Вы находитесь на странице: 1из 241

Nanostructured Materials for Electrochemical

Energy Production and Storage


Nanostructure Science and Technology
Series Editor: David J. Lockwood, FRSC
National Research Council of Canada
Ottawa, Ontario, Canada

For other titles published in this series, go to


www.springer.com/series/6331
Edson Roberto Leite
Editor

Nanostructured Materials
for Electrochemical Energy
Production and Storage

123
Edson Roberto Leite
Universidade Federal de Sao Carlos
Centro de Ciencias Exatas e de Tecnologia
Caixa Postal 676
Sao Carlos-SP
Brazil
derl@power.ufscar.br

Series Editor
David J. Lockwood
National Research Council of Canada
Ottawa, Ontario
Canada

ISBN: 978-0-387-49322-0 e-ISBN: 978-0-387-49323-7


DOI: 10.1007/978-0-387-49323-7

Library of Congress Control Number: 2008 941657

c Springer Science+Business Media, LLC 2009


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.com
Preface

The major problem facing new energy conversion and storage technologies remains
device efficiency. Projects based on nanostructured materials can yield improved
performance in devices involving electrochemical reactions and heterogeneous
catalysis, such as fuel and solar cells, batteries, etc. Nanoscale structures dramati-
cally alter the surface reaction rates and electrical transport throughout the material,
causing a dramatic improvement in energy storage, conversion, and generation.
Furthermore, the design of nanoscale materials to be applied in alternative energy
devices is a predictable way to develop a wide range of new technologies for a more
sustainable future. Therefore, the goal of this book is to present basic fundamentals
and the most relevant properties of nanostructured materials in order to improve
alternative energy devices.
This book begins with a chapter by Gratzel summarizing the use of mesoscopic
thin films and hybrid materials in the development of new kinds of regenerative
photoelectrochemical devices. Applications include high-efficiency solar cells.
In chapter two, Ribeiro and Leite describe assembly and properties of nanopar-
ticles. The chapter presents a review on the properties and main features of
nanoscale materials, emphasizing the dependence of key properties on size for
energy purposes. A general description is also given of nanoparticle synthesiza-
tion methods (mainly oxides), focusing on advances in tailoring controlled shape
nanostructures.
Bueno and Gabrielli present the basic principles of nanotechnology in general
and integrate fundamental electrochemistry with nanostructured materials in partic-
ular. The main focus of this chapter, therefore, is on novel strategies that exploit
nanoscale architectures to enhance the efficiency of alternative energy conversion
and storage devices and on the basic principles of electrochemistry governing the
effects of nanoscale structures on electrodes and electrolytes.
Heinzel and Konig summarize the impact of nanostructured materials on fuel
cell technology, mainly in the area of polymer electrolyte membrane fuel cells.
This chapter illustrates how nanostructured materials can modify component
performance such as electrocatalyst materials and membrane.

v
vi Preface

Dong and Dunn describe advances resulting from the use of sol-gel technology
on energy storage devices. This chapter reviews the importance of aerogel nanoar-
chitecture in achieving high performance electrochemical properties. Results ob-
tained with vanadium oxide aerogels are highlighted as these materials exhibit a
number of desirable characteristics for secondary lithium batteries.
A chapter focusing on the use of nanocomposites in electrochemical devices is
presented by Schoonman, Zavyalov, and Pivkina. A wide range of metal (metal ox-
ide)/polymer nanocomposites has been synthesized using Al, Sn, Zn, Pd, and Ti as
a metal source and poly-para-xylylene (PPX) as a polymeric matrix. The properties
of the nanocomposites were studied by comparing structure, morphology, electrical
properties, oxidation kinetics, and electrochemical parameters.
As the rapid development of nanostructured materials continues, this book illus-
trates the impact of this class of materials on performance improvements of alter-
native energy devices, particularly those based on electrochemical processes. The
authors make a powerful case for nanomaterials and nanotechnology as a way to
transform such alternative energy sources into significant contributors to the future
global energy mix.
Contents

Recent Applications of Nanoscale Materials: Solar Cells . . . . . . . . . . . . . . . 1


Michael Gratzel

Assembly and Properties of Nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . 33


Caue Ribeiro and Edson R. Leite

Electrochemistry, Nanomaterials, and Nanostructures . . . . . . . . . . . . . . . . 81


Paulo Roberto Bueno and Claude Gabrielli

Nanotechnology for Fuel Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151


Angelika Heinzel and Uwe Konig

Vanadium Oxide Aerogels: Enhanced Energy Storage in Nanostructured


Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
Winny Dong and Bruce Dunn

Nanostructured Composites: Structure, Properties, and Applications


in Electrochemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
Joop Schoonman, Sergey Zavyalov, and Alla Pivkina

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219

vii
Contributors

Paulo Roberto Bueno, Instituto de Qumica, Departamento de Fsico-Qumica,


Universidade Estadual Paulista, C. Postal 355, 14801-907, Araraquara, Sao Paulo,
Brazil, prbueno@mail.iq.unesp.br
Winny Dong, Chemical and Materials Engineering, California State Polytechnic
University, Pomona, CA 91768
Bruce Dunn, Department of Materials Science and Engineering, University of
California, Los Angeles, CA 90095, USA, bdunn@ucla.edu
Claude Gabrielli, UPR 15 du CNRS, Physique des Liquides et Electrochimie,
Universite Pierre et Marie Curie, 4 place Jussieu, 75252 Paris, France
Michael Gratzel, Laboratory of Photonics and Interfaces, Ecole Polytechnique
Federale de Lausanne, CH-1015, Lausanne, Switzerland, michael.graetzel@epfl.ch
Angelika Heinzel Centre for Fuel cell Technology (ZBT gGmbH), Carl-Benz-Str.
201, 47057 Duisburg, Germany
Uwe Konig Centre for Fuel cell Technology (ZBT gGmbH), Carl-Benz-Str. 201,
47057 Duisburg, Germany
Edson R. Leite, LIEC Universidade Federal de Sao Carlos, Departamento de
Qumica, Rod. Washington Luiz, km 235 13565-905, Sao Carlos, SP, Brazil
Alla Pivkina, Semenov Institute of Physical Chemistry, Russian Academy of
Science, Kosygin st. 4, 119991, Moscow, Russia
Caue Ribeiro, EMBRAPA Instrumentaca o Agropecuaria, Rua XV de Novembro,
1452 13560-970, CP 741, Sao Carlos, SP, Brazil
Joop Schoonman, Delft University of Technology, Delft Institute for Sustainable
Energy, P.O. Box 5045, 2600 GA Delft, The Netherlands, J.Schoonman@tudelft.nl
Sergey Zavyalov, Karpov Institute of Physical Chemistry, Vorontsovo Pole, 10,
103064 Moscow, Russia

ix
Recent Applications of Nanoscale Materials:
Solar Cells

Michael Gratzel

Abstract Photovoltaic cells have been dominated so far by solid state p-n junc-
tion devices made from silicon or gallium arsenide wavers or thin film embod-
iments based on amorphous silicon, CdTe and copper indium gallium diselenide
(CIGS) profiting from the experience and material availability of the semiconduc-
tor industry. Recently there has been a surge of interest for devices that are based
on nanoscale inorganic or organic semiconductors commonly referred to as bulk
junctions due to their interconnected three-dimensional structure. The present chap-
ter describes the state of the art of the academic and industrial development of
nanostructured solar cells, with emphasis in the development of the dye-sensitized
nanocristalline solar cell.

1 Introduction

Photovoltaic cells have been dominated so far by solid state pn junction devices
made from silicon or gallium arsenide wavers or thin film embodiments based on
amorphous silicon, CdTe and copper indium gallium diselenide (CIGS) profiting
from the experience and material availability of the semiconductor industry. Re-
cently there has been a surge of interest for devices that are based on nanoscale
inorganic or organic semiconductors commonly referred to as bulk junctions
due to their interconnected three-dimensional structure. Research in this field has
gained significant momentum with hundreds of groups being active to develop
new mesoscopic solar cell variants and improve their performance. These devices
are formed from junctions of, for example, nanocrystalline, inorganic oxides and
chalcogenides or fullerenes with organic electrolytes, ionic liquids, or inorganic and
organic hole conductors and conducting polymers, offering the prospect of very

M. Gratzel
Laboratory of Photonics and Interfaces, Ecole Polytechnique Federale de Lausanne, CH-1015,
Lausanne, Switzerland
e-mail: michael.graetzel@epfl.ch
E.R. Leite (ed.), Nanostructured Materials for Electrochemical Energy Production 1
and Storage, Nanostructure Science and Technology, DOI 10.1007/978-0-387-49323-7 1,
c Springer Science+Business Media LLC 2009
2 M. Gratzel

low cost fabrication without expensive and energy-intensive high-temperature and


high-vacuum processes. They are compatible with flexible substrates and a variety
of embodiments and appearances to facilitate market entry, both for domestic de-
vices and in architectural or decorative applications. Thus, it appears possible now
to depart completely from the concept of classical, flat pn junction cells to replace
them with interpenetrating network junctions. The nanoscale morphology produces
an interface with a huge area, endowing these systems with extraordinary opto-
electronic properties. Despite their disordered structure these novel solar cells have
shown strikingly high conversion efficiencies competing with those of conventional
devices while at the same time offering advantages of ease of production, lower cost
and shorter energy payback times.
The first embodiment and prototype of this family of devices is the dye-sensitized
solar cell (DSC), invented in the authors laboratory at the Ecole Polytechnique
Federale de Lausanne [13]. At the heart of the DSC is a mesoscopic film com-
posed by, for example, nanoparticles, nanorods or nanotubes of a wide band gap
semiconductor oxide (typically ZnO, SnO2 or TiO2 ) that is covered by a monolayer
of sensitizer. During the illumination of the cell the photo-excited dye molecules
adsorbed on the nanoparticle surface inject electrons into the conduction band of
the oxide. The sensitizer is regenerated by hole injection into a redox electrolyte
or a solid-state p-type conductor. Although many organic or inorganic sensitizers,
including semiconductor quantum dots are known to date and their performance
is rapidly improving, ruthenium poly-pyridyl complexes have maintained so far a
lead as the most efficient and stable sensitizers. Published first in 1993 [4], cis-
RuL2 (SCN)2 (L = 2, 2 -bipyridyl-4,4 -dicarboxylic acid), coded N3 or N719 for
the fully protonated or half protonated form, respectively, has become the paradigm
of a charge transfer sensitizer capturing a large domain of the visible spectrum. The
structure of the N3 dye is shown in Fig. 1.

Fig. 1 Structure of the N3 dye cis RuL2 (SCN)2 (L = 2,2-bipyridyl-4,4 dicarbo-xylate). See
Color Plates
Recent Applications of Nanoscale Materials: Solar Cells 3

Incident photon-to-electron conversion efficiencies (IPCEs) attaining almost


unity have been obtained with this sensitizer in the wavelength range near its ab-
sorption maximum corresponding to quantum yields for electric current generation
near 100% when light losses due to the conducting glass current collector are taken
into account [48].
The DSC is the only solar cell that uses a molecule to absorb photons and convert
them into electric charges without the need of excitonic transport. It also is the only
photovoltaic device that separates the two functions of solar light harvesting and
charge carrier transport, whereas conventional cells perform both operations simul-
taneously resulting in stringent demands on the purity of materials and much higher
production costs. In the DSC, the molecular sensitizer or semiconductor quantum
dot is placed at the interface between an electron and hole conducting material.
Upon photo-excitation the negative charge carriers are injected by the excited dye
into an electron (n) conductor and the positive charges in a hole (p) conductor or
an electrolyte. Therefore, the electrons and holes are charges that move in different
phases to the front and back contacts of the photocell. Their recombination occurs
across the interface separating the electron and the hole conductor materials. The
open circuit photo-voltage developed by the cell corresponds to the difference of the
quasi-Fermi level of the electrons and holes in their respective transport medium.
The great advantage of this cell configuration is that the solar energy conver-
sion process involves only majority charge carriers. Electron and holes are gener-
ated in different phases and their recombination across the interface is blocked to
a significant extent by the presence of the sensitizer. However, in order to harvest
solar light efficiently, the use of a bulk or interpenetrating network junction of meso-
scopic dimension is necessary since the optical cross section of a dye or QD is much
smaller than the area it occupies. The challenge one faces with these systems is to
make adequate provisions for contacting a large area of the junction by a suitable
hole conductor or electrolyte in order to regenerate the sensitizer following light-
induced electron injection and conduct the positive carriers to the back contact of
the photovoltaic cell.
The validated solar to electric power conversion efficiency of the liquid
electrolyte-based DSC stands currently at 11.1% under standard reporting con-
ditions (AM1.5 global sunlight at 1,000 W/m2 intensity, 298 K temperature) [9]
rendering it a credible alternative to conventional pn junction photovoltaic devices.
Solid-state DSC equivalents using organic hole conductors have reached over 5%
efficiency with ruthenium-based sensitizers [10] and over 4% with organic dyes
[11] whereas nanocomposite films composed of inorganic materials, such a TiO2
and CuInS2 have achieved efficiencies between 5 and 6% [12, 13]. New dyes show-
ing increased optical cross sections and capable of absorbing longer wavelengths
are currently under development. Similarly, the performance of mesoscopic TiO2
films employed as electron collectors is benefiting greatly from recent advances in
nanomaterial research. Taking advantage of the highly transparent nature of the sen-
sitized nanocrystalline oxide film, tandem structures employing a DSC and CIGS
top and bottom cell reached a conversion efficiency of >15% [14].
4 M. Gratzel

Obtaining long-term stability for DSCs at temperatures of 8085 C had re-


mained a major challenge for many years and has only been achieved in 2003
[15]. Solvent-free electrolytes such as ionic liquids or solid polymers have been
introduced to provide systems that are suitable for practical applications. Stabiliza-
tion of the interface by using self-assembly of hydrophobic sensitizers alone or in
conjunction with amphiphilic co-adsorbents has been particularly rewarding. Stable
operating performance under both prolonged thermal stress (at 85 C) and AM 1.5
light soaking conditions (at 60 C) has now been demonstrated [1618]. These re-
cent devices retained 98% of their initial power conversion efficiency after 1,000 h of
high-temperature aging. Long-term accelerated testing shows that DSCs can func-
tion in a stable manner for over 20 years, if sealing and the interfacial engineering
issues are properly addressed. The present review is on the state of the art of the
academic and industrial development of nanostructured solar cells, with emphasis
being placed on the work performed in the authors laboratory.

2 Band Diagram and Operational Principle of Nanocrystalline


Solar Cells

Figure 2 shows a band diagram of the dye-sensitized nanocrystalline solar cell


(DSC) and explains its operational principle. Sunlight is harvested by the sensi-
tizer that is attached to the surface of a large band gap semiconductor, typically a
film constituted by titania nanoparticles. Photo-excitation of the dye results in the

Conducting
glass TiO2 Dye Electrolyte Cathode

Injection
S*
0.5
Maximum
Voltage
0 h
E vs
NHE Red Mediator Ox
(V) 0.5
Diffusion

1.0
S /S+
e- e-

Fig. 2 Energy band diagram of the DSC. Light absorption by the dye (S) produces an excited state
(S ) that injects an electron into the conduction band of a wide band gap semiconducting oxide,
such as TiO2 . The electrons diffuse across the oxide to the transparent current collector made of
conducting glass. From there they pass through the external circuit performing electrical work
and re-enter the cell through the back contact (cathode) by reducing a redox mediator (ox). The
reduced form of the mediator (red) regenerates the sensitizer closing the cyclic conversion of light
to electricity. See Color Plates
Recent Applications of Nanoscale Materials: Solar Cells 5

injection of electrons into the conduction band of the oxide. The dye is regenerated
by electron donation from an organic hole conductor or an electrolyte that is infil-
trated into the porous films. The latter contains most frequently the iodide/triiodide
couple as a redox shuttle although other mediators such as cobalt(II/III) complexes
[19, 20] or the TEMPO/TEMPO + redox couple [21] have also been developed
recently as an alternative to the I /I +
3 system. Reduction of S by iodide regenerates
the original form of the dye under production of triiodide ions. This prevents any
significant buildup of S+ , which could recapture the conduction band electron at the
surface. The iodide is regenerated in turn by the reduction of the triiodide ions at the
counter electrode, where the electrons are supplied by migration through the exter-
nal load completing the cycle. Thus, the device is regenerative producing electricity
from light without any permanent chemical transformation. The voltage produced
under illumination corresponds to the difference between the quasi-Fermi level of
the electron in the solid and the redox potential of the electrolyte or the work func-
tion of the hole conductor.

3 The Importance of the Nanostructure

The nanocrystalline morphology of the semiconductor film is essential for the


efficient operation of all mesoscopic photovoltaic devices. A whole range of nanos-
tructures has been tested so far ranging from simple assemblies of nanoparticles to
nanorods [19], tetrapods [20] and nanotubes [21, 22]. The use of mesocopic inter-
penetrating network junctions has also overcome the fundamental problem posed
by the notoriously short diffusion length of excitons or charge carriers that are en-
countered in organic and hybrid photovoltaic cells. For example in the well-known
fullerene/polyhexylthiophene (P3HT) [23] or CdTe/P3HT [24] cells the excitons
produced by light excitation of the polymer diffuse only over a distance of a few
nanometers during their lifetime while the light absorption length in these films is
several hundred nanometers. In a flat junction geometry the excitons recombine be-
fore they reach the junction where they dissociate into electrons and holes that pro-
duce the photocurrent. On the other hand, in a mesoscopic junction, the diffusion
path to reach the interface is shortened to a few nanometers allowing generation of
charge carriers from the excitons before they recombine.
The use of interpenetrating network junctions is essential for the DSC. On a flat
surface a monolayer of dye absorbs at most a few percent of light because it occupies
an area that is several hundred times larger than its optical cross section.
Using multi-layers of sensitizer does not offer a viable solution to this prob-
lem. Only the molecules that are in direct contact with the oxide surface would be
photoactive the remainder filtering merely the light. Apart from poor light harvest-
ing a compact semiconductor film would need to be n-doped to conduct electrons. In
this case energy transfer quenching of the excited sensitizer by the electrons in the
semiconductor would inevitably reduce the photovoltaic conversion efficiency. For
this reason the conversion yields obtained from the sensitization of flat electrodes
6 M. Gratzel

100 nm

Fig. 3 Transmission electron microscope picture of a mesoscopic TiO2 (anatase) film. Note the
bipyramidal shape of the particles having (101) oriented facets exposed. The average particle size
is 20 nm

have been notoriously low. The use of nanocrystalline films to support the dye and
collect the photo-injected electrons has permitted to overcome these problems re-
sulting in a dramatic improvement of the performance of sensitized hetero-junction
devices.
Figure 3 shows a scanning electron microscopy picture of a mesoscopic TiO2
(anatase) layer. The particles have an average size of 20 nm and the facets exposed
have mainly (101) orientation, corresponding to the anatase crystal planes with the
lowest surface energy (ca. 0.5 J/m2 ). Employing such oxide nanocrystals covered by
a monolayer of sensitizer as light harvesting units allows overcoming the notorious
inefficiency problems, which have haunted all solar energy conversion devices based
on the sensitization of wide band gap semiconductors.

3.1 Light Harvesting by a Sensitizer Monolayer Adsorbed


on a Mesoscopic Semiconductor Film

Consider a 10-m thick mesoscopic oxide film composed of 20-nm-sized particles


whose real surface area is over 1,000 times greater than the projected one. Because
of the small size of the particles such films show high transparency and negligible
Recent Applications of Nanoscale Materials: Solar Cells 7

light scattering. Lambert Beers law can be applied to describe the light absorption
by the adsorbed dye monolayer yielding for the reciprocal absorption length.

= c. (1)
Here, and c are the optical absorption cross section of the sensitizer and its
concentration in the mesoporous film, respectively. The value of can be derived
from the decadic extinction coefficient of the sensitizer using the relation:

= 1, 000 (cm2 /mol). (2)


For example, the optical cross section of the N3 dye cisRuL2 (SCN)2 (L = 2,2-
bipyridyl-4,4 dicarboxylate) is 1.4 107 cm2 /mole at 530 nm, where it has an
absorption maximum, and its concentration in the nanocrystalline film at full
monolayer coverage is typically 2 104 mol/cm3 . Hence = 2.8 103 cm1
and the absorption length 1/ is 3.6 m for 530-nm light. The light harvesting
efficiency LHE is derived from the reciprocal absorption length via:

LHE( ) = 1 10d , (3)


where d is the thickness of the nanocrystalline film. Using d = 10 m and =
2.8 103 cm1 one obtains LHE = 99.8%. On a flat surface the N3 dye would have
only absorbed 0.3% of the incident 530-nm light. The dramatic difference of the
light harvesting efficiency is illustrated by the deep coloration of the nanocrystalline
TiO2 layers shown in Fig. 4, despite of the fact that they are covered only by a
monolayer of sensitizer. On a flat surface the N3 dye would have remained invisible
to the eye. A film exhibiting ordered mesoporous structure, such as shown in Fig. 5
has an even higher internal surface area than one that is composed of randomly
associated nanoparticles [24]. Because more sensitizer is adsorbed for the same film

Fig. 4 Uptake of N3 dye by a nanocrystalline TiO2 film, which is immersed in the dye solution.
The resulting deeply red-colored film is the photoactive part of the DSC. See Color Plates
8 M. Gratzel

Fig. 5 Top scanning electron microscope view of an ordered mesoporous TiO2 (anatase) film
produced by a block-copolymer templating method [25]

Fig. 6 Scanning electron microscope view of a film constituted of titania nanotubes. The length of
the tubes is about 5 m

thickness its concentration in the mesoscopic TiO2 layer is increased, reducing the
absorption length and enhancing the absorption of solar light over that of a dye-
covered layer of randomly associated colloidal TiO2 particles of the same thickness.
The problem with mesoporous films of the type shown in Fig. 5 is that ordered
structures can only be realized so far up to a thickness of 1 m, which is not enough
to produce cells with conversion efficiencies over 5%.
One-dimensional nanostructures such as the titania nanotubes shown in Fig. 6
and ZnO nanorods have been the focus of much recent interest [19,21,22,26]. These
studies are motivated by the expectation that the transport of charge carriers along
the tubes is more facile than within a random network of nanoparticles where the
Recent Applications of Nanoscale Materials: Solar Cells 9

electrons have to cross many particle boundaries. Hence, one-dimensional nanos-


tructures should produce a lower diffusion resistance than the nanoparticle films
facilitating the collection of photo-generated charge carriers.

3.2 Enhanced Red and Near IR Response by Light Containment

The light harvesting by the surface adsorbed sensitizer can be further improved
by exploiting light localization and optical enhancement effects. This increases the
absorption of solar light in particular in the red and near IR region of the spectrum
where currently used ruthenium complexes show only weak light absorption. For
example, incorporating 200400-nm-sized anatase particles enhances significantly
the absorption of red or near infrared photons by the film. A scanning electron mi-
crograph of such particles is shown in Fig. 7. These light management strategies
employ scattering and photonic band gap effects [2729] to localize light in the
mesoporous structure augmenting the optical pathway significantly beyond the film
thickness and enhancing the harvesting of photons in a spectral region where the op-
tical cross section of the sensitizer is small. The benefits from using such a photon
capture strategies are clearly visible below, where the light scattering layer is shown
to enhance the photocurrent response of the DSC in the near IR and visible region
of the solar spectrum. The gain in short circuit photocurrent through these optical
containment effects can be as high as 30%.

Fig. 7 Scanning electron micrograph showing anatase crystals of ca. 400 nm size, employed as
light scattering centres to enhance the red response of the DSC (courtesy of Dr. Tsuguo Koyanagi,
Catalysts & Chemicals Ind. Co. Ltd., Japan). See Color Plates
10 M. Gratzel

3.3 Light-Induced Charge Separation and Conversion of Photons


to Electric Current

The incident photon to current conversion efficiency (IPCE) sometimes referred to


also as external quantum efficiency (EQE) corresponds to the number of electrons
measured as photocurrent in the external circuit divided by the monochromatic pho-
ton flux that strikes the cell. The following product expresses this key parameter:

IPCE( ) = LHE( )inj coll . (4)

Here, LHE( ) is the light harvesting efficiency for photons of wavelength , inj is
the quantum yield for electron injection from the excited sensitizer in the conduction
band of the semiconductor oxide and coll is the electron collection efficiency. Hav-
ing analyzed above the light harvesting efficiency of dye-loaded mesoscopic films
we discuss now the other two parameters.
The quantum yield of charge injection (inj ) denotes the fraction of the pho-
tons absorbed by the dye that are converted into conduction band electrons. Charge
injection from the electronically excited sensitizer into the conduction band of the
semiconductor is in competition with other radiative or radiationless deactivation
channels. Taking the sum of the rate constants of these nonproductive channels
together as kdeact gives:
(inj ) = kinj /(kdeact + kinj ). (5)
Typical kdeact values lie in the range from 103 to 1010 s1 . Hence, injection rates in
the picosecond range may have to be attained in order to obtain inj values close
to 1. The currently used sensitizers satisfy this requirement. These dyes incorporate
functional groups e.g. carboxylate, hydroxamate or phosphonate moieties [30] that
attach the sensitizer to the oxide surface. Figure 8 shows the side and top view of the
RuL2 (NCS)2 (N3) sensitizer anchored to the (101) surface plane of TiO2 through
coordinative binding of two carboxyl groups to the titanium ions. The green and
red spheres present titanium and oxygen, respectively. Note that the left carboxylate
group straddles two Ti(IV) surface ions from adjacent titanium rows corresponding
to a bidentate bridging configuration while the right one forms a monodentate ester
bond with one Ti(IV) ion. The structure shown represents the lowest energy con-
figuration derived from molecular dynamics calculations [30] yielding for the area
occupied by one adsorbed N3 molecule a value of 1.64 nm2 .
By undergoing strong coordinative bonding with the titanium surface ions, these
groups enhance electronic coupling of the sensitizer LUMO with the Ti(3d) orbitals
forming the conduction band of the semiconductor. The lowest energy electronic
transition for ruthenium polypyridyl complexes, such as the N3 dye is of MLCT
(metal to ligand charge transfer) character, albeit with significant delocalization of
the highest occupied molecular orbital (HOMO) over the SCN groups [27]. Upon
optical excitation an electron is shifted from the Ru (SCN)2 moiety of the complex
to the lowest unoccupied molecular orbital (LUMO) of the carboxylated bipyri-
dine ligands moving in close vicinity to the titania surface. In a second step, the
electron is injected from the LUMO into the conduction band of the nanocrystalline
Recent Applications of Nanoscale Materials: Solar Cells 11

Side view Top view

Fig. 8 Side and top view of the RuL2 (NCS)2 (N3) sensitizer anchored to the (101) TiO2 anatase
surface through coordinative binding of two carboxyl groups to surface titanium ions. The green
and red spheres present titanium and oxygen, respectively. Note that the left carboxylate group
straddles two Ti(IV) surface ions from adjacent surface titanium rows while the right one forms an
ester bond. The structure shown represents the lowest energy configuration derived from molecular
dynamics calculations and the area occupied by one adsorbed N3 molecule being 1.64 nm2 . See
Color Plates

1e

Fig. 9 Calculated shift in electron density during optical excitation and charge injection from the
N3 sensitizer into the conduction band of a TiO2 (anatase) cluster [31] consisting of 38 titanium
ions. The surface of the cluster corresponds to the (101) plane. See Color Plates

titania particles. Figure 9 shows the results from time-dependant DFT calculations
[32] indicating the vectorial charge displacement from the HOMO of the sensitizer
to the T(3d) orbitals of the oxide during the optical excitation and electron injec-
tion process while Fig. 10 presents a schematic of the energy levels involved in the
sensitization.
Shown in Fig. 11 is the transient absorption signal following femtosecond laser
excitation of the N-719 dye adsorbed on the surface of nanocrystalline titania
[33]. The formation of the oxidized sensitizer and conduction band electrons due
12 M. Gratzel

Fig. 10 Interfacial electron transfer involving a ruthenium complex bound to the surface of TiO2
via a carboxylated bipyridyl ligand. Orbital diagram for the forward electron injection (rate con-
stant kf ) from the orbital of the bipyridyl ligand into the empty t2g orbitals forming the TiO2
conduction band and the backward electron transfer from the conduction band of the oxide into the
Ru(III) d orbitals. See Color Plates

to heterogeneous charge transfer from the excited ruthenium complex into the
conduction band of the oxide occurs on a femtosecond time scale. Figure 11 in-
dicates that the reaction is completed within the femtosecond laser excitation pulse.
Fitted data provide a cross-correlation time of 57 fs that is consistent with the
instrument response measured by Kerr gating in a thin glass window. Hence, this
temporal resolution does not allow determination of the rate of the injection pro-
cess accurately but its time constant can be estimated as being definitely shorter
than 20 fs corresponding to a rate constant kinj > 5 1013 s1 . Such high rate can
be rationalized in terms of electronic coupling of the sensitizer LUMO with the
t2g wavefunction of the Ti(3d) conduction manifold and a large density of acceptor
states in the semiconductor. Since nuclear motion in the molecule and its environ-
ment takes place within a time frame of at least 20 fs, the observed charge injection
dynamics is certainly beyond the scope of vibration-mediated electron transfer mod-
els [3439]. The process rate is therefore likely to be limited only by the electron
dephasing in the solid. Interestingly, much slower injection kinetics extending into
Recent Applications of Nanoscale Materials: Solar Cells 13

Fig. 11 Transient absorption signal for N719 adsorbed on nanocrystalline titania (open circle)
(pump wavelength 535 nm, probe 860 nm). Fitted instrument response is 57 fs (straight line). Sim-
ulated exponential rises with time constants of 20 fs (dashed line) and 50 fs (dotted line) and con-
voluted with the same instrument response are shown

the picosecond time domain were observed when the sensitizer was present in an
aggregated form at the surface of the titania films [40].
As the next step of the conversion of light into electrical current, a complete
charge separation must be achieved. On thermodynamic grounds, the preferred pro-
cess for the electron injected into the conduction band of the titanium dioxide films
is the back reaction with the oxidized sensitizer. Naturally this reaction is undesir-
able, since instead of electrical current it merely generates heat. For the characteri-
zation of the recombination rate, an important kinetic parameter is the rate constant
kb . It is of great interest to develop sensitizer systems for which the value of kinj is
high and that of kb low.
While for the N3-type sensitizer the forward injection is a very rapid process
occurring in the femtosecond time domain, the back reaction of the electrons with
the oxidized ruthenium complex occurs on a much longer timescale of micro- to
milliseconds. One of the reasons for this striking behaviour is that the electronic
coupling element for the back reaction is one to two orders of magnitude smaller
for the back electron transfer. As the recapture of the electrons by parent sensitizer
involves a d-orbital localized on the ruthenium metal the electronic overlap with the
TiO2 conduction band is small and is further reduced by the spatial contraction of
the wavefunction upon oxidation of Ru(II) to Ru(III).
A second important contribution to the kinetic retardation of charge recombina-
tion arises from the fact that this process is characterized by a large driving force and
small reorganization energy the respective values for N-719 being 1.5 and 0.3 eV,
respectively. This places the electron recapture clearly in the inverted Marcus region
reducing its rate by several orders of magnitude [41, 42]. For the same reason the
14 M. Gratzel

interfacial redox process is almost independent of temperature and is surprisingly


insensitive to the ambience that is in contact with the film [43].
Charge recombination is furthermore inhibited by the existence of an electric
dipole field at the surface of the titanium dioxide film. While the depletion layer
field within the oxide is negligible due to the small size of the particles and their
low doping level, a dipole layer is established at the surface by proton transfer from
the carboxylic acid groups of the ruthenium complex to the oxide surface. In aprotic
media, Li+ or Mg2+ are potential determining ions for TiO2 [44] as they charge
the oxide positively. The local potential gradient from the negatively charged sensi-
tizer to the positively charged oxide drives the injection in the desired direction and
inhibits the electrons from re-exiting the solid.
Finally, the back reaction dynamics are strongly influenced by the trapping of the
conduction band in the mesoscopic film. If the diffusion of trapped electrons to the
particle surface is rate determining, the time law for the back reaction is a stretched
exponential [45]. If, by contrast the interfacial back electron transfer is so slow that
it becomes rate determining then the back reaction follows first order kinetics.

3.4 Charge Carrier Collection

The question of charge carrier percolation over the mesoscopic particle network is
presently attracting a great deal of attention. This important process leads to nearly
quantitative collection of electrons injected by the sensitizer. The large band gap
semiconductor oxide films used in dye sensitized solar cells are insulating in the
dark, however, a single electron injected in a 20-nm-sized particle produces an elec-
tron concentration of 2.4 1017 cm3 . This corresponds to a specific conductivity
of 1.6 104 S/cm if a value of 104 cm2 /s is used for the electron diffusion coeffi-
cient [46]. In reality the situation is more complex as the transport of charge carriers
in these films involves trapping unless the Fermi level of the electron is so close to
the conduction band that all the traps are filled and the electrons are moving freely.
Therefore, the depth of the traps that participate in the electron motion affects the
value of the diffusion coefficient. This explains the observation [47, 48] that the dif-
fusion coefficient increases with light intensity. Recent Monte Carlo modelling gives
an excellent description of the intricacies of the electron transport in such meso-
scopic semiconductor films [49]. Of great importance for the operation of the DSC
is the fact that charges injected in the nanoparticles can be screened on the meso-
scopic scale by the surrounding electrolyte, facilitating greatly electron percolation
[50]. The electron charge is screened by the cations in the electrolyte, which elimi-
nates the internal field, so no drift term appears in the transport equation. Figure 12
illustrates this local screening effect.
The electron motion in the conduction band of the mesoscopic oxide film is cou-
pled with interfacial electron-transfer reaction and with ion diffusion in the elec-
trolyte. Bisquert [51] has introduced a transmission line description to model these
processes. The mesoscopic film is thought to be composed of a string of oxide
nanoparticles (Fig. 12). Apart from recapture by the oxidized dye, the electrons can
Recent Applications of Nanoscale Materials: Solar Cells 15

Dye Dye Dye Dye Dye Dye Dye


Dye Dye Dye Dye Dye
Dye Dye Dye Dye
RS rtrans rtrans rtrans rtrans
Dye

Dye
Dye
rct rct rct Dye
Dye
RFTO/EL cch cch Dye
cch RCE
Dye Dye

CFTO/EL Electrolyte Zd

Dye Dye
CCE
Dye Dye Dye Dye Dye Dye Dye Dye Dye
Dye Dye
Dye Dye
Dye Dye

Dye

Dye
Dye
Dye Dye
Dye Dye Dye
Dye Dye Dye Dye Dye Dye Dye
Dye Dye Dye Dye

Fig. 12 Equivalent electric circuit diagram of a solar cell based on a nanocrystaline semiconductor
film in contact with an electrolyte. Two transmission lines are used to model the motion of the con-
duction band electron motion through a network of mesoscopic semiconductor particles and the
charge compensating flow of redox electrolyte. The electrical equivalent circuit treats each particle
as a resistive element. Interfacial electron transfer from the conduction band of the nanoparticle to
the triodide is modelled by a charge transfer resistance rct connected in parallel with their chemical
capacitance cch . The latter is defined as the electric charge (measured in Coulomb) that is required
to move the Fermi level of the of the semiconductor nanoparticles by 1 eV. Zd is the Warburg dif-
fusion resistance describing the motion of triiodide ions through the porous network to the counter
electrode while RCE and CCE are the charge transfer resistance for the reduction of triodiodide and
the double-layer capacitance of the counter electrode, respectively The red dots present cations
from the electrolyte. See Color Plates

be lost to the electrolyte by the reaction with the oxidized from of the redox media-
tor, e.g. triiodide ions:
I
3 + 2e cb(TiO2 ) 3I .

(6)
The equivalent electrical circuit shown in the lower part of the figure treats each
particle as a resistive element coupled to the electrolyte through the interface. The
latter is presented by the chemical capacitance (cch ) connected in parallel with
the resistance (rct ) for interfacial electron transfer. The red dots denote electrolyte
cations.
It is clear from Fig. 12 that the movement of electrons in the conduction band of
the mesoscopic films must be accompanied by the diffusion of charge-compensating
cations in the electrolyte layer close to the nanoparticle surface. The cations screen
the Coulomb potential of the electrons avoiding the formation of uncompensated
local space charge, which would impair the electron motion through the film.
This justifies using an ambipolar or effective diffusion coefficient, which contains
both contributions from the electrons and charge-compensating cations [48, 52] to
16 M. Gratzel

e
+ *
3 S /S


1 2
Oxidation
e
Potential
5

+ 4 Red/Ox
TiO2
S+/S Couple

Fig. 13 Photo-induced processes occurring during photovoltaic energy conversion at the surface of
the nanocrystalline titania films. 1: sensitizer (S) excitation by light, 2: radiative and nonradiative
deactivation of the sensitizer, 3: electron injection in the conduction band followed by electron
trapping and diffusion to the particle surface, 4: recapture of the conduction band electron by
the oxidized sensitizer (S+ ), 5: recombination of the conduction band electrons with the oxidized
form of the redox couple regenerating the sensitizer and transporting the positive charge to the
counterelectrode. Grey spheres: titania nanoparticles, red dots: sensitizer, green and blue dots:
oxidized and reduced form of the redox couple. See Color Plates

describe charge transport in such mesoscopic interpenetrating network solar cells


although at the high electrolyte concentrations employed the electron diffusion is
the dominating factor.
Figure 13 summarizes the injection and recombination processes. Mastering the
interface to impair the unwanted back reactions remains a key target of current
research [53]. The efficient interception of recombination by the electron donor,
e.g. iodide:
2S + 3I 2S + I
3 (7)
is crucial for obtaining good collection yields and high cycle life of the sensitizer. In
the case of N3 or its amphiphilic analogue Z-907 time-resolved laser experiments
have shown the interception to take place with a rate constant of about 105 107 s1
at the iodide concentrations that are typically applied in the solar cell [54]. This is
more than a hundred times faster than the recombination rate and >108 times faster
than the intrinsic lifetime of the oxidized sensitizer in the electrolyte in absence
of iodide.
Recent Applications of Nanoscale Materials: Solar Cells 17

To reach IPCE values close to 100%, provisions must be made to collect all
photo-generated charge carriers. A key parameter is the electron diffusion length:

Ln = De r , (8)

where De and tr are the diffusion coefficient and lifetime of the electron, respectively.
Collection of charge carriers is quantitative if the electron diffusion length exceeds
the film thickness (d):
Ln > d (9)
The film in turn needs to be significantly thicker than the light absorption length
(1/ ) in order to ascertain nearly quantitative harvesting of the light in the spectral
absorption range of the quantum dot or the molecular sensitizer:

d > 1/ . (10)

The thickness of the nanocrystalline layer required to satisfy the last conditions is
typically of the order of a few microns depending on the optical cross section of the
sensitizer and its concentration in the film as discussed earlier. A simple considera-
tion shows that the electron collection efficiency is related to the electron transport
(t ) and recombination time (tr ) or the respective electron transfer and recombination
resistances (Rt and Rct ) by the equation [55]:
 
1 1 1 t Rt
cc = + = 1 = 1 . (11)
t t r t + r Rct + Rt

The transport time of conduction band electron across a 10-m thick nanocrystalline
titania film is typically a few milliseconds while for good cells the recombination
time is in the range of seconds. This explains why practically all the injected elec-
trons reach the current collector before they recombine. These time constants or
resistances can be measured by impedance spectroscopy, which provides a power-
ful tool for analyzing the circuit elements of nanocrystalline solar cells [56, 57].

3.5 Quantum Dot Sensitizers

Semiconductor quantum dots (QDs) can replace dyes as light harvesting units in the
DSC [58, 59]. Light absorption produces excitons or electronhole pairs in the QD.
The electron is subsequently injected in the semiconducting oxide support while
the hole is transferred to a hole conductor or an electrolyte present in the pores of
the nanocrystalline oxide film. Efficient and rapid hole injection from PbS quantum
dots into triarylamine hole conductors has already been demonstrated and IPCE val-
ues exceeding 50% have been reached without attempting to optimize the collector
structure and retard interfacial electron hole recombination [59]. QDs have much
higher optical cross sections than molecular sensitizers, depending on their size.
However, they also occupy a larger area on the surface of the mesoporous electrode
18 M. Gratzel

decreasing the QD concentration in the film. As a result, the value of the absorption
length is similar to that observed for the dye-loaded films.
A recent exciting discovery shows that multiple excitons can be produced from
the absorption of a single photon by a quantum dot via impact ionization (IMI) if the
photon energy is three times higher than its band gap [60, 61]. The challenge is now
to find ways to collect the excitons before they recombine. As recombination occurs
on a femtosecond time scale, the use of mesoporous oxide collector electrodes to
remove the electrons presents a promising strategy opening up research avenues
that ultimately may lead to photovoltaic converters reaching efficiencies beyond the
Shockley Queisar limit of 31%.

4 Photovoltaic Performance of the DSC

Having dealt with the fundamental features of operation of a DSC we present now
recent performance data obtained with this new type of thin film photovoltaic cell.
A cross sectional view of the cell structure used in these experiments is shown in
Fig. 14. Both the front and back contact are made of sodium lime float glass covered
by a transparent conducting oxide. The latter material is fluorine-doped tin dioxide
(FTO) having a sheet resistance of 1015 /square and has an optical transmission
of 8090% in the visible including reflection losses. The back contact is coated
with a small amount of Pt to catalyze the interfacial electron transfer from the SnO2
electrode to triiodide the typical loading being 50 mg/m2 . The nanocrystalline TiO2
film is deposited by screen printing onto the FTO glass serving as front electrode
followed by a brief sintering in air at 450 C to remove organic impurities and en-
hance the interconnection between the nanoparticles. Adsorption of the sensitizer
monolayer occurs from solution by self-assembly. The cell is sealed using a Bynel
(Dupont) hot melt. Redox electrolyte is introduced by injection through a hole on
the back contact.

Fig. 14 Cross-sectional view of the embodiment of DSC used in the laboratory for photovoltaic
performance measurements. See Color Plates
Recent Applications of Nanoscale Materials: Solar Cells 19

4.1 Photocurrent Action Spectra

Mesoscopic TiO2 films are currently prepared mainly by hydrothermal meth-


ods, which have been standardized to yield films composed of 1520 nm-sized
anatase. The mesocopic morphology has a dramatic effect on the performance of
a dye-sensitized solar cell. Figure 15 compares the photocurrent action spectrum
obtained from a single crystal of anatase to that of a nanocrystalline film, both being
sensitized by the standard N-719 ruthenium dye i.e. cis-RuL2 (SCN)2 (L = 2,2-
bipyridyl-4,4 dicarboxylate). The incident photon-to-current conversion efficiency
(IPCE) or external quantum efficiency is plotted as a function of wavelength. The
IPCE value obtained with the single crystal electrode is only 0.13% near 530 nm,
where the N-719 sensitizer has an absorption maximum, while it reaches 88% with
the nanocrystalline electrode. As a consequence, in sunlight the photocurrent aug-
ments more than 1,000 times when passing from a single crystal to a nanocrystalline
electrode. This striking improvement defies expectation as such large-area junctions

Fig. 15 Conversion of light to electric current by dye-sensitized solar cells. The incident photon
to current conversion efficiency is plotted as a function of the excitation wavelength. Left: single
crystal anatase cut in the (001) plane. Right: nanocrystalline anatase film. Pictures of the two elec-
trodes used as current collectors are also presented. The electrolyte consisted of a solution of 0.3 M
LiI and 0.03 M I2 in acetonitrile
20 M. Gratzel

should fare poorly in photovoltaic energy conversion the presence of defects at the
disordered surface enhancing recombination of photo-generated charge carriers.
Taking into account the optical losses in the FTO glass that serve as a front
contact, the conversion of incident photons is practically quantitative in the 500
600 nm range were the sensitizer has an absorption maximum. It is apparent from
(5) that the light harvesting, electron injection- and charge carrier collection effi-
ciency must be close to unity to achieve this result. Impedance studies have shown
the diffusion length of the conduction band electrons in the DSC to be typically in
the 20100-m range. This exceeds the thickness of the nanocrystalline TiO2 film
explaining why all photo-induced charge carriers can be collected.

4.2 Overall Conversion Efficiency Under Global AM 1.5 Standard


Reporting Condition

The overall conversion efficiency of the dye-sensitized cell is determined by the


photocurrent density measured at short circuit (JSC ), the open-circuit photo-voltage
(Voc ), the fill factor of the cell (FF) and the intensity of the incident light (Is ).

global = Jsc Voc f f /Is . (12)

Under full sunlight (air mass 1.5 global, intensity Is = 1, 000 W/cm2 ), short circuit
photo-currents ranging from 16 to 22 mA/cm2 are reached with state-of-the art
ruthenium sensitizers, while Voc is 0.70.86 V and the fill factor values 0.650.8.
A certified overall power conversion efficiency of 10.4% was attained [62] in
2001. A new record efficiency over 11.2% was achieved recently [3], and Fig. 16
shows the current voltage curve obtained with this cell.

Fig. 16 Photocurrent density vs. voltage curve for a DSC employing the N-719 dye adsorbed on
a double layer of nanocrystalline TiO2 and scattering particles. The iodide/triiodide-based redox
electrolyte employed a mixture of acetonitrile and valeronitrile as a solvent. The conversion effi-
ciency in AM 1.5 sunlight was 11.18%
Recent Applications of Nanoscale Materials: Solar Cells 21

4.3 Increasing the Open Circuit Photovoltage

We have identified additives, such as guanidinium ions, which are able to suppress
the dark current at the titania electrolyte junction. Although these effects remain
yet to be fully understood it appears that these ions assist the self-assembly of dye
molecules at the TiO2 surface, rendering it more impermeable and reducing in this
fashion the dark current of the cell. In addition guanidinium butyric acid was found
to suppress the number of surface states acting as a recombination centers [63].
The ruthenium dye N-719 i.e. cis-RuL2 (SCN)2 (L = 2,2-bipyridyl-4,4 dicarbo-
xylate) is adsorbed at the TiO2 surface via two of the four carboxylate groups. The
spatial configuration of the adsorbed dye at the (111) oriented surface of the TiO2
nanocrystals has been assessed by FTIR analysis and molecular dynamics calcula-
tion. The dye monolayer is disordered and the lateral repulsion of the negatively
charged dye molecules is attenuated by spontaneous co-adsorption of cations. It is
desirable to increase the order of the dye monolayer at the interface and render it
denser. The goal here is to make the dye layer insulating in order to block the dark
current across the interface. The resulting gain in open circuit voltage can be calcu-
lated from the diode equation:

Voc = (nRT /F) ln[(isc /io ) 1], (13)

where n is the ideality factor, whose value is between 1 and 2; io is the reverse sat-
uration current and R and F are the ideal gas and Faradays constant, respectively.
Increasing the injection and lowering the recombination rates are critical for maxi-
mizing the open circuit voltage of a cell as shown by (11). Using 1.5 as a value for
the ideality factor in DSC, the reduction of the dark current by a factor of 10 would
result in a voltage increase of 90 mV, boosting the conversion efficiency of the cell
by at least 15%. The fact that the dye itself blocks the dark current of the DSC has
been confirmed recently [64].

5 Development of New Sensitizers and Redox Systems

While the improvements in DSC performance obtained recently are remarkable, it


would be very difficult to reach much higher efficiencies with the standard N-719
sensitizer unless the redox system is changed. Because of the mismatch of the redox
levels of the N-719 and the iodide/triiodide couple the regeneration reaction of the
sensitizer consumes too large a fraction of the absorbed photon energy as is apparent
from the band diagram shown in Fig. 2. Work on alternative redox systems whose
Nernst potential is better adapted to that of the N-719 dye is currently being done
[65, 66] and should ultimately lead to DSCs exhibiting Voc values above 1 V.
Alternatively, if the present iodide-based redox system is maintained, intro-
ducing panchromatic sensitizers or dye mixtures can boost the efficiency of cells
further. To give 15% conversion efficiency, these should be designed to yield at least
24 mA/cm2 short circuit current under full sunlight and fill factor as well as open
22 M. Gratzel

COOH
O

HOOC
N
N N
Ru
N N
C N
S
C
S O

Scheme 1 K-19 sensitizer with an extended p-system in one of its ligands. See Color Plates

circuit voltage values similar to those that are presently obtained. To achieve such
photocurrents the light harvesting in the 650900-nm range needs to be significantly
improved.
Scheme 1 shows the structure of a heteroleptic ruthenium complex coded K-19,
which due to the extension of the -system in one of its ligands has an enhanced
absorption coefficient. An analogue of this dye with long alkyl side chains on the
bipyridyl group, named Z-907, showed excellent light conversion performance and
cell stability [16]. These dyes of Z-series have proved themselves to be very use-
ful to solid-state DSC, where their hydrophobic nature indeed became a helpful
factor. Subsequently, they enhanced the performance of systems containing ionic
electrolytes and hole conductors. The K-19 dye also exhibits excellent conversion
yield and stability [17, 18].

6 Solid-State DSCs

Solid-state DSCs employing the lithium-coordinating K-67 dye and the hole-
conductor spiro-OMeTAD [67] in conjunction with additives such as Li(CF3 SO2 )2N,
and t-butyl pyridine have shown 54% energy conversion efficiency under AM 1.5
global illumination [10]. Here again the self-assembly of the dye molecules to a
dense layer on the TiO2 surface plays an important role, with the COOH groups
serving as anchors and the lithium coordination to the sensitizer affording local
electrostatic screening assisting charge separation.

7 DSC Stability

While long-term accelerated light soaking carried tests have confirmed the intrinsic
stability of current DSC embodiments [67] stable operation under high-temperature
stress 8085 C has been achieved only recently by judicious molecular engineering
of the sensitizer used in conjunction with a robust and non-volatile electrolyte.
Recent Applications of Nanoscale Materials: Solar Cells 23

7.1 Criteria for Long-Term Stability of the Dye

Figure 17 shows the coupling of two redox cycles involved in the solar energy con-
version process. In analogy to natural photosynthesis, light acts as an electron pump
initiating charge flow from the sensitizer via the conduction band of the oxide semi-
conductor to the external circuit. The dye is regenerated by electron donation from
iodide producing iodine or triiodide. The latter diffuses to the counter electrode
where the electrons injected into the circuit by the sensitizer reduce it back to io-
dide, thus closing the two redox cycles involved in the energy conversion process.
The turnover frequency of the sensitizer is 25 s1 in full sunshine and during 20
years of outdoor service it must support 100 million turnovers.
Scheme 2 illustrates the catalytic cycle that the sensitizer performs during cell
operation. Critical for stability are any side reactions that may occur from the excited
state S* or the oxidized state of the dye (S+ ), which would compete with electron
injection from the excited dye into the conduction band of the mesoscopic oxide
and with the regeneration of the sensitizer. These destructive channels are assumed
to follow first or pseudo-first order kinetics and are assigned the rate constants k1 and
k2 . By introducing the two branching ratios, P1 = kinj /(k1 + kinj ) and P2 = kreg /(k2 +
kreg ) where kinj and kreg are the first order or pseudo-first order rate constants for
the injection and regeneration processes, respectively. The fraction of the sensitizer
molecules that survive one cycle can be calculated as the product P1 P2 . Also,
the upper limit for the sum of the two branching ratios can be calculated for a cell
operation of 20 years and is shown to be 1 108 . The turnover frequency, averaged
over seasons and daynight time, of the dye has been derived as 0.155 s1 .

