., Yoshinaka, M., Hirota, K., Yamaguchi, O.: Hot isostatic pressing and charac-
terization of sol-gel-derived chromium(III) Oxide. J. Am. Ceram. Soc. 78, 2271 (1995)
57. Han, W.Q., Wu, L., Stein, A., Zhu, Y., Misewich, J., Warren, J.: Oxygen-deficiency-induced
superlattice structures of chromia nanobelts. Angew. Chem. Int. Ed. 45, 6554 (2006)
58. Pop, I., Andrecut, M., Burda, I., Andrecut, C., Pop, O., Ivan, I., Nazarenco, I., Oprea, C.: X-ray
K edge absorption and magnetic behaviour of chromium in stoichiometric and
nonstoichiometric a-Cr2O3. Mater. Chem. Phys. 47, 85 (1997)
59. Gu, G., Zhang, B., Han, W.Q., Roth, S., Liu, J.: Tungsten oxide nanowires on tungsten
substrates. Nano Lett. 2, 849 (2005)
Chapter 12
Oxide Nanostructures for Energy Storage

Yuan Yang, Jang Wook Choi, and Yi Cui

12.1 Introduction

Energy is becoming a critical societal issue nowadays, which greatly impacts world
economy, environment, and human life. The global energy consumption will keep
growing in the following decades. Although traditional combustion-based energy
technologies, including coal, oil, and natural gas, dominate in energy needs,
alternative energy sources and technologies are imperative due to the disadvantages
of their traditional counterparts, such as the emission of greenhouse gases and long-
term environmental consequences. Moreover, the rapid depletion of fossil fuels
could result in severe economic and political conflictions between countries. As a
result, solving energy issues becomes one of the greatest challenges in the twenty-
first century. To conquer this challenge, alternative sustainable and clean energy
sources and technology should be explored and utilized. In general, energy tech-
nology can be divided into two categories, energy conversion and energy storage.
The former one deals with how to efficiently convert energy from one form to
another, such as solar cells, wind power plants, and nuclear power plants, which
convert various kinds of energy to electricity. The latter one stores energy in a
steady state. The most widely used energy storage devices include batteries and
electrochemical capacitors. Energy can be stored chemically in these devices for a
long time and can be converted directly into electrical energy with little or no
impact on the environment. Furthermore, the efficiency of such devices could go
beyond the Carnot limit.
Batteries and electrochemical capacitors have wide applications, such as porta-
ble devices, power tools, and electric vehicles. However, devices with better

Y. Yang • J.W. Choi • Y. Cui (*)


Department of Materials Science and Engineering, Stanford University,
Stanford, CA, USA
e-mail: yicui@stanford.edu

J. Wu et al. (eds.), Functional Metal Oxide Nanostructures, Springer Series 269


in Materials Science 149, DOI 10.1007/978-1-4419-9931-3_12,
# Springer Science+Business Media, LLC 2012
270 Y. Yang et al.

performance, such as higher energy density, longer cycle life, and faster charge/
discharge rate, are still desired or necessary for certain applications. Design of new
materials and techniques are crucial to achieving these goals. In view of this,
nanostructured materials and nanotechnology offer great promise because of the
unusual properties resulting from confining their dimensions and the combination
of bulk and surface properties to the overall behavior. Dramatic improvement in
power rate, cycling life, energy density, and other aspects has been observed [1, 2].
In the following part, progress in nanosized materials, especially nanostructured
oxides, for lithium-ion batteries and electrochemical capacitors will be presented.

12.2 Nano Oxides for Li-Ion Batteries

Rechargeable batteries play an important role in energy storage and have


found applications at different scales [3], including on-chip power supplies
(106–103 W h), portable devices (100–102 W h), vehicle electrification
(104–105 W h), and storage for electric grid and buildings (106–108 W h). Among
different types of rechargeable batteries, lithium-ion batteries can offer the highest
energy density and power density and show potential for further improvement [4].
As a result, much attention has been given to lithium-ion batteries. In commercial
Li-ion batteries, LiCoO2 and graphite are the most common cathode and anode,
respectively. As shown in Fig. 12.1, LiCoO2 and graphite both have a layered
structure and lithium ions can move readily in the two-dimensional interlayer
spacing. During charge, lithium is removed from LiCoO2 and intercalates into
graphite. The reverse process occurs during discharge. The practical capacity of
LiCoO2 is 140 mAh/g, as only half the lithium can be used to avoid cracking of the
layer structure. Graphite has a practical capacity of 320–360 mAh/g, which is close
to the theoretical capacity of 372 mAh/g, corresponding to the conversion of
graphite to the LiC6 phase [5]. The voltage of LiCoO2 and graphite are ~3.9 V
and ~0.1 V vs. Li/Li+, respectively. More details on the operation principles of
Li-ion batteries can be found in other references [3–6].
The LiCoO2/graphite system has been successfully utilized in the field of
portable electronics, such as cell phones and laptops. However, lithium-ion

Fig. 12.1 Illustration of the


structure of a lithium-ion
battery with graphite and
LiCoO2 as the anode and
cathode, respectively [1]
12 Oxide Nanostructures for Energy Storage 271

batteries are still far from optimization, and significant improvement is necessary
for use in applications such as vehicle electrification. Necessary improvements
include higher energy density, faster discharge/charge rates, longer lifetimes, better
environmental benignity, and safety. Much research has been dedicated to
accomplishing these goals by either synthesizing new materials or designing new
material structures and morphologies [7].
Scaling the size of existing materials down to the nanoscale offers a variety of
advantages. The relatively small size of nanomaterials leads to shorter diffusion
length and higher surface area for reaction, which can enhance the kinetic behavior
and thus the power density of batteries. Generally, nanomaterials can also accom-
modate larger stress without fracturing, which enables the use of new materials that
undergo large volume changes during reaction with lithium, such as silicon and tin.
Moreover, many exotic nanostructures have been created, including nanowires
[8–10], nanorings [11, 12], and mesoporous materials [13, 14], providing various
choices for the rational design of electrodes. For example, mesoporous LiMn2O4 as a
cathode material could trap dissolved Mn ions inside the porous network to suppress
capacity fading, which is the bottleneck for this material [15, 16]. As another
example, silicon nanowires used as an anode material provide one-dimensional
electron transport pathways and facile relaxation of stress along the radial direction
[17, 18]. These examples illustrate the promise of nanomaterials used in Li-ion
batteries.
It is worth pointing out that nanomaterials also have certain disadvantages in
battery applications. For instance, the high surface area of nanomaterials also
promotes side reactions, which have a parasitic effect on a battery’s performance.
Also, the density of nanopowder films is typically less than that for micrometer-
sized particles. As a result, the volumetric energy density is lower in nanomaterials
based electrodes compared to conventional battery electrodes, which usually con-
tain micron-sized powder. However, these problems may be minimized by utilizing
specific techniques, such as surface coatings to reduce the side reactions [19]. In the
following sections, we will review recent progress in the development of
nanostructured oxides for lithium batteries.

12.2.1 Spinel LiMn2O4

Spinel LiMn2O4 is a promising cathode material due to its low cost, low toxicity,
and high stability with respect to thermal runaway. The typical capacity of this
material is 110–120 mAh/g [6]. LiMn2O4-based battery products have been
fabricated [20, 21] and several companies are exploring its use as cathode material
for batteries designed for electric vehicles. However, its power density is not high
since the diffusivity of lithium ions in LiMn2O4 (109–1011 cm2/s) [22, 23] is
much lower than their diffusivity in LiCoO2 (107–109 cm2/s) [24, 25]. As a
result, shrinking the size of LiMn2O4 down to the nanoscale can shorten the
diffusion distance of lithium and result in enhanced power density. Recent work
272 Y. Yang et al.

b c 140
4.5 Discharge capacity/mAhg−1
120
Potentail/V (vs. Li/Li+)

Our sample 0.1A/g


4
Our sample 5A/g 100
3.5
Our sample 10A/g 80
3 60
Our sample 20A/g
2.5 40
Aldrich 20A/g
2 Honjyo 20A/g 20
Mitsui 20A/g
1.5 0
0 20 40 60 80 100 120 0 5 10 15 20
Discharge capacity/mAhg−1 Current density/Ag−1

Fig. 12.2 (a) A SEM image of as-synthesized single crystalline LiMn2O4 nanowires. (b) The
discharge curve of LiMn2O4 nanowire-based cathodes and commercial samples at different
current rates. (c) The rate-dependent second discharge capacity of LiMn2O4: single crystalline
nanowires (red circle), Honjyo Chemical (green box), Mitsui Metal (blue triangle), and Aldrich
(black box) [30]

has shown that the power performance of LiMn2O4 can be improved by utilizing
nanostructures, such as nanowires, nanorods, and nanoparticles [26–30].
Hosono et al. have synthesized LiMn2O4 nanowires by a three-step method that
consists of hydrothermal synthesis of Na0.44MnO2 nanowires, ion exchange with
lithium ions in molten salt, and high-temperature annealing [30]. The as-
synthesized nanowires have a diameter of 50–100 nm and are hundreds of
micrometers in length (Fig. 12.2a). These nanowires exhibit excellent high-power
performance: the discharge capacity is 118 mAh/g at 0.1 A/g (~0.7 C), and when the
current rate increases to 20 A/g (charge/discharge in ~16 s), 75% of the initial
capacity (88 mAh/g) still remains (Fig. 12.3b). In comparison, commercial micron-
sized LiMn2O4 particles show capacity less than 40 mAh/g at 20 A/g. Besides
nanowires, LiMn2O4 nanorods also show much better power performance than
micron-sized LiMn2O4 particles [27].
12 Oxide Nanostructures for Energy Storage 273

Fig. 12.3 (a) SEM images of nanorod devices in the organic electrolyte (1 M LiPF6 in EC/DEC).
(b) The evolution of the normalized conductance of nanorods in the electrolyte [44]

Another issue related to the use of LiMn2O4 as a cathode material is its fast
capacity fading, especially at high temperatures (e.g., 50–60 C). There are three
primary reasons for this fade in capacity [31–33]. The first is the dissolution of Mn2+
ions into the electrolyte after the disproportionation of LiMn2O4 [34, 35]. The trace
amount of water inside the electrolyte can react with LiPF6 to form HF. The acidic
environment promotes the disproportionation of Mn3+ ions in LiMn2O4 (Mn3+ !
Mn2+ + Mn4+). The second reason is the decomposition of the electrolyte, which also
generates H2O and leads to the disproportionation of LiMn2O4 [32, 36]. The last
reason is Jahn–Teller distortion. It should be noted that the high surface area inherent
in nanomaterials can enhance the dissolution of Mn ions and side reactions with
274 Y. Yang et al.

the electrolyte, which lead to capacity fading. These facts raise concerns regarding
the use of LiMn2O4 nanostructures in cathodes even though better power perfor-
mance can be realized. However, recent research has shown that nanostructured
LiMn2O4 can also exhibit good cycling performance [15, 16, 28, 29]. Though the
reasons are not yet clear, some have reported that the good cycling data might result
from the high crystallinity of the incorporated nanostructures; the degree of crystal-
linity plays an important role in the stability of LiMn2O4 [37, 38]. For example,
stoichiometric LiMn2O4 nanoparticles show 83% capacity retention after 200 cycles
at 55 C [29], which is much better than some reports on micron-sized particles [39].
Moreover, excellent cycling retention has been demonstrated in mesoporous
LiMn2O4 as dissolved Mn ions can be trapped inside the porous materials. This
study shows that at room temperature, 94% of the initial capacity is retained after 500
cycles and the fading per cycle is less than 0.005% after the first 100 cycles [15]. At
50 C, the capacity retention is over 95% after 100 cycles [16].
In addition to providing better performance, nanostructures also facilitate the use
of new methods to study battery materials. Common techniques used to study battery
materials are based on the electrochemical testing of ensemble electrodes. However,
the heterogeneous nature of ensemble electrodes averages all information and cannot
provide a direct correlation of electrochemical properties with the local morphology,
structures and chemical composition. Investigations at the single-particle level can
significantly expand the scope of understanding for specific battery materials. The
difficulties of studying a single particle include distinguishing a particle in the mixed
film-like electrode and contacting probes, such as metallic electrodes, onto such small
particles (typically less than 20 mm for battery materials). Nanostructures, especially
one-dimensional nanorods and nanowires, can solve these problems due to the
following reasons: (1) In contrast to micrometer-sized particles, it is feasible to
make contact with metallic probes to study their transport properties [40]. (2) The
single-nanostructure devices can be characterized by a variety of electron microscopy
and in situ techniques [40, 41]. (3) Nanomaterials can be highly crystalline [42, 43],
which provides well-defined nanoscale domains for testing intrinsic properties.
Recently, direct observation of the dissolution of LiMn2O4 nanostructures in an
electrolyte has been observed [44]. Electron beam lithography was used to deposit
metallic electrode contacts to fix and identify single LiMn2O4 nanorods. SEM images
show that pure LiMn2O4 is etched in organic electrolytes (1 M LiPF6 in EC/DEC)
after a few hours at 60 C while Al-doped LiMn2O4 is much more stable, as shown in
Fig. 12.3a. I–V measurements also record that the resistivity of pure LiMn2O4
nanorods increase significantly while that of the Al-doped sample remains basically
the same (Fig. 12.3b). These results correlate well with their electrochemical perfor-
mance: pure LiMn2O4 nanorods show only 69% capacity after 100 cycles at
55 C while LiAl0.1Mn1.9O4 nanorods exhibit a capacity retention of 80% after 100
cycles [44].
12 Oxide Nanostructures for Energy Storage 275

12.2.2 Manganese Dioxide

Spinel LiMn2O4 is favored as a cathode material because of its low cost, environmental
compatibility, and superior safety characteristics. However, the capacity of this
material is limited to ~120 mAh/g. Meanwhile, manganese dioxide (MnO2),
which is widely used in primary lithium batteries [45], could accommodate up to
~0.7 lithium [46], corresponding to a capacity of ~210 mAh/g. However, due to
structural changes and volume expansion, the reaction between lithium and MnO2 is
not fully reversible, which makes this material unsuitable for rechargeable batteries.
Recently, research has shown that the reversibility and capacity of MnO2 can be
improved by utilizing nanostructured MnO2, which is likely due to better accommo-
dation of strain during structural changes.
Manganese dioxide has many different polymorphs, including the a, b, g, d, e,
and l phases. The structures of MnO2 polymorphs can be described as various
stacking arrangements of linked Mn–O octahedra, as shown in Fig. 12.4 [47]. One-
dimensional tunnels exist in the a, b, and g phases. The tunnel sizes are different for
different phases of MnO2; these tunnel sizes largely determine the electrochemical
properties of the material. For example, b-MnO2 contains the so-called 1  1
tunnel, which is too small for the transport of lithium ions. As a result, this phase
is commonly considered inactive towards lithium [48]. The a and g phases have
2  2 and 1  2 channels, respectively [47]. The corresponding channels are of
sizes that are suitable for lithium transport, and these phases are electrochemically
active in lithium-based batteries [49]. Besides pure MnO2, there is another family of
layered MnO2-related materials, AxMnO2 (A ¼ H+ or cations). These materials
have a layered structure similar to d – MnO2 (birnessite) with a basal spacing of
~0.7 nm, allowing high mobility of the interlayer cations with fast kinetics and
slight structural changes [50].
MnO2 nanostructures have been synthesized by various methods, especially
hydrothermal reactions, and they exhibit good electrochemical performance
[49–55]. For example, a and g-MnO2 nanorods have been synthesized and show
reversible capacity [49]. Ma et al. synthesized nanorods of layered MnO2
(birnessite) nanobelts [50]. The nanobelts are 5–15 nm in width and several to
tens of micrometers in length. The initial discharge capacity reaches 375 mAh/g in
the voltage range of 1.0–4.8 V vs. Li/Li+, corresponding to 1.3 lithium per unit of
MnO2. A considerable capacity loss (85 mAh/g) was observed in the second cycle,
which is similar to bulk birnessite. However, the capacity loss of the following
cycles is only ~0.7% per cycle, much less than bulk birnessite [50]. Moreover, the
layered MnO2 nanobelts do not undergo the phase transformation to spinel structure
(l phase) during cycling, which occurs in bulk birnessite MnO2 [56] and should be
avoided in practical applications.
b-MnO2 is the most stable phase at room temperature [57, 58], but it can only
accommodate a small amount of lithium due to the narrow 1  1 channels. However,
Feng Jiao and Peter G. Bruce reported that a considerable amount of lithium can be
reversibly inserted into b-MnO2 by utilizing mesoporous structures [48]. In their
276 Y. Yang et al.

Fig. 12.4 Schematics of a, b, g, d, and l-MnO2. a, b, and g-MnO2 present channels with different
sizes. d-MnO2 has a layered structure. l-MnO2 has a cubic structure [47]

study, mesoporous MnO2 is formed by using a KIT-6 mesoporous silica template


(space group Ia3d). The as-synthesized mesoporous MnO2 has a highly ordered
three-dimensional pore structure (Fig. 12.5) with a lattice parameter of 24.9 nm.
The reported pore size was 3.65 nm and the Brunauer–Emmett–Teller (BET) surface
area was 127 m2/g. The mesoporous MnO2 exhibits an initial discharge capacity of
284 mAh/g (0.92 lithium per MnO2) at 15 mA/g. After some initial decay, the
capacity stabilizes around 200 mAh/g. Furthermore, the capacity only decreases by
19% when the rate is increased from 15 to 300 mA/g. In contrast, bulk MnO2 shows a
capacity less than 10 mAh/g, indicating very poor electrochemical activity.
12 Oxide Nanostructures for Energy Storage 277

Fig. 12.5 Characterization of mesoporous b-MnO2. Left: TEM images of (a) and (b) as-prepared
mesoporous b-MnO2; (c) and (d) after first discharge. Right: capacity retention for mesoporous
b-MnO2 cycled at (a) 15 mA/g, (b) 30 mA/g, (c) 300 mA/g, and (d) bulk b-MnO2 cycled at
16 mA/g [48]

12.2.3 Vanadium Pentoxide

Vanadium pentoxide (V2O5) was one of the first oxides investigated for use in
rechargeable lithium batteries [6]. Its orthorhombic crystal structure can be
visualized as layers (ab plane) of VO5 square pyramids that share edges and
corners. The sixth V–O bond in the c-direction consists of weak electrostatic
interactions, which facilitates the insertion of lithium ions between layers [59]. In
theory, this material can intercalate up to three lithium ions per unit of V2O5,
corresponding to a capacity of about 440 mAh/g. However, the drastic structural
transformation during intercalation is complicated and not fully reversible [60].
Several phases of LixV2O5 that contain different amounts of lithium have been
observed. The a and e phases, with x < 0.1 and 0.35 < x < 0.7, respectively,
maintain the orthorhombic phase while the interlayer distance increases with
increasing x. At x ¼ 1, the d-phase exists, resulting from one layer gliding out of
two. At x > 1, the irreversible g-phase is formed. When x is further increased to 3,
the structure of LixV2O5 transforms to the rock salt o-phase. The delithiation of
Li3V2O5 is limited to the o-phase and is not fully reversible in bulk materials,
where x cannot reach lower than 0.4.
While fully reversible insertion of lithium in the range of 0 < x < 3 cannot be
accomplished in bulk V2O5, studies show that it is likely to be realized in
nanostructured V2O5. For example, Chan et al. synthesized V2O5 nanoribbons by
chemical vapor deposition and observed that the chemical lithiation process is fully
reversible in the nanoribbons [59]. These nanowires grew along the <020> direc-
tion and the c-axis is parallel to the height. In their study, n-butyllithium is used to
chemically lithiate the nanoribbons while Br2 is employed as the reagent for
delithiation. This chemical process is considered to give nearly the same results
278 Y. Yang et al.

Fig. 12.6 Chemical lithiation in V2O5. (a) Schematic of the insertion/deinsertion process in
nanoribbon and bulk materials. (b) TEM image of as-synthesized nanoribbons. (c) TEM image
of LixV2O5 nanoribbons treated with Br2. Inset: electron diffraction shows the orthorhombic
structure of pristine V2O5, indicating full delithiation [59]

as electrochemical lithiation and is used to determine the full lithiation capacity of


the nanoribbons [61]. Ex situ TEM techniques were used to track the phase
transformation. During lithiation, the width of nanoribbons affects the phase trans-
formation. When the width is on the order of hundreds of nanometers, o-Li3V2O5
dominates the diffraction pattern, but another intermediate phase, g-Li2V2O5, is
also identified. Meanwhile, when the width increases to several micrometers, the
pristine V2O5 phase and o-Li3V2O5 show similar intensity in the diffraction pattern
even after lithiation times of several days. These results indicate a diffusion barrier
that was more easily overcome in the narrower regions than the wider region. In the
process of delithiation, a complete removal of lithium by Br2 was observed and the
structure transforms back to the pristine orthorhombic V2O5 phase (Fig. 12.6). In
comparison, using Br2 can only remove the lithium to x ¼ 0.1 in bulk V2O5, and the
resulting material maintained a disordered cubic structure [60]. In addition, the
effect of the thickness of nanoribbons (the c-axis dimension) on the Li insertion was
studied, which was overlooked before as lithium is considered to diffuse along the
ab plane. When 100 nm thick and 400 nm wide nanoribbons were lithiated for only
10 s, only the pristine V2O5 phase was observed, suggesting no significant amount
of Li in the NR layers. In contrast, in a nanoribbon with a width of 740 nm but
smaller height (~20 nm), o-Li3V2O5 was clearly observed as the dominant phase.
These observations indicate that the thickness along the c-direction has a significant
impact on the phase transformation. The authors suggest that the small thickness helps
overcometheactivationbarrierofdistortionduringthesimultaneousstructuraltransfor-
mationfromtheorthorhombicphasetotherocksaltcubicstructure[59].
In another study, Patrissi et al. synthesized V2O5 nanowire films by a template-
assisted method [62]. The nanowire film exhibits a higher capacity and better power
performance than a V2O5 thin film made by a sol–gel method. The nanowire
film can accommodate about 0.05 more lithium (0.95 vs. 0.90) than the thin film.
12 Oxide Nanostructures for Energy Storage 279

Also, the capacity retention at a rate of 1,021 C is 40% compared to C/20 for the
nanowire film. In comparison, the capacity retention is less than 20 % at 793 C in a
V2O5 thin film. A V2O5/carbon nanotube nanocomposite was synthesized by
Sakamoto et al. and also showed improved rate performance [63].

12.2.4 Titanium Oxide

Titanium oxide has advantages for use in batteries such as ample availability, low
toxicity, and safety. Titanium oxide also has a theoretical capacity of 335 mAh/g or
1.0 lithium per TiO2 unit [64]. However, bulk TiO2 materials always exhibit limited
capacity and poor cycling performance, which is partially attributed to the low
diffusivity of lithium in TiO2 [64]. Recently, it has been found that nanostructured
TiO2 can enhance the capacity and reversibility of this material significantly [64].
There are many polymorphs of TiO2, such as rutile, anatase, brookite, and TiO2-
B (bronze), as summarized in Table 12.1. The typical charge/discharge voltage of
these phases is 1.5–2.0 V vs. Li/Li+. Rutile is believed to be the most stable phase,
but can only accommodate limited amounts of lithium (<0.1 Li per TiO2) [20, 68].
The diffusivity of lithium is anisotropic inside rutile TiO2: 106 cm2/s along the
c-axis but only 1015 cm2/s along the ab-plane [69–71]. Therefore, the low diffu-
sivity along the ab-plane restricts the reaction process and limits the intercalation of
lithium. In contrast, nanostructured rutile shows significantly improved capacity.
Hu et al. reported more than 0.8 lithium (about 270 mAh/g) insertion per TiO2 in
rutile nanorods (10 nm  40 nm) while only around 0.1 lithium (~30 mAh/g) can
be inserted into particles with diameter of 20 mm at a current rate of C/20, as shown
in Fig. 12.7 [72]. Moreover, the capacity remains around 220 mAh/g at a 1 C rate
and over 100 mAh/g at a 10 C rate. The success is mainly due to the very short
diffusion length along the ab plane in the synthesized nanorods. The ab plane is
parallel to the cross section of the nanorods and thus the diffusion length is only
~5 nm. Furthermore, the authors found that the surface storage of lithium is also
favored in nanoparticles.
The Bruce group has also investigated TiO2-B nanostructures [73–76]. TiO2-B
has a layer structure that is more open than the structures of rutile, anatase, or
brookite, making it a better host for Li+ intercalation [76]. TiO2-B nanowires and

Table 12.1 The structure of TiO2 polymorphs


Structure Space group Density (g/cm3) Unit cell (Å) Reference
Rutile P42/mnm 4.13 a ¼ 4.59, c ¼ 2.96 [64–66]
Anatase I41/amd 3.79 a ¼ 3.79, c ¼ 9.51 [64–66]
Brookite Pbca 3.99 a ¼ 9.17, b ¼ 5.46, c ¼ 5.14 [64, 66, 67]
TiO2 (B) C2/m 3.64 a ¼ 12.17, b ¼ 3.74, c ¼ 6.51, [64, 66, 67]
b ¼ 107.29
TiO2 (R) Pbnm 3.87 a ¼ 4.9, b ¼ 9.46, c ¼ 2.96 [64, 66, 67]
280 Y. Yang et al.

a b X in Lix TiO2
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
3 20 μm Rutile
0.5α 5 μm Rutile
f⬘ 500 nm Rutile
a⬘ 10x40 nm Rutile

Voltage (V)
100 nm
Rutile [110] 2 e⬘
0.322 nm

b⬘
c⬘
1 d⬘
0 30 60 90 120 150 180 210 240 270 300
Capacity (mA h g−1)
10 nm

Fig. 12.7 TEM and electrochemical characterization of rutile TiO2 nanorods. (a) HRTEM and
electron diffraction pattern of TiO2 nanorods. (b) Voltage profiles of rutile TiO2 with different
sizes: 20 mm (–), 0.5  5 mm (┅), 500 nm (–•–), and 10  40 nm (—) [72]

nanotubes can be synthesized by mixing TiO2 powder and concentrated NaOH


aqueous solution (e.g., 15 M) together and heating to 150–170 C [75, 76]. TiO2-B
nanowires exhibit an initial capacity around 200 mAh/g with stable capacity upon
cycling. A capacity loss of no more than 0.1% per cycle occurred during the first
100 cycles, and only 0.06% from cycle 20 to 100 [76]. The TiO2-B nanotubes even
reach a capacity as high as 328 mAh/g in the first discharge at a rate of 10 mA/g,
corresponding to Li0.98TiO2 [75].

12.2.5 Metal Oxides with Displacement Mechanism

In the metal oxides already discussed, the reaction mechanism with lithium is
mainly based on intercalation. Meanwhile, many metal oxides can react with
lithium by a displacement mechanism involving the formation of lithium oxide
(Li2O). This mechanism can be summarized by the half reaction 2x Li+ + MOx + 2x
e ! x LiO2 + M0, and the composition can be written as MxOy/M0/Li2O [77].
Though some kinds of displaced metals can further react with lithium to form alloy
compounds LixMy, such as SnO2 [78] and ZnO [79, 80], the displacement reaction
is universal in these metal oxides. These metal oxides include Fe3O4, CoO, SnO2,
ZnO, CuO, Cu2O, and MoO2. Some oxides, such as MnOx, can also react with
lithium by this mechanism [81, 82], but with less interesting characteristics.
Displacement reactions of metal oxides with lithium have a typical voltage of
1–2 V vs. Li/Li+, as summarized in Table 12.2 [82]. As a result, these oxides can act
as the anode in lithium-ion batteries. These oxides have theoretical capacities
between 600 and 1,200 mAh/g, which is two to three times that of graphite, the
most common anode material [4, 82]. Moreover, many metal oxides, such as iron
12 Oxide Nanostructures for Energy Storage 281

Table 12.2 The equilibrium potential of some metal oxides by displacement mechanism [82]
Material Voltage vs. Li/Li+ (V) Capacity (mAh/g) Material EMF (V) Capacity (mAh/g)
Fe2O3 1.631 1007 CoO 1.802 715
SnO2 1.582 711 CuO 2.248 674
Ga2O3 1.188 858 ZnO 1.252 659

oxide and copper oxide, are very cheap. These advantages make them attractive for
use in lithium-ion batteries, and much research has been dedicated to this end.
However, the insulating nature of Li2O and the strong Li–O bonds result in marked
hysteresis in voltage and significant capacity fading during cycles [77, 83]. By
reducing the size of materials down to nanoscale, the spatial dimensions of
insulating Li2O can be significantly reduced. In addition, the pathways for electrons
and ions become much shorter. As a result, by designing rational and efficient
nanoarchitectures, such as metal/metal oxide hybrids and porous structures, the
electrochemical performance can be improved dramatically [77, 84–86].
Taberna et al. developed a strategy to form metal/metal oxide core/shell
nanowire arrays. In their design, the metal core acts as the efficient electron
collector and the thin oxide shell facilitates the transport of ions. A Cu/Fe3O4
core/shell nanowire array was demonstrated as the example. The nanowire arrays
were fabricated by a two-step method. First, Cu nanowires were electrodeposited on
a Cu foil with the confinement of an anodic aluminum oxide (AAO) template. Then
Fe3O4 was uniformly electrodeposited onto the Cu nanowires. The dimension of the
copper nanowires and the thickness of magnetite coating were tuned by the size of
the AAO template and the deposition conditions. Typically, the copper nanowires
had a diameter of 200 nm and uniform length of several micrometers. These hybrid
core/shell nanowire arrays exhibit significantly improved power performance
(Fig. 12.8). For instance, the sample with Fe3O4 deposited for 150 s results in a
much higher capacity than Fe3O4 powders at high rate. At a 1 C rate (8 Li+/h), more
than 85% capacity is retained compared to C/32 (0.25 Li+/h). Meanwhile, the
capacity retention at a 1 C rate is only ~45% for Fe3O4 powders. Even at a higher
rate, such as 8 C, the capacity retention for the Cu/Fe3O4 core/shell nanowires is still
75%, while that of Fe3O4 powder is only about 25%. These nanowire arrays also
show good capacity retention. No decay in capacity was observed after 50 cycles.
Another advantage of such core/shell nanowires is the higher mass loading due to the
nanowire geometry. The Fe3O4/Cu nanowires exhibit similar rate performance as the
planar Fe3O4 film [87]. However, the mass loading of the planar film was only
0.136 mg, which corresponds to a capacity of ~0.05 mAh/cm2. In contrast, the
nanowire structure showed a mass of 0.820 mg and a capacity of about 0.3 mAh/cm2.
Among oxides that react according to the displacement mechanism, tin oxide
(SnO2) is an especially attractive material. SnO2 has a high theoretical capacity of
1,494 mAh/g due to the two-step reaction with lithium. First, lithium replaces tin to
form lithium oxides (4 Li + SnO2 ! 2 Li2O + Sn, ~1.6 V vs. Li/Li+). Second,
lithium alloys with tin to form Li4.4Sn (4.4 Li + Sn ! Li4.4Sn, <1.0 V vs. Li/Li+).
As a result, one unit of SnO2 could accommodate 8.4 lithium. However, the first
282 Y. Yang et al.

c 120 C/32C/16 C/8 C/4 C/2 C 2C 4C 8C

100
Fe304–Cu

Normalized capacity
b, t2
80 c, t3
d, t4
With Fe3O4 60 a, t1
e, t5
40

20
Fe3O4 Powder
0
2 μm 1 μm 0.01 0.1 1 10
C rate
b1,200 d 0.5
C/32C/16 C/8 C/4 C/2 C 2C 4C 8C
Specific capacity (mA h g−1)

1,000 1 Li+/4h 1 Li+/2h 0.4

Capacity (mA h cm−2)


b,t2, m=0.820 mg
800
0.3
600
0.2
400
Fe304–Cu, m=0.136 mg
0.1
200

0 0
0 10 20 30 40 50 0.01 0.1 1 10
Cycle number C rate

Fig. 12.8 Fe3O4/Cu core/shell nanostructure for lithium-ion batteries. (a) Cross section of Cu-
nanostructured current collector before (left) and after (right) Fe3O4 coating. (b) The cycling
performance of a Fe3O4 film electrodeposited onto nanoarchitectured copper for 150 s. The
electrode was cycled first at C/32 for 15 cycles followed by a higher rate of C/16. (c) The rate-
dependent capacity of Fe3O4/Cu core/shell nanostructure with different time of electrodeposition:
120 s (t1), 150 s (t2), 180 s (t3), 230 s (t4), 300 s (t5). Fe3O4–Cu indicates a Fe3O4 film on planar
copper substrate. Fe3O4 powder is made by 60 wt% Fe3O4, 22 wt% carbon black, and 18 wt%
PVDF. (d) The rate-dependent capacity (mAh/cm2) of a Fe3O4-based Cu-nanostructured electrode
(deposition time: 150 s) and a Fe3O4-based Cu planar electrode [77]

step is highly polarized and not reversible due to the formation of lithium oxide.
Moreover, the second step involves large volume expansion. To overcome these
issues, various nanostructures have been designed and significant improvement has
been observed.
Kim et al. studied the size effect of SnO2 nanoparticles for lithium-ion batteries
[85]. They synthesized SnO2 nanoparticles of various sizes (3, 4, and 8 nm) by
hydrothermal methods. 3 nm nanoparticles show a discharge capacity of 740 mAh/g
in the voltage window of 0–1.2 V with capacity loss less than 10 mAh/g over 60
cycles. In comparison, 4 nm nanoparticles show a capacity fade over 100 mAh/g
under the same conditions. When the particle size further increases to 8 nm, the
corresponding discharge capacity decreases to less than 100 mAh/g after 30 cycles.
The authors attribute the good cycling performance of 3 nm particles to the fact that
these smaller nanoparticles can undergo reversible volume changes without aggre-
gation into larger Sn clusters during cycling.
To minimize the influence of pulverization of SnO2 during charge/discharge,
structures that can undergo large volume changes have been fabricated, such as
12 Oxide Nanostructures for Energy Storage 283

macroporous/mesoporous and hollow SnO2. Lou et al. synthesized hollow SnO2


nanospheres with diameter of 50–200 nm and wall thicknesses of ~10 nm [88]. The
hollow nanospheres exhibit an initial discharge capacity of nearly 1,200 mAh/g,
which is much higher than SnO2 nanoparticles. Lee et al. fabricated macroporous
SnO2/carbon materials from a close-packed PMMA nanosphere template; the
porous composites exhibit moderate capacity retention with cycling [89]. Besides
porous nanostructures, one-dimensional structures of SnO2, such as nanowires and
nanorods, have also been utilized and improved electrochemical performance has
been observed [78, 84, 90–93].

12.2.6 Nano Oxide Coatings

Besides acting as the active material in electrodes of lithium ion-batteries,


nanostructured oxides can also be employed as a surface modification agent for
battery materials, especially for cathodes. Typically, these oxide coatings have a
thickness of several tens of nanometers and are distributed uniformly on the surface
of active material particles. This inert oxide layer can reduce the contact between
active electrode and the electrolyte. There are two common types of coatings, and
they are distinguished by their functions and targeting materials: (1) coatings to
enhance the stability of layered oxides (e.g., LiCoO2, Li1 + xNiyMnzCowO2) at a
high potential and (2) coatings to suppress the dissolution of active materials, such
as LiMn2O4 and LiFePO4.
As mentioned at the beginning of this chapter, only half the lithium atoms in
LiCoO2 are utilized in commercial lithium-ion batteries due to the chemical
instability between LixCoO2 and the organic electrolyte when x < 0.5. One way
to increase the stability is to coat or modify the surface of LiCoO2 with an inert
material, which could reduce or prevent the direct contact between LiCoO2 and the
electrolyte. Nano oxide coatings on the surface of LiCoO2, including Al2O3
[94–96], TiO2 [97, 98], MgO [96, 99], ZnO [100], SnO2 [101], and ZrO2 [97,
102], have been realized and better stability of LixCoO2 with the electrolyte at high
voltage has been observed. For example, after coating LiCoO2 with a ~5 nm thick
amorphous Al2O3 film, the reversible capacity reaches 190 mAh/g over 20 cycles in
the voltage range of 3.0–4.5 V vs. Li/Li+, corresponding to ~0.7 Li atoms utilized
per formula unit instead of 0.5 Li atoms per LiCoO2 formula unit [94]. In compari-
son, bare LiCoO2 shows a capacity of only 118 mAh/g after 20 cycles, which is just
60% of the initial capacity.
Mixed layer oxide cathodes, such as Li1 + xNiyCozMnwO2, have recently become
a much-studied systems as they exhibit a capacity higher than 180 mAh/g [103].
For instance, the solid solution between Li2MnO3 (Li[Li1/3Mn2/3]O2) and
LiNixCoyMnzO2 can reach a capacity of about 250 mAh/g, corresponding to
0.8–0.9 lithium utilized per unit formula of LiNixCoyMnzO2 [104, 105]. However,
such high capacity requires charging up to about 4.8 V, which corresponds to the
oxidation of O2 to neutral oxygen. This process is irreversible during the following
284 Y. Yang et al.

Fig. 12.9 (a) TEM image of Al2O3-coated Li(Li0.2Mn0.54Ni0.13Co0.13)O2. The light color at the
interface indicates an Al2O3 coating. (b) First charge–discharge profiles of the layered (1  x) Li
[Li1/3Mn2/3]O2–x Li[Ni1/3Mn1/3Co1/3]O2 solid solutions before and after surface modification with
3 wt% nanostructured alumina [106]

discharge/charge. Consequently, an irreversible capacity loss related to the release


of oxygen occurs. This capacity loss can be as large as 50–80 mAh/g, resulting in a
coulomb efficiency less than 80% [104, 105]. Moreover, the release of oxygen can
also lead to safety concerns [1].
Diminution of this capacity loss has been achieved by nano oxide coatings.
For example, Wu et al. coated the solid solution of (1 – x) Li[Li1/3Mn2/3]O2–x Li
[Ni1/3Mn1/3Co1/3]O2 (0.3  x  0.7) with a layer of amorphous alumina less than
5 nm thick [106]. After coating, the irreversible capacity loss in the first cycle decreased
markedly. In the material with x ¼ 0.4 (Li[Li0.2Mn0.54Ni0.13Co0.13]O2), the capacity
loss was 41 mAh/g after coating, while a 75 mAh/g capacity loss was observed for the
bare samples (Fig. 12.9). The initial discharge capacity after coating was as high as
280 mAh/g, which is twice that of LiCoO2. Moreover, the capacity loss during the
following cycles was reduced from 0.53 to 0.34 mAh/g per cycle after coating.
In addition to stabilizing layered oxides, oxide coatings can also suppress the
dissolution of LiMn2O4 and LiFePO4. The main bottleneck to the commercial use
of LiMn2O4 is its dissolution in organic electrolytes, which leads to the dissolved
Mn3+ ions chemically attacking the graphite anode, especially at high temperatures
(discussed in Sect. 12.2.1). The trace amount of water in the electrolyte can react
with the commonly used solute LiPF6 to generate HF, which results in an acidic
environment to disproportionate Mn3+ ions in LiMn2O4 and thus cause the dissolu-
tion of this material [107, 108]. Previous studies have shown that when temperature
increases from 25 to 55 C, the concentration of Mn ions in the electrolyte increases
dramatically from 7.6 to ~400 ppm [109]. Moreover, many reports have shown that
the capacity of LiMn2O4 fades much faster at high temperatures [29, 107].
12 Oxide Nanostructures for Energy Storage 285

Accordingly, surface coating can the reduce contact between LiMn2O4 and the
electrolyte and thus the reduce dissolution of Mn. Various oxide coatings, including
Al2O3 [110, 111], TiO2 [112], ZrO2 [113], SiO2 [114], MgO [111], ZnO [19, 115,
116], VOx [26], CeO2 [117], and LiCo1  xNixO2 [111, 118, 119], have been
realized and better cycling performance has been observed. The battery material
LiFePO4 undergoes similar dissolution mechanisms; Fe ions can also be dissolved
in the electrolyte and commonly deposit on the carbon anode. Recent reports have
shown that oxide coatings, such as ZnO and TiO2, can improve the cycling behavior
of LiFePO4 at high temperatures [120, 121].
For example, Han et al. coated doped LiMn2O4 (Li1.05Al0.1Mn1.85O3.95F0.05)
with ZnO using aqueous zinc acetate solution as a precursor [19]. By introducing a
2 wt% ZnO coating, the capacity retention of LiMn2O4/graphite full cells improves
significantly at high temperatures (55 C). In bare LiMn2O4/graphite cells, the
retained capacity after 300 cycles was only about 40% of its initial capacity. Bare
Li1.05Al0.1Mn1.85O3.95F0.05/graphite full cell showed a retention of ~50% after 300
cycles, although bare Li1.05Al0.1Mn1.85O3.95F0.05/lithium half cell exhibited 97%
capacity retention after 50 cycles at 55 C. In comparison, In the case of 2% ZnO-
coated Li1.05Al0.1Mn1.85O3.95F0.05/graphite cells, the capacity retention increased to
about 80% after 500 cycles. Furthermore, there is little specific capacity loss
resulting from the dead weight of the added ZnO coating layer. ZnO-coated
Li1.05Al0.1Mn1.85O3.95F0.05/graphite cells also exhibited a smaller overpotential
and better performance at high power rates. For instance, in half cell tests with
lithium metal as the anode, the capacity of ZnO-coated Li1.05Al0.1Mn1.85O3.95F0.05
half cells at a 5 C rate is over 75 mAh/g, while the capacities of bare LiMn2O4 and
bare Li1.05Al0.1Mn1.85O3.95F0.05 are less than 60 mAh/g .
Though oxide coatings can significantly enhance the stability of cathode
materials, it is worth pointing out that the coating layer can have long-term
durability issues. As a result, developing stable and robust coatings remaining
intact even under aggressive charge and discharge conditions is the key for the
success of this technique [1].

12.3 Nano Oxides for Electrochemical Capacitors

Various metal oxide nanostructures have also been used as another type of energy
storage device, electrochemical capacitors (ECs). Applications of ECs have
expanded in recent years, most of which utilize them as complementary energy
storage devices to batteries, as they can operate at fast charge–discharge rates.
Specific examples that have either been demonstrated already or which progres-
sively will rely more on ECs in the future include hybrid electrical vehicles (HEV),
heavy duty cranes, and uninterruptible power supply (UPS) [122].
In general, ECs are divided into two classes, double layer ECs and pseudoca-
pacitors, depending on their charge storage mechanism [123]. Double layer ECs
store charges in the electrical double layer on the electrode surface via electrostatic
286 Y. Yang et al.

interactions. Thus, only the surfaces of electrodes contribute to capacitance, and


materials with large surface areas, such as activated carbons [122, 124–129], have
attracted the most attention. On the other hand, the primary charge storage mecha-
nism in pseudocapacitors involves faradaic redox reactions that control the oxida-
tion states of host materials. Because pseudocapacitors can store charge at a certain
depth within the active materials, rather than merely on their surfaces, pseudocapacitor
specific capacitances are typically much larger than those of double layer ECs. These
higher specific capacitances have attracted increasingly more attention to pseudoca-
pacitors, as both scientific and nonscientific communities are constantly exhibiting
higher demand for environmentally friendly energy storage devices with large energy
densities.
Pseudocapacitance has been observed mainly with conductive polymers and
metal oxides. In the pseudocapacitance process, ions in the electrolyte intercalate
into the lattices of host materials to a certain extent and facilitate redox reactions.
Pseudocapacitors are similar to batteries in that charge storage is based on redox
reactions in the electrode materials. However, different aspects also exist: (1)
Charging and discharging in batteries take place mostly at relatively constant
voltages that are determined by the redox potentials of the materials; in the case
of pseudocapacitance, the potential between both electrodes changes continuously
during charging and discharging. (2) Most pseudocapacitors operate only in aque-
ous electrolytes, whereas batteries can operate in both aqueous and organic
electrolytes. This is mainly because redox reactions in pseudocapacitors are
associated with protons and other ions soluble only in the aqueous phase.
In general, metal oxide pseudocapacitors exhibit the following characteristics:
(1) Most metal oxides have reasonably good electrical conductivities due to oxygen
vacancies. Nevertheless, the conductivities of some metal oxides such as manga-
nese oxide are not sufficiently high, so composite electrodes with conductive
nanomaterials are often used. (2) Transition metals have multiple oxidation states
between which electron hopping can take place. (3) Protons can freely intercalate
into oxide lattices to allow reduction of O2 $ OH. (4) The redox reaction can
reversibly take place over many cycles.
Similar to Li-ion batteries, nanodimensional electrodes provide several important
advantages in pseudocapacitor operations. In most cases, the large surface-to-volume
ratio inherent in nanostructures allows for easier ionic access into the active materials,
and thus, nanostructured pseudocapacitors show higher capacities and better rate
capabilities compared with equivalent bulk-scale materials [130]. The benefits of
nanostructures are more prominent for materials whose electrical or ionic conductivities
are poor. The decreased dimensions provide shorter pathways for electron and ion
diffusion and therefore lead to significantly improved specific capacitances and power
densities. In the case that metal oxides are integrated with conductive nanomaterials
in the form of composites [122, 131–134], nanostructures allow for more uniform
mixing, and therefore, the effect of the conductive nanomaterials becomes more
remarkable. Although pseudocapacitor performance can be improved monumentally
by using nanostructured materials, some disadvantages cannot be ignored. One of the
most critical problems is that electrode materials can be dissolved in the electrolyte
12 Oxide Nanostructures for Energy Storage 287

due to the large surface areas of the nanostructures which are exposed to the electrolyte.
This dissolution problem results in a limited cycle life, which is one of the most
important parameters in EC operations.

12.3.1 Ruthenium Oxide

Among a number of pseudocapacitor materials, ruthenium oxide (RuO2) has shown


the highest specific capacitances up to 800 F/g [129]. Both hydrated and crystalline
forms of RuO2 have been tested, but the hydrated form has shown the highest
specific capacitances, which are about twice as high as those of the crystalline form.
Typically, hydrous RuO2 is a mixed conductor that conducts both protons and
electrons in acidic solution, while crystalline RuO2 conducts only electrons. During
the charging–discharging process, the protons and electrons are transferred between
RuO2 and the electrolyte solution. Thus, it is desirable to have high conductivities
for both electrical and protonic transports. In addition, the stable potential range of
RuO2, ~1.4 V, is relatively wider than other materials [135–138]. However, the
excellent performance of the hydrated form has been demonstrated only in highly
acidic electrolytes such as sulfuric acid. The superior specific capacitance is
attributed to the high availability of protons in strong acids because protons have
better access to both the surface and interior of the electrode than other larger alkali
ions such as K+ and Na+. In fact, compared to these larger alkali ions, proton
diffusion in hydrated RuO2 is exceptionally fast, leading to higher power densities
[123]. The chemical diffusion coefficient of hydrogen in bulk crystalline RuO2 is
about 5  1014 cm2/s, and thus the penetration of protons during a typical charging
process must still be shallow [139, 140]. Therefore, nanostructures [141] in the
hydrous form are expected to improve the gravimetric capacitance because the
surface area is significantly larger. Another critical drawback of RuO2 is the scarcity
of the underlying metal, which makes it too expensive for commercialization. Due to
the cost burden, current research is focused rather on other abundant metal oxides.
A variety of RuO2 structures have been prepared using different methods. As
mentioned above, RuO2 is prepared in either its hydrous or crystalline form. These
structures are typically determined by the oxidation condition. Preparation of the
hydrous RuO2 is usually initiated from a hydrous ruthenium precursor such as hydrous
ruthenium chloride (RuCl3 · xH2O). First, precursors are crushed into fine particles
and then precipitated in the form of RuOxHy. Next, the precipitants are transformed
into a hydrous RuO2 solid through annealing steps at 150–400 C under oxygen flow.
The oxidation processes for hydrous RuO2 can be also accomplished by electrochem-
ical oxidation steps such as cyclic voltammetry sweeps in a sulfuric acid solution
[142]. On the other hand, crystalline RuO2 is initiated from anhydrous precursors and
transformed into the crystalline structure by longer annealing steps [143].
As in most EC characterization processes, RuO2 EC performance is typically
tested by galvanostatic, cyclic voltammetric, and impedance measurements. Rep-
resentative data from a galvanostatic measurement is presented in Fig. 12.10a. The
288 Y. Yang et al.

a b
1000

Specific Capacitance (F/g)


1.0
800
0.8
Voltage (volt)

Cycle # 154 Cycle # 2496 600


0.6

0.4 400

0.2 200

0 0
0 2 4 6 8 30 40 50 60 70 80 90 100
Time (hour) Specific Surface Area (m2/g)

Fig. 12.10 (a) Galvanostatic (5 mA constant current) charging–discharging curves for RuO2 
xH2O pseudocapacitors. Each electrode contains ~0.145 g of RuO2  xH2O powder that was
annealed at 150 C. (b) Specific capacitances for RuO2  xH2O with different BET surface areas.
The specific capacitance is constant in these BET surface area ranges. Both graphs are from [129]

areal capacitance was obtained by measuring the BET surface area. Hydrous RuO2
has shown areal capacitances as high as 260 mC/cm2, which is about ten times larger
than those of activated carbons. This significantly increased specific capacitance
verifies that charges are stored in the bulk of electrodes through insertion reactions.
Although RuO2 nanostructures have been developed using various methods, EC
performance is more directly related to whether the structure is hydrous or crystal-
line. Indeed, the highest specific capacitance values (700–900 F/g) are achieved
from hydrous RuO2 in both micrometer [129] and nanometer [144] dimensions.
Figure 12.10b from ref. [129] shows that a certain range of surface area increases
does not improve the gravimetric capacitance. In addition, the insertion depth of
guest ions into host structures is often directly correlated to the cycle life. The deep
insertion of guest ions during cycling can lead to volume changes of electrodes and
thus impair the cycle life.

12.3.2 Manganese Oxide

Despite their excellent capacitances, the high material cost of RuO2 has been a
primary limitation in its development toward commercial products. To overcome
the cost issue, research has been driven to investigate relatively low cost materials
such as manganese oxide (MnO2) and nickel oxide (NiO). Manganese oxides
especially have been low cost useful materials in various applications including
catalysis [143, 145] and energy storage [55, 141, 146]. However, their typically
poor electrical conductivities need to be enhanced [147, 148] for EC operations.
As covered in Sect. 12.2, manganese oxides have been synthesized as various
nanostructures, including dendritic clusters [149], nanocrystals [150–153] with
different shapes, nanowires [26, 30, 154], nanotubes [155], nanobelts [156, 157],
and nanoflowers [132]. Also, various electrochemical and chemical methods have
12 Oxide Nanostructures for Energy Storage 289

been developed for thin film MnO2 electrodes [158, 159]. Among numerous
chemical methods, the chemical reduction of KMnO4 is one of the most well-
known processes to produce MnO2 nanostructures. Kim et al. chemically reduced
KMnO4 with Mn/Ni/Pb acetate solutions [160] to generate both pure MnO2 and Pb,
Ni-mixed MnO2 nanostructures. Similarly, KMnO4 was also reduced using potas-
sium borohydride (KBH4), sodium dithionate (Na2S2O6), and sodium
hypophosphite (NaPO2H2) [161]. MnO2 nanostructures were also prepared by
simple precipitation methods. The precipitation is enabled by mixing aqueous
solutions, such as KMnO4–MnSO4 [162] and Mn(CH3COO)2–KMnO4 [163]. Lee
et al. also developed a simple thermal decomposition of finely ground KMnO4
powders at different temperatures ranging 300–1,000 C [164]. Among various
MnO2 pseudocapacitors, one of the most representative structures and operation
data are presented in Fig. 12.11 [156]. Similar to RuO2 cases, as hydrothermal
reaction times are increased, the sizes of structures become smaller and thus BET
surface areas increase. From a crystal structure perspective, longer hydrothermal
reaction times lead to a crystal structure transition from amorphous to crystalline
and due to this transition, specific capacitances decrease from 150 to 70 F/g. SEM
images in Fig. 12.11a show the nanostructure evolution at different hydrothermal
reaction points. The nanostructure generated after a 6-h hydrothermal reaction was
electrochemically characterized, and a galvanostatic curve is shown in Fig. 12.11b.
Contrary to RuO2 pseudocapacitors, MnO2 pseudocapacitors can also be operated
in various aqueous electrolytes beyond acid, such as sodium sulfate (Na2SO4).

12.3.3 Other Metal Oxides

Beyond RuO2 and MnO2, other metal oxides such as iridium oxide (IrO2), nickel
oxide (NiO), and cobalt oxide (CoO) have also been tested as pseudocapacitor
electrode materials.
NiO has been prepared not only as crystalline nanoparicles [165] and electro-
chemically deposited films [166], but also as mesoporous particles [167]. In most
cases, nickel hydroxide (Ni(OH)2) is made by heating initial nickel precursors or by
electrochemically inducing precipitation. Then, Ni(OH)2 is transformed into NiO by
calcination processes typically around 300 C. The mesoporous particles are prepared
via similar procedures, but templates such as sodium dodecyl sulfate (C12H25SO4Na)
are included in initial mixtures for uniform pore size. The template-based methods
yield uniform pore sizes usually around several nanometers. The highest specific
capacitance of around 300 F/g was reported [165] from crystalline nanoparticle
electrodes. Other mesoporous particles or film electrodes have exhibited relatively
lower specific capacitances of 50–100 F/g [17, 166–168].
CoO has also shown capacitive behavior when tested as pseudocapacitor
electrodes. However, CoO has not been studied as much as the aforementioned
metal oxide materials. Still, CoO was prepared in the form of xerogel [169], or film
[170, 171] using sol–gel processes or electrochemical deposition. The preparation
290 Y. Yang et al.

Fig. 12.11 (a) SEM images of MnO2 nanostructures after different hydrothermal reaction times.
An aqueous solution of MnSO4  H2O and KMnO4 underwent a hydrothermal reaction at 140 C.
(b) Galvanostatic charging–discharging data for the sample after a 6-h hydrothermal reaction. The
current rate was 200 mA/g. (a) and (b) are from [156]
12 Oxide Nanostructures for Energy Storage 291

steps are similar to those of other metal oxide nanostructures. In the sol–gel
reactions, precursors such as cobalt chloride (CoCl2) undergo a couple of reactions
at increased temperatures to form cobalt hydroxide (Co(OH)2) precipitants. The
precipitants are calcinated at ~150 C to form the final CoO nanostructures. As in
other metal oxide cases, longer or higher temperature calcinations can lead to
further oxidation toward crystalline Co3O4 and thus reduced specific capacitance.
Lin et al. [169] prepared xerogel electrodes that showed excellent specific
capacitances of ~300 F/g. In addition, cobalt has been used to form binary metal
oxides with other metal elements such as Mn and Ni. These binary metal oxides
have also been prepared through similar processes and tested as pseudocapacitor
electrodes [102, 172–175]. With these binary metal oxides, specific capacitances of
50–300 F/g have been demonstrated.
Iridium oxide (IrO2) has been also studied as a material for pseudocapacitor
electrodes since the 1970s [176]. However, material scarcity and its relatively small
potential range have prevented its research from being active. Currently, the
research focus is rather on other metal oxides.

12.3.4 Hierarchical Metal Oxide–Carbon Composites

In some pseudocapacitor metal oxides, poor electrical conductivity is one of the


most limiting factors in the EC performance. In fact, due to this limited conductiv-
ity, the gravimetric capacitance strongly depends on the film thickness. In the case
of MnO2 films, while thin (10–100 nm) films exhibit very high specific capacitances
of 700–1,300 F/g [158, 162, 177], thicker films or larger structure electrodes show
relatively lower specific capacitances of 150–250 F/g. In order to overcome the
poor conductivity, various carbon nanostructures have been integrated with metal
oxide nanostructures in the form of nanocomposites. Thus, carbon nanostructures
such as nanoparticles, nanotubes, nanofibers, and mesoporous materials not only
contribute to increasing the conductivity, but also function as double-layer EC
electrodes utilizing inherently large surface areas. Fischer et al. [141] integrated
MnO2 with porous carbon nanofoams using electroless self-limiting deposition, but
specific capacitance (60–120 F/g) was not improved significantly. Jang et al. [142]
impregnated a ruthenium precursor into mesoporous carbon and then electrochemi-
cally oxidized the precursor to RuO2. Their specific capacitance was as high as
250 F/g even with a large mass loading. One-dimensional nanostructures such as
carbon nanotubes [132, 178–180] and carbon nanofibers [181] are also incorporated
with metal oxide nanostructures to utilize their unique advantages of large surface
areas to improve the film conductivities. Zhang et al. [132] developed more
organized carbon-MnO2 composites by electrochemically depositing MnO2
nanoflower structures onto aligned carbon nanotube arrays. As shown in
Fig. 12.12a, MnO2 nanoflowers were attached to the carbon nanotube surfaces.
Their specific capacitances reached 200 F/g. Moreover, the composite structures
exhibited an excellent cycle life such that the capacitance loss after 20,000 cycles
292 Y. Yang et al.

Fig. 12.12 (a) An SEM image of a MnO2–carbon nanotube composite electrode. MnO2
nanoflowers are electrochemically attached to aligned carbon nanotube arrays. (b) Capacity
retention data for the cell shown in (a). Only 3% of the initial capacitance was lost after 20,000
cycles. (c) An SEM image of MnO2–carbon nanotube energy textiles. MnO2 nanostructures were
electrochemically attached to carbon nanotube networks that were integrated into cotton textiles.
(d) Capacity retention data for the cell shown in (c). No visible capacitance decay was observed
after 40,000 cycles. (e) An SEM image of a RuO2–carbon nanofiber composite electrode. RuO2 
xH2O is conformally coated onto carbon nanofibers. (a), (b), and (c) [132, 161, 181] respectively
12 Oxide Nanostructures for Energy Storage 293

was only 3% relative to the original (Fig. 12.12b). Our group also developed
carbon–MnO2 composite cells [161]. But, we used cotton textiles, instead of flat
metal substrates, to reduce the overall cell weight and also to improve the electro-
lyte accessibility to the electrode surface, which was driven by capillary forces
within the three-dimensional porous structures of textile fibers. We first prepared a
carbon nanotube ink and then dipped the textile into the ink. After several cycles of
a dipping and drying process, a sheet resistance as low as 1 Ω/sq was reached.
Then, MnO2 nanoflowers were electrochemically deposited onto the carbon nano-
tube networks. Not only was a high specific capacitance of 230 F/g achieved, but
the compatibility of energy storage devices with stretchable platforms was also
demonstrated. An excellent capacity retention was demonstrated up to 40,000
cycles (Fig. 12.12d). Lee et al. [181] first transformed RuCl3 to ruthenium ethoxide
(Ru(OC2H5))3, and then added carbon fibers to the (Ru(OC2H5))3 solution. As this
solution was annealed at 180 C, RuO2 films were formed conformally on the
carbon nanofibers (Fig. 12.12e). They also compared their composite cells with
control samples without carbon nanofibers. The improved specific capacitances
(>1,000 F/g vs. ~400 F/g) verified the importance of carbon nanostructures as
electrically conductive pathways.

12.4 Summary

This chapter reviews how nanostructured oxide materials can impact the perfor-
mance of energy storage devices, such as lithium-ion batteries and electrochemical
capacitors. Recent developments in designing new materials and structures, and
improvement on known materials are discussed. Significant effects have been
observed in utilizing nanostructured materials. Higher power rate, better cycling
performance and larger capacity have been observed in many studies for both
lithium-ion batteries and electrochemical capacitors.
In lithium-ion batteries, since the transport lengths of both ions and electrons are
reduced significantly in nanostructured materials, much higher power rate is
achieved in nanostructured materials, such as LiMn2O4, TiO2, and V2O5. Some
inactive materials even turn to active toward lithium by reducing the size down to
nanoscale, such as b-MnO2. Besides improving electrochemical properties of
materials, nanosized oxides also act as a protective layer to enhance the stability
of active materials in the electrolyte.
Nanostructured metal oxides also result in improved performance in electro-
chemical capacitor operations. Similar to lithium-ion batteries, nanostructures
decrease electrical and ionic diffusion lengths significantly so that power and
energy densities can be improved compared to bulk-scale electrodes. Also,
nanostructures allow uniform mixing with conductive carbon nanomaterials in the
metal oxide–carbon nanocomposites. These composite structures exhibit further
improved specific capacitances compared to the bare metal oxide cases.
294 Y. Yang et al.

In summary, nanostructured materials offer great opportunities for the next


generation energy storage devices. Novel design of composition, structures, and
synthetic conditions are crucial for the success of nanostructured materials.

Acknowledgments The authors acknowledge support from King Abdullah University of Science
and Technology (KAUST) and Global Climate and Energy Project (GCEP) at Stanford University.
Yuan Yang acknowledges support from Stanford Graduate Fellowship.

References

1. Manthiram, A., Murugan, A.V., Sarkar, A., Muraliganth, T.: Nanostructured electrode
materials for electrochemical energy storage and conversion. Energy Environ. Sci. 1,
621–638 (2008)
2. Bruce, P.G., Scrosati, B., Tarascon, J.M.: Nanomaterials for rechargeable lithium batteries.
Angew. Chem. Int. Ed. 47, 2930–2946 (2008)
3. Linden, D., Reddy, T.B.: Handbook of Batteries. McGraw-Hill, New York (2001)
4. Tarascon, J.M., Armand, M.: Issues and challenges facing rechargeable lithium batteries.
Nature 414, 359–367 (2001)
5. Yoshio, M., Brodd, R.J., Kozawa, A.: Lithium-Ion Batteries: Science and Technology.
Springer, New York (2009)
6. Whittingham, M.S.: Lithium batteries and cathode materials. Chem. Rev. 104, 4271–4301
(2004)
7. Cheng, F., Tao, Z., Liang, J., Chen, J.: Template-directed materials for rechargeable lithium-
ion batteries. Chem. Mater. 20, 667–681 (2008)
8. Morales, A.M., Lieber, C.M.: A laser ablation method for the synthesis of crystalline
semiconductor nanowires. Science 279, 208–211 (1998)
9. Hu, J.T., Odom, T.W., Lieber, C.M.: Chemistry and physics in one dimension: Synthesis and
properties of nanowires and nanotubes. Acc. Chem. Res. 32, 435–445 (1999)
10. Xia, Y.N., Yang, P.D., Sun, Y.G., Wu, Y.Y., Mayers, B., Gates, B., Yin, Y.D., Kim, F., Yan,
Y.Q.: One-dimensional nanostructures: Synthesis, characterization, and applications. Adv.
Mater. 15, 353–389 (2003)
11. Kong, X.Y., Ding, Y., Yang, R., Wang, Z.L.: Single-crystal nanorings formed by epitaxial
self-coiling of polar nanobelts. Science 303, 1348–1351 (2004)
12. Cho, K.S., Talapin, D.V., Gaschler, W., Murray, C.B.: Designing PbSe nanowires and
nanorings through oriented attachment of nanoparticles. J. Am. Chem. Soc. 127,
7140–7147 (2005)
13. Fan, J., Wang, T., Yu, C.Z., Tu, B., Jiang, Z.Y., Zhao, D.Y.: Ordered, nanostructured tin-
based oxides/carbon composite as the negative-electrode material for lithium-ion batteries.
Adv. Mater. 16(16), 1432 (2004)
14. Holland, B.T., Blanford, C.F., Do, T., Stein, A.: Synthesis of highly ordered, three-dimen-
sional, macroporous structures of amorphous or crystalline inorganic oxides, phosphates, and
hybrid composites. Chem. Mater. 11(3), 795–805 (1999)
15. Luo, J.Y., Wang, Y.G., Xiong, H.M., Xia, Y.Y.: Ordered mesoporous spinel LiMn2O4by a
soft-chemical process as a cathode material for lithium-ion batteries. Chem. Mater. 19,
4791–4795 (2007)
16. Jiao, F., Bao, J.L., Hill, A.H., Bruce, P.G.: Synthesis of ordered mesoporous Li-Mn-O spinel
as a positive electrode for rechargeable lithium batteries. Angew. Chem. Int. Ed. 47(50),
9711–9716 (2008)
12 Oxide Nanostructures for Energy Storage 295

17. Chan, C.K., Peng, H.L., Liu, G., McIlwrath, K., Zhang, X.F., Huggins, R.A., Cui, Y.: High-
performance lithium battery anodes using silicon nanowires. Nat. Nanotechnol. 3(1), 31–35
(2008)
18. Cui, L.F., Ruffo, R., Chan, C.K., Peng, H.L., Cui, Y.: Crystalline-amorphous core-shell
silicon nanowires for high capacity and high current battery electrodes. Nano Lett. 9(1),
491–495 (2009)
19. Han, J.M., Myung, S.T., Sun, Y.K.: Improved electrochemical cycling behavior of
ZnO-coated Li1.05Al0.1Mn1.85O3.95F0.05 spinel at 55 degrees C. J. Electrochem. Soc. 153(7),
A1290–A1295 (2006)
20. Ohzuku, T., Takehara, Z., Yoshizawa, S.: Non-aqueous lithium-titanium dioxide cell.
Electrochim. Acta 24(2), 219–222 (1979)
21. Zheng, Z.S., Tang, Z.L., Zhang, Z.T., Shen, W.C.: Review of cathode material LiMn2O4 for
lithium ion batteries. J. Inorg. Mater. 18(2), 257–263 (2003)
22. Dickens, P.G., Reynolds, G.J.: Transport and equilibrium properties of some oxide insertion
compounds. Solid State Ion. 5, 331–334 (1981)
23. Guyomard, D., Tarascon, J.M.: Li metal-free rechargeable LiMn2O4/carbon cells – their
understanding and optimization. J. Electrochem. Soc. 139(4), 937–948 (1992)
24. Kikkawa, S., Miyazaki, S., Koizumi, M.: Electrochemical aspects of the deintercalation of
layered AMO2 compounds. J. Power Sources 14(1–3), 231–234 (1985)
25. Thomas, M., Bruce, P.G., Goodenough, J.B.: AC impedance of the Li(1-X)CoO2 electrode.
Solid State Ion. 18–19, 794–798 (1986)
26. Cho, J.: VOx-coated LiMn2O4 nanorod clusters for lithium battery cathode materials.
J. Mater. Chem. 18(19), 2257–2261 (2008)
27. Kim, D.K., Muralidharan, P., Lee, H.W., Ruffo, R., Yang, Y., Chan, C.K., Peng, H., Huggins,
R.A., Cui, Y.: Spinel LiMn2O4 nanorods as lithium ion battery cathodes. Nano Lett. 8(11),
3948–3952 (2008)
28. Luo, J.Y., Xiong, H.M., Xia, Y.Y.: LiMn2O4 nanorods, nanothorn microspheres, and hollow
nanospheres as enhanced cathode materials of lithium ion battery. J. Phys. Chem. 112(31),
12051–12057 (2008)
29. Shaju, K.M., Bruce, P.G.: A stoichiometric nano-LiMn2O4 spinel electrode exhibiting high
power and stable cycling. Chem. Mater. 20(17), 5557–5562 (2008)
30. Hosono, E., Kudo, T., Honma, I., Matsuda, H., Zhou, H.S.: Synthesis of single crystalline
spinel LiMn2O4 nanowires for a lithium ion battery with high power density. Nano Lett. 9(3),
1045–1051 (2009)
31. Gummow, R.J., Dekock, A., Thackeray, M.M.: Improved capacity retention in rechargeable
4V lithium/lithium-manganese oxide (spinel) cells. Solid State Ion. 69(1), 59–67 (1994)
32. Amatucci, G., Tarascon, J.M.: Optimization of insertion compounds such as LiMn2O4 for Li-
ion batteries. J. Electrochem. Soc. 149(12), K31–K46 (2002)
33. Thackeray, M.M., Shao-Horn, Y., Kahaian, A.J., Kepler, K.D., Vaughey, J.T., Hackney, S.
A.: Structural fatigue in spinel electrodes in high voltage (4V) Li/LixMn2O4 cells.
Electrochem. Solid State Lett. 1(1), 7–9 (1998)
34. Xia, Y.Y., Zhou, Y.H., Yoshio, M.: Capacity fading on cycling of 4V Li/LiMn2O4 cells.
J. Electrochem. Soc. 144(8), 2593–2600 (1997)
35. Yoshio, M., Xia, Y.Y., Kumada, N., Ma, S.H.: Storage and cycling performance of Cr-
modified spinel at elevated. J. Power Sources 101(1), 79–85 (2001)
36. Chromik, R., Beck, F.: A quantitative discrimination between reversible Li+-insertion and
irreversible solvent oxidation at a lithium/manganese-spinel electrode. Electrochim. Acta 45
(14), 2175–2185 (2000)
37. Gao, Y., Dahn, J.R.: Synthesis and characterization of Li1+xMn2-xO4 for Li-ion battery
applications. J. Electrochem. Soc. 143(1), 100–114 (1996)
38. Blyr, A., Sigala, C., Amatucci, G., Guyomard, D., Chabre, Y., Tarascon, J.M.: Self-discharge
of LiMn2O4/C Li-ion cells in their discharged state – Understanding by means of three-
electrode measurements. J. Electrochem. Soc. 145(1), 194–209 (1998)
296 Y. Yang et al.