Light

Conduction
Band
I
S S* e
e e e e

1
2 I2 S+

Semiconducting
e Membrane

Electrical Work

Fig. 17 The two coupled redox cycles involved in the generation of electricity from light in a
dye-sensitized solar cell. See Color Plates
24 M. Gratzel

Scheme 2 The catalytic cycle of the sensitizer during cell operation

7.2 Kinetic Measurements

As indicated earlier, for most of the common sensitizers the rate constant for electron
injection from the excited state to the conduction band of the TiO2 particles is in the
femtosecond range. Assuming kinj = 1 1013 s1 , a destructive side reaction with
k1 < 105 s1 could be tolerated. Ruthenium sensitizers of the N3 type readily satisfy
this condition. They can undergo photo-induced loss or exchange of the thiocyanate
ligand, which however occurs at a much lower rate than the 105 s1 limit. It is also
debatable whether this pathway is destructive as the product formed still acts as
a charge transfer sensitizer. In ethanolic solution prolonged photolysis of N3 dye
leads to sulphur loss and formation of the cyanato-ruthenium complex probably via
photoxidation by oxygen. However, this reaction is not observed when the dye is
adsorbed on oxide surfaces.
Precise kinetic information has also been gathered for the second destructive
channel involving the oxidized state of the sensitizer, the key parameter being the
ratio k2 /kreg of the rate constants for the degradation of the oxidized form of the
sensitizer and its regeneration. The S+ state of the sensitizer can be readily produced
by chemical or electrochemical oxidation and its lifetime determined independently
by absorption spectroscopy. Data from a recent study of Z-907 shows that the for-
mation of its oxidized form occurs over the first 810 min after the addition of an
oxidant. The subsequent decay occurs with a lifetime of 75 min corresponding to
k2 = 2.2 104 s1 . The regeneration rate constant for this sensitizer in a typical
iodide/triiodide redox electrolyte is at least 2 105 s1 . Hence the branching ratio is
about 109 that is well below the limit of 108 admitted to achieve the 100 million
turnovers and a 20-year lifetime for the sensitizer.

7.3 Recent Experimental Results on DSC Stability

Many long-term tests have been performed with the N3-type ruthenium complexes
confirming the extraordinary stability of these charge transfer sensitizers. For exam-
ple, a European consortium financed under the Joule program [41,42] has confirmed
Recent Applications of Nanoscale Materials: Solar Cells 25

cell photocurrent stability during 10,000 h of light soaking at 2.5 sun corresponding
to ca. 56 million turnovers of the dye without any significant degradation. These
results corroborate the projections from the kinetic considerations made earlier. A
more difficult task has been to reach stability under prolonged stress at higher tem-
peratures, i.e. 8085 C. Recent stabilization of the interface by using self-assembly
of sensitizers in conjunction with amphiphilic coadsorbents has been particularly
rewarding by allowing the DSC to meet for the first time the specifications laid out
for outdoor applications of silicon photovoltaic cells. For example, the new am-
phiphlic sensitizer K-19 shows increased extinction coefficients due to extension
of the conjugation of the hydrophobic bipyridyl and the presence of electron-
donating alkoxy groups. Taking advantage of the enhanced optical absorption of
this new sensitizer and using it in conjunction with decylphoshponic acid (DPA) as
co-adsorbent and a novel electrolyte formulation, a 8% efficiency DSC has been
realized showing strikingly stable performance under both prolonged thermal stress
and light soaking [16].
Hermetically sealed cells were used for long-term thermal stress test of cells
stored in the oven at 80 C. The VOC of such a device drops only by 25 mV dur-
ing 1,000-h aging at 80 C while there was a 70-mV decline for a device stained
with the K-19 sensitizer alone. The stabilizing effect of the DPA is attributed to
the formation of a robust and compact molecular monolayer at the mesoscopicTiO2
surface, reducing the amount of adsorbed water and other interfering impurities.
This stabilization of the VOC allowed the solar cell to sustain the high conversion
efficiency during extended heat exposure [16].
Figure 18 shows results from a recent long-term illumination experiment carried
out at Dyesol with two cells over a period of close to 14 months. After 10,000 h
of continuous illumination 0.56 million coulombs of charge had passed per square
centimeter of electrode surface corresponding to a turnover number of 60 million.
During this time the measured JSC increased from initially 12 to 15 mA/cm2 while

Fig. 18 Temporal evolution of short circuit photocurrent and open circuit photo-voltage under
long-term light soaking of a Z907-sensitized DSC using a non-volatile electrolyte (courtesy of
Dyesol)
26 M. Gratzel

the Voc decreased slightly from 0.72 to 0.65 V. The opposite change of Jsc and Voc
reflects probably a small positive shift of flat-band potential of the mesoporous tita-
nia film under the thermal stress, which can result in a net enhancement of photo-
induced charge separation efficiency in the DSC.

8 First Large-Scale Field Tests and Commercial Developments

During recent years industrial interest in the dye-sensitized solar cell has surged and
the development of commercial products is progressing rapidly. A number of indus-
trial corporations, such as G24 Innovations (http://www.g24i.com), Aisin Seiki in
Japan, and well as Solaronix in Switzerland are actively pursuing the development
of both flexible- and glass-based modules. Particularly interesting are applications
in building integrated photovoltaic elements such as electric power-producing glass
tiles. The Australian company Dyesol (http://www.dyesol.com) has produced such
tiles on a large scale for field testing and several buildings have been equipped with
a wall of this type. Aisin Seiki in Japan in collaboration with Toyota Research labo-
ratory has started DSC prototype production. The layout of these modules is shown
in Fig. 19. Note the monolithic design is using carbon as interconnect and cathode
material to keep the cost down.

Fig. 19 Production of DSC prototypes by Aisin Seiki in Japan. Note the monolithic design of the
PV modules and the use of carbon as interconnect and counter electrode material. The red dye is
related to N-719 while the black dye has the structure RuL (NCS)3 where L = 2, 2 , 2 -terpyridyl-
4,4, 4 tricarboxylic acid. See Color Plates
Recent Applications of Nanoscale Materials: Solar Cells 27

Field tests of such modules have already started several years go and the results
of these tests revealed advantages of the DSC with regards to silicon panels under
realistic outdoor conditions. Thus, for equal rating under standard test conditions
(STC) the DSC modules produced 2030% more energy than the polycrystalline
silicon (pc-Si) modules [68]. The superior performance of the DSC can be ascribed
to the following factors:
The DSC efficiency is practically independent of temperature in the range of
2565 C while that of mono and pc-Si declines by ca. 20% over the same range.
Outdoor measurements indicate that the DSC exhibits lower sensitivity to light
capture as a function of the incident angle of the radiation, although this needs to
be further assessed.
The DSC shows higher conversion efficiency than pc-Si in diffuse light or cloudy
conditions.
While it is up to the commercial supplier to set the final price for such modules
it is clear that the DSC shares the cost advantage of all thin film devices. In addition
it uses only cheap and readily available materials [26]. Finally, in contrast to amor-
phous silicon and CIGS cells the DSC avoids high vacuum production steps that are
very cost intensive. Given these additional advantages at comparable conversion ef-
ficiency, module costs well below 1 are realistic targets even for plants having well
below GW capacity. The DSC has thus become a viable contender for large-scale
future solar energy conversion systems on the bases of cost, efficiency, stability and
availability as well as environmental compatibility.
These DSC panels have been installed in the walls of the Toyota dream house
shown in Fig. 20, offering a building-integrated source of solar power to the
inhabitants.

The Toyota Dream House

DSC
made by
AISIN -SEIKI

Fig. 20 The Toyota Dream House featuring DSC panels made by Aisin Seiki. For details see
web announcement http://www.toyota.co.jp/jp/news/04/Dec/nt04 1204.html. See Color Plates
28 M. Gratzel

Fig. 21 First commercial flexible lightweight cell produced by G24 Innovation on a large scale for
us as telephone chargers. See Color Plates

G24 innovation has been the first to realize large-scale, role-to-role production
of lightweight flexible cells, which are sold presently on the market for mobile tele-
phone charging. A photograph of such a cell is shown in Fig. 21.

9 Future Prospects

Reaching much beyond 12% conversion efficiency for DSC, by relying mainly on
panchromatic and IR absorbing dyes or surface modifications will require enhanced
light collection in the 700900-nm region. An alternative and promising approach
will be the use of a tandem concept, where the top and bottom cells are judiciously
chosen to absorb complimentary components of the available light including the
IR region. Such a device was recently tried in our laboratory and obtained 15%
conversion efficiency

10 Summary

Using a principle derived from natural photosynthesis, mesoscopic injection solar


cells and in particular the DSC have become a credible alternative to solid-state
pn junction devices. Conversion efficiencies over 11% and 15% have already been
Recent Applications of Nanoscale Materials: Solar Cells 29

obtained with single junction and tandem cells, respectively, on the laboratory scale,
but there is ample room for further amelioration. Future research will focus on im-
proving the Jsc by extending the light response of the sensitizers in the near IR spec-
tral region. Substantial gains in the Voc are expected from introducing ordered oxide
mesostructures and controlling the interfacial charge recombination by judicious
engineering on the molecular level. Hybrid cells based on inorganic and organic
hole conductors are an attractive option in particular for the flexible DSC embodi-
ment. Nanostructured devices using purely inorganic components will be developed
as well. The mesoscopic cells are well suited for a whole realm of applications
ranging from the low power market to large-scale applications. Their excellent per-
formance in diffuse light gives them a competitive edge over silicon in providing
electric power for stand-alone electronic equipment both indoor and outdoor. Ap-
plication of the DSC in building integrated PV has already started and will become
a fertile field of future commercial development.

Acknowledgements Financial support from the EU and Swiss sources (ENK6-CT2001-575 and
SES6-CT-2003-502783), as well as the United States Airforce (USAF contract No. FA8655-03-
13068) is acknowledged.

References

1. B.O Regan, M. Gratzel, Nature 1991, 335, 7377


2. M. Gratzel, Nature 2001, 414, 338344
3. M. Gratzel, Chem. Lett. Chem. Lett. 2005, 34, 813
4. M.K. Nazeeruddin, A. Kay, I. Rodicio, R. Humphrey-Baker, E. Muller, P. Liska,
N. Vlachopoulos, M. Gratzel, J. Am. Chem. Soc. 1993, 115, 6382
5. M. Gratzel, J. Photochem. Photobiol. A: Chem. 2004, 164, 314
6. M. Gratzel, Inorg. Chem. 2005, 44, 6841
7. M.K. Nazeeruddin, F. De Angelis, S. Fantacci, A. Selloni, G. Viscardi, P. Liska, S. Ito,
B. Takeru, M. Gratzel, J. Am. Chem. Soc. 2005, 127, 16835
8. Q. Wang, S. Ito, M. Gratzel, F. Fabregat-Santiago, I. Mora-Sero, J. Bisquert, T. Bessho,
H. Imai, J. Phys. Chem. B 2006, 110, 25210
9. Y. Chiba, A. Islam, Y. Watanabe, R. Komiya, N. Koide, L. Han, Jpn. J. Appl. Phys. Part 2,
2006, 45, 2428
10. H.J. Snaith, M. Gratzel, Nano. Lett. 2007, 7, 33723376
11. U. Schmidt-Mende, R. Bach, T. Humphry-Baker, H. Horiuchi, S. Miura, S. Ito, M. Uchida,
Adv. Mat. 2005, 17, 813815
12. M. Nanu, J. Schoonman, A. Goossens, Adv. Mat. 2004 16, 453.
13. M. Nanu, J. Schoonman, A. Goossens, Adv. Funct. Mat. 2005, 15, 95
14. P. Liska, R. Thampi, D. Bremaud, D. Rudmann, H.M. Upadhyaya, A.N. Tiwari, M. Gratzel.
Appl. Phys. Lett. 2006, 88, 203103
15. P. Wang, S.M. Zakeeruddin, J.E. Moser, M.K. Nazeeruddin, T. Sekiguchi, M. Gratzel Nat.
Mat. 2003, 2, 402
16. P. Wang, C. Klein, R. Humphry-Baker, S.M. Zakeeruddin, M. Gratzel, Appl. Phys. Lett. 2005,
86, 123508
17. D. Kuang, C. Klein, S. Ito, J.E. Moser, R. Humphry-Baker, N. Evans, F. Duriaux, C. Gratzel,
S. M. Zakeeruddin, M. Gratzel, Adv. Mater. 2007, 19, 1133
18. D. Kuang, C. Klein, Z. Zhang, , S. Ito, J-E. Moser, S.M. Zakeeruddin, M. Gratzel, Small 2007,
3, 20942102
30 M. Gratzel

19. E. Galoppini, J. Rochford, H. Chen, G. Saraf, Y. Lu, A. Hagfeldt, G. Boschloo, J. Phys. Chem.
B 2006, 110, 1615916161
20. L. Fang, J.Y. Park, Y. Cui, P. Alivisatos, J. Shcrier, B. Lee, L.W Wang, M. Salmeron, J. Chem.
Phys. 2007, 127, 184704/1184704/6
21. K. Shankar, J. Bandara, M. Paulose, H. Wietasch, O.K. Varghese, G.K. Mor, T.J. LaTempa,
M. Thelakkat, C.A. Grimes, Nano Lett. 2008, 8, 16541659
22. J.M. Macak, A. Ghicov, R. Hahn, H. Tsuchiya, P. Schmuki, J. Mat. Res. 2006, 21, 28242828
23. G. Dennler, N.S. Sariciftci, C.J. Brabec, Semiconducting Polym. (2nd Edition) 2007, 2
455530
24. W.U. Huyn, J.J. Dittmer, A.P. Alivisatos, Science 2002, 295, 2425
25. M. Zoukalova, A. Zoukal. L. Kavan, M.K. Nazeeruddin, P. Liska, M. Gratzel Nanolett. 2005,
5, 17891792
26. M. Law, L.E. Greene, J.C. Johnson, R. Saykally, P. Yang, Nat. Mat. 2005, 4, 455459
27. S. Nishimura, N. Abrams, B. Lewis, L.I. Halaoui, Th.E. Mallouk, K.D. Benkstein, J. van de
Lagemaat, J A.J. Frank, J. Am. Chem. Soc. 2003, 125, 6306
28. L.I. Halaoui, N.M. Abrams, T.E. Mallouk, J. Phys. Chem. 2005, 109, 6334
29. S. Hore, P. Nitz, C. Vetter, C. Prahl, M. Niggemann, R. Kern, Chem. Comm. 2005, 15, 2011
30. V. Shklover, M.K. Nazeeruddin, M. Gratzel, Yu.E. Ovshinikov. Appl. Organometallic Chem.
2002, 16, 635647
31. F. De Angelis, S. Fantacci, A. Seloni, Md.K. Nazeeruddin, M. Gratzel, J. Am. Chem. Soc.
2007, 129, 14156
32. A. Hagfeldt, M. Gratzel, Acc. Chem. Res. 2000, 33, 269277
33. D. Kuang, S. Ito, B. Wenger, C. Klein, J.-E. Moser, R. Humphry-Baker, S.M. Zakeeruddin,
M. Gratzel, J. Am. Chem. Soc. 2006, 128, 41464154
34. Y. Tachibana, J.E. Moser, M. Gratzel, D.R. Klug, J.R. Durrant, J. Phys. Chem. 1996, 100,
20056
35. J.E. Moser, M. Gratzel, Chimia 1998, 52, 160
36. G. Benko, J. Kallioinen, J.E.I. Korppi-Tommola, A.P. Yartsev, V. Sundstrom, J. Am. Chem.
Soc. 2002, 124, 489
37. J. Kallioinen, G. Benko, V. Sundstrom, J.E.I. Korppi-Tommola, A.P. Yartsev, J. Phys. Chem.
B, 2002, 106, 4396
38. G. Benko, J. Kallioinen, P. Myllyperkio, F. Trif. J.E.I. Korppi-Tommola, A.P. Yartsev,
V. Sundstrom, J. Phys. Chem. B. 2004, 108, 2862
39. K. Schwarzburg, R. Ernstorfer, S. Felber, F. Willig, Coord. Chem. Rev. 2004, 248(1314),
12591270
40. B. Wenger, M. Gratzel, J.E. Moser, Am. Chem. Soc. 2005, 127, 1215012151
41. J.M. Lanzafame, R.J.D. Miller, A.A. Muenter, B.A. Parkinson, J. Phys. Chem. 1992, 96, 2820
42. R. Huber, J.E. Moser, M. Gratzel, J. Wachtveitl, Proc. SPIE 2003, 5223, 121
43. J.E. Moser, and M. Gratzel, Chem. Phys. 1993, 176, 493
44. G. Redmond, and D. Fitzmaurice, J. Phys. Chem. 1993, 97, 11081
45. J. Nelson, R.E. Chandler, Coord. Chem Rev. 2004, 248, 11811194
46. H.G. Agrell, G. Boschloo, A. Hagfeldt, J. Phys. Chem. B 2004, 108, 12388
47. M.J. Cass, A.B. Walker, D. Martinez, L.M. Peter, J. Phys. Chem. B 2005, 109, 5100
48. A.J. Frank, N. Kopidakis, J. van de Lagemaat, Coord. Chem. Rev. 2004, 248, 1165
49. M.J. Cass, F.L. Qiu, A.B. Walker, A.C. Fisher, and L.M. Peter J. Phys. Chem. B 2003, 107 (1),
113119
50. W. Kubo, T. Kitamura, K. Hanabusa, Y. Wada, S. Yanagida, Chem. Commun. 2002, 374-375
51. J. Bisquert, J. Phys. Chem. B 2002, 106, 325333
52. S. Nakade, W. Kubo, Y. Saito, T. Kanzaki, T. Kitamura, Y. Wada, S.J. Yanagida, Phys. Chem.
B 2003, 107, 43744381
53. P. Wang, S.M. Zakeeruddin, J.E. Moser, R. Humphry-Baker, P. Comte, V. Aranyos,
A. Hagfeldt, M.K. Nazeeruddin, M. Gratzel, Adv. Mat. 2004, 16, 1806
54. P. Wang, B. Wenger, R. Humphry-Baker, J.-E. Moser, J. Teuscher, W. Kantlehner, J. Mezger,
E.V. Stoyanov, S.M. Zakeeruddin, M. Gratzel, J. Am. Chem. Soc. 2005, 127, 18
Recent Applications of Nanoscale Materials: Solar Cells 31

55. Q. Wang, Z. Zhang, S. M. Zakeeruddin, M. Gratzel J. Phys. Chem. C 2008, 112 (17),
70847092
56. F. Fabregat-Santiago, J. Bisquert, G. Garcia-Belmonte, G. Boschloo, A. Hagfeldt, Solar En-
ergy Mat. Solar Cells 2005, 87, 117
57. Q. Wang, J.-E. Moser, M. Gratzel, J. Phys. Chem. B 2005, 109, 1494514953
58. R. Plass, S. Pelet, J. Kruger, M. Gratzel, U. Bach. J. Phys. Chem. B 2002, 106, 75787580
59. A.J. Nozik, Quantum dot solar cells, Next Generation Photovoltaics 2004, 196222
60. R. Schaller, V.I. Klimov, Phys. Rev. Lett. 2004, 92, 186601
61. M.C. Beard, K.P. Knutsen, P. Yu, J.M. Luther, Q. Song, W.K. Metzger, R.J. Ellingson,
A.J. Nozik, Nano. Lett. 2007, 7, 25062512
62. M.K. Nazeeruddin, P. Pechy, T. Renouard, S.M. Zakeeruddin, R. Humphry-Baker, P. Comte,
P. Liska, L. Cevey, E. Costa, V. Shklover, L. Spiccia, G.B. Deacon, C.A. Bignozzi, M. Graet-
zel, J. Am. Chem. Soc. 2001, 123, 16131624
63. Z. Zhang, S.M. Zakeeruddin, B.C. ORegan, R. Humphry-Baker, M. Gratzel, J. Phys. Chem.
B 2005, 109, 2181821824
64. S. Ito, P. Liska, P. Comte, R. Charvet, P. Pechy, U. Bach, L. Schmidt-Mende, S.M.
Zakeeruddin, A. Kay, M.K. Nazeeruddin, M. Gratzel, Chem. Comm. (Cambridge, UK) 2005,
43514353
65. H. Nusbaumer, S.M. Zakeeruddin, J.-E. Moser, M. Gratzel, Chem. Europ. J. 2003, 9,
37563763
66. M. Brugnati, S. Caramori, S. Cazzanti, L. Marchini, R. Argazzi, C.A. Bignozzi, Int.
J. Photoenergy 2007, 2, 80756/180756/10
67. A. Hinsch, J.M. Kroon, R. Kern, I. Uhlendorf, J. Holzbock, A. Meyer, J. Ferber, Prog. Photo-
voltaics 2001, 9, 425438
68. T. Toyoda, T. Motohiro, Seramikkusu 2004, 39, 465468
Assembly and Properties of Nanoparticles

Caue Ribeiro and Edson R. Leite

Abstract A short review is presented on the properties and main features of


nanoscale materials, emphasizing the dependence of key properties on size for en-
ergy purposes. A general description is also given about nanoparticle synthesization
methods (mainly oxides), focusing on advances in tailoring controlled shape nano-
structures.

1 Introduction

Most of todays energy needs are still met by fossil fuels (finite reserves). However,
fossil fuels may be abandoned far earlier than generally believed in favor of clean
renewable energy sources, as soon as the latter become environmentally and eco-
nomically more attractive alternatives.
The main problem of new energy conversion and storage technologies remains
the efficiency of devices. Designs based on nanoscale-range materials can provide
new or improved technologies for devices involving electrochemical reactions and
heterogeneous catalysis such as fuel and solar cells, batteries, etc. Nanoscale struc-
tures dramatically alter surface reaction rates and electrical transport throughout the
material, considerably improving its ability to store, convert, and generate energy.
Furthermore, the design of nanoscale materials for application in alternative en-
ergy devices is a predictable way to develop a wide range of new technologies for
an environmentally friendlier future.
The physical and chemical properties of nanoscale materials (usually defined
in the 1100 nm range) are of immense interest and increasing importance for fu-
ture technological applications. Nanoparticles or nanocrystals (in this work, the
terms nanoparticles and nanocrystals are synonymous) generally display properties

E.R. Leite ()


LIEC Universidade Federal de Sao Carlos, Departamento de Qumica, Rod. Washington Luiz,
km 235 13565-905, Sao Carlos, SP, Brazil
e-mail: derl@power.ufscar.br

E.R. Leite (ed.), Nanostructured Materials for Electrochemical Energy Production 33


and Storage, Nanostructure Science and Technology, DOI 10.1007/978-0-387-49323-7 2,
c Springer Science+Business Media LLC 2009
34 C. Ribeiro and E.R. Leite

that differ from those of bulk material. The literature provides several examples of
properties, such as magnetic and optical properties, melting point, specific heat, and
surface reactivity, which can be affected by particle size [16]. A materials prop-
erties are usually very substantially modified in the 110 nm sizes. These changes
are known as quantum size effects, and their origin is directly related to the type of
chemical bond in the crystal [7]. The correlation between properties and particle size
has been known since the nineteenth century, when M. Faraday demonstrated that
the color of colloidal Au particles can be modified, changing the Au particle size [8].
However, despite the subjects long history, interest in nanoparticles has only grown
considerably over the last decade. The driving force for this increase in research
activities is the ability to control a materials properties by controlling the size and
shape of crystals and the arrangement of such particles. These developments can
lead to new technologies, including energy conversion [914], catalysis and sensors
[15, 16], ultrahigh density data storage media [1719], nanoparticle light-emitting
diodes [20,21], and special pigments [22]. From the standpoint of energy, nanostruc-
tured materials offer a way for alternative energy devices such as solar and fuel cells
to become truly feasible and for the performance of batteries and super-capacitors
for energy storage to be dramatically improved. The future of these new technolo-
gies is strictly dependent on the development of synthetic routes to process metal,
metal oxides, and semiconductor nanoparticles, as well as processes that allow such
nanoparticles to be manipulated and controlled.
This introductory chapter discusses basically two topics, i.e., the properties of
nanoparticles or, more precisely, how size modifies a materials properties, and how
to synthesize nanocrystals with a controlled morphology. Special attention is given
to fundamental subjects such as quantum confinement, phase transformation and
nucleation, and growth processes.

2 Nanoparticles Surfaces

The surface of a material plays a key role in many of its characteristics, ranging from
chemical reaction rates to optical properties [2]. Simply stated, this is due to the fact
that an inner atom in the particle will interact with other atoms in its surroundings,
whereas the surface atoms do not have neighbors in a given direction [23]. This
interaction enables nanoparticles to contribute substantially to the materials surface
disorder and, hence, to its properties. Gilbert and coworkers [24] applied wide-angle
X-ray scattering technique (WAXS) on 3.4 nm ZnS nanoparticles, attributing pair
distribution functions (PDF) to the nanoparticles. Their results revealed structural
disorder caused by nanoparticle strain and contraction of the bond lengths at the
surface. These interesting results confirm that any observation of nanoparticles must
take into account the surface effects, especially in very small particles.
The dependence of the surface area on size is, therefore, the simplest way to
observe the modulation of properties in nanoparticles. The correlation between the
surface area and volume of a spherical particle can be determined by the formula
Assembly and Properties of Nanoparticles 35

Area 4 R2p 3
= = , (1)
Volume 4/3 Rp 3 Rp

where Rp is the particle radius. To normalize the relation for a given mass (to obtain
the surface area in area/weight), it is useful multiply the result by the materials
density, . Since nanoparticles are not true spheres, we can derive a geometric ex-
pression to give the number of atoms in a near-spherical shape, as follows [25]:

R3p
N= , (2)
ra3

where ra is the radius of an atom deduced from the atomic volume. Since the
nanoparticle volume is proportional to N and (by analogy) the number of atoms
in surface scales is quadratically proportional to Rp , a similar result is obtained for
the ratio of surface atoms and inner atoms in a given particle. Thus, (1) can be inter-
preted as being related to the proportion of atoms on the particles surface. There-
fore, the relation assumes a constant value (close to zero) only for large particles;
however, in very small particles, the relation tends to infinity, i.e., the majority of
the atoms forming a particle are at the surface.
Practical conclusions can be observed in properties intrinsically related to the
coordination of atoms in space. One of the first properties studied extensively as a
function of particle size was the solidliquid transition, i.e., the variation of melting
point with size [2632]. A decrease in the melting temperature has been observed
with decreasing nanocrystal size in a wide range of materials. In CdS nanocrystals,
Goldstein and coworkers [31] observed a temperature depression of over 50% for
nanocrystals in the 15 nm range. Comparing the final results with the reported bulk
melting temperature about 1,690 C the smaller nanoparticles melted at 600 C.
This phenomenon must be interpreted in light of the fact that given their high surface
energy, surface atoms tend to be unsaturated. This surface energy is always lower in
the liquid phase than in the solid, and in the liquid phase, these atoms tend to move to
minimize energy. In the rigid geometry of a solid, the surface atoms are constrained,
and melting is a way to reduce the total surface energy. In higher surface areas,
the contribution of surface energy will be higher, and melting temperature will be
reduced.
According to the liquid-drop model, the total cohesive energy (Eb ) of a nanopar-
ticle with N atoms is equal to the volumetric or bulk energy av N minus the surface
energy, the latter term arising from the presence of atoms on the surface. Hence, the
cohesive energy per atom, i.e., Eb /N = av,Rp , is given by

4 ra2
av,Rp = av = av as N 1/3 , (3)
N 1/3
where av represents the bulks cohesive energy and is the materials surface energy.
This expression is the same as that of the binding energy per nucleon obtained from
the liquid-drop model. Since the number of atoms in a near-spherical nanoparticle
with radius Rp is known (2), we can write
36 C. Ribeiro and E.R. Leite

120
av,Rp = av , (4)
Rp

where 0 is the atomic volume. This expression gives a qualitative view of the size
dependence of the amount of energy required to remove an atom from a cluster.
Since empirical studies [26] have established a linear relation between the melting
point (Tm ) and the cohesive energy av , we can write a relation of Tm /T0 and the
nanoparticle radius, as follows [25]:

Tm 2E, ,T0
= 1 (5)
T0 Rp

where E, ,T0 is a constant dependent upon surface energy, atomic volume, and
the bulk melting temperature, T0 . Similar results have been obtained from classical
thermodynamic treatments. The linear relationship between Tm and R1 was com-
pared with experimental data for Pb and In nanoparticles, and for several metals in
moleculardynamic simulations [25]. The results showed that this simple approach
efficiently demonstrates the melting points dependence on size. Figure 1 gives an
example of melting temperature depression in sodium clusters [32], where the ap-
proach seems to be valid.
Other aspects of the high surface area in nanoparticles appear in catalytic stud-
ies. In fact, it is well known that, in heterogeneous catalysis, the rate of reaction
is assumed to be proportional to the surface coverage [33]. Therefore, the greater

Fig. 1 Melting of sodium nanoparticles (from Martin et al. [32]). The metal has bulk melting
temperature Tm,bulk = 371 K
Assembly and Properties of Nanoparticles 37

the materials surface area the greater is its catalytic activity. Several reports discuss
this enhancement of catalytic activity, mainly in metallic nanoparticles such as Pt,
Rh, Pd, and Co [3438]. However, it is not easy to separate the effects intrinsically
dependent on size from those dependent on the shape of nanoparticles [39]. Irreg-
ular shapes will strongly interfere in the vicinity of the surface atoms, especially
at the corners and edges. In fact, the sphere is the most stable geometry, and any
other shape will have a higher surface area to a given volume. Also, the distribution
of crystallographic planes differs in each shape. Narayanan and El-Sayed [3840]
compared the activation energy in the electron transfer reaction between hexacyano-
ferrate (III) ions and thiosulfate ions catalyzed by Pt nanoparticles, mainly tetra-
hedral, cubical, and spherical. Tetrahedral Pt nanoparticles are composed of (111)
facets, with sharp edges and corners, while cubical nanoparticles comprise (100)
facets whose edges and corners are less sharp than the tetrahedral ones, and spher-
ical (or near-spherical) nanoparticles consist of numerous (111) and (100) facets
with smooth corners and edges. The authors observed lower activation energy in the
above described reaction in tetrahedral nanoparticles (14.0 0.6 kJ mol1 ) than in
cubical (26.4 1.3 kJ mol1 ) and spherical (22.6 1.2 kJ mol1 ) nanoparticles. The
lower activation energy of spherical particles was attributed to the lower particle
size when compared with the cubical (4.9 against 7.1 nm). However, the comparison
with tetrahedral particles (4.8 nm) is consistent and clearly shows the dependence
on shape, which ultimately means the corners/edges and the crystallographic planes
at the surface. Another interesting feature to be pointed out is the instability of these
anisotropic shapes [41]. The dissolution and poisoning of atoms in heterogeneous
catalysis has been found to occur primarily in the corners and edges because of the
higher activity attributed to these atoms in the structure. This fact allows one to con-
clude that, when it comes to nanoparticles of comparable sizes, the spherical ones
display the lowest catalytic activity but the greatest stability.
Another aspect of this topic is the colloidal stability of nanoparticles, mainly in
metal oxides. In water, the most common liquid medium, metal oxide surface chem-
istry is controlled by the surface hydroxyl groups [4244]. The following surface
equilibrium condition must therefore be considered:

MOH+ 
2  MOH + H
+
(6)
 MO + H+
MOH  (7)

These two conditions of equilibrium are described by pK1 and pK2 values, respec-
tively. The surface charge, together with the zero point of charge or zeta potential
(pH ), are important properties that determine the stability of a colloid. In princi-
ple, the surface area interferes in the absolute number of dispersed charges, but does
not affect the zeta potential [45]. However, two effects on this scale are not negligi-
ble: the adsorption of any counter ion is enhanced and can shift pH in more than
two units [45] (this effect was explored by several authors to manipulate nanoparti-
cles by attaching them to organic molecules [46, 47]); and dipole interactions may
affect the particle-to-particle interaction [4850]. The classical equation for the en-
ergy of dipole attraction in spheres (aligned dipoles), E = D /20 x(x2 4R2p )
38 C. Ribeiro and E.R. Leite

(D is dipole moment and x is the distance between the spheres) shows that this
energy can assume significant values in small particles. Estimating these values for
CdTe nanoparticles ranging from 2.5 to 5.6 nm in diameter, Tang and coauthors [48]
reported values of about 810 kJ mol1 . These values are substantially higher than
the regular molecular dipoledipole attraction (1.5 kJ mol1 ). The authors pointed
out that this effect can explain relative orientation in nanoparticulate colloids (such
as those reported in the same paper and in others), like pearl necklaces. On the
one hand, in fact, some papers report ordered agglomeration or significant enhance-
ment of viscosity or rheological properties in nanoparticulate colloids [51,52] when
compared with microparticulate in the same solid load [44].
On the other hand, Tohver et al. [53,54] reported that the same charge localization
can interfere in the agglomeration of larger particles by repelling others. They ob-
served this behavior in dynamic viscosity measurements in binary colloids of SiO2
microspheres (0.285 m) and ZrO2 nanoparticles (3 nm). The authors proposed
that the total interparticle energy Vtotal would have an additional contribution beyond
the van der Waals interaction and the repulsive electrostatic potential, a depletion
term due to the interaction between the micro and nanoparticles. This interaction
implies that the nanoparticles avoided agglomeration of the larger ones, reducing
the expected viscosity.

3 Quantum Size Effects

Nanoparticles have few atoms that suffice only to form an identifiable (crystal) inte-
rior and the interactions among these atoms in a small limited size bring the proper-
ties close to the discrete conditions displayed by isolated molecules or atomic pairs.
These phenomena are commonly related to quantum size effects, and are revealed
mainly in optical and electrical properties.
Basic quantum mechanics provides an excellent example of energy level de-
pendence on size: the particle-in-a-box model. For an one-dimensional system,
Schrodingers equation is:
 
h 2 d2
+V = E, (8)
2m dx2

where m is the particle mass, V is the potential energy barrier, E is the energy of the
particle, and is the associated wave function. In a box with length L, the general
solution for the equation (according to Levine [55]) is given by
  n
= 2/L sin x , (9)
L
and the energy of the particle is given by

n2 2 h 2
E= , (10)
2mL2
Assembly and Properties of Nanoparticles 39

Fig. 2 The energy of the three first levels (n = 13) of a particle with mass = electron mass in a
1-dimensional box with length L. Inside the box V = 0 and outside V = (the particle is confined
in the box)

where n is an integer. The expression shows a strong dependence of the energy


level on the size (as seen in Fig. 2): at large L values, all the energy levels tend to-
ward a single level (level superposition), eliminating the quantum regime. Enhance-
ment of the particle energy is easily observed in small sizes. Although the model
is commonly referred to as only theoretical, its implications were recently observed
through scanning tunneling microscopy (STM) by Folsch et al. [56] in linear Cu
chains and by Nilius and coworkers [57] in linear Pd chains. The dependence of
the length L was observed in both studies, as illustrated in Fig. 2, by the number of
atoms in the chain.
Another way to explain the influence of size on electronic properties can
be through the semiempirical methods utilized to estimate interactions between
molecules [55]. In these methods, the contribution of each atom is postulated as ,
while the contribution of the interaction with neighboring atoms is postulated as
, and the interactions of nonneighbors are neglected (assumed to be zero). The
wave functions will have a number of solutions equal to the number of atoms in the
molecule. The general solution for the energies is given by

k
Ek = + 2 cos , (11)
N +1
where N is an integer and k = 1 N. Let us assume a particle with N equal atoms
and two energy levels a fundamental level and an excited one, with ,  and ,  ,
40 C. Ribeiro and E.R. Leite

respectively. The Fermi energy depends only on the density of electrons n, following
a free-electron model (EF = h 2 (3 2 ne /2m)2/3 ) and is therefore independent of the
particle size [58]. However, small particles have few conduction electrons ne and a
smaller number of electronic states at T = 0. Since the density of electronic states
is proportional to the volume, the space between the filled states, , is inversely
proportional to the particle size for spherical particles, the level spacing scales
with d 3 . More precisely, = 4EF /3ne . Now let us take a single particle at a given
temperature where kB T
(where kB is the Boltzmann constant). In this situation,
the Fermi energy would probably be found in a gap between adjacent levels, and the
particle would reveal a nonmetallic state.
For a semimetal (such as carbon), this feature is clearly visible in a qualitative
plot of the (11) for k = 1 and k = N in the two cases (fundamental and excited),
as shown in Fig. 3. As can be seen, the metallic behavior appears only in a cluster
with a significant number of atoms; in the presence of fewer atoms, the particles
behavior is similar to that of a semiconductor, i.e., it shows a forbidden band depen-
dent on the number of atoms in the particle. The discreteness of the electronic level
structure was discussed in detail by Kubo and Kawabata [59, 60], and reviewed by
Halperin [58].
In semiconductors, the existence (even in the bulk) of the characteristic band gap
requires a better understanding of the conduction mechanism, which differs entirely
from that of metals, to describe their quantum size effects [6163]. By definition,
the band gap is the energy necessary to create an electron (e ) and hole (h+ ). Thus,
the ionization potential Ei of a semiconductor can be treated as Ei = Eg + Ea , where
Eg is the band gap energy and Ea is the ionization energy from the edge of the
conduction band [61].

Fig. 3 Diagram of levels in a single-atom metal according to semiempirical methods


Assembly and Properties of Nanoparticles 41

The S eigenfunctions (in spherical coordinates) and their energy levels in a


charged particle (electron or hole) with effective mass m confined in a spherical
well of infinite depth and radius Rw are defined in classical quantum mechanics by
the equations

sin(n r/Rw )
n (r) = C , (12)
r
h 2 2 n2
En = . (13)
2m R2w

The internal motion in a small particle can be interpreted by a specific eigenfunction,


, given by a direct expansion of n , as defined by (12). Assuming some error, we
can write 1 + n , or, for the total energy

h 2 2 n2 e2 
E 2
+ P, (14)
2m Rw 2Rw

where P is an average over the 1S function [61]. This interesting result is equal to
the ionization potential or electron affinity of a small particle with radius Rp (making
Rw = Rp ). Both terms decrease with the size Rp of the particle, and the charge the
electron or the hole is less stable in small sizes. Consequently, the barrier to the
ionization Ea lowers with size.
However, the small size implies the approximation of the carriers (e and h+ ,
with radius re and rh , respectively), which may form a bond state, an exciton [64].
On the basis of this assumption, the pair can be described approximately by a hy-
drogenic Hamiltonian [6264]:
2 2

= h 2h h 2e e2
H , (15)
2mh 2me |re rh |

where mh , me are the effective masses of the hole and electron, and is the semicon-
ductor dielectric constant. By analogy, the band gap energy is the ionization limit of
the hydrogenic electronhole bound states. In small sizes (assumed to be spherical),
one must consider polarization terms due to the Coulomb interaction in the presence
of the crystalline surface. Hence, the Schrodinger equation is written as follows
 
h 2 h 2
+V0 (Se , Sh ) ((Se , Sh ) = E(Se , Sh ), (16)
2me 2mh

where Se and Sh are the electron and hole positions inside the sphere (V0 is con-
sidered infinity outside the sphere). At small values of Rp , the eingenfunction will
be dominated by the carrier confinement, and the solution can be obtained by the
variational method using the S wave function for a particle in a sphere (12 and 13)
as a tentative function. Thus, the simple uncorrelated function may be an acceptable
approximation (especially for direct-gap semiconductors):
42 C. Ribeiro and E.R. Leite

0 = 1 (Se )1 (Sh ). (17)

In this condition, and using the reduced mass of the electronhole pair , one has
the effective band gap energy, E eff (keeping in mind that the energy of the lowest
excited state E is the shift of Eg , i.e., the bandgap energy):

h 2 2 n2 1.8e2 e2
E eff = Egap + E = Egap + + 1 . (18)
2 R2 R R

The first term represents the quantum energy of carrier confinement or localization,
and the second term represents the Coulomb attraction. The third term is the energy
loss given by the solvation in the surrounding medium, and will depend not only
on the semiconductors dielectric constant, , but also on that of the surrounding
medium. 1 is an average sum over a 1 function [62].
Finally, the diameter-range where these size effects are expected can be deter-
mined by an exciton Bohr radius, as follows [64]:

h 2
aB = . (19)
e2
When the crystal size approaches aB , the pair becomes confined and the effects are
observable. This model is a simplification, since other correlation functions (with
more complex solutions) are possible. However, several experimental works have
demonstrated the validity of the expression for near-spherical nanoparticles [6579].
More complex solutions may be necessary for anisotropic shapes, as experimentally
demonstrated by Buhro and Covin [80] for InAs. In general, the third term of (18)
is neglected and the second term is only significant when aB Rp [79]. In these
particles, the mode is known as strong confinement, whereas for larger, albeit still
small particles, the mode is known as weak confinement [64].
Rossetti and coworkers [65, 66] compared experimental results of absorbance
in the UVVis range for CdS and ZnS nanocrystals in colloidal suspensions with
predictions from classical Mie scattering theory [81], using bulk crystal proper-
ties. This theory provides an exact solution to Maxwells equations, assuming that
small crystallites are characterized by the same wavelength-dependent dielectric
constant as that of the bulk material. For particles that are much smaller than
the optical wavelength in the external medium, only the electric dipole term
needs to be considered and the crystallite absorption cross section is given by
= (8 2 R3 )/ [(  1)/(  )].  is the ratio of the complex dielectric coeffi-
cient of the bulk crystal to the real dielectric coefficient of the external medium. Al-
though the authors found a general shape congruent with the spectra obtained with
the Mie curves, some unpredicted plateaus or peaks were observed, and the curves
were displaced to higher energies, indicating that the small particles had higher band
gap energies than the bulk. The peak was subsequently reported to be a resonance
peak resulting from the creation of the electronhole pairs [63, 64, 82]. The authors
concluded that the classical treatment of the Mie theory was invalid, since is a col-
lective property of the macroscopic material (the small crystal does not support the
Assembly and Properties of Nanoparticles 43

continuous distribution of electron or hole momenta due to the boundary conditions


imposed by the surface). Using (18), the authors estimated the effective band gaps
of some particle sizes ranging from 2 to 7.5 nm, concluding that the observed spec-
tra was the sum of widely differing spectra corresponding to various sizes. They
pointed out that the model predicts that the absorption coefficient is independent
of size, and that the crystallite size distribution itself should be used to predict the
experimental absorption threshold data.
In fact, the broadening of particle size distribution can interfere sufficiently in
the absorbance spectrum near the onset to deviate gap measurements significantly
by the onset of extrapolation. Pesika and collaborators [77, 78] proposed the inverse
observation, i.e., particle size distribution by absorbance spectra measurements. The
absorbance A at any wavelength in the quantum regime is related to the total vol-
ume of particles with radius greater than or equal to the size corresponding to the
onset of absorption, in a diluted concentration limit (absorbance will occur contin-
uously since the critical size is reached). For spherical particles, assuming that the
absorption coefficient is independent of particle size, we have

4
A(r) R3p n(Rp )dRp , (20)
Rp 3

where n(Rp ) is the particle size distribution. The expression can be rearranged, keep-
ing in mind that when r , n(Rp ) = 0:
dA
dRp
n(Rp ) 4 . (21)
3 Rp
3

The derivative dA/dRp is obtained by the slope of the absorbance curve (convert-
ing the energy axis to Rp using (18)). The authors tested the assumptions on ZnO
nanoparticles, comparing them with direct size measurements by TEM. A good
agreement with the statistical data was achieved; however, considerable deviations
were observed in very small radii. The effect was explained by the resonance ex-
citon peak, more pronounced for sharp particle size distributions. On the basis of
these observations, the authors concluded that only small fractions of the overall
distribution are affected and, in many cases, the excitons contribution is negligible.
This deviation is observed in SnO2 nanocrystals: the very small calculated Bohr
radius (aB 2.7 nm) implies that, in most cases, the experimentally detectable ef-
fects are related to a weak confinement regime. Absorbance measurements in water
were carried out on nanoparticles synthesized by the hydrolysis of hydrated SnCl2
and were compared with size distribution measured by TEM (at least 200 particles),
as illustrated in Fig. 4 [83]. An acceptable agreement was found for particles over
1.3 nm, but the smaller ones were not detected. According to Pesika, the results
were probably strongly affected by the sharp exciton peak (at 283 nm, in the inset
of the graph).
The case of SnO2 illustrates the experimental problems observed in width gap
materials (as is the case of many oxides), for it is sometimes difficult to make a
44 C. Ribeiro and E.R. Leite

Fig. 4 Comparison between statistical measurement of SnO2 particle sizes by TEM and ab-
sorbance spectra derived distribution (21). On the inset, the obtained absorbance spectra, showing
the sharp exciton peak [83]

good estimate of the band gap. The confinement effect in this oxide was studied by
photoluminescence measurements, with better results for the average particle size in
several conditions. The photoluminescence peak was considered a good estimate of
the band gap, since it relates to the emission of the lower excited level to the ground
state, whereas absorbance of the photon will occur continuously, as discussed earlier
herein. The results depicted in Fig. 5 show a good congruence with the theory and
also the weak confinement regime in this oxide.

4 Phase Stability and Transformation

Something commonly seen in the synthesis of nanomaterials is that many metastable


structures appear stable in the nanometric range. A typical case is the synthesis of
TiO2 polymorphs. TiO2 has three crystalline polymorphs, anatase, brookite, and ru-
tile [84]. Although several papers report on the synthesis of nanocrystalline anatase
[12, 8589], few report on nanocrystalline rutile [90, 91] as an example. However,
several papers state that the rutile formation passes through the three metastable
phases, and it has been established that rutile is the most stable TiO2 polymorph
(observations of micrometric anatase are scarce) [9296].
Assembly and Properties of Nanoparticles 45

Fig. 5 Comparison between application of (18) to tin dioxide and experimental data Eeff
g obtained
by photoluminescence emission and particle radius by TEM measurements (from Lee et al. [79]).
The inset in the figure shows an scheme of the photoluminescence emission (adapted from [82])

It is well known that crystallization generally follows a sequence of metastable


phases before the most stable phase is attained (known as Ostwald step rule) [94,97].
Although this final crystalline polymorph is the most thermodynamically stable, the
others are often only slightly metastable by a few kilojoules per mole. One must
keep in mind that the stability of a given nucleus or small cluster (in homogeneous
nucleation) is given by the balance between the free energy of formation, GV , and
the work given by the new surface, A (the product of the surface free energy and
the surface area), as follows for a sphere [23, 97, 98]
4
G = R3p .GV + 4 R2p . (22)
3
Making the first derivative of the expression equal to zero, d(G)/dR = 0, we have
the minimum ratio for a stable nucleus (Rp,crit ):

2
Rp,crit = . (23)
GV
At this moment, we can consider that GV is independent of size (we will further
analyze the equation later on herein). On the basis of this assumption, phases with
higher values will need larger nucleus to become stable in solution or melt. Thus,
the immediate problem is to determine the surface energy of polymorphs.
46 C. Ribeiro and E.R. Leite

Table 1 Surface enthalpies and transformation enthalpies relative to bulk stable polymorph for
oxides (adapted from Navrotsky [94])
Oxide Surface enthalpy Transformation
(J m2 ) enthalpy (kJ mol1 )
-Al2 O3 2.6 0.2 0
-Al2 O3 1.7 0.1 13.4 2.0
TiO2 rutile 2.2 0.2 0
TiO2 brookite 1.0 0.2 0.7 0.4
TiO2 anatase 0.4 0.1 2.6 0.4
ZrO2 monoclinic 6.5 0.2 0
ZrO2 tetragonal 2.1 0.05 9.5 0.4
ZrO2 amorphous 0.5 0.05 34 4
Zeolitic SiO2 0.09 0.01 815
SiO2 amorphous 0.1 9

Calorimetric measurements are a way to determine this property [84, 99]. At


room pressure, the effect of the volume variation, PV , is expected to be small, and
the surface entropy is also expected to be slight. On the basis of this assumption, the
surface free energy can be approximated to the surface enthalpy (the variable that
is properly measured in calorimetry). Table 1 shows surface enthalpies and trans-
formation enthalpy data for some oxides and their polymorphs. An initial analysis
of the data reveals (in accordance with the literature) that the surface enthalpy (or
energy) decreases as the phase becomes more metastable (higher transformation en-
thalpy relative to the bulk stable polymorph), i.e., smaller surface energies lead to
lower barriers to stabilization [93, 94]. This analysis (albeit not at all correct) indi-
cates that metastable phases tend to nucleate more easily than stable ones, so it is
coherent with Ostwalds step rule. We can interpret the same result by stating that
metastable phases have a lower activation energy G , obtained by substituting the
value Rp,crit in (22) [23, 97]:
16 3
G = (24)
3G2V
In a reactional medium (such as water, for several materials), the competition
between dissolution and reprecipitation renders this interpretation more complex.
Even though the main ideas remain valid, the first stages of the formation of a crys-
tal may be the coalescence of two fairly large clusters, eliminating water, protons,
and OH groups from the surface [2, 94]. This step often referred to as polycon-
densation [43,100] has no clear activation energy and may differ significantly from
classical nucleation. Hence, a conclusion dictated by common sense may hold true
in some cases: the metastable structure of the nanoparticle comes from a memory of
the precursor or, in the case of amorphous nanoparticles, the precursor nanostructure
constrains the atoms in positions such that crystallization is impossible.
At the surface, the low coordination of atoms unbalances the bonding forces
and generally causes any phase transformation to be easier at the surface. Under
such conditions, phase transformations will depend on particleparticle contact,
Assembly and Properties of Nanoparticles 47

while the kineticrate constant of the process may also be size-dependent (as shown
experimentally) [92, 101107]. The unbalanced forces in small particles can be un-
derstood as an excess of pressure upon the whole solid particle, and may be approx-
imated using such expressions as Laplace-Youngs equation [92, 108, 109]:
2
Peff = , (25)
Rp

where is the surface tension. Zhang and Banfield [92] assumed that the increase
in activation energy would be proportional to the excess pressure, Ea = E +CPeff =
E +C Rp , where Ea is the effective activation energy, E is the bulk energy activa-
tion, and C, C are proportionality constants. The authors confirmed this assumption
for the transition of TiO2 -anatase to rutile in the 5100 nm range, with a good corre-
lation. The difference in Ea to the total range was 60 kJ mol1 , a significant value
in view of the estimated bulk activation energy (E = 185 kJ mol1 ).
Excess pressure is also observable in pressure-driven phase transformations. Tol-
bert and Alivisatos [2, 110, 111] showed that a significant increase in pressure in-
duced wurtzite to transform into rock salt (from the less dense to the denser phase)
in CdSe nanocrystals, following a scaling law of the type Ptransf 1 + C /Rp .
They observed an increase of 35% in transition pressure in 10 nm nanoparti-
cles in relation to 21 nm nanoparticles (3.64.9 GPa). The transformations were
found to be fully reversible, albeit with some hysteresis, showing an energy bar-
rier to direct transformation, which is coherent with the excess pressure at the
surface.
However, other factors may predominate and may also be related to the afore-
mentioned phenomena. Solid surfaces of different crystallographic orientations have
different surface energies and different affinities for absorbed ions and molecules
[112] (as exemplified for SnO2 , in Table 2 [113116]). The fact is that shape is an
important variable in the stability of nanocrystals, since the total amount of surface
energy depends on the exposed crystallographic planes [117].
To adapt this feature, Barnard and Zapol [109] proposed a general model for the
phase stability of any nanoparticle based on the Gibbs free energy of an arbitrary
particle. According to the authors, the correct treatment of the free energy must
include contributions from the edges and corners rather than only from the bulk and
surface. As an example, a Si cubic nanocrystal with 200 atoms will have 9% of the

Table 2 Surface energies for SnO2 calculated for several crystallographic planes in vacuum
(adapted from Beltran et al. [113])
Surface Ref. [113] Ref. [114] Ref. [115, 116]

(110) 1.20 1.04 1.301.40


(010), (100) 1.27 1.14 1.661.65
(101), (101) 1.43 1.33 1.55
(201) 1.63
(001) 1.84 1.72 2.36
48 C. Ribeiro and E.R. Leite

atoms in edges; with 103 atoms, 4% will be in the edges and with 105 only 0.3% will
remain. These estimates highlight the importance of the small terms, in some cases.
For a given x nanoparticle, the free energy can be expressed as a sum of individual
contributions, i.e.,

G0x = Gbulk
x + Gsurface
x + Gedge
x + Gcorners
x . (26)

The first term is defined as the standard free energy of formation, Gbulk x =
G0x (T ), which is dependent on the temperature. The second term is expressed
in terms of surface energy i for each i plane on the surface and molar surface area
A. Using the relations of density , molar mass M and surface to volume ratio q,
one has
M
Gsurface = (T )A = q fi i (T ), (27)
x
i
where fi is a weight factor of the facets i in the crystal ( fi = 1). In the above
formulation, the expression takes into account the crystallographic alignment of the
properties and, indirectly, the shape. The edge and corner energies can be described
by similar expressions:
M

Gedge
x = (T )L = p g j j (T ), (28)
j
M

Gcorner
x = (T )W = w hk k (T ), (29)
k

where (T ), (T ) are the edge and corner free energies, L is the total length of edges
and W is the total number of corners, p, w the edge and the corner to volume ratios,
and g j , hk weight factors. Substituting and rearranging the terms, (26) becomes

M

G0x = G0x (T ) + q fi i (T ) + p g j j (T ) + w hk k (T ) . (30)

However, as proposed by Zhang and Banfield [92], the effective pressure must be
taken into account. The volume dilation ed is given as
V
= ed = Peff V , (31)
V
where V is the materials compressibility. Peff can be estimated by (25) above. Al-
though it is known that = + A( / A), when the dependence of on A is small,
the approximation = is acceptable. The anisotropy should be included in the
determination of , as done in (27). Using these approximations, (30) becomes
 
M 2V
G0x = G0x + 1 q fi i + p g j j + w hk k . (32)
R
Assembly and Properties of Nanoparticles 49

Table 3 Surface free energies (in J m2 ), for the clean, partially hydrogenated and fully hy-
drogenated low index surfaces of TiO2 anatase and rutile (adapted from Barnard and Zapol [96])

Anatase Rutile
Surface Clean Partial Full Surface Clean Partial Full
(001) 0.51 0.86 0.84 (100) 0.60 0.71 1.82
(100) 0.39 0.55 0.65 (110) 0.47 0.56 0.84
(101) 0.35 0.51 0.63 (011) 0.95 1.02 1.19

Although the LaplaceYoung equation (25) is only applicable to spherical particles,


the approach was tested successfully in simulations for Si, Ge, nanodiamonds, and
TiO2 polymorphs. The authors predicted faceted shapes (e.g., cubes or tetrakaidec-
ahedrons) as preferred shapes in very small sizes for most cases, which is contradic-
tory to common sense (i.e., spheres).
This strong dependence on surface, corner, and edge energies can induce unex-
pected phase transformations through interactions with the medium. Gilbert et al.
[118120] reported the reversible transition of ZnS wurtzite to a structure close to
spharelite by the gradual substitution of methanol medium (from synthesis) by wa-
ter. The authors explained the water-driven transformation through MD simulations,
showing that interactions between water and ZnS reduce the surface energy. This as-
pect can be explored by including the contribution of absorbed ions or molecules at
each surface i, i.e. [96, 121],
1 
i = A ENsurf NE bulk Ny y , (33)
2
where ENsurf and E bulk are the total energy for the surface and bulk for a given area
A, N is number of units in the stoichiometric cell (considering a minimal slab),
Ny is the number of absorbed molecules or ions and y is the chemical potential
of the y molecule or ion. Taking in account these parameters, the surface energies
for some crystallographic planes of clean, partially and fully hydrogenated TiO2
anatase and rutile were obtained by density functional calculations (as shown in
Table 3 [96]). The slight variations of some values caused by ion adsorption can be
extremely significant in small particles and in anisotropic shapes, as experimentally
observed.