39. Kamarulzaman, N., et al.: Investigation of cell parameters, microstructures and electrochem-
ical behaviour of LiMn2O4 normal and nano powders. J. Power Sources 188(1), 274–280
(2009)
40. Peng, H.L., Xie, C., Schoen, D.T., Cui, Y.: Large anisotropy of electrical properties in layer-
structured In2Se3 nanowires. Nano Lett. 8(5), 1511–1516 (2008)
41. Meister, S., Schoen, D.T., Topinka, M.A., Minor, A.M., Cui, Y.: Void formation induced
electrical switching in phase-change nanowires. Nano Lett. 8(12), 4562–4567 (2008)
42. Duan, X.F., Lieber, C.M.: General synthesis of compound semiconductor nanowires. Adv.
Mater. 12(4), 298–302 (2000)
43. Lieber, C.M.: One-dimensional nanostructures: chemistry, physics & applications. Solid
State Commun. 107(11), 607–616 (1998)
44. Yang, Y., Xie, C., Ruffo, R., Peng, H., Kim, D.K., Cui, Y.: Single nanorod devices for battery
diagnostics: a case study on LiMn2O4. Nano Lett. 9(12), 4109–4114 (2009)
45. Linden, D., Reddy, T.B.: Handbook of Batteries, pp. 14.55–14.71. McGraw-Hill, New York
(2001)
46. Linden, D., Reddy, T.B.: Handbook of Batteries, pp. 34.8–34.12. McGraw-Hill, New York
(2001)
47. Devaraj, S., Munichandraiah, N.: Effect of crystallographic structure of MnO2 on its electro-
chemical capacitance properties. J. Phys. Chem. 112(11), 4406–4417 (2008)
48. Jiao, F., Bruce, P.G.: Mesoporous crystalline beta-MnO2 – a reversible positive electrode for
rechargeable lithium batteries. Adv. Mater. 19(5), 657 (2007)
49. Cheng, F.Y., Zhao, J.Z., Song, W., Li, C.S., Ma, H., Chen, J., Shen, P.W.: Facile controlled
synthesis of MnO2 nanostructures of novel shapes and their application in batteries. Inorg.
Chem. 45(5), 2038–2044 (2006)
50. Ma, R.H., Bando, Y., Zhang, L.Q., Sasaki, T.: Layered MnO2 nanobelts: hydrothermal
synthesis and electrochemical measurements. Adv. Mater. 16(11), 918–922 (2004)
51. Xiong, Y.J., Xie, Y., Li, Z.Q., Wu, C.Z.: Growth of well-aligned gamma-MnO2 monocrys-
talline nanowires through a coordination-polymer-precursor route. Chem. Eur. J. 9(7),
1645–1651 (2003)
52. Wang, X., Li, Y.D.: Selected-control hydrothermal synthesis of alpha- and beta-MnO2 single
crystal nanowires. J. Am. Chem. Soc. 124(12), 2880–2881 (2002)
53. Wang, X., Li, Y.D.: Synthesis and formation mechanism of manganese dioxide nanowires/
nanorods. Chem. Eur. J. 9(1), 300–306 (2003)
54. Wu, C.Z., Xie, Y., Wang, D., Yang, J., Li, T.W.: Selected-control hydrothermal synthesis of
gamma-MnO2 3D nanostructures. J. Phys. Chem. 107(49), 13583–13587 (2003)
55. Zhu, S.M., Zhou, H.A., Hibino, M., Honma, I., Ichihara, M.: Synthesis of MnO2 nanoparticles
confined in ordered mesoporous carbon using a sonochemical method. Adv. Funct. Mater. 15
(3), 381–386 (2005)
56. Ammundsen, B., Paulsen, J.: Novel lithium-ion cathode materials based on layered manga-
nese oxides. Adv. Mater. 13(12–13), 943 (2001)
57. Wang, L., Maxisch, T., Ceder, G.: A first-principles approach to studying the thermal stability
of oxide cathode materials. Chem. Mater. 19, 543–552 (2007)
58. Chabre, Y., Pannetier, J.: Structural and electrochemical properties of the proton gamma-
MnO2 system. Prog. Solid State Chem. 23(1), 1–130 (1995)
59. Chan, C.K., Peng, H.L., Twesten, R.D., Jarausch, K., Zhang, X.F., Cui, Y.: Fast, completely
reversible Li insertion in vanadium pentoxide nanoribbons. Nano Lett. 7(2), 490–495 (2007)
60. Delmas, C., Cognacauradou, H., Cocciantelli, J.M., Menetrier, M., Doumerc, J.P.: The
LiXV2O5 system - an overview of the structure modifications induced by the lithium interca-
lation. Solid State Ion. 69(3–4), 257–264 (1994)
61. Whittingham, M.S., Dines, M.B.: Normal-butyllithium – effective, general cathode screening
agent. J. Electrochem. Soc. 124(9), 1387–1388 (1977)
12 Oxide Nanostructures for Energy Storage 297

62. Patrissi, C.J., Martin, C.R.: Sol-gel-based template synthesis and Li-insertion rate perfor-
mance of nanostructured vanadium pentoxide. J. Electrochem. Soc. 146(9), 3176–3180
(1999)
63. Sakamoto, J.S., Dunn, B.: Vanadium oxide-carbon nanotube composite electrodes for use in
secondary lithium batteries. J. Electrochem. Soc. 149(1), A26–A30 (2002)
64. Yang, Z.G., Choi, D., Kerisit, S., Rosso, K.M., Wang, D.H., Zhang, J., Graff, G., Liu, J.:
Nanostructures and lithium electrochemical reactivity of lithium titanites and titanium
oxides: A review. J. Power Sources 192(2), 588–598 (2009)
65. Cromer, D.T., Herrington, K.: The structures of anatase and rutile. J. Am. Chem. Soc. 77(18),
4708–4709 (1955)
66. Banfield, J.F., Veblen, D.R.: Conversion of perovskite to anatase and TiO2 (B) – a TEM study
and the use of fundamental building-blocks for understanding relationships among the TiO2
minerals. Am. Mineral. 77(5–6), 545–557 (1992)
67. Kavan, L., Gratzel, M., Gilbert, S.E., Klemenz, C., Scheel, H.J.: Electrochemical and
photoelectrochemical investigation of single-crystal anatase. J. Am. Chem. Soc. 118(28),
6716–6723 (1996)
68. Takai, S., Kamata, M., Fujine, S., Yoneda, K., Kanda, K., Esaka, T.: Diffusion coefficient
measurement of lithium ion in sintered Li1.33Ti1.67O4 by means of neutron radiography. Solid
State Ion. 123(1–4), 165–172 (1999)
69. Koudraichova, M.V., Harrison, N.M., de Leeuw, S.W.: Diffusion of Li-ions in rutile. An ab
initio study. Solid State Ion. 157(1–4), 35–38 (2003)
70. Koudriachova, M.V., Harrison, N.M., de Leeuw, S.W.: Effect of diffusion on lithium
intercalation in titanium dioxide. Phys. Rev. Lett. 86(7), 1275–1278 (2001)
71. Stashans, A., Lunell, S., Bergstrom, R., Hagfeldt, A., Lindquist, S.E.: Theoretical study of
lithium intercalation in rutile and anatase. Phys. Rev. 53(1), 159–170 (1996)
72. Hu, Y.S., Kienle, L., Guo, Y.G., Maier, J.: High lithium electroactivity of nanometer-sized
rutile TiO2. Adv. Mater. 18(11), 1421 (2006)
73. Armstrong, G., Armstrong, A.R., Canales, J., Bruce, P.G.: TiO2(B) nanotubes as negative
electrodes for rechargeable lithium batteries. Electrochem. Solid State Lett. 9(3), A139–A143
(2006)
74. Armstrong, G., Armstrong, A.R., Bruce, P.G., Reale, P., Scrosati, B.: TiO2(B) nanowires as
an improved anode material for lithium-ion batteries containing LiFePO4 or LiNi0.5Mn1.5O4
cathodes and a polymer electrolyte. Adv. Mater. 18(19), 2597–2600 (2006)
75. Armstrong, G., Armstrong, A.R., Canales, J., Bruce, P.G.: Nanotubes with the TiO2-B
structure. Chem. Commun. 19, 2454–2456 (2005)
76. Armstrong, A.R., Armstrong, G., Canales, J., Garcia, R., Bruce, P.G.: Lithium-ion intercala-
tion into TiO2-B nanowires. Adv. Mater. 17(7), 862 (2005)
77. Taberna, L., Mitra, S., Poizot, P., Simon, P., Tarascon, J.M.: High rate capabilities Fe3O4-
based Cu nano-architectured electrodes for lithium-ion battery applications. Nat. Mater. 5(7),
567–573 (2006)
78. Li, N.C., Martin, C.R.: A high-rate, high-capacity, nanostructured Sn-based anode prepared
using sol-gel template synthesis. J. Electrochem. Soc. 148(2), A164–A170 (2001)
79. Wang, H.B., Pan, Q.M., Cheng, Y.X., Zhao, J.W., Yin, G.P.: Evaluation of ZnO nanorod
arrays with dandelion-like morphology as negative electrodes for lithium-ion batteries.
Electrochim. Acta 54(10), 2851–2855 (2009)
80. Yan, G.F., Fang, H.S., Li, G.S., Li, L.P., Zhao, H.J., Yang, Y.: Improved Electrochemical
Performance of Mg-doped ZnO Thin Film as Anode Material for Lithium Ion Batteries. Chin.
J. Struct. Chem 28(4), 409–413 (2009)
81. Reddy, A.L.M., Shaijumon, M.M., Gowda, S.R., Ajayan, P.M.: Coaxial MnO2/carbon nano-
tube array electrodes for high-performance lithium batteries. Nano Lett. 9(3), 1002–1006
(2009)
82. Li, H., Balaya, P., Maier, J.: Li-storage via heterogeneous reaction in selected binary metal
fluorides and oxides. J. Electrochem. Soc. 151(11), A1878–A1885 (2004)
298 Y. Yang et al.

83. Obrovac, M.N., Dahn, J.R.: Electrochemically active lithia/metal and lithium sulfide/metal
composites. Electrochem. Solid State Lett. 5(4), A70–A73 (2002)
84. Wang, Y., Lee, J.Y.: Molten salt synthesis of tin oxide nanorods: morphological and
electrochemical features. J. Phys. Chem 108(46), 17832–17837 (2004)
85. Kim, C., Noh, M., Choi, M., Cho, J., Park, B.: Critical size of a nano SnO2 electrode for Li-
secondary battery. Chem. Mater. 17(12), 3297–3301 (2005)
86. Shi, Y., Guo, B., Corr, S.A., Shi, Q., Hu, Y.-S., Heier, K.R., Chen, L., Seshadri, R., Stucky, G.
D.: Ordered mesoporous metallic MoO2 materials with highly reversible lithium storage
capacity. Nano Lett. 9(12), 4215–4220 (2009)
87. Mitra, S., Poizot, P., Finke, A., Tarascon, J.M.: Growth and electrochemical characterization
versus lithium of Fe3O4 electrodes made via electrodeposition. Adv. Funct. Mater. 16(17),
2281–2287 (2006)
88. Lou, X.W., Wang, Y., Yuan, C.L., Lee, J.Y., Archer, L.A.: Template-free synthesis of SnO2
hollow nanostructures with high lithium storage capacity. Adv. Mater. 18(17), 2325–2329
(2006)
89. Lee, K.T., Lytle, J.C., Ergang, N.S., Oh, S.M., Stein, A.: Synthesis and rate performance of
monolithic macroporous carbon electrodes for lithium-ion secondary batteries. Adv. Funct.
Mater. 15(4), 547–556 (2005)
90. Liu, Y., Dong, H., Liu, M.L.: Well-aligned “nano-box-beams” of SnO2. Adv. Mater. 16(4),
353 (2004)
91. Wang, Y., Lee, J.Y., Zeng, H.C.: Polycrystalline SnO2 nanotubes prepared via infiltration
casting of nanocrystallites and their electrochemical application. Chem. Mater. 17(15),
3899–3903 (2005)
92. Wang, Y., Zeng, H.C., Lee, J.Y.: Highly reversible lithium storage in porous SnO2 nanotubes
with coaxially grown carbon nanotube overlayers. Adv. Mater. 18(5), 645 (2006)
93. Park, M.S., Wang, G.X., Kang, Y.M., Wexler, D., Dou, S.X., Liu, H.K.: Preparation and
electrochemical properties of SnO2 nanowires for application in lithium-ion batteries.
Angew. Chem. Int. Ed. 46(5), 750–753 (2007)
94. Liu, L.J., Wang, Z.X., Li, H., Chen, L.Q., Huang, X.J.: Al2O3-coated LiCoO2 as cathode
material for lithium ion batteries. Solid State Ion. 152, 341–346 (2002)
95. Kosova, N., Devyatkina, E., Slobodyuk, A., Kaichev, V.: Surface chemistry study of LiCoO2
coated with alumina. Solid State Ion. 179(27–32), 1745–1749 (2008)
96. Wang, Z.X., Liu, L.J., Chen, L.Q., Huang, X.J.: Structural and electrochemical character-
izations of surface-modified LiCoO2 cathode materials for Li-ion batteries. Solid State Ion.
148(3–4), 335–342 (2002)
97. Fey, G.T.K., Lu, C.Z., Huang, J.D., Kumar, T.P., Chang, Y.C.: Nanoparticulate coatings for
enhanced cyclability of LiCoO2 cathodes. J. Power Sources 146(1–2), 65–70 (2005)
98. Fey, G.T.K., Lu, C.Z., Kumar, T.P., Chang, Y.C.: TiO2 coating for long-cycling LiCoO2:
A comparison of coating procedures. Sur. Coat. Technol. 199(1), 22–31 (2005)
99. Zhao, H.L., Ling, G., Qiu, W.H., Zhang, X.H.: Improvement of electrochemical stability of
LiCoO2 cathode by a nano-crystalline coating. J. Power Sources 132(1–2), 195–200 (2004)
100. Fang, T., Duh, J.G., Sheen, S.R.: LiCoO2 cathode material coated with nano-crystallized ZnO
for Li-ion batteries. Thin Solid Films 469, 361–365 (2004)
101. Cho, J., Kim, C.S., Yoo, S.I.: Improvement of structural stability of LiCoO2 cathode during
electrochemical cycling by sol-gel coating of SnO2. Electrochem. Solid State Lett. 3(8),
362–365 (2000)
102. Kim, Y.J., Cho, J.P., Kim, T.J., Park, B.: Suppression of cobalt dissolution from the LiCoO2
cathodes with various metal-oxide coatings. J. Electrochem. Soc. 150(12), A1723–A1725
(2003)
103. Wang, L., Li, J.G., He, X.M., Pu, W.H., Wan, C.R., Jiang, C.Y.: Recent advances in layered
LiNi (x) CoyMn1-x-y O-2 cathode materials for lithium ion batteries. J. Solid State
Electrochem. 13(8), 1157–1164 (2009)
12 Oxide Nanostructures for Energy Storage 299

104. Arunkumar, T.A., Wu, Y., Manthiram, A.: Factors influencing the irreversible oxygen loss
and reversible capacity in layered Li[Li/sub 1/3/Mn/sub 2/3/]O/sub 2/-Li[M]O/sub 2/(M ¼
Mn/sub 0.5-y/Ni/sub 0.5-y/Co/sub 2y/and Ni/sub 1-y/Co/sub y/) solid solutions. Chem.
Mater. 3067, 73 (2007)
105. Lu, Z.H., Beaulieu, L.Y., Donaberger, R.A., Thomas, C.L., Dahn, J.R.: Synthesis, structure,
and electrochemical behavior of Li[NixLi1/3-2x/3Mn2/3-x/3]O-2. J. Electrochem. Soc. 149
(6), A778–A791 (2002)
106. Wu, Y., Manthiram, A.: High capacity, surface-modified layered Li[Li(1-x)/3Mn(2-x)/3Nix/3
Cox/3]O2 cathodes with low irreversible capacity loss. Electrochem. Solid State Lett. 9(5),
A221–A224 (2006)
107. Inoue, T., Sano, M.: An investigation of capacity fading of manganese spinels stored at
elevated temperature. J. Electrochem. Soc. 145(11), 3704–3707 (1998)
108. Wohlfahrt-Mehrens, M., Vogler, C., Garche, J.: Aging mechanisms of lithium cathode
materials. J. Power Sources 127(1–2), 58–64 (2004)
109. Nishiwaki, Y., Terada, Y., Nakai, I.: Study of capacity loss mechanism of lithium manganate
cathode during electrochemical cycles at high temperature by in situ TXRF and XAFS
analyses. Electrochemistry 71(3), 163–168 (2003)
110. Eftekhari, A.: Aluminum oxide as a multi-function agent for improving battery performance
of LiMn2O4 cathode. Solid State Ion. 167(3–4), 237–242 (2004)
111. Kannan, A.M., Manthiram, A.: Surface/chemically modified LiMn2O4 cathodes for lithium-
ion batteries. Electrochem. Solid State Lett. 5(7), A167–A169 (2002)
112. Lai, C.E., Ye, W.Y., Liu, H.Y., Wang, W.J.: Preparation of TiO2-coated LiMn2O4 by carrier
transfer method. Ionics 15(3), 389–392 (2009)
113. Lim, S., Cho, J.: PVP-Assisted ZrO2 coating on LiMn2O4 spinel cathode nanoparticles
prepared by MnO2 nanowire templates. Electrochem. Commun. 10(10), 1478–1481 (2008)
114. Arumugam, D., Kalaignan, G.P.: Synthesis and electrochemical characterizations of Nano-
SiO2-coated LiMn2O4 cathode materials for rechargeable lithium batteries. J. Electroanal.
Chem. 624(1–2), 197–204 (2008)
115. Sun, Y.K., Hong, K.J., Prakash, J.: The effect of ZnO coating on electrochemical cycling
behavior of spinel LiMn2O4 cathode materials at elevated temperature. J. Electrochem. Soc.
150(7), A970–A972 (2003)
116. Liu, D.Q., Liu, X.Q., He, Z.Z.: Surface modification by ZnO coating for improving the
elevated temperature performance of LiMn2O4. J. Alloy. Compd. 436(1–2), 387–391 (2007)
117. Ha, H.W., Yun, N.J., Kim, K.: Improvement of electrochemical stability of LiMn2O4 by
CeO2 coating for lithium-ion batteries. Electrochim. Acta 52(9), 3236–3241 (2007)
118. Park, S.C., Kim, Y.M., Kang, Y.M., Kim, K.T., Lee, P.S., Lee, J.Y.: Improvement of the rate
capability of LiMn2O4 by surface coating with LiCoO2. J. Power Sources 103(1), 86–92
(2001)
119. Park, S.C., Kim, Y.M., Han, S.C., Ahn, S., Ku, C.H., Lee, J.Y.: The elevated temperature
performance of LiMn2O4 coated with LiNi1-XCoXO2 (X ¼ 0.2 and 1). J. Power Sources 107
(1), 42–47 (2002)
120. Chang, H.H., Chang, C.C., Su, C.Y., Wu, H.C., Yang, M.H., Wu, N.L.: Effects of TiO2
coating on high-temperature cycle performance of LiFePO4-based lithium-ion batteries.
J. Power Sources 185(1), 466–472 (2008)
121. Leon, B., Vicente, C.P., Tirado, J.L., Biensan, P., Tessier, C.: Optimized chemical stability
and electrochemical performance of LiFePO4 composite materials obtained by ZnO coating.
J. Electrochem. Soc. 155(3), A211–A216 (2008)
122. Miller, J.R., Simon, P.: Materials science – electrochemical capacitors for energy manage-
ment. Science 321(5889), 651–652 (2008)
123. Conway, B.E.: Electrochemical Supercapacitors: Scientific Fundamentals and Technological
Applications. Kluwer Academic/Plenum, New York (1999)
124. Qu, D.Y., Shi, H.: Studies of activated carbons used in double-layer capacitors. J. Power
Sources 74(1), 99–107 (1998)
300 Y. Yang et al.

125. Wang, D.W., Li, F., Liu, M., Lu, G.Q., Cheng, H.M.: 3D aperiodic hierarchical porous
graphitic carbon material for high-rate electrochemical capacitive energy storage. Angew.
Chem. Int. Ed. 47(2), 373–376 (2008)
126. Balducci, A., Dugas, R., Taberna, P.L., Simon, P., Plee, D., Mastragostino, M., Passerini, S.:
High temperature carbon-carbon supercapacitor using ionic liquid as electrolyte. J. Power
Sources 165(2), 922–927 (2007)
127. Pandolfo, A.G., Hollenkamp, A.F.: Carbon properties and their role in supercapacitors.
J. Power Sources 157(1), 11–27 (2006)
128. Frackowiak, E., Beguin, F.: Carbon materials for the electrochemical storage of energy in
capacitors. Carbon 39(6), 937–950 (2001)
129. Zheng, J.P., Cygan, P.J., Jow, T.R.: Hydrous ruthenium oxide as an electrode material for
electrochemical capacitors. J. Electrochem. Soc. 142(8), 2699–2703 (1995)
130. Hu, C.C., Wang, C.C.: Nanostructures and capacitive characteristics of hydrous manganese
oxide prepared by electrochemical deposition. J. Electrochem. Soc. 150(8), A1079–A1084
(2003)
131. Chen, Z., Qin, Y.C., Weng, D., Xiao, Q.F., Peng, Y.T., Wang, X.L., Li, H.X., Wei, F., Lu, Y.
F.: Design and synthesis of hierarchical nanowire composites for electrochemical energy
storage. Adv. Funct. Mater. 19(21), 3420–3426 (2009)
132. Zhang, H., Cao, G.P., Wang, Z.Y., Yang, Y.S., Shi, Z.J., Gu, Z.N.: Growth of manganese
oxide nanoflowers on vertically-aligned carbon nanotube arrays for high-rate electrochemical
capacitive energy storage. Nano Lett. 8(9), 2664–2668 (2008)
133. Zhang, Y.P., Sun, X.W., Pan, L.K., Li, H.B., Sun, Z., Sun, C.P., Tay, B.K.: Carbon nanotube-
ZnO nanocomposite electrodes for supercapacitors. Solid State Ion. 180(32–35), 1525–1528
(2009)
134. Zhang, H., Cao, G.P., Yang, Y.S.: Carbon nanotube arrays and their composites for electro-
chemical capacitors and lithium-ion batteries. Energy Environ. Sci. 2(9), 932–943 (2009)
135. Galizzio, D., Tantardi, F., Trasatti, S.: Ruthenium dioxide – new electrode material. 1. Behav-
ior in acid solutions of inert electrolytes. J. Appl. Electrochem. 4(1), 57–67 (1974)
136. Hadzijordanov, S., Angersteinkozlowska, H., Conway, B.E.: Surface oxidation and H depo-
sition at ruthenium electrodes – resolution of component processes in potential-sweep
experiments. J. Electroanal. Chem. 60(3), 359–362 (1975)
137. Hadzijordanov, S., Angersteinkozlowska, H., Vukovic, M., Conway, B.E.: Reversibility and
growth-behavior of surface oxide-films at ruthenium electrodes. J. Electrochem. Soc. 125(9),
1471–1480 (1978)
138. Trasatti, S., Buzzanca, G.: Ruthenium dioxide – new interesting electrode material – solid
state structure and electrochemical behaviour. J. Electroanal. Chem. 29(2), A1 (1971)
139. Weston, J.E., Steele, B.C.H.: Proton diffusion in crystalline ruthenium dioxide. J. Appl.
Electrochem. 10(1), 49–53 (1980)
140. Arikado, T., Iwakura, C., Tamura, H.: Electrochemical behavior of ruthenium oxide electrode
prepared by thermal-decomposition method. Electrochim. Acta 22(5), 513–518 (1977)
141. Fischer, A.E., Pettigrew, K.A., Rolison, D.R., Stroud, R.M., Long, J.W.: Incorporation of
homogeneous, nanoscale MnO2 within ultraporous carbon structures via self-limiting elec-
troless deposition: Implications for electrochemical capacitors. Nano Lett. 7(2), 281–286
(2007)
142. Jang, J.H., Han, S., Hyeon, T., Oh, S.M.: Electrochemical capacitor performance of hydrous
ruthenium oxide/mesoporous carbon composite electrodes. J. Power Sources 123(1), 79–85
(2003)
143. Sugimoto, W., Iwata, H., Yasunaga, Y., Murakami, Y., Takasu, Y.: Preparation of ruthenic
acid nanosheets and utilization of its interlayer surface for electrochemical energy storage.
Angew. Chem. Int. Ed. 42(34), 4092–4096 (2003)
144. Long, J.W., Swider, K.E., Merzbacher, C.I., Rolison, D.R.: Voltammetric characterization of
ruthenium oxide-based aerogels and other RuO2 solids: The nature of capacitance in
nanostructured materials. Langmuir 15(3), 780–785 (1999)
12 Oxide Nanostructures for Energy Storage 301

145. Shen, X.F., Ding, Y.S., Liu, J., Cai, J., Laubernds, K., Zerger, R.P., Vasiliev, A., Aindow, M.,
Suib, S.L.: Control of nanometer-scale tunnel sizes of porous manganese oxide octahedral
molecular sieve nanomaterials. Adv. Mater. 17(7), 805 (2005)
146. Kim, J.K., Manthiram, A.: A manganese oxyiodide cathode for rechargeable lithium
batteries. Nature 390(6657), 265–267 (1997)
147. Shinomiya, T., Gupta, V., Miura, N.: Effects of electrochemical-deposition method and
microstructure on the capacitive characteristics of nano-sized manganese oxide. Electrochim.
Acta 51(21), 4412–4419 (2006)
148. Cheng, J., Cao, G.P., Yang, Y.S.: Characterization of sol-gel-derived NiOx xerogels as
supercapacitors. J. Power Sources 159(1), 734–741 (2006)
149. Adelkhani, H., Ghaemi, M., Jafari, S.M.: Novel nanostructured MnO2 prepared by pulse
electrodeposition: characterization and electrokinetics. J. Mater. Sci. Technol. 24(6),
857–862 (2008)
150. Yuan, J.K., Li, W.N., Gomez, S., Suib, S.L.: Shape-controlled synthesis of manganese oxide
octahedral molecular sieve three-dimensional nanostructures. J. Am. Chem. Soc. 127(41),
14184–14185 (2005)
151. Yin, M., O’Brien, S.: Synthesis of monodisperse nanocrystals of manganese oxides. J. Am.
Chem. Soc. 125(34), 10180–10181 (2003)
152. Zhong, X.H., Xie, R.G., Sun, L.T., Lieberwirth, I., Knoll, W.: Synthesis of dumbbell-shaped
manganese oxide nanocrystals. J. Phys. Chem. 110(1), 2–4 (2006)
153. Zhang, L.C., Liu, Z.H., Lv, H., Tang, X.H., Ooi, K.: Shape-controllable synthesis and
electrochemical properties of nanostructured manganese oxides. J. Phys. Chem. 111(24),
8418–8423 (2007)
154. Wu, M.S., Chiang, P.C.J., Lee, J.T., Lin, J.C.: Synthesis of manganese oxide electrodes with
interconnected nanowire structure as an anode material for rechargeable lithium ion batteries.
J. Phys. Chem. 109(49), 23279–23284 (2005)
155. Cheng, F.Y., Chen, J., Gou, X.L., Shen, P.W.: High-power alkaline Zn-MuO(2) batteries
using gamma-MnO2 nanowires/nanotubes and electrolytic zinc powder. Adv. Mater. 17(22),
2753 (2005)
156. Subramanian, V., Zhu, H.W., Vajtai, R., Ajayan, P.M., Wei, B.Q.: Hydrothermal synthesis
and pseudocapacitance properties of MnO2 nanostructures. J. Phys. Chem. 109(43),
20207–20214 (2005)
157. Oaki, Y., Imai, H.: One-pot synthesis of manganese oxide nanosheets in aqueous solution:
Chelation-mediated parallel control of reaction and morphology. Angew. Chem. Int. Ed. 46
(26), 4951–4955 (2007)
158. Pang, S.C., Anderson, M.A., Chapman, T.W.: Novel electrode materials for thin-film
ultracapacitors: comparison of electrochemical properties of sol-gel-derived and electrode-
posited manganese dioxide. J. Electrochem. Soc. 147(2), 444–450 (2000)
159. Hu, C.C., Tsou, T.W.: Ideal capacitive behavior of hydrous manganese oxide prepared by
anodic deposition. Electrochem. Commun. 4(2), 105–109 (2002)
160. Kim, H., Popov, B.N.: Synthesis and characterization of MnO2-based mixed oxides as
supercapacitors. J. Electrochem. Soc. 150(3), D56–D62 (2003)
161. Hu, L., Pasta, M., La Mantia, F., Cui, L., Jeong, S., Deshazer, H.D., Choi, J.W., Han, S.M.,
Cui, Y.: Stretchable, porous, and conductive energy textiles. Nano Lett. 10(2), 708–714
(2010)
162. Toupin, M., Brousse, T., Belanger, D.: Charge storage mechanism of MnO2 electrode used in
aqueous electrochemical capacitor. Chem. Mat. 16(16), 3184–3190 (2004)
163. Lee, H.Y., Manivannan, V., Goodenough, J.B.: Electrochemical capacitors with KCl electro-
lyte. C R Acad. Sci. Ser. II C 2(11–13), 565–577 (1999)
164. Lee, H.Y., Goodenough, J.B.: Supercapacitor behavior with KCl electrolyte. J. Solid State
Chem. 144(1), 220–223 (1999)
165. Zhang, F.B., Zhou, Y.K., Li, H.L.: Nanocrystalline NiO as an electrode material for electro-
chemical capacitor. Mater. Chem. Phys. 83(2–3), 260–264 (2004)
302 Y. Yang et al.

166. Srinivasan, V., Weidner, J.W.: Studies on the capacitance of nickel oxide films: Effect of
heating temperature and electrolyte concentration. J. Electrochem. Soc. 147(3), 880–885
(2000)
167. Xing, W., Huang, C.C., Zhuo, S.P., Yuan, X., Wang, G.Q., Hulicova-Jurcakova, D., Yan, Z.
F., Lu, G.Q.: Hierarchical porous carbons with high performance for supercapacitor
electrodes. Carbon 47(7), 1715–1722 (2009)
168. Srinivasan, V., Weidner, J.W.: An electrochemical route for making porous nickel oxide
electrochemical capacitors. J. Electrochem. Soc. 144(8), L210–L213 (1997)
169. Lin, C., Ritter, J.A., Popov, B.N.: Characterization of sol-gel-derived cobalt oxide xerogels as
electrochemical capacitors. J. Electrochem. Soc. 145(12), 4097–4103 (1998)
170. Srinivasan, V., Weidner, J.W.: Capacitance studies of cobalt oxide films formed via electro-
chemical precipitation. J. Power Sources 108(1–2), 15–20 (2002)
171. Armelao, L., Barreca, D., Gross, S., Martucci, A., Tieto, M., Tondello, E.: Cobalt oxide-based
films: sol-gel synthesis and characterization. J. Non Cryst. Solids 293, 477–482 (2001)
172. Wang, Y., Yang, W.S., Zhang, S.C., Evans, D.G., Duan, X.: Synthesis and electrochemical
characterization of Co-Al layered double hydroxides. J. Electrochem. Soc. 152(11),
A2130–A2137 (2005)
173. Chuang, P.Y., Hu, C.C.: The electrochemical characteristics of binary manganese-cobalt
oxides prepared by anodic deposition. Mater. Chem. Phys. 92(1), 138–145 (2005)
174. Prasad, K.R., Miura, N.: Electrochemically synthesized MnO2-based mixed oxides for high
performance redox supercapacitors. Electrochem. Commun. 6(10), 1004–1008 (2004)
175. Hu, C.C., Cheng, C.Y.: Ideally pseudocapacitive behavior of amorphous hydrous cobalt-
nickel oxide prepared by anodic deposition. Electrochem. Solid State Lett. 5(3), A43–A46
(2002)
176. Wu, N.L.: Nanocrystalline oxide supercapacitors. Mater. Chem. Phys. 75(1–3), 6–11 (2002)
177. Broughton, J.N., Brett, M.J.: Investigation of thin sputtered Mn films for electrochemical
capacitors. Electrochim. Acta 49(25), 4439–4446 (2004)
178. Kim, I.H., Kim, J.H., Kim, K.B.: Electrochemical characterization of electrochemically
prepared ruthenium oxide/carbon nanotube electrode for supercapacitor application.
Electrochem. Solid State Lett. 8(7), A369–A372 (2005)
179. Wu, Y.T., Hu, C.C.: Effects of electrochemical activation and multiwall carbon nanotubes on
the capacitive characteristics of thick MnO2 deposits. J. Electrochem. Soc. 151(12),
A2060–A2066 (2004)
180. Hughes, M., Shaffer, M.S.P., Renouf, A.C., Singh, C., Chen, G.Z., Fray, J., Windle, A.H.:
Electrochemical capacitance of nanocomposite films formed by coating aligned arrays of
carbon nanotubes with polypyrrole. Adv. Mater. 14(5), 382–385 (2002)
181. Lee, B.J., Sivakkumar, S.R., Ko, J.M., Kim, J.H., Jo, S.M., Kim, D.Y.: Carbon nanofibre/
hydrous RuO2 nanocomposite electrodes for supercapacitors. J. Power Sources 168(2),
546–552 (2007)
Chapter 13
Metal Oxide Resistive Switching Memory

Shimeng Yu, Byoungil Lee, and H.‐S. Philip Wong

13.1 Introduction

Information storage device is a key component of nanoelectronic systems.


Conventional memories such as SRAM, DRAM, and FLASH are facing formidable
device scaling challenges. Emerging memory device concepts have been actively
pursued both in industry and academia in hopes of finding solutions for future
information storage needs [1, 2]. The ideal characteristics for a memory device
include fast programming speed (~ns), long retention time (>10 years), low power
consumption, good reliability, high integrated density, and continued scalability.
Several candidates have been proposed to achieve the above goals, such as magne-
toresistive random access memory (MRAM) [3], ferroelectric random access
memory (FeRAM) [4]. In recent years, memory devices based on the electrically
switchable resistance phenomenon have been extensively studied. The basic prin-
ciple of this kind of memory is to use a high resistance state (HRS) or low resistance
state (LRS) to store the information data “0” or “1,” and the transition between the
two states can be triggered by electrical inputs. Roughly speaking, resistive switching
memories can be classified into three groups [5]: (1) phase-change memory (PCM)
based on chalcogenides [6–9], which relies on the temperature-induced change
between the crystalline phase (corresponding to LRS) and the amorphous phase
(corresponding to HRS) (2) programmable-metallization-cell (PMC) memory based
on solid electrolytes or polymers [10–13], which relies on the formation
(corresponding to LRS) or the rupture (corresponding to HRS) of a metallic
conducting bridge between two electrodes, and (3) resistance random access memory
(RRAM) based on metal oxides. Among these resistive switching memories, metal
oxide memory is promising for practical applications due to its compatibility with

S. Yu • B. Lee • H.‐S.P. Wong (*)


Department of Electrical Engineering, Center for Integrated Systems,
Stanford University, Stanford, CA, USA
e-mail: hspwong@stanford.edu

J. Wu et al. (eds.), Functional Metal Oxide Nanostructures, Springer Series 303


in Materials Science 149, DOI 10.1007/978-1-4419-9931-3_13,
# Springer Science+Business Media, LLC 2012
304 S. Yu et al.