5 Synthetic Methods

After reviewing the fundamentals of nanoparticle properties, in the following sec-


tions we discuss some of the fundamentals involved in nanoparticle synthesis, and
review some methods commonly employed to produce such particles, with empha-
sis on methods for synthesizing nanoparticles for energy purposes.
50 C. Ribeiro and E.R. Leite

5.1 Nucleation and Growth

Many of the synthesization methods to produce nanoparticles are based on coprecip-


itation steps, or nucleation and growth in reactional media [98, 108]. Precipitation
reactions involve the simultaneous occurrence of these steps, as well as coarsen-
ing and agglomeration processes [122, 123]. Because of the difficulties in isolating
each process for independent study, the fundamental mechanisms of precipitation
are still not entirely understood. However, a good understanding of the nucleation
step is fundamental for grasping the nature of nanosize particles.
In the section below, we discuss the correlation between the critical nucleus and
surface energy (23). A more complete analysis of precipitation must take into ac-
count the chemical potential of the nucleus formed in equilibrium with the reactional
media. In such a condition, the nucleus possesses free Gibbs energy, as follows
[23, 98, 108, 124]:
dG = 0 dn + i dSi , (34)
i

where 0 is the chemical potential of the bulk nucleus, n is the number of moles,
i is the surface energy and Si is the area for each surface i. The equation can be
rearranged to
dSi
= 0 +Vm i . (35)
i dV
since dn = dV /Vm (where Vm is the molar volume). For a particle of any shape, the
surface area Si and the volume V can be written as generic equations for a given Z
characteristic dimension, as follows:

Si = i Z 2 , (36)
V = Z . 3
(37)

where i are are geometric constants. Taking the derivative of the two expressions
as Z, and applying it to a chain rule, one has:

dSi /dZ dSi 2Si


= = . (38)
dV /dZ dV 3V

As proposed by (27), the surface energy can be inserted as an average surface ten-
sion. Here we will not consider the contributions from edges and corners, although
this contribution (as discussed) may be important. For the sake of simplicity, we can
rearrange the expression as follows:

i Si
= . (39)
Si
Inserting the (38) and (39) into (35) and rearranging them, one has:

2Vm
= 0 + F , (40)
3Z
Assembly and Properties of Nanoparticles 51

where F is a shape factor defined as i / . This equation describes the chemi-


cal potential of the formed nucleus. The chemical potential of the substance  (or
reactional product) in solution or melt is

 = 0 + RT ln a, (41)

where R is the universal gas constant, a is the activity in solution, and 0 is the
standard chemical potential in solution. We can assume that the standard chemical
potential 0 in the particle is 0 = 0 + RT ln a0 , where a0 is the saturation activity.
Comparing the equilibrium condition for precipitation, i.e., =  , and rearranging
it, one has:
a 2Vm
ln = F . (42)
a0 3ZRT
This equation is remarkable, since it describes a general form to express the rela-
tion among surface energies, chemical potential and dimension Z. For a spherical
particle, Z = Rp and F = 3, (42) reduces to

2Vm
Rp = , (43)
RT ln aa0

which is easily compared with (23). The approximation in the equation of activity to
the concentration, a c, is usual and commonly accepted as valid. We can correct
the set of equations to consider the reduction of surface energy by contact with other
surfaces (which represents a heterogeneous nucleation process) by substituting
to eff , an effective surface energy obtained by contact with another surface [125].
Applying the geometric relations of the contact with a cluster and a surface, we
can consider the decrease of activation energy to nucleation (as given in (24)), i.e.,
Ghet = G (2 + cos )(1 cos )2 )/4, where is the contact angle with the two
surfaces. The heterogeneous nucleation is preferable only if the geometric relation
is less than 1.
While the nucleation process takes place, the growth process can be concurrent
in some ways: transport of reactive species in solution, adsorption in the crystal
solution interface, interface reactions in each case, the growth corresponds to
the ordering of ions or molecular species (monomers) over the nucleus surface
[42, 124, 126128]. The growth rate is governed by empirical power laws, which
are described as a function of the slowest mechanism present [129, 130]. To obtain
nanoparticles in solution, it is usually necessary to stop the growth mechanisms or at
least to control them to prevent uncontrollable growth and, hence, undesirable par-
ticle sizes. All the reactional parameters can be controlled by the proper selection
of reactant relations [43,131]. As an example, in precipitation by hydrolysis, a large
excess of water in relation to the metal source reactant leads to nanoparticles due to
the fact that all the monomers present in solution are captured in primary nucleus
also, in these cases, the initial particle size is closer to Rp,crit .
Special cases that must be analyzed are the growth mechanisms in equilibrium,
i.e., not dependent on reactional processes but dependent on diffusional parameters
52 C. Ribeiro and E.R. Leite

and on the particles relative mobility. Initially, (43) can be rearranged assuming
the activity a is equal to the solubility of the formed particle Sp , and the saturation
activity a0 as the bulk solubility Sb,0 [70]:
 
2Vm 1
Sp = Sb,0 exp . (44)
RT Rp

This relation, widely known as the Ostwald-Freundlich equation, describes the


dependence of the solubility of a formed particle on its size. This dependence is
particularly important in very small particles, since dissolution and reprecipitation
phenomena can easily occur [132]. Usually the argument in exponential terms is
too small, and the equation can be linearized to Sp Sb,0 (1 + 2Vm /RT ) 1/Rp . If
particles dissolve and grow readily without being limited by the rate of interfacial
reactions, the growth rate of the particles is likely to be limited by diffusion through
a surrounding medium and can be described by Ficks first law. This supposition is
clearly valid in colloids and in reactional media. A convenient form of Ficks first
law for a particle in a diffusion field is given as (in spherical coordinates) [133]:
dRp dc
4 R2p = D4 r2 , (45)
dt dr
where D is the diffusivity and dc/dr is the gradient in concentration at distance r.
Considering that, in a distance r Rp , the solubility is the same as that of the average
particle size (Rp ), (45) can be rewritten for evaluation at r = Rp , after integration of
the right-hand side:
  
dRp D c0 D 2Vm 1 1
= c(Rp ) c(Rp ) = . (46)
dt Rp Rp RT Rp Rp

The concentration values are substituted by using (44) linearized, assuming c = Sp


and c0 = Sb,0 . Figure 6 shows a schematic distribution of growth rates for some ar-
bitrary values of particle radii. Clearly, the maximum growth rate will occur in a
defined range of particles. If we take the second derivative d2 Rp /dt 2 = 0, we dis-
cover that, when Rp = 2Rp , we have the condition of maximum growth rate, as
plotted in Fig. 6 by the straight line. If we assume that growth (in a closed system)
is governed by the fastest growing particles, we can write dRp = dRp . Substituting
these values in (46) and integrating them, we obtain [23, 125, 134]:

3 3 3c0 DVm
Rp Rp,0 = t. (47)
4RT
This growth mechanism, known as Ostwald ripening, provides a good descrip-
tion of the growth behavior of a wide range of nanoparticles [71, 76, 135137].
A thorough analysis of the fundamental equations (44 and 45) leads to a general
expression of the type [138140]
n n
Rp Rp,0 t, (48)
Assembly and Properties of Nanoparticles 53

Fig. 6 Growth rate in Ostwald ripening of particles with arbitrary radius. The straight line refers
to the maximum growth rate, when Rp = 2Rp

where n is an exponent dependent on the boundary conditions assumed in the growth


[137, 139]: for n = 2, it is inferred that crystal growth is controlled by ion diffusion
throughout the particle to its vicinity (the solution of a matrix, in solid compounds);
for n = 3, the growth is controlled by the volume diffusion of ions in the vicinity;
and when n = 4, it is deduced that growth is controlled by dissolution kinetics at
that particlevicinity interface.
Despite the good applicability of the Ostwald ripening model, recent studies
have demonstrated that this mechanism cannot be considered responsible for the
growth process in some systems [141145], since the main postulates of the theory
are frequently neglected. The oriented attachment mechanism was proposed as an-
other significant process, which may occur during nanocrystal growth [146150].
By this mechanism, nanocrystals can grow by the alignment and coalescence of
neighboring particles, by eliminating a common boundary. The driving force for
this mechanism is clearly the decrease in the surface and grain boundaries free
energies. By the localized nature of oriented attachment, the mechanism leads to
the formation of nanoparticles with irregular morphologies, which are not expected
in precipitation-based growth. Several studies indicate that oriented attachment is
very significant, even in the early stages of nanocrystal growth, and may lead to
the formation of anisotropic nanostructures in suspensions, such as nanorods, by
the consumption of nanoparticles as building blocks [48, 151154]. This mech-
anism has already been theoretically studied [155158] and experimentally ob-
served in micrometer-sized metallic systems for several years [159162]. Recently,
it was modeled by Moldovan et al. [163166], investigated by molecular dynamics
54 C. Ribeiro and E.R. Leite

studies [102, 128, 167, 168], and confirmed experimentally [169172]. In all of the
above-mentioned theoretical studies, the authors assumed that the nanoparticles
were in contact with each other. Oriented attachment occurred by means of rel-
ative rotations between the particles or by plastic deformation associated with dis-
placement motion, until a thermodynamically favorable interface configuration (i.e.,
crystallographic alignment) was reached. Figure 7 shows the oriented attachment of
a SnO2 nanoparticle on the surface of a SnO2 nanoribbon. Edge discordances in the
nanoparticlenanoribbon interface evidence the attachment.
However, in nanoparticle synthesis, an aspect of particular interest is controlled
growth under colloidal conditions. Considerable efforts have reportedly been ded-
icated to building adequate models to describe the coalescence of nanoparticles
in suspension and at surfaces [173]. Penn and Banfield proposed that dispersed
nanoparticles can be treated as molecules or molecular clusters [142] in solution.
This treatment was already used by Huang et al. [104, 174] in the development of a
kinetic model serving to explain ZnS nanoparticle growth induced by hydrothermal
treatments. Penn also developed a kinetic model for oriented attachment growth,
considering the electrostatic interaction between particles in solution [175].
Ribeiro et al. [176] proposed another mechanism for oriented attachment growth
in dispersed nanoparticles. The authors considered that coalescence may also occur

Fig. 7 Example of a SnO2 nanoparticle attached to a single-crystalline surface (a SnO2 nanorib-


bon) by the oriented attachment mechanism
Assembly and Properties of Nanoparticles 55

when particles with similar crystallographic orientations (or with slight differences)
collide, which explains the growth behavior in SnO2 colloidal suspensions. This
mechanism is based on the assumption that nanoparticles dispersed in a liquid
medium present a very high degree of freedom for rotation and translation mo-
tions. Hence, in suspensions where agglomeration does not take place, growth by
means of oriented collisions should be more effective than by surface mechanisms
(i.e., coalescence induced by relative rotations between particles in contact). Dis-
persed nanoparticles should present a high velocity in response to the Brownian
motion, so nanoparticles in suspension are expected to present a high frequency of
collisions. Therefore, growth by coalescence may be interpreted statistically, since
collisions can be considered effective (i.e., leading to coalescence) or ineffective
(i.e., an elastic event). This mechanism is similar to Smoluchowskys coagulation
model [126,177182], which is extensively used to explain polycrystalline colloidal
growth and aggregation mechanisms in suspension. If we assume that all of the
aforementioned considerations are valid, the coalescence of two particles in suspen-
sion may be interpreted as the following chemical equation 49:
k
A + A B (49)

where A is a primary nanoparticle and B is the product of coalescence of two


nanoparticles. Figure 8 shows a generic scheme of the two proposed ways in which
the oriented attachment mechanism works.
Several papers have discussed the role of oriented attachment in the anisotropic
growth of nanocrystals. A recent paper by Cho and coauthors [50] discusses the
oriented attachment process as the main mechanism involved in the construction of
anisotropic PbSe nanocrystals in several shapes. The authors used the proposition of

Fig. 8 Scheme of oriented attachment mechanism: a attachment by collision of two particles with
similar crystallographic orientations; b attachment induced by rotation and alignment of particles
in contact
56 C. Ribeiro and E.R. Leite

the alignment of preformed nanoparticles by dipoledipole interactions as the mech-


anism responsible for the anisotropy associated with the attachment (as discussed
before). The process was observed concomitantly to the synthesis of the nanocrys-
tals. A similar experiment was proposed by Yu et al. [183] for the synthesis of
ZnS. The authors suggested that the formation of ZnS nanorods occurred through
the oriented attachment of spherical nanoparticles formed in a precursor solution to
preformed nanorods in solution. An interesting feature was the subsequent coarsen-
ing (Ostwald ripening) of the final nanorods, smoothing the nanoribbon surface.
Anisotropic nanoparticle shapes are often desirable in charge-carrier devices
such as solar cells or electrodes [184, 185]. A strategy developed by Vayssieres
[186195], called purpose-built materials (PBM), appears to be a promising route
for growth-oriented 3D crystalline nanorods in several kinds of substrates. This
strategy allows one to obtain 3D arrays of several semiconducting metal oxides
using controlled aqueous chemical processing at low temperatures (usually in the
range of 9095 C) and inexpensive precursors [194]. The PBM strategy is based
on the heterogeneous nucleation of the desired phase on a substrate. The substrate
decreases the activation energy in the crystallization process, promoting heteroge-
neous nucleation at the surface of the solid. The kinetics of nucleation and growth
is controlled by the temperature and the hydrolysis rate, whereby the nucleation and
growth processes can be separated. The experimental conditions are adjusted to ob-
tain the most stable thermodynamic structure. The growth mechanism is controlled
basically by a dissolutionprecipitation process, and the direction of crystal growth
is controlled by the surface energy of the oxide crystalline phase. This argument ex-
plains the epitaxial growth of single-crystalline metal oxide nanorods perpendicular
to the substrate. This growth process is usually associated with the Ostwald-ripening
mechanism, but the oriented attachment mechanism may be involved.

5.2 Synthesis of Transition Metal Nanocrystals

The development of metal nanocrystals is fundamental for devices where catalytic


heterogeneous reactions take place, such as fuel cells. The bottom-up methods of
wet chemical nanocrystal synthesis are based on the chemical reduction of the salts,
or the controlled decomposition of metastable organometallic compounds in an or-
ganic or water solution. These reactions are always carried out in the presence of a
large variety of stabilizers, which are used basically to control the growth of the ini-
tial nanocluster and to prevent particle coagulation or agglomeration. As discussed,
the mechanism of nanoparticle formation is generally based on a process of nucle-
ation, growth, and agglomeration. This process was proposed by Turkevich and is
based on the synthesis of metal nanoparticles by salt reduction [34,196]. This model
still is valid and has recently been refined. A recent review authored by Bonnemann
and Richards [197] contains a good discussion about the refined model, and supple-
mentary references on this subject are also available.
Assembly and Properties of Nanoparticles 57

Since nanocrystals are unstable from the standpoint of agglomeration and bulk,
coagulation and agglomeration are the paths that nanoparticles follow to decrease
their high surface area, thus becoming more stable. In the absence of any extrinsic
impediment, the unprotected particle coagulates, basically under the action of van
der Waals forces. To prevent the coagulation process from occurring, the particle
surface can be protected by electrostatic stabilization and/or steric stabilization [44].
Electrostatic stabilization is based on the Coulombic repulsion between particles,
promoted by a double layer composed of ions adsorbed on the particle surface. The
electrostatic stabilization process can be modified by several parameters, such as
ionic strength of the dispersing media, ion concentration, and the presence of neu-
tral adsorbate, which may replace the adsorbed ion on the particle surface. Steric
stabilization is based on the steric hindrance caused by organic molecules that are
attached to the particle surface, forming a protective layer that prevents particle
coagulation or agglomeration. This type of stabilizing system can be viewed as a
nanocomposite material, since the organic layer forms a nanometric scale second
phase [198, 199]. Several kinds of protective groups can be used as steric stabi-
lization agents, among them polymers and block-polymers, P, N, and S donors
(phosphanes, amines, thioethers), surfactants, organometallic compounds, and sol-
vents. A detailed description of the several types of steric stabilizers used during the
synthesis of metal nanocrystals is given by Bradley [42].
The synthesis of transition metal nanocrystals can be divided basically into two
major groups: salt reduction and decomposition method. Examples of these methods
are described below.
The salt reduction method is a process whereby a reduction agent reduces the
metal salt, in solution, to metal. These reactions can be done in water or in an or-
ganic solution. In an organic solution, the solvent can also act as a reduction agent.
Alcohols are generally useful reduction agents, particularly hydrogen-containing al-
cohols. In this process, the alcohol is oxidized to the corresponding carbonyl group.
An example of this kind of synthesis is the processing of palladium nanoparticles
through the reduction of palladium acetate by methanol [200]. Teranishi and Miyake
[37] reported on the reduction of H2 PdCl4 by alcohols to synthesize Pd nanoparti-
cles, demonstrating that the mean diameter of Pd nanocrystals can be controlled
from 1.7 to 3.0 nm in a one-step process by changing the amount of protective poly-
mer, poly(N-vynil-2-pyrrolidone) (PVP) and the kind and/or concentration of al-
cohol in the solvent. The solvent they used was water. They also showed that the
reduction rate of [PdCl4 ] ions is an important factor in the production of smaller
Pd particles. The reduction rate was controlled using different kinds of alcohol.
The reduction of metal salts by the addition of a reducing agent in a nonre-
ducing solvent is a well-established synthetic route for the preparation of aque-
ous suspensions of metal nanocrystals. Faraday, for instance, used phosphorous
vapor to promote the reduction of [AuCl4 ] in aqueous solution to synthesize
gold nanoparticles [8]. Different kinds of reducing agents have been used to pro-
cess gold nanocrystals, allowing for the processing of particles ranging from 1 to
100 nm in diameter. Turkevitch and coworkers [196] established the first repro-
ducible standard protocol for the synthesis of gold nanoparticles. Their processing
58 C. Ribeiro and E.R. Leite

of gold nanoparticles by the reduction of [AuCl4 ] with sodium citrate, for example,
became a standard for histological staining applications [201] and for undergradu-
ate experiments in surface and nanomaterials chemistry [202]. Platinum nanoparti-
cles can also be synthesized by the reduction of metal salts, using a reducing agent
[36, 203]. Van Rheenen et al. [203] demonstrated that the morphology of platinum
particles could be controlled by controlling the synthetic parameters, such as tem-
perature, protective polymer, time, pH, reagent concentration, and the sequence of
reagent additions. These authors used various reducing agents and chloroplatinic
acid as platinum salt.
An interesting synthetization route was recently developed based on the reduc-
tion of organometallic compounds by dihydrogen at low pressure and temperature
[204208]. The organometallic compounds used were low-valent alkene or a poly-
ene complex of the desired metal. Using this process, well-dispersed nanoparticles
of Ru, Pt, Ni, and Co with a narrow size distribution were synthesized. The particles
were stabilized by the presence of poly(vinylpyrrolidone) (PVP). On the basis of a
similar process, Ould Ely et al. [209] synthesized nanoscale bimetallic Cox Pt1 x
particles, using Co( 3C8 H13 )( 4C8 H12 ) and Pt2 (dba)3 (dba = bis-dibenzylidene
acetone) as organometallic compounds. They found that the alloys composition was
determined by the initial ratio of the two organometallic precursors.
Recently, the so-called polyol process [210] has been used successfully to process
magnetic nanoparticles with a very narrow particle size distribution [211,212]. This
process is based on the reduction of metallic salt in solution, at a high temperature
(100 < T < 300 C), by the addition of a polyol (such as ethylene glycol), resulting
in nanometric particles. In this process, surfactants such as oleic acid are used to
control particle growth and stabilize the nanoparticles.
Park and Cheon [213] discussed an interesting synthetization route to process
solid solution and core-shell type cobalt-platinum nanoparticles via redox transmet-
allation reaction, reporting they had obtained nanoparticles of solid solution and
core-shell structures smaller than 10 nm. These alloys were formed by redox trans-
metallation reactions between the reagents without the addition of reducing agents.
The reaction between Co2 (CO)8 and Pt(hfac)2 (hfac = hexafluoroacetylacetonate)
resulted in the formation of solid solution, while the reaction between Co nanopar-
ticles and Pt(hfac)2 in solution resulted in Co-core Pt-shell type nanoparticles.
Narrow particle size distributions were achieved in both processes.
The organometallic compounds of transition metals usually display low ther-
mal stability, decomposing into their respective metals even under mild condi-
tions. Owing to these properties, organometallic compounds can be considered good
sources to process metal nanoparticles. Metal carbonyl pyrolysis has been used for
the synthesis of several metal nanoparticles, although a broad particle size distri-
bution is usually obtained [127, 214]. Park et al. [215] reported on the synthesis
of iron nanorods and spherical nanoparticles using the thermal decomposition of
Fe(CO)5 , in the presence of surfactant. They found that rod-like particles, with a
higher aspect ratio, could be obtained by changing the concentration of didode-
cyldimethylammonium bromide (DDAB) during the reaction process. Puntes et al.
[216] recently reported on the control of the size and shape of Co nanocrystals.
Assembly and Properties of Nanoparticles 59

A synthetization route, based on the principles applied to the synthesis and control
of CdSe nanocrystals, was used [143]. These authors discussed the synthesis of Co
nanoparticles with high crystallinity, narrow particle size distribution and a high de-
gree of shape control. The nanocrystals are produced by injecting an organometallic
precursor (Co2 (CO)8 ) into a hot (T 180 C) surfactant mixture (oleic acid and
trioctylphosphine oxide (TOPO)) under an inert atmosphere.
An interesting approach to synthesize metal alloy nanocrystals is the use of si-
multaneous salt reduction and thermal decomposition processes. Sun et al. [18]
reported on the synthesis of ironplatinum (FePt) nanoparticles through the reduc-
tion of platinum acetylacetonate by a diol, and decomposition of iron pentacarbonyl
(Fe(CO)5 ) in the presence of a surfactant mixture (oleic acid and oleyl amine). On
the basis of a similar approach, Chen and Nikles [217] synthesized ternary alloy
nanoparticles (Fex Coy Pt100xy ), using a simultaneous reduction of acetylacetonate
and platinum acetylacetonate and thermal decomposition of Fe(CO)5 and obtaining
an average particle diameter of 3.5 nm and narrow particle size distribution.
The general process of metal nanocrystal synthesis can be divided, for didactic
purposes, into five steps. The first step (step I) consists of the reduction of the metal-
lic precursor (M+ X ), which results in metal atoms (Mo ). These metallic atoms,
ions, and metallic clusters will interact (step II), resulting in a metallic cluster growth
process. Steps I and II are reversible. When the cluster grows to a critical size (step
III), the process becomes irreversible (thermodynamic condition). Particle size can
be controlled with the aid of stabilizers (step IV). The presence of only one stabilizer
can result in a spherical particle. The origin of this morphology is thermodynamic.
In fact, and as extensively discussed earlier, the cluster will grow in a geometri-
cal arrangement to minimize the surface energy. The presence of two simultaneous
stabilizers may give rise to a preferential growth process caused by the preferen-
tial adsorption of one of the stabilizers. This process, which leads to the formation
of anisotropic particles such as nanorods, occurs under a kinetic condition. Parti-
cle agglomeration, basically, is prevented by steric stabilization under the influence
of the molecules attached to the particles surface (step V). Step V is essential to
control nanoparticle deposition. Thus, colloidal metal dispersion can be used as a
building block to produce functional materials [218]. The nanocrystal self-assembly
process requires a monodispersion system (particle size deviating by less than 10%
from the average size) [219] and can be achieved by solvent evaporation [18, 217]
or polymer-mediated nanocrystal assembly [220].

5.3 Metal Oxide Nanocrystals

Metal oxides represent an important class of materials with a variety of technologi-


cal applications. In general, transition metal oxides are vital for a series of technolo-
gies, such as solar cells and Li ion batteries. Several reports in the literature describe
the effects of size on the various properties of this class of materials. Nanocrystalline
metal oxide semiconductors such as TiO2 , SnO2 and ZnO, for example, display
60 C. Ribeiro and E.R. Leite

a quantum confinement effect, with enlargement of the band gap as the particle
size decreases [73, 135, 221]. Colloidal nanocrystals with quantum size effects are
promising building blocks for novel electrical and optoelectronic devices [2, 222].
Based on the above analysis, the development of metal oxides of nanometric
dimensions can result in devices and materials with superior performance. However,
these developments are directly related to the development of synthetic methods that
allow for controlled particle size, particle morphology, and deposition. Once again,
the bottom-up methods of wet chemical nanocrystal synthesis are apparently the
most viable approach to achieve such control. Compared with the control attained
in the synthesis of metal and II-IV semiconductor nanocrystals, the control of metal
oxide nanocrystals is still incipient, particularly insofar as the synthesis of complex
metal oxide nanocrystals (oxides formed of more than one cation) is concerned.
The synthesis of metal oxide nanocrystals by wet chemical processes can be
divided basically into two major groups: (a) chemical synthesis method based on
the hydrolysis of metal alkoxides or metal halides; (b) chemical synthesis based on
the nonhydrolytic method. Examples of these methods are described below.
The chemical synthesis of metal oxide nanocrystals based on hydrolysis falls
into two major groups: hydrolysis of metal alkoxides; hydrolysis of metal halides,
and other inorganic salts. Metal alkoxide compounds are defined as compounds that
have metaloxygencarbon bonds. Si(OC2 H5 )4 (tetraethyl orthosilicate or TEOS),
for instance, is an alkoxide compound. This class of compound is highly reactive
with water. Because the hydroxyl ion (OH ) becomes bonded to the metal of the
organic precursor, this reaction is called hydrolysis. Equation (50) shows a typical
hydrolytic reaction of an alkoxide compound [100]:

M OR + H2 O M OH + ROH (50)

where M represents Si, Ti, Zr, Al, and other metals, R is a ligand such as an alkyl
group, and ROH is an alcohol. Hydrolytic reactions are strongly dependent on wa-
ter content and catalysts. Because of the high reactivity of alkoxide compounds with
water, hydrolytic reactions must be carried out in an atmosphere devoid of water va-
por and the solvents used must have very low water content. A partially hydrolyzed
metal alkoxide molecule can react with other partially hydrolyzed molecules by a
polycondensation reaction, as described in (51) and (52):

M OH + M OR M O M + ROH (51)
M OH + M OH M O M + ROH (52)

This type of reaction leads to the formation of an inorganic polymer or a three-


dimensional network formed of metal oxianions. The above-described process is
called metal alkoxide-based solgel. The literature contains excellent reports pro-
viding in-depth analyses of this method [100, 223]. The solgel process allows for
very good chemical homogeneity and offers the possibility of obtaining metastable
phases, including the amorphous phase. This process normally promotes the forma-
tion of amorphous metal oxides, which require thermal or hydrothermal treatment to
promote crystallization. Several factors affect the solgel process, including the kind
Assembly and Properties of Nanoparticles 61

of metal alkoxide, pH of the reaction solution, water/alkoxide ratio, temperature,


nature of the solvent and stabilizers [100]. By varying these parameters, particles
can be synthesized with controlled size, morphology, and agglomeration. When the
metal alkoxides hydrolytic reaction rate is too fast, particle size and morphology
are more difficult to control. A good alternative to overcome this problem is to use
organic additives, which act as chelating ligands (carboxylic acids, -diketones, and
others) and decrease the precursors reactivity [224].
The solgel process can generally be divided into three steps: (1) precipitation
of hydrous oxide particles, (2) control of hydrous oxide particle coagulation, and
(3) crystallization of the hydrous oxide particle. Thus, the solgel process requires
control of the particle size and morphology during the precipitation and coagulation
steps and during the heat or hydrothermal treatment to promote crystallization. The
precipitation of amorphous metal oxide (step 1) is well described by the LaMer and
Dinegar theory [225]. Following this model, the supersaturation of hydrous oxides
increases continuously (by a change in temperature or pH) until a critical concen-
tration is reached. In this condition, nucleation occurs very rapidly and leads to pre-
cipitation. Then, the applicability of the nucleation theory discussed above, in this
case, is only achieved by considering the as-mentioned monomer, i.e., the solvated
metallic ion. Precipitation decreases the supersaturation to levels below the critical
concentration, preventing further nucleation and precipitation. After nucleation oc-
curs, the nuclei thus formed grow, reducing the concentration until an equilibrium
concentration is achieved. The growth of particles may then occur especially by the
Ostwald ripening mechanism (48).
Controlled hydrolysis is one of the most popular methods for processing silica
spheres in the range of 101,000 nm. The method was developed by Stober, Fink,
and Bohn (SFB) [226229] and is based on the hydrolysis of TEOS in a basic so-
lution of water and alcohol. Particle size depends on the reactant concentration, i.e.,
the TEOS/alcohol ratio, water concentration, and pH (>7). This method has been
extended to other metal oxide systems with similar success, particularly for TiO2
synthesis [85, 230]. The hydrous oxide particles precipitated by the hydrolysis of
an alkoxide compound have the same tendency to agglomerate as that described for
metal colloid systems. Different stabilizers can be used to stabilize these particles
and prevent coagulation (step 2). These stabilizers control coagulation by electro-
static repulsion or by steric effects [44], similarly to the metal colloid systems.
There is not only a similarity but also a fundamental difference between the ap-
proach used to control coagulation in the solgel process and that used for metal
nanocrystal systems. As previously discussed, in metal oxide particles the surface
charge is controlled by the protonation or deprotonation of their hydrous oxide sur-
faces (M-OH). Thus, the charge-determining ions are H+ and OH . The ease with
which protonation or deprotonation occurs will depend on the metal atoms and can
be controlled by the pH.
Electrostatic stabilization is most commonly employed in the water solution sys-
tem, while steric stabilization can be more effective in organic media [231, 232]. A
steric stabilizer can be used to control the condensation reaction during the precip-
itation of hydrous oxides. In this case, the stabilizer is added during the hydrolysis
62 C. Ribeiro and E.R. Leite

step [86, 87, 233, 234]. Peiro and coworkers [235] recently reported on the synthesis
of TiO2 anatase phase with a nanorod morphology (9 5 nm size) using controlled
hydrolysis of tetraisopropyl orthotitanate (TIP) and tetrabutylammonium hydroxide
((TBA)OH) as a steric stabilizing agent.
Finally, the crystallization process (step 3) can be considered as the critical step
in the solgel process when a crystalline phase is desirable. If an amorphous phase
is the final target, as in the case of SiO2 nanoparticle processing, the synthesis is
complete in step 2. However, for crystalline materials, a heat or hydrothermal treat-
ment is often necessary to promote the crystallization of the generally amorphous
hydrous oxide which is formed during hydrolysis. Such subsequent treatments can
lead to particle growth and modify the particles morphology.
In the case of crystallization by heat treatment, the hydrous oxide colloidal sus-
pension must be dry before the treatment. During the heat treatment, normally done
in an electric furnace, crystallization occurs by a nucleation growth process into the
generated nucleus and can be described by the discussed nucleation and growth the-
ory [100, 236]. Since the amorphous phase crystallizes via the nucleation-growth
process, particle size and growth can be controlled based on the separation of the
nucleation phenomena from the growth process. However, during crystallization,
each hydrous metal oxide particle can generate several nuclei, rendering it very dif-
ficult to control particle morphology and shape. The polynuclei process generates
polycrystalline particles rather than freestanding ones. Controlling the generation of
polynuclei in a single amorphous particle is the main challenge involved in obtain-
ing crystalline metal oxide freestanding nanoparticles through the solgel process
or by any other process that requires crystallization by heat treatment at high tem-
peratures. The origin of the polynuclei process occurring during solgel amorphous
precursor crystallization is assumed to be related to the preferential heterogeneous
nucleation process (at the surface and interface nucleation) in detriment to homoge-
neous nucleation. The presence of hydroxyl groups and other defects on the particle
surface can contribute to reduce the Gibbs free energy for crystallization, render-
ing the surface crystallization more favorable than the bulk crystallization. Since
crystallization occurs in a scenario of high driving force (high temperature of heat
treatment), surface crystallization must occur first, followed or not by bulk crystal-
lization, giving rise to a particle with several nuclei. A possible way to avoid this
problem is to suppress surface crystallization by using an inhibitory surface layer. If
crystallization occurs at a temperature that favors bulk crystallization, a single nu-
cleus can be generated, resulting in a freestanding particle. This approach was used
recently to process freestanding lead zirconate titanate (PZT) nanoparticles. Liu and
coworkers [237] used a solgel process based on controlled hydrolysis and a two-
step heat treatment. They first applied a 12-h treatment in Ar atmosphere at 700 C,
which formed a surface layer rich in carbonaceous materials on the nanoparticles,
inhibiting surface nucleation. A second treatment was carried out at 500600 C, in
air, to burn out the carbon residue. Freestanding PZT nanoparticles with a mean
particle size of 17 nm were reported.
Freestanding particles are desirable in a variety of fundamental studies and
in some technologies, particularly for ferroelectric metal oxides such as PbTiO3
Assembly and Properties of Nanoparticles 63

(PT), Pb(Zr,Ti)O3 (PZT), BaTiO3 (BT), among others. Freestanding and single
crystallinenanorods of BT and SrTiO3 (ST) were recently obtained [238, 239]. The
approach used in both these studies to obtain this type of material was the injection
of a bimetallic alkoxide compound into a solvent at high temperature (100280 C),
in which the hydrolysis took place (injection-hydrolysis method). OBrien et al.
[238] synthesized BT nanoparticles with diameters ranging from 612 nm based
on this approach, controlling the particle size by the ratio BaTi(OR)6 /oleic acid.
Urban and coworkers [239] reported on the synthesis of BT and ST single crys-
talline nanorods using a similar process. The origin of nanorod morphology is not
yet well understood. Nonetheless, the above-described approach appears to promote
and control crystallization with no extra heat treatment, allowing for good control
of particle size and morphology.
Another alternative approach to avoid the heat treatment process is to promote
crystallization under hydrothermal conditions, a process that is widely used in the
synthesis of zeolites [240, 241]. In hydrothermal conditions, the solubility of the
amorphous oxide particle is significantly enhanced, and the crystallization may oc-
cur concurrently to growth processes, i.e., redissolution and reprecipitation of the
particles but in crystalline nuclei. Ying and Wang [242] used hydrothermal treat-
ment to promote the crystallization of anatase and rutile phases, using an alkoxide
solgel route and achieving the crystallization of anatase TiO2 phase with a mean
particle size of 10 nm at 180 C, as well as the synthesis of ultra fine rutile TiO2
phase obtained by hydrothermal treatment in an acidic medium.
The hydrolysis of metal halides and other inorganic salts is a method widely
employed to process metal oxide nanoparticles, such as TiO2 [90, 243], doped and
undoped SnO2 [244248], ZnO temeulenkamp98, meulenkamp981 , wong98, ZrO2
[245, 250], Y2 O3 [251], and others. This process is less sensitive to water content,
requiring less control than the hydrolysis of metal alkoxide. In fact, the hydrolytic
process normally occurs in a water solution. In solution, the metallic salt generates
the anion (Cl , F , NO n+
3 , and others) and the cation (M ). The cation is normally
hydrolyzed by pH changing. Hydrolysis promotes the precipitation of an insoluble
amorphous hydrous metal oxide. Thus, the steps used to describe the metal alkoxide
hydrolysis-based solgel process can also be used to describe the solgel process
based on inorganic salts. The synthesis based on this approach requires the same
control described earlier for the solgel method related to the hydrolysis of metal
alkoxide. However, control of the atmosphere and water content in the solvents is
much less demanding.
The crystallinity of the formed particle is, in any case, determined by thermody-
namic and kinetic parameters. Leite and coworkers [252,253] recently demonstrated
that well-crystallized SnO2 nanocrystals could be produced at room temperature
with no hydrothermal treatment. This process is based on the hydrolysis of SnCl2 in
an ethanol solution, followed by dialysis to remove the Cl ions. The result of this
dialysis is a transparent colloidal suspension formed by near-spheric particles, as
illustrated in Fig. 9 later. Zinc Oxide (ZnO) nanocrystals have also been synthesized
at room temperature. The process developed by Bahnemann et al. [67] consists of
hydrolyzing zinc acetate dihydrate dissolved in 2-propanol by the addition of NaOH
64 C. Ribeiro and E.R. Leite

Fig. 9 Tin dioxide (SnO2 ) particles synthesized by hydrolysis of SnCl2 at room temperature. The
final particles (after elimination of residual chloride by dialysis) are near-spherical nanocrystals

in a 2-propanol solution. A colloidal suspension of crystalline ZnO nanoparticles is


obtained without hydrothermal treatment. Similar results were obtained by Span-
hel and Anderson [254] and by Meulenkamp emeulenkamp98, meulenkamp981;
however, they dissolved the zinc acetate dihydrate (Zn(Ac)2 xH2 O) in ethanol and
used LiOH to promote the Zn+2 hydrolysis. Particles in the range of 36 nm were
reportedly obtained by both these processes.
However, the most common product of synthesis is amorphous, like in metal
alkoxide hydrolysis and, again, the major problem of the metal salts hydrolysis ap-
proach is the crystallization step. An adequate route to obtain good crystallinity
at low temperatures with minimum particle growth is the hydrothermal treatment.
Nutz and Haase [244] synthesized well-crystallized Sb-doped SnO2 nanocrystals,
with particles in the range of 49 nm, using a hydrothermal treatment of colloidal
gel. The gel was treated in an autoclave at temperatures in excess of 250 C. The
authors used a solution of SnCl4 and SbCl3 or SbCl5 in fuming HCl as precur-
sors and promoted hydrolysis by increasing the pH (using aqueous ammonium).
Goebbert et al. [247] also reported on the synthesis of well-crystallized Sb-doped
SnO2 , using the hydrothermal process. However, they used a solution of SnCl4 and
SbCl3 or SbCl5 in ethanol, promoting hydrolysis by raising the pH (using aqueous
ammonium). The hydrothermal treatment was carried out at 150 C using 10 bar of
pressure. This synthesization route produced nanocrystals in the range of 5 nm.
Assembly and Properties of Nanoparticles 65

Once again, control of growth after nucleation is necessary to obtain desirable


nanoparticles. Rusakova et al. [255] used an interesting approach to control the par-
ticle growth of hydrous metal oxide gels. They showed that the growth could be in-
hibited by replacing the surface hydroxyl group, before the crystallization step, with
a functional group that does not condense and that can produce small secondary-
phase particles, which restrict boundary mobility at high temperatures. These au-
thors reported that fully crystalline SnO2 , TiO2 , and ZrO2 nanocrystals (ranging
in size from 1.5 to 5 nm) can be obtained after heat treating the precipitate gel at
500 C, by replacing the hydroxyl group with the methyl siloxyl group before firing.
The development of metal oxide nanocrystals by nonhydrolytic synthesization
routes results in materials whose surfaces are free of OH groups and produces
nanocrystals with different properties particularly suitable for catalytic and sensor
applications. Several nonhydrolytic processes have been developed to process metal
oxides, and the molecular chemistry of these various methods is discussed by Vioux
[256]. On the basis of the strategy used to process II-IV semiconductor nanocrys-
tals, using the rapid decomposition of molecular precursor in the presence of strong
coordinating agents, Trentler and coworkers [88] proposed an interesting route to
process TiO2 nanocrystals, based on the following reactions:

TiX4 + Ti(OR)4 2TiO2 + RX (53)


TiX4 + 2ROR TiO2 + 4RX (54)

where X is a halide ion (Cl , F , Br , I ) and R an alkyl group. The synthetic


route involved injecting the metal alkoxide (Ti(OR)4 ) into a titanium halide mixed
with trioctylphosphine oxide (TOPO) and a solvent at high temperature (300 C).
Nanoparticles with a mean particle size of 7.3 nm and anatase phase were obtained.
Alivisatos and collaborators [257] demonstrated that transition metal oxide
nanocrystals ( -Fe2 O3 , Mn3 O4 , Cu2 O) could be prepared using a nonhydrolytic
process based on the thermal decomposition of metal Cupferron complexes Mx Cupx
(M = metal ion, Cup = C6 H5 N(NO)O ) in a hot solvent with surfactant. Their re-
sults suggest that a good level of control can be achieved when this approach is used
to process metal oxide nanoparticles.
Camargo et al. [258261] recently developed a new route to synthesize lead-
based perovskite nanoparticles, such as PT [260], PZT [259], PbZrO3 (PZ) [258],
and PbHfO3 (PH) [261]. This method, which apparently involves no hydrolytic
reaction and is carbon and halide-free, is called the oxidant peroxo method (OPM)
because it is based on the oxidationreduction reaction between Pb(II) ion and
water-soluble metal-peroxide complexes with high pH. This process results in an
inorganic amorphous precursor that requires subsequent thermal treatment to pro-
mote crystallization of the desired phase. The low crystallization temperatures
(400450 C for the PT phase) of the amorphous precursor suggest that the OPM
method favors the formation of a homogeneous inorganic compound.
An important nonhydrolytic chemical process is the so-called Pechini process or
in situ polymerizable complex (IPC) [262]. This process is based on the ability of
polycarboxylic acids, particularly citric acid (CA), to form very stable water-soluble
66 C. Ribeiro and E.R. Leite

chelate complexes. Even cations with a high tendency to become hydrolyzed, such
as Ti+4 and Nb+5 , can be chelated by CA in a water solution, preventing the hydrol-
ysis and precipitation of hydrous metal oxide. The CA complex thus formed can
be immobilized in a solid organic resin through a polyesterification reaction with
ethylene glycol (EG). This process leads to the formation of a polymeric precursor
with the cations of interest randomly distributed in a three-dimensional solid lat-
tice, avoiding precipitation or phase segregation during the synthesis of the metal
oxide compound [263]. This method is widely used to process titanates [264268],
niobates [269271] and other kinds of polycationic or single cationic metal oxides
[272, 273]. In the last five years, this method has also proved suitable to process ox-
ide thin films with superior performance [273277]. Using this process, PbTiO3 thin
film [277] and nanometric powder [278], for example, can be synthesized at temper-
atures as low as 450 C, resulting in a metastable cubic PbTiO3 phase. Crystallization
was observed at a temperature at which long-range diffusion had to be constrained,
and the thermodynamic equilibrium configuration was kinetically suppressed [278].
The ability to form complex metal oxides at low crystallization temperatures and
metastable phases is not yet well understood, but it is generally assumed to be as-
sociated with the tendency of a poly-cationic CA complex to develop during the
chelation step in water solution [279], and/or the tendency to form an inorganic
amorphous phase, with a local symmetry close to that of the crystalline phase, dur-
ing the crystallization step [280].
The major problem with this process is maintaining control over the particle size
and morphology. During the crystallization process, it is very difficult to keep the
nucleation and the growth processes separate, resulting in agglomerates made up
of nanocrystals. The particle growth process, which was studied in the final stage
of the crystallization of nanometric powder processed by the IPC method, showed
that growth occurs in two different stages [281]. At heat treatment temperatures of
<800 C, the growth process is associated with the surface diffusion mechanism,
with an activation energy in the range of 4080 kJ mol1 . At a temperature of >
800 C, particle growth is controlled by densification of the agglomerate formed by
nanometric particles and by the neck-size-controlled growth mechanism [282].
Basically, two methodologies have been used to control the particle size of metal
oxides processed by the IPC method. Quinelato et al. [272] demonstrated that the
particle size and morphology of CeO2 -doped ZrO2 could be controlled by con-
trolling the metal/CA ratio. A high concentration of CA leads to smaller particles
with a soft agglomeration. Leite and collaborators [283, 284] showed that the par-
ticle size and morphology of SnO2 could be controlled by the addition of dopants
such as Nb2 O5 and rare earths. The same authors [284] also showed that doped
SnO2 nanocrystals are highly stable against particle growth, even at high temper-
atures. The technique used to achieve this high stability was to process supersat-
urated solid solution between the SnO2 and the dopant. Segregation of the dopant
on the nanocrystal surface occurs during the heat treatment, decreasing the particle
boundary mobility or the surface energy. This approach was originally developed to
control the particle growth of metal nanocrystals [5, 285] and was used successfully
to control the growth of metal oxide nanocrystals.
Assembly and Properties of Nanoparticles 67

6 Summary

As one can see, several properties can be explored on a nanoscale and an introduc-
tory view of the subject was discussed here. Advances in understanding nanoparticle
formation mechanisms and the nature of nanoparticle properties undoubtedly offer
the best pathway for developing viable nanotechnology and for augmenting the ben-
efits of its use.

Acknowledgments The authors gratefully acknowledge the financial support of the Brazilian re-
search funding agencies FAPESP and CNPq.