Fig. 13.1 (a) Schematic of metal–insulator–metal (MIM) structure for metal oxide memory cell,
and schematic of metal oxide memory’s I–V curves, showing two modes of operation: (b) unipolar
and (c) bipolar

silicon CMOS fabrication technology. Here, we review the research progress of metal
oxide memory including the physical mechanism of switching, state-of-the-art device
performances, as well as device cell structure for integration into a memory array.
The negative differential resistances phenomenon in oxides was first reported in
the 1960s [14, 15], then it was reviewed by Dearnaley et al. [16] in 1970. Recent
work on the resistive switching metal oxide memory can be traced back to the
discovery of hysteresis I–V characteristics in perovskite oxides [17–19] such as
Pr0.7Ca0.3MnO3, SrTiO3, SrZrO3, etc., in the late 1990s and the early 2000s. Since
2004, the research activities have focused on binary metal oxides [20–26] such as
NiO, TiO2, ZrO2, ZnO, Cu2O, Al2O3, HfO2, etc., because of the simplicity of the
material and good compatibility with silicon CMOS fabrication process. Recently,
the performance of perovskite oxide memories and related physical mechanism
of resistive switching were reviewed by Waser et al. [27, 28] and Sawa [29].
Therefore, here we focus this review on binary metal oxide memory.

13.1.1 Device Operation

It is necessary to first introduce some basic concepts and terminologies about metal
oxide memory. The typical metal oxide memory cell is a simple metal–insula-
tor–metal (MIM) structure, as shown in Fig. 13.1a. The switching event from high
resistance state (HRS) to low resistance state (LRS) is referred to as the “set”
process. Conversely, the switching event from LRS to HRS is referred to as the
“reset” process. In some cases, for the fresh samples in its initial resistance state
(IRS), a larger voltage is needed to trigger on the resistive switching behaviors for
the subsequent cycles. This is called the “electroforming” or “forming” process.
The switching modes of metal oxide memory can be broadly classified into two
switching modes: unipolar and bipolar. Unipolar switching means the switching
direction depends on the amplitude of the applied voltage but not on the polarity,
thus set/reset can occur at the same polarity, and usually it can symmetrically occur
13 Metal Oxide Resistive Switching Memory 305

at both forward and reversed voltages, as illustrated in Fig. 13.1b. Bipolar switching
means the switching direction depends on the polarity of the applied voltage, thus set
can only occur at one polarity and reset can only occur at the reverse polarity,
as illustrated in Fig. 13.1c. To avoid a hard dielectric breakdown in the set process,
it is recommended to enforce a set compliance current, which is usually provided by
the semiconductor parameter analyzer, or more practically, by a memory cell
selection (or access) transistor/diode or a series resistor. To read the data from the
cell, a small voltage is applied that does not affect the state of the memory cell.
For nonvolatile application, the cell should retain its state at standby mode and free
from disturb from reading/writing of neighboring cells.

13.1.2 Device Characteristics

To further understand how the metal oxide memory works, the device characteristics
of HfO2-based memory cell [30, 31] from Industrial Technology Research Institute
(ITRI) are presented as an example. Figure 13.2a shows the transmission electron
microscopy (TEM) image of the concave structure device of TiN/Ti/HfOx/TiN stack
with 30 nm cell size; here 5 nm HfOx layer is used for the main switching layer,
TiN is used for the electrode material, and a thin Ti buffer layer is used for improving
the switching stability. Figure 13.2b shows the typical I–V curve of such memory cell.
A 200 mA set compliance current is enforced, and the device exhibits bipolar
switching. Figure 13.2c shows the switching endurance testing result. The set/reset
programming condition is +1.5 V/1.4 V pulse with 500 ms width, and it is seen that
after 106 switching cycles, the HRS/LRS resistance ratio is still larger than 100,
although there is some degradation. Figure 13.2d shows the data retention testing
result, the measurement is performed at 150 C, and it is seen that 10 years lifetime is
expected using a simple linear extrapolation. Besides, fast switching speed (~5 ns),
and multi-bit storage potential is also demonstrated in the HfO2 memory.
Besides HfO2, other binary metal oxides such as NiO, TiO2, ZrO2, ZnO, Cu2O,
Al2O3 have been extensively explored. Up to now, dozens of binary metal oxides
have been found to exhibits electric-field-induced bistable resistance switching
behavior. Most of the metals are transition metals, and some are lanthanide series
metals. The materials for switching layer and electrode layer are summarized in
Table 13.1. Note that some nitrides, e.g., TiN are also used for the electrode layer.

13.2 Possible Physical Mechanism for Resistive


Switching in Metal Oxides

The physical mechanism for resistive switching in metal oxide memory is


still under a heated debate. According to ITRS 2007 [32], resistive switching
mechanism can be roughly classified into three groups: thermal effect, electronic
306 S. Yu et al.

Fig. 13.2 (a) Transmission electron microscopy (TEM) image of the concave structure device
of TiN/ Ti/ HfOx/ TiN stack; (b) typical I–V curve of the device with the cell size of 30 nm;
(c) endurance testing by 500 ms pulse: successive 106 switching cycles is achieved; (d) retention
testing at 150 C: 10 years lifetime is expected. Reprinted from [30, 31]

effect, or ionic effect. Thermal effect refers to the thermal dissolution of conductive
filaments triggered by local Joule heating and works like a traditional household
fuse albeit at the nanoscale [33, 34]. It is also referred to as fuse/anti-fuse switching
type. The electronic effect mechanism conjectures that injected charges are trapped
by interface defects. These trapped charges modify the Schottky barrier height
between electrodes and oxides and thereby changing the conductance through the
MIM structure [35, 36]. Another case of electronic effect is Mott metal–insulator
transition due to the strong electron–electron correlation in some transition metal
oxides [37, 38]. Ionic effect refers to the migration of ions with related electro-
chemical reactions to form a conducting bridge between electrodes [39, 40].
It is also referred to as reduction/oxidation type. Actually, this ionic effect is
just the operating principle of programmable-metallization-cell memory in solid
electrolytes or polymers as mentioned before. Metal oxides can also serve as fast
13 Metal Oxide Resistive Switching Memory 307

Table 13.1 Summary of the materials that has been used for binary metal oxide memory. Yellow
metals corresponding binary oxides are used for switching layer; blue metals are used for electrode
layer

ions conductors for some particular metal ions like Ag+ and Cu2+ [41, 42]. Here we
plan to focus on the intrinsic properties of metal oxides, thus we would not include
those metal oxides memory with Ag or Cu electrodes in the following discussions.
So far, no single model mentioned above can explain all the experimental results
obtained in various materials. It seems that the thermal dissolution model can address
parts of the unipolar switching characteristics, while the ionic migration model can
explain most of the bipolar switching characteristics. However, the distinction
between the two switching behaviors is ambiguous. Here we discuss several key
issues such as the conducting mechanism, the nature of electroforming/set/reset
process, the effect of electrode materials on switching modes, with the aim of seeking
a rational mechanism for both unipolar and bipolar switching characteristics.

13.2.1 Conduction Mechanism

A lot of conductive atomic force microscopy (C-AFM) measurements [43–47] have


been reported in the literature, revealing that conductive filaments (CFs) with
nanoscale diameters are formed in the oxide layer during the set process, and
large conducting current passes through these filaments in LRS. Figure 13.3
shows one of these C-AFM measurements results [45]. Typically, CFs are sparsely
308 S. Yu et al.

Fig. 13.3 Conductive atomic force microscopy (C-AFM) of LRS (a) and HRS (b) conductance in
Al2O3 memory, showing filamentary conduction dominates in LRS (scanned area: 2 mm  2 mm;
maximum y scale: (a) 105 nA, (b) 5 nA). Reprinted from [45]

and nonhomogeneously distributed under the electrodes. This means that the
conducting area in LRS is an extremely small portion of the entire electrode area,
which is typically quite large in most experiments. Therefore, the conductance in
LRS would not decrease as much as the conductance in HRS does when the
electrode area is decreased, resulting a larger HRS/LRS resistance ratio, which is
a benefit of scaling down memory cell size. It has been experimentally
demonstrated that these CFs can be formed by C-AFM tip individually [48], but
in most cases multiple CFs exist in the oxide layer, which is believed to be
responsible for the multilevel switching behaviors exhibited in some metal oxide
memory devices [49, 50]. Although the filamentary conduction mechanism is
widely recognized, some researchers argued that a bulk conduction mechanism
dominates the LRS resistance in their TiO2 devices [51]. They proposed that
charged dopants (mainly oxygen vacancies) migrate under a voltage bias to the
TiO2/electrode interface. The interfacial changes modify the Schottky barrier
height at the interface and lead to the resistive switching in their devices. In essence,
such arguments do not contradict with the filamentary conduction mechanism,
provided that the so-called bulk conduction is regarded as many parallel CFs. It
should be noted that if CFs are dense enough in ultra-scaled devices, the bulk
conduction behavior may arise. A usual experiment that aims at distinguishing
filamentary or bulk conduction is to measure the trend of the HRS/LRS resistance
ratio vs. the cell area. If the ratio goes up when the cell area is scaled down,
filamentary conduction prevails. If the ratio remains almost constant when the
cell is scaled down, bulk conduction prevails. However, so far, these experiments
have not been performed for devices with truly nanometer scale size (<100 nm)
memory cells and therefore the results are not yet conclusive.
The CFs are conjectured to be made up of oxygen vacancies (Vo) in most metal
oxide memory devices, and this assumption was confirmed by micro X-ray fluores-
cence (XRF) and X-ray absorption near-edge spectroscopy (XANES) [52]. It is
well-known that Vo can act as an effective donor in n-type metal oxides. Ab initio
13 Metal Oxide Resistive Switching Memory 309

calculation of the rutile TiO2 electronic structure [53] reveals that Vo can produce a
defect state within the band gap. This state is occupied by two electrons that are
localized on Ti 3d orbital of the nearest Ti atoms to Vo. If a chain of such Vo forms,
it is expected that electrons delocalization occurs and the hopping probability can
increase remarkably along these CFs and ultimately lead to the LRS. Although Vo
plays an important role in the switching behaviors of most metal oxide memory
devices, the composition of CFs is not limited to Vo. There are quite a few reports
that the CFs in NiO memory are composed of excess metallic Ni in NiO. This
suggestion is confirmed by measurement of the dependence of the LRS resistance
on temperature [54] and X-ray photoelectron spectroscopy (XPS) composition
analysis [55]. It was further verified by electron energy loss spectroscopy (EELS)
that there exists a Ni-rich phase at the grain boundaries, which implies that the CFs
of Ni are formed as precipitates at the grain boundaries [56]. A simple way to
distinguish whether the CFs are metallic or semiconducting is to measure the LRS
resistance temperature dependency. If the LRS resistance goes up with the increase
of temperature, the CFs are metallic and may consist of metal precipitates. On the
contrary, if the LRS resistance drops with the increase of temperature, CFs are
semiconducting and may consist of Vo.
There are many efforts to fit the I–V characteristics of HRS and LRS to current
conduction models in the literature. Most of the reports show an Ohmic relationship
in the LRS. And in the semiconducting CFs, Mott variable hopping conduction [57]
is proposed to dominate in LRS, and this assumption was supported by some
temperature varying measurements [58, 59] and AC conductance measurement
[60]. But, the conduction fitting results in HRS are quite diverse, Poole–Frenkel
emission (I ~ V*exp(V1/2)) [61, 62], Schottky emission (I ~ exp(V1/2)) [63, 64], the
space charge limited current (SCLC) characteristic (the Ohmic region I ~ V,
followed by the child’s square law region I ~ V2) [65, 66], were observed in various
metal oxide memories. The diverse I–V characteristics in HRS involving different
leakage mechanism may be associated with the different dielectric properties or
different fabrication processes conditions, e.g., annealing temperature, annealing
ambient, and the properties of the interface between the oxides and the electrodes.
Nevertheless, the key issue of metal oxides memory modeling is not the leakage
current mechanism in HRS, but to investigate the mechanism that triggers the
resistive switching. So in the next section, we will discuss the forming/set and
reset process, respectively.

13.2.2 Electroforming/Set/Reset Process with Oxygen Migration

Dielectric breakdown is a process of local materials transitioning from insulating to


conductive under a high external electric field. This phenomenon is well-known for
gate dielectrics in MOSFETs and has been studied for a long time. Recently, by
site-specific structural analysis using high resolution transmission electron micros-
copy (HR-TEM) with EELS, Li et al. [67] revealed that after dielectric breakdown,
310 S. Yu et al.

a b
2+ 0 106
Ni Ni Electrical impulse
(854.3 eV) (852.8 eV)
105 Solid: before
Ni
Open: after
Intensity (arb. inits)

104

Intensity(cps)
Pt
103

102 Oxygen
Deposited at 5% O2
101
Deposited at 30% O2
100
860 858 856 854 852 850 25 50 75 100 120
Binding energy (eV) Length (nm)

Fig. 13.4 (a) X-ray photoelectron spectroscopy (XPS) analysis of NiO memory; the sample
deposited at an oxygen partial pressure of 5% shows the coexistence of Ni (852.8 keV) and NiO
(854.3 keV) peaks, which exhibited bistable resistive switching; the sample deposited at an oxygen
partial pressure of 30% did not exhibit any stable resistive switching phenomenon, suggesting CFs
may consist of metallic Ni precipitates. (b) Secondary-ion mass spectroscopy (SIMS) data in NiO
memory, showing the migration of oxygen atoms toward the electrodes after electrical switching.
Reprinted from [71]

oxygen atoms in gate dielectric SiO2 were missing. Substoichiometric silicon oxide
SiOx (with x < 2) was formed, and the local energy gap was lowered with interme-
diate bonding state of silicon atoms Si1+, Si2+, and Si3+ in the percolation leakage
path. Also, chemical bond breakage and the local Joule heating due to large current
flowing through the percolation path are believed to be the main driving forces
leading to the oxygen dissociation.
Similarly, in metal oxides, the electroforming/set process is interpreted to be
some kind of soft dielectric breakdown [60, 68, 69]. Yang et al. [70] claimed that
they observed oxygen gas bubbles at the electrode surface of their bipolar TiO2
memory devices in the electroforming process. They regarded the electroforming
process in metal oxide memory as an electro-reduction and Vo creation process
caused by high electric field, which is then enhanced by local Joule heating. During
electroforming, Vo are created and drift toward the cathode, forming localized CFs
in the oxide. Simultaneously, oxygen ions drift toward the anode where they
discharge to evolve oxygen gas. Lee et al. [71] investigated the composition change
in their unipolar NiO memory devices during the set process by secondary ion mass
spectroscopy (SIMS), as shown in Fig. 13.4. The authors suggested that the oxygen
atoms within the NiO layer migrated toward the Pt electrode after the switching,
thus leaving the Ni rich filaments in the NiO layer.
Therefore, in both unipolar and bipolar switching metal oxide memory devices,
the electroforming/set process is conjectured to be associated with the generation of
Vo and the migration of the oxygen atoms, generating CFs consist of either Vo or
metal precipitates. In fresh samples, usually the amount of Vo is small, thus a high
voltage is needed to trigger the electroforming process. After electroforming, the
devices have switched to LRS, and a sufficient amount of Vo is present. In subsequent
13 Metal Oxide Resistive Switching Memory 311

Fig. 13.5 XPS spectra of Ti/MnO2 interface. (a) The O 1s core levels spectra of the lattice peak
(529.45 eV) and an additional nonlattice peak (531.33 eV), the change of nonlattice oxygen
indicates oxygen ions migration during switching; (b) the spectra of Ti 2p, the shift of Ti 2p
peak indicates an formation of interfacial TiOx layer, which can be regarded as an oxygen
reservoir. Reprinted from [79]

switching cycles, only a portion of the Vo, the ones near the anode, can be recombined
during the reset process. This is why the set voltage would be smaller than forming
voltage and the resistance in HRS would be much smaller than the resistance in IRS.
Through fitting of electrical parameters, the formation and rupture of the CFs is
estimated to occur in a localized region (3–10 nm) thick near the anode [72]. The
remaining Vo rich region is referred to as the “virtual cathode” [28].
Obviously, it is not desirable to have a large electroforming voltage in practical
applications. Thus significant efforts have been made to achieve the so-called
forming-free devices. The forming voltage is linearly dependent on the thickness of
the oxide film [30, 73–75]. So a thinner oxide film is effective in reducing the forming
voltage. Lee et al. [30] claimed that the as-fabricated atomic layer deposition (ALD)
HfO2 device is free from the electroforming process as the thickness of the film is
thinned to be 3 nm. Besides, controlling the annealing ambient during deposition to
obtain oxygen-deficient films is also helpful in reducing the forming voltage [76–78].
So far, we have stated that the nature of dielectric breakdown is the formation of
oxygen-deficient percolation leakage paths. Note that gate dielectric breakdown is a
permanent and irreversible process, while the switching process in metal oxide
memory is a reversible process. So the memory device should have a location to
store the missing oxygen atoms during the set process and then drive them back
during the reset process. The concept of an “oxygen reservoir” is thus envisioned.
Usually, the anode electrode materials act as the oxygen reservoir, and it also
prevents the oxygen from escaping from the device to the ambient. The anode
electrode material then provides a source of oxygen for reset process. Figure 13.5
illustrates how the oxygen reservoir works in a Ti/MnO2/Pt memory device [79]. XPS
reveals that amount of nonlattice oxygen became less in LRS than in HRS, and the
corresponding shift of Ti 2p peak indicates that an interfacial TiOx layer was formed.
This observation suggests that during the set process, oxygen ions moved toward the
Ti anode when a positive bias was applied, where they reduced the Ti and formed
312 S. Yu et al.

TiOx. This interfacial TiOx layer is believed to act as an oxygen reservoir that stores
oxygen atoms. When a negative bias was applied to the Ti anode, oxygen ions moved
away from the interfacial layer toward the oxide bulk layer and recombine with Vo.
Therefore, the bipolar switching characteristics are assumed to be the formation and
rupture of the CFs consisting of Vo associated with oxygen ion migration.
In the bipolar switching case, it is straightforward to conceive of oxygen ions
migrating away from the interface, since oxygen ions are negative charged. Under a
negative bias, oxygen ions can be pushed toward the oxide bulk layer and then
recombine with Vo and rupture the CFs if the CFs are made up of Vo. Yoshida
et al. [80] performed C-AFM writing experiments in 18O tracer gas atmosphere and
related composition analysis by SIMS on the NiO films. Their results suggest that
external oxygen can penetrate into the oxide layer, and the composition change
of the surface region may be responsible for the resistance change. Jeong et al. [81]
performed EELS oxygen element mapping experiments in HRS and LRS of
Al/TiO2/Al structure. The results suggest that the drift of oxygen ions by the applied
bias is the microscopic origin of the bistable resistivity switching. So the reset
mechanism for bipolar switching is quite unambiguous in the literature [82–84].
Recently, the nonlinear dynamical properties of bipolar switching TiO2 were
modeled by the drift and diffusion of ionized dopants (oxygen vacancies) in the
oxide thin film [85–87]. In essence, the description using oxygen vacancies is
equivalent to the description using oxygen ions here.
In the unipolar switching case, the mainstream viewpoint of reset is due to a
thermal dissolution of CFs by local Joule heating [88–90]. Electro-thermal calcula-
tion [88] suggests that the local temperature can rise to around 600 K during the
switching process. But this model is only phenomenological and does not reveal the
microscopic nature of the rupture of CFs. First, it does not address whether the loss
of oxygen during set process can be recaptured in subsequent switching cycles or
not, and the model does not address the question of whether the set process is a
reversible process. Second, if the CFs are assumed to be simply “dissolved” by local
Joule heating, then the CFs should rupture in the middle of the filament where the
temperature is highest [91], since the metal electrodes work as a large heat sink. The
local Joule heating conjecture contradicts with the many experimental observations
that switching occurs in the region near the anode as mentioned before. Alterna-
tively, Lee et al. [71] regarded the reset process of their unipolar NiO memory as an
oxidation process of the metallic CFs, which is a consequence of the thermally
activated oxygen atoms diffusing from anode side followed by oxidation of the
Ni-rich CFs. This argument acknowledges that the reset in unipolar switching is a
Joule heating-assisted process. At the same time, this argument is essentially
consistent with the electrochemical ionic migration model for bipolar switching
devices. The model provides an inherent link between unipolar and bipolar
switching. In the next section, we discuss another important question of why
some materials exhibit unipolar characteristics, while some other materials exhibit
bipolar characteristics before we make the final conclusion.
13 Metal Oxide Resistive Switching Memory 313

Fig. 13.6 (a) I–V characteristics of Pt/ZnO/Pt and TiN/ZnO/Pt devices, (b) Pt/NiO/Pt and Pt/NiO/
SrRuO3 devices, showing the electrode materials effect on switching modes. Data are collected
from [Seo 04] [20], [Chang 08] [61], [Xu 08] [95], [Choi 09] [96]

13.2.3 The Effect of Electrode Materials on Switching Modes

There are many reports [92–94] on the effect of electrode materials on the resistive
switching characteristics of metal oxide memory. In most cases, the switching
mode is not an inherent property of the oxide film but an effect of the interaction
between the oxide and electrode materials. Waser et al. pointed out that for bipolar
switching the MIM structure should have some asymmetry, such as different
electrode materials, while for uniploar switching the MIM structure may be sym-
metric [27]. Figure 13.6 shows two examples: the Pt/ZnO/Pt [61] and Pt/NiO/Pt [20]
314 S. Yu et al.

structure show uniplolar switching, while the TiN/ZnO/Pt [95] and Pt/NiO/SrRuO3
[96] structure show bipolar switching. Besides, many other metal oxides can show
either unipolar or bipolar behavior depending on what electrode materials are used.
Pt/ZrO2/Pt [97], Pt/TiO2/Pt [21], Pt/HfO2/Pt [62] all show unipolar switching, and
Ti/ZrO2/Pt [97], Pt/TiO2/TiN [98], TiN/HfO2/Pt [99] all show bipolar switching.
It should be noted that the unipolar switching I–V curves are usually symmetric,
since there is no distinction of either of the electrodes. As a result, for unipolar
switching devices, bipolar switching behaviors can also be demonstrated [100],
which means the set can be triggered at one polarity and reset can be triggered at
the reversed polarity. Such case is also referred to “nonpolar” switching [74]. But in
bipolar switching devices, there is always one polarity that reset cannot occur when
the same polarity bias as set process is applied.
The electrode materials can be roughly classified into two groups: one is noble
metals, such as Pt, Au, Ru, etc., which are resistant to oxidation; the other one is non-
noble metals, such as Ti, Al, Cu, Ni, W, TiN, TaN, etc., which may suffer oxidation
during the switching process. As observed above, devices using noble metals always
show unipolar behaviors. Even with the Pt/TiO2/TiN [98] structure, unipolar
switching can occur at the Pt side. However, the unipolar switching cannot occur at
TiN side, and only bipolar switching is obtained at the TiN side. As mentioned
before, during the set process, the oxygen ions migrate toward the anode. Thus, the
anode materials may react with the oxygen and form an interfacial layer if the anode
materials are oxidizable. Zhou et al. [101] compared the different roles played by the
Pt electrode and the TaN electrode on the switching modes of CuxO memory.
Figure 13.7a, b shows depth profile by Auger Electron Spectroscopy (AES) for Pt/
CuxO/Cu structure and TaN/CuxO/Cu structure, we can see that the oxygen concen-
tration has an obvious shift toward the TaN layer, which indicates that a reaction
occurs between the TaN electrode and oxide film. This interfacial TaON layer was
further confirmed by XPS binding energy depth spectra. The bright gray areas under
the TaN electrode in the TEM picture of Fig. 13.7c are assumed to be the TaON ultra-
thin layer. And Fig. 13.7d illustrates that with Pt as top electrode, reset can occur at
both bias polarities, showing a unipolar switching behavior, while with TaN as top
electrode, reset can occur only under the reversed bias, showing a bipolar switching
behavior. Therefore, we can infer that if noble metals like Pt are used for the
electrode, it hardly forms such interfacial oxide layer. And without the diffusion
barrier, the thermally activated oxygen ions may diffuse back during the reset process
to the oxide bulk layer since there is a large concentration gradient of oxygen across
the interface region. This may account for the unipolar reset mechanism. If oxidizable
materials are used for the electrode, it may form an interfacial oxide layer between
the electrode and oxide films, which may act as an oxygen diffusion barrier. Thus, it
is difficult for the oxygen ions to diffuse back through the thermal activation. Then
only by the acceleration of a reversed electric field can the oxygen ions drift back to
rupture the CFs. This may account for the bipolar reset mechanism. Numerical
simulation [102] based on the oxygen ions nonlinear transport model is employed
to support the above assumptions.
13 Metal Oxide Resistive Switching Memory 315

Fig. 13.7 (a) Auger electron spectroscopy (AES) depth profile of (a) Pt/CuxO/Cu structure and
(b) TaN/CuxO/Cu structure, indicating the existence of an interfacial TaON layer; (c) TEM picture
of the TaN/CuxO/Cu structure, showing the interfacial TaON layer; (d) I–V characteristics of
Pt/CuxO/Cu structure with unipolar switching behavior and TaN/CuxO/Cu structure with bipolar
switching behavior. Reprinted from [101]

13.2.4 Summary of the Physical Mechanism


for Resistive Switching in Metal Oxide Memory

Here, we would like to summarize the physical resistive switching mechanism in


metal oxide memory. It should be noted, however, while a self-consistent physical
picture can be constructed through the discussions above, the resistive switching
phenomenon in metal oxides involves a large variety of materials, so the model
described here may not apply to all combinations of materials.
Basically, resistive switching is conjectured to be due to the formation and
dissolution of conductive filaments localized at the anode interface, and is a
reversible switching process. The conductive filaments may consist of oxygen
316 S. Yu et al.

vacancies or excess metallic precipitates. Multiple, parallel filaments may be


formed. The filaments may preferentially be located at the grain boundaries.
Experimental evidences suggest that the switching process is an electrochemical
process associated with oxygen migration, oxidation and reduction. As shown in
Fig. 13.8, during the electroforming process, soft dielectric breakdown occurs and
oxygen ions drift to the anode interface by the high electric field, where they are
discharged as neutral nonlattice oxygen if the anode materials are noble metals or
react with the oxidizable anode materials to form an interfacial oxide layer.
Meanwhile in the bulk, the resultant oxygen vacancies or metal precipitates form
the highly conducting paths. During the reset process, oxygen ions migrate back to
the bulk either to recombine with the oxygen vacancies or to oxidize the metal
precipitates. In the unipolar switching case, the Joule heating thermally activates
the diffusion of oxygen ions. Oxygen ions diffuse against the electric field from the
anode due to the concentration gradient. In the bipolar switching case, usually
interfacial layer between the anode and oxide acts as a diffusion barrier, and oxygen
ions can only drift back aided by the electric field. Usually only a part of the
conducting paths is ruptured, leaving the bottom part of the conducting paths to be a
virtual cathode that extend partially into the film. Then during the subsequent set
process, a process similar to the electroforming occurs but only in the region near
anode. The physical picture presented here does not contradict with the previous
thermal model for unipolar switching or ionic model for bipolar switching. And it
can at best be viewed as a phenomenological description of experimental
observations. Details of the physics of switching remain an area of active research.

13.3 Performances of Metal Oxide Memory Devices

First, we would like to discuss the fabrication techniques briefly. Currently, the
fabrication of metal oxide memory devices uses mainly conventional semiconduc-
tor fabrication processes. Sputtering of metals followed by annealing in oxygen
ambient or reactive sputtering in oxygen ambient are the most common deposition
techniques for metal oxide layers. Other methods used are ALD [30, 43], pulsed
laser deposition (PLD) [103], metal organic chemical vapor deposition (MOCVD)
[26] and the sol–gel method [104]. Several papers report resistive switching
behavior of self-assembly grown nanostructure, e.g., NiO nanowires [105, 106].
The thickness of the metal oxide layer is usually around 10–50 nm. The substrate of
the devices is mainly silicon; however, there are several efforts to fabricate metal
oxide memory on transparent substrate, e.g., glass [107] or flexible substrate,
e.g., polyethersulfone (PES) [108, 109]. In the rest of this section, we will discuss
the characteristics of metal oxide memory devices including noise margin, scalabil-
ity, power consumption, speed, reliability, and uniformity.
The noise margin for the read operation is characterized by the HRS/LRS
resistance ratio. Generally, a ratio >10 is desired for easier sense amplifier design.
13
Metal Oxide Resistive Switching Memory

Fig. 13.8 Schematic of the switching process in metal oxide memory


317
318 S. Yu et al.

Fig. 13.9 HRS and LRS resistance vs. cell area of metal oxide memory devices. Data are
collected from [Lee 08] [30], [Baek 04] [73], [Yang 09] [79], [Xu 08] [83], [Sun 09] [103],
[Kim 09] [109], [Lv 08] [122]

Almost all the metal oxide memory devices in the literature exceed this
requirement.
Scalability is one of the key concerns for any kind of memory. Because of
the filamentary conduction nature, metal oxide memory devices may be scaled
down to the nanoscale. Sub-100 nm feature size cells have been fabricated by
193-nm lithography with a contact hole structure [78], as well as by nanoimprint
lithography or e-beam lithography with a crossbar arrays structure [51, 110, 111].
Recently, aggressively scaled 30 nm  30 nm HfO2 memory devices in 1 Kb array
have been demonstrated with excellent yield [31]. Resistive switching behavior has
been successfully triggered by C-AFM tip on a 10 nm  10 nm region of NiO thin
films [71], suggesting the potential to scale the cell size even down to the sub-10 nm
regime. Figure 13.9 plots the general scaling trend of HRS and LRS resistance. The
leakage current in HRS is mainly due to bulk leakage current. Thus, the resistance of
HRS increases as the inverse of the cell area, roughly following the Ohm’s law. The
conduction current in LRS is mainly filamentary conduction current, as discussed
before. So the resistance of LRS has only a slight dependency on the cell area. This
trend of increasing HRS/LRS resistance ratio as cell area decreases is a benefit of
device scaling.
Although the read noise margin is improved in the scaled memory cell, the
filamentary conduction nature of the memory cell results in increasing power
density as devices scale down. The typical switching voltage in literature is around
1–5 V, while the switching current can be many orders different, depending on the
material and the device structure. The key to reducing power consumption is to
reduce the reset switching current. The peak value of current in metal oxide
memory is the LRS current at the point when the reset process occurs, which is
usually referred to as the reset current. Recently, Wu et al. [112] have demonstrated
13 Metal Oxide Resistive Switching Memory 319

Fig. 13.10 The peak value of reset current and corresponding current density vs. cell area. Data
are collected from [Rohde 05] [21], [Lee 08] [30], [Kim 06] [45], [Kim 07] [72], [Baek 04] [73],
[Yoshida 07] [98], [Sun 09] [103], [Shin 08] [138], [Lee 09] [142]

extremely low reset current (~1 mA) for ALD grown Al2O3 memory cell. Yet, most
reports in the literature show a typical reset current for a single device on the order
of mA or hundreds of mA. Figure 13.10 plots the general scaling trend of reset
current and corresponding reset current density. It is seen that the reset current
reduce only slightly when the devices are scaled down, thus leading to a remarkable
increase of the current density required for reset. This presents a significant
challenge for the memory cell selection devices, which need to provide such a
large current density for ultra-scaled cells, e.g., ~MA/cm2 even for a relatively large
100 nm  100 nm cell area. Although there is contrary viewpoint that the reset
current would decrease with reduced cell area [113], it should be noted that it is not
a fair comparison because in those experiments smaller set compliance was applied
for the smaller memory cell. We will see in the following discussions that the
reset current value can be modulated by controlling of the set compliance value.
Figure 13.11 plots the relationship between reset current and set compliance
current. It is seen that reset current is usually only a little larger than the set
compliance current, and more importantly, the reset current decreases almost
linearly with the decrease of set compliance current. A possible reason for this
relationship is that with a smaller set compliance current, less oxygen atoms
migrate to the anode and weaker CFs are formed during the set process. Thus a
smaller reset current is needed to rupture the CFs. In the literature the reset current
of single memory cell is usually difficult to be reduced down to sub-mA regime,
even if sub-mA set compliance current is applied. The deviation from the linear
relationship between reset current and set compliance is most probably due to the
parasitic capacitance in the measurement setup as discussed in [114]. Thus the reset
current tends to be elevated at the mA level even if the set compliance of the
320 S. Yu et al.

Fig. 13.11 The reset current vs. set compliance current. Data are collected from [Seo 04] [20],
[Rohde 05] [21], [Kinoshita 08] [114], [Tsunoda 07] [121], [Shima 08] [143]

parameter analyzer used in the measurement is reduced down to the mA level.