References

1. L. V. Interrante and M. J. Hampden-Smith. Chemistry of Advanced Materials. Wiley, New


York, NY, 1998
2. A. P. Alivisatos. Perspectives on the physical chemistry of semiconductor nanocrystals.
J. Phys. Chem., 100:1322613239, 1996
3. S. Morup. Nanomagnetism, chapter Studies of Superparamagnetism in Samples of Ultrafine
Particles, pages 9399. Kluwer, Boston, MA, 1993
4. K. OGrady and R. W. Chantrell. Magnetic Properties of Fine Particles, chapter Magnetic
Properties: Theory and Experiments. Elsevier, Amsterdan, 1992
5. H. Gleiter. Nanostructured materials: Basic concepts and microstructure. Acta Materialia,
48:129, 2000
6. M. Gratzel. Photoelectrochemical cells. Nature, 414:338344, 2001
7. P. Mulvaney. Not all thats gold does glitter. MRS Bull., 26:10091014, 2001
8. M. Faraday. The bakerian lecture: Experimental relations of gold (and other metals) to light.
Philos. Trans. R. Soc., 147:145181, 1857
9. B. Oregan and M. Gratzel. A low-cost, high-efficiency solar-cell based on dye-sensitized
colloidal TiO2 films. Nature, 353:737740, 1991
10. W. J. Li, H. Osora, L. Otero, D. C. Duncan, and M. A. Fox. Photoelectrochemistry of a
substituted-Ru(bpy)(3)(2+)-labeled polyimide and nanocrystalline SnO2 composite formu-
lated. J. Phys. Chem. A, 102:53335340, 1998
11. I. Bedja P. V., Kamat X., Hua A., Lappin G., and Hotchandani S. Photosensitiza-
tion of nanocrystalline ZnO films by bis(2, 2 -bipyridine)(2, 2 -bipyridine-4,4 -dicarboxylic
acid)ruthenium(ii). Langmuir, 13:23982403, 1997
12. L. Kavan, K. Kratochvilova, and M. Gratzel. Study of nanocrystalline TiO2 (anatase) elec-
trode in the accumulation regime. J. Electroanal. Chem., 394:93102, 1995
13. F. Croce, G. B. Appetecchi, L. Persi, and B. Scrosati. Nanocomposite polymer electrolytes
for lithium batteries. Nature, 394:456458, 1998
14. U. Bach, D. Lupo, P. Comte, J. E. Moser, F. Weissortel, J. Salbeck, H. Spreitzer, and
M. Gratzel. Solid-state dye-sensitized mesoporous TiO2 solar cells with high photon-to-
electron conversion efficiencies. Nature, 395:583585, 1998
15. A. Hagfeldt and M. Gratzel. Light-induced redox reactions in nanocrystalline systems. Chem.
Rev., 95:4968, 1995
16. N. Yamazoe. New approaches for improving semiconductor gas sensors. Sens. Actuators
B-Chem., 5:719, 1991
17. D. Weller and A. Moser. Thermal effect limits in ultrahigh-density magnetic recording. IEEE
Trans. Magn., 35:44234439, 1999
68 C. Ribeiro and E.R. Leite

18. S. H. Sun, C. B. Murray, D. Weller, L. Folks, and A. Moser. Monodisperse FePt nanoparticles
and ferromagnetic FePt nanocrystal superlattices. Science, 287:19891992, 2000
19. J. B. Dai, J. K. Tang, S. T. Hsu, and W. Pan. Magnetic nanostructures and materials in mag-
netic random access memory. J. Nanosci. Nanotechnol., 2:281291, 2002
20. W. Chen, D. Grouquist, and J. Roark. Voltage tunable electroluminescence of CdTe nanopar-
ticle light-emitting diodes. J. Nanosci. Nanotechnol., 2:4753, 2002
21. V. L. Colvin, M. C. Schlamp, and A. P. Alivisatos. Light-emitting-diodes made from cad-
mium selenide nanocrystals and a semiconducting polymer. Nature, 370:354357, 1994
22. C. Feldmann. Preparation of nanoscale pigment particles. Adv. Mater., 13:13011303, 2001
23. Y. M. Chiang, D. P. Birnie III, and W. D. Kingery. Physical Ceramics. Wiley, New York, NY,
1997
24. B. Gilbert, F. Huang, H. Z. Zhang, G. A. Waychunas, and J. F. Banfield. Nanoparticles:
Strained and stiff. Science, 305:651654, 2004
25. K. K. Nanda, S. N. Sahu, and S. N. Behera. Liquid-drop model for the size-dependent melting
of low-dimensional systems. Phys. Rev. A, 66:01320818, 2002
26. C. J. Coombes. The melting of small particles of lead and indium. J. Phys. F: Metal. Phys.,
2:441449, 1972
27. P. Buffat and J. P. Borel. Size effect on melting temperature of gold particles. Phys. Rev. A,
13:22872298, 1976
28. T. Castro, R. Reifenberger, E. Choi, and R. P. Andres. Size-dependent melting temperature
of individual nanometer-sized metallic clusters. Phys. Rev. B, 42:85488556, 1990
29. C. L. Cleveland and U. Landman. The energetics and structure of nickel clusters size de-
pendence. J. Chem. Phys., 94:73767396, 1991
30. F. Ercolessi, W. Andreoni, and E. Tosatti. Melting of small gold particles mechanism and
size effects. Phys. Rev. Lett., 66:911914, 1991
31. A. N. Goldstein, C. M. Echer, and A. P. Alivisatos. Melting in semiconductor nanocrystals.
Science, 256:14251427, 1992
32. T. P. Martin, U. Naher, H. Schaber, and U. Zimmermann. Evidence for a size-dependent
melting of sodium clusters. J. Chem. Phys., 100:23222324, 1994
33. J. B. Mitchel and J. T Schwartz. Preparation of Catalysts IV: Scientific Bases for the Prepa-
ration of Heterogeneous Catalysts (Studies in Surface Science and Catalysis), volume 31,
chapter Catalyst preparation science. Elsevier, Amsterdan, 1987
34. J. Turkevich and G. Kim. Palladium preparation and catalytic properties of particles of
uniform size. Science, 169:873, 1970
35. H. Hirai, Y. Nakao, and N. Toshima. Colloidal rhodium in polyvinyl-alcohol as hydrogena-
tion catalyst of olefins. Chem. Lett., 1976:905910, 1976
36. D. N. Furlong, A. Launikonis, W. H. F. Sasse, and J. V. Sanders. Colloidal platinum sols
preparation, characterization and stability towards salt. J. Chem. Soc.-Farad. Trans. I, 80:571,
1984
37. T. Teranishi and M. Miyake. Size control of palladium nanoparticles and their crystal struc-
tures. Chem. Mater., 10:594600, 1998
38. R. Narayanan and M. A. El-Sayed. Catalysis with transition metal nanoparticles in colloidal
solution: Nanoparticle shape dependence and stability. J. Phys. Chem. B, 109:1266312676,
2005
39. C. Burda, X. B. Chen, R. Narayanan, and M. A. El-Sayed. Chemistry and properties of
nanocrystals of different shapes. Chem. Rev., 105:10251102, 2005
40. R. Narayanan and M. A. El-Sayed. Shape-dependent catalytic activity of platinum nanopar-
ticles in colloidal solution. Nano Lett., 4:13431348, 2004
41. T. S. Ahmadi, Z. L. Wang, T. C. Green, A. Henglein, and M. A. ElSayed. Shape-controlled
synthesis of colloidal platinum nanoparticles. Science, 272:19241926, 1996
42. J. S. Bradley. Clusters and Colloids: From Theory to Applications, chapter The Chemistry of
Transition Metal Colloids, pages 459544. VCH, Weinheim, 1994
43. E. R. Leite. Encyclopedia of Nanoscience and Nanotechnology, chapter Nanocrystals assem-
bled from bottom-up, pages 537550. American Scientific Publishers, 2004
Assembly and Properties of Nanoparticles 69

44. R. J. Pugh. Surface and Colloid Chemistry in Advanced Ceramics Processing, chapter Dis-
persion and Stability of Ceramic Powders in Liquids, pages 127192. Marcel Dekker, NY,
New York, 1994
45. G. A. Parks. Isoelectric points of solid oxides solid hydroxides and aqueous hydroxo complex
systems. Chem. Rev., 65:177, 1965
46. Y. Yin and A. P. Alivisatos. Colloidal nanocrystal synthesis and the organic-inorganic inter-
face. Nature, 437:664670, 2005
47. X. Wang, J. Zhuang, Q. Peng, and Y. Li. A general strategy for nanocrystal synthesis. Nature,
473:121124, 2005
48. Z. Y. Tang, N. A. Kotov, and M. Giersig. Spontaneous organization of single CdTe nanopar-
ticles into luminescent nanowires. Science, 297:237240, 2002
49. Y. Volkov, S. Mitchell, N. Gaponik, Y. P. Rakovich, J. F. Donegan, D. Kelleher, and
A. L. Rogach. In-situ observation of nanowire growth from luminescent cdte nanocrystal
in a phosphate buffer solution. Chem. Phys. Chem., 5:16001602, 2004
50. K. S. Cho, D. V. Talapin, W. Gaschler, and C. B. Murray. Designing PbSe nanowires and
nanorings through oriented attachment of nanoparticles. J. Am. Chem. Soc., 127:71407147,
2005
51. N. J. Wagner and R. Klein. The rheology and microstructure of charged colloidal suspen-
sions. Colloid Polym. Sci., 269:295319, 1991
52. J. Nemeth and I. Dekany. The effect of nanoparticle growth on rheological properties of silica
and silicate dispersions. Colloid Polym Sci., 278:211219, 2000
53. V. Tohver, A. Chan, O. Sakurada, and J. A. Lewis. Nanoparticle engineering of complex fluid
behavior. Langmuir, 17:84148421, 2001
54. V. Tohver, J. E. Smay, A. Braem, P. V. Braun, and J. A. Lewis. Nanoparticle halos: A new
colloid stabilization mechanism. Proc. Natl Acad. Sci. U S A., 98:89508954, 2001
55. I. N. Levine. Quantum Chem. Prentice Hall, 5th edition, 1999
56. S. Folsch, P. Hyldgaard, R. Koch, and K. H. Ploog. Quantum confinement in monatomic Cu
chains on cu(111). Phys. Rev. Lett., 92:05680314, 2004
57. N. Nilius, T. M. Wallis, and W. Ho. Realization of a particle-in-a-box: Electron in an atomic
pd chain. J. Phys. Chem. B, 109:2065720660, 2005
58. W. P. Halperin. Quantum size effects in metal particles. Rev. Mod. Phys., 58:533606, 1986
59. A. Kawabata and R. Kubo. Electronic properties of fine metallic particles.2. plasma reso-
nance absorption. J. Phys. Soc. Jpn., 21:1765, 1966
60. R. Kubo, A. Kawabata, and S. Kobayashi. Electronic-properties of small particles. Ann. Rev.
Mater. Sci., 14:4966, 1984
61. L. E. Brus. A simple-model for the ionization-potential, electron-affinity, and aqueous redox
potentials of small semiconductor crystallites. J. Chem. Phys., 79:55665571, 1983
62. L. E. Brus. Electron electron and electron-hole interactions in small semiconductor crystal-
lites the size dependence of the lowest excited electronic state. J. Chem. Phys., 80:4403
4409, 1984
63. L. Brus. Electronic wave-functions in semiconductor clusters experiment and theory. J.
Phys. Chem., 90:25552560, 1986
64. S. V. Gaponenko. Optical Properties of Semiconductor Nanocrystals. Cambridge University
Press, Cambridge, 1998
65. R. Rossetti, S. Nakahara, and L. E. Brus. Quantum size effects in the redox potentials, reso-
nance Raman-spectra, and electronic-spectra of cds crystallites in aqueous-solution. J. Chem.
Phys., 79:10861088, 1983
66. R. Rossetti, J. L. Ellison, J. M. Gibson, and L. E. Brus. Size effects in the excited electronic
states of small colloidal cds crystallites. J. Chem. Phys., 80:44644469, 1984
67. D. W. Bahnemann, C. Kormann, and M. R. Hoffmann. Preparation and characterization of
quantum size zinc-oxide a detailed spectroscopic study. J. Phys. Chem., 91:37893798,
1987
68. A. P. Alivisatos, A. L. Harris, N. J. Levinos, M. L. Steigerwald, and L. E. Brus. Electronic
states of semiconductor clusters homogeneous and inhomogeneous broadening of the
optical-spectrum. J. Chem. Phys., 89:40014011, 1988
70 C. Ribeiro and E.R. Leite

69. C. B. Murray, D. J. Norris, and M. G. Bawendi. Synthesis and characterization of nearly


monodisperse cde (e = s, se, te) semiconductor nanocrystallites. J. Am. Chem. Soc.,
115:87068715, 1993
70. X. G. Peng, J. Wickham, and A. P. Alivisatos. Kinetics of IIVI and IIIV colloidal semi-
conductor nanocrystal growth: Focusing of size distributions. J. Am. Chem. Soc., 120:5343
5344, 1998
71. G. Oskam, Z. S. Hu, R. L. Penn, N. Pesika, and P. C. Searson. Coarsening of metal oxide
nanoparticles. Phys. Rev. E, 66:01140314, 2002
72. N. S. Pesika, Z. S. Hu, K. J. Stebe, and P. C. Searson. Quenching of growth of ZnO nanopar-
ticles by adsorption of octanethiol. J. Phys. Chem. B, 106:69856990, 2002
73. N. Chiodini, A. Paleari, D. DiMartino, and G. Spinolo. SNO2 nanocrystals in SiO2: A wide-
band-gap quantum-dot system. Appl. Phys. Lett., 81:17021704, 2002
74. Z. S. Hu, G. Oskam, R. L. Penn, N. Pesika, and P. C. Searson. The influence of anion on the
coarsening kinetics of ZnO nanoparticles. J. Phys. Chem. B, 107:31243130, 2003
75. Z. S. Hu, G. Oskam, and P. C. Searson. Influence of solvent on the growth of ZnO nanopar-
ticles. J. Colloid Interface Sci., 263:454460, 2003
76. G. Oskam, A. Nellore, R. L. Penn, and P. C. Searson. The growth kinetics of TiO2 nanopar-
ticles from titanium(IV) alkoxide at high water/titanium ratio. J. Phys. Chem. B, 107:
17341738, 2003
77. N. S. Pesika, K. J. Stebe, and P. C. Searson. Determination of the particle size distribution of
quantum nanocrystals from absorbance spectra. Adv. Mater., 15:1289, 2003
78. N. S. Pesika, K. J. Stebe, and P. C. Searson. Relationship between absorbance spectra and
particle size distributions for quantum-sized nanocrystals. J. Phys. Chem. B, 107:10412
10415, 2003
79. E. J. H. Lee, C. Ribeiro, T. R. Giraldi, E. Longo, E. R. Leite, and J. A. Varela. Photolumines-
cence in quantum-confined SnO2 nanocrystals: Evidence of free exciton decay. Appl. Phys.
Lett., 84:17451747, 2004
80. W. E. Buhro and V. L. Colvin. Semiconductor nanocrystals Shape matters. Nat. Mater.,
2:138139, 2003
81. A. A. Kokhanovsky and E. P. Zege. Optical properties of aerosol particles: A review of
approximate analytical solutions. J. Aerosol Sci., 28:121, 1997
82. G. Blasse and B. C. Grabmaier. Luminescent Materials. Springer, Berlin, 1 edition, 1994
83. C. Ribeiro, E. J. H. Lee, T. R. Giraldi, E. Longo, and E. R. Leite. Internal report, Federal
University of Sao Carlos, unpublished work, 2003
84. A. Navrotsky and O. J. Kleppa. Enthalpy of anatase-rutile transformation. J. Am. Ceramic
Soc., 50:626, 1967
85. E. A. Barringer and H. K. Bowen. High-purity, monodisperse TiO2 powders by hydrolysis
of titanium tetraethoxide.1. synthesis and physical-properties. Langmuir, 1:414420, 1985
86. J. H. Jean and T. A. Ring. Nucleation and growth of monosized TiO2 powders from alcohol
solution. Langmuir, 2:251255, 1986
87. T. E. Mates and T. A. Ring. Steric stability of alkoxy-precipitated TiO2 in alcohol-solutions.
Colloids Surf., 24:299313, 1987
88. T. J. Trentler, T. E. Denler, J. F. Bertone, A. Agrawal, and V. L. Colvin. Synthesis of TiO2
nanocrystals by nonhydrolytic solution-based reactions. J. Am. Chem. Soc., 121:16131614,
1999
89. G. Garnweitner, M. Antonietti, and M. Niederberger. Nonaqueous synthesis of crystalline
anatase nanoparticles in simple ketones and aldehydes as oxygen-supplying agents. Chem.
Comm., 3:397399, 2005
90. J. Ragai and W. Lotfi. Effect of preparative ph and aging media on the crystallographic trans-
formation of amorphous TiO2 to anatase and rutile. Colloids Surf., 61:97109, 1991
91. S. Han, S. H. Choi, S. S. Kim, M. Cho, B. Jang, D. Y. Kim, J. Yoon, and T. Hyeon. Low-
temperature synthesis of highly crystalline TiO2 nanocrystals and their application to photo-
catalysis. Small, 1:812816, 2005
92. H. Z. Zhang and J. F. Banfield. Kinetics of crystallization and crystal growth of nanocrys-
talline anatase in nanometer-sized amorphous titania. Chem. Mater., 14:41454154, 2002
Assembly and Properties of Nanoparticles 71

93. M. R. Ranade, A. Navrotsky, H. Z. Zhang, J. F. Banfield, S. H. Elder, A. Zaban, P. H. Borse,


S. K. Kulkarni, G. S. Doran, and H. J. Whitfield. Energetics of nanocrystalline TiO2 . Proc.
Natl Acad. Sci. U S A, 99:64766481, 2002
94. A. Navrotsky. Energetic clues to pathways to biomineralization: Precursors, clusters, and
nanoparticles. Proc. Natl Acad. Sci. U S A, 101:1209612101, 2004
95. A. S. Barnard and P. Zapol. Predicting the energetics, phase stability, and morphology evo-
lution of faceted and spherical anatase nanocrystals. J. Phys. Chem. B, 108:1843518440,
2004
96. A. S. Barnard and P. Zapol. Effects of particle morphology and surface hydrogenation on the
phase stability of TiO2 . Phys. Rev. B, 70:235403113, 2004
97. A. Paul. Chemistry of Glasses. Chapman and Hall, London, 1982
98. B. L. Cushing, V. L. Kolesnichenko, and C. J. OConnor. Recent advances in the liquid-phase
syntheses of inorganic nanoparticles. Chem. Rev., 104:38933946, 2004
99. S. V. Ushakov and A. Navrotsky. Direct measurements of water adsorption enthalpy on hafnia
and zirconia. App. Phys. Lett., 87:164103, 2005
100. C. J. Brinker and G. W. Scherrer. SolGel Science. Academic, Boston, MA, 1990
101. D. V. Talapin, A. L. Rogach, M. Haase, and H. Weller. Evolution of an ensemble of nanopar-
ticles in a colloidal solution: Theoretical study. J. Phys. Chem. B, 105:1227812285, 2001
102. H. Z. Zhang, F. Huang, B. Gilbert, and J. F. Banfield. Molecular dynamics simulations, ther-
modynamic analysis, and experimental study of phase stability of zinc sulfide nanoparticles.
J. Phys. Chem. B, 107:1305113060, 2003
103. H. Z. Zhang and J. F. Banfield. Aggregation, coarsening, and phase transformation in zns
nanoparticles studied by molecular dynamics simulations. Nano Lett., 4:713718, 2004
104. F. Huang, B. Gilbert, H. H. Zhang, and J. F. Banfield. Reversible, surface-controlled structure
transformation in nanoparticles induced by an aggregation state. Phys. Rev. Lett., 92:15550
14, 2004
105. F. Huang, B. Gilbert, H. Zhang, M. P. Finnegan, J. R. Rustad, C. S. Kim, G. A. Waychunas,
and J. F. Banfield. Interface interactions in nanoparticle aggregates. Geochim. Cosmochim.
Acta, 68:A222A222, 2004
106. H. Z. Zhang and J. F. Banfield. Size dependence of the kinetic rate constant for phase trans-
formation in TiO2 nanoparticles. Chem. Mater., 17:34213425, 2005
107. F. Huang and J. F. Banfield. Size-dependent phase transformation kinetics in nanocrystalline
zns. J. Am. Chem. Soc., 127:45234529, 2005
108. S. Toschev. Crystal Growth an introduction, chapter Homogeneous nucleation, pages 149.
Elsevier, Amsterdam 1973
109. A. S. Barnard and P. Zapol. A model for the phase stability of arbitrary nanoparticles as a
function of size and shape. J. Chem. Phys., 121:42764283, 2004
110. S. H. Tolbert and A. P. Alivisatos. Size dependence of a first-order solid-solid phase-
transition the wurtzite to rock-salt transformation in CdSe nanocrystals. Science, 265:
373376, 1994
111. S. H. Tolbert and A. P. Alivisatos. The wurtzite to rock-salt structural transformation in CdSe
nanocrystals under high-pressure. J. Chem. Phys., 102:46424656, 1995
112. C. Herring. Some theorems on the free energies of crystal surfaces. Phys. Rev., 82:8793,
1951
113. A. Beltran, J. Andres, E. Longo, and E. R. Leite. Thermodynamic argument about SnO2
nanoribbon growth. Appl. Phys. Lett., 83:635637, 2003
114. J. Oviedo and M. J. Gillan. Energetics and structure of stoichiometric SnO2 surfaces studied
by first-principles calculations. Surf. Sci., 463:93101, 2000
115. P. A. Mulheran and J. H. Harding. The stability of SnO2 surfaces. Model. Simul. Mater. Sci.
Eng., 1:3943, 1992
116. B. Slater, C. R. A. Catlow, D. H. Gay, D. E. Williams, and V. Dusastre. Study of surface seg-
regation of antimony on SnO2 surfaces by computer simulation techniques. J. Phys. Chem.
B, 103:1064410650, 1999
72 C. Ribeiro and E.R. Leite

117. M. J. Yacaman, J. A. Ascencio, H. B. Liu, and J. Gardea-Torresdey. Structure shape and


stability of nanometric sized particles. J. Vac. Sci. Technol. B-an Int. J. Devoted to Micro-
electron. Nanometer Struct.-Process Meas. Phenomena, 19:10911103, 2001
118. B. Gilbert, H. Z. Zhang, F. Huang, M. P. Finnegan, G. A. Waychunas, and
J. F. Banfield. Special phase transformation and crystal growth pathways observed in
nanoparticles. Geochem. Trans., 4:2027, 2003
119. B. Gilbert, H. Z. Zhang, F. Huang, J. F. Banfield, Y. Ren, D. Haskel, J. C. Lang, G. Srajer,
A. Jurgensen, and G. A. Waychunas. Analysis and simulation of the structure of nanoparticles
that undergo a surface-driven structural transformation. J. Chem. Phys., 120:1178511795,
2004
120. H. Z. Zhang, B. Gilbert, F. Huang, and J. F. Banfield. Water-driven structure transformation
in nanoparticles at room temperature. Nature, 424:10251029, 2003
121. A. S. Barnard, P. Zapol, and L. A. Curtiss. Modeling the morphology and phase stability of
TiO2 nanocrystals in water. J. Chem. Theory Comput., 1:107116, 2005
122. P. P. von Weimarn. The precipitation laws. Chem. Rev., 2:217242, 1925
123. M. Zurita-Gotor and D. E. Rosner. Aggregate size distribution evolution for brownian co-
agulation sensitivity to an improved rate constant. J. Colloid Interface Sci., 274:502514,
2004
124. B. Mutaftschiev. Handbook of Crystal Growth, chapter Nucleation, page 187. Elsevier Sci-
ence, Amsterdam, 1993
125. B. K. Chakraverty. Crystal Growth An Introduction, chapter Heterogeneous nucleation and
condensation on substrates, pages 50104. Elsevier, Amsterdam, 1973
126. G. H. Weiss. Overview of theoretical models for reaction rates. J. Stat. Phys., 42:336, 1986
127. D. P. Dinega and M. G. Bawendi. A solution-phase chemical approach to a new crystal
structure of cobalt. Angewandte Chemie-international Edition, 38:17881791, 1999
128. P. Jensen. Growth of nanostructures by cluster deposition: Experiments and simple models.
Rev. Mod. Phys., 71:16951735, 1999
129. M. Aoun, E. Plasari, R. David, and J. Villermaux. A simultaneous determination of nucle-
ation and growth rates from batch spontaneous precipitation. Chem. Eng. Sci., 54:11611180,
1999
130. R. Zauner and A. G. Jones. Determination of nucleation, growth, agglomeration and disrup-
tion kinetics from experimental precipitation data: the calcium oxalate system. Chem. Eng.
Sci., 55:42194232, 2000
131. S. A. Kukushkin and S. V. Nemna. The effect of pH on nucleation kinetics in solutions.
Doklady Phys. Chem., 377(6):792796, 2000
132. E. A. Meulenkamp. Size dependence of the dissolution of ZnO nanoparticles. J. Phys. Chem.
B, 102:77647769, 1998
133. G. W. Greenwood. The growth of dispersed precipitates in solutions. Acta Metallurgica,
4:243248, 1956
134. S. A. Kukushkin and A. V. Osipov. Crystallization of binary melts and decay of supersatu-
rated solid solutions at the ostwald ripening stage under non-isothermal conditions. J. Phys.
Chem. Solids, 56:12591269, 1995
135. E. M. Wong, J. E. Bonevich, and P. C. Searson. Growth kinetics of nanocrystalline ZnO
particles from colloidal suspensions. J. Phys. Chem. B, 102:77707775, 1998
136. E. M. Wong, P. G. Hoertz, C. J. Liang, B. M. Shi, G. J. Meyer, and P. C. Searson. Influence
of organic capping ligands on the growth kinetics of ZnO nanoparticles. Langmuir, 17:8362
8367, 2001
137. F. Huang, H. Z. Zhang, and J. F. Banfield. Two-stage crystal-growth kinetics observed during
hydrothermal coarsening of nanocrystalline ZnS. Nano Lett., 3:373378, 2003
138. I. M. Lifshitz and V. V. Slyozov. The kinetics of precipitation from supersaturated solid
solutions. J. Phys. Chem. Solids, 19:3550, 1961
139. C. Wagner. Theorie der alterung von niederschlagen durch umlosen (ostwald-reifung).
Z. Elektrochem., 65:581591, 1961
140. H. Gratz. Ostwald ripening: New relations betwenn particle growth and particle size distri-
bution. Scrip. Mater., 37:916, 1997
Assembly and Properties of Nanoparticles 73

141. R. L. Penn, G. Oskam, T. J. Strathmann, P. C. Searson, A. T. Stone, and D. R. Veblen.


Epitaxial assembly in aged colloids. J. Phys. Chem. B, 105:21772182, 2001
142. J. F. Banfield, S. A. Welch, H. Z. Zhang, T. T. Ebert, and R. L. Penn. Aggregation-based crys-
tal growth and microstructure development in natural iron oxyhydroxide biomineralization
products. Science, 289:751754, 2000
143. X. G. Peng, L. Manna, W. D. Yang, J. Wickham, E. Scher, A. Kadavanich, and
A. P. Alivisatos. Shape control of CdSe nanocrystals. Nature, 404:5961, 2000
144. R. L. Penn and J. F. Banfield. Morphology development and crystal growth in nanocrystalline
aggregates under hydrothermal conditions: Insights from titania. Geochim. Cosmochim. Acta,
63:15491557, 1999

145. E. J. H. Lee. Sntese e caracterizaca o de nanopartculas de Oxido de estanho obtidas a partir
de suspensoes coloidais. Masters thesis, Univ. Fed. S. Carlos, 2004
146. M. Nespolo and G. Ferraris. The oriented attachment mechanims in the formation of twins
a survey. Eur. J. Mineral., 16:401406, 2004
147. M. Nespolo, G. Ferraris, S. Durovic, and Y. Takeuchi. Twins vs. modular crystal structures.
Z. Kristallogr., 219:773778, 2004
148. M. Nespolo and G. Ferraris. Applied geminography symmetry analysis of twinned crystals
and definition of twinning by reticular polyholohedry. Acta Cryst. A, 60:8995, 2004
149. M. Nespolo and G. Ferraris. Hybrid twinning a cooperative type of oriented crystal associ-
ation. Z. Kristallogr., 220:317323, 2005
150. R. L. Penn and J. F. Banfield. Oriented attachment and growth, twinning, polytypism, and
formation of metastable phases: Insights from nanocrystalline TiO2 . Am. Mineralogist, 83:
10771082, 1998
151. Y. Chushkin, M. Ulmeanu, S. Luby, E. Majkova, I. Kostic, P. Klang, V. Holy, Z. Bochnicek,
M. Giersig, M. Hilgendorff, and T. H. Metzger. Structural study of self-assembled Co
nanoparticles. J. Appl. Phys., 94:77437748, 2003
152. J. Polleux, N. Pinna, M. Antonietti, and M. Niederberger. Ligand-directed assembly of pre-
formed titania nanocrystals into highly anisotropic nanostructures. Adv. Mater., 16:436439,
2004
153. J. Polleux, N. Pinna, M. Antonietti, C. Hess, U. Wild, R. Schlogl, and M. Niederberger.
Ligand functionality as a versatile tool to control the assembly behavior of preformed titania
nanocrystals. Chem.-a Europ. J., 11:35413551, 2005
154. E. J. H. Lee, C. Ribeiro, E. Longo, and E. R. Leite. Oriented attachment: An effective mech-
anism in the formation of anisotropic nanocrystals. J. Phys. Chem. B, 109:2084220846,
2005
155. W. W. Mullins. 2-dimensional motion of idealized grain boundaries. J. Appl. Phys., 27:900
904, 1956
156. P. Feltham. Grain growth in metals. Acta Metallurgica, 5:97105, 1957
157. M. Hillert. On theory of normal and abnormal grain growth. Acta Metallurgica, 13:227, 1965
158. N. P. Louat. Theory of normal grain-growth. Acta Metallurgica, 22:721724, 1974
159. G. Herrmann, H. Gleiter, and G. Baro. Investigation of low-energy grain-boundaries in metals
by a sintering technique. Acta. Met., 24:353359, 1976
160. H. Sautter, H. Gleiter, and G. Baro. Effect of solute atoms on energy and structure of grain-
boundaries. Acta. Met., 25:467473, 1977
161. U. Erb and H. Gleiter. Effect of temperature on the energy and structure of grain-boundaries.
Scrip. Metall., 13:6164, 1979
162. H. Kuhn, G. Baero, and H. Gleiter. Energy-misorientation relationship of grain-boundaries.
Acta. Met., 27:959963, 1979.
163. D. Moldovan, D. Wolf, and S. R. Phillpot. Theory of diffusion-accommodated grain rotation
in columnar polycrystalline microstructures. Acta Materialia, 49:35213532, 2001
164. D. Moldovan, V. Yamakov, D. Wolf, and S. R. Phillpot. Scaling behavior of grain-rotation-
induced grain growth. Phys. Rev. Lett., 89:2061011, 2002
165. D. Moldovan, D. Wolf, S. R. Phillpot, and A. J. Haslam. Role of grain rotation during
grain growth in a columnar microstructure by mesoscale simulation. Acta Materialia, 50:
33973414, 2002
74 C. Ribeiro and E.R. Leite

166. A. J. Haslam, D. Moldovan, V. Yamakov, D. Wolf, S. R. Phillpot, and H. Gleiter. Stress-


enhanced grain growth in a nanocrystalline material by molecular-dynamics simulation. Acta
Materialia, 51:20972112, 2003
167. H. L. Zhu and R. S. Averback. Sintering of nano-particle powders: Simulations and experi-
ments. Mater. Manuf. Process., 11:905923, 1996
168. H. L. Zhu and R. S. Averback. Sintering processes of two nanoparticles: A study by
molecular-dynamics. Philos. Mag. Lett., 73:2733, 1996
169. M. Yeadon, J. C. Yang, R. S. Averback, J. W. Bullard, D. L. Olynick, and J. M. Gibson. In-situ
observations of classical grain growth mechanisms during sintering of copper nanoparticles
on (001) copper. Appl. Phys. Lett., 71:16311633, 1997
170. M. Yeadon, M. Ghaly, J. C. Yang, R. S. Averback, and J. M. Gibson. Contact epitaxy ob-
served in supported nanoparticles. Appl. Phys. Lett., 73:32083210, 1998
171. K. E. Harris, V. V. Singh, and A. H. King. Grain rotation in thin films of gold. Acta Materi-
alia, 46:26232633, 1998
172. C. Ribeiro, E. J. H. Lee, T. R. Giraldi, R. Aguiar, E. Longo, and E. R. Leite. In situ oriented
crystal growth in a ceramic nanostructured system. J. Appl. Phys., 97:02431314, 2005
173. S. A. Kukushkin and A. V. Osipov. New phase formation on solid surfaces and thin film
condensation. Prog. Surf. Sci., 5:1107, 1996
174. F. Huang, H. Z. Zhang, and J. F. Banfield. The role of oriented attachment crystal growth in
hydrothermal coarsening of nanocrystalline ZnS. J. Phys. Chem. B, 107:1047010475, 2003
175. R. L. Penn. Kinetics of oriented aggregation. J. Phys. Chem. B, 108:1270712712, 2004
176. C. Ribeiro, E. J. H. Lee, T. R. Giraldi, J. A. Varela, E. Longo, and E. R. Leite. Study of
synthesis variables in the nanocrystal growth behavior of tin oxide processed by controlled
hydrolysis. J. Phys. Chem. B, 108:1561215617, 2004
177. M. von Smoluchowski. Versuch einer mathematischen theorie der koagulationkinetik kol-
lider losungen. Z. Phys. Chem., Stoechiom. Verwandtschaftsl, 29:129168, 1917
178. D. Mozyrsky and V. Privman. Diffusional growth of colloids. J. Chem. Phys., 110:9254
9258, 1999
179. N. J. Wagner. The smoluchowski equation for colloidal suspensions developed and analyzed
through the generic formalism. J. Non-newtonian Fluid Mech., 96:177201, 2001
180. O. A. Linnikov. Spontaneous crystallization of potassium chloride from aqueous and
aqueous-ethanol solutions part 1: Kinetics and mechanism of the crystallization process.
Cryst. Res. Technol., 39:516528, 2004
181. O. A. Linnikov. Spontaneous crystallization of potassium chloride from aqueous and
aqueous-ethanol solutions part 2: Mechanism of aggregation and coalescence of crystals.
Cryst. Res. Technol., 39:529539, 2004
182. T. Arita, O. Kajimoto, M. Terazima, and Y. Kimura. Experimental verification of the smolu-
chowski theory for a bimolecular diffusion-controlled reaction in liquid phase. J. Chem.
Phys., 120:70717074, 2004
183. J. H. Yu, J. Joo, H. M. Park, S. I. Baik, Y. W. Kim, S. C. Kim, and T. Hyeon. Synthesis of
quantum-sized cubic ZnS nanorods by the oriented attachment mechanism. J. Am. Chem.
Soc., 127:56625670, 2005
184. D. V. Talapin, E. V. Shevchenko, C. B. Murray, A. Kornowski, S. Forster, and H. Weller.
Cdse and CdSe/CdS nanorod solids. J. Am. Chem. Soc., 126:1298412988, 2004
185. M. Adachi, Y. Murata, J. Takao, J. T. Jiu, M. Sakamoto, and F. M. Wang. Highly efficient
dye-sensitized solar cells with a titania thin-film electrode composed of a network structure
of single-crystal-like TiO2 nanowires made by the oriented attachment mechanism. J. Am.
Chem. Soc., 126:1494314949, 2004
186. L. Vayssieres, C. Chaneac, E. Tronc, and J. P. Jolivet. Size tailoring of magnetite particles
formed by aqueous precipitation: An example of thermodynamic stability of nanometric ox-
ide particles. J. Colloid Interface Sci., 205:205212, 1998
187. L. Vayssieres, A. Hagfeldt, and S. E. Lindquist. Purpose-built metal oxide nanomaterials.
The emergence of a new generation of smart materials. Pure Appl. Chem., 72:4752, 2000
188. L. Vayssieres, K. Keis, A. Hagfeldt, and S. E. Lindquist. Three-dimensional array of highly
oriented crystalline ZnO microtubes. Chem. Mater., 13:43954398, 2001
Assembly and Properties of Nanoparticles 75

189. L. Vayssieres, K. Keis, S. E. Lindquist, and A. Hagfeldt. Purpose-built anisotropic metal


oxide material: 3D highly oriented microrod array of zno. J. Phys. Chem. B, 105:33503352,
2001
190. L. Vayssieres, N. Beermann, S. E. Lindquist, and A. Hagfeldt. Controlled aqueous chemi-
cal growth of oriented three-dimensional crystalline nanorod arrays: Application to iron(III)
oxides. Chem. Mater., 13:233235, 2001
191. L. Vayssieres, L. Rabenberg, and A. Manthiram. Aqueous chemical route to ferromagnetic
3-d Arrays of iron nanorods. Nano Lett., 2:13931395, 2002
192. L. Vayssieres and A. Manthiram. 2-d mesoparticulate arrays of alpha-Cr2 O3 . J. Phys. Chem.
B, 107:26232625, 2003
193. L. Vayssieres. Growth of arrayed nanorods and nanowires of ZnO from aqueous solutions.
Adv. Mater., 15:464466, 2003
194. L. Vayssieres and M. Graetzel. Highly ordered SnO2 nanorod arrays from controlled aqueous
growth. Angewandte Chemie-international Edition, 43:36663670, 2004
195. L. Vayssieres, C. Sathe, S. M. Butorin, D. K. Shuh, J. Nordgren, and J. H. Guo. One-
dimensional quantum-confinement effect in alpha-Fe2 O3 ultrafine nanorod arrays. Adv.
Mater., 17:2320, 2005
196. J. Turkevich, P. C. Stevenson, and J. Hillier. A study of the nucleation and growth processes
in the synthesis of colloidal gold. Discuss. Farad. Soc., 11:55, 1951
197. H. Bonnemann and R. M. Richards. Nanoscopic metal particles Synthetic methods and
potential applications. Europ. J. Inorg. Chem., 10:24552480, 2001
198. E. Bourgeat-Lami. Organic-inorganic nanostructured colloids. J. Nanosci. Nanotechnol.,
2:124, 2002
199. C. Sanchez, G. J. D. A. Soler-Illia, F. Ribot, T. Lalot, C. R. Mayer, and V. Cabuil.
Designed hybrid organic-inorganic nanocomposites from functional nanobuilding blocks.
Chem. Mater., 13:30613083, 2001
200. J. S. Bradley, J. M. Millar, and E. W. Hill. Surface-chemistry on colloidal metals a high-
resolution nuclear-magnetic-resonance study of carbon-monoxide adsorbed on metallic pal-
ladium crystallites in colloidal suspension. J. Am. Chem. Soc., 113:40164017, 1991
201. M. A. Hayat. Colloidal Gold: Methods and Applications. Academic, San Diego, CA, 1989
202. C. D. Keating, M. D. Musick, M. H. Keefe, and M. J. Natan. Kinetics and thermodynamics
of Au colloid monolayer self-assembly undergraduate experiments in surface and nanoma-
terials chemistry. J. Chem. Education, 76:949955, 1999
203. P. R. Vanrheenen, M. J. Mckelvy, and W. S. Glaunsinger. Synthesis and characterization of
small platinum particles formed by the chemical-reduction of chloroplatinic acid. J. Solid
State Chem., 67:151169, 1987
204. T. Ould Ely, C. Amiens, B. Chaudret, E. Snoeck, M. Verelst, M. Respaud, and J. M. Broto.
Synthesis of nickel nanoparticles. Influence of aggregation induced by modification of
poly(vinylpyrrolidone) chain length on their magnetic properties. Chem. Mater., 11:526+,
1999
205. F. Dassenoy, K. Philippot, T. Ould-Ely, C. Amiens, P. Lecante, E. Snoeck, A. Mosset,
M. J. Casanove, and B. Chaudret. Platinum nanoparticles stabilized by CO and octanethiol
ligands or polymers: FT-IR, NMR, HREM and WAXS studies. New J. Chem., 22:703711,
1998
206. J. Osuna, D. deCaro, C. Amiens, B. Chaudret, E. Snoeck, M. Respaud, J. M. Broto, and A.
Fert. Synthesis, characterization, and magnetic properties of cobalt nanoparticles from an
organometallic precursor. J. Phys. Chem., 100:1457114574, 1996
207. A. Rodriguez, C. Amiens, B. Chaudret, M. J. Casanove, P. Lecante, and J. S. Bradley. Synthe-
sis and isolation of cuboctahedral and icosahedral platinum nanoparticles. ligand-dependent
structures. Chem. Mater., 8:19781986, 1996
208. A. Duteil, R. Queau, B. Chaudret, R. Mazel, C. Roucau, and J. S. Bradley. Preparation of
organic solutions or solid films of small particles of ruthenium, palladium, and platinum
from organometallic precursors in the presence of cellulose derivatives. Chem. Mater., 5:
341347, 1993
76 C. Ribeiro and E.R. Leite

209. T. Ould Ely, C. Pan, C. Amiens, B. Chaudret, F. Dassenoy, P. Lecante, M. J. Casanove, A.


Mosset, M. Respaud, and J. M. Broto. Nanoscale bimetallic CoxPt1-x particles dispersed
in poly(vinylpyrrolidone): Synthesis from organometallic precursors and characterization.
J. Phys. Chem. B, 104:695702, 2000
210. M. Figlarz. Chimie-douce a new route for the preparation of new materials some exam-
ples. Chemica Scripta, 28:37, 1988
211. G. Viau, F. Ravel, O. Acher, F. Fievetvincent, and F. Fievet. Preparation and microwave
characterization of spherical and monodisperse co-ni particles. J. Magn. Magn. Mater., 140:
377378, 1995
212. C. B. Murray, S. H. Sun, H. Doyle, and T. Betley. Monodisperse 3d transition-metal (Co, Ni,
Fe) nanoparticles and their assembly into nanoparticle superlattices. MRS Bull., 26:985991,
2001
213. J. I. Park and J. Cheon. Synthesis of solid solution and core-shell type cobalt-platinum
magnetic nanoparticles via transmetalation reactions. J. Am. Chem. Soc., 123:57435746,
2001
214. J. R. Thomas. Preparation and magnetic properties of colloidal cobalt particles. J. Appl.
Phys., 37:2914, 1966
215. S. J. Park, S. Kim, S. Lee, Z. G. Khim, K. Char, and T. Hyeon. Synthesis and magnetic
studies of uniform iron nanorods and nanospheres. J. Am. Chem. Soc., 122:85818582, 2000
216. V. F. Puntes, K. M. Krishnan, and A. P. Alivisatos. Colloidal nanocrystal shape and size
control: The case of cobalt. Science, 291:21152117, 2001
217. M. Chen and D. E. Nikles. Synthesis, self-assembly, and magnetic properties of
FexCoyPt100-x-y nanoparticles. Nano Lett., 2:211214, 2002
218. S. R. Whaley, D. S. English, E. L. Hu, P. F. Barbara, and A. M. Belcher. Selection of pep-
tides with semiconductor binding specificity for directed nanocrystal assembly. Nature, 405:
665668, 2000
219. L. O. Brown and J. E. Hutchison. Formation and electron diffraction studies of ordered
2-D and 3-D superlattices of amine-stabilized gold nanocrystals. J. Phys. Chem. B, 105:
89118916, 2001
220. S. H. Sun, S. Anders, H. F. Hamann, J. U. Thiele, J. E. E. Baglin, T. Thomson, E. E. Fullerton,
C. B. Murray, and B. D. Terris. Polymer mediated self-assembly of magnetic nanoparticles.
J. Am. Chem. Soc., 124:28842885, 2002
221. A. L. Roest, J. J. Kelly, D. Vanmaekelbergh, and E. A. Meulenkamp. Staircase in the electron
mobility of a ZnO quantum dot assembly due to shell filling. Phys. Rev. Lett., 89:3680114,
2002
222. H. Weller. Quantized semiconductor particles a novel state of matter for materials science.
Adv. Mater., 5:8895, 1993
223. J. Y. Ying. Preface to the special issue: Sol-gel derived materials. Chem. Mater., 9:2247
2248, 1997
224. C. Sanchez, J. Livage, M. Henry, and F. Babonneau. Chemical modification of alkoxide pre-
cursors. J. Non-Cryst. Solids, 100:6576, 1988
225. V. K. Lamer and R. H. Dinegar. Theory, production and mechanism of formation of monodis-
persed hydrosols. J. Am. Chem. Soc., 72:48474854, 1950
226. W. Stober, A. Fink, and E. Bohn. Controlled growth of monodisperse silica spheres in micron
size range. J. Colloid Interface Sci., 26:62, 1968
227. H. Boukari, J. S. Lin, and M. T. Harris. Probing the dynamics of the silica nanostructure
formation and growth by saxs. Chem. Mater., 9:23762384, 1997
228. H. Boukari, J. S. Lin, and M. T. Harris. Small-angle x-ray scattering study of the formation
of colloidal silica particles from alkoxides: Primary particles or not? J. Colloid Interface Sci.,
194:311318, 1997
229. H. Boukari, G. G. Long, and M. T. Harris. Polydispersity during the formation and growth
of the stober silica particles from small-angle x-ray scattering measurements. J. Colloid In-
terface Sci., 229:129139, 2000
230. M. T. Harris and C. H. Byers. Effect of solvent on the homogeneous precipitation of titania
by titanium ethoxide hydrolysis. J. Non-Cryst. Solids, 103:4964, 1988
Assembly and Properties of Nanoparticles 77

231. T. Sato and R. Ruch. Stabilization of Colloidal Dispersion by Polymer Adsorption. Marcel
Dekker , New York, 1980
232. D. H. Napper. Polymer Stabilization of Colloidal Dispersion. Academic, New York, 1983
233. J. L. Look and C. F. Zukoski. Alkoxide-derived titania particles use of electrolytes to con-
trol size and agglomeration levels. J. Am. Ceramic Soc., 75:15871595, 1992
234. J. L. Deiss, P. Anizan, S. ElHadigui, and C. Wecker. Steric stability of TiO2 nanoparticles in
aqueous dispersions. Colloids Surf. A-physicochem. Eng. Aspects, 106:5962, 1996
235. A. M. Peiro, J. Peral, C. Domingo, X. Domenech, and J. A. Ayllon. Low-temperature deposi-
tion of TiO2 thin films with photocatalytic activity from colloidal anatase aqueous solutions.
Chem. Mater., 13:25672573, 2001
236. R. W. Schwartz. Chemical solution deposition of perovskite thin films. Chem. Mater.,
9:23252340, 1997
237. C. Liu, B. S. Zou, A. J. Rondinone, and Z. J. Zhang. Solgel synthesis of free-standing
ferroelectric lead zirconate titanate nanoparticles. J. Am. Chem. Soc., 123:43444345, 2001
238. S. OBrien, L. Brus, and C. B. Murray. Synthesis of monodisperse nanoparticles of barium
titanate: Toward a generalized strategy of oxide nanoparticle synthesis. J. Am. Chem. Soc.,
123:1208512086, 2001
239. J. J. Urban, W. S. Yun, Q. Gu, and H. Park. Synthesis of single-crystalline perovskite
nanorods composed of barium titanate and strontium titanate. J. Am. Chem. Soc., 124:
11861187, 2002
240. R. M. Barrer. Hydrothermal Chemistry of Zeolites. Academic, London, 1982
241. S. Mintova, N. H. Olson, V. Valtchev, and T. Bein. Mechanism of zeolite a nanocrystal growth
from colloids at room temperature. Science, 283:958960, 1999
242. C. C. Wang and J. Y. Ying. Sol-gel synthesis and hydrothermal processing of anatase and
rutile titania nanocrystals. Chem. Mater., 11:31133120, 1999
243. L. X. Cao, H. B. Wan, L. H. Huo, and S. Q. Xi. A novel method for preparing ordered
SnO2 /TiO2 alternate nanoparticulate films. J. Colloid Interface Sci., 244:97101, 2001
244. T. Nutz and M. Haase. Wet-chemical synthesis of doped nanoparticles: Optical properties
of oxygen-deficient and antimony-doped colloidal SnO2 . J. Phys. Chem. B, 104:84308437,
2000
245. G. S. Pang, S. G. Chen, Y. Koltypin, A. Zaban, S. H. Feng, and A. Gedanken. Controlling
the particle size of calcined SnO2 nanocrystals. Nano Lett., 1:723726, 2001
246. R. S. Hiratsuka, S. H. Pulcinelli, and C. V. Santilli. Formation of SnO2 gels from dispersed
sols in aqueous colloidal solutions. J. Non-crystalline Solids, 121:7683, 1990
247. C. Goebbert, R. Nonninger, M. A. Aegerter, and H. Schmidt. Wet chemical deposition of
ATO and ITO coatings using crystalline nanoparticles redispersable in solutions. Thin Solid
Films, 351:7984, 1999
248. A. P. Rizzato, L. Broussous, C. V. Santilli, S. H. Pulcinelli, and A. F. Craievich. Structure of
SnO2 alcosols and films prepared by sol-gel dip coating. J. Non-Crystalline Solids, 284:61
67, 2001
249. E. A. Meulenkamp. Synthesis and growth of ZnO nanoparticles. J. Phys. Chem. B, 102:
55665572, 1998
250. M. Z. C. Hu, R. D. Hunt, E. A. Payzant, and C. R. Hubbard. Nanocrystallization and phase
transformation in monodispersed ultrafine zirconia particles from various homogeneous pre-
cipitation methods. J. Am. Ceramic Soc., 82:23132320, 1999
251. M. D. Fokema, E. Chiu, and J. Y. Ying. Synthesis and characterization of nanocrystalline yt-
trium oxide prepared with tetraalkylammonium hydroxides. Langmuir, 16:31543159, 2000
252. E. R. Leite, T. R. Giraldi, F. M. Pontes, E. Longo, A. Beltran, and J. Andres. Crystal growth
in colloidal tin oxide nanocrystals induced by coalescence at room temperature. Appl. Phys.
Lett., 83:15661568, 2003
253. E. R. Leite, E. J. H. Lee, T. R. Giraldi, F. M. Pontes, and E. Longo. A simple and novel
method to synthesize doped and undoped SnO2 nanocrystals at room temperature. J. Nanosci.
Nanotechnol., 4:774778, 2004
78 C. Ribeiro and E.R. Leite

254. L. Spanhel and M. A. Anderson. Semiconductor clusters in the solgel process quantized
aggregation, gelation, and crystal-growth in concentrated zno colloids. J. Am. Chem. Soc.,
113:28262833, 1991
255. N. L. Wu, S. Y. Wang, and I. A. Rusakova. Inhibition of crystallite growth in the solgel
synthesis of nanocrystalline metal oxides. Science, 285:13751377, 1999
256. A. Vioux. Nonhydrolytic sol-gel routes to oxides. Chem. Mater., 9:22922299, 1997
257. J. Rockenberger, E. C. Scher, and A. P. Alivisatos. A new nonhydrolytic single-precursor
approach to surfactant-capped nanocrystals of transition metal oxides. J. Am. Chem. Soc.,
121:1159511596, 1999
258. E. R. Camargo, M. Popa, J. Frantti, and M. Kakihana. Wet-chemical route for the prepara-
tion of lead zirconate: An amorphous carbon- and halide-free precursor synthesized by the
hydrogen peroxide based route. Chem. Mater., 13:39433948, 2001
259. E. R. Camargo, J. Frantti, and M. Kakihana. Low-temperature chemical synthesis of lead
zirconate titanate (PZT) powders free from halides and organics. J. Mater. Chem., 11:1875
1879, 2001
260. E. R. Camargo and M. Kakihana. Peroxide-based route free from halides for the synthesis of
lead titanate powder. Chem. Mater., 13:11811184, 2001
261. E. R. Camargo and M. Kakihana. Lead hafnate (PbHfO3) perovskite powders synthesized
by the oxidant peroxo method. J. Am. Ceramic Soc., 85:21072109, 2002
262. M. Kakihana. solgel preparation of high temperature superconducting oxides. J. Sol-gel
Sci. Technol., 6:755, 1996
263. P. A. Lessing. Mixed-cation oxide powders via polymeric precursors. Am. Ceramic Soc.
Bull., 68:10021007, 1989
264. S. G. Cho, P. F. Johnson, and R. A. Condrate. Thermal-decomposition of (sr, ti) organic
precursors during the pechini process. J. Mater. Sci., 25:47384744, 1990
265. E. R. Leite, C. M. G. Sousa, E. Longo, and J. A. Varela. Influence of polymerization
on the synthesis of srtio3.1. characteristics of the polymeric precursors and their thermal-
decomposition. Ceramics Int., 21:143152, 1995
266. E. R. Leite, J. A. Varela, E. Longo, and C. A. Paskocimas. Influence of polymerization on
the synthesis of srtio3.2. particle and agglomerate morphologies. Ceramics Int., 21:153158,
1995
267. M. Kakihana, T. Okubo, M. Arima, O. Uchiyama, M. Yashima, M. Yoshimura, and
Y. Nakamura. Polymerized complex synthesis of perovskite lead titanate at reduced tem-
peratures: Possible formation of heterometallic (Pb,Ti)-citric acid complex. Chem. Mater,
9:451456, 1997
268. M. Cerqueira, R. S. Nasar, E. R. Leite, E. Longo, and J. A. Varela. Synthesis and characteriza-
tion of PLZT (9/65/35) by the Pechini method and partial oxalate. Mater. Lett., 35:166171,
1998
269. H. Takahashi, M. Kakihana, Y. Yamashita, K. Yoshida, S. Ikeda, M. Hara, and K. Domen.
Synthesis of NiO-loaded KTiNbO5 photocatalysts by a novel polymerizable complex
method. J. Alloys Compounds, 285:7781, 1999
270. M. A. L. Nobre, E. Longo, E. R. Leite, and J. A. Varela. Synthesis and sintering of ultra fine
NaNbO3 powder by use of polymeric precursors. Mater. Lett., 28:215220, 1996
271. E. R. Camargo, E. Longo, and E. R. Leite. Synthesis of ultra-fine columbite powder
MgNb2O6 by the polymerized complex method. J. Solgel Sci. Technol., 17:111121, 2000
272. A. L. Quinelato, E. Longo, E. R. Leite, and J. A. Varela. Synthesis of nanocrystalline tetrag-
onal zirconia by a polymeric organometallic method. Appl. Organomet. Chem., 13:501507,
1999
273. S. M. Zanetti, E. R. Leite, E. Longo, and J. A. Varela. Cracks developed during SrTiO3 thin-
film preparation from polymeric precursors. Appl. Organomet. Chem., 13:373382, 1999
274. V. Bouquet, M. I. B. Bernardi, S. M. Zanetti, E. R. Leite, E. Longo, J. A. Varela, M. G. Viry,
and A. Perrin. Epitaxially grown LiNbO3 thin films by polymeric precursor method. J. Mater.
Res., 15:24462453, 2000
Assembly and Properties of Nanoparticles 79

275. F. M. Pontes, J. H. G. Rangel, E. R. Leite, E. Longo, J. A. Varela, E. B. Araujo, and J. A. Eiras.


Low temperature synthesis and electrical properties of PbTiO3 thin films prepared by the
polymeric precursor method. Thin Solid Films, 366:232236, 2000
276. F. M. Pontes, E. R. Leite, E. Longo, J. A. Varela, E. B. Araujo, and J. A. Eiras. Effects of the
postannealing atmosphere on the dielectric properties of (Ba, Sr)TiO3 capacitors: Evidence
of an interfacial space charge layer. Appl. Phys. Lett., 76:24332435, 2000
277. F. M. Pontes, E. R. Leite, E. J. H. Lee, E. Longo, and J. A. Varela. Dielectric properties and
microstructure of SrTiO3 /BaTiO3 multilayer thin films prepared by a chemical route. Thin
Solid Films, 385:260265, 2001
278. E. R. Leite, E. C. Paris, E. Longo, and J. A. Varela. Direct amorphous-to-cubic perovskite
phase transformation for lead titanate. J. Am. Ceramic Soc., 83:15391541, 2000
279. S. M. Zanetti, E. R. Leite, E. Longo, and J. A. Varela. Preparation and characterization of
SrBi2Nb2O9 thin films made by polymeric precursors. J. Mater. Res., 13:29322935, 1998
280. E. R. Leite, E. C. Paris, E. Longo, F. Lanciotti, C. E. M. Campos, P. S. Pizani, V. Mastelaro,
C. A. Paskocimas, and J. A. Varela. Topotatic-like phase transformation of amorphous lead
titanate to cubic lead titanate. J. Am. Ceramic Soc., 85:21662170, 2002
281. E. R. Leite, M. A. L. Nobre, M. Cerqueira, E. Longo, and J. A. Varela. Particle growth
during calcination of polycation oxides synthesized by the polymeric precursors method.
J. Am. Ceramic Soc., 80:26492657, 1997
282. C. Greskovich and K. W. Lay. Grain-growth in very porous a12o3 compacts. J. Am. Ceramic
Soc., 55:142, 1972
283. E. R. Leite, I. T. Weber, E. Longo, and J. A. Varela. A new method to control particle size
and particle size distribution of SnO2 nanoparticles for gas sensor applications. Adv. Mater.,
12:96, 2000
284. E. R. Leite, A. P. Maciel, I. T. Weber, P. N. Lisboa, E. Longo, C. O. Paiva-Santos, A. V. C. An-
drade, C. A. Pakoscimas, Y. Maniette, and W. H. Schreiner. Development of metal oxide
nanoparticles with high stability against particle growth using a metastable solid solution.
Adv. Mater., 14:905908, 2002
285. J. Weissmuller. Alloy thermodynamics in nanostructures. J. Mater. Res., 9:47, 1994
Electrochemistry, Nanomaterials,
and Nanostructures

Paulo Roberto Bueno and Claude Gabrielli

Abstract This chapter deals with the development of new methods for the design of
more efficient electrochemical cells destined specifically for energy conversion and
storage based on synthesis and design of functional electrodes and electrolytes. The
main focus of this chapter is on novel strategies that exploit nanoscale architectures
to enhance the efficiency of alternative energy conversion and storage devices as
well as on the basic principles of electrochemistry governing the effects of nanoscale
design on electrodes and electrolytes. In addition, the chapter provides a review of
fundamental electron transfer concepts of relevance to electrochemistry in general
and alternative energy devices in particular.

1 Introduction

An in-depth understanding of the processes involved in the operation of


electrochemical cells is crucial in the development of new methods for the design
of more efficient electrochemical cells destined specifically for energy conversion
and storage. An indispensable aspect of this body of knowledge is a thorough grasp
of the synthesis and design of functional electrodes and electrolytes with unusual
and valuable properties which, today, are provided to a large extent by the nanosize
effect of these cells functional components. Indeed, future developments in this
field will necessarily depend on nanoscience and nanotechnology, as this chapter
intends to demonstrate.
The main focus of this chapter, therefore, is on novel strategies that exploit
nanoscale architectures to enhance the efficiency of alternative energy conversion
and storage devices and on the basic principles of electrochemistry governing the

P.R. Bueno ()


Instituto de Qumica, Departamento de Fsico-Qumica, Universidade Estadual Paulista, C. Postal
355, 14801-907, Araraquara, Sao Paulo, Brazil
e-mail: prbueno@mail.iq.unesp.br

E.R. Leite (ed.), Nanostructured Materials for Electrochemical Energy Production 81


and Storage, Nanostructure Science and Technology, DOI 10.1007/978-0-387-49323-7 3,
c Springer Science+Business Media LLC 2009
82 P.R. Bueno and C. Gabrielli

effects of nanoscale on electrodes and electrolytes. In addition, this chapter intro-


duces basic principles of nanotechnology to guide the reader through this incipient
interdisciplinary boundary. The scope of ensembles of nanosized objects and prod-
ucts that have been and can be created is truly remarkable. Moreover, the potential
impact of these products on basic science and applications is so great that the subject
is inexhaustible, particularly in view of the enormous diversity of nanosized objects.
Therefore, this chapter must perforce be limited to objects and systems that already
show real promise for a variety of applications.
The electrochemical manufacture of small objects ranks high in the design of new
materials and systems with special properties. However, although the mechanism of
formation or removal of small individual portions of solids plays a decisive role in
the field of electrochemical nanotechnology, the use of electrochemistry as a tool
for the preparation of nanostructures for electrode and electronic applications lies
outside the scope of this chapter. In other words, the absence of this and other related
topics clearly indicates, from the start, that no attempt will be made to cover the field
comprehensively here.

2 Electrochemistry and Nanoscale Materials

2.1 Electrochemistry and Size Effects

Electrochemistry and nanoscience (and/or nanotechnology) are interdisciplinary


fields, both of which are gaining increasing importance in the development of high
performance and reliable alternative energy devices (conversion or storage) [13].
To begin to understand how these areas are interrelated to improve the performance
of such devices, a brief explanation about both areas must be given. Electrochem-
istry, in turn, can be understood as the interaction between chemistry and electricity,
although it encompasses far more than mere knowledge of chemical systems and the
physics of electric fields or potential. As the name itself suggests, electrochemistry
is a field of science that deals with the relationship between electrical current or
potential and chemical systems. From a more specific standpoint, electrochemistry
addresses the chemical and physical transformations underlying chemical energy
storage and conversion and their relationship to limitations in the performance of
electrochemical systems [26]. Modern electrochemistry covers all phenomena in
which a chemical change is the result of electric forces and vice versa, where an
electric force is generated by chemical processes. It includes the properties and be-
havior of electrolytic conductors in liquid or solid form. Moreover, modern electro-
chemistry can be divided into interfacial electrochemistry (electrochemistry of het-
erogeneous systems) and bulk electrochemistry (electrochemistry of homogeneous
systems). On the one hand, interfacial electrochemistry differs from bulk electro-
chemistry by dealing with topics such as the nature of an electrodeelectrolyte in-
terphase, the thermodynamics and kinetics of reactions occurring in the interphase
and mass-transport effects throughout it [26]. On the other hand, bulk electrochem-
Electrochemistry, Nanomaterials, and Nanostructures 83

istry deals with ionsolvent and ionion interactions, activity coefficient, ionic mo-
bility, ionic conductivity, and so on [26]. Bulk electrochemistry describes the per-
formance of electrolytes and is the key to future improvements in electrochemical
components used as ionic sources and membranes in solar energy devices for elec-
tric power production, energy storage, e.g., batteries and supercapacitors, as well as
advanced electroanalytical sensor devices [26]. Bulk electrochemistry is discussed
briefly here, in the context of ionic conductivity of solid-state electrolytes. However,
most of this chapter focuses on the fundaments of interfacial electrochemistry and
its relationship to nanostructures, which is of vital importance in electrochemical
alternative energy devices.
The term nanotechnology has been defined in a variety of ways; however, we will
define it here as the technology of design, manufacture and application of nanostruc-
tures and nanomaterials [7, 8]. The study of the fundamental relationships between
physical and chemical phenomena and materials dimensions on a nanometric scale
are also referred to as nanoscience. Nanotechnology involves the fabrication, char-
acterization, and manipulation of materials with at least one dimension in the range
of 1100 nm. Thanks to their diminutive size, nanoscale objects will lead to ground-
breaking discoveries and are expected to be at the forefront of technological inno-
vation for the next decade. The interrelation between interfacial electrochemistry
and nanoscience gives rise to new possibilities for designing chemical surfaces by
controlling surface structures at the molecular level, leading to innovative metal or
semiconductor surfaces and charge transfer lengths.
Improved nanostructured devices based on nanostructured electroactive material
can be designed in different ways [915]. For instance, the interfacial electrochem-
ical properties of specific materials can govern the performance of charge accumu-
lating at the interface, leading to highly efficient double layer capacitors [1624].
In bulk electrochemistry, a proper understanding of the high performance of ion-
conduction properties on a nanolength scale is crucial [2529]. Other examples will
be discussed later herein.
Micrometric-scale materials generally display the same physical properties as
those in bulk form; however, nanometric-scale materials may exhibit physical prop-
erties that are distinctively unlike those of bulk. Materials in this size range possess
remarkable specific properties deriving from the transition from atoms or molecules
to bulk form that takes place in this size range. On the one hand, the interfaces in
polycrystalline microstructured materials are considered as defects that influence the
macroscopic properties. On the other hand, in polycrystalline nanostructured mate-
rials, the interface dominates and the bulk plays a totally different role. Numerous
experimental studies have shown that, if a materials particle size is less than the crit-
ical size of about 10 nm, its bulk properties change noticeably. A particle of about
10 nm contains 104 105 atoms, 15% of which are on the surface of the particle and
contribute substantially to the materials physicochemical properties [30]. Ensem-
bles of such nanostructures will be shown here to be important in electrochemical
applications such as electrodes [8, 9, 12, 13, 3138].
Another example of the influence of nanometric effects is on the Debye tem-
perature, which decreases in nanocrystalline systems, and is lower when the same
84 P.R. Bueno and C. Gabrielli

system is on a micrometric scale. Thus, the nanometric scale gives an additional


contribution to the low-temperature heat capacity, and this contribution increases as
the scale decreases [39].
This effect can be explained by changes in the nanocrystals vibrational status
(an increase in the number of low frequency modes in response to a decrease in
the number of high frequency modes) as the size decreases. Therefore, nanometric-
scale crystals have a low melting point (sometimes below 900 C) and reduced lattice
constants, since the number of surface atoms or ions represents a significant fraction
of the total number of atoms or ions. In this case, surface energy plays a significant
role in the thermal stability, while bulk represents the defect in the properties. From
this standpoint, nanotechnology can be considered new, but nanometer scale re-
search is not new at all. Nanometric-scale engineering of many materials such as
colloidal dispersion, metallic quantum dots and catalysts [37, 4049] has existed for
decades. Nanotechnology is therefore not a novelty; indeed, you can see it all around
you if you just know where to look.
Actually, nanotechnology is a combination of existing technologies, and our new-
found ability to observe and manipulate on the atomic scale makes nanotechnology
highly compelling from the scientific, business and political points of view. This
aspect of nanotechnology was already foreseen in a conference presented by Nobel
Prize winner Richard Feynman in 1959.
One can visualize another example of the nanoscale effect by picturing a bulk
material, e.g., a metal or n-type semiconductor metal oxide, simplistically, as con-
sisting of charged particles, positive ions, and electrons in the first situation, and
positive and negative ions in the second. In both cases, the total positive charge bal-
ances the total negative charge, so that there is no net electric charge in the materials
bulk. Now imagine reducing a piece of this material to the nanoscale by applying a
top-down approach such as any lithographic technique in a vacuum medium. What
electronic properties will such a small, strongly charged material have? Consider-
ing the vacuum as the environment in which these properties are embedded, there
is a small degree of spillover of the electrons from the material into the vacuum,
resulting in a disturbance of the former balance of electrical charge in the surface
region of the materials, which in this case represents a substantial percentage of
the total volume. Note that interfaces in micrometric or millimetric materials can
also be electrified, but there are some differences. A totally different functionality
arises on the nanoscale, because of the size range involved [30]. Indeed, changes
in the electronic properties of nanoscale materials due to the size effect must be
interpreted in the domain of quantum chemistry [50]. If the size of the nanocrystal
is compared with the de Broglie wavelength of elementary excitations, the quanti-
zation conditions of electron energy are changed and the energy bands split into an
energy system of energy level and, in this case, the pattern of the absorption spec-
trum becomes similar to that of the spectra of individual clusters. Semiconductor
nanocrystals possessing these properties are called quantum dots [50].
The physicochemical properties of nanometric-scale materials are part of the do-
main of nanoscience but, in specific situations, nanomaterials can converge with
electrochemistry. For instance, if our previously pictured nanomaterials (nanocrys-
Electrochemistry, Nanomaterials, and Nanostructures 85

talline particles) are immersed in an appropriate solution, a colloidal system is likely


to arise [51, 52]. The use of nanoparticles to form colloidal systems has a long his-
tory, as exemplified by a comprehensive study about the preparation and properties
of colloidal gold, which was first published in the mid-nineteenth century [53]. The
colloidal dispersion of gold prepared by Michael Faraday in 1857 [54] remained
stable for almost a century, showing no sign of spontaneous chemical degrada-
tion before it was destroyed in a Word War II air raid. Colloidal properties result
from the combination of the properties of an electrified surface and nanomaterials/
solution, the structure of the double layer thus formed being the most important
factor.
In the emerging field of nanotechnology applied to electrode design, the goal
is to create nanostructures or nanoarrays with special properties unlike those of
bulk or single particle nanosize elements. Thus, if nanoparticles are combined in
any way under a conductor substrate and inserted into an electrolyte, the resulting
properties will differ from those produced under the same conditions but without
nanoparticles or nanoscale materials [10, 12, 30, 37, 47]. Nanoparticles themselves,
or an electrode composed of nanoparticles, can exhibit unique chemical proper-
ties because of the limited size and high density of corner or edge surface sites
[7, 30, 55, 56], but as electrode-forming components, the most important properties
are those arising from the charge transfer rate, electronic and ionic transport, etc.
[8, 15, 3032, 40, 42, 47, 57, 58].
It is a known fact that, from the standpoint of materials and surface science, many
of the properties of systems are controlled by interfaces and contact between differ-
ent materials [7]. However, knowledge of how to control and design new contact
(electrodic or electronic) properties on a nanometer scale is new and is very rapidly
gaining ground. Such contacts constitute the electrodes of galvanic cells which can
be used for the generation of chemical products by electric power (electrolysis), as
will be discussed later herein.
It is of primary importance to know how to differentiate between electrochem-
ical responses and events that result from changes in charge transfer length relat-
ing to the functionalized surface properties of nanoscale materials and those that
arise from transport properties through nanostructures. It is easy to demonstrate that
chemical reactions occurring in nanostructured electrodes (electrode reactions) and
mass and electrical transportation in such electrodes, and even in electrolytes based
on nanocomposites or other kinds of nanomaterials, will cause a tremendous rev-
olution in every aspect of applied electrochemistry [8, 12, 13, 15, 30, 3234, 36, 38,
42, 43, 57, 5970]. The main focus here is to show how these nanoscale effects re-
sulting from the use of designed nanostructured electrodes will inevitably contribute
to the development of alternative energy devices based principally on concepts of
electrochemistry.
We, therefore, need to know more about how to deal with and take advantage of
the continuous flow of electrons across electronic nanostructures and ionic species
in a nanoscale ionic conductor (electrolyte). Furthermore, it is also important to un-
derstand how charge transfer lengths are influenced on the nanoscale and how size
affects electrochemical properties [8, 30]. Accordingly, we will consistently address
86 P.R. Bueno and C. Gabrielli

the processes and factors affecting the transportation of charges across interfaces
between nanostructured chemical phases. In the latter case, electrochemical reac-
tions are involved and the transfer of electrons from the electronic phase to the ionic
conducting phase is classified as an oxidationreduction (redox) reaction. In the
case of electrochemical conversion systems, it is also important to deal with charge
transfers occurring from molecules to semiconducting nanoparticles [71].

2.2 Challenges of Charge Transfer

It must also be emphasized that oxidationreduction reactions always involve charge


transfers, without which we can no longer exist [8]. Electron transfers are vital to
our daily life. Indeed, we owe our very existence to photosynthetic and biochemical
electron transfer events, and we would be unable to function properly without the
myriad transistor-based electronic devices that control and amplify current flows.
Electron transfer and charge storage are correlated in a series of events in our body.
Another example of the importance of electron transfer reactions is photosynthe-
sis, which photoelectrochemical systems designed for energy conversion attempt to
imitate [3, 72]. In photosynthesis, a photon triggers an electrical current in the leaf,
activating the leafs electrons to produce H2 while the holes produce O2 .
Therefore, electron transfer events are basic to the understanding and develop-
ment of novel generations of electrochemical energy conversion and storage systems
based on the imitiation of the most efficient biological systems. Electrochemical en-
ergy conversion and storage devices form an important branch of alternative energy
emerging in the twenty-first century [1] that, like nanotechnology, has crucial scien-
tific, technological, and political implications.
At this point, however, it is important to stress that despite the importance
of properly grasping the charge transfer phenomenon, molecular and bulk level
charge transfer processes are still not fully understood and the characterization of
nanoscale (1100 nm) processes is still incipient [8, 73]. Nanoscale charge trans-
fer is important both to the frontiers of fundamental science and to applications
in molecular electronics, including problems such as electrocatalysis [8] and solar
photoconversion.
Progress in the area of nanoscale charge transfer requires interdisciplinary col-
laboration, combining a wide range of materials synthesis and electrochemical char-
acterization, a challenging range of experimental techniques to probe charge trans-
fer processes, and theory for their interpretation [8]. Current interest ranges from
the use of single or small groups of molecules (usually organic) as components in
electronic devices to the exploitation of semiconductor and metal nanoparticles be-
cause of their high surface areas and other size-dependent properties. The nature
of the attachment of such components to bulk metal and semiconductor surfaces
and the control of their properties are overarching concerns [8]. The experimen-
tal measurements used to characterize nanoscale charge-transfer properties include
Electrochemistry, Nanomaterials, and Nanostructures 87

rate constants, spectroscopy, and conductance/resistance measurements, depending


on the nature of the system studied [30].
Predictions of the kinetics of electrons taking into account all size-dependent
factors are possible only when adequate ion-molecular models of reaction layers
are built. For a number of systems, this problem can be solved successfully by
employing quantum-chemical methods based on quantum mechanical theory of
the charge-transfer elementary act [74, 75] along with the classical effects of the
cation size, which are manifested in the reduction of anions on a negatively charged
surface [74, 75].
In general, the characteristic dimensions of conventional components of
electrochemical systems (electrodes, electrolytes and membranes) cover 510
orders of magnitude and the lower boundaries of the corresponding intervals
approach typical sizes of specific space regions (layers) that arise only at the
electrodeelectrolyte junctions, generally called Electrical Double Layer (EDL)
[36]. According to modern notions of the interface structure, the adsorbate layers
and the edges of surface atoms on electrodes are spatially separated by gaps. In elec-
trochemical systems, the properties of this gap depend on the electrode potential
(charge).
The prospect of using size effects to intensify electrode processes cannot be con-
templated without including nanotechnology and electrode design. The latter will
be introduced in Sect. 2.3.

2.3 Nanomaterials and Nanostructured Films as Electroactive


Electrodes

Earlier herein we mentioned several challenges posed by nanoscience and electro-


chemistry. Some aspects relating to electrified surfaces in contact with electrolyte,
the so-called electrified interfaces in electrodics, metal or semiconductor
electrolyte junctions whose properties can control charge transfer length, charge
transfer kinetics, and the EDL structure, have been discussed briefly. These topics
will be discussed in greater detail in other sections of this chapter.
Also briefly mentioned earlier is the fact that the physical properties of the
interface of nanoparticles in solution/solvent or electrolytes may lead not only to
colloidal behavior but also to particleparticle interaction or particlesolvent inter-
action. Self-supporting colloid network structures allow for the coexistence of high
conductivity with mechanical stability, enabling colloidal gels to be used as elec-
trolytes [7678].
Despite all these aspects of electrochemistry and nanoscale discussed so far, un-
doubtedly the most important question to be answered is Why is nanoscience or
nanotechnology so beneficial to electrochemical energy devices?
The answer lies in understanding the fundamental principles underpinning the
electrochemical generation of electricity. The conversion of energy necessarily in-
volves some kind of energy transfer step, whereby the energy from the source is
88 P.R. Bueno and C. Gabrielli

passed along to electrons, which constitute the electric current. This transfer occurs
at a finite rate and must take place at an interface or reaction surface. Thus, the
amount of electricity produced scales with the amount of reaction surface area or
interfacial area available for the energy transfer. In other words, the efficiency of the
reaction and the amount of reaction depend, respectively, on the features of the sur-
face and on the extent of the surface. In this perspective, how can nanotechnology
and electrochemical devices (e.g., energy conversion and storage devices) converge
from a nanoscopic standpoint of electronic transfer, transport and kinetic reactions
occurring in these devices?
To begin answering this question, we must first offer a basic introduction to the
term electroactive materials so popular in the electrochemical literature. Elec-
troactive materials are materials that possess redox properties, and they play an
increasingly central role in nanotechnology and electrochemistry. At present, the
wide range of applications for these materials include electrodes and membranes
for electrochemical energy conversion and storage, electroceramic electrode devices
and sensors, organic diodes, and magnetic and optical devices [1, 7, 8, 43, 7983].
Further along in this chapter we will discuss the concept and definition of elec-
trodes and electrochemical cells. For now, an electrode can be considered a source or
a sink of electrons to be interplayed with the ionic conductor in contact with it, while
the electrochemical cell is composed of a source and a sink electrode separated by
an ionic conductor (electrolyte). Therefore, all electroactive materials are potential
electrodes. From the standpoint of nanotechnology, this material can be prepared
on a nanoscale (to enlarge the surface area) mainly in three distinct forms (a) zero-
dimensional, (b) one-dimensional and (c) two-dimensional nanostructures. There-
fore, it is not surprising that the desire for large surface areas has led researchers to
focus on nanomaterials and electrochemistry in the search for nanosized structures
for electrode applications. Nanomaterials include molecular wires, nanoparticles
both semiconducting and metallic, nanotubes, electrodes, and the connectors that
link these objects together to create larger structures (Fig. 1). Other products, such
as filled zeolites, aerogels, dendrites, and layered polymers, also offer enormous
potential for useful electrochemical functions [8].
All these nanosized structures [from (a) to (c)] are synthesized using either one
of two manufacturing concepts, i.e., top-down or bottom-up approaches. Based on
these two approaches, several methods have been developed in recent years for the
preparation of novel nanostructures [7].
In the case of (a), the zero-dimensional (0-D) nanostructure is composed of
nanoparticles whose fabrication requires the control of more than merely their
diminutive size. For any practical applications such as electroactive components
in electrodes, the processing conditions must be controlled so that the resulting
nanoparticles have the following characteristics (a) identical size of all particles
(also referred to as monosized or quasi-monosized), (b) identical shape or mor-
phology, and (c) identical or at least very similar chemical composition and crys-
tal structure. Single crystalline nanoparticles are often referred to in the litera-
ture as nanocrystals. When the characteristic dimensions of nanoparticles are suffi-
ciently small and quantum effects are observed, these nanoparticles are commonly
Electrochemistry, Nanomaterials, and Nanostructures 89

Fig. 1 Nanostructured electrodes can be prepared by several methodologies, many of them based
on bottomup approaches. This figure illustrated the ordered (left) and disordered (right) prepa-
ration of brushlike structures obtained from nanowires and/or nanotubes. We are indebted to
Dr. Ednan Joanni for providing some parts of this schematic representation

described as quantum dots. From the electrochemical standpoint we are interested in


the redox properties of such nanoparticles, which are generally metallic, inorganic
or compounds.
One-dimensional (1-D) nanostructures (b) are composed mainly of nanowires,
nanorods and nanobelts. Spontaneous growth, template-based synthesis and elec-
trospinning are considered bottom-up approaches, while lithography is a top-down
technique [7, 84]. Two-dimensional (2-D) nanostructures (c) involve thin films,
which have been the object of intensive study for almost a century and for which
many methods have been developed and improved.
Insofar as the concept of electrode is concerned, it is important to stress
here that most electrodes for electrochemical application in alternative energy
devices are manufactured mainly by interfacing semiconducting nanostructures
with conducting substrates. By combining different electrodes, one can attain a
three-dimensional (3-D) integrated electrochemical cell [8587]. Electrodes can
90 P.R. Bueno and C. Gabrielli

also be developed based on a 3-D concept, whereby they are assembled using 0-D,
1-D, or 2-D nanostructures. 2-D-based electrodes can be fabricated as homoge-
neous thin films or by an assembly of 0-D and 1-D nanostructures; however, the
films thickness must be limited to the nanoscale. 3-D electrodes can be similarly
manufactured, but their thickness is not limited to nanolengths. Therefore, the
assembly and synthesis of these nanostructures with multiple dimensions is com-
mon in electrochemistry. They are usually referred to as nanostructured electrodes
and most of them are highly porous.
The nanostructuring of electroactive materials introduces major changes in the
electrochemical properties of the devices of which they are a part. For instance,
dye-sensitized solar cells (DSSC) is an example of 2-D and 3-D nanometer-sized
structures assembled with quasi 0-D structures or building blocks (nanoparticles).
The charge transfer kinetics is often influenced and high surface areas provide
greater numbers of electroactive charge transfer sites. A more specific example is the
extremely large internal surface area, which is made possible by the nanocrystalline
particle nature of nanostructured semiconductor electrodes. This huge surface area
is crucial for the proper functioning of DSSCs for a number of reasons. Some of
these reasons are (1) sufficient adsorption of the monolayer of dye molecules to
achieve efficient light absorption, (2) the huge surface causes charge carrier perco-
lation across the nanoparticulate lattice, and (3) the very rapid and highly efficient
interfacial charge transfer between the oxide and each and every one of the dye
molecules anchored to the particle surface.
We have already shown here that large surface areas translate into enhanced elec-
trochemical performance and, from this perspective, nanostructures are also very
important for fuel cell devices [8,70,8890], particularly as regards the catalyst part
of the cell where the oxygen reduction reaction can only occur in spatially confined
regions [8, 70, 8890].
Returning to the subject of nanostructures designed and used as electrodes,
it is easily observed from the electrochemical literature that 2-D or 3-D struc-
tured electrodes composed of architectures possessing high surface-to-volume ratio
nanostructures, which are constituted by the arrangement of multiple dimensional
nanostructures (0-D, 1-D, and 2-D), are extremely useful in energy storage or con-
version device applications because they improve storage capacity and conversion
[1, 8, 15, 37, 70, 86]. We will provide specific examples in this chapter, and several
other examples are given in the handbook.
As mentioned earlier, thin films can be considered 2-D nanostructures. In the
recent past, thin film-based technologies have been responsible for the design of
an enormous variety of thin film-based electrochemical devices. 2-D nanostruc-
tures [91] composed of 0-D or 1-D materials are those that are associated with the
interfacial properties of electrodes. In electrochemistry they are known as porous
electrodes, and they sometimes possess an effective surface more than 1,000 times
greater than the geometric area expected for a compact and homogeneous 2-D struc-
tured electrode, e.g., porous thin film-related electrodes [9296].
As can be seen, therefore, the porous effect is an important and significant as-
pect of nanostructures applied as electrodes. It is so important to note that, in the
Electrochemistry, Nanomaterials, and Nanostructures 91

electrochemical literature, this aspect leads to the use of terms such as porous elec-
trodes when referring to nanostructured electrodes [92101]. Porous, or nanostruc-
tured, electrodes form an electrochemical system when in contact with an elec-
trolyte, leading to the simultaneous transport of electronic and ionic species. Owing
to the minute scale of the constituents of the porous lattice, charge ionic carriers in
transit are always close to the surface, implying that transport and heterogeneous
transfer processes are strongly coupled in these systems [92101]. We will discuss
porous electrode theory further along in this chapter.

2.4 Nanomaterials as Electrolytes

Nanoscience and nanotechnology influence not only the design of nanoscale


electrodes but also that of nanoscale electrolytes [9, 28, 102105]. The continu-
ous research of new materials for electrolyte applications has led to advances in the
design of nanostructured materials to improve ionic conductivity [25, 28, 102]. The
development of new solid materials for electrolyte applications is creating opportu-
nities for new types of electrical power generation and storage systems. Progress to
date in solid-state ionics is largely the result of fast ionic conductors [29,106]. There
are several new classes of materials based on the concept of nanoionics (nanostruc-
tured materials for use primarily as electrolytes). Solid ionic materials of this kind
are based on nanoceramics, polymeric compounds, and hybrid organicinorganic
nanocomposites [102]. All these materials are being investigated to improve their
transport properties for the ionic species of interest, e.g., lithium, proton, or oxygen.
Therefore, enhancing the ionic conductivity of nanoscale materials is of great in-
terest, for it offers a predictable way to design a wide range of new ionic materials
with practical applications. An obvious technological target is the preparation of
nanoscale lithium-ion conductors to make more efficient rechargeable batteries for
portable electronics [107]. Many problems still remain to be solved, such as the
scale-up of laboratory techniques for mass production. Nonetheless, the results
achieved so far suggest that the emerging field of nanoionics has a healthy future
[28, 107].
The concept behind nanoionic materials is based on the use of nanoscale effects
to improve overall ionic conductivity [28]. Overall conductivity, which is the prod-
uct of defect concentration and mobility, is low in simple classic inorganic solid
ionic materials such as NaCl. The electrical conductivity at the melting point is
eight orders of magnitude lower than in metal, and is almost impossible to measure
at room temperature. The conductivities of a few materials even approach those of
aqueous electrolyte solutions. Solid state electrolyte material began to attract seri-
ous attention with the discovery of new compounds and nanocompounds with high
ionic conductivities, and with the demonstration of feasible devices, particularly
solid-state batteries, fuel cells, and sensors [108,109]. Fuel cells based on these ma-
terials known as fast-ion conductors or solid electrolytes offer a highly efficient
and clean method of large-scale energy production [109].
92 P.R. Bueno and C. Gabrielli

An example of classical solid state electrolyte is electrolytes based on zirconia


polycrystalline material that conducts oxygen ions, with hydrogen as the fuel, air as
the oxidant, and water as the waste product [109]. The fossil fuel crisis and the need
for better energy management will lead to increased activity in this area, and interest
in nanosized ionic materials continues to grow [1]. For example, the properties of
electrolyte ceramic materials can be improved by the nanoscale surface properties.
The surface properties of an ionic crystal create a space-charge region, in which
the concentration of defects at the surface is governed by the energies of individual
defects and one type of defect in a pair can dominate. As a result, the surface
has an excess of one type of defect, as in the case of chemical doping, and ionic
conduction in this region is enhanced accordingly [28, 110]. Some experiments and
theories indicate that dramatic increases in ionic conductivity should be expected
when the spacing of interfaces is comparable to, or smaller than, the space-charge
layer [28, 110].
However, it is evident that lowering the length scale may lead to enhanced de-
fect densities and, therefore, to enhanced ionic conductivities in the space charge
regions, irrespective of yet unknown mobility effects. With regard to the safe stor-
age of (sustainable) hydrogen, exciting new avenues are being opened for storage
using nanosized and nanostructured materials [102].
While the quantum confinement regime has been explored for some ionic ma-
terials, it is evident that thus far only optical properties have been related to the
quantum confinement regime. Electrical properties of solid electrolytes or mixed
ionicelectronic conducting materials in the quantum confinement regime have not
yet been described fully and represent a challenge to this field. The achieving of high
ionic conductivity in nanostructure-based electrolytes requires a better understand-
ing of the fundamentals of ion dissociation and transport in such kinds of nanosized
materials [28].

2.5 Nanoscale Electronic and Ionic Transport

Finally, considering what was discussed previously, when dealing with nanosized
materials and nanostructured electrodes for electrochemistry, it is important to sep-
arate the different effects of interface on the electronic and ionic transport: the
kinetics and mechanisms of transport along and across interfaces. The literature
commonly considers transport along interfaces as grain boundary transport, corre-
sponding to diffusion parallel to interfaces, as in grain boundaries of polycrystalline
materials or in nanoscale materials, as across nanostructures limited to a thin layer
of nanometric thickness. In contrast, transport across interfaces involves transport
perpendicular to the interface.
It is possible to show that, if one considers the correct description of ionic dif-
fusion coefficient suction of a crystal, the value obtained for the ionic coefficient
on the surface is abnormally high in the nanoscale compared with the microscale
[111]. It has also been stated [72] that the coefficients of oxygen diffusion in TiO2
Electrochemistry, Nanomaterials, and Nanostructures 93

nanocrystals are 56 orders of magnitude greater than in bulk crystals and that the
coefficients of diffusion of dissolved hydrogen in palladium nanocrystals are greater
than in the bulk specimen. Oxygen conductivity can also be improved in the same
way for ZrO2 -based solid electrolyte [102].
The same concept can be applied to interfaces between the components of
rechargeable solid-state lithium ion batteries, thereby improving the performance
and reducing the resistivities of dc batteries [102].

2.6 Energy Conversion and Storage in Electrochemistry

The conversion and storage of energy by electrochemical devices is a highly


important technological issue of debate, for it represents an entirely new source
of energy generation affecting every aspect of our daily lives, and its efficiency
is tied in with the concept of free energy (the capacity to do work, in the formal
physical definition), which is the most fundamental resource. With an inexhaustible
supply of free energy, almost anything can be done. According to the first law of
thermodynamics, energy cannot be created or destroyed, but the critical constraint
lies in the second law of thermodynamics. In the way it is usually expressed, the
second law states that, in an isolated system, a quantity called entropy, having units
of energy/temperature, must always increase in any spontaneous process. It there-
fore follows that increasing entropy, in turn, implies that free energy cannot be fully
recycled. Although the total energy remains unchanged, less energy is available to
perform useful work. A certain amount is irretrievably lost, in practice generally
as low-grade waste heat, which from Earth is ultimately radiated out into space.
Hence, ongoing sources of free energy (or just energy, loosely) are required for all
biological and technological activities. Sunlight is the source for the biosphere. In
the case of conventional technology, a whole suite of sources is used, but nowadays
the predominant source is fossil fuels. These have a considerably higher energy den-
sity than that of sunlight. However, as fossil fuel supplies are expected to become
increasingly scarce, expensive, and environmentally impacting, increasing depen-
dence on energy conservation and alternative sources of energy are expected to be-
come the most obvious solution, especially energy from the sun, i.e., solar energy.
When the sun is directly overhead and the sky is clear, radiation on a horizontal sur-
face is about 1, 000 W m2 [1, 71]. To give an example, photovoltaics and solar cells
have, in fact, provided reliable electrical power to space missions for many years.
Sunlight can be used for heating, lighting, and electricity generation, and it can be
concentrated to provide steam to run turbines. The principal disadvantages of solar
energy are still that, at present, the conversion efficiency of sunlight to electric power
is not high, and sunlight varies according to the time of day, weather conditions and
seasons.
Closely related to solar cells are fuel cells, which convert fuel or chemical
energy directly to electric energy. The main chemical reactions involve the oxida-
tion of CO and H2 . Although fuel cells have been operating reliably and efficiently
94 P.R. Bueno and C. Gabrielli

in space missions for more than 40 years, they have not yet been widely used
on Earth, largely because of their cost. To make this technology commercially
competitive, better anode, cathode, and electrolyte materials and processes are
needed [1].
However, what is the real importance of electrochemical alternative energy de-
vices? There are several energy sources, as presented earlier. All of them involve
electron transfer reactions, and their development and applications are growing
rapidly [1, 71, 109, 112], especially because conventional energy supply technology
is exhausting its resource base at an accelerating rate, exacerbated by the revolution
of rising expectations in the less-developed world due to the global communications
revolution. This rapid expansion in the study and development of electrochemical
energy generation and conversion-based devices is due mainly to the fact that most
of the worlds energy is supplied by fossil and nuclear sources, which face continual
and increasingly severe issues of resource limitation and environmental pollution.
To meet the growing global demand for energy and to compensate for the depletion
of fossil fuel supplies in the coming years, alternative renewable sources of energy
that do not depend on fossil fuels and that cause only marginal environmental im-
pacts must be developed, offering low costs, safety, and higher efficiency. There is
also a great demand for devices based on the concept of energy storage, i.e., on
the electrochemical concept of storage of chemical energy. These devices are also
required to supply mobile energy for the portable electronic devices upon which
our modern lifestyle is so strongly dependent, for they are at the heart of modern
mobility and convenience [107].
Indeed, electrochemical nanostructure-based devices offer an important solution
to this dilemma, with the prospect of providing something approaching a sustain-
able standard of living for the entire world (so far enjoyed only by industrialized
countries), since they provide the capability for developing more efficient devices in
all areas involving electrochemical alternative energy while causing a low environ-
mental impact (sustainable development).

3 Overview of the Principles of Operation of Energy Conversion


and Storage Devices

To understand the basic principles of operation of an energy conversion or storage


device it is important to know what an electrochemical cell is. Basically, it is a
device in which a chemical reaction either generates or is caused by an electric
current. A galvanic cell is an electrochemical cell in which a spontaneous chemical
reaction is used to generate an electric current. An electrolytic cell, in turn, is an
electrochemical cell in which a reaction is driven in its nonspontaneous direction
by an externally applied electric current. There are three types of galvanic cell: the
primary, the secondary, and the fuel cell [5, 6].
Primary galvanic cells are those in which the reactants are built-in during the
manufacturing process, while secondary cells are those that must be charged (i.e.,
Electrochemistry, Nanomaterials, and Nanostructures 95

1.00 20.4 nM

0.75 10.2 nM

0.50 0.0 nM

0.25

0
0 0.25 0.50 0.75 1.00
m

Fig. 2 Design of 2-D and 3-D nanostructures from 0-D nanostructures. Both 2-D and 3-D nanos-
tructures are very commonly applied as electroactive materials in the manufacture of nanostruc-
tured porous electrodes

used as electrolytic cells) before the reactants are present for the reverse, sponta-
neous reaction [5, 6]. Fuel cells are the type of galvanic cells in which the reactants
are supplied continuously as they produce current.
An electrochemical cell comprises two electrodes in contact with an electrolyte,
i.e., an ionic conductor. These electrodes are called cathodes and anodes. A cath-
ode is the electrode that acts as a source of site for reduction reactions in the elec-
trochemical cell, while an anode is the electrode that acts as a source of site for
oxidation reactions. Therefore, the removal of electrons that occurs in the anode can
be described as follows:
Ared Aox + e , (1)
96 P.R. Bueno and C. Gabrielli

in which Ared is the reduced form of a site or species and Aox is the oxidized site or
species in the anode. The following reaction occurs in the cathode:

Cox + e Cred , (2)

in which, in turn, Cox corresponds to the oxidized form of a site or species in


the cathode, while Cred is the reduced site or species in the cathode. Many oxi-
dation and/or reduction reactions are also accompanied by the transfer of atoms or
molecules.
Batteries and fuel cells operate according to the same fundamental principle. In
other words, under appropriate conditions, the chemical free energy associated with
a particular reaction can be converted into electrical energy, often with extremely
high efficiency. Considering that both are electrochemical cells, it is possible to
demonstrate that during operation, electrons flow from the anode to the cathode,
constituting an electrical current that can be used to drive an external device, e.g.,
see Figs. 3.3 and 3.5. The fuel is an inherent part of the battery; a battery, in other
words, carries its fuel around with it, whereas fuel must be supplied to a fuel cell
from an external source. Unlike a battery, a fuel cell cannot become depleted and
so long as fuels are supplied it will generate electricity. Recharging problems are
therefore peculiar to batteries [107].
These systems can therefore be understood in terms of parts, since the overall
chemical reactions can be broken down into components of oxidation and reduction
reactions, the former involving the transfer of electrons from one set of reactants
to the anode and the latter involving the transfer of electrons from the cathode to a

CHARGE

Power
supply

e
Load

O
DISCHARGE
Co

Li
ELECTROLYTE

Carbon

CATHODE ANODE
Li1-xCoO2 Graphite

Fig. 3 Schematic of a rechargeable lithium battery in discharge/charge mode. The lithium ion is
intercalated in the anode during charging and in the cathodes during discharging. The layered host
lattice in the cathode and anode is also illustrated, considering a cathode composed of a lithium
cobalt host and an anode composed of a crystalline structure of hexagonal graphite. See Color
Plates
Electrochemistry, Nanomaterials, and Nanostructures 97

second set of reactants. The overall electrochemical device may be set up in such a
way that electron-transfer reactions are rapid and, ideally, completely reversible in
chemical terms [6].

3.1 Lithium Ion Batteries

Lithium-ion rechargeable battery solid state devices rank high among the alterna-
tive energy devices in which nanostructured cathodes or anodes make a dramatic
difference in improving rate capabilities [79, 81, 82, 86, 107, 113117]. Despite its
low density (0.53 g cm3 ), low electronegativity and high electron/atom mass ratio,
lithium is the preferred choice for the active element of the anode, whose discharge
functions as an electron donor can be expressed as:
discharge

xLi charge xLi+ (electrolyte) + xe , (3)

where Li+ enters the electrolyte, and the electron exits the anode, moving to the
external circuit to power the load. Other types of materials containing lithium (lithi-
ated host materials) are also used as anodes [118120]. The most classic example is
lithiated carbon (Lix C6 ). The anode reaction for this type of material is as follows:
discharge

Lix C charge C + xLi+ (electrolyte) + xe . (4)

The carbon host materials that have been studied include natural and synthetic
graphites, carbon fibers, and mesocarbons, all of which differ in degree of
crystallization and stacking order, but all of which have the characteristic struc-
tural feature of graphite, namely, planar layers of carbon atoms forming fused six-
membered rings and separated by intercalate layers, supplying an electrochemical
potential close to that of metallic lithium electrodes [121]. The large number of
boundaries resulting from the use of graphite nanostructures has proved useful for
improving lithium intercalation capacity [102, 107, 121]. Indeed, the interfacial
boundary area can accommodate lithium to form Lix C with x > 1, and hence, an
increased reversible capacity [102, 121].
The following reaction generally takes place at the cathode:
discharge
xLi+ (electrolyte) + xe + MO2 Lix MO2 , (5)

which clearly indicates that to operate effectively, Lix MO2 (which represents a
lithiated transition-metal oxide usually applied in commercial cells) must conduct
electronically, or at least be miscible with, an appropriate inert conducting ionic
electronic nanocomposite, in which the diffusion of Li ions must be reasonably
98 P.R. Bueno and C. Gabrielli

facile and, ideally, highly reversible to enable the battery to be rechargeable. The
latter requirement suggests that layered structures, which allow Li+ ions to diffuse
easily throughout the bulk of the crystal, are likely the most effective structure. In the
cathode, Lix MO2 is usually mixed with carbon black to raise the electronic conduc-
tivity. Lithium cobalt and nickel oxides are commonly used as cathodes (Li1y CoO2
and Li1y NiO2 ), since these oxides possess the transition metal in a highly oxidized
state, allowing a high cell voltage to be developed. Considering the lithium cobalt
oxide, one may have
discharge
xLi+ (electrolyte) + xe + LiCoO2 Li1+x CoO2 . (6)

Therefore, in the cathode, Li+ ions engage in an electron-transfer reaction that


decreases the chemical potential of lithium in relation to its value in the anode and
calls for the compensating electron. Upon discharge, the cathode functions as an
electron acceptor and the previous reaction (6) can be expressed alternatively as
follows:
xCo4+ + xe xCo3+ .
cathode
(7)
If the battery is to be rechargeable, the reactions must be reversible. The importance
of lithiated transition-metal oxides (Lix MO2 ) is that they are capable of accommo-
dating large quantities of lithium per formula unit and have low relative molecular
masses, giving rise to high power and energy densities.
Materials for cathodes and anodes (insertion electrodes) for rechargeable lithium
batteries are called intercalation compounds and constitute a special class of elec-
troactive material [122]. The intercalation refers to the reversible insertion of mobile
guest species into a crystalline host lattice, which contains an interconnected system
of empty lattice sites of appropriate size, while the structural integrity of the host
lattice is formally conserved [122].
Considering electrodes composed of nanoscale electroactive materials, high en-
ergy density and high power density (the same as rate capabilities) can be achieved
simultaneously. This requires a large electrodeelectrolyte interfacial area coupled
with short diffusion distances within the electrodes themselves [1013, 86, 113,
118120]. Nanoscale materials for lithium ion storage devices are emerging as a
successful solution for improving the rate capability [86, 113, 123125]. A batterys
rate capability is its ability to deliver a large capacity when discharged at high C
rates (a rate of C/1 corresponds to the current required to completely discharge an
electrode in 1 h). During high-power pulses required for transmission of digitized
and compressed voice data, the batterys delivery capacity decreases to a fraction of
its low rate value. It is widely believed that these limitations in the rate capabilities of
Li ion batteries are caused by slow solid-state diffusion of Li+ within the electrode
materials. For this reason, tremendous interest currently focuses on the research
and development of nanostructured Li ion battery electrodes, whose nanostructure
clearly restricts the distance that Li+ must diffuse, which may be as small as 50 nm
[1013,86,113,118120]. Two kinds of geometry are commonly applied to achieve
a faster solid-state diffusion, as depicted in Fig. 4. The first geometry is based on
Electrochemistry, Nanomaterials, and Nanostructures 99

a <100 nm

<100 nm

0 L

Fig. 4 Schematic of a highly porous nanostructured electrode protruding from a current collector.
a Nanofiber structure or a template-like membrane, and b spherical-like particle-based nanostruc-
ture. Note that real nanostructures are not so regular, but both nanostructured phase and electrolyte,
as discussed herein, are continuous along the length L [113]. The length L determines if the nanos-
tructured electrode will be considered 2-D or 3-D and is very important in modeling electronic and
ionic transport in the nanostructured electrode [86, 113]

connected spherical-like nanoparticles while the second is based on nanofibers that


protrude, brushlike, from the current collector; in fact, this type of nanostructured
electrode geometry allows for improved rate capabilities. The use of nanostructured
materials for batteries is not restricted to lithium ion batteries. Nickel and metal
hybrid nanocrystalline materials have also been developed, and nanostructured ma-
terials offer improvements in terms of power density and durability by controlling
the charge diffusion and oxidation state on a nanoscale level.
All the electrochemical cell devices discussed in this section are based on 2-D
electrode configurations in which electrodes can be assembled from 0-D or 1-D
nanostructures (see Figs. 2 and 4). In other words, commercial lithium ion batteries
are based on a layer-by-layer construction of the cell, whether or not the layers are
based on nanoscale materials. The use of novel nanoscale materials will offer im-
provements in gravimetric and volumetric energy densities (W h g1 and W h L1 ,
respectively), but the same configurations will prevail. Furthermore, cells designed
in 3-D nanostructures are also emerging as a new possibility for specific applications
in the category of lithium ion batteries, i.e., as batteries to power microelectrome-
chanical systems, or MEMS. Thus, thin-film batteries, despite their excellent energy
100 P.R. Bueno and C. Gabrielli

per unit volume, fall far short of being able to power a smart dust mote for 1 day,
and the consequences of the 2-D nature of thin-film batteries are easily overlooked.
Three-dimensional configurations offer a means to keep transport distances short
and yet provide enough material so that the batteries are able to power MEMS de-
vices for extended periods of time [85, 126].
It is important to stress that 3-D cells are different from the 3-D electrodes con-
cept. Examples of prospective 3-D cell architectures for charge-insertion batteries
are (a) array of interdigitated cylindrical cathodes and anodes, (b) interdigitated
plate array of cathodes and anodes, (c) rod array of cylindrical anodes coated with a
thin layer of ion-conducting dielectric (electrolyte) with the remaining free volume
filled with the cathode material, (d) aperiodic sponge architectures in which the solid
network of the sponge serves as the charge insertion cathode, which is coated with
an ultrathin layer of ion-conducting dielectric (electrolyte), and the remaining free
volume is filled with an interpenetrating, continuous anode [85, 126].