By using an on-chip 1 transistor-1 resistor (1T-1R) structure rather than the single
resistor (1R) structure, the parasitic capacitance is greatly reduced, thus the reset
current tends to follow the linear relationship with set compliance current. Sub-
100 mA reset current has been successfully demonstrated in several 1T-1R cell
structures by adjusting the selection transistor’s gate voltage to deliver a small
set compliance current [30, 115–117]. As a consequence, however, the smaller set
compliance may result in a smaller HRS/LRS resistance ratio. Figure 13.12 plots the
dependence of the LRS resistance on the set compliance current. The LRS resistance
increases almost inversely with the decreasing set compliance current. Smaller set
compliance forms weaker CFs, thus leading to a larger LRS resistance. Given certain
HRS resistance, the requirement for a sufficient ratio limits the smallest set compli-
ance that can be used. Remember that HRS resistance rises up rapidly with the
decrease of cell area, as mentioned before. This means for the scaled memory cell, a
larger LRS resistance can be tolerated, and a smaller set compliance can be utilized to
achieve lower reset current. With this methodology, by using 5 mA set compliance, a
greatly reduced reset current (~50 mA) is demonstrated in a 100 nm  100 nm NiO
memory cell with sufficient HRS/LRS ratio (~104) [71, 113].
An attractive feature of metal oxide memory is the nanosecond switching speed,
which is comparable to DRAM and is substantially faster than Flash on a single bit
programming basis. It is at least three to ten times faster than the emerging PCM.
Figure 13.13 plots the HRS and LRS resistance vs. the programming pulse width.
It is seen that with smaller pulse width, HRS resistance drops while LRS resistance
rises, leading to a smaller HRS/LRS resistance ratio. The shorter pulses result in a
less complete switching process. Thus, the requirement to maintain a sufficient
noise margin limits the switching speed. The measurement of switching speed
13 Metal Oxide Resistive Switching Memory 321

Fig. 13.12 LRS resistance vs. set compliance current. Data are collected from [Lin 07] [25],
[Lee 07] [111], [Tsunoda 07] [121], [Lee 06] [144]

Fig. 13.13 HRS and LRS resistance vs. the programming pulse width. Data are collected from
[Chen 07] [119], [Choi 06] [120], [Tsunoda 07] [121], [Muraoka 07] [145], [Do 09] [146]

involves a delicate procedure [118]: a load resistor is connected in series with the
memory cell; the programming pulse (Vp) is applied to the stacked resistor and cell
structure through a pulse generator, while an oscilloscope is used to monitor the
voltage (Vcell) across the memory cell. Note that Vcell ¼ Vp  Rcell / (Rcell + RL),
thus the time when Vcell suddenly rises or drops can be estimated as the switching
time of reset or set process for the applied cell voltage Vcell. The magnitude of the
current can be calculated from the measured cell voltage at the oscilloscope divided
by load resistor, Icell ¼ Vcell/Rcell. With this methodology, the relationship between
322 S. Yu et al.

the switching time and the pulse amplitude can be investigated. There are reports
indicating that in order to achieve faster switching, a larger pulse amplitude is
required [119]. Therefore, a trade-off exists between fast switching speed and low
power consumption. Bipolar switching devices have shown ultra-fast switching
speed using reasonable voltage pulse amplitudes, e.g., +2 V/10 ns for set and
1.5 V/10 ns for reset in TaOx memory [64], +3.2 V/5 ns for set and 2.7 V/5 ns
for reset in HfO2 memory [30]. Some earlier works showed that for unipolar
switching devices, the reset process needs a much longer pulse width than set
process, e.g., +1 V/10 ns for set and +0.5 V/5 ms for reset in NiO memory [73],
+3 V/10 ns for set and +2.5 V/5 ms for reset in TiO2 memory [120]. It is suggested
that a certain amount of time (~ms) is necessary to sufficiently heat up the CFs to
trigger the rupture [120], since the unipolar reset mechanism involves a thermally
assisted oxygen diffusion process. But recently, several works demonstrated that
ultra-fast reset switching speed that is comparable to the set switching speed is also
achievable in unipolar switching devices, e.g., +2.8 V/10 ns for set and +1.8 V/5 ns
for reset in Ti-doped NiO memory [121], +1.5 V/10 ns for set and +1 V/10 ns for
reset in scaled NiO memory [113]. It is shown that the reset switching time
decreases significantly with decreasing cell area, from 1 ms (30 mm  30 mm
cell) to 10 ns (100 nm  100 nm cell) for unipolar NiO memory [71]. If individual
CFs are similar in size or resistivity and only the number of CFs varies between
different cell areas, then the switching time should be insensitive to the cell
area, because CFs are connected in parallel and ruptured independently. Therefore,
it is inferred the CFs should be weaker or have larger resistivity for the smaller
cells [71]. This trend that smaller cells have faster switching speed can also be due
to the smaller parasitic capacitance of the smaller cell.
The reliability characteristics include the endurance and retention. Current Flash
technology shows a maximum number of programming cycles between 103 and
107, depending on the device type. Metal oxide memory should provide at least the
same endurance, preferably a better one. Excellent endurance records in the litera-
ture are shown for unipolar NiO memory (>106) [73], bipolar HfO2 memory
(>106) [30], and bipolar TaOx memory (>109) [64] at the single cell level. The
failure of switching may be due to grain structure change near the anode region [56]
and insufficient nonlattice oxygen ions to rupture the CFs [83], since during the
repeated switching cycles, oxygen may escape from the device. Most nonvolatile
memory applications specify at least 10 years data retention time at typical
operating temperatures (e.g., 85 C). Promising retention records in the literature
are shown for unipolar Ti-doped NiO memory (>1,000 h at 150 C) [121], and
bipolar TaOx memory (>3,000 h at 150 C) [64]. A practical method to estimate the
retention capability is to draw the Arrhenius plot of retention time vs. the reciprocal
of temperature, and extrapolate the data obtained in high temperature regime
to the operating temperature (e.g., 85 C) [64, 122]. The loss of data in LRS or
HRS may be associated with the diffusion of oxygen ions or oxygen vacancies due
to their concentration gradient, and this process can be accelerated at elevated
temperature because of the temperature-dependent diffusion coefficient. This
is the basis of any temperature varying experiments to estimate the retention
13 Metal Oxide Resistive Switching Memory 323

time. A more complicated statistical method to estimate the retention capability by


temperature-accelerated measurement is developed by Chen et al. [123].
Uniformity of device characteristics is a key problem that hinders the metal
oxide memory from entering large-scale manufacturing. Significant parameter
fluctuations exist in terms of variations of the switching voltages as well as the
resistance in HRS and LRS. These parameters fluctuate from programming cycle to
programming cycle and from device to device. Many efforts have been expended to
improve the uniformity. Kim et al. [124] proposed to use IrO2 as the top electrode for
the NiO memory to help stabilize the local oxygen migrations for the formation and
rupture of CFs. Chang et al. [125] proposed to embed Pt nanocrystals into the TiO2
memory to provide an easy path for the formation and rupture of CFs by local
enhancement of the electric field adjacent to the Pt nanocrystals. Yu et al. [126]
proposed to stack a thin Al buffer layer on HfO2 memory to diffuse Al impurity
atoms into the switching layer, aiming to stabilize the formation and rupture of CFs
along the impurities due to a lower Vo formation energy caused by the impurities.
Liu et al. [127] proposed to implant Ti into ZrO2 memory to provide a source of Vo
and play the role of a seed for forming CFs. The key idea of these techniques is to
induce the formation and rupture of CFs in specific locations during the switching
instead of allowing them to occur randomly. Besides materials solutions mentioned
above, novel programming methods are also helpful in reducing parameter
fluctuations. Using multiple pulses rather than a single pulse [128] or using a ramped
series of pulses [129] can remarkably improve the cycle to cycle uniformity.
To summarize, the scalability of metal oxide memory devices is the most
competitive characteristics compared with other emerging memories. It is foresee-
able to achieve sufficient noise margin (>10), low energy consumption (<pJ per
switching cycle), fast switching speed (~ns) for nanoscale (10 nm  10 nm) cells in
the near future. However, the reliability and uniformity of metal oxide memory
needs to be further improved before this technology can be commercialized.

13.4 Cell Structure of Metal Oxide Memory Arrays

For random access application, each metal oxide memory cell is usually connected
with a cell selection transistor). This cell structure is referred to as 1 transistor-1
resistor (1T-1R). By controlling the selection transistor’s gate voltage, varying
current can be provided to control the switching characteristics of memory cell.
As mentioned in Sect. 13.3, a smaller set compliance results in a larger LRS
resistance, which is preferred to reduce reset current. At the same time, the set
compliance should be larger enough to guarantee a sufficient HRS/LRS resistance
ratio. By adjusting the selection transistor’s gate voltage, it is easy to optimize
the programming current for the set and reset processes [115, 116, 119]. However,
the area penalty of using a selection transistor is apparent, even with a delicate
layout design of integrating the memory cell into the via contact of the transistor.
The 1T-1R cell area can range from 6F2 (F: feature size used for patterning the cell)
324 S. Yu et al.

Fig. 13.14 The crossbar matrix (a) without cell selection elements and (b) with cell selection
elements

using aggressive DRAM-like design rules with borderless contacts and zero gate to
source/drain spacing to 8F2 (with F/2 gate to source/drain spacing) and 18F2
(contacts with borders).
For high-density integration, a crossbar matrix with 4F2 area is preferred. The
simplest way is to connect the word and bit lines at each node by a memory cell.
The crossbar matrix can be a set of orthogonal nanoscale metal wires with the metal
oxide as the memory element at the crossbar junction. TiO2 memory with junction
area (50 nm  50 nm) [51], NiO memory with junction area (70 nm  70 nm)
[110] have been reported. Lee et al. [111] demonstrated a novel structure for NiO
memory which utilizes the sidewalls between the two nanoscale metal wires as the
active switching region. Thus, the active memory cell area is even smaller than the
layout cell area (48 nm  48 nm).
A cross-talk problem with the above simple crossbar matrix may arise in read
operation [130]. As shown in Fig. 13.14a, if the cell to be read out happens to be in
HRS with surrounding cells in LRS, the reading current can easily flow through the
surrounding cells in the LRS and thus a LRS data will be mistakenly read out. In order
to avoid these parasitic leakage paths, a cell selection element with large I–V
nonlinearity should be added at each node. The cell selection element is usually
envisioned to be a diode (with an exponential I–V characteristics). Therefore, this cell
structure refers to 1 diode-1 resistor (1D-1R). The reverse biased diodes effectively
cuts off the leakage current paths, thus the interference between neighboring cells is
prevented, as shown in Fig. 13.14b. The diode on/off ratio required ranges from 104
to 106 depending on the memory array size, the on/off ratio of the memory cell, and
the resistance of the interconnect wire, etc. It should be noted for the 1D-1R structure,
the resistive switching element must be unipolar switching devices because the
diode limits the reversed current that a bipolar switching device requires.
13 Metal Oxide Resistive Switching Memory 325

With such crossbar structure, the cell area can in principle be scaled down to 4F2.
Furthermore, by stacking the crossbar structure in the third dimension, the effective
cell size can be further reduced to 2F2, 1F2, and so on [131]. The 1D-1R structure
can be stacked for 3-D integration. Thus there are successive attempts to fabricate
stacked nanowire arrays for NiO memory with different cell selection elements
[132–135].
Here we would like to discuss the cell selection element material. A p–n diode is
the most common device for the cell selection element. Although high performance
p–n diode is easily fabricated with current epitaxial silicon technology for the
planar device structures, it is not feasible to implement epitaxial silicon-based
p–n diode into the stacked crossbar metal oxide memory structures because it is
difficult to grow epitaxial silicon on a metal layer and high processing temperatures
are required. On the other hand, amorphous silicon allows for lower processing
temperatures. But it does not meet the requirement for current density for memory
cell programming. Therefore, new materials need to be explored for the cell
selection element, which should both allow for low processing temperatures and
also provide high current drive. Compared with silicon diode, oxide diode is
attractive because it offers better flexibility in processing technology because it
can be fabricated over any substrate even at room temperature [131]. If the oxide
material is oxygen deficient with sufficient amount of oxygen vacancies, such as
TiO2, ZrO2, ZnO, and indium tin oxide (ITO), it is n-type; while if the oxide
material is metal deficient with sufficient amount of metal vacancies, such as
NiO, it is p-type. Thus a combination of p-type oxide and n-type oxide essentially
forms a p-n diode. The research on oxide diode started earlier than that on the metal
oxide memory even in the 1990s, e.g., p-NiO/n-Al: ZnO [136], p-NiO/n-ITO [137].
Recently, several kinds of oxide diodes [133], such as p-NiO/n-TiO2, p-NiO/n-
ZnO, p-NiO/n-InZnO, p-CuO/n-InZnO, have been stacked with Pt/NiO/Pt structure
in series, among which p-CuO/n-InZnO is regarded as the best candidate. Besides
the p-n oxide diode, single oxide materials can also be applied to rectify the current.
Shin et al. [138] demonstrated a Schottky-type cell structure consisting of Pt/ITO/
TiO2/Pt stacked layers. The low and high potential barrier at the ITO/TiO2 and
TiO2/Pt junctions, respectively, constitute the rectifying properties of the stacked
structure, where TiO2 acts as the resistive switching layer. VO2 is found to exhibit
electric field-induced metal–insulator transition [139]. Unlike other metal oxide
memory devices, the resistive switching behavior in VO2 is not bistable, so it is
sometimes referred to threshold switching. The sharp transition of current in VO2
when turned on provides an ideal behavior for switching element. Lee et al. [140]
demonstrated a prototype cell structure consisting of Pt/VO2/Pt/NiO/Pt stacked
layers. Table 13.2 compares several switching elements mentioned here for the
metal oxide memory cell in aspects of forward current density, forward/reverse
current ratio, and turn on voltage. Generally, high forward current density, high
forward/reverse current ratio, and low turn on voltage are preferred. Programming
current density of the order of 10 MA/cm2 is required for metal oxide memory
today. Thus, the oxide-based diodes reported to date are several orders of magni-
tude too low in current drive for the crossbar memory application.
326 S. Yu et al.

Table 13.2 Comparison of several oxide switching elements for the metal oxide memory cell.
Forward current density is extracted at +2 V bias, forward/reverse current ratio is defined at the
current ratio at +2 V and 2 V, and turn on voltage is defined as J ¼ 100 A/cm2
p-NiO/n- p-NiO/n- p-NiO/n- p-CuO/n- ITO VO2
TiO2 [133] ZnO [133] IZO [133] IZO [133] [138] [140]
Forward current ~20 ~300 ~10,000 ~30,000 ~1,400 ~4,000a
density (A/cm2)
Forward/reverse 4.7  105 2.6  105 3  104 3  104 1  104 N/A
current ratio
Turn on voltage (V) 2.72 1.55 0.73 0.66 0.62 0.6
Data are collected from [133, 138, 140]
a
Estimated

Fig. 13.15 (a) Scanning electron microscopy (SEM) image of an 8  8 1D-1R arrays integrated
with GaInZnO selection TFTs. Top image is the cross-section view of a row in the 8  8 1D-1R
arrays. (b) Image of the integrated devices shown over a glass substrate. (c) I–V characteristics of a
NiO cell, a 1D-1R cell, and a 1D-1R cell with TFT selection transistor. (d) Schematic diagram of a
1D-1R cell with selection transistor. Reprinted from [141]
13 Metal Oxide Resistive Switching Memory 327

Recently, Lee et al. [141] demonstrated a 8  8 1D-1R array with word/bit line
selection transistors, as shown in Fig. 13.15. The memory cell is aPt/NiO/Pt/p-CuO/
n-InZnO/Pt stacked structure, and the word line selection transistor is amorphous
gallium indium zinc oxide (GaInZnO) thin film transistor (TFT). Oxide TFT rather
than silicon transistor is utilized for the low processing temperature required for
3-D integration.

13.5 Summary

A review of recent research progress of metal oxide memory is presented here.


Possible physical mechanism of resistive switching in metal oxides is discussed.
Filamentary conductive paths dominate the conduction in LRS. Electrochemical/
electro-thermal oxygen migration, oxidation, and reduction may be the microscopic
origin of resistive switching behavior. Electrode materials play an important role in
determining the switching modes: bipolar or unipolar. The device characteristics of
metal oxide memory cells reported in the literature are summarized. Sufficient
noise margin, excellent scalability, low power consumption, ultra-fast speed,
good endurance and retention make metal oxide memory a competitive candidate
for future nonvolatile memory. Device uniformity must be further improved in
order to meet the large-scale manufacturing requirements. Cell structure for high-
density memory array is discussed. The 1D-1R crossbar array structure is promises
ultra-high integration density with the potential for 3-D integration. Low process
temperature memory cell selection element with the required current drive and on/
off ratio is the key to realizing 3-D stackable crossbar memory arrays.

References

1. Burr, G.W., Kurdi, B.N., Scott, J.C., Lam, C.H., Gopalakrishnan, K., Shenoy, R.S.: Overview
of candidate device technologies for storage-class memory. IBM J. Res. Dev. 52, 449 (2008)
2. Kryder, M.H., Chang, S.K.: After hard drives – what comes next? IEEE Trans. Magn. 45,
3406 (2009)
3. Zhu, J.-G.: Magnetoresistie random access memory: the path to competitiveness and scal-
ability. Proc. IEEE 96, 1786 (2008)
4. Arimoto, Y., Ishiwara, H.: Current status of ferroelectric random-access memory. MRS Bull.
29, 823 (2004)
5. Meijer, G.I.: Who wins the nonvolatile memory race? Science 319, 1625 (2008)
6. Wuttig, M., Yamada, N.: Phase-change materials for rewriteable data storage. Nat. Mater. 6,
824 (2007)
7. Pirovano, A., Lacaita, A.L., Benvenuti, A., Pellizzer, F., Bez, R.: Electronic switching in
phase-change memories. IEEE Trans. Electron Devices 51, 452 (2004)
8. Lee, S.H., Jung, Y., Agarwal, R.: Highly scalable non-volatile and ultra-low-power phase-
change nanowire memory. Nat. Nanotechnol. 2, 626 (2007)
9. Kim, S., Zhang, Y., McVittie, J.P., Jagannathan, H., Nishi, Y., Wong, H.-S.P.: Integrating
phase-change memory cell with Ge Nanowire diode for crosspoint memory – experimental
demonstration and analysis. IEEE Trans. Electron Devices 55, 2307 (2008)
328 S. Yu et al.

10. Kozicki, M.N., Park, M., Mitkova, M.: Nanoscale memory elements based on solid-state
electrolytes. IEEE Trans. Nanotechnol. 4, 331 (2005)
11. Wang, Z., Griffin, B., McVittie, J., Wong, S., McIntyre, C., Nishi, Y.: Resistive switching
mechanism in ZnxCd1-xS nonvolatile memory devices. IEEE Electron Device Lett. 28, 14 (2007)
12. Scott, J.C., Bozano, L.D.: Nonvolatile memory elements based on organic materials. Adv.
Mater. 19, 1452 (2007)
13. Hong, S.H., Kim, O., Choi, S., Ree, M.: Bipolar resistive switching in a single layer memory
device based on a conjugated copolymer. Appl. Phys. Lett. 91, 093517 (2007)
14. Hickmott, T.W.: Low-frequency negative resistance in thin anodic oxide films. J. Appl. Phys.
33, 2669 (1962)
15. Gibbons, J.F., Beadle, W.E.: Switching properties of thin NiO Films. Solid State Electron.
7, 785 (1964)
16. Dearnaley, G., Stoneham, A.M., Morgan, D.V.: Electrical phenomena in amorphous oxide
films. Rep. Prog. Phys. 33, 1129 (1970)
17. Asamitsu, A., Tomioka, Y., Kuwahara, H., Tokura, Y.: Current switching of resistive states in
magnetoresistive manganites. Nature 388, 50 (1997)
18. Beck, A., Bednorz, J.G., Gerber, C., Rossel, C., Widmer, D.: Reproducible switching effect in
thin oxide films for memory applications. Appl. Phys. Lett. 77, 139 (2000)
19. Zhuang, W.W., Pan, W., Ulrich, B.D., Lee, J.J., Stecker, L., Burmaster, A., Evans, D.R., Hsu,
S.T., Tajiri, M., Shimaoka, A., Inoue, K., Naka, T., Awaya, N., Sakiyama, K., Wang, Y., Liu,
S.Q., Wu, N.J., Ignatiev, A.: Novel colossal magnetoresistive thin film nonvolatile resistance
random access memory (RRAM). In: Electron devices meeting, 2002. IEDM ’02. Digest.
International, San Francisco, CA, p. 193 (2002)
20. Seo, S., Lee, M.J., Seo, D.H., Jeoung, E.J., Suh, D.-S., Joung, Y.S., Yoo, I.K., Hwang, I.R.,
Kim, S.H., Byun, I.S., Kim, J.-S., Choi, J.S., Park, B.H.: Reproducible resistance switching in
polycrystalline NiO films. Appl. Phys. Lett. 85, 5655 (2004)
21. Rohde, C., Choi, B.J., Jeong, D.S., Choi, S., Zhao, J.-S., Hwang, C.S.: Identification of a
determining parameter for resistive switching of TiO2 thin films. Appl. Phys. Lett. 86, 262907
(2005)
22. Lee, D., Choi, H., Sim, H., Choi, D., Hwang, H., Lee, M.-J., Seo, S.-A., Yoo, I.K.: Resistance
switching of the nonstoichiometric zirconium oxide for nonvolatile memory applications.
IEEE Electron Device Lett. 26, 719 (2005)
23. Villafuerte, M., Heluani, S.P., Juárez, G., Simonelli, G., Braunstein, G., Duhalde, S.: Electric-
pulse-induced reversible resistance in doped zinc oxide thin films. Appl. Phys. Lett. 90,
052105 (2007)
24. Dong, R., Lee, D.S., Xiang, W.F., Oh, S.J., Seong, D.J., Heo, S.H., Choi, H.J., Kwon, M.J.,
Seo, S.N., Pyun, M.B., Hasan, M., Hwang, H.: Reproducible hysteresis and resistive
switching in metal-CuxO-metal heterostructures. Appl. Phys. Lett. 90, 042107 (2007)
25. Lin, C.-Y., Wu, C.-Y., Wu, C.-Y., Hu, C., Tseng, T.-Y.: Bistable resistive switching in
AL2O3 memory thin films. J. Electrochem. Soc. 154, G189 (2007)
26. Lee, S., Kim, W.-G., Rhee, S.-W., Yong, K.: Resistance switching behaviors of hafnium
oxide films grown by MOCVD for nonvolatile memory applications. J. Electrochem. Soc.
155, H92 (2008)
27. Waser, R., Aono, M.: Nanoionics-based resistive switching memories. Nat. Mater. 6, 833
(2007)
28. Waser, R., Dittmann, R., Staikov, G., Szot, K.: Redox-based resistive switching memories –
nanoionic mechanisms, prospects, and challenges. Adv. Mater. 21, 2632 (2009)
29. Sawa, A.: Resistive switching in transition metal oxides. Mater. Today 11, 28 (2008)
30. Lee, H.Y., Chen, S., Wu, T.Y., Chen, Y.S., Wang, C.C., Tzeng, J., Lin, C.H., Chen, F., Lien,
C.H., Tsai, M.-J.: Low power and high speed bipolar switching with a thin reactive Ti buffer
layer in robust HfO2 based RRAM. In: Electron devices meeting, 2008. IEDM 2008. IEEE
International, San Francisco, CA, p. 297 (2008)
13 Metal Oxide Resistive Switching Memory 329

31. Chen, Y.S., Lee, H.Y., Chen, S., Gu, Y., Chen, C.W., Lin, W.P., Liu, W.H., Hsu, Y.Y., Sheu,
S.S., Chiang, C., Chen, W.S., Chen, F.T., Lien, C.H., Tsai, M.-J.: Highly scalable hafnium
oxide memory with improvements of resistive distribution and read disturb immunity. In:
Electron devices meeting (IEDM), 2009 IEEE International, Baltimore, MD, p. 105 (2009)
32. International Technology Roadmap for Semiconductors, 2007 edition. http://www.itrs.net/
(2007)
33. Sato, Y., Kinoshita, K., Aoki, M., Sugiyama, Y.: Consideration of switching mechanism of
binary metal oxide resistive unctions using a thermal reaction model. Appl. Phys. Lett. 90,
033503 (2007)
34. Russo, U., Ielmini, D., Cagli, C., Lacaita, A.L., Spiga, S., Wiemer, C., Perego, M., Fanciulli,
M.: Conductive-filament switching analysis and self-accelerated thermal dissolution model
for reset in NiO-based RRAM. In: Electron devices meeting, 2007. IEDM 2007. IEEE
International, Washington, DC, p. 775 (2007)
35. Fujii, T., Kawasaki, M., Sawa, A., Akoh, H., Kawazoe, Y., Tokura, Y.: Hysteretic current-
voltage characteristics and resistance switching at an epitaxial oxide Schottky junction
SrRuO3/SrTi0.99Nb0.01O3. Appl. Phys. Lett. 86, 012107 (2005)
36. Xia, Y., He, W., Chen, L., Meng, X., Liu, Z.: Filed-induced resistive switching based on
space-charge-limited current. Appl. Phys. Lett. 90, 022907 (2007)
37. Rozenberg, M.J., Inoue, I.H., Sanchez, M.J.: Strong electron correlation effects in nonvolatile
electronic memory devices. Appl. Phys. Lett. 88, 033510 (2006)
38. Sanchez, M.J., Inoue, I.H., Rozenberg, M.J.: A mechanism for unipolar resistance switching
in oxide nonvolatile memory devices. Appl. Phys. Lett. 91, 252101 (2007)
39. Nian, Y.B., Strozier, J., Wu, N.J., Chen, X., Ignatiev, A.: Evidence for an oxygen diffusion
model for the electric pulse induced resistance change effect in transition-metal oxides. Phys.
Rev. Lett. 98, 146403 (2007)
40. Jeong, D.S., Schroeder, H., Waser, R.: Mechanism for bipolar switching in a Pt/TiO2/Pt
resistive switching cell. Phys. Rev. B 79, 195317 (2009)
41. Tsunoda, K., Fukuzumi, Y., Jameson, J.R., Wang, Z., Griffin, B., Nishi, Y.: Bipolar resistive
switching in polycrystalline TiO2. Appl. Phys. Lett. 90, 113501 (2007)
42. Sakamoto, T., Banno, N., Iguchi, N., Kawaura, H., Sunamura, H., Fujieda, S., Terabe, K.,
Hasegawa, T., Aono, M.: A Ta2O5 solid-electrolyte switch with improved reliability. In:
2007 IEEE symposium on VLSI technology, Kyoto, p. 38 (2007)
43. Choi, B.J., Jeong, D.S., Kim, S.K., Rohde, C., Choi, S., Oh, J.H., Kim, H.J., Hwang, C.S.,
Szot, K., Waser, R., Reichenberg, B., Tiedke, S.: Resistive switching mechanism of TiO2 thin
films grown by atomic-layer deposition. J. Appl. Phys. 98, 33715 (2005)
44. Szot, K., Speier, W., Bihlmayer, G., Waser, R.: Switching the electrical resistance of
individual dislocations in single-cyrstalline SrTiO3. Nat. Mater. 5, 312 (2006)
45. Kim, K.M., Choi, B.J., Koo, B.W., Choi, S., Jeong, D.S., Hwang, C.S.: Resistive switching in
Pt/Al2O3/TiO2/Ru stacked structures. Electrochem. Solid State Lett. 9, G343 (2006)
46. Lee, D., Seong, D.-J., Jo, I., Xiang, F., Dong, R., Oh, S., Hwang, H.: Resistance switching of
copper doped MoOx films for nonvolatile memory applications. Appl. Phys. Lett. 90, 122104
(2007)
47. Son, J.Y., Shin, Y.-H.: Direct observation of conduction filaments on resistive switching of
NiO thin films. Appl. Phys. Lett. 92, 222106 (2008)
48. Muenstermann, R., Dittmann, R., Szot, K., Mi, S., Jia, C.-L., Meuffels, P., Waser, R.:
Realization of regular arrays of nanoscale resistive switching blocks in thin films of
Nb-doped SrTiO3. Appl. Phys. Lett. 93, 023110 (2008)
49. Liu, Q., Dou, C., Wang, Y., Long, S., Wang, W., Liu, M., Zhang, M., Chen, J.: Formation of
multiple conductive filaments in the Cu/ZrO2:Cu/Pt device. Appl. Phys. Lett. 95, 023501
(2009)
50. Liu, M., Abid, Z., Wang, W., He, X., Liu, Q., Guan, W.: Multilevel resistive switching with
ionic and metallic filaments. Appl. Phys. Lett. 94, 233106 (2009)
330 S. Yu et al.

51. Yang, J.J., Pickett, M.D., Li, X., Ohlberg, D.A.A., Stewart, D.R., Williams, R.S.: Memristive
switching mechanism for metal/oxide/metal nanodevices. Nat. Nanotechnol. 3, 429 (2008)
52. Janousch, M., Meijer, G.I., Staub, U., Delley, B., Karg, S.F., Andreasson, B.P.: Role of
oxygen vacancies in Cr-doped SrTiO3 for resistance-change memory. Adv. Mater. 19, 2232
(2007)
53. Park, S.-G., Kope, B.M., Nishi, Y.: First-principles study of resistance switching in rutile
TiO2 with Oxygen vacancy. In: NVMTS 2008. 9th Annual, Monterey, CA, p. 1 (2008)
54. Jung, K., Seo, H., Kim, Y., Im, H., Hong, J., Park, J.-W., Lee, J.-K.: Temperature dependence
of high- and low-resistance bistable states in polycrystalline NiO films. Appl. Phys. Lett. 90,
052104 (2007)
55. Lee, M.J., Park, Y., Ahn, S.E., Kang, B.S., Lee, C.B., Kim, K.H., Xianyu, W.X., Lee, J.H.,
Chung, S.J., Kim, Y.H., Lee, C.S., Choi, K.N., Chung, K.S.: Comparative structural and
electrical analysis of NiO and Ti doped NiO as materials for resistance random access
memory. J. Appl. Phys. 103, 013706 (2008)
56. Park, G.-S., Li, X.-S., Kim, D.-C., Jung, R.-J., Lee, M.-J., Se, S.: Observation of electric-field
induced Ni filament channels in polycrystalline NiOx film. Appl. Phys. Lett. 91, 222103
(2007)
57. Mott, N.F., Davis, E.A.: Electronic processes in non-crystalline materials. Clarendon, Oxford
(1979)
58. Ho, C.H., Lai, E.K., Lee, M.D., Pan, C.L., Yao, Y.D., Hsieh, K.Y., Liu, R., Lu, C.Y.: A highly
reliable self-aligned graded oxide WOx resistance memory: conduction mechanisms and
reliability. In: 2007 IEEE symposium VLSI technology, Kyoto, p. 228 (2007)
59. Liu, Q., Guan, W., Long, S., Liu, M., Zhang, S., Wang, Q., Chen, J.: Resistance switching of
Au-implanted-ZrO2 film for nonvolatile memory application. J. Appl. Phys. 104, 114514
(2008)
60. Xu, N., Liu, L.F., Sun, X., Liu, X.Y., Han, D.D., Wang, Y., Han, R.Q., Kang, J.F., Yu, B.:
Characteristics and mechanism of conduction/set process in TiN/ZnO/Pt resistance switching
random-access memories. Appl. Phys. Lett. 92, 232112 (2008)
61. Chang, W.Y., Lai, Y.-C., Wu, T.-B., Wang, S.-F., Chen, F., Tsai, M.-J.: Unipolar resistive
switching characteristics of ZnO thin films for nonvolatile memory applications. Appl. Phys.
Lett. 92, 022110 (2008)
62. Kim, Y.-M., Lee, J.-S.: Reproducible resistance switching characteristics of hafnium oxide-
based nonvolatile memory devices. J. Appl. Phys. 104, 114115 (2008)
63. Lin, C.-Y., Wang, S.-Y., Lee, D.-Y., Tseng, T.Y.: Electrical properties and fatigue behaviors
of ZrO2 resistive switching thin films. J. Electrochem. Soc. 155, H615 (2008)
64. Wei, Z., Kanzawa, Y., Arita, K., Katoh, Y., Kawai, K., Muraoka, S., Mitani, S., Fujii, S.,
Katayama, K., Iijima, M., Mikawa, T., Ninomiya, T., Miyanaga, R., Kawashima, Y., Tsuji,
K., Himeno, A., Okada, T., Azuma, R., Shimakawa, K., Sugaya, H., Takagi, T., Yasuhara, R.,
Horiba, K., Kumigashira, H., Oshima, M.: Highly reliable TaOx ReRAM and direct evidence
of redox reaction mechanism. In: IEEE international electron devices meeting, 2008. IEDM
2008. IEEE International, San Francisco, CA, p. 293 (2008)
65. Lee, H.Y., Chen, P.-S., Wu, T.-Y., Chen, Y.S., Chen, F., Wang, C.-C., Tzeng, P.-J., Lin, C.H.,
Tsai, M.-J., Lien, C.: HfOx Bipolar resistive memory with robust endurance using AlCu as
buffer electrode. IEEE Electron Device Lett. 30, 703 (2009)
66. Liu, Q., Guan, W., Long, S., Jia, R., Liu, M., Chen, J.: Resistive switching memory effect of
ZrO2 films with Zr+ implanted. Appl. Phys. Lett. 92, 012117 (2008)
67. Li, X., Tung, C.H., Pey, K.L.: The nature of dielectric breakdown. Appl. Phys. Lett. 93,
072903 (2008)
68. Fujiwara, K.K., Nemoto, T., Rozenberg, M.J., Nakamura, Y., Takagi, H.: Resistance
switching and formation of a conductive bridge in metal/binary oxide/metal structure for
memory devices. Jpn. J. Appl. Phys. 47, 6266 (2008)
69. Buh, G.-H., Hwang, I., Park, B.H.: Time-dependent electroforming in NiO resistive
switching devices. Appl. Phys. Lett. 95, 142101 (2009)
13 Metal Oxide Resistive Switching Memory 331

70. Yang, J.J., Miao, F., Pickett, M.D., Ohlberg, D.A.A., Stewart, D.R., Lau, C.N., Williams, R.
S.: The mechanism of electroforming of metal oxide memristive switches. Nanotechnology
20, 215201 (2009)
71. Lee, M.-J., Han, S., Jeon, S.H., Park, B.H., Kang, B.S., Ahn, S.-E., Kim, K.H., Lee, C.B.,
Kim, C.J., Yoo, I.-K., Seo, D.H., Li, X.-S., Park, J.-B., Lee, J.-H., Park, Y.: Electrical
manipulation of nanofilaments in transition-metal oxides for resistance-based memory.
Nano Lett. 9, 1476 (2009)
72. Kim, K.M., Choi, B.J., Shin, Y.C., Choi, S., Hwang, C.S.: Anode-interface localized fila-
mentary mechanism in resistive switching of TiO2 thin films. Appl. Phys. Lett. 91, 012907
(2007)
73. Baek, I.G., Lee, M.S., Seo, S., Lee, M.J., Seo, D.H., Suh, D.-S., Park, J.C., Park, S.O., Kim, H.
S., Yoo, I.K., Chung, U.-I., Moon, I.T.: Highly scalable nonvolatile resistive memory using
simple binary oxide driven by asymmetric unipolar voltage pulses. In: Electron devices
meeting, 2004. IEDM Technical Digest. IEEE International, Washington, DC, p. 587 (2004)
74. Inoue, I.H., Yasuda, S., Akinaga, H., Takagi, H.: Nonpolar resistance switching of metal/
binary-transition-metal oxides/metal sandwiches: homogeneous/inhomogeneous transition of
current distribution. Phys. Rev. B 77, 035105 (2008)
75. Walcyzk, C., Wenger, C., Sohal, R., Lukosius, M., Fox, A., Da˛browski, J., Wolansky, D.,
Tillack, B., Mussig, H.-J., Schroeder, T.: Pulse-induced low-power resistive switching in
HfO2 metal0insulator-metal diodes for nonvolatile memory applications. J. Appl. Phys. 105,
114103 (2009)
76. Hsiung, C.-P., Gan, J.-Y., Tseng, S.-H., Tai, N.-H., Tzeng, P.-J., Lin, C.-H., Chen, F., Tsai,
M.-J.: Resistance switching characteristics of TiO2 thin films prepared with reactive
sputtering. Electrochem. Solid State Lett. 12, G31 (2009)
77. Cao, X., Li, X., Gao, X., Yu, W., Liu, X., Zhang, Y., Chen, L., Cheng, X.: Forming-free
colossal resistive switching effect in rare-earth-oxide Gd2O3 films for memristor applications.
J. Appl. Phys. 106, 073723 (2009)
78. Goux, L., Lisoni, J.G., Wang, X.P., Jurczak, M., Wouters, D.J.: Optimized Ni oxidation in 80-
nm contact holes for integration of forming-free and low-power Ni/NiO/Ni memory cells.
IEEE Trans. Electron Devices 56, 2363 (2009)
79. Yang, M.K., Park, J.-W., Ko, T.K., Lee, J.-K.: Bipolar resistive switching behavior in
TiMnO2/Pt structure for nonvolatile memory devices. Appl. Phys. Lett. 95, 042105 (2009)
80. Yoshida, C., Kinoshita, K., Yamasaki, T., Sugiyama, Y.: Direct observation of oxygen
movement during resistance switching in NiO/Pt film. Appl. Phys. Lett. 93, 042106 (2008)
81. Jeong, H.Y., Lee, J.Y., Choi, S.-Y., Kim, J.W.: Microscopic origin of bipolar resistive
switching of nanoscale titanium oxide thin films. Appl. Phys. Lett. 95, 162108 (2009)
82. Fujimoto, M., Koyama, H., Konagai, M., Hosoi, Y., Ishihara, K., Ohnishi, S., Awaya, N.:
TiO2 anatase nanolayer on TiN thin film exhibiting high-speed bipolar resistive switching.
Appl. Phys. Lett. 89, 223509 (2006)
83. Xu, N., Gao, B., Liu, L.F., Sun, B., Liu, X.Y., Han, R.Q., Kang, J.F., Yu, B.: A unified
physical model of switching behavior in oxide-based RRAM. In: 2008 Symposium on VLSI
technology, Honolulu, HI, p. 100 (2008)
84. Bruchhaus, R., Waser, R.: Bipolar resistive switching in oxides for memory. In: Ramanathan,
S. (ed.) Thin film metal-oxides: fundamentals and applications in electronics and energy, pp.
131–168. Springer, Heidelberg (2009)
85. Strukov, D.B., Snider, G.S., Stewart, D.R., Williams, R.S.: The missing memristor found.
Nature 453, 80 (2008)
86. Strukov, D.B., Borghetti, J.L., Williams, R.S.: Coupled ionic and electronic transport model
of thin-film semiconductor memristive behavior. Small 5, 1058 (2009)
87. Pickett, M.D., Strukov, D.B., Borghetti, J.L., Yang, J.J., Snider, G.S., Stewart, D.R.,
Williams, R.S.: Switching dynamics in titanium dioxide memristive devices. J. Appl. Phys.
106, 074508 (2009)
332 S. Yu et al.