3.2 Fuel Cells

As mentioned earlier, the fuel cell is a particular type of galvanic cell, as briefly
described later. There are high-temperature and low-temperature fuel cells [11, 35,
109, 127]. Here, we will simply describe two common types of fuel cell to illustrate
their main principles of operation. At sufficiently high temperatures it is possible to
use solid ceramic-oxide ion conductors that have very high conductivities exceed-
ing 900 C. The typical electrolyte used is ZrO2 with 810 mol% of Y2 O3 , which is
particularly O2 ion-conducting with a minimum of electronic conductivity. Elec-
trolyte for this kind of fuel cell must exhibit good chemical stability with respect to
the electrodes and inlet gases; it must have a high density to inhibit the crossover of
fuel, and its thermal expansion must be compatible with other components. Anode
and cathode must be porous to physically offer low resistance to the transport of fuel.
The anode requirements are that it must be an effective oxidation catalyst, have high
electronic conductivity, and be stable in the reducing environment. For this kind of
cell, the material traditionally used as anode is 35% Ni ZrO2 /Y2 O3 cermet, i.e., a
well-mixed combination of ceramic and metal. The anode reaction is as follows:

H2 + O2 H2 O + 2e .
anode
(8)

The cathode must also display good electrocatalyst activity for the reduction of O2
and offer good electronic conductivity, since it must serve as the current collec-
tor. The cathode material most commonly used is porous or mesoporous perovskite
manganite with the formula La1x Srx MnO3 (0.10 < x < 0.15). The cathode reaction
is as follows:
1
O2 + 2e O2 .
cathode
(9)
2
Electrochemistry, Nanomaterials, and Nanostructures 101

a Load b Load

H2 O2 H2 O2

H2 H2 O H2 O O2

ANODE CATHODE ANODE CATHODE


diffusion path diffusion path

H O O2- H
+
e

Fig. 5 Schematic cross section of the simplified planar anodeelectrodecathode structure of two
typical fuel cells: a polymer-electrolyte membrane fuel cell and b solid oxide fuel cell. See Color
Plates

Cells that operate using the aforementioned cathode, anode, and electrolyte are
called solid-oxide fuel cells (SOFC) (see Fig. 3.5b for a schematic representation
of the operation of such cells). Another important fuel cell that can be built in anal-
ogy to the solid oxide-based fuel cell is based on solid proton conductors. This
cell is a low-temperature fuel cell (about 150 C), because high-temperature solid
proton conductors are still difficult or nigh impossible to obtain. In fact, they may
not even exist at all, in view of the obvious fact that, at high temperatures, all hy-
drated oxides tend to lose water and if conduction takes place at all, it must do
so through the metal or the oxide ion. Several inorganic or organic solid proton
conductors exist. When an organic conductor is used, it leads to the so-called solid-
polymer-electrolyte fuel cell (SPEFE) or polymer-electrolyte-membrane fuel cell
(PEMFC) (see Fig. 3.5a). In the case of SPEFE or PEMFC, the electrolyte is criti-
cal and is composed of a polymeric-membrane proton conductor based on polymers
containing perfluorinated sulfonic-acid groups. In their simplest designs, anodes and
cathodes are formed either directly from metal particles or from catalyzed carbon
particles bound to the membrane. Current collectors are porous plates of carbon or
graphite and the cell reactions are as follows:

H2 2H+ + 2e ,
anode
(10)
1
O2 + 2H+ + 2e H2 O.
cathode
(11)
2
Low operating temperatures dictate the use of noble-metal catalysts, and particular
problems are experienced with the cathodes. Nanomaterials, e.g., clusters of Pt or
Pt/Ru or even other noble metals, are used in catalytic electrode layers, e.g., the
membrane electrode of fuel cells [90]. It is important to mention that ZrO2 -based
electrolyte, such as nanoscale materials in SOFC systems, can be used to decrease
102 P.R. Bueno and C. Gabrielli

the operating temperature of the fuel cell by increasing the ionic conductivity [102].
Hence, intermediate temperature SOFC development will be tremendously influ-
enced by nanostructured electroceramics.

3.3 Photoelectrochemical Solar Cells

We have seen how lithium ion batteries and fuel cells generally operate and how
they are dependent on the prior conversion of energy into electrical energy. Solar
energy conversion devices are able to complete the cycle by finding ways of directly
converting solar energy into chemical or electrical energy, thereby breaking free
of dependence on fossil fuels. Here, our purpose is to show the operating princi-
ples of DSSC cells and how nanostructures influence their efficiency and operation
[71, 128].
In the manufacture of photoelectrochemical solar cell it is important to establish
contact between the n-type semiconductor and an electrolyte in the presence of a re-
dox couple of appropriate reversibility. Illumination of this junction then generates
holes and electrons that are separated by the field, giving rise to an electrical cur-
rent. In other words, the energy contained in the light allows current to flow through
the circuit from the semiconductor to the metal counterelectrode, again generating
an electrical current [71]. With this type of photoelectrochemical cell, although it is
possible to make relatively efficient devices, they are frequently plagued by stability-
related problems because the semiconductor is so easily corroded by the electrolyte
under the strongly oxidizing or reducing conditions at the interface. Another variant
is a photogalvanic cell, which can be built provided its configuration is combined
with materials called sensitizers, i.e., dyes, in which case it is possible to obtain a
photoelectrochemical cell called a dye-sensitized solar cell (DSSC). According to
Fig. 6, which illustrates the mode of operation of a DSSC, the dye (D) absorbs a
photon, with the concomitant excitation of an electron from the highest occupied
molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO).
Provided the LUMO lies just above the conduction-band edge in energy, electrons
are injected into the conduction band of the semiconductor, leaving an oxidized form
of the dye on the surface. If a kinetically facile redox couple is present in solution,
this oxidized form of the dye can be rereduced, allowing the cycle to begin again.
It is important to keep in mind that electrolyte is based on organic solvent contain-
ing iodide-tri-iodide at the counterelectrode, and that the iodide is regenerated by
reduction of tri-iodide at the counterelectrode (i.e., I
3 + 2e 3I ) and the circuit

is completed by diffusion of the I back to the dye-sensitized electrode (see Fig. 6
for details) [71].
The semiconductor nanoparticle most commonly used is the anatase form of
TiO2 ; however, other mesoscopic oxides can also be used, such as SnO2 , ZnO,
Nb2 O5 , WO3 , and Ta2 O5 . Dye molecules are chemisorbed onto the semiconduc-
tor (Fig. 6b), forming close to a monolayer. This dyesemiconductor assembly
is arranged over a transparent conducting oxide and the set is put into contact
Electrochemistry, Nanomaterials, and Nanostructures 103

a D+/D* 4

h 4 6

h
Vph 5
O/R O/R
1 3

SnO2/Pt
2
D+/D

b nanostructured semiconducting
+
N oxide glass substrate

N+ OH
P
O
O

OH
+ +
N N P O

OH
O
+ P
N O

+
adsorbed dye molecules
N

Fig. 6 Schematic of DSSC operation for conversion of light into electrical energy. a Represents the
different steps involved in energy conversion. The horizontal line represents the Fermi level/redox
potential/electrochemical potential of the electron. Step 1 represents the regeneration of iodide by
the reduction of tri-iodide at the counter electrode, step 2: diffusion of I to dye-sensitized elec-
trode, step 3: restoration of dye to the original state through electron donation by the electrolyte,
step 4: photoexcitation of the dye and the resulting injection into the conduction band on the semi-
conductiong oxide, and finally, step 5: recombination. b Represents the dye molecule adsorbed on
nanoparticles (based on appropriate semiconducting oxide) of the nanostructured electrode

with a redox electrolyte (as mentioned earlier, the electrolyte most commonly em-
ployed is the iodide-tri-iodide redox system embedded in an organic solvent). This
serves to close the electrical circuit with a second transparent conducting electrode
(counter electrode) in which a catalyst is deposited to facilitate the redox reaction
[71, 79, 80, 87].
As can be seen, the heart of the device is the semiconducting mesoporous or
nanocrystalline oxide, composed of a network of nanoparticles that have been sin-
tered together to establish electronic conduction, which is characterized by the total
depletion of the semiconductor due to the small size of the nanoparticles and porous
structure [65]. The Fermi level in the dark is therefore near the bandgap center,
allowing for the generation of high photovoltages upon illumination [71].
Another important point to be emphasized regarding the nanosize effect in pho-
toelectrochemical processes is the photoelectrolysis of water for hydrogen produc-
tion [71] through solar energy. Unmistakable nanosize effects can be observed in
photoelectrochemical processes involving semiconducting particles separated from
the metal surface by barrier layers. The size effects are triggered by a shift of the
absorption edge when the particle diameter ranges from 10 to 20 nm [129, 130].
Photoelectrochemical processes are substantially accelerated when they pass from
massive to nanosize semiconductors.
104 P.R. Bueno and C. Gabrielli

3.4 Electrochemical Double-Layer Capacitors

Electrochemical double-layer capacitors (EDLC) or electrochemical supercapaci-


tors (double-layer structures are explained in the next section) are like lithium ion
batteries, i.e., energy storage devices [1624]. However, in terms of energy stor-
age capacity, EDLCs fall in the gap between batteries and conventional dielectric
devices, although EDLCs have much higher power and a much longer cycle life
(at least two orders of magnitude) [131]. An important point about EDLCs is the
optimization of pore-size distribution for easy electrolyte access [132134]. Nan-
otechnology is therefore evidently important not only from this standpoint but also
because it enables large surface areas to be obtained to optimize performance in
terms of capacitance and ionic and electronic resistance. The optimization of the
operation depends on the electrolyte-supported voltages to reach the highest possi-
ble level; the amount of energy stored is given by the classical equation:

CV 2
E= , (12)
2
where V is the voltage and C the capacitance. Certain electrolytes display higher
specific resistance than aqueous ones, resulting in an increase in the equivalent series
resistance, which, in turn, affects the maximum power according to

V2
P= , (13)
4R
where R is the equivalent series resistance. Carbon materials contribute consider-
ably to equivalent series resistances, which include electronic and ionic components
for charging small pores rather than the ionic resistance between electrodes. There-
fore, to yield high-capacitance and low-resistance electrodes, it is very important to
develop materials with selected pore-size distribution [135137].
Nanostructured carbons have been developed to optimize high surface area in
terms of capacitance and ionic and electronic resistance. In carbon nanotubes, the
main focus has been optimization of interparticle contact resistance and electrolyte
wettability of the pores [17, 138140]. Since the pores are open spaces in the entan-
gled network, they are all connected and easily accessible surface areas possessing
low charging resistance. Nanostructured metal oxide materials have also been inves-
tigated for supercapacitor applications, especially RuO2 and RuO2 xH2 O nanotube
composites [20, 22, 24].

4 Concepts of Electrochemistry

The general concepts and techniques used in more classical electrochemistry can be
employed when electrodes or electrolytes are nanomaterials or nanostructured films.
However, it is important to consider appropriate boundary conditions. The goal
Electrochemistry, Nanomaterials, and Nanostructures 105

of this section is to present the basic aspects of electrochemistry and techniques


based on electrochemistry taking into account the classical boundary condition.
The concepts could then be extended to other boundary conditions, for instance,
when nanostructures are required. As in classical electrochemistry using solid elec-
trodes and aqueous solutions of electrolytes, the initial aim of experiments is to
separate charge transfer at the electrode from the mass transport of reactive species.
After that, the reaction mechanism involved in the electrode must be detailed. To
understand how the current changes over time, for time techniques or frequency
techniques are related to the elementary processes, some fundamentals of elec-
trochemistry and an understanding of the methods theoretical backgrounds are
helpful. We will begin by summarizing the fundamental concepts and then go on
to the classical techniques: chronopotentiometry, chroamperometry, cyclic voltam-
metry, impedance spectroscopy, . . . , which will be briefly reviewed in this section
[97, 141153].

4.1 Fundamental Concepts

An electrochemical reaction is a heterogeneous chemical process involving the


transfer of electrons between a metal or a semiconductor the electrode, and an
ionic conductor the electrolyte [141, 146]. The electrode reaction may be an an-
odic (or cathodic) process when a species is oxidized (or reduced) by the loss (or
gain) of electrons to (or from) the electrode, according to the short picture given pre-
viously. When electrolysis occurs, in addition to electron transfer at the electrode,
ions must pass through the solution. The current I is a convenient measure of the rate
of the global reaction and the charge q, passed during a period t, indicates the total
amount of chemical reaction that has taken place. The charge required to convert m
moles of material in an electrode reaction involving the transfer of n electrons per
molecule is calculated using Faradays law:
t
q= I dt = mnF, (14)
0

where F is the Faraday (96, 500 C mol1 ).


As explained earlier, an electrochemical cell requires at least two electrodes an
anode and a cathode to enable a current to flow through it. The rate of electrolysis
will depend on the kinetics of the two electrode reactions. It is usually essential to
have an overpotential, , to increase the rate at which an electrode reaction occurs.
The total cell voltage required to bring about chemical changes by electrolysis is
given by:
V = E0C E0A |A | |C | IRcell , (15)
where E0C E0A is the difference between the reversible potentials of the cathode
and the anode in the cell, and Rcell is the resistance of the electrolyte solution.
In practice, the potential of the studied electrode is measured with respect to a
106 P.R. Bueno and C. Gabrielli

reference electrode immersed in the cell. Its value is generally imposed by using
a potentiostat whose details are given in all textbooks, which is used to polarize this
three-electrode cell.

4.1.1 Kinetics of Electron Transfer Reaction at Interfaces

The faradaic current relative to the following electron transfer reaction

k1
O + ne

R, (16)
+k2

can be written as follows:


I j
= = k2 cR (0) k1 cO (0), (17)
nFA nF
where j is the current density.

i.e., j = ja + jc ,

where ja = nFk2 cR (0,t) and jc = nFk1 cO (0,t) are the anodic and cathodic com-
ponents of the total faradaic current.
The concentrations involved in the expression of the rate are those existing at
the location of the reaction, i.e., at the electrode surface, which is x = 0, while ka
and kc are the rate constants characteristic of the two processes of oxidation (anodic
direction ka ) and reduction (cathodic direction kc ), (unit: cm s1 ). Their value
changes with the potential as follows:
 
(1 )nF
k1 = k exp (E E ) , (18)
RT
 
nF
kc = k exp (E E ) , (19)
RT

where is the charge transfer coefficient.


With such laws, when E increases, the oxidation rate constant increases expo-
nentially, whereas the reduction rate constants tend to 0 and vice-versa when E
decreases. This behavior corresponds to the variation of the current density with
respect to the potential E, such as:
    
nF (1 ) nF
j = nFk cR (0,t) exp (E E O ) cOx (0,t) exp (E E O )
RT RT
(20)

At equilibrium potential, Eeq , one has I = 0, therefore,


Electrochemistry, Nanomaterials, and Nanostructures 107
   
nF (1 )nF
cR (0,t) exp (Eeq E ) = cO (0,t) exp
O
(Eeq E )
O
(21)
RT RT

In addition, as there is no current, the concentrations at the electrode of species R


and O are equal to their value in the solution, so:

RT cO
Eeq = E O + ln . (22)
nF cR

At this potential, | j1 | = | j2 | = jeq , which is the exchange current at equilibrium


potential (E = Eeq ).
The current density, j, can be written in the ButlerVolmer form with respect to
the overvoltage = E Eeq :
    
cR (0,t) nF cO (0,t) (1 )nF
j = jeq exp exp . (23)
cR RT cO RT

Figure 7 shows the currentvoltage curve j = f ( ) and the partial current densities
ja = f ( ) and jc = f ( ). Hence, j > 0 when | ja | > | jc | (for > 0), and j < 0 for
| ja | < | jc | (for < 0). At equilibrium where j = 0, | ja | = | jc | = jeq .
At low overvoltages:
nF
j jeq . (24)
RT

I/
Ie
Ii
1.0

0.8
Ia
0.6

0.4 Total current I


0.2
100 200 300 400

400 300 200 100 h, mV


Eeq 0.2 Ic

Ieq 0.4

0.6

0.8

1.0
Ii

Fig. 7 Currentvoltage curve j = f ( ) for mass transport limitation (continuous curve) and partial
current densities ja = f ( ) and jc = f ( ) (dashed line), limiting diffusion current densities: jal
and jcl . For comparison, the curve j = f ( ) where kinetics is the only rate-limiting step, i.e.,
without mass transport limitation is plotted (dotted line)
108 P.R. Bueno and C. Gabrielli

In the linear range of the currentpotential relationship, the charge transfer is equiv-
alent to a resistance called charge transfer resistance at the equilibrium potential,
(Rt )eq
RT 1
(Rt )eq = . (25)
nF jeq
In contrast, far from the equilibrium potential (large overvoltages), one of the two
exponential functions is preponderant. Hence:
For  0, the rate of the reduction process (cathodic) is practically 0 and the
oxidation process (anodic) is dominant.
 
cR (0,t) nF
j ja = jeq exp . (26)
cR RT

This relationship called Tafel relationship was proposed empirically in 1905.


For
0, on the contrary, it is the reduction process (cathodic) that is practically
preponderant:
 
cO (0,t) (1 ) nF
j jc = jeq exp . (27)
cO RT

The system is considered fast if jeq is large, i.e., a large value of k and a low value of
the charge transfer resistance are associated. Conversely, when jeq is low, k is also
low while the charge transfer resistance is large, so the system is considered slow.

4.1.2 Mass Transport Phenomena Involved in Electrochemical Processes

Because electrochemical reactions are localized at the contact between the


electrolyte and the electrode, an electrochemical reaction can evolve only if the
electroactive species involved is present on the surface. This means that the dis-
solved species arrives at the electrode surface by some transport process through
the electrolyte in which it is initially homogeneously distributed. Various transport
modes can occur [147, 149]:
For ionic species, an electric field in the solution produces transport by ionic
electromigration.
The consumption or the production of a species at the electrode surface leads to
a decrease or an increase of the concentration of this species in the vicinity of the
electrode. This concentration gradient produces transport by diffusion.
In liquid electrolytes, a movement of the ensemble of the electrolyte produces
transport by convection (i.e., stirring or flowing of the electrolyte or electrode
movement).
In the permanent regime, equilibrium is established between the fluxes imposed
by the various transport processes on the one hand, and the flux of the electric
Electrochemistry, Nanomaterials, and Nanostructures 109

charges transferred across the interface to produce the electrode reaction on the
other hand. This equilibrium is traduced by a relationship between the mass flux
and the current density corresponding to the electrochemical reaction. This relation-
ship implies that the global rate of the electrochemical process depends not only on
the kinetics of the reaction occurring at the electrode but also on the mass transport
kinetics. Therefore, the rate-determining step of the global process can be controlled
by either the reaction kinetics or the mass transport, or both.


Ionic electromigration in bulk electrolytes. The electric field E = (where
is the potential in the bulk electrolyte) is the driving force for the electric charges.
Under its influence each ion gains a speed:



v i = ui , (28)

where ui is the electric mobility of the ion, positive for a cation, and negative for an
anion.
Because of the opposite signs of their charge, anions and cations migrate
in opposite directions: anions migrate toward the anode and cations toward the
cathode.
This ionic movement of all the mobile ions maintains the electroneutrality of the
solution.
The flux density of the ion, Ji , is defined as the number of moles of the ion
crossing a unit surface perpendicular to the transport direction per time unit.



J i = ui ci (Ji in mol cm2 s1 ). (29)

The electric quantity brought by the flux of the ion i (with charge zi ) is the current
density (charge flux):




j i = Fzi J i = Fzi ui ci . (30)
If the migration of each ion is independent of the migration of the others, which is
valid at infinite dilution, then:




j = j i = F (zi ui ci ) , (31)
i i




i.e., j = E , where = F i zi ui ci is the conductivity of the electrolyte (S or 1 ),
which is the sum of the partial conductivities i = Fzi ui ci for each species i.
Transport by diffusion. The electrolytic current can be related to the mass flux
densities at the electrode surface, Jid . For an electrochemical reaction, one has

O + ne R, (32)
j = nFJOd = nFJRd , (33)

by counting the fluxes leaving the electrode positively and the fluxes going toward
the electrode negatively.
110 P.R. Bueno and C. Gabrielli

The flux density of species i transported by diffusion is equal to:



d

J i = Di ci . (34)

For a one-dimensional diffusion, Ficks first law is obtained:


d ci
J i = Di . (35)
x
In the same way, if Di is independent of ci , Ficks second law is as follows:

ci 2 ci
= Di 2 . (36)
t x
Transport by diffusion-convection. If transport by convection (natural or mechani-
cally imposed), which imposes a speed v on the electrolyte, is taken into account,
the convective diffusion law is obtained. When a diffusion gradient is superimposed
on this convection regime, the density of the global flux is as follows:

dc
d c

J i = J i + J i = Di ci + ci

v. (37)

In this condition, Ficks second law for one-dimensional transport is as follows:

ci 2 ci ci
= Di 2 v . (38)
t x x
The solution for these equations requires complete knowledge of the velocity of the
liquid, which depends on the mechanical device used to impose the convection.
The differential equation of the steady-state regime is obtained for ci / t 0 in
the general equation, which becomes as follows:

2 ci ci
Di =v . (39)
x2 x
A very simplified model of the convective diffusion was introduced in electrochem-
istry by Nernst (1904), which is based on the hypothesis of the formation, at the
electrode surface, of a motionless limiting layer with a thickness N where diffusion
occurs.
Hence, in the steady-state regime, the concentration gradients and the diffusion
fluxes remain restricted in a layer, called the Nernst diffusion layer, close to the
electrode. In this layer, Ficks equations are easily solved:

2 ci
=0 (40)
x2
leads to
ci c0 ci
= cte = i , (41)
x N
Electrochemistry, Nanomaterials, and Nanostructures 111

where ci is the concentration at the external limit of the diffusion layer (x N ),


i.e., the concentration of the species in the homogeneous solution, and c0i = c(x = 0)
is the concentration at the other limit, at the electrode. Hence, the steady-state flux
is equal to:
c0 ci
d
Ji,stat = Di i (42)
N
( ci / x > 0 and Ji < 0 for a species consumed at the electrode and ci / x <
0 and Ji > 0 for a species produced).
The steady-state current density related to this diffusion flux is as follows:
   
 nF d   nF d 0 
stat    
| j | =  Ji,stat  =  ki (ci ci ) , (43)
i i

where kid = Di /N (cm s-1) is the diffusion transport rate across the Nernst layer.
The profile of the concentration is given in Fig. 8.
In reality, instead of the angular point at x = N , the concentration profile changes
progressively from a linear variation close to the electrode to c = c far from the
electrode.

4.1.3 The Electric Double Layer at Interfaces

The accumulation of electric charges (nontransferable across the interface in the


absence of electrochemical process) on each part of the interface results from the
existence of mobile charge carriers in the two phases in contact with each other
and an interfacial potential difference ( El/sol = El sol ) [153]. The charge
accumulated on one side of the interface is counterbalanced by the charge accu-
mulated on the other side:
qsol = qEl , (44)
where q represents the charge per surface unit (C cm2 ). The charge brought by
the electrode itself (qEl ) is constituted either by an excess of electrons (negative
charge) or by a shortage of electrons (positive charge), depending on the sign of .
The compensating charge on the electrolytic solution side is due either to an ex-
cess of cations compared with anions, for < 0 (when the electrode is negatively
charged), or to an excess of anions compared with the cations (when the electrode
is positively charged).
the thickness of the
If the charged layer at the electrode surface is thin (< 0.1 A),
layer in the solution where the ionic distribution is not electrically neutral is much
larger because of the size of the solvated ions.
Double-layer structure. It is generally assumed that the ions can approach the
electrode only at a distance of a few angstroms. Their centers are located in a plane
(called Helmholtz plane). The Helmholtz layer (or compact layer) would contain
only solvent molecules oriented by the electric field.
112 P.R. Bueno and C. Gabrielli

a C(x)

Cs

(real profile)

C* steady-state
flux

0 N x

steady-state homogeneous
diffusion layer solution
(NERNST hypothesis)

b C(x)

steady-state
C* flux

Cs

0 N x

Fig. 8 Concentration profile of the electroactive species dissolved in the diffusion layer. a Species
consumed by the electrochemical reaction, b species produced by the electrochemical reaction.
The concentration profiles are represented in the Nernst hypothesis (continuous curve), and in real
values

However, this is actually more complex. The ions that are not adsorbed (gen-
erally cations) are kept at a certain distance from the electrode by their solvation
shell and by a layer of solvent molecules adsorbed on the electrode. Conversely,
those that are specifically adsorbed (anions) are directly in contact with the elec-
trode surface. Therefore, there are two Helmhotz planes. On the other hand, the ions
accumulated close to the electrode are under the influence of the thermal movement.
They constitute the diffuse layer (or GouyChapman layer).
Electrochemistry, Nanomaterials, and Nanostructures 113

The ensemble Helmholtz layer/GouyChapman layer constitutes the electro-


chemical double layer. Its thickness is in the order of a few tens of Angstroms. This
layer is generally represented by the series combination of two capacitances relative
to the diffuse and compact layers, Cdiff and Ccomp . The capacity of the double layer,
Cd , is thus equal to:
1 1 1
= + . (45)
Cd Cdiff Ccomp
Capacitive current. The charging of the electrode/solution interface leads to a ca-
pacitive current when a variation of the interfacial charge occurs:

dqEl
jcapa = . (46)
dt
This phenomenon occurs when the electrode potential is changed as qEl changes
with E. The capacitive current shows an exponential decrease relating to the values
of the double layer capacity, Cd , and the cell resistance, Rcell . The decrease is very
short (lower than 1 ms) since the time constant, = Cd Rcell , is in the order of a few
tens of microseconds.
If a continuous change of the potential is imposed at the electrode, a permanent
capacitive current is observed:

dqEl dE dE
jcapa = = Cd . (47)
dE dt dt
If E varies like E = Ei + bt, a continuous capacitive current appears, which is pro-
portional to the rate b.

4.2 Techniques Used for Investigating Electrode Reactions

To determine the elementary processes involved in a reaction mechanism occurring


at an electrode/electrolyte interface (mass transport, chemical, and/or electrochemi-
cal reactions) requires the use of techniques to control the state of the electrode and
to analyze the behavior of the interface. One begins by studying the steady-state
regime. Although this study sometimes suffices for simple processes, it proves in-
adequate as the degree of complexity of the processes and their coupling increases.
Nonsteady-state techniques must then be used [148, 151, 153].
In electrochemistry, because electrical quantities are easy to use and provide in-
formation directly relating to the behavior of the interface, they are particularly
useful to identify interfacial processes. Contrary to other techniques, which re-
quire a vacuum chamber [low-energy electron diffraction (LEED), Auger electron
spectroscopy, etc.] or electromagnetic radiation (optical: ellipsometry, or X-rays:
EXAFS), which need no alteration of the electrode surface, electrical techniques
can be used in situ on any surface state of the electrode. In addition, thanks to
the advances in electronics, experimentalists can use more and more sophisticated
114 P.R. Bueno and C. Gabrielli

instrumentation. These electrical techniques will be only briefly presented, since de-
tails can be found in reference books. They can be transposed to other quantities:
pressure, surface, rotation velocity of the electrode, temperature, etc., which can be
used to clarify the phenomena involved at the interface between an electrode and an
electrolyte.
The principle of the nonsteady-state techniques (also called relaxation tech-
niques) is based on the fact that the steady state depends on some quantities
(potential, pressure, temperature). A perturbation of these quantities by the experi-
mentalist changes the state of the system. The rate at which it tends to a new steady
state depends on its characteristic parameters: reaction rate constants, diffusion co-
efficient, and their various couplings. Analyses of the transient regime can lead to
information on these parameters and on the phenomenological equations that relate
to them. However, due to the nonlinearity inherent to electronic transfer, the deriva-
tion and the exploitation of the response of the interface to a large amplitude signal
are often inextricable in complex processes. Therefore, low-amplitude signals are
often used in electrochemistry. However, even in this condition, the time transient
response is generally very complicated. In contrast, the frequency response is usu-
ally simpler, allowing for easy exploitation of experimental results.
By controlling the electrochemical reactions, the use of electrochemical quanti-
ties allows for kinetic studies whereby the various elementary phenomena can be
dissociated. In this way, the monoelectronic steps of the reaction mechanisms can
be distinguished and the often unstable intermediates involved in the reactions can
be counted. Although these techniques do not lead to an identification of the chem-
ical bonds or the intermediates in the chemical sense, they give information on the
rates of the reactions occurring at the electrochemical interface and provide a certain
characterization of the intermediates.
The plot of the steady state currentvoltage curve allows reliable data to be quan-
titatively obtained only for the slowest step in the overall reaction scheme. This may
suffice for very simple systems (i.e., oxidoreduction reaction without mass trans-
port limitation) but is largely insufficient for multistep mechanisms with or without
mass transport limitation. In potential step experiments, the potential of the work-
ing electrode is changed instantaneously and the currenttime response is recorded.
These techniques are known as chronoamperometry and chronopotentiometry when
the potential response to a current step is recorded. Sometimes it is appropriate to
use chronocoulometry by plotting the total charge (determined by integration of the
current) with respect to time. These techniques are restricted to pure electron trans-
fer under mass transport limitation coupled, or not, with homogeneous chemical
reactions. For more complex systems, particularly where multistep electrochemical
reactions occur, one is well advised to use small signal techniques and particularly
impedance spectroscopy.
In some very favorable situations, several techniques can have comparable ef-
fectiveness. However, when complex heterogeneous reactions interact with mass
transport, analyses of the current or potential time transients lead to poor results if
a reaction mechanism has to be resolved. A frequency analysis is then more effi-
cient. Therefore, impedance measurements, by means of a perturbation sine wave
Electrochemistry, Nanomaterials, and Nanostructures 115

signal with low amplitude in a wide frequency range, have been widely developed
[97,141153]. Examples of applications of impedance analyses to various problems
can be found in the special issues of Electrochimica Acta, which follow the Inter-
national Symposia on Electrochemical Impedance Spectroscopy (EIS) that are held
every 3 years: Electrochimica Acta 35, n 10 (1990); 38, n 14 (1993); 41, n 7/8
(1996); 44, n 24 (1999); 47, n 13/14 (2002); 51, n 8/9 (2006). On the other hand,
a special issue devoted to impedance techniques was published by J. Electroanal.
Chem. 572, n 2 (2004).
Among the nonsteady state techniques, impedance techniques are increasingly
used in electrochemistry as well as in academic studies or industrial applications to
characterize electrode processes to dissect overall electrochemical processes. These
techniques are based on an analysis of the current response (or potential) to a low-
amplitude perturbation, often sinusoidal, of the potential (or current). The measure-
ment is carried out at constant polarization potential, as soon as the electrode has
reached a steady state, by varying the analyzing frequency in a large-frequency do-
main (often from 50 kHz to 0.001 Hz). This measurement is then repeated all along
the currentvoltage curve in the studied potential range. The electrochemical system
is polarized using a potentiostat or a galvanostat. The impedance is measured with
a frequency response analyzer.
The impedance gives information on the processes occurring at the interface
(electrochemical and chemical reactions and diffusion) and on the structure of the
interface. Plotted in the complex plane (Re[Z( f )]),-Im[Z( f )), the high-frequency
limit generally gives the electrolyte resistance and the low-frequency limit the
polarization resistance (inverse of the slope of the currentvoltage curve). In the
high-frequency range, a capacitive loop is related to the parallel arrangement of
the charge transfer resistance and the double-layer capacity. In the lower frequency
range, one can observe capacitive or inductive semicircles, which represent the re-
laxations of the reaction intermediates, capacitive loops related to diffusion char-
acterized by a 45 part with respect to the real axis, and negative resistance for
passivation.
Impedance is a quantity defined for a linear system. In electrochemistry, where
nonlinear behavior occurs because of the reaction rate constants, which depend ex-
ponentially on the potential, the use of low-amplitude signal allows the system to
be linearized. Therefore, the electrochemical nonlinear system is approximated by
a linear system around the polarization point. The experimental results can be in-
terpreted in two ways. On the one hand, one can look for an equivalent circuit hav-
ing the same impedance. An example will be given in the following paragraph,
using transmission lines to describe the behavior of porous electrodes. On the other
hand, one can look for a model involving kinetic equations describing the reactions
and mass transport and, after linearization, calculate a theoretical impedance, which
can be compared with the experimental data. The latter approach will be developed
later.
To illustrate the various techniques, the same redox process limited by mass
transport will be analyzed by three different methods, namely linear voltammetry,
chronoamperometry, and impedance spectroscopy.
116 P.R. Bueno and C. Gabrielli

If an oxidoreduction reaction occurs between species O and R, with concentra-


tions cO and cR , such as:
k1
O + ne

R,
+k2

where the reaction rate constants are


   
nF nF
k1 = k10 exp (1 ) E and k2 = k20 exp E , (48)
RT RT

the concentrations of species O and R should diffuse with coefficients DO and DR in


the direction x perpendicular to the electrode surface, and their concentrations obey
the following:

cO (x,t) 2 cO (x,t)
= DO (49)
t x2
cR (x,t) cR (x,t)
2
= DR . (50)
t x2
The initial and boundary conditions describe the experimental conditions.
(a) The initial conditions, at t = 0, express the homogeneity of the electrochemical
solution before switching on the current, and the concentrations of O and R
species are then equal to cO and cR :
for t = 0
one has cO (x, 0) = cO and cR (x, 0) = cR .
(b) The boundary conditions impose, at a certain distance from the electrode sur-
face, that the concentrations of the species are equal to those found in the bulk
electrolyte, cO and cR :
for t 0 and x one has cO (x,t) cO and cR (x,t) cR .
(c) The boundary conditions at the electrode surface are imposed by the charge and
mass balance. The fluxes of the O and R species are the same and are equal to
the faradaic current (Ficks first law):
for t 0 and x = 0 one has
cO cR IF (t)
DO (0,t) = DR (0,t) = , (51)
x z nFA
where the faradaic current is given by the law of heterogeneous kinetics:

IF (t) = nFA[k2 cR (0,t) k1 cO (0,t)]. (52)

Then, at the electrode surface, (x = 0):


   
cO (x,t) cR (x,t)
DO + DR = 0. (53)
t x=0 t x=0
Electrochemistry, Nanomaterials, and Nanostructures 117

4.2.1 Linear Sweep Voltammetry

For linear sweep potential voltammetry and chronoamperometry, a reversible elec-


trochemical reaction, i.e., a redox process for which the reaction rate constants k1
and k2 are assumed to be very large, is considered. Thus, the potential can be written
as follows:
RT cO (0,t)
E = EO + ln (54)
nF cR (0,t)
 
cO (0,t) nF
or = f (t) = exp (E E ) .
O
(55)
cR (0,t) RT

For linear sweep voltammetry, the potential changes are expressed as follows:

E(t) = Ei vt, (56)

where is the rate of the potential sweep.


The flux at the electrode is proportional to the current intensity:
 
I(t) CO (x,t)
JO (0,t) = = DO . (57)
nFA x x=0

Hence, it can be shown that [146]:


1
CO (0,t) = CO [nFA( DO )1/2 ]1 i( )(t )d . (58)
O

The change of the current with respect to the potential can be obtained from the
numerical integration of:
t
i(z)dz nFAc0 ( D0 )1/2
=  1/2 , (59)
0 ( t z)1/2
1 + DDR0 exp [ (Ei vt E 0 )]

where = (nF/RT ). The results of this integration are given in Fig. 9.


If, for a reversible redox process limited by diffusion, linear cyclic voltammetry
allows quantitative results to be obtained in a relatively short time, the interpretation
of the voltammograms relative to more complicated reaction mechanisms such as
processes involving adsorbed intermediates becomes much more difficult.

4.2.2 Chronoamperometry

In the case of chronoamperometry, E(t) is a potential step. If E(t) is imposed on an


electrochemical interface where a reversible redox reaction occurs from a potential
118 P.R. Bueno and C. Gabrielli

Ep
ip
E1/2
0.4

0.3
Ep/2
Current

0.2

0.1

0
+100 0 100
E-E1/2

Fig. 9 Voltamperogram with linear sweep of potential

where there is no current to a more negative potential, the current response, I(t), can
be obtained by using the Laplace transform:

nFAD0 c0 1
1/2
I(t) = (60)
t 1 +

where 
 
nF(E E O ) DO
= exp and = . (61)
RT DR
The limiting diffusion current, Il , obtained for E , is equal to:

nFAD1/2 c
Il = O O, (62)
t

i.e., I = Il /(1 + ) and E = E1/2 + (RT /nF) ln(Il I)/I, where the half-wave
potential is equal to 
RT DR
E1/2 = E O + ln . (63)
nF DO
The concentration repartition with respect to the distance x is given by:
Electrochemistry, Nanomaterials, and Nanostructures 119
  
1 x
cO (x,t) = cO 1 erfc , (64)
1 + 2 DOt
 
x
cR (x,t) = cO erfc , (65)
1 + 2 DRt

where the function erfc(z) is defined by:


z  
2
erfc(z) = 1 exp y2 dy. (66)
1/2 0

Therefore, the concentrations at the electrode surface are as follows:


 
1
cO (0,t) = cO 1 and cR (x,t) = cO .
1 + 1 +

The solution of nonlinear evolution equations in the time domain is known ana-
lytically only in very simple cases such as reversible redox processes limited by
diffusion. For electrochemical nonlinear systems, the treatment of nonsteady-state
techniques generally requires calculations that are at least partially numerical. In
addition, the solutions found to express the response to a perturbing signal depend
specifically on the form of the perturbation. These drawbacks are largely eliminated
if the amplitude perturbation is limited to a sufficiently low value to allow the
equations to be linearized. In this case, analyses in the frequency domain are very
powerful.

4.2.3 Electrochemical Impedance

Here, the redox process is supposed to be irreversible, i.e., the process is not con-
sidered fast (k1 and k2 are not infinite, as in the previous examples). The faradaic
current associated to this process is then:

IF (t) = nFA(k2 cR (0,t) k1 cO (0,t)). (67)

When a small amplitude perturbation E exp ( j t) is applied to the interface around


the polarization potential, the corresponding current response, IF exp( j t), is ob-
tained by differentiating the equations describing the value of the faradaic current
and mass transport. By eliminating the terms in exp( j t) on the two sides of the
following relationship, one has the following when there is a transport limitation by
diffusion:
IF (1 )nF nF
= k1 cO (0)E k1 cO (0) + k2 cR (0)E + k2 cR (0),
nFA RT RT
2 ci (x)
j ci (x) = Di , (68)
x2
120 P.R. Bueno and C. Gabrielli

where ci represents a small variation of ci and i represents the indices O and R.


ci (x) is the concentration of the species, at a distance x from the electrode, for a
steady state polarization, such as:

ci (x,t) = ci (x) + ci (x)exp( j t). (69)

These equations lead to two important concepts: the charge transfer resistance and
the Warburg impedance. It must be kept in mind that, even if simply for the sake of
derivation convenience, the quantities exp( j t) are only taken into account implic-
itly, the preexponential terms IF and Ci (x) also depend on the pulsation .
Charge transfer resistance Rt . The charge transfer resistance is defined by:
 
1 IF
= . (70)
Rt E ci

The linearized expression of the faradaic current leads to:

1 n2 F 2
=A [(1 )k1 cO + k2 cR ] , (71)
Rt RT

where ci = ci (0,t ).
Warburg impedance. The resulting perturbation of the concentration ci (0,t) is de-
duced from the general solution of the equation:

2 ci (x)
j ci (x) = Di
x2
obtained after the terms in exp( j t) have been cancelled:
     
j j
ci (x) = Mi exp x + Ni exp x . (72)
Di Di

The integration constants Mi and Ni result from the boundary conditions and depend
on the hypothesis of the diffusion layer thickness.
Diffusion layer of infinite thickness (e.g., motionless solution). In this case,
Mi = 0; otherwise ci when x , and one has:
  
j
ci (x) = Ni exp x . (73)
Di

By putting this value in the boundary conditions at the electrode surface, one has:
 
j j
IF (t) = NO nFADO exp( j t) = NR nFADR exp( j t). (74)
DO DR
After elimination of the integration constant Ni between the last two equations:
Electrochemistry, Nanomaterials, and Nanostructures 121

cO (x = 0) 1
= , (75)
IF nFA j DO

cR (x = 0) 1
= . (76)
IF nFA j DR

Thus, the change of the faradaic current with respect to the potential is obtained:
 
1 k1 k2 I
IF = E + F (77)
Rt DO DR j

and, therefore, the impedance is equal to:


 
E
ZF ( ) = = Rt 1 + , (78)
IF j
   
where = kd / DOx + ki / DRed .
In the expression of the faradaic impedance, the term Rt / j is usually called
the Warburg impedance. The limiting value of the impedance when the frequency
tends to infinity is equal to Rt .
By taking into account the double-layer capacity, Cd , and the electrolyte resis-
tance, Re , one obtains the Randles equivalent circuit [150] (Fig. 10), where the
faradaic impedance ZF is represented by the transfer resistance Rt in series with
the Warburg impedance W . It can be shown that the high-frequency part of the
impedance diagram plotted in the complex plane (Nyquist plane) is a semicir-
cle representing Rt in parallel with Cd and the low-frequency part is a Warburg
impedance.
As Fig. 10 indicates, the extrapolation of the 45 straight line, which represents
the Warburg impedance, intercepts the real axis at:

R0 = Re + Rt R2t 2Cd . (79)

Various shapes can be obtained for the impedance diagrams, depending on the
relative values of the parameters describing the charge transfer and species diffu-
sion. Hence, obtaining the kinetic quantities from the simple extrapolation of the
45 straight line can be difficult, except if
1, since, in that case, the charge
transfer and diffusion phenomena are well separated.
Diffusion layer of finite thickness (diffusion + convection). We now use the Nernst
hypothesis, which assumes that the concentration of the reacting species that diffuse
changes linearly in a layer of thickness N and is constant thereafter.
x
ci (x,t) = ci (0,t) + [ci ci (0,t)] for x < N ,
N
ci (x,t) = ci for x N ,
122 P.R. Bueno and C. Gabrielli

a Cd

Re

W
Rc t

Re Re + Rct

Rct 2 2 Cd

Fig. 10 Electrochemical impedance for a diffusion layer of infinite thickness. a Randles equivalent
circuit; b Scheme of the impedance in the complex plane

where N is the thickness of the Nersnt diffusion layer. Thus, the general solution of
the diffusion equation leads to the following equation for the boundary condition at
x = N :
     
j j
ci (x = N ) = Mi exp N + Ni exp N = 0 for x N
Di Di
  
j
so Mi = Ni exp 2N
N
     
j j
and ci (x) = 2Ni exp N sinh (x N ) for x < N
Di Di
Electrochemistry, Nanomaterials, and Nanostructures 123

.02

.03

Im(Z) .05

.07
.07 .05
.14 .11 .09
.18 .03
.24 .02
.40
1 .01
4
20 .002

Rt kf N
Rt Re (Z)
D0

Fig. 11 Faradaic impedance ZF ( ) plotted for a finite thickness of the diffusion layer (Nernst
hypothesis) (frequency in Hz). For comparison, the impedance is also plotted (dotted lines) for an
infinite thickness of the diffusion layer


     
j j j
and IF = 2nFANi Di exp N cosh N , then
Di Di Di

j
cR (0) 1 1 ki tanh N DR
= W ( ) = (80)
IF nFA N nFA j DR
and the faradaic impedance is equal to:
 
tanh N DjO tanh N DjR
ZF ( ) = Rt 1 + k1 + k2 . (81)
j DO j DR

This impedance is represented in Fig. 11. It is noticeable that:


(a) When , the Warburg impedance is found in the high-frequency range, as
follows: 
j
tanh N D
tanh(y) 1 when y , therefore: j D i j1 D , and
i i
(b) When N , the Warburg impedance is obviously found again.
(c) If DO = DR , then:

tanh N DjO
ZF ( ) = Rt 1 + (k1 + k2 ) . (82)
j DO

Therefore, if the double-layer capacity is considered, as for the diffusion layer


of infinite thickness, there is a coalescence of the 45 straight line and the high-
124 P.R. Bueno and C. Gabrielli

frequency circle, and the straight line crosses the real axis at:

R0 = Re + Rt R2t 2Cd . (83)

In Fig. 11, the oscillation above the Warburg straight line (at frequencies from 0.4
to 0.14 Hz)) is related to the Nernst hypothesis. A numerical calculation taking into
account the convection term in the transport equation showed that the impedance
diagram is below the Warburg straight line.
Within the electrochemical framework of this classical example of a redox pro-
cess whose rate is limited by the transport by diffusion, it was shown that, even for a
reversible redox process, the derivation of the current response in the time domain is
far from simple. In contrast, the impedance approach allows the more difficult case
of an irreversible (finite reaction rate constants) redox process to be derived. Using
the same approach, we will now examine the case of a multistep reaction, which is
very difficult to investigate using techniques of potential step cyclic voltammetry.

4.2.4 Two-Step Charge Transfer with an Adsorbed Intermediate

A species A in the solution should react at the electrode to give a species, C, through
a reaction intermediate B, which is adsorbed at the electrode surface. Thus, the
reaction mechanism follows:
k
First step: Asol + e
1
Bads ,
k
Second step: Bads + e
2
Csol .
The reaction rate constants are equal to:

k1 = k10 exp (b1 E) and k2 = k20 exp (b2 E) ,

where the Tafel coefficients, b1 and b2 , are positive.


The kinetic parameter whose evolution governs the impedance is the concen-
tration of the intermediate Bads in the adsorbed phase. If is the fraction of the
electrode surface covered by this adsorbate and is the superficial concentration
for a complete coverage by Bads , by assuming a Langmuir adsorption isotherm, the
surface concentration of Bads is cB = .
The evolution equation describes the balance of and expresses the mass con-
servation:
(d /dt) = formation rate consumption rate.
The formation rate of Bads is equal to k1 cA for a first order reaction in A. By
taking into account the available area fraction, Bads becomes (1 )k1 cA .
The consumption rate of Bads is equal to k2 . By taking into account the area
fraction occupied by Bads , then Bads becomes k2 .
Electrochemistry, Nanomaterials, and Nanostructures 125

The evolution equation for is then:

d
= (1 )k1 cA k2 .
dt
The potential E, which is perturbed, does not explicitly appear in this equation but
is implicitly involved through k1 and k2 .
Steady-state solution of the evolution equation. The steady-state solution, s , of
the evolution equation at a polarization point Es , is obtained by putting d/dt 0,
which leads to:
(1 )k1 cA k2 = 0,
i.e., s = k1 cA /(k1 cA + k2 ).
Solution of the linearized nonsteady-state equation. The small amplitude changes
of the quantities , E, and I will be denoted by . Thus, the linearized evolution
equation is written as:

d
= ((1 s )b1 k1 cA s b2 k2 )E (k1 cA + k2 ) , (84)
dt
where the values of the rate constants k1 and k2 are taken at the potential Es . The so-
lution of this equation for a sinewave potential perturbation of E = |E| exp( j t)
centered on Es is as follows:

(1 s )b1 k1 cA s b2 k2
= . (85)
E j + k1 cA + k2

For frequencies much greater than the rate constants, the kinetics of the variations
of , imposed by its evolution equation, does not allow the perturbation of the po-
tential to be followed, which means that /E 0. At these high frequencies,
the coverage is said to be frozen. The coverage rate is not usually directly mea-
surable. The impedance is the observable quantity, and will be calculated from the
electrochemical current I.
The faradaic current I is the sum of the elementary currents of each step:

I = F((1 )k1 cA + k2 ).

By considering = s in this equation, the steady-state current is obtained:

IF = 2F s k2 ,

i.e., by taking into account the steady-state value of

k1 k2 cA
Is = 2F .
k1 cA + k2

The differentiation of I around the polarization point (s , Es ) gives:

I = F{((1 s )b1 k1 cA + b2 k2 s )E + (k1 cA + k2 ) }.


126 P.R. Bueno and C. Gabrielli

Then, by substituting /E and s for their values, one obtains:

1 k1 k2 cA (b1 + b2 ) k1 k2 cA (b1 b2 )
= + (k1 cA + k2 ) .
FZF k1 cA + k2 (k1 cA + k2 )( j + k1 cA + k2 )

When ZF = (k1 cA + k2 )/[F k1 k2 cA (b1 + b2 )], which is the charge transfer


resistance.
When 0 ZF (0) = Rp , which is the polarization resistance (i.e., the slope of
the currentvoltage curve).
The variations with the frequency of the faradaic impedance are determined by
the sign of the numerator of the fraction coming from /E. This sign depends
on the relative rate of the two steps of the total reaction.
Figure 12 shows the currentvoltage curve and the impedance for b1 > b2 . For
potentials lower than Ec , the impedance has an inductive loop, whereas there is a
capacitive loop for potentials greater than Ec . Ec is defined when the two steps have
equal rates. At this particular potential, where s = 0.5, the impedance is reduced
to the charge transfer resistance Rt . It is possible to ascertain that the polarization
resistance, Rp , which is the limit of ZF when 0, always remains positive and
equal to the slope of the currentvoltage curve.
For a two-step monoelectronic reaction mechanism, the Rt Is product is a constant
equal to:
2
Rt Is = .
b1 + b2

s 1
0.5

log I
E>Ec
Im[Z]

= =0
Rt Rp Re[Z]
Im[Z]

= =0
E<Ec Rt =Rp
Re[Z]
Im[Z]

Rp Rt
= Re[Z]
=0
Ec E

Fig. 12 Scheme in the Tafel plane (log I, E) of the steady state behavior and the impedance of the
two-step reaction mechanism when b1 > b2 . I: total current. I1 and I2 : partial currents of the first
and second steps. Ec is the potential where I1 = I2 and = 0.5. Upper part: variation of with
respect to E
Electrochemistry, Nanomaterials, and Nanostructures 127

5 Porous (Nanostructured) Electrode Geometry

As the use of nanotechnology in design electrodes and cells for energy conversion
and storage devices continues to grow, it is important to take into account the
porous effect and how to characterize the electrodes response to such porous
and nanostructured architectures. The porosity and nanostructures are very im-
portant in energy conversion and storage devices. Therefore, important aspects
involving porous electrode theory will be presented and discussed in this sec-
tion. The transport and reaction processes occurring in nanostructured electrode
can be understood by considering the electrochemical theory of porous electrodes.
Although the subject of porous theory has been extensively examined in the litera-
ture [92, 95, 96, 98101, 154184], it continues to be of permanent interest, partic-
ularly now that nanostructured electrodes are so widely applied and the capability
to design different porous materials and electrodes for electrochemical applications
continues to grow.
The porous electrode theory is based on the fact that electrodes operate in contact
with an electrolyte by simultaneous transport of electronic and ionic species in the
solid and liquid phase, respectively. The solid phase in contact with the conduct-
ing substrate provides a continuous path for the transport of electrons (or holes, or
polarons), but the dimension of the structural elements is extremely small, i.e., on
the nanometric scale. The electrolyte penetrates the voids in the solid phase up to
the substrate, and again the dimension of the liquid channel is very small. There-
fore, the system is characterized by the existence of two closely mixed phases in the
electroactive nanostructured layer, with narrow channels for transport, displaying a
large degree of disorder. The transport of charge carriers in both phases is believed
to be influenced by complex mechanisms. Note that, to apply the model, one might
deal with the ability of the electrolyte to penetrate the pores; thus, contact with the
very high surface area of the semiconductor is critical.
The porous electrode theory was developed by several authors for dc conditions
[185188], but the theory is usually applied in the ac regime [92,100,101,189199],
where mainly small signal frequency-resolved techniques are used, the best exam-
ple of which are ac theory and impedance spectra representation, introduced in the
previous section. The porous theory was first described by de Levi [92], who as-
sumed that the interfacial impedance is independent of the distance within the pores
to obtain an analytical solution. Because the dc potential decreases as a function of
depth, this corresponds to the assumption that the faradaic impedance is indepen-
dent of potential or that the porous model may only be applied in the absence of dc
current. In such a context, the effect of the transport and reaction phenomena and
the capacitance effects on the pores of nanostructured electrodes are equally impor-
tant, i.e., the effects associated with the capacitance of the ionic double layer at the
electrode/electrolyte-solution interface. For instance, with regard to energy storage
devices, the desirable specifications for energy density and power density, etc., are
related to capacitance effects. It is a known fact that energy density decreases as the
power density increases. This is true for EDLC or supercapacitors as well as for sec-
ondary batteries and fuel cells, particularly due to the distributed nature of the pores
128 P.R. Bueno and C. Gabrielli

in nanostructured electrodes. The usable energy stored in EDLCs or Li-ion batter-


ies diminishes as it is extracted at higher discharge rates. Therefore, it is important
to understand not only the porous effect but also the distribution characteristics in-
volved and to grasp their rationale to maximize the energy density at desired power
densities in such devices.
In addition, it has already been proven than nanostructured lithium-ion battery
electrodes based on nanofibers or nanotubes of the electrode material protruding
from an underlying current-collector surface like the bristles of a brush, for exam-
ple, or nanoparticles (see Fig. 4a) are able to dramatically improve the rate capa-
bilities because, in nanostructured electrodes, the distance that Li+ must traverse to
diffuse through the solid state (the current- and power-limiting step in Li-ion bat-
tery electrodes) is significantly smaller [86, 91, 113, 123125, 200]. Two kinds of
geometry are commonly employed to improve rate capabilities and achieve a faster
solid-state diffusion, as depicted in Fig. 4. The first geometry is based on connected
spherical-like nanoparticles and the second on nanofibers that protrude, brushlike,
from the current collector [113].
With regard to energy conversion devices, as discussed earlier herein, porous
nanocrystalline materials for electrodes possess extraordinary physical and chemical
properties thanks to their ultrafine structure (i.e., grain size < 50 nm), resulting in
very important surface effects that render them appropriate for use as electrodes in
DSSC devices. The role of the pores is to increase the screening of the electrons
in the electrode, via adsorbed ionic solution species, increasing the rate of charge
transfer between the oxide and each of the dye molecules adsorbed on the particle
surface. The nanostructured electrode also provides sufficient monolayer adsorption
of dye molecules to increase the efficiency of light adsorption [201, 202]. Although
this picture of the operation of such films is generally accepted, a consensus has
not yet been reached regarding the description of the transport and recombination
mechanisms of electrons inside the porous matrix, nor is the origin of photopotential
in DSSC completely understood. Impedance spectroscopy is one of the techniques
exploited to increase knowledge in processes that occurs in DSSC devices [22].

5.1 Transmission Line Description of Porous Electrodes

To help elucidate some of the aspects involved in the boundary conditions of


nanostructured electrodes, a proper grasp of the theory of ac porous impedance
is important to extract relevant information on the kinetics of specially designed
nanostructured electrodes. For a general view of this subject, the features expected
in a homogeneous electrode are compared, later, with those of a porous electrode
in which two phases, liquid and solid, become mixed inside the electrode region. A
homogeneous electrode with a macroscopically flat surface is shown schematically
in Fig. 13. The impedance porous model is adequately represented by the transmis-
sion line approach, according to Fig. 14. In general, the transmission line method
is strictly one-dimensional; the equipotential surfaces are planes, and such models
Electrochemistry, Nanomaterials, and Nanostructures 129

flat
electrode

Zc Zi Ze
electrolyte
conducting
substrate

Fig. 13 Schematic of a cross section of a macroscopically flat surface electrode, indicating the
impedance elements. Three different impedances are represented: Zc is the impedance of the
contact between the current collector (conduction substrate) and the electrode, Zi is interfacial
impedance of the electrodeelectrolyte interface, and Ze is the impedance corresponding to the
properties of the electrolyte, which is generally given by a pure resistance element such as the
one indicated in Fig. 10 (Re ) and (14) (Rs ) or as Rcell in (15), representing the ohmic drop in the
electrolyte

electroactive
nanostructured
electrode
1 1 1 1

Rs
conducting
2 2 2
substrate electrolyte

x
0 L
0 L

Fig. 14 Scheme of a porous electrode with the equivalent general circuit model according to the
theory explained earlier. Note that in this picture Rs is the electrolyte resistance, which in the
previous section was Rcell

can hardly represent special types of pores. The different models based on the trans-
mission line approach differ from each other by the choice of these elements and
partly by the possible addition of further discrete elements of the transmission line.
Therefore, in this kind of approach of porous electrode representation, basic electri-
cal properties such as capacitance, resistance, and inductance are defined indepen-
dently of frequency, for idealized cases. Also, their impedances have well-defined
frequency behaviors and the majority of methods usually presuppose that all the
130 P.R. Bueno and C. Gabrielli

elements throughout the layer are the same. As can be seen, the equivalent circuit
modeling of the cell impedance involves the connection of a series of elements de-
scribing the division of the applied small-signal ac voltage in three parts: one at the
bulk of the layer, another at the solid/liquid interface, and the third, which includes
the potential drop at the bulk electrolyte and metal contacts (see Fig. 14).
It is well known from studies of the interfacial impedance Zi of planar semi-
conductor electrodes that several effects occur at the semiconductor/liquid inter-
face, including the capacitance and resistance of the space-charge region, the effect
of surface states, the capacity and charge transfer resistance across the Helmholtz
layer, and the diffusion of reacting species (details were given previously in Sect. 4).
Therefore, the form of Zi may become quite complex, but nonetheless, the vari-
ous processes involved in Zi are localized in the sense that the potential difference
driving these processes obeys two conditions: it resides essentially at the semicon-
ductor/electrolyte interface, and it is essentially independent on the position on the
surface. Or, to put it another way, the flux of carriers is always normal to the plane
of the surface, and the current density is the same at any point in the surface. Con-
sequently, Zi can be described by the series and/or parallel combinations of resistive
and capacitive components. Figure 14 schematically illustrates the cases in which
the two phases are closely mixed in space because the layer is porous on a small
scale. Both media are considered as effectively homogeneous and continuously con-
nected phases. A standard equivalent circuit model also is represented in Fig. 14,
describing the essential features of electrical transport along each phase and the
exchange of charge through the inner surface. This model assumes that the predom-
inant contribution to the current is electrical field-driven rather than diffusive, but
different situations arise according to the systems to be considered. Now the porous
structure gives rise to the spread of electrical current in various directions. First,
electrical charge can flow along each medium; the resulting ac currents are termed
here i1 and i2 (the subscripts 1 and 2 denote the liquid and solid phase, respectively)
and they follow the x direction in the scheme of Fig. 14, i.e., both i1 and i2 are
parallel to the inner surface shown in the figure. Moreover, current can flow in the
normal direction to the inner surface due to electrochemical reactions and/or ca-
pacitive charging. Therefore, at a given location, i1 may decrease (increase) with a
corresponding increase (decrease) of i2 , the constraint being obeyed that iT = i1 + i2
(the total current flowing through the external circuit) is independent of position.
In agreement with this description of electrical current distribution, the equivalent
circuit branches at each point in each medium into an element that continues in the
same medium, 1 or 2 , and another impedance element that crosses the interface.
The impedance elements 1 and 2 describe a local ohmic drop at each point of the
transport channels, depending on media conductivity and more generally on trans-
port properties, whereas the element describes an exchange of electrical charge at
the interface, owing to faradaic currents and polarization at the pore surface. (Ob-
viously, the interfacial impedance itself might consist of a complex equivalent
circuit, as mentioned earlier with regard to Zi ). The branching in the equivalent cir-
cuit model is intended to occur continuously.
Electrochemistry, Nanomaterials, and Nanostructures 131

The classical model for porous or mixed-phase electrodes is therefore formulated


in terms of equations that describe the local variation of electric current and potential
in each phase in the layer of thickness L by the following equations:

1 1
i1 = , (86)
1 x
1 2
i2 = , (87)
2 x
i1 1
= (1 2 ), (88)
x
i2 1
= (1 2 ). (89)
x

The quantities or elements 1 and 2 are impedances per unit length ( cm1 ) cor-
responding to the whole electrode area A, and is an impedance length ( cm1 ).
The overall impedance is isomorphous to that of a transmission line. Regarding the
electrical potential distribution, the simple assumption is made that an ac potential
can be defined in each phase 1 and 2 which is, at each frequency, a unique func-
tion of position x; no radial distribution of potential into the pores or solid particles
is considered. It follows that the ac potential difference between the two phases at a
point x, i.e., the overvoltage in interfacial reactions, is 2 1 .
In addition to the differential equations, the boundary conditions must also be
taken into account. For this case, the ionic current is usually assumed to vanish at
the end of the liquid channel, whereas the electronic current vanishes at the outer
edge of the electrode, so that i1 (L) = 0 and i2 (0) = 0. For such boundary conditions,
the generalized solution of the model leads to the following impedance function
[92, 95, 97, 160, 190]:
 
1 2 2 2 + 22
Z= L+ + 1 coth(L/ ), (90)
1 + 2 sinh(L/ ) 1 + 2

in which  1/2

= , (91)
1 + 2
and L represents the thickness of layered electroactive material over the current
collector. Or, in other words, L stands for the porous thickness length.

5.2 Macrohomogeneous Concept (Two-Phase Model)

It is important to emphasize that there are different types of assumptions regard-


ing geometry and microstructure that lead to (5). One arrives at this result by
considering the original perfect cylinder geometry described in de Levis original
132 P.R. Bueno and C. Gabrielli

proposal of porous electrode geometry, which is normally used provided that pore
is long compared with its diameter, or from an effective macrohomogeneous de-
scription of two closely mixed phases, as described by Paasch et al., who consid-
ered an effective macrohomogeneous mixture of two phases in the electrode region
[101]. Keiser et al. also extended the transmission line model to a noncylindrical
pore [14].
The models that consider this approach are largely based on the assumption of
effectively homogeneous local relaxation processes related to transport in each of
the phases and electrical charge exchange between them. Thus, the complex prob-
lem of an uneven distribution of electrical current and potential inside the elec-
trode can be described analytically, and impedances can be calculated. Furthermore
the models may be conveniently pictured as a double-channel transmission line
(Fig. 3.5). In several papers, the theory of the impedance of porous electrodes has
been extended to cover those cases in which a complex frequency response arises in
the transport processes [100] or at the inner surface [194, 203].
Therefore, if the electrodes nanostructure composed of two mixed phases is of a
macrohomogeneous nature, (5) can be used even if the pores have a noncylindrical
geometry. In this case, one can consider a solid phase having the form of connected
channels of spherical-like particles or nanofibers that protrude from the current col-
lector substrate and a liquid phase consisting of a penetrating electrolyte that reaches
the current collector substrate. In specific cases in which a solid electrolyte is used,
this macrohomogeneous medium can be seen as a mixture of two solid phases with-
out compromising the result of the analysis. It should be noted that if the percolating
phase emerging from the current collector substrate is primarily a mixed ionic and
electronic conductor, the other phase (such as a liquid) that emerges from the elec-
trolyte (which can be seen as an ion collector penetrating up to the current collector
substrate) must be a purely ionic conductor. Furthermore, because of the previous
picture, the transmission line approach is easily used to envision a physical model
based on frequency-dependent phenomena.
A macrohomogeneous electrode can be established in different dimensional
structures and the resulting models, which can present analytical or numerical
solutions, could relate the global performance of the cathodic or anodic layer to
unmeasurable local distributions of reactants, electrode potential, and reaction rates.
These unmeasurable local distributions define a penetration depth of the active zone
and suggest an optimum range of current density and electroactive layer thick-
ness with minimal performance losses and highest electroactive effectiveness. In
addition, the macrohomogeneous theory can be extended to include concepts of
percolation theory.
As the macrohomogeneous electrode theory has proven its worth in electrode
diagnostics and design, so the finer details of electroactive layer structure and elec-
trocatalytic mechanisms are moving to the fore. A useful concept is to consider
agglomerates as structural units of the electroactive layer. Ideal locations of elec-
troactive particles are at the true two- or even three-phase boundary. This approach
is capable of and vital for showing that micropores inside agglomerates are filled
with liquid water to keep the particles active. Even for well defined and extensively
Electrochemistry, Nanomaterials, and Nanostructures 133

studied electrode geometry, the essentials of the kinetic mechanisms are not totally
settled. Each of the key steps (adsorption, surface mobility, charge transfer, and des-
orption) still constitutes a huge scientific problem involving the application of the
macrohomogeneous concept.
Therefore, the macrohomogeneous concept can also be adequately extended to
the whole cell. For instance, a framework for macrohomogeneous modeling of
porous SOFC electrodes is possible by taking into account multicomponent dif-
fusion, multiple electrochemical and chemical reactions, and electronic and ionic
conduction. The concept applies to both porous anodes and cathodes. The derivation
of the model is illustrated by considering different chemical and electrochemical re-
action schemes. The framework is general enough so that additional chemical and
electrochemical reactions can be accounted for.
Moreover, recent studies have revealed drastic differences in the kinetics of
nanoparticle surfaces. Generally, small particle sizes and highly dispersed electroac-
tive sites lead to high specific activities (per total mass of electroactive sites such as
Pt in the catalytic layer of fuel cells). Considering fuel cell applications, it is im-
portant to know the real effect of nanoparticles on the kinetics because extremely
small particle sizes (below 3 nm) affect the electronic structure of the system and
render the catalyst surface rather heterogeneous. This begs several questions: What
is the optimum nanoparticle size? What are the best properties of nanoparticle ar-
rays? Which substrate is the best? Ultimately, the following questions should be
addressed: What is the benefit of nanoparticles? Is it an intrinsic size effect or an
effect of surface heterogeneity? What is the role of the substrate? All of these ques-
tions can be addressed in the future by the use of the macrohomogeneous concept
of electrodes and cells.

5.3 Transport in the Solid and Electrolyte Phases

1 , 2 and elements must be specified in the context of (90) to be used as


impedance measurements in a defined kinetic model. Independently of the specific
systems, the general basic models can be obtained considering some reasonable and
generally valid assumptions about the basic elements in the general transmission
line, featuring the transport characteristics in each of the phases and their interfa-
cial behavior. For instance, the solid phase impedance 1 can be simply treated as
possessing an ohmic behavior (Ohms law). Therefore, the solid channel (channel
1) consists of distributed resistances such as:
 
1
1 = r1 = , (92)
Ae

where A is the geometric surface of the phase and e is its electronic conductivity.
Assuming that the NernstEinstein relation is obeyed, the conductivity is related to
the diffusion coefficient, De , by the equation:
134 P.R. Bueno and C. Gabrielli

q2 n
e = De , (93)
kB T
where n stands for the dc concentration of electronic charge carriers, q is the ele-
mentary charge, kB is Boltzmanns constant, and T is the temperature. In this simple
model for electronic conductivity in the solid phase channel of the porous electrode,
the conductivity is merely a function of the local concentration of the carriers and
the diffusion coefficient of the material. Therefore, the total resistance can be cal-
culated easily based on the distributed resistance by R1 = r1 L. This means that the
dc behavior of the charge transport processes is independent of frequency, which is
useful mainly for very high conductive solid phases. The impedance of the liquid
phase or electrolyte can similarly be described by a resistive element and, therefore,
the impedance-related element takes the form of 2 = r2 = 1/Al , in which A is
now the geometric area of the liquid or electrolyte and l is the ionic conductivity.
Furthermore, similar to (8), we obtain l = (q2 c/k B T )Dl , where c is the concentra-
tion of ionic charge carriers and Dl is the diffusion coefficient of the ionic species in
the liquid phase or electrolyte. In the same way, the total resistance in the electrolyte
contained inside the pores (electrolyte channel) can be given by R2 = r2 L.

5.4 Polarization and Charge Transfer at the Porous Interface

The interfacial impedance element, , offers many possibilities. This is the region
of transition between the solid phase (electroactive material or electrode) and the
electrolyte. Therefore, in this region, charge transfer and double-layer effects are
evidenced. Potential differences may sustain charge storage and charge transfer ki-
netics in this region. For an ideally polarizable interface, the differential relationship
between the charge at the boundary and the electrical potential is the interfacial
capacitance. An ideally polarizable distributed interface can be described by a ca-
pacitance, ci , the distributed interfacial capacitance, whose interfacial impedance is
described by:
1
= . (94)
jci
The total capacitance in the walls of the pores is given by Ci = ci L. This capacitance
is attributed to double-layer effects, so it is usually a function of the potential. It can
also be used to describe the space-charge polarization at the semiconductorliquid
junction if the spatial distribution of electrical charge as a function of potential is
known. An ideally polarizable interface with charge transfer can be described by
considering the charge transfer as a resistance, rct , which goes in parallel to the
capacitance so that the impedance element yields an impedance such as:
rct
= , (95)
1 + j /i
Electrochemistry, Nanomaterials, and Nanostructures 135

with the characteristic frequency of charge transfer defined as = 1/RctCi = 1/rct ci ,


with Rct = rct /L being the total wall charge transfer resistance. It is not easy to
interpret rct , but among a variety of situations, it could be the concentration of
reactant species in the electrolyte, characteristic reaction rates, etc. In any case,
the reaction or rct resistance originates the faradaic currents at the surface of the
electrode.

5.5 Distributed Features and Dispersion

Ideally, the polarizable interface is the exception rather than the rule. In many types
of interfaces the capacitance of the interface has been found to be a function of
the frequency, which is known as capacitive dispersion. In other words, the elec-
trical properties of real circuit elements only approach the ideal within a limited
frequency range. For instance, EDLCs show frequency-dependent capacitance even
though a capacitance should be independent of frequency. This abnormal frequency
dependency is called a distributed characteristic or frequency dispersion of elec-
trical properties. A circuit element with distributed characteristics (distributed el-
ement) cannot be exactly expressed as a combination of a finite number of ideal
circuit elements except in certain limited cases. Distributed characteristics result
from two origins [204]. First, they appear nonlocally when a dimension of a sys-
tem under study (e.g., electrode thickness or pore length) is longer than a charac-
teristic length (e.g., diffusion length or penetration depth), which is a function of
frequency. This type of distributed characteristic exists even when all system prop-
erties are homogeneous and space-invariant. This category includes diffusion in a
diffusion-limited system [189, 190], double-layer charging of a porous electrode
[92], and sluggish processes such as adsorption of anions, surface reconstruction,
and transformation in the layer. Secondly, a distributed characteristic is attributed to
various heterogeneities: geometric heterogeneity such as roughness or distribution
of pore size [205] and crystallographic heterogeneity such as anisotropic surface
metal structure and a surface disorder of polycrystalline platinum or gold [192].
Porous material with deep pores is an extreme example of the influence of geometry
on frequency dispersion, unlike rough surfaces (with shallow pores) [192]. This is
especially true when other heterogeneities are suppressed by careful experimental
conditions.
The frequency dispersion of porous electrodes can be described based on the
finding that a transmission line equivalent circuit can simulate the frequency re-
sponse in a pore. The assumptions of de Levis model (transmission line model)
include cylindrical pore shape, equal radius and length for all pores, electrolyte con-
ductivity, and interfacial impedance, which are not the function of the location in a
pore, and no curvature of the equipotential surface in a pore is considered to exist.
The latter assumption is not applicable to a rough surface with shallow pores. It has
been shown that the impedance of a porous electrode in the absence of faradaic re-
actions follows the linear line with the phase angle of 45 at high frequency and then
136 P.R. Bueno and C. Gabrielli

a vertical line at low frequency in the Nyquist plot [190,206]. Several particularized
examples of blocking interface or nonblocking dispersive boundary conditions are
studied in detail in different papers [190, 207]. These models all state that, for ide-
alized interfaces, the phase angle approaches 90 at low frequencies. In the Nyquist
plot, a vertical line is shown at low frequencies. However, many experimental data
of impedance for porous materials show that the phase angle is not 90 even at very
low frequencies, which is shown as an inclined line in the Nyquist plot.
In this case, the dispersive capacitance can be described by another interfacial
element capable of dealing with such low-frequency dispersion. A blocking capaci-
tive interface response that takes into account a frequency dependency can generally
be modeled by an interfacial impedance element such as:
1
= ( j )i , (96)
qi
which is known as constant phase element (CPE), with a prefactor for the whole
interface given by Qi = qi L.
In contrast with the voltage dependence of the interfacial capacitance, it is ex-
tremely difficult to justify the observed frequency dependence on a theoretical basis.
For instance, with regard to dispersive behavior, the porous carbon electrodes widely
used in EDLCs have two types of distributed characteristics mentioned previously;
one from pore lengths longer than the penetration depth of the ac signal and the other
from the pore size distribution. Therefore, the microstructure (represented as pore
size distribution) of the porous electrode should be optimized to obtain the desirable
energy and power density. With the introduction of a CPE (96) into the porous model
represented by (5), it is possible to simulate the nonvertical behavior of impedance
at low frequencies [93,94,189,190,208]. Various origins of CPE behavior have been
discussed, depending on different systems [189, 195199, 207, 209212].
Some authors believe that the inclined line of impedance at low frequencies
comes from the pore size distribution of porous materials [171, 182], and a few
attempts have been made to consider the effect of pore size distributions (PSD) on
the impedance of a porous electrode [171, 182], although the PSD must contribute
considerably to the distributed characteristics [171, 182]. The impedance curve in
the Nyquist plot is observed to change with the shape of a pore in the intermedi-
ate frequency region, despite its similarity to a cylindrical pore at extremely low
or high frequencies. Some authors have reported that the real part of the reduced
impedance (the ratio of impedance of a pore to electrolyte resistance in a pore) ap-
proached one-third at low frequency, irrespective of the shape of a pore [171, 182].
The PSD effect is difficult to take into account, particularly because of the time-
consuming calculations required by this method, while a parametric study is dif-
ficult because of too many parameters (sizes of different pores), but some ana-
lytical solutions are being used to represent the pore size distribution of a porous
electrode [171, 182].
In other words, the pore size distribution considers that the elements of the
impedance of the layer are not constant along the length. Thus, both porosity and re-
sistivities can vary in the course of the preparation. In other experimental situations,
Electrochemistry, Nanomaterials, and Nanostructures 137

during the preparation of a nanostructured porous layer, the dependence of the po-
tential sweep (electrochemically prepared nanostructured porous layer) causes vari-
ation in the degree of oxidation, and this degree of oxidation can vary throughout
the layer, or variations in the composition may lead to gradients of resistivities, etc.
Such problems have been considered in several cases, e.g., in [189, 190, 213] by
using a CPE description.
Another way of dealing with such problems is by using transfer matrix meth-
ods [167, 193], which allow for their solution. Eloot et al. [214, 215] developed the
matrix method to calculate the impedance of a noncylindrical pore by extending
Keisers model [206]. According to these authors, it appears that the limit of the real
part of the reduced impedance depends on the geometry of a pore.
The matrix model was used recently as the basis for considering the elements in
the channels dependent on the frequency, and they can also vary across the layer
[167, 193]. The examples show clearly that, independently of the parameters of the
system, inhomogeneity can lead to qualitative modifications of the impedance re-
sponse, the most important of which is the influence on the low-frequency pseudo-
capacitive behavior of the impedance, which is transformed into a CPE-like form
(which, as discussed previously here, is usually introduced only phenomenologi-
cally by replacing j( /c ) to j( /c ) with < 1). Some of the dependencies
demonstrated by the authors resemble those already obtained with simpler equiv-
alent circuits (but including the phenomenological CPE). Therefore, as discussed
earlier and as is well known for simpler models, the use of transfer matrix methods
and porous theory for data fitting and microscopic characterization requires further
experimental information (e.g., on porosity, changing composition) as well as the
impedance data itself. Therefore, experimental support is still needed to validate the
model and show what the influence of porous distribution is.

5.6 Charge Transport in Nanostructured Electrodes

The electrical response of nanostructured electrodes has been extensively studied, as


discussed earlier here. We know, from solid-state semiconductor physics, that dis-
ordered structures in polycrystalline film electrodes, for instance, normally perform
poorly compared with electrodes made of highly ordered materials such as single
crystals, a fact that raises several questions. How can electrons be collected and
transported efficiently through a poor and disordered structure such as a nanostruc-
tured film electrode? What are the mechanisms of charge transport? Although exten-
sive work has been done, there are still questions to be answered and the mechanism
of charge transport in mesoporous electrodes or disordered systems is currently un-
der intense debate in the literature [201, 202, 216220]. Nanostructured compound
electrolytes or ionically conductive electrodes for lithium ion batteries may also
present nonpattern charge transport [210, 221, 222].
Furthermore, with regard to the ac electrical response, it is a well-established fact
that the conductivity in many solid materials displays a universal pattern consisting
138 P.R. Bueno and C. Gabrielli

of a power-law domain at high frequencies and a constant conductivity in the low-


frequency wing. The possible consequences of such behavior on the impedance re-
sponse of porous electrodes have been explored [223, 224]. Therefore, a description
of the anomalous transport effect is readily incorporated into the standard double-
channel transmission line model for porous electrodes, on the basis of a macroscopic
phenomenological theory for transport in disordered solids [223,224]. The influence
of the power-law behavior in the solid network gives rise to such familiar features
as a curvature in the Warburg part of the spectra, or to an arc at high frequency, de-
pending on the relative magnitudes of the conductivities in both solid and liquid or
electrolyte phases. It was suggested that a similar description could be attempted for
the ionic transport through fine pores, so the transport in solid and electrolyte phase
exemplified previously by (92) and (93) must be reviewed to consider the disorder
effects [24, 83, 222, 225251].
Another aspect that should be taken into account in the interpretation of experi-
ments is the fact that the diffusion mechanism need not be unique at all the scales of
frequencytime. It has been observed that the normal behavior in many disordered
systems consists in an anomalous diffusion mechanism at high frequencies and short
times, which changes to ordinary diffusion as the frequency decreases or the time
scale increases [221]. So the resulting diffusion would be either anomalous or ordi-
nary, depending on whether the distance traveled by the random walker during the
observation time was shorter or longer than the cutoff scale of the structure [252].
These situations could be treated by introducing a characteristic crossover time or
frequency that separates the diffusion regimes. Such an analysis was presented for
impedance of porous electrodes [223].
Example of the uses of porous model approach in energy conversion porous elec-
trode is given by the use of general transmission line model of DSSC. From this
model it is possible to calculate, for instance, charge-transfer resistance of the charge
recombination process between electrons in the mesoporous TiO2 film, the chemical
capacitance of the electroactive film, transport resistance of the electrons in the TiO2
film, and Nernst diffusion in the electrolyte among other parameters. By using such
approach it is possible to understand characteristics of high efficiency of DSSC to
conclude, at this moment, that the high efficiency of the DSSC cells can be ascribed
to the high transport and low recombination rate of the electrons in TiO2 -electrolyte
interface [253].

6 Future Prospects

The development of alternative storage and conversion energy devices was shown to
be strongly correlated to and in some aspects dependent on nanotechnology. There-
fore, new components for these devices will continue to be designed in the coming
years, especially aiming for better performance and scale mass production. The fu-
ture economic development of our modern society will depend to a great extent on
such development, since the era of fossil fuel is coming to an end.
Electrochemistry, Nanomaterials, and Nanostructures 139

Conversion and storage of energy requires efficient energy capture, charge


separation and transport, and, finally, efficient utilization or storage by chemical or
electrochemical means. The high surface area of nanostructures has been proven to
enhance efficiency tremendously. In addition, to improve the performance of these
devices, the role of interfaces is crucial for highly efficient nanostructured elec-
trodes. The efficiencies of catalytic sequences for utilization and storage are also
enhanced by the high surface area. Thus, nanoscale systems are expected to advance
conversion, and any headway made in understanding the relevant charge transfer
process and the strongly related energy transfer will be central to this progress.
To produce materials and systems with improved properties we must be able
to modify the amount of interface and its properties in a controlled manner. The
engineering of nanostructured electrodes depends on a more in-depth understand-
ing of the nanoscale size effect properties of compounds and on the characteri-
zation and assembly of 2-D and/or 3-D nanostructured electrodes or cells. Any
strategy for modifying these nanostructured electrodes in a specific way should
involve the design of nanostructured electrodes with controlled electrochemical
properties.
Despite recent technological successes through the clever incorporation of the
use of nanosized materials in electrodes (nanostructured electrodes) in alternative
energy devices, we are still far away from possessing a solid scientific understand-
ing of what really goes on at the nanoscale. Many critical questions remain to be
answered. For instance, what are the characteristic dimensions over which energy
transfer or charge transfer reactions can effectively occur in devices? How does
the nanosize material surface influence this dimension? How exactly do kinetic
properties scale in small dimension? Is there more than simple surface area scaling
at work?
Because of these and other questions that are still open, the study of the kinetic
scaling behavior of nanostructured systems is quite complex and must be the focus
of further development of appropriate nanostructured electrodes.

References

1. Dresselhaus, M.S. and I.L. Thomas, Alternative energy technologies. Nature, 2001. 414:
pp. 332337
2. Goodisman, J., Electrochemistry: Theoretical Foundation. 1987, Chichester: Wiley
3. Schmickler, W., Interfacial Electrochemistry. 1995, Oxford: Oxford University Press
4. Bard, A.J. and L.R. Faulkner, Electrochemical Methods: Fundamentals and Applications.
2000, Chichester: Wiley
5. Bockris, J.O.M. and S.U.M. Khan, Surface Electrochemistry: A Molecular Level Approach.
1993, New York: Springer
6. Bockris, J.O.M. and A.K.N. Reddy, Modern Electrochemistry. Vol.1. 1973, New York: A
Plenum/Rosetta Edition
7. Cao, G., Nanostructures and Nanomaterials: Synthesis, Properties and Applications. 2004,
Singapore: Imperial College Press
8. Adams, D.M., L. Brus, C.E.D. Chidsey, S. Creager, C. Creutz, C.R. Kagan, P.V. Kamat,
M. Lieberman, S. Lindsay, R.A. Marcus, R.M. Metzger, M.E. Michel-Beyerle, J.R. Miller,
140 P.R. Bueno and C. Gabrielli

M.D. Newton, D.R. Rolison, O. Sankey, K.S. Schanze, J. Yardley, and X.Y. Zhu, Charge
transfer on the nanoscale: Current status. Journal of Physical Chemistry B, 2003. 107(28):
pp. 66686697
9. Zukalova, M., A. Zukal, L. Kavan, M.K. Nazeeruddin, P. Liska, and M. Gratzel, Orga-
nized mesoporous TiO2 films exhibiting greatly enhanced performance in dye-sensitized so-
lar cells. Nano Letters, 2005. 5(9): pp. 17891792
10. Li, N.C., C.R. Martin, and B. Scrosati, Nanomaterial-based Li-ion battery electrodes. Journal
of Power Sources, 2001. 9798: pp. 240243
11. Liu, Y., C. Compson, and M.L. Liu, Nanostructured and functionally graded cathodes for
intermediate temperature solid oxide fuel cells. Journal of Power Sources, 2004. 138(12):
pp. 194198
12. Sides, C.R., N.C. Li, C.J. Patrissi, B. Scrosati, and C.R. Martin, Nanoscale materials for
lithium-ion batteries. Mrs Bulletin, 2002. 27(8): pp. 604607
13. Singhal, A., G. Skandan, G. Amatucci, F. Badway, N. Ye, A. Manthiram, H. Ye, and J.J. Xu,
Nanostructured electrodes for next generation rechargeable electrochemical devices. Journal
of Power Sources, 2004. 129(1): pp. 3844
14. Bishop, D., Nanotechnology and the end of Moores Law? Bell Labs Technical Journal, 2005.
10(3): pp. 2328
15. Lai, L.B., D.H. Chen, and T.C. Huang, Preparation and electrocatalytic activity of Pt/Ti
nanostructured electrodes. Journal of Materials Chemistry, 2001. 11(5): pp. 14911494
16. An, K.H., K.K. Jeon, J.K. Heo, S.C. Lim, D.J. Bae, and Y.H. Lee, High-capacitance superca-
pacitor using a nanocomposite electrode of single-walled carbon nanotube and polypyrrole.
Journal of the Electrochemical Society, 2002. 149(8): pp. A1058A1062
17. Frackowiak, E. and F. Beguin, Carbon materials for the electrochemical storage of energy in
capacitors. Carbon, 2001. 39: pp. 937950
18. Fuertes, A.B., F. Pico, and J.M. Rojo, Influence of pore structure on electric double-layer
capacitance of template mesoporous carbons. Journal of Power Sources, 2004. 133(2):
pp. 329336
19. Kim, I.H., J.H. Kim, Y.H. Lee, and K.B. Kim, Synthesis and characterization of electro-
chemically prepared ruthenium oxide on carbon nanotube film substrate for supercapacitor
applications. Journal of the Electrochemical Society, 2005. 152(11): pp. A2170A2178
20. Kim, I.H., J.H. Kim, and K.B. Kim, Electrochemical characterization of electrochemically
prepared ruthenium oxide/carbon nanotube electrode for supercapacitor application. Elec-
trochemical and Solid State Letters, 2005. 8(7): pp. A369A372
21. Liu, C.G., M. Liu, M.Z. Wang, and H.M. Cheng, Research and development of carbon ma-
terials for electrochemical capacitors II The carbon electrode. New Carbon Materials,
2002. 17(2): pp. 6472
22. Wang, Q., J.E. Moser, and M. Gratzel, Electrochemical impedance spectroscopic analysis of
dye-sensitized solar cells. Journal of Physical Chemistry B, 2005. 109(31): pp. 1494514953
23. Xiao, Q.F. and X. Zhou, The study of multiwalled carbon nanotube deposited with conducting
polymer for supercapacitor. Electrochimica Acta, 2003. 48(5): pp. 575580
24. Lee, K., R. Menon, C.O. Yoon, and A.J. Heeger, Reflectance of conducting polypyrrole:
Observation of the metal insulator transition driven by disorder. Physical Review B, 1995.
52: pp. 4779
25. Gray, F.M., Polymer Electrolytes. 1997, Cambridge: Royal Society of Chemistry
26. Bujdak, J., E. Hackett, and E.P. Giannelis, Chemical Materials, 2000. 12: pp. 2168
27. Solomon, M.J., A.S. Almusallam, K.F. Seefeldt, A. Somwangthanaroj, and P. Varadan,
Macromolecules, 2001. 34: p. 7219
28. Maier, J., Prog. Solid State Chem., 1995. 23: pp. 171265
29. Rhodes, C.P., J.W. Long, M.S. Doescher, J.J. Fontanella, and D.R. Rolison, Nanoscale poly-
mer electrolytes: Ultrathin electrodeposited poly(phenylene oxide) with solid-state ionic con-
ductivity. Journal of Physical Chemistry B, 2004. 108(35): pp. 1307913087
30. Petrii, O.A. and G.A. Tsirlina, Size effects in electrochemistry. Uspekhi Khimii, 2001. 70(4):
pp. 330344
Electrochemistry, Nanomaterials, and Nanostructures 141

31. Hagfeldt, A., G. Boschloo, H. Lindstrom, E. Figgemeier, A. Holmberg, V. Aranyos, E. Mag-


nusson, and L. Malmqvist, A system approach to molecular solar cells. Coordination Chem-
istry Reviews, 2004. 248(1314): pp. 15011509
32. Li, N.C., C.J. Patrissi, G.L. Che, and C.R. Martin, Rate capabilities of nanostructured
LiMn2O4 electrodes in aqueous electrolyte. Journal of the Electrochemical Society, 2000.
147(6): pp. 20442049
33. Schultze, J.W., A. Heidelberg, C. Rosenkranz, T. Schapers, and G. Staikov, Principles of elec-
trochemical nanotechnology and their application for materials and systems. Electrochimica
Acta, 2005. 51(5): pp. 775786
34. Davies, T.J., M.E. Hyde, and R.G. Compton, Nanotrench arrays reveal insight into graphite
electrochemistry. Angewandte Chemie International Edition, 2005. 44(32): pp. 51215126
35. Waje, M., C. Wang, J. Tang, and Y.S. Yan, Nanostructured electrodes for hydrogen fuel cells.
Abstracts of Papers of the American Chemical Society, 2004. 227: pp. U1082U1082
36. Cai, C.D., J.Z. Zhou, L. Qi, Y.Y. Xi, B.B. Lan, L.L. Wu, and Z.H. Lin, Conductance of
a single conducting polyaniline nanowire. Acta Physico-Chimica Sinica, 2005. 21(4): pp.
343346
37. Park, S., Y. Xie, and M.J. Weaver, Electrocatalytic pathways on carbon-supported platinum
nanoparticles: Comparison of particle-size-dependent rates of methanol, formic acid, and
formaldehyde electrooxidation. Langmuir, 2002. 18(15): pp. 57925798
38. Vinodgopal, K., M. Haria, D. Meisel, and P. Kamat, Fullerene-based carbon nanostructures
for methanol oxidation. Nano Letters, 2004. 4(3): pp. 415418
39. Sun, N.X. and K. Lu, Physical Review B, 1997. 54: pp. 6058
40. Park, S., P.X. Yang, P. Corredor, and M.J. Weaver, Transition metal-coated nanoparticle
films: Vibrational characterization with surface-enhanced Raman scattering. Journal of the
American Chemical Society, 2002. 124(11): pp. 24282429
41. Park, S., A. Wieckowski, and M.J. Weaver, Electrochemical infrared characterization of CO
domains on ruthenium-decorated platinum nanoparticles. Journal of the American Chemical
Society, 2003. 125(8): pp. 22822290
42. Park, S. and M.J. Weaver, A versatile surface modification scheme for attaching metal
nanoparticles onto gold: Characterization by electrochemical infrared spectroscopy. Jour-
nal of Physical Chemistry B, 2002. 106(34): pp. 86678670
43. Park, S., S.A. Wasileski, and M.J. Weaver, Some interpretations of surface vibrational
spectroscopy pertinent to fuel-cell electrocatalysis. Electrochimica Acta, 2002. 47(2223):
pp. 36113620
44. Park, S., Y.T. Tong, A. Wieckowski, and M.J. Weaver, Infrared spectral comparison of elec-
trochemical carbon monoxide adlayers formed by direct chemisorption and methanol disso-
ciation on carbon-supported platinum nanoparticles. Langmuir, 2002. 18(8): pp. 32333240
45. Park, S., Y. Tong, A. Wieckowski, and M.J. Weaver, Infrared reflection-absorption prop-
erties of platinum nanoparticle films on metal electrode substrates: control of anomalous
opticaleffects. Electrochemistry Communications, 2001. 3(9): pp. 509513
46. Park, S., P.K. Babu, A. Wieckowski, and M.J. Weaver, Electrochemical infrared characteri-
zation of CO domains on ruthenium decorated platinum nanoparticles. Abstracts of Papers
of the American Chemical Society, 2003. 225: pp. U619U619
47. Weaver, M.J. and S.H. Park, Vibrational and electrocatalytic characterization of Pt-group
nanoparticle films. Abstracts of Papers of the American Chemical Society, 2002. 223: pp.
U387U387
48. Weaver, M.J., Surface-enhanced Raman spectroscopy as a versatile in situ probe of
chemisorption in catalytic electrochemical and gaseous environments. Journal of Raman
Spectroscopy, 2002. 33(5): pp. 309317
49. Brinker, C.J. and C.W. Scherer, SolGel Science: The Physics and Chemistry of Sol-Gel
Processing. 1990, San Diego: Academic
50. Alivisatos, A.P., Science, 1996. 271: p. 933
51. Tamashiro, M.N., V.B. Henriques, and M.T. Lamy, Aqueous suspensions of charged spher-
ical colloids: Dependence of the surface charge on ionic strength, acidity, and colloid con-
centration. Langmuir, 2005. 21(24): pp. 1100511016
142 P.R. Bueno and C. Gabrielli

52. Manciu, M. and E. Ruckenstein, The polarization model for hydration/double layer inter-
actions: the role of the electrolyte ions. Advances in Colloid and Interface Science, 2004.
112(13): pp. 109128
53. Turkevish, J., Gold Bull., 1985. 18: p. 86.
54. Faraday, M., Philos. Trans., 1857. 147: p. 145.
55. Grabar, K.C., P.C. Smith, M.D. Musik, J.A. Davis, D.G. Walter, M.A. Jackson, A.P. Guthrie,
and M.J. Natan, Journal of American Chemical Society, 1996. 118: p. 1148
56. Hodes, G., Electrochemistry of Nanomaterials. 2001, Weinheim: Wiley
57. Agrios, A.G. and P. Pichat, State of the art and perspectives on materials and applications
of photocatalysis over TiO2 . Journal of Applied Electrochemistry, 2005. 35(7): pp. 655663
58. Wallace, J.M., B.M. Dening, K.B. Eden, R.M. Stroud, J.W. Long, and D.R. Rolison, Silver-
colloid-nucleated cytochrome c superstructures encapsulated in silica nanoarchitectures.
Langmuir, 2004. 20(21): pp. 92769281
59. Xiang, J., B. Liu, S.T. Wu, B. Ren, F.Z. Yang, B.W. Mao, Y.L. Chow, and Z.Q. Tian, A con-
trollable electrochemical fabrication of metallic electrodes with a nanometer/angstrom-sized
gap using an electric double layer as feedback. Angewandte Chemie International Edition,
2005. 44(8): pp. 12651268
60. Wasileski, S.A. and M.J. Weaver, Influence of solvent co-adsorption on the bonding and
vibrational behavior of carbon monoxide on Pt(111) electrodes. Abstracts of Papers of the
American Chemical Society, 2003. 225: p. U682
61. Schindler, W., M. Hugelmann, and P. Hugelmann, In situ scanning probe spectroscopy at
nanoscale solid/liquid interfaces. Electrochimica Acta, 2005. 50(15): pp. 30773083
62. Han, D.H. and S.M. Park, Electrochemistry of conductive polymers. 32. Nanoscopic exami-
nation of conductivities of polyaniline films. Journal of Physical Chemistry B, 2004. 108(37):
pp. 1392113927
63. Gutierrez-Tauste, D., I. Zumeta, E. Vigil, M.A. Hernandez-Fenollosa, X. Domenech, and
J.A. Ayllon, New low-temperature preparation method of the TiO2 porous photoelectrode for
dye-sensitized solar cells using UV irradiation. Journal of Photochemistry and Photobiology
Chemistry, 2005. 175(23): pp. 165171
64. Altair, Hosokawa, Rutgers work on nanostructured electrodes. American Ceramic Society
Bulletin, 2004. 83(10): pp. 33
65. Gomez, R. and P. Salvador, Photovoltage dependence on film thickness and type of illumina-
tion in nanoporous thin film electrodes according to a simple diffusion model. Solar Energy
Materials and Solar Cells, 2005. 88(4): pp. 377388
66. Gomez, R., J. Solla-Gullon, J.M. Perez, and A. Aldaz, Nanoparticles-on-electrode approach
for in situ surface-enhanced Raman spectroscopy studies with platinum-group metals: exam-
ples and prospects. Journal of Raman Spectroscopy, 2005. 36(67): pp. 613622
67. Lakard, B., J.C. Jeannot, M. Spajer, G. Herlem, M. de Labachelerie, P. Blind, and B. Fahys,
Fabrication of a miniaturized cell using microsystern technologies for electrochemical ap-
plications. Electrochimica Acta, 2005. 50(9): pp. 18631869
68. Jang, S.Y., M. Marquez, and G.A. Sotzing, Writing of conducting polymers using nanoelec-
trochemistry. Synthetic Metals, 2005. 152(13): pp. 345348
69. Sides, C.R. and C.R. Martin, Nanostructured electrodes and the low-temperature perfor-
mance of Li-ion batteries. Advanced Materials, 2005. 17(1): pp. 125128
70. Xu, Q. and M.A. Anderson, J. Am. Ceram. Soc., 1994. 77: p. 1939.
71. Gratzel, M., Photoelectrochemical cells. Nature, 2001. 414: pp. 338344
72. Uvarov, N.F. and V.V. Boldyrev, Size effects in chemistry of heterogeneous systems. Russian
Chemical Review, 2001. 70(4): pp. 265284
73. Rhodes, C.P., J.W. Long, M.S. Doescher, B.M. Dening, and D.R. Rolison, Charge in-
sertion into hybrid nanoarchitectures: mesoporous manganese oxide coated with ultrathin
poly(phenylene oxide). Journal of Non-Crystalline Solids, 2004. 350: pp. 7379
74. Kuznetsov, A.M. and J. Ulstrup, Electrochemica Acta, 2000. 45: p. 2339
75. Fawcett, W.R., J. Lipkowski and P.N. Ross, Editors. Electrocatalysis 1998, Wiley: New York.
p. 323
Electrochemistry, Nanomaterials, and Nanostructures 143

76. Aral, B.K. and D.M. Kalyon, Effects of temperature and surface roughness on time-
dependent development of wall slip in torsional flow of concentrated suspensions. Journal
of Rheology, 1994. 38: p. 957972
77. Roberts, G.P. and H.A. Barnes, New measurements of the flow-curves for Carbopol disper-
sions without slip artifacts. Rheological Acta, 2001. 40: p. 499
78. Leger, L., H. Hervert, G. Massey, and E. Durlist, Journal of Physics: Condensed Matter,
1997. 9: p. 7719
79. Arico, A.S., P. Bruce, B. Scrosati, J.M. Tarascon, and W. Van Schalkwijk, Nanostructured
materials for advanced energy conversion and storage devices. Nature Materials, 2005. 4(5):
pp. 366377
80. Nanu, M., J. Schoonman, and A. Goossens, Solar-energy conversion in TiO2/CuInS2
nanocomposites. Advanced Functional Materials, 2005. 15(1): pp. 95100
81. Sides, C.R., F. Croce, V.Y. Young, C.R. Martin, and B. Scrosati, A high-rate, nanocomposite
LiFePO4/carbon cathode. Electrochemical and Solid State Letters, 2005. 8(9): pp. A484
A487
82. Finke, A., P. Poizot, C. Guery, and J.M. Tarascon, Characterization and Li reactivity of elec-
trodeposited coppertin nanoalloys prepared under spontaneous current oscillations. Journal
of the Electrochemical Society, 2005. 152(12): pp. A2364A2368
83. Zuppiroli, L., M.N. Bussac, S. Paschem, O. Chauvet, and L. Forro, Hopping in disordered
conducting polymers. Physical Review B, 1994. 50: p. 5196
84. Xia, Y., P. Yang, Y. Sun, Y. Wu, B. Mayers, B. Gates, Y. Yin, F. Kim, and H. Yan, Advanced
Materials, 2003. 15: p. 353
85. Hart, R.W., H.S. White, B. Dunn, and D.R. Rolison, 3-D microbatteries. Electrochemistry
Communications, 2003. 5(2): pp. 120123
86. Bueno, P.R., E.R. Leite, T.R. Giraldi, L.O.S. Bulhoes, and E. Longo, Nanostructured Li ion
insertion electrodes. 2. Tin dioxide nanocrystalline layers and discussion on Nanoscale Ef-
fect. Journal of Physical Chemistry B, 2003. 107: pp. 88788883
87. Nanu, M., J. Schoonman, and A. Goossens, Nanocomposite three-dimensional solar cells
obtained by chemical spray deposition. Nano Letters, 2005. 5(9): pp. 17161719
88. Pietron, J.J., R.M. Stroud, and D.R. Rolison, Using three dimensions in catalytic mesoporous
nanoarchitectures. Nano Letters, 2002. 2(5): pp. 545549
89. Rolison, D.R., Catalytic nanoarchitectures The importance of nothing and the unimpor-
tance of periodicity. Science, 2003. 299(5613): pp. 16981701
90. Loffler, M.-S., H. Natter, R. Hempelmann, and K. Wippermann, Electrochemica Acta, 2003.
48: p. 3047
91. Li, N., L. Shi, W. Lu, X. Huang, and L. Chen, Journal of Electrochemical Society, 2001. 147:
p. A915
92. de Levie, R., On porous electrodes in electrolyte solutions. Electrochimica Acta, 1963. 8:
pp. 751780
93. Gabrielli, C., O. Haas, and H. Takenoutti, Impedance analysis of electrodes modified with a
reversible redox polymer film. Journal of Applied Electrochemistry, 1987. 17: p. 82
94. Gabrielli, C., H. Takenoutti, O. Haas, and A. Tsukada, Impedance investigation of the charge
transport in film-modified electrodes. Journal of Electroanalytical Chemistry, 1991. 302: pp.
5989
95. Gassa, L.M., J.R. Vilche, M. Ebert, K. Juttner, and W.J. Lorenz, Electrochemical impedance
spectroscopy on porous electrodes. Journal of Applied Electrochemistry, 1990. 20: pp.
677685
96. Lasia, A., Hydrogen evolution/oxidation reactions on porous electrodes. Journal of Electro-
analytical Chemistry, 1998. 454: pp. 115121
97. Macdonald, J.R., Impedance Spectroscopy. 1987, New York: Wiley
98. Ohmori, T., T. Kimura, and H. Masuda, Impedance measurements of a platinum cylindri-
cal porous electrode replicated from anodic porous alumina. Journal of the Electrochemical
Society, 1997. 144: p. 1286
99. Rangarajan, S.K., Theory of flooded porous electrodes. Journal of Electroanalytical Chem-
istry, 1969. 22: pp. 89104
144 P.R. Bueno and C. Gabrielli

100. Raistrick, I.D., Impedance studies of porous electrodes. Electrochimica Acta, 1990. 35: pp.
15791586
101. Paasch, G., K. Micka, and P. Gersdorf, Theory of the electrochemical impedance of macro-
homogeneous porous electrodes. Electrochimica Acta, 1993. 38(18): pp. 26532662
102. Schoonman, J., Nanoionics. Solid State Ionics, 2003. 157: pp. 319326
103. Guo, Y.G., J.S. Lee, and J. Maier, AgI nanoplates with mesoscopic superionic conductivity
at room temperature. Advanced Materials, 2005. 17(23): p. 28152819
104. Bhattacharyya, A.J., J. Fleig, Y.G. Guo, and J. Maier, Local conductivity effects in polymer
electrolytes. Advanced Materials, 2005. 17(21): p. 2630
105. Snaith, H.J., S.M. Zakeeruddin, L. Schmidt-Mende, C. Klein, and M. Gratzel, Ion-
coordinating sensitizer in solid-state hybrid solar cells. Angewandte Chemie International
Edition, 2005. 44(39): pp. 64136417
106. Rhodes, C.P., J.W. Long, and D.R. Rolison, Direct electrodeposition of nanoscale solid poly-
mer electrolytes via electropolymerization of sulfonated phenols. Electrochemical and Solid
State Letters, 2005. 8(11): pp. A579A584
107. Tarascon, J.-M. and M. Armand, Issues and challenges facing rechargeable lithium batteries.
Nature, 2001. 414: pp. 359367
108. Chadwick, A.V., Solid progress in ion conduction. Nature, 2000. 408: pp. 925926
109. Steele, B.C.H. and A. Heinzel, Material for fuel-cell technologies. Nature, 2001. 414: pp.
345352
110. Sata, N., K. Eeberman, K. Eberl, and J. Maier, Nature, 2000. 408: pp. 946949
111. Hoffler, H.J., R.S. Averback, H. Hahn, and H. Gleiter, Journal of Applied Physics, 1993. 74:
p. 3832
112. Schlapbach, L. and A. Zuttel, Hydrogen-storage materials for mobile applications. Nature,
2001. 414: pp. 353358
113. Bueno, P.R. and E.R. Leite, Nanostructured Li ion insertion electrodes. 1. Discussion on
fast transport and short path for ion diffusion. Journal of Physical Chemistry B, 2003. 107:
pp. 88688877
114. Grugeon, S., S. Laruelle, L. Dupont, F. Chevallier, P.L. Taberna, P. Simon, L. Gireaud, S.
Lascaud, E. Vidal, B. Yrieix, and J.M. Tarascon, Combining electrochemistry and metal-
lurgy for new electrode designs in Li-ion batteries. Chemistry of Materials, 2005. 17(20): pp.
50415047
115. Delacourt, C., L. Laffont, R. Bouchet, C. Wurm, J.B. Leriche, M. Morcrette, J.M. Taras-
con, and C. Masquelier, Toward understanding of electrical limitations (electronic, ionic)
in LiMPO4 (M=Fe, Mn) electrode materials. Journal of the Electrochemical Society, 2005.
152(5): pp. A913A921
116. Scrosati, B., Power sources for portable electronics and hybrid cars: Lithium batteries and
fuel cells. Chemical Record, 2005. 5(5): pp. 286297
117. Thackeray, M.M., C.S. Johnson, J.T. Vaughey, N. Li, and S.A. Hackney, Advances in
manganese-oxide composite electrodes for lithium-ion batteries. Journal of Materials
Chemistry, 2005. 15(23): pp. 22572267
118. Courtney, I.A. and J.R. Dahn, Journal of Power Sources, 1997. 144: p. 2045
119. Besenhard, J.O., J. Yang, and M. Winter, Journal of Power Sources, 1997. 68: p. 87
120. Idota, Y., T. Kubota, A. Matsufuji, Y. Maekawa, and T. Miyasaki, Tin-based amorphous
oxide: A high capacity lithium-ion storage material. Science, 1997. 276: p. 1395
121. Huang, H., Anode materials for lithium-ion batteries. 1999, Delft University of Technology
122. Winter, M., J.O. Besenhard, M.E. Spahr, and P. Novak, Insertion electrode materials for
rechargeable lithium batteries. Advanced Materials, 1998. 10(10): pp. 725763
123. Li, N., C.R. Martin, and B. Scrosati, Electrochem. Solid-State Letters, 2000. 3: p. 316
124. Li, N. and C.R. Martin, Journal of Electrochemical Society, 2001: p. A164
125. Martin, C.R., N. Li, and B. Scrosati, Nanomaterial-based Li-ion battery electrodes. Journal
of Power Sources, 2001. 9798: pp. 240243
126. Long, J.W., B. Dunn, D.R. Rolison, and H.S. White, Three-dimensional battery architectures.
Chemical Reviews, 2004. 104(10): pp. 44634492
Electrochemistry, Nanomaterials, and Nanostructures 145