88. Russo, U., Ielmini, D., Cagli, C., Lacaita, A.L.: Filament conduction and reset mechanism in
NiO-based resistive-switching memory (RRAM) devices. IEEE Trans. Electron Devices 56,
186 (2009)
89. Russo, U., Ielmini, D., Cagli, C., Lacaita, A.L.: Self-accelerated thermal dissolution model
for reset programming in unipolar resistive-switching memory (RRAM) devices. IEEE
Trans. Electron Devices 56, 193 (2009)
90. Cagli, C., Nardi, F., Ielmini, D.: Modeling of set/reset operations in NiO-based resistive-
switching memory devices. IEEE Trans. Electron Devices 56, 1712 (2009)
91. Jeong, D.S., Choi, B.J., Hwang, C.S.: Study of the negative resistance phenomenon in
transition metal oxide films from a statistical mechanics point of view. J. Appl. Phys. 100,
113724 (2006)
92. Seo, S., Lee, M.J., Kim, D.C., Ahn, S.E., Park, B.-H., Kim, Y.S., Yoo, I.K., Byun, I.S.,
Hwang, I.R., Kim, S.H., Kim, J.-S., Choi, J.S., Lee, J.H., Jeon, S.H., Hong, S.H., Park, B.H.:
Electrode dependence of resistance switching in polycrystalline NiO films. Appl. Phys. Lett.
87, 263507 (2005)
93. Lee, C.B., Kang, B.S., Lee, M.J., Ahn, S.E., Stefanovich, G., Xianyu, W.X., Kim, K.H., Hur,
J.H., Yin, H.X., Park, Y., Yoo, I.K., Park, J.-B., Park, B.H.: Electromigration effect of Ni
Electrodes on the resistive switching characteristics of NiO thin films. Appl. Phys. Lett. 91,
082104 (2007)
94. Lee, C.B., Kang, B.S., Benayad, A., Lee, M.J., Ahn, S.-E., Kim, K.H., Stefanovich, G., Park,
Y., Yoo, I.K.: Effects of metal electrodes on the resistive memory switching property of NiO
thin films. Appl. Phys. Lett. 93, 042115 (2008)
95. Xu, N., Liu, L.F., Sun, X., Chen, C., Wang, Y., Han, D.D., Liu, X.Y., Han, R.Q., Kang, J.F.,
Yu, B.: Bipolar switching behavior in TiN/ZnO/Pt resistive nonvolatile memory with fast
switching and long retention. Semicond. Sci. Technol. 23, 075019 (2008)
96. Choi, J.S., Kim, J.-S., Hwang, I.R., Hong, S.H., Jeon, S.H., Kang, S.-O., Park, B.H., Kim, D.
C., Lee, M.J., Seo, S.: Different resistance switching behaviors of NiO thin films deposited on
Pt and SrRuO3 electrodes. Appl. Phys. Lett. 95, 022109 (2009)
97. Lin, C.-Y., Wu, C.-Y., Wu, C.-Y., Lee, T.-C., Yang, F.-L., Hu, C., Tseng, T.-Y.: Effect of top
electrode material on resistive switching properties of ZrO2 film memory devices. IEEE
Electron Device Lett. 28, 366 (2007)
98. Yoshida, C., Tsunoda, K., Noshiro, H., Sugiyama, Y.: High speed resistive switching in Pt/
TiO2/TiN film for nonvolatile memory application. Appl. Phys. Lett. 91, 223510 (2007)
99. Gao, B., Zhang, H.W., Yu, S., Sun, B., Liu, L.F., Liu, X.Y., Wang, Y., Han, R.Q., Kang, J.F.,
Yu, B., Wang, Y.Y.: Oxide-based RRAM: Uniformity improvement using a new material-
oriented methodology. In: 2009 Symposium on VLSI technology, Honolulu, HI, p. 30 (2009)
100. Lin, C.-Y., Wu, C.-Y., Wu, C.-Y., Tseng, T.Y., Hu, C.: Modified resistive switching behavior
of ZrO2 memory films based on the interface layer formed by using Ti top electrode. J. Appl.
Phys. 102, 094101 (2007)
101. Zhou, P., Yin, M., Wan, H.J., Lu, H.B., Tang, T.A., Lin, Y.Y.: Role of TaON interface for
CuxO resistive switching memory based on a combined model. Appl. Phys. Lett. 94, 053510
(2009)
102. Yu, S., Wong, H.-S.P.: A phenomenological model of oxygen ion transport for metal oxide
resistive switching memory. In: Memory Workshop (IMW), 2010 IEEE International, Seoul
(2010)
103. Sun, X., Sun, B., Liu, L.F., Xu, N., Liu, X.Y., Han, R.Q., Kang, J.F., Xiong, G., Ma, T.P.:
Wavelength tuning of GaAs LED’s through surface effects. IEEE Electron Device Lett. 30,
334 (2009)
104. Sun, B., Liu, Y.X., Liu, L.F., Xu, N., Wang, Y., Liu, X.Y., Han, R.Q., Kang, J.F.: Highly
uniform resistive switching characteristics of TiN/ZrO2/Pt memory devices. J. Appl. Phys.
105, 061630 (2009)
105. Kim, S.I., Lee, J.H., Chang, Y.W., Hwang, S.S., Yoo, K.-H.: Reversible resistive switching
behaviors in NiO nanowires. Appl. Phys. Lett. 93, 033503 (2008)
13 Metal Oxide Resistive Switching Memory 333

106. Oka, K., Yanagida, T., Nagashima, K., Tanaka, H., Kawai, T.: Nonvolatile bipolar resistive
memory switching in single crystalline NiO heterostructured nanowires. J. Am. Chem. Soc.
131, 3434 (2009)
107. Seo, J.W., Park, J.-W., Lim, K.S., Yang, J.-H., Kang, S.J.: Transparent resistive random
access memory and its characteristics for nonvolatile resistive switching. Appl. Phys. Lett.
93, 223505 (2008)
108. Kim, S., Choi, Y.-K.: Resistive switching of aluminum oxide for flexible memory. Appl.
Phys. Lett. 92, 223508 (2008)
109. Kim, S., Moon, H., Gupta, D., Yoo, S., Choi, Y.-K.: Resistive switching characteristics of sol-
gel zinc oxide films for flexible memory applications. IEEE Trans. Electron Devices 56, 696
(2009)
110. Yun, D.K., Kim, K.-D., Kim, S., Lee, J.-H., Park, H.-H., Jeong, J.-H., Choi, Y.-K., Choi, D.-
G.: Mass fabrication of resistive random access crossbar arrays by step and flash impring
lithography. Nanotechnology 20, 445305 (2009)
111. Lee, B., Wong, H.-S.P.: NiO resistance change memory with a novel structure for 3D
integration and improved confinement of conduction path. In: 2009 Symposium on VLSI
technology, Honolulu, HI, p. 28 (2009)
112. Wu, Y., Lee, B., Wong, H.-S.P.: Ultra-low power Al2O3-based RRAM with 1mA reset
current. In: 2010 International symposium on VLSI Technology Systems and Applications
(VLSI-TSA), Hsinchu, p. 136 (2010)
113. Ahn, S.-E., Lee, M.-J., Park, Y., Kang, B.S., Lee, C.B., Kim, K.H., Seo, S., Suh, D.-S., Kim,
D.-C., Hur, J., Xianyu, W., Stefanovich, G., Yin, H., Yoo, I.-K., Lee, J.-H., Park, J.-B., Baek,
I.-G., Park, B.H.: Write current reduction in transition metal oxide based resistance change
memory. Adv. Mater. 20, 924 (2008)
114. Kinoshita, K., Tsunoda, K., Sato, Y., Noshiro, H., Yagaki, S., Aoki, M., Sugiyama, Y.:
Reduction in the reset current in a resistive random access memory consisting of NiOx
brought about by reducing a parasitic capacitance. Appl. Phys. Lett. 93, 033506 (2008)
115. Chen, A., Haddad, S., Wu, Y.-C., Fang, T.-N., Lan, Z., Avanzino, S., Pangrle, S., Buynoski,
M., Rathor, M., Cai, W., Tripsas, N., Bill, C., VanBuskirk, M., Taguchi, M.: Non-volatile
resistive switching for advanced memory applications. In: IEDM Technical Digest,
Washington, DC, p. 765 (2005)
116. Chen, A., Haddad, S., Wu, Y.C., Fang, T.N., Kaza, S., Lan, Z.: Erasing characteristics of
Cu2O metal-insulator-metal resistive switching memory. Appl. Phys. Lett. 92, 013503 (2008)
117. Sato, Y., Tsunoda, K., Kinoshita, K., Noshiro, H., Aoki, M., Sugiyama, Y.: Sub-100 mA reset
current of nickel oxide resistive memory through control of filamentary conductance by
current limit of MOSFET. IEEE Trans. Electron Devices 55, 1185 (2008)
118. Ielmini, D., Cagli, C., Nardi, F.: Resistance transition in metal oxides induced by electronic
threshold switching. Appl. Phys. Lett. 94, 063511 (2009)
119. Chen, A., Haddad, S., Wu, Y.C., Lan, Z., Fang, T.N., Kaza, S.: Switching characteristics of
Cu2O metal-insulator-metal resistive memory. Appl. Phys. Lett. 91, 123517 (2007)
120. Choi, B.J., Choi, S., Kim, K.M., Shin, Y.C., Hwang, C.S., Hwang, S.-Y., Cho, S.-S., Park, S.,
Hong, S.-K.: Study on the resistive switching time of TiO2 thin films. Appl. Phys. Lett. 89,
012906 (2006)
121. Tsunoda, K., Kinoshita, K., Noshiro, H., Yamazaki, Y., Iizuka, T., Ito, Y., Takahashi, A.,
Okano, A., Sato, Y., Fukano, T., Aoki, M., Sugiyama, Y.: Low power and high speed
switching of Ti-doped NiO ReRAM under the unipolar voltage source of less than 3 V. In:
Electron devices meeting, 2007. IEDM 2007. IEEE International, Washington, DC, p. 767
(2007)
122. Lv, H.B., Yin, M., Fu, X.F., Song, Y.L., Tang, L., Zhou, P., Zhao, C.H., Tang, T.A., Chen, B.
A., Lin, Y.Y.: Resistive memory switching of CuxO films for a nonvolatile memory applica-
tion. IEEE Electron Device Lett. 29, 309 (2008)
334 S. Yu et al.

123. Chen, A., Haddad, S., Wu, Y.-C.: A temperature-accelerated method to evaluate data
retention of resistive switching nonvolatile memory. IEEE Electron Device Lett. 29, 38
(2008)
124. Kim, D.C., Lee, M.J., Ahn, S.E., Seo, S., Park, J.C., Yoo, I.K., Baek, I.G., Kim, H.J., Yim, E.
K., Lee, J.E., Park, S.O., Kim, H.S., Chung, U.-I., Moon, J.T., Ryu, B.I.: Improvement of
resistive memory switching in NiO using IrO2. Appl. Phys. Lett. 88, 232106 (2006)
125. Chang, W.-Y., Cheng, K.-J., Tsai, J.-M., Chen, H.-J., Chen, F., Tsai, M.-J., Wu, T.-B.:
Improvement of resistive switching characteristics in TiO2 thin films with embedded Pt
nanocrystals. Appl. Phys. Lett. 95, 042104 (2009)
126. Yu, S., Gao, B., Dai, H.B., Sun, B., Liu, L.F., Liu, X.Y., Han, R.Q., Kang, J.F., Yu, B.:
Improved uniformity of resistive switching behaviors in HfO2 thin films with embedded Al
layers. Electrochem. Solid State Lett. 13, H36 (2010)
127. Liu, Q., Long, S., Wang, W., Zuo, Q., Zhang, S., Chen, J., Liu, M.: Improvement of resistive
switching properties in ZrO2-based ReRAM with implanted Ti ions. IEEE Electron Device
Lett. 30, 1335 (2009)
128. Yoo, I.K., Kang, B.S., Park, Y.D., Lee, M.J., Park, Y.: Interpretation of nanoscale conducting
paths and their control in nickel oxide (NiO) thin films. Appl. Phys. Lett. 92, 202112 (2008)
129. Yin, M., Zhou, P., Lrav, H.B., Xu, J., Song, Y.L., Fu, X.F., Tang, T.A., Chen, B.A., Lin, Y.Y.:
Improvement of resistive switching in CuxO using new RESET mode. IEEE Electron Device
Lett. 29, 681 (2008)
130. Liang, J., Wong, H.-S.P.: Size limitation of cross-point memory array and its dependence on
data storage pattern and device parameters. In: Interconnect Technology Conference (IITC),
2010 International, Burlingame, CA (2010)
131. Lee, M.-J., Seo, S., Kim, D.-C., Ahn, S.-E., Seo, D.H., Yoo, I.-K., Baek, I.-G., Kim, D.-S.,
Byun, I.-S., Kim, S.-H., Hwang, I.-R., Kim, J.-S., Jeon, S.-H., Park, B.H.: A low-temperature-
grown oxide diode as a new switch element for high-density nonvolatile memories. Adv.
Mater. 19, 73 (2007)
132. Baek, I.G., Kim, D.C., Lee, M.J., Kim, H.-J., Yim, E.K., Lee, M.S., Lee, J.E., Ahn, S.E., Seo,
S., Lee, J.H., Park, J.C., Cha, Y.K., Park, S.O., Kim, H.S., Yoo, I.K., Chung, U-I., Moon J.T.,
Ryu, B.I.: Multi-layer cross-point binary oxide resistive memory (OxRRAM) for post-NAND
storage application. In: Electron devices meeting, 2005. IEDM Technical Digest. IEEE
International, Washington, DC, p. 750 (2005)
133. Lee, M.-J., Park, Y., Kang, B.-S., Ahn, S.-E., Lee, C., Kim, K., Xianyu, W., Stefanovich, G.,
Lee, J.-H., Chung, S.-J., Kim, Y.-H., Lee, C.-S., Park, J.-B., Baek, I.-G., Yoo, I.-K.: 2-Stack
1D-1R cross-point structure with oxide diodes as switch elements for high density resistance
RAM applications. In: Electron devices meeting, 2007. IEDM 2007. IEEE International,
Washington, DC, p. 771 (2007)
134. Lee, M.-J., Lee, C.B., Kim, S., Yin, H., Park, J., Ahn, S.E., Kang, B.S., Kim, K.H.,
Stefanovich, G., Song, I., Kim, S.-W., Lee, J.H., Chung, S.J., Kim, Y.H., Lee, C.S., Park, J.
B., Baek, I.G., Kim, C.J., Park, Y.: Stack friendly all-oxide 3D RRAM using GaInZnO
peripheral TFT realized over glass substrates. In: Electron devices meeting, 2008. IEDM
2008. IEEE International, San Francisco, CA, p. 85 (2008)
135. Yoon, H.S., Baek, I.-G., Zhao, J., Sim, H., Park, M.Y., Lee, H., Oh, G.-H., Shin, J.C., Yeo,
I.-S., Chung, U.-I.: Vertical cross-point resistance change memory for ultra-high density non-
volatile memory applications. In: 2009 Symposium on VLSI technology, Honolulu, HI, p. 26
(2009)
136. Ohya, Y., Uedo, M., Takahasi, Y.: Oxide thin film diode fabricated by liquid-phase method.
Jpn. J. Appl. Phys. 35, 4738 (1996)
137. Lee, W.Y., Mauri, D., Hwang, C.: High-current-density ITOx/NiOx thin-film diodes. Appl.
Phys. Lett. 72, 1584 (1998)
138. Shin, Y.C., Song, J., Kim, K.M., Choi, B.J., Choi, S., Lee, H.J., Kim, G.H., Eom, T., Hwang,
C.S.: (In,Sn)2O3/TiO2/Pt Schottky-type diode switch for the TiO2 resistive switching mem-
ory array. Appl. Phys. Lett. 92, 162904 (2008)
13 Metal Oxide Resistive Switching Memory 335

139. Okimura, K., Sakai, J.: Time-dependent characteristics of electric field-induced metal-insulator
transition of planer VO2/c-Al2O3 structure. Jpn. J. Appl. Phys. 46, L813 (2007)
140. Lee, M.-J., Park, Y., Suh, D.-S., Lee, E.-H., Seo, S., Kim, D.-C., Jung, R., Kang, B.-S., Ahn,
S.-E., Lee, C.B., Seo, D.H., Cha, Y.-K., Yoo, I.-K., Kim, J.-S., Park, B.H.: Two series oxide
resistors applicable to high speed and high density nonvolatile memory. Adv. Mater. 19, 3919
(2007)
141. Lee, M.-J., Kim, S.I., Lee, C.B., Yin, H., Ahn, S.-E., Kang, B.S., Kim, K.H., Park, J.C., Kim,
C.J., Song, I., Kim, S.W., Stefanovich, G., Lee, J.H., Chung, S.J., Kim, Y.H., Park, Y.: Low-
temperature-grown transition metal oxide based storage materials and oxide transistors for
high-density non-volatile memory. Adv. Funct. Mater. 19, 1587 (2009)
142. Chang, W.-Y., Ho, Y.-T., Hsu, T.-C., Chen, F., Tsai, M.-J., Wu, T.-B.: Influence of crystalline
constituent on resistive switching properties of TiO2 memory films. Electrochem. Solid State
Lett. 12, H135 (2009)
143. Shima, H., Takano, F., Muramatsu, H., Akinaga, H., Tamai, Y., Inque, I.H., Takagi, H.:
Voltage polarity dependent low-power and high-speed resistance switching in CoO resistance
random access memory with Ta electrode. Appl. Phys. Lett. 93, 113504 (2008)
144. Lee, D., Seong, D.-J., Choi, H.J., Jo, I., Dong, R., Xiang, W., Oh, S., Pyun, M., Seo, S.-O.,
Heo, S., Jo, M., Hwang, D.-K., Park, H.K., Chang, M., Hasan, M., Hwang, H.: Excellent
uniformity and reproducible resistance switching characteristics of doped binary metal oxides
for non-volatile resistance memory applications. In: IEDM Technical Digest. IEEE Interna-
tional, San Francisco, CA, p. 1 (2006)
145. Muraoka, S., Osano, K., Kanzawa, Y., Mitani, S., Fujii, S., Katayama, K., Katoh, Y., Wei, Z.,
Mikawa, T., Arita, K., Kawashima, Y., Azuma, R., Kawai, K., Shimakawa, K., Odagawa, A.,
Takagi, T.: Fast switching and long retention Fe-O ReRAM and its switching mechanism.
In: IEDM Technical Digest. IEEE International, Washington, DC, p. 779 (2007)
146. Do, Y.H., Kwak, J.S., Bae, Y.C., Jung, K., Im, H., Hong, J.P.: Hysteretic bipolar resistive
switching characteristics in TiO2 /TiO2−x multilayer homojunctions. Appl. Phys. Lett. 95,
093507 (2009)
Chapter 14
Nanostructured Metal Oxides
for Li-Ion Batteries

Juchen Guo and Chunsheng Wang

Nanoscale and nanostructured metal oxides have drawn tremendous interests from
the researchers working in the field of energy storage and energy conversion
technologies in recent years. In this chapter, the state of the art of nano metal
oxide materials in Li-ion batteries will be discussed as a comprehensive overview.
Lithium-ion battery has been a very important category of rechargeable batteries
since its first commercialization by Sony in 1991. It has been widely used in
portable consumer electronics such as laptops, digital cameras, small power tools,
etc. However, its potential is not limited to such small devices due to several unique
merits: Li-ion battery has the highest energy density among all types of recharge-
able batteries that are currently on the market. It also has a relatively low self-
discharge rate. Because of these virtues, interests in Li-ion batteries keep growing
for defense, aerospace, smart grid system, and automotive applications. To satisfy
the demands of these emerging applications, the next generation of Li-ion batteries
must achieve a holistic and striking advancement from the current technology,
specifically in four criteria: energy density, discharging and charging rate (power
density), safety feature, and cycle stability. The energy density, power density, and
cycle stability of Li-ion batteries are mainly determined by electrode materials and
structures. Enhancement of the safety feature ultimately depends on the develop-
ment of nonflammable electrolyte and solid electrolyte to replace the current liquid
electrolyte consisting of highly flammable organic solvents. The use of nano metal
oxides (nanoscale or nanostructured) as anode materials, cathode materials, and
electrolyte additives has greatly enhanced the performance of Li-ion batteries due
to their unique chemical and structural properties.

J. Guo • C. Wang (*)


Department of Chemical and Biomolecular Engineering,
University of Maryland, College Park, MD 20742, USA
e-mail: cswang@umd.edu

J. Wu et al. (eds.), Functional Metal Oxide Nanostructures, Springer Series 337


in Materials Science 149, DOI 10.1007/978-1-4419-9931-3_14,
# Springer Science+Business Media, LLC 2012
338 J. Guo and C. Wang

14.1 Classification of Electrode Materials for Li-Ion Batteries

Before going into further discussion, it is necessary to briefly introduce how Li-ion
battery works. For instance, the most common cathode material is lithium cobalt
oxide (LiCoO2) and the anode material is graphite as shown in Fig. 14.1. The
typical electrolyte consists of lithium salts like lithium hexafluorophosphate
(LiPF6) or lithium tetrafluoroborate (LiBF4) dissolved in an organic solvent such
as a mixture of propylene carbonate and diethyl carbonate. During the battery
charging process, Li atoms in LiCoO2 become ions, migrating to the graphite
anode across the electrolyte and inserting into the gaps between the graphene
layers. The reverse process takes place during the battery discharge: Li atoms
stored in the layered graphite become ions, migrating to the LiCoO2 cathode and
inserting into the layers of octahedral lattices formed by cobalt and oxygen atoms.
This type of lithiation/delithiation mechanism is referred as intercalation in which
Li is stored in layer-structured materials such as graphite and LiCoO2. Nowadays,
the intercalation mechanism has been generalized to refer to all topotactic reactions
of Li-ions inserting into the interior of the lattice of the host materials of which the
structures are not limited to be layered. Another lithiation/delithiation mechanism is
based on the reversible redox reaction between metal oxide and Li (Li+), and is
referred to as conversion reaction. According to the conversion mechanism,
lithiation takes place through the reduction of metal oxide by Li to produce metal
and lithium oxide, and delithiation takes place through the oxidation of the formed
metal by lithium oxide. Besides, these two mechanisms, in a few binary

Fig. 14.1 Schematic of Li-ion battery with graphite anode and LixMO2 cathode in the state of
discharge
14 Nanostructured Metal Oxides for Li-Ion Batteries 339

intermetallic AB compounds, Li will reversibly displace A to form LixB, and the


formed LixB has a strong structural relationship with the parent AB compound. This
mechanism is referred to as displacement reaction. All current metal oxide elec-
trode materials for Li-ion batteries can be sorted into these three categories, except
tin dioxide (SnO2)-based materials in which the lithiation/delithiation process
combines conversion and alloying reactions. Therefore, for the purpose of articula-
tion, all the metal oxide electrode materials will be discussed with respect to their
different lithiation/delithiation mechanisms.

14.2 Advantages and Disadvantages


of Nanoelectrode Materials

The advantages of nanoelectrode materials come from their nanometer characteris-


tic length and their tremendously large surface area. Generally speaking, the
smaller size can shorten the Li-ion/electron transport pathways and enhance
phase transformation. The large surface area can also speed up the charge transfer
reaction kinetics due to the increased contact area with the electrolyte. Enhanced Li
insertion/extraction kinetics can lead to higher rate performance, even novel
lithiation/delithiation mechanism. Higher surface area can also enhance the capac-
ity through the surface Li storage mechanism [1]. Moreover, nanoelectrode
materials can better accommodate the mechanical strain induced by the concomi-
tant volume change in the lithiation/delithiation process, thus improving the cycle
stability.
Unfortunately, the disadvantages of nanoelectrode materials are also from their
nano characteristic scale and large surface area. The nanoscale materials will lower
the packing density of the electrode thus resulting in a low overall energy density of
the batteries. Also, the large surface area will promote larger amount of side
reactions at the electrolyte/electrode interfaces. Therefore, the development of
nanoelectrode materials should focus on the direction of the optimized properties
balancing the advantages and disadvantages.

14.3 Nanometal Oxide Anode Materials

14.3.1 Intercalation Metal Oxides

Materials with a layered structure (graphite for anode and LiCoO2 for cathode) are
the natural choice as the Li host material [2], but not the only ones. Materials with
tunneled structures (such as spinel) can also be used as Li storage hosts with
intercalation mechanism. Among them, Li4Ti5O12 and TiO2 are the two most
intensively studied metal oxide anode materials.
340 J. Guo and C. Wang

The concept of using the B2O4 framework of an AB2O4 (A ¼ Li) spinel material
as host structure for Li-ion storage was originally proposed by Thackeray et al. in
the 1980s [3]. Lithium titanium oxide, Li4Ti5O12, a ceramic material having a
defect tunneled ([Li1/3Ti5/3]O4) structure, was initially proposed as an anode mate-
rial by Colbow et al. in 1989 [4] and tested by Ferg et al. [5] and Ohzuku et al.
[6] in the early 1990s. Li4Ti5O12 can be lithiated over the composition range of
Li4+xTi5O12 (0 < x < 3) at a potential of about 1.55 V versus Li/Li+, and its
theoretical lithiation capacity is 175 mAh g1. Despite its moderate lithiation
capacity, the particular advantage of Li4Ti5O12 comparing to other spinel anodes
was that it is a “zero-strain” intercalation material. The defect Li1/3Ti5/3O4 spinel
framework exhibits minimal volume change during Li-ion insertion and extraction
so that the crystal structure is better retained, thus resulting in better cycle life.
Another characteristic of Li4Ti5O12 is its lithiation voltage of 1.55 V versus Li/Li+,
which is considered to have two-faced effects. On one hand, the 1.55 V lithiation
voltage is higher than the decomposition voltage of the organic solvents in the
electrolyte. Therefore, using Li4Ti5O12 as the anode material can eliminate
the formation of solid electrolyte interface (SEI) film, which a considerable cost
efficiency factor. Also, the higher lithiation voltage significantly reduces the possi-
bility of the lithium metal plating at the anode, so that the safety can be enhanced.
On the other hand, using Li4Ti5O12 anode sacrifices the full cell working voltage
because of its higher lithiation voltage compared to graphite (0.1–0.2 V versus Li/Li+).
Since pure Li4Ti5O12 is an electric insulator, the advantage of nanoscale
Li4Ti5O12 is the extraordinary enhancement of Li insertion/extraction kinetics.
The mean Li-ion diffusion time in an ideal anode particle can be approximately
expressed using the following equation, assuming Fickian diffusion:

L2

2D

where L is the diffusion distance and D is the Li-ion diffusivity in the material.
Based on this equation, the advantage of nanoscale electrode material is obvious:
the resultant short diffusion distance can reduce the diffusion time significantly. For
instance, if the particle size is reduced to 100 nm from 1 mm, Li-ion diffusion time
can be decreased 100 times. Another advantage of nanoscale electrode materials for
charge transfer kinetics enhancement is their large surface area which results in
large contact surface between the electrode and electrolyte. As an electrically
insulating material, the electronic conductivity of Li4Ti5O12 increases during the
lithiation reaction from the outer surface directing inward, which is not critically
problematic for lithiation, since the Li+/e transport takes place at the outer layer
anyway. During delithiation, as Li is being extracted, the conductivity starts to
decrease from the outer layer of the Li4Ti5O12 particle. Therefore, the delithiation
process has worse kinetics than lithiation. Fast separation of Li+ and e is critical
to achieve fast charge/discharge rate, which can be achieve by reducing the Li+/e
transport pathway using nanoscale Li4Ti5O12 materials.
14 Nanostructured Metal Oxides for Li-Ion Batteries 341

Fig. 14.2 (a) TEM image of the Li4Ti5O12 nanorods, (b) rate capacity test of the Li4Ti5O12
nanorods at different C rates [7]

Because of these advantages, nanoscale Li4Ti5O12 anode materials have been


extensively studied. Among them, Kim and Cho [7] reported the synthesis and
electrochemical performance of Li4Ti5O12 nanorods. The diameter of the reported
Li4Ti5O12 nanorods is about 100 nm as shown in Fig. 14.2a. The notable merit of
this material is its very promising discharge rate capacity. As shown in Fig. 14.1b,
the reversible first discharge capacity was 165 mAh g1 under a cycling rate of 0.l C
(16 mA g1), and no capacity fading was observed up to 30 cycles between 1 and
2.5 V. At rates of 0.5 C, the first capacity at 0.5 C was identical to that at 0.1 C. Very
small capacity decreases with increasing current were observed at 5 and 10 C
(1,600 mA g1) rates, the capacity retention was 95 and 93%, showing 157 and
155 mAh g1, respectively. As a comparison, the electrochemical performance of
Li4Ti5O12 particles with 700 nm diameter is distinctly worse. Though this cannot be
used as a direct evidence of the superiority of nanorods over nanoparticles due to
their different characteristic sizes, it clearly demonstrates the significant advantage
of nanoscale materials by enhancing charge transfer kinetics.
In light of the great promise of Li4Ti5O12 as Li intercalation anode materials,
researchers naturally began to investigate titanium dioxides as candidates of anode
materials because of their higher theoretical lithiation capacity. The Li intercalation
reaction to TiO2 can be generally expressed as the following reaction:

xLiþ þ TiO2 þ xe $ Lix TiO2

Full lithiation should lead to the formation of lithium titanium oxide with a formula
of LiTiO2 (x ¼ 1) with 335 mAh g1 theoretical capacity. This reaction takes place
in the voltage range 1.5–1.8 V. Therefore, like Li4Ti5O12, using TiO2 as the anode
material can avoid anode passivation and also enhance the safety feather. The
investigation on TiO2 was actually not a recent idea: it started in the 1980s and
continued in the 1990s [8–11]. However, the sluggish performance of the earlier
TiO2 materials had merely attracted lukewarm attention. The TiO2 research really
took off in virtue of the development of nanotechnology. To date, four types of
titanium dioxides have been reported to have lithiation capacity: are rutile,
342 J. Guo and C. Wang

Table 14.1 Data for TiO2 polymorphs for anode materials [12]
Type Density (g cm3) Unit-cell data (nm) Structure
Rutile 4.13 a ¼ 0.459
c ¼ 0.296

Brookite 3.99 a ¼ 0.917


b ¼ 0.546
c ¼ 0.514

Anatase 3.79 a ¼ 0.379


c ¼ 0.951

Bronze TiO2(B) 3.64 a ¼ 1.216


b ¼ 0.374
c ¼ 0.651
b ¼ 107.29º

brookite, anatase, and bronze (TiO2(B)). These polymorphs are the only known
naturally occurring TiO2 forms to date, even bronze is rare. Rutile is the most
common and the most thermodynamically stable form of TiO2. Anatase and brookite
can be converted to rutile upon heating in a temperature range of 700–1,000 C. The
basic physical and structural properties of these TiO2 polymorphs are listed in
Table 14.1 [12].
The main problem of TiO2 polymorphs as anode materials is their poor Li+ and
electron conductivity so that the Li lithiation/delithiation reaction kinetics was
largely hindered. Recent studies suggested that the TiO2 lithiation/delithiation
reaction kinetics, ultimately the rate performance, is closely related to its crystal
structural properties such as site occupation, local coordination, and energetics
[13–15]. For instance, the thermodynamically stable positions for Li insertion in
rutile are the octahedral sites in the ab planes. It has been proved that the Li+
14 Nanostructured Metal Oxides for Li-Ion Batteries 343

Fig. 14.3 (a) Galvanostatic cycling curves of rutile TiO2 samples using a 30 mAh g1 current
between 3 and 1 V at 20 C; (b) the capacity retention for these samples

diffusion in rutile is anisotropic: the theoretical diffusion coefficient of Li-ion along


ab plane is 8 orders of magnitude lower than that along the c axis (1014 cm2 s1 and
106 cm2 s1, respectively) [16–22]. The repulsive interaction between the Li ions
diffusing along the c axis may slow down the diffusion, and the Li-ion pairs in the ab
plane may also block the Li-ion diffusion along the c axis [18, 22]. Therefore, the low
Li insertion capacity of rutile is mainly restrained by poor Li-ion transport kinetics.
Recent studies suggested that the lithiation capacity of rutile could be increased by
reducing the structural size. Jiang et al. reported full lithiation in the first cycle using
nano-sized needlelike rutile particles (15 nm) for the first time [23]. Also, 0.7 Li per
unit of rutile can be extracted in the first delithiation. Reducing the rutile particle size
has threefold advantages. First, it decreases the Li+ and electron diffusion pathway.
Second, the mechanical strain during Li insertion is reduced. Finally, because of the
enormously enhanced surface area, Li surface storage capacity is increased [24].
Other remarkable rutile works include the rutile nanorods (10 nm  40 nm) reported
by Hu et al. [24] and nanowires (10 nm  200 nm) reported by Baudrin et al. [25].
The later rutile nanowires demonstrated much superior capacity than bulk rutile and
even nano-sized rutile particles as shown in Fig. 14.3.
Although the brookite has been less reported for its Li storage capacity com-
pared to other TiO2 polymorphs, the reported performance of brookite indicated
strong dependence on the particle size. The 10-nm-sized brookite particles deliv-
ered reversible capacity of 170 mAh g1 for more than 40 cycles, reported by
Reddy et al. [26, 27]. For anatase, the Li+ diffusion in it is more facile compared to
rutile because of its looser lattice structure. Upon the Li uptake into anatase, its
original lattice structure of tetragonal body-centered I41/amd space group changes
to orthorhombic pmn21 space group when 0.5 Li per unit of anatase is inserted [28].
Also, the lithiation results in 4% volume expansion thus causing rapid capacity fade
for bulk anatase material [29]. Reducing the anatase particle size again can shorten
the Li-ion diffusion pathway and increase the Li surface storage capacity due to the
344 J. Guo and C. Wang

Fig. 14.4 (a) TEM image of TiO2(B) wires [33]; (b) TEM image of TiO2(B) tubes [34]; (c)
variation of potential with Li content for TiO2(B) nanowires and TiO2(B) nanotubes cycled under
identical conditions [34]

large surface area [30, 31]. For example, Gao et al. reported first discharge and
charge capacities of 340 and 200 mAh g1, respectively, for the anatase nanotubes
with diameter of 10–15 nm and length of 200–400 nm [32].
TiO2(B) (bronze) is rare in nature so, all reported TiO2(B) anode materials to date
were synthesized. The advantage of bronze compared to other TiO2 polymorphs is
its more open lattice structure that facilitates the Li insertion. Armstrong et al.
synthesized TiO2(B) nanowires (20–40 nm diameter and 2–10 mm length) [33]
and nanotubes (10–20 nm outer diameter, 5–8 nm inner diameter and ~1 mm length)
[34] as shown in Fig. 14.4a, b, respectively. The TiO2(B) nanowires demonstrated
superior lithiation capacity (Li0.91TiO2, 305 mAh g1 specific capacity) to the bulk
material (Li0.71TiO2, 240 mAh g1 specific capacity). During the lithiation process,
there is no detectable volume change taking place due to the more open lattice
structure. Compared to nanoscale TiO2(B) particles with similar diameter, even both
showed similar first cycle lithiation capacity, the capacity retention of the nanowires
was far more better. The TiO2(B) nanotubes showed marginally higher lithiation
capacity (Li0.98TiO2, 325 mAh g1 specific capacity) than the nanowires. However,
the TiO2(B) nanowires demonstrated better kinetics in spite of their large diameter.
As shown in Fig. 14.4c, the plateaus of the charge/discharge curves of TiO2(B)
nanowires are flatter and closer to each other, which indicates a small overpotential.
The irreversible capacity of the nanowires is also smaller.
TiO2 polymorphs have demonstrated very attractive electrochemical properties
as anode materials, such as higher lithiation voltage avoiding electrode passivation
and enhancing safety feature. The nanoscale TiO2 further improved the charge
transfer reaction kinetics by shortening the Li+ and electron transport pathway.
However, there are still a few intrinsic disadvantages of TiO2 that may need further
investigation. As it has relatively higher lithiation potential versus the Li/Li+ redox
couple, the full cells with TiO2 anode and typical cathode are subject to lower cell
voltages. However, exceptions may be possible if a high-potential cathode material
can be found. One example is TiO2(B) nanotubes/Li[Ni0.5Mn1.5]O4 cell reported by
14 Nanostructured Metal Oxides for Li-Ion Batteries 345

Armstrong et al. [35], which could achieve a 3 V overall cell potential. The other
serious problem of nanoscale TiO2 materials is the continuously irreversible capac-
ity on every cycle, except one report by Armstrong et al. [33]. The irrecoverable
capacity is mainly attributed to the electric insulating nature of the TiO2. As
previously mentioned in this chapter, the electric insulation can hurt Li extraction
more than insertion. Because the outer layer of the particle becomes electrically
insulating, it will be more difficult to extract all inserted Li. The current develop-
ment indicates that even after the dimension of TiO2 was reduced to the scale of
tens of nanometers, the transport pathway was still not efficient enough for fast
electron and Li+ separation to completely deplete the inserted Li. This problem can
definitely jeopardize the real application of TiO2 as Li-ion battery anode. One
solution is to incorporate electric conductive nano-sized composite into TiO2 to
form nanostructured material. One great example is that Guo et al. recently reported
a mesoporous RuO2-anatase TiO2 composite [36]. The RuO2 nanoparticles formed
an electric conductive network in the mesoporous TiO2 structure so that the
electrochemical performance was enhanced. Liu and coworkers reported a hybrid
nanostructure of rutile TiO2 and graphene [37]. Both works demonstrated reduced
irreversible capacity and largely improved fast charge/discharge performance.
Other reported intercalation metal oxides include oxides of vanadium, niobium
from Group 5B and molybdenum, tungsten from Group 6B in the Periodic Table of
the Elements. The concept of using these oxides as Li-ion battery electrode taking
the advantage of their layered lattice structures was proposed by Whittingham et al.
in the 1970s [38]. Binary vanadium oxides with octahedral or distorted octahedral
coordination are known for all oxidation states between V5+ and V2+. The typical
high lithiation–delithiation voltages of most of the vanadium oxides makes them
suitable as cathode materials. One exception is VO2(B), a vanadium dioxide
formula having monoclinic metastable shear structure. Its advantage as Li storage
material is its structural stability arising from an increased edge sharing and the
consequent resistance to lattice shearing during cycling [39]. VO2(B) received
particular interest as an anode material for aqueous electrolyte Li-ion batteries
because of its proper electrode potential of 2.5 V versus Li/Li+ [40]. Flower-like
VO2(B) nanoparticles have been synthesized by Zhang et al. [41], and tested as
anode material coupled with LiMn2O4 cathode. The flowerlike VO2(B)
demonstrated superior electrochemical properties to VO2(B) nanobelt and caram-
bola-like nanoparticles [41]. However, the Li-ion batteries with aqueous phase
electrolyte based on VO2(B) suffers a very low energy density, which is about
75 mAh g1 at the first cycle. It can be attributed to the cell voltage restrict to avoid
H2O decomposition. Besides VO2(B), some vanadium oxide-based compounds
such as LiVO2 [42], LiV3O8 [43], MnV2O6 [44], RVO4 (R ¼ In, Cr, Fe) [45],
and LiMVO4 (M ¼ Cd, Co, Zn, Ni, Cu and Mg) [46] were also studied as anode
materials. The niobium oxides have similar properties as vanadium oxides. Ternary
niobium oxide such as Ag nanoparticle-doped LiNbO3 [47] was investigated as the
anode materials, as well as KNb5O13 and K6Nb10.8O30 [48]. LiNbO3 demonstrated
low lithiation voltage (<0.5 V) and high first cycle capacity. However, the first
delithiation capacity is only 12–13% of the first lithiation one. These niobium
346 J. Guo and C. Wang

oxides demonstrated capacities between 100 and 200 mAh g1 within the voltage
range of 1.0–1.5 V versus Li/Li+. Though there has not been nanoscale or
nanostructured niobium oxides reported to date, it can be speculated that these
types of materials can possess enhanced electrochemical performance.
Molybdenum dioxide (MoO2) and tungsten dioxide (WO2) as Li-ion battery
anode materials were systematically investigated by Auborn and Barberio in 1987
for the first time [48]. MoO2 nanoparticles were synthesized by Yang et al. by the
reduction of molybdenum trioxide (MoO3) [49]. This rutile-like MoO2 material
demonstrated a reversible capacity of 318 mAh g1 for 20 cycles at 5 mA cm2
current density. Also, 85% of the reversible capacity was within the range of
1.0–2.0 V. The same research group also synthesized MoO2 tremella-like
nanoparticles consisting of nanosheets, which showed reversible capacity of
about 600 mAh g1 at 0.5 mA cm2 current density [50]. Superior electrochemical
performance was achieved by Dillon and coworkers based on the nanoscale MoOx
consisting of Mo, MoO2, and MoO3 phases [51]. Reversible capacity of
600 mAh g1 over 50 cycles at C/5 charge/discharge rate. MoO3 and tungsten
trioxide (WO3) are also promising Li intercalation materials for their layered
structure. The lithiation/delithiation potential of WO3 is around 4 V so that it is a
cathode material [52]. On the other hand, MoO3 has proper Li intercalation
potential as anode material and relatively high theoretical capacity of
1,117 mAh g1. MoO3 has an orthorhombic crystal structure composed of distorted
MoO6 octahedral layers. The reported electrochemical performance of the nano-
scale MoO3 materials varies depending on different morphologies. Lithiated MoO3
nanobelt (200 nm wide and 2–6 mm long) was reported by Mai et al. [53]. Compared
with non-lithiated MoO3 nanobelt, the lithiated one showed lower lithiation capac-
ity, but better cycle stability and less lithiation–delithiation hysteresis. However,
the reported capacity (~250 mAh g1) was significantly lower than the theoretical
capacity and the plateaus of the lithiation–delithiation voltage plot are above
2.25 V, which was not ideal for anode due to the resultant low full cell potential.
Recently, a similar MoO3 nanobelt with a uniform carbon coating was reported by
Hassan et al. [54]. The reported result was very attractive: the carbon-coated MoO3
nanobelt could retain capacity of 1,064 mAh g1 after 50 cycles with 0.1 C between
3.0 and 0.05 V, and no trend of fade was observed. The first cycle lithiation capacity
(~1,300 mAh g1) exceeded the theoretical capacity which could be attributed to the
formation of the SEI film or the surface Li storage. Also, the volume change-induced
MoO3 pulverization was observed from the SEM image. MoO3 nanoparticles
(~20 nm) were synthesized by Lee and coworkers [55]. These MoO3 nanoparticles
demonstrated superior cycle stability and great capacity: Above 600 mAh g1
capacity could be delivered after 150 cycles between 3 V and 0.005 V with C/2
charge/discharge rate. The performance was compared to micron-sized MoO3
particles of which the performance is far inferior. Therefore, the excellent perfor-
mance of MoO3 nanoparticles can be attributed to their nano size.
As for the intercalation metal oxide anode materials, they all possess layered or
tunneled structures. Typical lithiation reaction always occurs with concomitant
lattice volume expansion which has negative influence on the cycle stability.
14 Nanostructured Metal Oxides for Li-Ion Batteries 347

The lithiation capacity of intercalation metal oxides is limited by the availability of


lattice sites and suppressed by the redox competition with other phases. Further-
more, most of the intercalation metal oxides have low Li+ diffusivity and electric
conductivity. Therefore, the slow ion/electron transport kinetics is really the bottle-
neck to achieve full capacity and fast charge/discharge rate. The advantages of
nanoscale materials mainly are the shortened Li+ and electron transport pathway
and the enhanced surface area. As a result, the kinetics can be greatly enhanced.
Higher grain interface area in nanoscale materials can also enhance the storage of
Li. The problem of the current intercalation metal oxide anode materials is the
irreversible capacity (low coulombic efficiency), which can still remain even if the
material size is reduced to tens of nanometers. This observation may indicate that
reducing size is not the panacea, the conductivity must also be enhanced. The
current solution for this problem is conductive layer coating and conductive
composite doping.

14.3.2 Conversion Metal Oxide Materials

In 2000, Poizot et al. reported a new lithiation–delithiation mechanism of Li storage


in transition metal oxides which can be used as anode materials [56]. This new
mechanism is different from the aforementioned Li insertion/extraction, and can
generally be expressed as

2xLi þ MOx $ M þ xLi2 O

ðM ¼ Fe; Ni; Co; Cu; Mn; Cr; Ru; ZnÞ

During the lithiation (metal oxide reduction) reaction, the metal oxide is reduced
by lithium at a certain potential. In the delithiation (metal oxide oxidation) reaction,
the metal is oxidized by Li2O to its original valence state. This mechanism is
referred to as conversion reaction. The lithiation through metal oxide reduction
has actually been well recognized and investigated for a long time. For instance, Li
and manganese dioxide (MnO2) is one anode/cathode pair for primary Li battery.
However, this reaction was considered as irreversible until the investigation of
Poizot et al. The detailed study shows that the lithiation resulted in 2–5 nm metal
grains embedded in an amorphous Li2O matrix [56]. This nanostructure greatly
reduces the Li+ and electron transport pathway, and facilitates the movement of Li+
through the Li2O phase and electrons through the metal phase, thus making
reversible lithiation/delithiation possible. Therefore, this conversion mechanism
is indeed one excellent example of the applications enabled by nanotechnology.
However, the delithiation capacity is not completely matched with the lithiation
capacity. Most of the results of conversion metal oxide anode materials showed
considerable amount of irreversible capacity in the first cycle. The only exception is
ruthenium dioxide reported by Balaya et al. [57], because RuO2 has an unusually
348 J. Guo and C. Wang

high electric conductivity. However, RuO2 is not a practical anode material because
of its rarity. The cause of the irreversibly capacity is the insufficient Li+ and
electron conductivity for complete delithiation. Another common problem for
metal oxide anode materials caused by the same reason is the large hysteresis of
charge/discharge voltage profile. The ideal electrode should have voltage hysteresis
as small as possible to have high energy density efficiency. The work of Li et al.
clearly demonstrated that reducing the particle size of the metal oxide could
effectively lower the potential hysteresis [58].
Since the original study by Poizot et al., numerous investigations on nanostruc-
tured metal oxides with conversion mechanism have been carried out. To date, the
most commonly investigated metals include iron, cobalt, nickel, copper, manga-
nese, and chromium. All valence states of these metal oxides can theoretically be
used as anode materials. However, the conversion of high-valence metal oxides is
complex compared to the low-valence ones with rock salt-like structure. Larcher
and coworkers demonstrated the Li-inserted intermediates in forms of LixCo3O4
and LixFe2O3 for the conversion of Co3O4 and a-Fe2O3, respectively [59, 60]. Both
Co3O4 and a-Fe2O3 possess tunneled spinel structures, thus Li initially inserting
into these materials following the intercalation mechanism. This process had been
long recognized. However, Larcher et al. found that further lithiation could result in
a deep structural modification to the rock-salt structure. Finally the metallic phase
was formed as nanograins dispersed in Li2O matrix. Similar process could be
applied to Fe3O4. The formation of the Li-intercalated intermediate phase is closely
related to the material size [60]. For instance, one Li could be taken into LixFe2O3
when nano-sized a-Fe2O3 was used, whereas x is only about 0.05 for micron-sized
a-Fe2O3. For anode application, the metal oxides should have high capacities which
make higher valence metal oxides such as Cr2O3, Mn2O3, and Fe2O3 a better choice .
They also should have low potentials, in terms of which the merit is in order
Cr2O3 > Mn2O3 > Fe2O3 based on theoretical calculations [61]. The spherical
Cr2O3 nanoparticle reported by Hu et al. exhibited a lithiation voltage lower than
0.5 V versus Li/Li+ and an initial lithiation capacity of 1,200–1,400 mAh g1, which
is higher than its theoretical capacity of 1,058 mAh g1 [62]. The excess capacity
can be attributed to the surface Li storage mechanism. The average delithiation
voltage of Cr2O3 is about 1.2 V which is also much lower than that of most of the
other metal oxides. Therefore, Cr2O3 seems a more suitable conversion metal oxide
anode than the others.
Besides nanospheres, a wide variety of nanoscale metal oxides have been
synthesized and tested as anode materials in the last decade. The geometries include
nanotube or nanowire arrays [63–73], nanoflakes [74], nanospindles [75], flower-
like [76, 77], hollow sea-urchin-like nanoparticles [78], and mesoporous structures
[79, 80]. Among them, Chen and coworkers reported the synthesis and electro-
chemical performance of a-Fe2O3 (hematite) nanotubes as anode material [63]. The
reported a-Fe2O3 nanotubes demonstrated very high initial lithiation capacity. The
a-Fe2O3 nanoflakes reported by Reddy et al. demonstrated stable charge–discharge
capacity above 600 mAh g1 up to 80 cycles despite the 70% irreversible capacity
at the first cycle [74]. Recently, Liu et al. reported carbon-coated a-Fe2O3 nanotube
14 Nanostructured Metal Oxides for Li-Ion Batteries 349

Fig. 14.5 (a) SEM image of the a-Fe2O3 nanotube arrays; (b) TEM image and SAED pattern of
a-Fe2O3 nanotubes; (c) cycle performance at C/5 rate for a-Fe2O3 nanotube arrays and carbon
coated a-Fe2O3 nanotube arrays

Fig. 14.6 (a) SEM image of the mesoporous Co3O4 nanoneedles; (b) cycle performance at current
density 150 mAh g1 for mesoporous Co3O4 nanoneedles prepared from different temperatures

arrays with impressive capacity retention as shown in Fig. 14.5 [73]. Lou et al.
reported a mesoporous nanoneedle structure of Co3O4 as shown in Fig. 14.6 [80].
The mesoporous structure could arguably enhance the charge transfer kinetics and
cycle stability, thus resulting in promising electrochemical performance. Besides
350 J. Guo and C. Wang

the aforementioned usual metal oxides, carbon-coated ZnO nanorod arrays were
also investigated by Liu et al. as anode material [81].
Despite some attractive characters of the metal oxide anode materials, there are
still some intrinsic disadvantages of these materials. First, most of them have a
relatively high lithiation potential (above 1 V versus Li/Li+) which will cause low
overall cell voltage. Secondly, most of the reported delithiation capacities of these
metal oxides can only be achieved in a wide potential window, typically from a near-
zero lower limit to an upper limit of 3 V. Obviously not entire such capacities can be
accounted for real-life battery applications. Besides these intrinsic problems, there
are also some formidable technical difficulties including large irreversible capacity
(low coulombic efficiency) and large charge/discharge hysteresis (typically about
1 V) both of which are due to the poor electronic properties. A number of techniques
have been used to enhance the kinetics through improving the electronic properties,
including carbon coating and metal doping [76, 82]. However, substantial improve-
ment has not been achieved. Therefore, metal oxide anode materials still remain as a
concept rather than a realistic choice for Li-ion batteries.

14.3.3 Displacement Metal Oxide Materials

The concept of displacement mechanism is to displace one metal A from a binary


intermetallic AB by lithium reduction. AB proves a host framework for the
inserting and extracting metal A and Li, respectively. Therefore, the intense volume
change by direct alloying can be limited. To date, there has been only one reported
metal oxide, namely Cu2.33V4O11, obeying the displacement mechanism [83].
In this material, Cu inserted in [V4O11]n layered structures. 5.6 Li per
Cu2.33V4O11 can be reversibly inserted and extracted into the layered structure
via displacement reaction. The total capacity of Cu2.33V4O11 is about 270 mAh g1,
However, the lithiation potential is pretty high at 2.5 V which makes it more
suitable as a cathode material.

14.3.4 Tin Dioxides-Based Anode Materials

SnO2 is another well-investigated anode material. Unlike other metal oxide anode
materials, the lithiation/delithiation process of SnO2 is a combination of conversion
and allaying mechanisms which can be described as follows: In the first lithiation
reaction, SnO2 is irreversibly reduced to metallic Sn by Li:

4Liþ þ 4e þ SnO2 ! 2Li2 O þ Sn

Further lithiation obeys the reversible allaying reaction:

xLiþ þ xe þ Sn $ Lix Sn


14 Nanostructured Metal Oxides for Li-Ion Batteries 351

Fig. 14.7 (a) SEM images of thin film of SnO2 based porous spheres; (b) cyclability of SnO2
based composite film of porous spheres

Therefore, SnO2 conversion only takes place as the initial part of the first
lithiation. The following cycles follow the allaying/dealloying reaction between
Sn and Li. Sn has a very high lithiation capacity: in theory, as many as 4.4 Li can be
inserted in 1 Sn. Such high capacity will induce a severe volume change which can
result in particle pulverization, thus causing rapid capacity fade. The advantage of
using SnO2 as anode material instead of Sn is that the initial conversion reaction
will produce nanoscale Sn grains dispersed in the Li2O matrix. The Li2O matrix is
inert in the alloying/dealloying reactions, and can function as the cushion structure
to accommodate the Sn volume change. In addition to the small grain size, Sn is an
electronic conductor that can facilitate high capacity and fast charge/discharge rate.
Therefore, using SnO2 as the anode material can significantly enhance the cycle
stability of the anode. However, it is achieved on the sacrifice of large irreversible
capacity in the first cycle. The reported nanoscale SnO2 anode materials include
nanospheres [84, 85], nanowires [86, 87], flower-like nanoparticles [88], and porous
cage-like nanospheres [89]. Figure 14.7 shows the morphology and performance of
the porous cage-like SnO2 anode materials reported by Yu and coworkers.

14.4 Nano Metal Oxide Cathode Materials

Lithium transition metal oxides and lithium transition metal phosphates represent
the most successful cathode electrode materials for Li-ion batteries. When using
lithium metal anode, some metal oxides such as MnO2 and V2O5 can be used as
cathodes. The power and cycling life of low-potential cathodes have been greatly
improved by simple shrinking the particle size from micro- to nanoscale due to
short diffusion length, high electrode/electrolyte interface, and fast phase transfor-
mation. However, high contact area between the electrode and electrolyte in
nanoscale high-potential cathodes also promote the decomposition of the
352 J. Guo and C. Wang

electrolyte and formation of a SEI layer on the surface of the particles, resulting in
fading of cycle life especially at high temperatures [90, 91]. For Li–Mn–O cathode,
the use of small particles also increases undesirable dissolution of Mn. Surface
coating with a nanolayer of inert oxides (SnO2, Al2O3, MgO, ZrO2) can alleviate
Mn dissolution and SEI formation but also decreases the reaction rate, scarifying
the benefit of nanoscale particle.

14.4.1 Nanoscale Cathode Materials

A great improvement in rate performance and cycling stability of V2O5 nanotubes


[92] and nanowires [93] and LiCoO2 nanowires have been reported [94]. High-
quality single crystalline cubic spinel LiMn2O4 nanowires were synthesized by
Hosono et al. [95] using Na0.44MnO2 nanowires as a self-template. These single
crystalline spinel LiMn2O4 nanowires show high thermal stability and excellent
performance at high rate charge–discharge with excellent cycle stability. Spinel
LiMn2O4 nanorods having an average diameter of 130 nm and length of 1.2 mm
were also synthesized by Cui’s group [96] using a simple solid-state reaction. The
LiMn2O4 nanorod cathodes have a high charge storage capacity at high power rates
compared with commercially available powders.
Nano-LiFePO4 cathodes have been extensively investigated in terms of rate
performance and cycling stability. LiFePO4 has a gravimetric capacity of
170 mAh g1, low cost, high thermal and chemical stability, less reactivity with
electrolytes due to its low potential, and very flat discharge potential, which make it
as promising cathode for hybrid electric vehicle batteries. The lithiation and
delithiation of bulk LiFePO4 involves a phase transformation between LixFePO4
(x ¼ 0.032) and Li1yFePO4 (y ¼ 0.038) at 3.48 V [97, 98] and Li-ion transport
mainly along the (010) direction. However, the miscibility gap and equilibrium
phase transformation potential are highly size dependent. This miscibility gap is
reduced to values of y ¼ 0.12 and x ¼ 0.06 when the particle size decreases to
40 nm, as shown in Fig. 14.8 [99]. Values of y ¼ 0.17 and x ¼ 0.12 are obtained for
34 nm particles [100]. The dramatic shrinking of the miscibility gap at nanoscale
particle size is clearly seen in Figs. 14.8 and 14.9. Stoichiometric 30–40 nm
particles of LiFePO4 exhibit two-phase behavior over 70% of composition range,
while highly defective materials with the same particle size demonstrates solid
solution behavior over the entire composition range [101]. The size dependence of
miscibility gap has been explained by the increasing contribution of elastic energy
induced by the formation of coherent two-phase interphases in small particles. The
coexistence of two crystallographic phases within one particle leads to a phase
boundary energy penalty, due to the difference in lattice parameters of the phases.
This strain-induced energy can destabilize the two-phase coexistence in small
particles, decreasing the energy gain from phase transformation and narrowing
the miscibility gap.
14 Nanostructured Metal Oxides for Li-Ion Batteries 353

4.4
4.2
200 nm
4.0 80 nm
3.8 40 nm
Voltage/V

3.6 3.5 V
3.4
3.2 3.3 V
3.0
2.8
2.6

0.00 0.10 0.70 0.80 0.90 1.00


x in Lix FePO4

Fig. 14.8 OCV curves measured for LixFePO4 at room temperature with various mean particles
sizes of 200, 80, and 40 nm. Reproduced from [99] with permission

Fig. 14.9 Temperature-dependence of LiFePO4 with different particle sizes. Reproduced from
[101] with permission

The particle size of LiFePO4 not only changes the miscibility gap, but also affects
the equilibrium potential and potential hysteresis [102, 103]. The particle size changes
the equilibrium phase diagram of LiFePO4, as demonstrated in Fig. 14.9 [104]. Our
results demonstrate that the equilibrium discharge potential of 40-nm LiFePO4 is
8 mV higher than bulk LiFePO4, as evidenced in Fig. 14.10. The narrowed miscibility
gap and increased equilibrium discharge potential of nano-LiFePO4 greatly reduces
the accommodation energy of phase transformation (Fig. 14.11) and increases the
interface mobility of phase transformation (Fig. 14.12). Therefore, the use of
354 J. Guo and C. Wang

Fig. 14.10 Equilibrium potentials of LiFePO4 with different particle sizes measured by GITT
[103]

Fig. 14.11 Discharge accommodation energies of LiFePO4 with different particle sizes [103]

nanoscale LiFePO4 not only reduces the Li-ion diffusion path but also enhances the
phase transformation kinetics, resulting in a great increase in the rate performance
of LiFePO4.

14.4.2 Nanostructured Cathode Materials

Although the use of nanoscale cathode materials enhances the power of cathode, the
tap density and energy density drop drastically as the particle size decreases [105].
14 Nanostructured Metal Oxides for Li-Ion Batteries 355

Fig. 14.12 Interface


mobilities of (a) bulk and
(b) nano-LiFePO4 obtained
from phase transformation
GITT [103]

So the use of nanoscale materials might lead to high power but could result in very
low volumetric energy storage. To avoid low volumetric energy density and high
reactive surface, but retain the advantage of the nanoscale, attention has turned to
nanostructured cathode materials. Nanostructured cathode for Li-ion battery has
been reviewed by Wang and Cao [106]. The benefit of using nanostructured cathode
materials was evidenced from LiMnO2 cathode. Li can be lithiated/delithiated in
LixMnO2 spinel over the range of 0 < x < 2. Cycling is usually confined to the
range 0 < x < 1 to avoid the transformation of cubic LiMn2O4 into tetragonal
Li2Mn2O4, which leads to a marked loss of capacity. However, the lattice stress
caused by Jahn–Teller distortion can be accommodated more easily in the case of
nanodomain structure. The entire nanodomains can spontaneously change between
cubic and tetragonal structures during lithiation/delithiation. Therefore, capacity
retention is greatly improved compared with the normal bulk LiMnO2 [107].
356 J. Guo and C. Wang

14.5 Nanometal Oxides in Electrolyte

The conventional electrolyte for Li-ion batteries is a Li salt solution based on


electrochemical stable organic solvents. The advantage of the conventional electro-
lyte is the high Li-ion conductivity, but it is undermined by the flammability of the
organic solvent that is a serious safety concern. There are several strategies
attempting to solve this problem, including using aqueous electrolytes, solid ceramic
electrolytes, ionic liquid electrolytes, and solid polymeric electrolytes [108]. To the
best knowledge of the authors, the first three methodologies do not involve applica-
tion of nanostructured metal oxides. Therefore, only the application of nanostructured
metal oxides in polymeric electrolytes will be discussed in this section. In addition to
the potential safety enhancement, solid polymeric electrolytes can also provide
the simplicity of manufacture and a wide variety of battery geometries. Strictly
speaking, the solid polymeric electrolytes only refer to those solvent-free membranes
based on the mixture of Li salts and polymers. Among them, the most attractive
membranes are the ones based on poly(ethylene oxide) (PEO) and a variety of Li salts
such as LiPF6 or LiCF3SO3 [109]. The conduct of Li ions in these membranes is
through the complexation between the ether groups in PEO and the Li ions. However,
the ionic conductivity is poor at room temperature so that the real-life application of
this type of electrolyte has not been achieved.
A very effective method to enhance the conductivity of the PEO electrolyte
is addition of nano-sized metal oxide particles as fillers, such as TiO2, Al2O3,
and SiO2 [110], as shown in Fig. 14.13. A long recognized effect of fillers,
traditionally referred to as plasticizers, is to lower the degree of crystallinity of
the polymer. PEO is a polymer that tends to crystallize, so that the crystalline
domains will block the Li-ion transport in the amorphous regions. The metal oxide
nanoparticle fillers can inhibit the PEO chain to crystallize at lower temperature.
Furthermore, the ionic conductivity enhancement upon addition of metal oxide
nanoparticles was also explained by the space charge theory [111]. Accordingly,
the ionic conductivity could be promoted by the Lewis acid–base interactions
between the surface state of the metal oxide nanoparticles with both the polymer
chains and the anion of the Li salt. This hypothesis could be proved by the close
relationship between the degree of conductivity and the filler surface modification.
It was demonstrated that the sulfate-promoted superacid zirconia (S-ZrO2) fillers
could greatly improve both the ionic conductivity and the Li-ion transference
number due to its high acidic surface state [112].

14.6 Conclusion and Outlook

The state of the art of application of nanoscale and nanostructured metal oxides in
the development of Li-ion batteries was summarized and discussed in this chapter.
A large number of metal oxides can be used as anode and cathode materials in
14 Nanostructured Metal Oxides for Li-Ion Batteries 357

T(°C)
143.35 111.14 84.01 60.20 39.37 20.99 4.65
10−2

10−3

10−4
Conductivity (S cm−1)

10−5

10−6

10−7 ceramic free


TiO2 added, cooling scan
TiO2 added, heating scan
Al2O3 added, cooling scan
10−8
Al2O3 added, 2nd heating scan
Al2O3 added, 1nd heating scan

10−9
2.4 2.6 2.8 3.0 3.2 3.4 3.6
1,000T −1(K−1)

Fig. 14.13 Arrhenius plots of the conductivity of filler-free PEO–LiClO4 and of nanocomposite
PEO–LiClO4, 10 wt% TiO2, or 10 wt% Al2O3 was used based on total weight (PEO:LiClO4 ¼ 8:1
in all cases) [110]

Li-ion battery according to their specific structural and chemical properties. The
advantages of metal oxides with nanodimension are undoubted: nano-sized metal
oxides can enhance the lithiation/delithiation kinetics owing to their smaller size
and larger surface area. Therefore, the rate capacity can be significantly enhanced.
Even new lithiation/delithiation mechanism, conversion reaction, can be enabled.
However, the smaller size and large surface area also bring some downside effects
including low pack density (low energy density) and large surface side reaction.
Future investigation should emphasize to compress these shortcomings without
sacrificing the merit of the nanostructured metal oxides.
One possible means to achieve this goal is to introduce mesoporous (pore size
between 2 and 50 nm) structures into bulk metal oxides to form hierarchical
nanostructures. For instance, the electric conductive additive, or dopant, should be
uniformly dispersed in the body of metal oxide and forms interconnected networks.
The aforementioned RuO2-incorporated anatase TiO2 mesoporous composite [36]
358 J. Guo and C. Wang

could serve a good example of this type of structure. The overall bulk size could lead
to higher packing density, and the uniformly mixed nanoscale metal oxide domains
and nanoscale conductive network can provide superior charge transfer kinetics. In
addition, mesoporous structure can facilitate the contact with electrolyte. Therefore,
this true three-dimensional structure could exhibit better performance than the current
majority zero-dimensional (particle) and one-dimensional (tube or wire) structures.

References

1. Maier, J.: Nanoionics: ion transport and electrochemical storage in confined systems. Nat.
Mater. 4, 805 (2005)
2. Whittingham, M.S., Chianelli, R.R.: Layered compounds and intercalation chemistry: an
example of chemistry and diffusion in solids. J. Chem. Educ. 57, 569 (1980)
3. Thackeray, M.M., Johnson, P.J., De Picciotto, L.A., Bruce, P.G., Goodenough, J.B.: Electro-
chemical extraction of lithium from LiMn2O4. Mater. Res. Bull. 19, 179 (1984)
4. Colbow, K.M., Dahn, J.R., Haering, R.R.: Structure and electrochemistry of the spinel oxides
LiTi2O4 and Li4/3Ti5/3O4. J. Power Sources 26, 397 (1989)
5. Ferg, E., Gummow, R.J., de Kock, A., Thacheray, M.M.: Spinel anodes for lithium-ion
batteries. J. Electrochem. Soc. 141, L147 (1994)
6. Ohzuku, T., Ueda, A., Yamamoto, N.: Zero-strain insertion material of Li[Li1/3Ti5/3]O4 for
rechargeable lithium cells. J. Electrochem. Soc. 142, 1431 (1995)
7. Kim, J., Cho, J.: Spinel Li4Ti5O12 nanowires for high-rate Li-ion intercalation electrode.
Electrochem. Solid State Lett. 10, A81 (2007)
8. Bonino, F., Busani, L., Lazzari, M., Manstretta, M., Rivolta, B., Scrosatti, B.: Anatase as a
cathode material in lithium – organic electrolyte rechargeable batteries. J. Power Sources
6, 261 (1981)
9. Ohzuku, T., Hirai, T.: An electrochromic display based on titanium dioxide. Electrochim.
Acta 27, 1263 (1982)
10. Zachou-Christiansen, B., West, K., Jacobsen, R., Atlung, S.: Lithium insertion in different
TiO2 modifications. Solid State Ionics 28–30, 1176 (1988)
11. Huang, S.Y., Kavan, L., Exnar, I., Gratzel, M.: Rocking chair lithium battery based on
nanocrystalline TiO2 (anatase). J. Electrochem. Soc. 142, L142 (1995)
12. Yang, Z., Choi, D., Kerisit, S., Rosso, K.M., Wang, D., Zhang, J., Graff, G., Liu, J.:
Nanostructures and lithium electrochemical reactivity of lithium titanites and titanium-
oxides: a review. J. Power Sources 192, 588 (2009)
13. Tielens, F., Calatayud, M., Beltran, A., Minot, C., Andres, J.: Lithium insertion and mobility
in the TiO2-antase/titanate structure: a periodic DFT Study. J. Electroanal. Chem. 581, 216
(2005)
14. Gligor, N.M., de Leeuw, S.W.: Lithium diffusion in rutile structured titania. Solid State
Ionics 177, 2741 (2006)
15. Lunell, S., Shashans, A., Ojamae, L., Lindstrom, H., Hagfeldt, A.: Li and Na diffusion in
TiO2 from quantum chemical theory versus electrochemical experiment. J. Am. Chem. Soc.
119, 7374 (1997)
16. Macklin, W.J., Neat, R.J.: Performance of titanium dioxide-based cathodes in a lithium
polymer electrolyte cell. Solid State Ionics 53, 694 (1992)
17. Koudriachova, M.V., Harrison, N.M., de Leeuw, S.W.: Effect of diffusion on lithium
intercalation in titanium dioxide. Phys. Rev. Lett. 86, 1275 (2001)
18. Koudriachova, M.V., Harrison, N.M., de Leeuw, S.W.: NH4Y and HY zeolites as electrolytes
in hydrogen sensors. Solid State Ionics 35, 157 (2003)
14 Nanostructured Metal Oxides for Li-Ion Batteries 359

19. Johnson, O.W.: One-dimensional diffusion of Li in rutile. Phys. Rev. 136, A284 (1964)
20. Gligor, F., de Leeuw, S.W.: Lithium diffusion in rutile structured titania. Solid State Ionics
177, 2741 (2006)
21. Koudriachova, M.V., Harrison, N.M., de Leeuw, S.W.: Effect of diffusion on lithium
intercalation in titanium dioxide. Phys. Rev. Lett. 86, 1275 (2001)
22. Shashans, A., Lunell, S., Bergstroem, R.: Theoretical study of lithium intercalation in rutile
and anatase. Phys. Rev. B 53, 159 (1996)
23. Jiang, C., Honma, I., Kudo, T., Zhou, H.: Nanocrystalline rutile TiO2 electrode for high-
capacity and high-rate lithium storage. Electrochem. Solid State Lett. 10, A127 (2007)
24. Hu, Y.-S., Lorenz, K., Guo, Y.-G., Maier, J.: High lithium electroactivity of nanometer-sized
rutile TiO2. Adv. Mater. 18, 1421 (2006)
25. Baudrin, E., Cassaignon, S., Koesch, M., Jolivet, J.-P., Dupont, L., Tarascon, J.M.: Structural
evolution during the reaction of Li with nano-sized rutile type TiO2 at room temperature.
Electrochem. Commun. 9, 337 (2007)
26. Reddy, M.A., Pralong, V., Varadaraju, U.V., Raveau, B.: Crystalline size constraints on
lithium insertion into brookite TiO2. Electrochem. Solid-State Lett. 11, A132 (2008)
27. Reddy, M.A., Kishore, M.S., Pralong, V., Varadaraju, U.V., Raveau, B.: Lithium intercala-
tion into nanocrystalline brookite TiO2. Electrochem. Solid-State Lett. 10, A29 (2007)
28. Cava, R.J., Murphy, D.W., Zahurak, S., Santoro, A., Roth, R.S.: The crystal structures of the
lithium-inserted metal oxides Li0.5 anatase, LiTi2O4 spinel, and Li2Ti2O4. J. Solid State
Chem. 53, 64 (1984)
29. Sudant, G., Baudrin, E., Larcher, D., Tarascon, J.-M.: Electrochemical lithium reactivity with
nanotextured anatase-type TiO2. J. Mater. Chem. 15, 1263 (2005)
30. Kavan, L., Kalbac, M., Zukalova, M., Exnar, I., Lorenzen, V., Nesper, R., Graetzel, M.:
Lithium storage in nanostructured TiO2 made by hydrothermal growth. Chem. Mater. 16, 477
(2004)
31. Zukalova, M., Kalbac, M., Kavan, L., Exnar, I., Graetzel, M.: Pseudocapacitive lithium
storage in TiO2(B). Chem. Mater. 17, 1248 (2005)
32. Gao, X.P., Lan, Y., Zhu, H.Y., Liu, J.W., Ge, Y.P., Wu, F., Song, D.Y.: Electrochemical
performance of anatase nanotubes converted from protonated titanate hydrate nanotubes.
Electrochem. Solid State Lett. 8, A26 (2005)
33. Armstrong, A.R., Armstrong, G., Canales, J., Garcia, R., Bruce, P.G.: Lithium-ion intercala-
tion into TiO2-B nanowires. Adv. Mater. 17, 862 (2005)
34. Armstrong, G., Armstrong, A.R., Canales, J., Bruce, P.G.: Nanotubes with the TiO2-B
structure. Chem. Comm. 41, 2454 (2005)
35. Armstrong, G., Armstrong, A.R., Bruce, P.G., Reale, P., Scrosati, B.: TiO2(B) nanowires as
an improved anode material for lithium-ion batteries containing LiFePO4 or LiNi0.5Mn1.5O4
or LiNi0.5Mn1.5O4 cathodes and a polymer electrolyte. Adv. Mater. 18, 2597 (2006)
36. Guo, Y.-G., Hu, Y.-S., Sigle, W., Maier, J.: Superior electrode performance of nanostructured
mesoporous TiO2 (Anatase) through efficient hierarchical mixed conducting networks. Adv.
Mater. 19, 2087 (2007)
37. Wang, D., Choi, D., Li, J., Yang, Z., Nie, Z., Kou, R., Wang, C., Saraf, L.V., Zhang, J.,
Aksay, I.A., Liu, J.: Self-assembled TiO2-graphene hybrid nanostructures for enhanced Li-
ion insertion. ACS Nano 3, 907 (2009)
38. Whittingham, M.S.: The role of ternary phases in cathode reactions. J. Electrochem. Soc.
123, 315 (1976)
39. Murphy, D.W., Christian, P.A., DiSalvo, F.J., Carides, J.N., Waszczak, J.V.: Lithium
incorporation by V6O13 and related vanadium (+4, +5) oxide cathode materials. J. Electrochem.
Soc. 128, 2053 (1981)
40. Li, W., Dahn, J.R., Wainwright, D.S.: Rechargeable lithium batteries with aqueous
electrolytes. Science 264, 1115 (1994)
360 J. Guo and C. Wang

41. Zhang, S., Li, Y., Wu, C., Zheng, F., Xie, Y.: Novel flowerlike metastable vanadium dioxide
(B) microanostructures: facile synthesis and application in aqueous lithium ion batteries.
J. Phys. Chem. C 113, 15058 (2009)
42. Choi, N.-S., Kim, J.-S., Yin, R.-Z., Kim, S.-S.: Electrochemical properties of lithium vana-
dium oxcide as an anode material for lithium-ion battery. Mater. Chem. Phys. 116, 603
(2009)
43. Kohler, J., Makihara, H., Uegaito, H., Inoue, H., Toki, M.: LiV3O8: characterization as anode
material for an aqueous rechargeable Li-ion battery system. Electrochim. Acta 46, 59 (2000)
44. Kim, S.-S., Ikuta, H., Wakihara, M.: Synthesis and characterization of MnV2O6 as a high
capacity anode material for a lithium secondary battery. Solid State Ionics 139, 57 (2001)
45. Denis, S., Baudrin, E., Touboul, M., Tarascon, J.-M.: Synthesis and electrochemical
properties of amorphous vanadates of general formula RVO4 (R ¼ In, Cr, Fe, Al, Y) vs Li.
J. Electrochem. Soc. 144, 4099 (1997)
46. Guyomard, D., Sigala, C., Le Gal la Salle, A., Piffard, Y.: New amorphous oxides as high
capacity negative electrodes for lithium batteries: the LixMVO4 (M ¼ Ni, Co, Cd, Zn;
1 < x  8). J. Power Sources 68, 692 (1997)
47. Son, J.T.: Novel electrode material for Li ion battery based on polycrystalline LiNbO3.
Electrochem. Commun. 6, 990 (2004)
48. Han, J.-T., Liu, D.-Q., Song, S.-H., Kim, Y., Goodenough, J.B.: Lithium ion intercalation
performance of niobium oxides: KNb5O13 and K6Nb10.8O30. Chem. Mater. 21, 4753 (2009)
49. Auborn, J.J., Barberio, Y.L.: Lithium intercalation cells without metallic lithium.
J. Electrochem. Soc. 134, 638 (1987)
50. Yang, L.C., Gao, Q.S., Tang, Y., Wu, Y.P., Holze, R.: MoO2 synthesized by reduction of
MoO3 with ethanol vapor as an anode material with good rate capability for the lithium ion
battery. J. Power Sources 179, 357 (2008)
51. Yang, L.C., Gao, Q.S., Zhang, Y.H., Tang, Y., Wu, Y.P.: Tremella-like molybdenum dioxide
consisting of nanosheets as an anode material for lithium ion battery. Electrochem. Commun.
10, 118 (2008)
52. Dillon, A.C., Mahan, A.H., Deshpande, R., Parilla, P.A., Jones, K.M., Lee, S.-H.: Metal oxide
nano-particles for improved electrochromic and lithium-ion battery technologies. Thin Solid
Film 516, 794 (2008)
53. Huang, K., Pan, Q., Yang, F., Ni, S., Wei, X., He, D.: Controllable synthesis of hexagonal
WO3 nanostructures and their application in lithium batteries. J. Phys. D: Appl. Phys. 41,
155417 (2008)
54. Mai, L., Hu, B., Chen, W., Qi, Y., Lao, C., Yang, R., Dai, Y., Lin Wang, Z.: Lithiated MoO3
nanobelts with greatly improved performance for lithium batteries. Adv. Mater. 19, 3712
(2007)
55. Hassan, M.F., Guo, Z.P., Chen, Z., Liu, H.K.: Carbon-coated MoO3 nanobelts as anode
materials for lithium-ion batteries. J. Power Sources 195, 2372 (2010)
56. Lee, S.-H., Kim, Y.-H., Deshpande, R., Parilla, P.A., Whitney, E., Gillaspie, D.T., Jones, K.M.,
Mahan, A.H., Zhang, S., Dillon, A.C.: Reversible lithium-ion insertion in molybdenum oxide
nanoparticles. Adv. Mater. 20, 3627 (2008)
57. Poizot, P., Laruelle, S., Grugeon, S., Dupont, L., Tarascon, J.-M.: Nano-sized transition-metal
oxides as negative-electrode materials for lithium-ion batteries. Nature 407, 496 (2000)
58. Balaya, P., Li, H., Kienle, L., Maier, J.: Fully reversible homogeneous and hetrogeneous Li
storage in RuO2 with high capacity. Adv. Funct. Mater. 13, 621 (2003)
59. Li, H., Balaya, P., Maier, J.: Li-storage via heterogeneous reaction in selected binary metal
fluorides and oxides. J. Electrochem. Soc. 151, A1878 (2004)
60. Larcher, D., Sudant, G., Leriche, J.-B., Chabre, Y., Tarascon, J.-M.: The electrochemical
reduction of Co3O4 in a lithium cell. J. Electrochem. Soc. 149, A234 (2002)
61. Larcher, D., Masquelier, C., Bonnin, D., Chabre, Y., Masson, V., Leriche, J.-B., Tarascon,
J.-M.: Effect of particle size on lithium intercalation into a-Fe2O3. J. Electrochem. Soc.
150, A133 (2003)
14 Nanostructured Metal Oxides for Li-Ion Batteries 361

62. Poizot, P., Laruelle, S., Grugeon, S., Tarascon, J.-M.: Rationalization of the low-potential
reactivity of 3d-metal-based inorganic compounds toward Li. J. Electrochem. Soc. 149,
A1212 (2002)
63. Hu, J., Li, H., Huang, X.: Cr2O3-based anode materials for Li-ion batteries. Electrochem.
Solid State Lett. 8, A66 (2005)
64. Chen, J., Xu, L., Li, W., Gou, X.: a-Fe2O3 nanotubes in gas sensor and lithium-ion battery
applications. Adv. Mater. 17, 582 (2005)
65. He, Y., Huang, L., Cai, J.-S., Zheng, X.-M., Sun, S.-G.: Structure and electrochemical
performance of nanostructured Fe3O4/carbon nanotube composites as anodes for lithium
ion batteries. Electrochim. Acta 55, 1140 (2010)
66. Reddy, A.L.M., Shaijumon, M.M., Gowda, S.R., Ajayan, P.M.: Coaxial MnO2/carbon nano-
tube array electrodes for high-performance lithium batteries. Nano Lett. 9, 1002 (2009)
67. Du, N., Zhang, H., Chen, B., Wu, J., Ma, X., Liu, Z., Zhang, Y., Yang, D., Huang, X., Tu, J.:
Porous Co3O4 nanotubes derived from Co4(CO)12 clusters on carbon nanotube templates:
a highly efficient material for Li-battery applications. Adv. Mater. 19, 4505 (2007)
68. Li, Y., Tan, B., Wu, Y.: Freestanding mesoporous quasi-single-crystalline Co3O4 nanowire
arrays. J. Am. Chem. Soc. 128, 14258 (2006)
69. Li, Y., Tan, B., Wu, Y.: Mesoporous Co3O4 nanowire arrays for lithium ion batteries with
high capacity and rate capability. Nano Lett. 8, 265 (2008)
70. Ryu, J., Kim, S.-W., Kang, K., Park, C.B.: Synthesis of diphenylalanine/cobalt oxide hybrid
nanowires and their application to energy storage. ACS Nano 4, 159 (2010)
71. Jiang, J., Liu, J., Ding, R., Ji, X., Hu, Y., Li, X., Hu, A., Wu, F., Zhu, Z., Huang, X.: Direct
synthesis of CoO porous nanowire arrays on Ti substrate and their application as lithium-ion
battery electrodes. J. Phys. Chem. C 114, 929 (2010)
72. Taberna, P.L., Mitra, S., Poizot, P., Simon, P., Tarascon, J.-M.: High rate-capabilities
Fe3O4-based Cu nano-architectured electrodes for lithium-ion battery applications. Nat.
Mater. 5, 567 (2006)
73. Wang, L., Yu, Y., Chen, P.C., Zhang, D.W., Chen, C.H.: Electrospinning synthesis of C/
Fe3O4 composite nanofibers and their application for high performance lithium-ion batteries.
J. Power Sources 183, 717 (2008)
74. Liu, J., Li, Y., Fan, H., Zhu, Z., Jiang, J., Ding, R., Hu, Y., Huang, X.: Iron oxide-based
nanotube arrays derived from sacrificial template-accelerated hydrolysis: large-area design
and reversible lithium storage. Chem. Mater. 22, 212 (2010)
75. Reddy, M.V., Yu, T., Sow, C.-H., Shen, Z.X., Lim, C.T., Subba Rao, G.V., Chowdari, B.V.
R.: a-Fe2O3 nanoflakes as an anode material for Li-ion batteries. Adv. Funct. Mater. 17, 2792
(2007)
76. Zhang, W.-M., Wu, X.-L., Hu, J.-S., Guo, Y.-G., Wan, L.-J.: Carbon coated Fe3O4
nanospindles as a superior anode material for lithium-ion batteries. Adv. Funct. Mater. 18,
3941 (2008)
77. Xiang, J.Y., Tu, J.P., Yuan, Y.F., Wang, X.L., Huang, X.H., Zeng, Z.Y.: Electrochemical
investigation on nanoflower-like CuO/Ni composite film as anode for lithium ion batteries.
Electrochim. Acta 54, 1160 (2009)
78. Xiang, J.Y., Tu, J.P., Zhang, L., Zhou, Y., Wang, X.L., Shi, S.J.: Self-assembled synthesis of
hierarchical nanostructured CuO with various morphologies and their application as anodes
for lithium ion batteries. J. Power Sources 195, 313 (2010)
79. Li, B., Rong, G., Xie, Y., Huang, L., Feng, C.: Low-temperature synthesis of a-MnO2 hollow
urchins and their application in rechargeable Li+ batteries. Inorg. Chem. 45, 6404 (2006)
80. Shaju, K.M., Jiao, F., Debart, A., Bruce, P.G.: Mesoporous and nanowire Co3O4 as negative
electrodes for rechargeable lithium batteries. Phys. Chem. Chem. Phys. 9, 1837 (2007)
81. Lou, X.W., Deng, D., Lee, J.Y., Archer, L.A.: Thermal formation of mesoporous single-
crystal Co3O4 nano-needles and their lithium storage properties. J. Mater. Chem. 18, 4397
(2008)
362 J. Guo and C. Wang

82. Liu, J., Li, Y., Ding, R., Jiang, J., Hu, Y., Ji, X., Chi, Q., Zhu, Z., Huang, X.: Carbon/ZnO
nanorod array electrode with significantly improved lithium storage capability. J. Phys.
Chem. C 113, 5336 (2009)
83. Needham, S.A., Wang, G.X., Konstantinov, K., Tournayre, Y., Lao, Z., Liu, H.K.: Electro-
chemical performance of Co3O4-C composite anode materials. Electrochem. Solid State Lett.
9, A315 (2006)
84. Morcrette, M., Rozier, P., Dupont, L., Mugnier, E., Sannier, L., Galy, J., Tarascon, J.M.:
A reversible copper extrusion-insertion electrode for rechargeable Li batteries. Nat. Mater.
2, 755 (2003)
85. Kim, C., Noh, M., Choi, M., Cho, J., Park, B.: Critical size of a nano SnO2 electrode for
Li-secondary battery. Chem. Mater. 17, 3297 (2005)
86. Lou, X.W., Chen, J.S., Chen, P., Archer, L.A.: One-pot synthesis of carbon-coated SnO2
nanocolloids with improved reversible lithium storage properties. Chem. Mater. 21, 2868
(2009)
87. Li, N., Martin, C.R.: A high-rate, high-capacity, nanostructured Sn-based anode prepared
using sol-gel template synthesis. J. Electrochem. Soc. 148, A164 (2001)
88. Meduri, P., Pendyala, C., Kumar, V., Sumanasekera, G.U., Sunkara, M.K.: Hybrid tin oxide
nanowires as stable and high capacity anodes for Li-ion batteries. Nano Lett. 9, 612 (2009)
89. Jiang, L.-Y., Wu, X.-L., Guo, Y.-G., Wan, L.-J.: SnO2-based hierarchical nanomicros-
tructures: facile synthesis and their applications in gas sensors and lithium-ion batteries.
J. Phys. Chem. C 113, 14213 (2009)
90. Yu, Y., Chen, C.-H., Shi, Y.: A tin-based amorphous oxide composite with a porous,
spherical, multideck-cage morphology as a highly reversible anode material for lithium-ion
batteries. Adv. Mater. 19, 993 (2007)
91. Liu, H.K., Wang, G.X., Guo, Z.P., Wang, J.Z., Konstantinov, K.: Nanomaterials for lithium-
ion rechargeable batteries. J. Nanosci. Nanotechnol. 6, 1 (2006)
92. Ye, S.H., Lv, J.Y., Gao, W.P., Wu, F., Song, D.Y.: Synthesis and electrochemical properties
of LiMn2O4 spinel phase with nanostructure. Electrochim. Acta 49, 1623 (2004)
93. Nordliner, S., Edstrom, K., Gustafsson, T.: The performance of vanadium oxide nanorolls
as cathode material in a rechargeable lithium battery. Electrochem. Solid State Lett. 4, A129
(2001)
94. Patrissi, C.J., Martin, C.R.: Sol-gel-based template synthesis and Li-insertion rate perfor-
mance of nanostructured vanadium pentoxide. J. Electrochem. Soc. 146, 3176 (1999)
95. Jiao, F., Shaju, K.M., Bruce, P.G.: Synthesis of nanowire and mesoporous low-temperature
LiCoO2 by a post-templating reaction. Angew. Chem. Int. Ed. 44, 6550 (2005)
96. Hosono, E., Kudo, T., Honma, I., Matsuda, H., Zhou, H.: Synthesis of single crystalline spinel
LiMn2O4 nanowires for a lithium ion battery with high power density. Nano Lett. 9, 1045
(2009)
97. Kim, D.K., Muralidharan, P., Lee, H.W., Ruffo, R., Yang, Y., Chan, C.K., Peng, H., Hu, R.,
Huggins, A., Cui, Y.: Spinel LiMn2O4 nanorods as lithium ion battery cathodes. Nano Lett.
8, 3948 (2008)
98. Yamada, A., Koizumi, H., Sonoyama, N., Kanno, R.: Phase change in LixFePO4.
Electrochem. Solid State Lett. 8, A409 (2005)
99. Yamada, A., Koizumi, H., Nishimura, S.I., Sonoyama, N., Kanno, R., Yonemura, M.,
Nakamura, T., Kobayashi, Y.: Room-temperature miscibility gap in LixFePO4. Nat. Mater.
5, 357 (2006)
100. Kobayashi, G., Nishimura, S.I., Park, M.S., Kanno, R., Yashima, M., Ida, T., Yamada, A.:
Isolation of solid solution phases in size-controlled LixFePO4 at room temperature. Adv.
Funct. Mater. 19, 395 (2009)
101. Meethong, N., Huang, H.Y.S., Carter, W.C., Chiang, Y.M.: Size-dependent lithium
miscibility gap in nanoscale Li1-xFePO4. Electrochem. Solid State Lett. 10, A134 (2007)
14 Nanostructured Metal Oxides for Li-Ion Batteries 363

102. Meethong, N., Kao, Y.H., Tang, M., Huang, H.-Y., Carter, W.C., Chiang, Y.M.: Electro-
chemically induced phase transformation in nanoscale olivines Li1-xMPO4 (M ¼ Fe, Mn).
Chem. Mater. 20, 6189 (2008)
103. Lee, K.T., Kan, W.H., Nazar, L.: Proof of intercrystallite ionic transport in LiMPO4
electrodes (M ¼ Fe, Mn). J. Am. Chem. Soc. 131, 6044 (2009)
104. Gibot, P., Casas-Cabanas, M., Laffont, L., Levasseur, S., Carlach, P., Hamelet, S., Tarascon,
J.-M., Masquelier, C.: Room-temperature single-phase Li insertion/extraction in nanoscale
LixFePO4. Nat. Mater. 7, 741 (2008)
105. Zhu, Y., Wang, C.: Galvanostatic intermittent titration technique for phase-transformation
electrodes. J. Phys. Chem. C 114, 2830 (2010)
106. Chen, Z., Dahn, J.R.: Reducing carbon in LiFePO4/C composite electrodes to maximize
specific energy, volumetric energy, and tap density. J. Electrochem. Soc. 149, A1184 (2002)
107. Wang, Y., Cao, G.: Developments in nanostructured cathode materials for high-performance
lithium-ion batteries. Adv. Mater. 20, 2251 (2008)
108. Robertson, A.D., Armstrong, A.R., Bruce, P.G.: Layered LixMn1-yCoyO2 intercalation
electrodes – influence of ion exchange on capacity and structure upon cycling. Chem.
Mater. 13, 2380 (2001)
109. Goodenough, J.B., Kim, Y.: Challenges for rechargeable Li batteries. Chem. Mater. 22, 587
(2010)
110. Lightfoot, P., Metha, M.A., Bruce, P.G.: Crystal structure of the polymer electrolyte poly
(ethylene oxide)3: LiCF3SO3. Science 262, 883 (1993)
111. Croce, F., Appetecchi, G.B., Persi, L., Scrosati, B.: Nanocomposite polymer electrolytes for
lithium batteries. Nature 394, 456 (1998)
112. Maier, J.: Ionic conduction in space charge regions. Prog. Solid State Chem. 23, 171 (1995)
113. Croce, F., Settimi, L., Scrosati, B.: Superacid ZrO2-added, composite polymer electrolytes
with improved transport properties. Electrochem. Commun. 8, 364 (2006)
Index

A CEMs. See Correlated electron materials


AFM. See Atomic force microscopy Ceria, 242–256, 265
ALD. See Atomic layer deposition CFs. See Conductive filaments
Anatase, 104, 105, 161, 167, 199, 256–258, Charge and bonding models, 95–97
260, 261, 265, 279, 342–345, 357 Charge neutrality level (CNL), 128, 139–142
Angular-resolved photoelectron spectroscopy, Charge order, 82, 89, 100, 210
309, 310 Charge transfer, 89, 90, 101, 106, 107, 157,
Anion, 39, 43, 131, 138, 142, 241–243, 356 210, 339–341, 344, 349, 358
Anodization, 160–161, 169 Chromium oxide, 261–266
Atomic force microscopy (AFM), 13, 14, 16, Clapeyron equation, 14
17, 71, 209, 225, 226, 231, 232, 307, CNL. See Charge neutrality level
308, 312, 318 Colossal magnetoresistance, 3, 4, 69, 210
Atomic layer deposition (ALD), 176, 179, Conductive filaments (CFs), 306–312, 314,
180, 183, 190, 191, 193, 194, 311, 315, 319, 320, 322, 323
316, 319 Confinement model, 91–93
Conformal-evaporated-film-by-rotation, 176,
193, 194
B Conversion reaction, 338, 347, 351, 357
Beetles, 183–186, 188, 189, 191, 195, 197, Core/shell structure, 54, 281, 282
200, 201 Correlated electron materials (CEMs), 3–5,
Bioinspiration, 186 8, 11, 13, 18–20, 99
Biotemplating, 176, 190–201 Crossbar array, 318, 327
Bipolar switching, 305, 307, 310, 312–316,
322, 324
Block copolymer lithography, 163, 164 D
Bound carrier excitation, 101–102 Defects, 13, 23–29, 31, 37–61, 88, 97,
Butterflies, 183, 184, 186–188, 191, 193–195, 103, 104, 110, 139–142, 161, 258,
200, 201 265, 306, 309
Density functional theory (DFT), 25, 43, 137,
149, 205, 209–212, 214, 217, 219, 224,
C 225, 228, 231, 232
Carrier concentration, 29, 44–46, 88, 131, 132, DFT+U, 233
134, 136, 137 Dielectric breakdown, 305, 309–311, 316
Catalysis, 108, 158, 175, 210, 211, 288 3-Dimensional fabrication, 179
Cation, 23, 31, 40, 58, 74, 138, 142, 197, 3-D integration, 325, 327
241, 242 Domain dynamics, 4, 8–11, 83
Cell selector, 305, 319, 323–325, 327 Domain organization, 6–8, 200

J. Wu et al. (eds.), Functional Metal Oxide Nanostructures, Springer Series 365


in Materials Science 149, DOI 10.1007/978-1-4419-9931-3,
# Springer Science+Business Media, LLC 2012
366 Index

Domain structure, 3, 6, 19, 75, 76, 98–100, Iridescence, 200


159, 355 Isobaric, 12, 13
Doping, 5, 24, 27, 28, 42, 47, 48, 60, 71, 74, 81, Isothermal, 12–14, 249
84, 88, 104, 142, 148, 168, 347, 350
4D transition metal oxides, 74, 148, 150, 151
J
Jahn-Teller distortion, 273, 355
E Joule-heating, 9, 306, 310, 312, 316
EELS. See Electron energy loss spectroscopy
Electrical double layer, 285
Electrical properties, 11, 17, 45, 131 K
Electrochemical deposition, 166, 289 Kerr effect, 52
Electrofilling, 160
Electroforming, 304, 307, 309–312, 316
Electrolyte stability, 283, 352 L
Electron energy loss spectroscopy (EELS), Latent heat, 14
209, 247, 259, 263–265, 309, 312 LEED. See Low-energy electron diffraction
Electronic inhomogeneity, 70 Lithium intercalation, 279, 338, 339, 341, 346
Electronic phase transition, 82 Lithium-ion, 270, 280–283, 293, 337
Electron microscopy, 79, 185, 187, 274 Lithium-ion batteries, 270, 280–283, 293
Electron transport, 73, 167, 271, 339, 344, 347 Lithium iron phosphate, 283–285, 352–355
Ellipsometry, 159, 160 Lithium metal oxide cathode materials,
Epitaxial growth, 210, 212, 213, 227 351–352
Lithium titanium oxide, 339–341
Low-dimensional materials, 69
F Low-energy electron diffraction (LEED),
Fermi-level pinning, 25–27, 30, 31 209, 214, 216, 217, 219–221,
Ferromagnetism, 45, 53–61, 88, 152, 211, 261 224–228, 231, 232
Finite length scale effects, 91, 97, 110
Fuel cell, 211, 243, 244
M
Macrostructured oxides, 175
G Magnéli phases, 242, 256–261
Galvanostatic intermittent titration, 287, 288 Manganites, 4, 20, 69, 70, 72, 74, 80–83,
88, 99, 100, 210, 211
Mesoporous structures, 275, 348
H Metal, 4, 26, 37, 69, 87, 127, 147, 157, 175,
Hall effect, 46, 129, 131, 133, 136 209, 241, 272, 303, 337
Heterojunction, 158 Metal-insulator-metal (MIM), 304, 306, 313
High energy materials, 101, 348 Metal-insulator transition, 9, 17, 69–84, 94,
High power, 26, 272, 285, 352, 355 147, 148, 211, 306, 325
Hybrid functionals, 25–27, 30, 31, 214, 222 Metal-organic frameworks (MOFs), 98,
Hydrogen binding, 108 108, 109
Hydrothermal growth, 167–169 Metal oxide, 37–47, 51–61, 83, 87–110,
127–142, 148, 151, 157, 161, 167,
175, 179, 182, 210, 211, 213, 232,
I 233, 241–266, 280–283, 285–293,
Impurities, 23, 24, 27–32, 47–49, 53, 57, 80, 303–327, 337–358
139–141, 147, 323 Metal oxide composite, 286, 291–293
Infrared reflectivity, 129, 136 Microtransfer molding, 162
Inhomogeneous media, 93–95 MIM. See Metal-insulator-metal
In-situ characterization, 159 Miscibility gap, 352, 353
Intercalation reaction, 341 Modulation structure, 242, 261–265
Index 367

MOFs. See Metal-organic frameworks Phase separation, 1, 4, 7, 8, 11, 19, 69, 70,
Mott transition, 11, 74–80, 83, 98–99 75, 76, 80–84
Multilayers, 183, 187, 213, 225 Phase transformation, 13, 275, 278, 339,
351–355
Photoluminescence, 47–50, 103
N Photonic band gap, 177–179, 181, 183, 197,
Nanobelts, 29, 242, 261–265, 275, 288 199–201
Nanocavities, 242, 258, 259, 265 Photonic crystals, 175–201
Nanoimprint lithography, 161, 318 Photovoltaics, 157, 158, 167
Nanomaterials, 37, 38, 60, 88, 89, 91–93, Plasmon, 46–47, 52, 95
95–97, 261, 271, 273, 274, 286, 293 Polar oxide thin film, 90, 98, 104–107, 110
Nanoparticles, 37, 40, 41, 49–56, 59, Polymer electrolyte, 356
91–93, 95–98, 103–107, 110, Polysynthetic twins, 16
159–161, 211, 244, 272, 274, 279, Poly(vinyl alcohol), 161
282, 283, 289, 291, 341, 345, 346, Porous materials, 89, 108, 271, 274, 291
348, 351, 356 Power consumption, 303, 316, 318, 322
Nanostructure, 3–20, 23, 26, 29, 37–61,
71–73, 91, 104, 127, 128, 142,
157–169, 195, 209–233, 241, 245–247, Q
256, 261, 269–294, 316, 337–358 Quantum size effects, 91, 95, 98, 103–104,
Nanotubes, 88, 90, 94, 102, 158, 160, 161, 110, 158
241, 280, 288, 291, 344, 348, 349, 352
Nanowires, 13, 16, 29, 83, 127, 129, 133,
160, 161, 167, 211, 241, 271, 272, R
274, 277, 279–281, 283, 288, 316, Raman spectroscopy, 16, 17, 51, 228
343, 344, 351, 352 Rechargeable batteries, 270, 275, 337
Native point defects, 23, 26, 29, 31, 43, 60 Redox, 253, 256, 286, 338, 344, 347
Non-stoichiometric, 41, 231, 241, 242, 256, 264 Reliability, 26, 73, 212, 303, 316, 322, 323
Non-volatile memory, 322, 327 Resistance random access memory
(RRAM), 303
Resistance ratio, 305, 308, 318, 320, 323
O Resistance switching, 88, 305
One-dimensional, 47, 183, 241–266, 271, 274, Resistive switching, 303–310, 313, 315–316,
275, 283, 291, 358 318, 324, 325, 327
One-dimensional nano-wires, 160 Ruthenates, 70, 74, 80, 83, 150
Optical absorption, 6, 106, 107, 129, 133–136, Rutile, 5, 15, 16, 29, 30, 41, 98, 104, 105,
259 138, 167, 168, 256, 261, 279, 280,
Optical properties, 11, 23, 46, 47, 51, 52, 87–110, 309, 341–343, 345, 346
176, 193, 195, 197, 200, 256, 257
Optical spectroscopy, 88–90, 105, 110
Orbital order, 80, 101, 111 S
Oxide nanolayers, 209, 213 Scalability, 303, 316, 318, 323, 327
Oxide semiconductors, 23–32, 52, 127–142, 181 Scanning tunneling microscopy (STM), 209
Oxygen deficient, 42, 43, 45, 57, 60, 241–266, SEI. See Solid electrolyte interface
311, 325 Self-assembly, 71, 159–160, 164, 169, 175,
Oxygen migration, 309–313, 316, 323, 327 178, 180–182, 190, 316
Semiconductor, 18, 23–32, 47, 49, 52, 75,
103, 127–142, 147, 148, 151, 157,
P 158, 179, 181, 241, 242, 256, 260,
Peierls transition, 11, 13 305, 316
Percolation transition, 148 Sol filling, 163
Phase-field modeling, 9, 10, 14 Sol gel, 161, 176, 182, 190, 194–201, 244,
Phase inhomogeneity, 3–8, 11, 19, 20 261, 264, 278, 289, 291, 316
368 Index

Sol-gel chemistry, 176, 190, 194–198, 201 Transition metal oxides, 43, 87–110,
Solid electrolyte interface (SEI), 340 127–142, 148, 151, 210, 211, 256,
Solution based process, 169 261, 306, 347, 351
Space charge theory, 356 Transmission electron microscopy, 6, 41,
Spin polarization, 57–58, 261 188, 245, 305, 306, 309
STM. See Scanning tunneling microscopy Transparent conducting oxides, 161, 168
Strain energy, 6, 8–10, 15, 232 Transparent conductors, 127
Strain engineering, 9, 11, 77 Two-dimensional electron gas, 133, 148
Structural phase transition, 5, 17, 19
Sub-stoichiometric, 242, 256–265, 310
Superelasticity, 5, 13–15, 20 U
Surface coating, 271, 285, 352 Uniformity, 194, 316, 323
Surface effects, 45, 142 Unintentional n-type conductivity, 23,
Surface electron accumulation layer, 134, 27–29, 31
135, 138 Unipolar switching, 304, 307, 312, 314–316,
Surface plasmon, 52, 95 322, 324
Surface redox reaction, 286
Surfaces, 7, 37, 74, 88, 127–142, 149, 157,
182, 209, 245, 270, 310, 339 V
Switching speed, 305, 320, 322, 323 Vanadium oxide nanoscrolls, 98,
101–103, 110

T
Thermoelectrics, 108, 241 W
Thin films, 3, 23, 41, 71, 88, 128, 148, 160, Water-gas shifting reaction, 242
210, 278, 312, 351 Water soluble template, 161
Three-dimensional nano-structures, 276, 293 Wide-band-gap semiconductors, 49
Threshold resistivity, 12, 13
Tin dioxide, 29, 195, 339, 350–351
Titania, 30, 157, 163, 166, 168, 181–183, 193, X
196–201, 258 X-ray absorption near edge structure
Titanium dioxide (TiO2), 157, 178, 181, 183, (XANES), 248, 250, 253, 308
194, 195, 200, 256, 341 X-ray diffraction, 6, 93, 248
Titanium oxide, 241, 256, 279, 340, 341 X-ray photoelectron spectroscopy (XPS),
Top-down and bottom-up fabrication, 129–131, 133, 134, 136, 138, 209, 214,
176, 179 309–311, 314

You might also like