127. Liu, Y., S.W. Zha, and M.L. Liu, Novel nanostructured electrodes for solid oxide fuel cells
fabricated by combustion chemical vapor deposition (CVD). Advanced Materials, 2004.
16(3): p. 256260
128. Gratzel, M., Solar energy conversion by dye-sensitized photovoltaic cells. Inorganic Chem-
istry, 2005. 44(20): pp. 68416851
129. Miyake, M., T. Torimoto, T. Sakata, H. Mori, and H. Yoneyama, Photoelectrochemical char-
acterization of nearly monodisperse CdS nanoparticles-immobilized gold electrodes. Lang-
muir, 1999. 15: pp. 15031507
130. Drouard, S., S.G. Hickey, and D.J. Riley, CdS nanoparticle-modified electrodes for photo-
chemical studies. Chemical Communications, 1: 1999: p. 67
131. Conway, B.E., Electrochemical Supercapacitors. 1999, New York: Kluwer/Prenum
132. Endo, M., T. Maeda, T. Takeda, Y.J. Kim, K. Koshiba, H. Hara, and M.S. Dresselhaus, Ca-
pacitance and pore-size distribution in aqueous and nonaqueous electrolytes using various
activated carbon electrodes. Journal of Electrochemical Society, 2001. 148: pp. A910A914
133. Qu, D. and H. Shi, Studies of activated carbon used in double-layer capacitors. Journal of
Power Sources, 1998. 74: pp. 99107
134. Salitra, G., A. Soffer, L. Eliad, Y. Cohen, and D. Aurbach, Carbon electrodes for double-
layer capacitors. I. Relations between ion and pore dimensions. Journal of Electrochemical
Society, 2000. 147: pp. 24862493
135. Shiraishi, S., H. Kurihara, H. Tsubota, A. Oya, Y. Soneda, and Y. Yamada, Electrochem. and
Solid State Letters, 2001. 4: pp. A5A8
136. Endo, M., Y.J. Kim, T. Takeda, T. Maeda, T. Hayashi, K. Kashiba, H. Hara, and M.S. Dres-
selhaus, Journal of Electrochemical Society, 2001. 148: pp. A1135A1140
137. Weng, T.-C. and H. Teng, Journal of Electrochemical Society, 2001. 148: pp. A368A373
138. Schmitt, C., H. Probstle, and J. Fricke, Journal of Non-Crystalline Solids, 2001. 285: pp.
277282
139. Niu, C., E.K. Sichel, R. Hoch, D. Moy, and H. Tennent, Applied Physics Letters, 1997. 70:
pp. 14801482
140. Diederich, L., E. Barborini, P. Piseri, A. Podest`a, P. Milani, A. Schneuly, and R. Gallay,
Applied Physics Letters. 75: pp. 26622664
141. Instrumental Method in Electrochemistry, Ed. Southampton Electrochemistry Group. 1985,
Chichester: Ellis Horwood Ltd
142. Epelboin, I. and M. Keddam, Journal of Electrochemical Society. 1970. 117: p. 1052
143. Gabrielli, G., Identification of Electrochemical Processes by Frequency Response Analysis.
1980, Farnborough U. K.: Solartron
144. Gabrielli, G., Use and Applications of Electrochemical Impedance Techniques. 1990, Farn-
borough U. K.: Solartron
145. Gerisher, H. and W. Mehl, Z. Elektrochem, 1955 59: p. 1049
146. Girault, H.H., Electrochimie physique et analytique. 2001, Presses Polytechniques et Univer-
sitaires Romandes: Lausanne, Suisse
147. Levich, V.D., Physicochemical Hydrodynamics. 1962, Englewood Cliffs, NJ: Prentice Hall
148. Macdonald, D.D., Transient Techniques in Electrochemistry. 1977, New York: Plenum
149. Newman, J., Electrochemical Systems. 1973, Englewood Cliffs, NJ: Prentice Hall
150. Randles, J.E.B., Transactions of Faraday Society, 1948. 44: p. 327
151. Rubinstein, I., Physical electrochemistry. 1995, New York: Marcel Dekker
152. Sluyters-Rembach, M. and J.H. Sluyters, in Electroanalytical Chemistry, A.J. Bard, Editor.
1970, Marcel Dekker: New York
153. Yeager, E., J. OM. Bockris, B.E. Conway, and S. Sarangapani, Comprehensive treatise of
Electrochemistry. 1984, New York: Plenum
154. Bisquert, J., G. Garcia Belmonte, and F. Fabregat Santiago. Coupled ion-electron transport
in illuminated TiO2 nanoporous electrodes. in 12th International Conference on Photoelec-
trochemical Conversion and Storage of Solar Energy. 1998. Berlin
155. Bisquert, J., G. Garcia-Belmonte, F. Fabregat-Santiago, and A. Compte, Anomalous trans-
port effects in the impedance of porous electrodes. Electrochemistry Communications, 1999.
1: pp. 429435
146 P.R. Bueno and C. Gabrielli

156. Candy, J.-P., P. Fouilloux, M. Keddam, and H. Takenouti, The characterization of porous
electrodes by impedance mesurements. Electrochimica Acta, 1981. 26: p. 1029
157. Fievet, P., M. Mullet, and J. Pagetti, Impedance measurements for determination of pore
texture of a carbon membrane. Journal of Membrane Science, 1998. 149: pp. 143150
158. Keddam, M., C. Rakotomavo, and H. Takenoutti, Impedance of a porous electrode with an
axial gradient of concentration. Journal of Applied Electrochemistry, 1984. 14: p. 437
159. Kramer, M. and M. Tomkiewicz, Porous electrodes. I. Numerical simulation using random
network and single-pore models. Journal of the Electrochemical Society, 1984. 131: pp.
12831288
160. Lasia, A., Impedance of porous electrodes. Journal of Electroanalytical Chemistry, 1995.
397: pp. 2733
161. Lasia, A., Porous electrodes in the presence of a concentration gradient. Journal of Electro-
analytical Chemistry, 1997. 428: pp. 155164
162. Lasia, A., Nature of two semicircles observed on the complex plane plots on porous elec-
trodes in the presence of a concentration gradient. MMM, Journal of Electroanalytical
Chemistry
163. Liu, M. and Z. Wu, Significance of interfaces in solid-state cells with porous electrodes of
mixed ionicelectronic conductors. Solid State Ionics, 1998. 107: pp. 105110
164. McHardy, J., J.M. Baris, and P. Stonehart, Investigation of hydrophobic porous electrodes. I.
Differential capacitance by a low frequency a. c. impedance technique. Journal of Applied
Electrochemistry, 1976. 6: pp. 371376
165. Meyers, J.P., M. Doyle, R.M. Darling, and J. Newman, The impedance response of a porous
electrode composed of intercalation particles. Journal of the Electrochemical Society, 2000.
147: pp. 29302940
166. Newman, J.S. and C.W. Tobias, Theoretical analysis of current distribution in porous elec-
trodes. Journal of the Electrochemical Society, 1962. 1962: p. 1183
167. Nguyen, P.H. and G. Paasch, Transfer matrix method for the electrochemical impedance of
inhomogeneous porous electrodes and membranes. Journal of Electroanalytical Chemistry,
1999. 460(18): pp. 6379
168. Posey, F.A. and T. Morozumi, Theory of potentiostatic and galvanostatic charging of the
double layer in porous electrodes. Journal of the Electrochemical Society, 1966. 113:
pp. 176184
169. Prins-Jansen, J.A., G.A.J.M. Plevier, K. Hemmes, and J.H.W. Wit, An ac-impedance study
of dense and porous electrodes in molten-carbonated fuel cells. Electrochimica Acta, 1996.
41: pp. 13231329
170. Rossberg, K., G. Paasch, L. Dunsch, and S. Ludwig, The influence of porosity and the na-
ture of the charge storage capacitance on the impedance behaviour of electropolymerized
polyaniline films. Journal of Electroanalytical Chemistry, 1998. 443: p. 49
171. Song, H.-K., Y.-H. Jung, K.-H. Lee, and L.H. Dao, Electrochemical impedance spectroscopy
of porous electrodes: the effect of pore size distribution. Electrochimica Acta, 1999. 44:
pp. 35133519
172. Bisang, J.M., K. Juttner, and G. Kreysa, Potential and current distribution in porous elec-
trodes under charge-transfer kinetic control. Electrochimica Acta, 1994. 39(8/9): pp. 1297
1302
173. Bisquert, J., Influence of the boundaries in the impedance of porous film electrodes. Physical
Chemistry Chemical Physics, 2000. 2: pp. 41854192
174. Choi, Y.-M. and S.-I. Pyun, Effects of intercalation-induced stress on lithium transport
through porous LiCoO2 electrode. Solid State Ionics, 1997. 99: pp. 173183
175. Hitz, C. and A. Lasia, Experimental study and modeling of impedance of the her on porous
Ni electrodes. Journal of Electroanalytical Chemistry, 2001. 500: pp. 213222
176. Lasia, A., Nature of the two semi-circles observed on the complex plane plots on porous elec-
trodes in the presence of a concentration gradient. Journal of Electroanalytical Chemistry,
2001. 500: pp. 3035
177. Lindbergh, G., Experimental determination of the effective electrolyte conductivity in porous
lead electrodes in the lead-acid battery. Electrochimica Acta, 1997. 42(8): pp. 12391246
Electrochemistry, Nanomaterials, and Nanostructures 147

178. Liu, C., J.E. Szecsody, J.M. Zachara, and W.P. Ball, Use of the generalized integral trans-
form method for solving equations of solute transport in porous media. Advances in Water
Resourses, 2000. 23: pp. 483492
179. Lundqvist, A. and G. Lindbergh, Kinetic study of a porous metal hydride electrode. Elec-
trochimica Acta, 1999. 44: pp. 25232542
180. Pell, W.G. and B.E. Conway, Analysis of power limitations at porous supercapacitor elec-
trodes under cyclic voltammetry modulation and dc charge. Journal of Power Sources, 2001.
96: pp. 5767
181. Perez, J., E.R. Gonzalez, and E.A. Ticianelli, Oxygen electrocatalysis on thin porous coating
rotating platinum electrodes. Electrochimica Acta, 1998. 44: pp. 13291339
182. Song, H.-K., H.-Y. Hwang, K.-H. Lee, and L.H. Dao, The effect of pore size distribution on
the frequency dispersion of porous electrodes. Electrochimica Acta, 2000. 45: pp. 22412257
183. Srikumar, A., T.G. Stanford, and J.W. Weidner, Linear sweep voltammetry in flooded porous
electrodes at low sweep rates. Journal of Electroanalytical Chemistry, 1998. 458: pp. 161
173
184. Yang, T.-H. and S.I. Pyun, A study of the hydrogen absorption reaction into alfa- and beta-
LaNi5Hx porous electrodes by using electrochemical impedance spectroscopy. Journal of
Power Sources, 1996. 62: pp. 175178
185. Bisang, J.M., K. Juttnerr, and G. Kreysa, Electrochemica Acta, 1994. 39: p. 1297
186. Posey, F.A., Journal of Electrochemical Society, 1964. 111: p. 1173
187. Scott, K., Journal of Applied Electrochemistry, 1983. 13: p. 709
188. Tilak, B.V., S. Vankatesh, and S.K. Rangarajan, Journal of Electrochemical Society, 1989.
136: p. 1977
189. Bisquert, J., G. Garcia-Belmonte, P.R. Bueno, E. Longo, and L.O.S. Bulhoes, Impedance of
constant phase element (CPE) -blocked diffusion in film electrodes. Journal of Electroanalyt-
ical Chemistry, 1998. 452: pp. 229234
190. Bisquert, J., G. Garcia-Belmonte, F. Fabregat-Santiago, and P.R. Bueno, Theoretical models
for ac impedance of diffusion layers exhibiting low frequency dispersion. Journal of Electro-
analytical Chemistry, 1999. 475: p. 152
191. de Levie, R., in Advances in Electrochemistry and Electrochemical Engineering, P. Delahay,
Editor. 1967, Interscience: New York
192. Grebenkov, Transport Laplacien Aux Interfaces Irregulieres: Etude Theorique, Numerique

Et Experimentale. 2004, Ecole Polytechnique: Paris
193. Paasch, G. and P.H. Nguyen, Electrochem. Appl., 1997. 1: p. 7
194. Pakossy, T., Solid State Ionics, 1997. 94: p. 123
195. Presa, M.J.R., R.I. Tucceri, M.I. Florit, and D. Posadas, Constant phase element behavior
in the poly(o-toluidine) impedance response. Journal of Electroanalytical Chemistry, 2001.
502: pp. 8290
196. Sadkowski, A., On the ideal polarisability of electrode displaying CPE-type capacitance
dispersion. Journal of Electroanalytical Chemistry, 2000. 481: pp. 222226
197. Sadkowski, A., Response to the Comments on the ideal polarisability of electrodes display-
ing CPE-type capacitance by G. Lang, K. E. Heusler. Journal of Electroanalytical Chem-
istry, 2000. 481: pp. 232236
198. Zoltowski, P., On the electrical capacitance of interfaces exhibiting constant phase element
behaviour. Journal of Electroanalytical Chemistry, 1998. 443: pp. 149154
199. Zoltowski, P., Comments on the paper On the ideal polarisability of electrodes displaying
CPE-type capacitance by A. Sadkowski. Journal of Electroanalytical Chemistry, 2000. 481:
pp. 230231
200. Beaulieu, L.Y., D. Larcher, R.A. Dunlap, and J.R. Dahn, Journal of Electrochemical Society,
2000. 147: p. 3206
201. Peter, L.M., E.A. Ponomarev, G. Franco, and N.J. Shaw, Electrochemica Acta, 1999. 45:
pp. 549560
202. Peter, L.M. and J. Vanmaekelbergh, in Advances in Electrochemical Science and Engineer-
ing, R.C. Alkire and D.M. Kolb, Editors. 1999, New York: Wiley
148 P.R. Bueno and C. Gabrielli

203. Motheo, A.J., A. Sadkowski, and R.S. Neves, Journal of Electroanalytical Chemistry, 1998.
455: p. 107
204. Macdonald, J.R. and D.R. Franceschetti, in Impedance Spectroscopy, J.R. Macdonald, Editor.
1987, Wiley: New York. pp. 84132
205. Sapoval, B., J.-N. Chazalviel, and J. Peyri`ere, Electrical response of fractal and porous in-
terfaces. Physical Review A, 1988. 38(11): pp. 58675887
206. Keiser, H., K.D. Beccu, and M.A. Gutjahr, Electrochimica Acta, 1976. 21: p. 539
207. Diard, J.P., B. Le Gorrec, and C. Montella, Linear diffusion impedance. General expression
and applications. Journal of Electroanalytical Chemistry, 1999. 471: pp. 126131
208. Deslouis, C., C. Gabrielli, M. Keddam, A. Khalil, R. Rosset, B. Tribollet, and M. Zidoune,
Impedance techniques at partially blocked electrodes by scale deposition. Electrochimica
Acta, 1997. 42(8): pp. 12191233
209. Lang, G. and K.L. Heusler, Comments on the ideal polarisability of electrodes displaying
CPE-type capacitance dispersion. Journal of Electroanalytical Chemistry, 2000. 481: pp.
227229
210. Kerner, Z. and T. Pajkossy, Impedance of rough capacitive electrodes: the role of surface
disorder. Journal of Electroanalytical Chemistry, 1998. 448: pp. 139142
211. Lang, G. and K.E. Heusler, Remarks of the energetics of interfaces exhibiting constant phase
element behaviour. Journal of Electroanalytical Chemistry, 1998. 457: pp. 257260
212. Liu, S.H., Fractal model for the ac response of a rough interface. Physical Review Letters,
1985. 55: pp. 529532
213. Gohr, H. and C.A. Schiller, Electrochimica Acta, 1993. 38: p. 1961
214. Eloot, K., F. Debuyck, M. Moors, and A.P. van Peterghem, Journal of Applied Electrochem-
istry, 1995. 25: p. 334
215. Eloot, K., F. Debuyck, M. Moors, and A.P. van Peterghem, Journal of Applied Electrochem-
istry, 1995. 25: p. 326
216. Sodergren, S., A. Hagfeldt, J. Olsson, and S.E. Lindquist, Journal of Physical Chemistry B,
1998. 98: pp. 55525556
217. Cao, F., G. Oskam, and P.C. Searson, Journal of Physical Chemistry B, 1996. 100: pp. 17021
17027
218. Vanmaekelbergh, J. and P.E. de Jongh, Journal of Physical Chemistry B, 1999. 103:
pp. 747750
219. de Jongh, P.E. and J. Vanmaekelbergh, Journal of Physical Chemistry B, 1997. 101:
pp. 27162722
220. de Jongh, P.E. and J. Vanmaekelbergh, Physical Review Letters, 1996. 77: pp. 34273440
221. Sidebottom, D.L., P.F. Green, and R.K. Brow, Anomalous-diffusion model of ionic transport
in oxide glasses. Physical Review B, 1995. 51: p. 2770
222. Maass, P., J. Petersen, A. Bunde, W. Dieterich, and H.E. Roman, Non-Debye relaxation in
structurally disordered ionic conductors: Effect of Coulombic interaction. Physical Review
Letters, 1991. 66: p. 52
223. Bisquert, J., G. Garcia-Belmonte, F. Fabregat-Santiago, and A. Compte, Anomalous trans-
port effects in the impedance of porous film electrodes. Electrochemistry Communications,
1999. 1: pp. 429435
224. Garcia-Belmonte, G., J. Bisquert, E.C. Pereira, and F. Fabregat-Santiago, Anomalous trans-
port on polymeric porous film electrodes in the dopant-induced insulator-to-conductor
tansition analyzed by electrochemical impedance. Applied Physics Letters, 2001. 78(13):
pp. 18851887
225. Bassler, H., Charge transport in disordered organic photoconductors. Physics Status Solidii
(b), 1993. 175: pp. 1556
226. Bassler, H., P.M. Borsenberger, and R.J. Perry, Charge transport in poly(methyl-
phenylsilane): the case of superimposed disorder and polaron effects. Journal of Polymer
Science: Part B: Polymer Physics, 1994. 32: pp. 16771685
227. Bernasconi, J., H.U. Beyeler, S. Strassler, and S. Alexander, Anomalous frequency-dependent
conductivity in disordered one-dimensional systems. Physical Review Letters, 1979. 42: p.
819
Electrochemistry, Nanomaterials, and Nanostructures 149

228. Bisquert, J. and G. Garcia-Belmonte, Scaling properties of thermally stimulated currents in


disordered systems. Journal of Non-Crystalline Solids, 1999. 260: pp. 109115
229. Borsenberger, P.M., L. Pautmeier, and H. Bassler, Charge transport in disordered molecular
solids. The Journal of Chemical Physics, 1991. 94: pp. 54475454
230. Bouchaud, J.P. and A. Georges, Anomalous diffusion in disordered media: statistical mecha-
nismis, models and physical applications. Physics Reports, 1990. 195: pp. 127293
231. Brown, R. and B. Esser, Kinetic networks and order-statistics for hopping in disordered
systems. Philosophical Magazine B, 1995. 72: pp. 125148
232. Bunde, A. and P. Maass, Diffusion in disordered systems: non-Debye relaxation due to long-
range interactions. Journal of Non-Crystalline Solids, 1991. 131133: p. 1022
233. Dieterich, W., D. Knodler, and P. Pendzig, Relaxation of charged particles in disordered
systems. Journal of Non-Crystalline Solids, 1994. 172: p. 1237
234. Dyre, J.C., A simple model of a.c. conductivity in disordered solids. Physics Letters, 1985.
108A: p. 457
235. Dyre, J.C., The random free-energy barrier model for ac conductivity in disordered solids.
Journal of Applied Physics, 1988. 64: p. 2456
236. Dyre, J.C. and T.B. Schroder, Universality of ac conduction in disordered solids. Reviews of
Modern Physics, 2000. 72: p. 873
237. Dyre, J.C. and T.B. Schroder, Effective one-dimensionality of universal ac hopping condution
in the extreme disorder limit. Physical Review B, 1996. 54: pp. 1488414887
238. Dyre, J.C. and J.M. Jacobsen, Universality of anomalous diffusion in extremely disordered
systems. Chemical Physics, 1996. 212: p. 61
239. Dyre, J.C., Universal low-temperature ac conductivity of macrossopially disordered non-
metals. Physical Review B, 1993. 48: pp. 1251112526
240. Dyre, J., Univeral ac conductivity in nonmetallic disordered solids at low temperatures.
Physical Review B, 1993. 47: p. 9128
241. Dyre, J.C., Some remarks on ac conduction in disordered solids. Journal of Non-Crystalline
Solids, 1991. 135: p. 219
242. Lee, P.A. and T.V. Ramakrishnan, Disordered electronic systems. Reviews of Modern
Physics, 1985. 57: pp. 287337
243. Macdonald, J.R., Analysis of ac conduction in disordered solids. Journal of Applied Physics,
1989. 65: pp. 48454853
244. Mott, N.F., Electrons in disordered structures. Advances in Physics, 1967. 16: p. 49
245. Moura, F.A.B.F. and M.L. Lyra, Delocalization in the 1D Anderson model with long-range
correlated disorder. Physical Review Letters, 1998. 81: p. 3735
246. Scher, H. and M. Lax, Stochastic transport in a disordered solid. I. Theory. Physical Review
B, 1973. 7: pp. 44914502
247. Scher, H. and M. Lax, Stochastic transport in a disordered solid. II. Impuruty conduction.
Physical Review B, 1973. 7: pp. 45024519
248. Ziman, J.M., Models of Disorder. 1979, Cambridge: Cambridge University. Press. pp. 370
385
249. Sheng, P. and J. Klafter, Hopping conductivity in granular disordered systems. Physical Re-
view B, 1983. 27: pp. 25832586
250. Schroder, T.B. and J.C. Dyre, Scaling and universality of ac conduction in disordered solids.
Physical Review Letters, 2000. 84: p. 310
251. Schirmacher, W., Anomalous diffusion in disordered systems: an effective medium descrip-
tion. Berichte der Bunsengesellschaft fur Physical Chemie, 1991. 95: pp. 368376
252. Gefen, Y., A. Aharony, and S. Alexander, Anomalous diffusion on percolating clusters. Phys-
ical Review Letters, 1983. 50: pp. 7780
253. Wang, Q., S. Ito, M. Gratzel, F. Fabregat Santiago, I. Mora-Sero, J. Bisquert, T. Bessho, and
H. Imai
Nanotechnology for Fuel Cells

Angelika Heinzel and Uwe Konig

Abstract There are noteworthy developments in nanotechnology and its relevance


to the energy field. Fuel cells especially benefit from electrodes and membrane
electrolytes with nanostructured and therefore enlarged surfaces. Fuel cells also
derive benefits from the development of nanoparticles and nanotubes for catalytic
application, allowing also study of the molecular electrochemical behaviour. In this
chapter we describe the impact of nanotechnology in the performance of the dif-
ferent components of the fuel cell as well as the impact of nanotechnology in the
electrochemistry process.

1 Introduction

1.1 What Relevance Has Nanotechnology for Fuel Cell


Systems [1]?

Energy research is becoming increasingly important, particularly as regards the role


it plays in support of a wide range of key policies (e.g. security and diversification
of energy supply, energy market liberalization, sustainable development). Nanotech-
nology shows promising potential in all segments of the energy sector: production,
storage, distribution and conversion.
There are noteworthy developments in nanotechnology and its relevance to the
energy field. Fuel cells especially benefit from electrodes and membrane electrolytes
with nanostructured and therefore enlarged surfaces. Fuel cells also derive benefits
from the development of nanoparticles and nanotubes for catalytic application, al-
lowing also study of the molecular electrochemical behaviour.

A. Heinzel ()
Fachgebiet Energietechnik, Universitat Duisburg, Lotharstr. 1-21, 47057 Duisburg, Germany
e-mail: a.heinzel@uni-duisburg.de

E.R. Leite (ed.), Nanostructured Materials for Electrochemical Energy Production 151
and Storage, Nanostructure Science and Technology, DOI 10.1007/978-0-387-49323-7 4,
c Springer Science+Business Media LLC 2009
152 A. Heinzel and U. Konig

However, the fuel cell catalysts have drawbacks; they are both expensive
and have limited efficiency. To solve this issue, research work is being done in
nanostructuring the carbon electrodes to avoid the deactivation of catalysts by
e.g. agglomeration and therefore reduce the amount of noble metals and increase
the catalyst performance.
In spite of the promising potentials of fuel cells, most analysts do not believe that
fuel cells will be widely used in the coming 20 years [1]. Fuel cells for portable ap-
plications are appraised as most promising. Although the detailed problems depend
on the type of fuel cell the main obstacles are the same:
The central problem inhibiting a wider market penetration of fuel cells is the high
manufacturing costs. The costs of fuel cells were approximately 20,000 Euro per
kilowatt power in 2002 [2].
The problems are mainly caused by expensive materials used in fuel cell tech-
nology.
Technical challenges such as thermal as well as mechanical expansion, seals,
lifetime and reproducible properties have also to be solved.
However, a growing number of companies are confident that they are now on the
verge of bringing prices for fuel cells down to levels where they can compete with
conventional electric-power generating equipment.
The most prominent nanostructured materials in fuel cells are the electrocata-
lysts of low- and medium-temperature fuel cells, which consist of carbon-supported
precious metal particles in the range of 15 nm. This structure is necessary to in-
crease the surface to volume ratio of the noble metals, thus reducing the costs of the
material.
A further approach to nanostructured materials is the introduction of nanoscale
hydrophilic (high affinity to water) inorganic materials such as silica into the
polymer membranes being used as electrolyte to improve water retention of the
membrane at elevated temperatures. The functioning of the cell using sulfonated
membranes such as NafionTM is related to the hydrogen ion conductivity of the
membrane, which decreases strongly if water content is not sufficient. Thus, alter-
native membranes with defined properties in the nanometer range are intensively
under development.

1.2 Fuel Cell Technology and Nanotechnology

Fuel cells as efficient energy conversion devices are one of the present R&D sub-
jects in energy technology. Hydrogen and oxygen from air may be converted directly
by an electrochemical process into electrical energy and heat. High efficiency val-
ues are achievable even in the lower power range. As fuel cells can be operated
continuously as long as fuel is available, various applications are interesting. These
range from combined heat and power supply systems in the 100 kW to the 1-kW
power segment, to mobile (electric traction) or even portable applications. The main
obstacle the use of hydrogen as fuel might be overcome by gas processing tech-
nologies, converting fossil or biogenic fuels into a hydrogen-rich gas mixture.
Nanotechnology for Fuel Cells 153

H2
AFC OH O2 100 C
H2O

PEFC H'
O2 80 C
H2
PAFC H 2O
200 C
H2 O2
MCFC CO2
H2O
CO22
CO2 650 C
H2
SOFC H2O O2 O2 1000 C

fuel gas oxygen

anode elektrolyte cathode


Fig. 1 Operation principle of the various types of fuel cells: PEMFC polymer electrolyte mem-
brane fuel cell, AFC alkaline fuel cell, PAFC phosphoric acid fuel cell, MCFC molten carbonate
fuel cell, SOFC solid oxide fuel cell

Different material combinations for practical realization of fuel cells have been
developed in the past few decades [3]; comprehensive overviews about the tech-
nology have been published [4]. Possible operation temperatures of fuel cells range
from ambient temperature to 1,000 C. The operation principle of the different fuel
cells is depicted in Fig. 1. The main components are the same for all types of fuel
cells and comprise an electrolyte, catalytically active electrodes, and a cell frame
for gas distribution and current collection. Regular nanostructures are not typically
used until now, but nanomaterials for preparation of layers are frequently the best
base materials. Some examples will be given in the description of the five types of
fuel cells in this introductory chapter.
The main electrochemical reactions for hydrogen-fed fuel cells are as follows:

Anode: H2 2H+ + 2e
1
Cathode: O2 + 2e O2
2
1
H2 + O2 H2 O
2
This reaction should theoretically lead to a cell voltage of 1.23 V, practically, cell
voltages of 1 V at zero current (open circuit voltage) and 0.5 V during operation
are achieved. These low-voltage values per cell lead to the requirement of series
connection of various single cells to form a so-called cell stack.
For most of the cell types, a layered bipolar construction is state of the art, shown
in Fig. 2, for the example of a polymer electrolyte membrane fuel cell.
154 A. Heinzel and U. Konig

Fig. 2 Components of a membrane fuel cell

The functions of the different layers are as follows:


Cell frames contain a system of gas channels integrated in the bipolar plate, the
so-called flow field for the distribution of fuel and air evenly over the entire active
electrode area of the anode and the cathode; the dimensions of channels and
ribs are usually in the millimetre range. The cell frames act as well as current
collectors, and thus a good electrical conductivity is required. In addition, the cell
frames can contain cooling channels, either for air cooling or for liquid cooling.
Various cooling concepts have already been realised.
The second layer is a gas diffusion layer (GDL), typically a carbon paper or car-
bon cloth. This GDL protects the membrane from mechanical damages, ensures
spatial electrical contact, and is important for gas distribution to the electrode
parts under the ribs of the flow field and for removal of product water. GDLs
are a porous system formed by hydrophobic and hydrophilic pores in the sub-
millimeter range. Its compressibility is important for stack construction, because
small deviations in thickness of sealings and bipolar plates can be compensated.
The third layer is the electrode, a thin catalyst layer that might be coated onto
the GDL or in most cases onto the membrane, then forming a so-called mem-
brane/electrode assembly MEA. For operation at low (ambient) temperatures,
noble metals are required as catalyst. The goal of cost reduction led to the devel-
opment of catalyst systems consisting of a carbon carrier material with a noble
metal loading of about 0.4 mg cm2 ; lower loadings are envisaged for future elec-
trode materials. A large electro-catalytically active surface area is important for
achieving high current densities, thus preparation processes are used leading to
nanoparticles of noble metals on the larger carbon particles (see Fig. 3). Another
issue is CO tolerance of the electrocatalyst, if the fuel cell shall be operated with
reformate generated from fossil fuels instead of pure hydrogen.
The fourth layer is the polymer membrane itself. The membrane is the elec-
trolyte. Proton conductivity is its most important property besides the safe
separation of the fuel in the anode compartment of the cell and the air in the cath-
ode compartment. Chemical stability is achieved by using a fluorinated polymer
Nanotechnology for Fuel Cells 155

Fig. 3 Electrocatalyst as prepared for use in PEMFC (source: MPI Mulheim)

backbone. Proton conductivity is realised by sulfonic acid groups attached to side


chains of this backbone. The first polymer electrolyte of this type was developed
in the early seventies [5]. The hydrophobic backbone and the hydrophilic acid
groups lead to a structure, showing hydrophilic channels in a hydrophobic ma-
trix. Proton conductivity additionally requires the presence of a sufficient amount
of liquid water [6] (see Fig. 4).
Thus, for membrane fuels cells, water management in the different layers is im-
portant to be properly controlled, ensuring a good humidity of the membrane but
avoiding flooding of the porous GDL. Another important aspect is the structure of
the so-called three-phase boundary in a gas diffusion electrode. The electrodes in
PEMFC operating with gaseous fuel in fact constitute a four-phase boundary, even
more complicating the facts. The first phase is the gaseous phase, containing either
the hydrogen or the oxygen, which shall be transported to the reaction zone. The sec-
ond phase is liquid water being generated as product and contributing to the ionic
conductivity of the membrane. Part of the water is necessary to conduct hydrated
protons from the anode to the cathode side; part of the water has to be removed in
order to avoid flooding of the gas diffusion structure. The third phase is the solid
electrolyte material, necessarily being hydrated with liquid water and also being in
close contact with the sites, where ions are generated. The fourth phase finally is the
electrocatalyst, also in close contact with the fuels in order to catalyse the electron
transfer reaction and to conduct the electrons to the outer circuits. This interface is
a unique structure of the membrane fuel cell, and much effort was undertaken to
optimize it [7].
156 A. Heinzel and U. Konig

0,12

0,1
Specific conductivity [S /cm]

0,08

0,06

0,04

0,02

0
0 5 10 15 20 25 30
water content = NH2O / NSO3H

Fig. 4 Specific conductivity of NafionTM 117 as function of water content

The same type of membrane fuel cell can be fed with a liquid solution of
methanol in water instead of hydrogen or reformate as fuel. The electrochemical
reactions are as follows:

Anode: CH3 OH + H2 O CO2 + 6H+ + 6e


3
Cathode: O2 + 6H+ + 6e 3H2 O
2
3
O2 + CH3 OH CO2 + H2 O
2
For methanol as fuel, the cell voltage calculated from thermodynamic data is
1.215 V, but here in practice open circuit voltages of about 0.7 V are achieved and
0.40.3 V is achieved during operation.
Because of the formation of the poisoning intermediate adsorbate CO, the achiev-
able current densities are limited, and higher noble metal catalyst loadings are
required, and CO tolerance also is an important issue. Special MEAs have been
developed for direct methanol fuel cells (DMFC). A polymer membrane in contact
with methanol shows significant transfer of methanol and water from the anode side
to the cathode side of the fuel cell, leading to losses due to formation of a mixed
potential at the air electrode. Typical MEAs are shown in Fig. 5, where the variety
of nanostructures is also pointed out. But the advantageous simplicity of the fuel
cell system and the high energy density of the liquid fuel make the DMFC attractive
despite these mentioned drawbacks. For a liquid-fed DMFC, the requirements for
the structure of the interfacial catalyst layer are different than for gaseous hydrogen
as fuel, which is thus also a matter for optimisation [8].
Nanotechnology for Fuel Cells 157

Fig. 5 Scanning electron microscopy (SEM) pictures of DMFC electrodes: upper E-Tek elec-
trode, lower electrode made by DLR

In addition to the earlier described state of the art, nanotechnology plays a role in
the development of micro fuel cells; the realisation of special properties of surfaces
and the enhancement of functionalities by nanostructures, nanolayers as coatings
and nanoparticles raise increasing interest. Nanostructured electrolytes, carbon sup-
ports or coatings for bipolar plates are examples.
The alkaline fuel cell (AFC) with its liquid alkaline electrolyte KOH uses gas
diffusion electrodes with a hydrophobic porous part, which is not flooded by the
alkaline electrolyte, and a hydrophilic part containing electrolyte and thus leading
to a three-dimensional three-phase boundary layer. As the electrode potentials in
alkaline electrolyte are shifted towards more negative values, corrosion is less prob-
lematic. Raney Nickel and silver are the state-of-the-art catalysts. The practical use
158 A. Heinzel and U. Konig

of alkaline fuel cells is limited due to the sensitivity of the electrolyte towards CO2 ,
which leads to potassium carbonate precipitation.
The phosphoric acid fuel cell (PAFC) has a quite similar construction and com-
ponents as the PEMFC; the electrolyte is liquid phosphoric acid in an inert matrix.
The operation temperature of 200 C avoids formation of liquid water and improves
CO tolerance of the electrocatalyst. For the catalyst properties, the same require-
ments are valid as for the PEMFC nanoparticles with a high surface area and
a good dispersion on the carbon carrier material are required. The application of
PAFC typically is the combined heat and power supply in the 200-kW power range.
The molten carbonate fuel cell (MCFC) with its operation temperature of 650 C
is developed since the sixties. Corrosion of the metallic cell frames in contact with
the molten salt electrolyte and electrolyte loss during operation still limit the lifetime
of the stacks. The materials used mainly have structures in the micrometer range;
the anode is made of Ni/Cr or Ni/Al with a mean pore size of 6 m; the cathode is
NiO, which is formed from Ni by in-situ oxidation of a porous Ni plate with pores
of a size of 8 m, and the liquid electrolyte is contained in a matrix (LiAlO2 ) and
must be held there by capillary forces. The pore size of the matrix must therefore
be carefully controlled and be smaller than the mean pore size of the electrodes.
Changes of pore diameters during lifetime have to be avoided. Thus, submicron
LiAlO2 is a nanomaterial used in MCFC, and the particle growth during tens of
thousands of operational hours is well investigated [9].
The solid oxide fuel cell (SOFC) consists of solid components, usually the three
layers, namely, anode, electrolyte and cathode, which are manufactured as MEA as
it is the case for the membrane fuel cell. Because of the high operation temperature,
mechanical stress due to different thermal expansion coefficients of the materials
used is a major challenge. Material development for adopted physical properties and
for electrolytes with high conductivity at reduced temperatures of about 500700 C
is the focus. To achieve a good performance, the anodes of SOFC are a well-defined
micro-structural layer of Ni/YZS (Yttria-stabilised Zirconia). Operation at high cur-
rent densities and at high fuel utilisation leads to a significant increase of anode
resistance, and a structural change of the cermet can be observed [10], mainly sin-
tering of the small Ni particles. For realizing a long-term stable high surface area
cathode, a special preparation method was recently examined [11]. Single YSZ par-
ticles are sintered onto the electrolyte, resulting in a high surface area. The surface
then is covered by a thin layer (approximately 100 nm) of (La, Sr)CoO3 by metal-
organic deposition. An increase in power density and in long-term thermal cycling
stability was observed. Another approach to improve the anode using nanolayers
was recently reported [12]. A better long-term stability was achieved with a chan-
nelled anode produced directly by solidification of a NiOYSZ eutectic mixture, by
laser zone melting of rods or plates and subsequent reduction of NiO. A lamellar
thickness between 200 nm and 1 m could be realised.
Summarising the activities using nanotechnology for fuel cells, the membrane
fuel cell is the most promising type. Thus, the focus of the subsequent chapters will
be on this type of fuel cell and will give more details on new approaches using
nanostructured electrocatalysts and membranes.
Nanotechnology for Fuel Cells 159

2 Nanostructures

2.1 General Properties of Electrolyte Membranes

The membrane in a membrane fuel cell fulfils several important functions as stated
in the introduction. NafionTM was the first commercially available membrane, which
lead to a breakthrough in fuel cell technology. Today, various companies are en-
gaged in membrane development especially for this purpose, aiming at improved
material properties. The goals are less sensitivity towards elevated temperature
and dry operation, better chemical and mechanical stability and reduced methanol
crossover for DMFC operation. A significant improvement of the mechanical sta-
bility was achieved by incorporation of a PTFE porous sheet as mechanical support
for the membrane material [13, 14].
Another approach was the synthesis of inorganic/organic composite materials to
influence the properties of the membrane. An overview on the state of the art of
composite perflourinated membranes is given in [15]. Infiltration of a polymer car-
rier material with various inorganic proton conductors is subject of a patent [16].
For operation at elevated temperature, several materials have been considered. In
an early work, NafionTM /H3 PO4 showed better conductivity at temperatures above
100 C compared with blank NafionTM and also reduced methanol permeability [17].
New types of polymers are also under development for better temperature stability;
one of the most advanced examples is the high-temperature material polybenzimi-
dazole, which usually is doped with phosphoric acid [18].
The use of low-cost basic polymers instead of NafionTM is an interesting alterna-
tive [19, 20].The development of new polymers for ionomer membranes including
perfluorinated ionomers, partially fluorinated ionomers, nonfluorinated ionomers,
high-molecular/low-molecular composite membranes as well as novel polymer
modification processes and novel membrane materials is summarised in [21].
The microstructure of NafionTM and sulfonated polyetherketones also was a
matter of recent investigations [22]. For NafionTM , the nanoseparation into hy-
drophobic and hydrophilic domains is well known, with the hydrophilic domain
being responsible for the transport of protons and the hydrophobic backbone pro-
viding the morphological stability and preventing the membrane from dissolving
(Fig. 6). The sulfonated polyetherketones showed a less pronounced nanoseparation
due to the less hydrophobic backbone and the less acidic sulfonic groups.
Nanomaterials come into consideration for various composite materials, in which
different materials contribute to specific advantageous properties.

2.2 Alternative Membranes

There are tremendous efforts in developing alternative polymeric membranes for


PEFC. The main challenge for new membranes is the realisation of high proton
160 A. Heinzel and U. Konig

O
F O
(CF 2) n
C O
(CF2)n H2O
O O
-
SO3 SO3-
CF2
Na+
F3C CF H2O
O
CF2
CF2 H2O
H2O SO3-
SO3- O
Na+
H2O SO3- O

H2O O
O
A C
B

Fig. 6 Structure of a NafionTM membrane according to H. Yeager, A. Eisenberg in Perflu-


orinated Ionomer Membranes, A. Eisenberg, H. Yeager, Eds., ACS Symp. Series No. 180
(American Chemical Society, Washington DC, 1982) pp 16, 4163/ (source: http://www.psrc.
usm.edu/mauritz/nafion.html). A: hydrophobic fluorocarbon region, B: interfacial region, C: hy-
drophilic region with ionic exchange groups, counter ions (Na+ in this picture instead of H+) and
water

mobility in a robust polymer matrix; a defined nanostructure often is a key issue.


Especially the transport properties of the protons and the water content have to be
taken into account. Furthermore, the contact to the catalysts is important.
Nanotechnological approaches will lead to defined structures on the molecular
level by implementing active side groups or isolated particles as well as by crosslink-
ing via side chains. Specially designed chemicals such as ionic liquids (ILs) allow
the immobilisation and as a consequence the defined distribution of active particles
such as catalysts. These kinds of chemicals can also be used to functionalise inor-
ganic structures such as zeolites or molecular sieves.
Nanoparticles will be incorporated into the membranes to enhance the in-situ
generation of hydrogen from e.g. methanol as well as to increase the basic fuel cell
reaction. Moreover, the proton conductivity can be improved by the incorporation
of nanoparticles.
Nanotechnology for Fuel Cells 161

Several types of other proton conducting membranes that incorporate quaternary


nitrogen atoms are presently under investigation:
Polymer blends leading to high-end polymers, e.g. from sulfonated polymers
(sPEEK sulfonated polyether-etherketone, sPPSU sulfonated polyphenyl-
sulfone) combined with alkaline components (amine, imidazole, polybenzim-
idazole): The combination results in ionic cross-linked phases. Commercially
available polymers can be modified by different sulfonation reagents. Another
possibility is to combine different monomers based on block co-polymers. The
conductivity can be controlled by the number of SO3H groups due to the depen-
dence of the water uptake from the number of groups ([23] and references cited
therein).
Novel side-chain polymers with heterocycles such as imidazole attached to
appropriate polymer backbones are used as proton-solvating moieties and for
achieving high proton mobility at high temperatures (>100 C), where poison-
ing effects of the used electrocatalysts are drastically reduced. As opposed to the
conventional membranes, these systems are aprotic; the high proton conductance
does not rely on the presence of water.
Advantages: High thermal stability, high conductivity and lower permeability of
methanol.
Disadvantages: Problems may arise from oxidation of aliphatic bonds; often dif-
ficult to synthesise and expensive; swellable by water uptake.
Dendrimer PTFE copolymers combining hydrophilic dendrimers with hydropho-
bic linear polymers
Advantages: High thermal stability and high conductivity.
Disadvantages: Problems may arise from oxidation of aliphatic bonds; often dif-
ficult to synthesize and expensive; swellable by water uptake.
Composite membranes formed by incorporation of inorganic nanoparticles or
polyacids
Solvated proton conducting polymers and composite membranes should govern
the transport of protons and water by micro structural control. The objective of
this attempt is to take advantage of the high proton conductance in watery sys-
tems and to control the proton conductivity by a sophisticated water management.
Nevertheless, the conductivity is strongly dependent on the water content, which
also influences the mechanical properties by e.g. swelling.
Advantage: High variety of properties possible; humidity systems also to temper-
atures above T > 100 C possible.
Disadvantage: Agglomeration of particles possible; often difficult to synthesize
and expensive; swellable by water uptake.
More details are given in Sect. 2.3.
162 A. Heinzel and U. Konig

Novel Brnstedt acidbase ionic liquids [24, 25], e.g. from organic amines with
(trifluoromethanesulfonyl) amide (HTFSI); electroactive for H2 oxidation and
O2 reduction at a Pt electrode under non-humidifying conditions at moderate
temperatures (ca. 130C).
Advantages: High variety of properties possible; operation at temperatures above
T > 100C possible; no water content; no swelling; high conductivity.
Disadvantages: Ionic liquids have to be supported by e.g. ceramic systems; often
difficult to synthesize; presently expensive due to scientific state.
Ionic liquid mixtures such as a 4:6 mixture of methyl and dimethyl ammonium
nitrate: The incorporation of imidazol derivates results in a better stability up
to T = 180 C with a conductivity of 0.1 S cm1 [26]. The stabilisation of the
ionic liquids is possible by introducing the substances into a porous silica matrix
[27,28]. A typical liquid is the 1-butyl-3-methyldiazonium derivate with addition
of BF4 and H3 PO4 to fit the acidity. The system shows a thermal stability up to
T = 300C with a conductivity up to 80 mScm1 [29].
Functional monomers are added to an adequate backbone, e.g. the addition of
alkylimidazilium to a vinyl- or allyl backbone. The resulting polymer form a
cationic structure where the anions can move at (Fig. 7).
Advantages: High variety of properties possible; operation at temperatures above
T > 100 C possible; no water content; no swelling; high conductivity; high
chemical and mechanical stability.
Disadvantages: Ionic liquids have to be supported by e.g. ceramic systems; often
difficult to synthesize; presently expensive due to scientific state.
In all cases disadvantages arise from the possibility of oxidation of the aliphatic
bonds and the fact that the systems are often difficult to synthesise and the costs
are high. Detailed information about the state of the art of ionic liquids is given in
Sect. 2.5.

R +
R + N N (CH2)n
N N N N
RX X Kat. X
R
R R +
R +
N N N N (CH2)n
N N
X
X

n = 0 und 1
Fig. 7 Cationic polymerisation of IL
Nanotechnology for Fuel Cells 163

2.3 Nanoparticles for Improved Membrane Properties


Composite Membranes

The modification of membrane properties by inorganic materials raised increased


interest in the past few years. Facilitating the water management especially for
low-humidity operation conditions and improving the mechanical and thermal prop-
erties usually is the goal; a reduced methanol crossover could also be a desired re-
sult. The homogeneous dispersion of small particles in the polymer matrix is of
outstanding importance for good membrane properties. Thus, manufacturing proce-
dures leading to nanoparticles are typically used. Two main preparation procedures
are known [30]: the in-situ growth of inorganic materials and the polymer in-situ
sol/gel reaction.
The incorporation of inorganic nanoparticles with a high affinity to water into the
electrolyte polymers, such as for example silica, improves the water retention of the
membrane. Different approaches have been pursued in the past. The incorporation of
hydrophilic compounds into recast NafionTM films was intensively investigated. Zir-
conium phosphate is one of the most prominent inorganic compounds that exhibit
proton conductivity and a layered structure. Hybrid membranes can be produced
[31]. It is important to maintain the nanoscale platelet structure of the inorganic
particles. For example in [32], layered phosphates of titanium and zirconium were
prepared and investigated with respect to proton conductivity. The inorganic addi-
tive was reported to increase the stiffness of the membrane and reduce the methanol
transfer. It finally leads to a slightly reduced proton conductivity but at the same
time to a smaller influence of temperature on conductivity so that the conductiv-
ity values for 130 C are nearly the same for the composite membrane and blank
NafionTM , due to reduced conductivity of the latter at elevated temperature caused
by drying out of the NafionTM membrane. Tailor-made layered structures of vari-
ous zirconium phosphates and zirconium-sulfophenyl phosphates as oriented lamel-
lar structures were also investigated [33]. As membranes, NafionTM and sulfonated
polytherketones were used as polymer materials, and it was found that the improve-
ment of membrane properties for polyetherketones was less pronounced than for
composites with NafionTM . The performance of Nafion-composite membranes at an
elevated temperature of 110 C could significantly be improved, as hydration of the
composite could be maintained at a high level.
Nanocomposite membranes consisting of SiO2 /polyethlenoxide (PEO) have
been synthesised [34]. The molecular design of the nanocomposites was achieved
by nanosized interfusion among organic, inorganic and acidic molecules. The or-
ganic and inorganic components were hydrolysed at a nanoscale. The resulting
hybrid materials can have quite different properties than a linear combination of the
individual bulk properties. The hybrids can be structurally and chemically modified
to form nanosized interconnecting networks.
An increasing interest exists in adding heteropoly acids (HPA) as proton-
conducting components to the sulfonated polymers [35]. The addition will enhance
the proton conductivity and the acceptance of CO. Because of their structural
164 A. Heinzel and U. Konig

diversity these materials can be incorporated into a wide variety of membrane ma-
terials. The HPA have interesting redox and catalytic properties, which are not fully
understood yet.
The addition of superacid metal (IV) phosphonates is particularly suitable for
the preparation of hybrid membranes. The proton conductivity in some cases
reaches values even higher than 0.1 S cm1 . The presence of nanoparticles of metal
phosphonates in the electrode interface Nafion/Pt already improves the electro-
chemical characteristics of fuel cells in the temperature range 80130 C [36].
With monodecylphosphate and phosphotungstic acid, flexible, transparent and
homogeneous hybrid membrane materials could be synthesised. The temperature
stability reached 250 C, and the conductivity of the humidified membrane reached
1 103 S cm1 at 80 C.
Nanosized ceramic powders with good adsorption capacity for acids together
with a polymer binder not an ionomer were also reported to be developed by Tel
Aviv University [37]: silicon dioxide (150 nm), alumina (50 nm) and titanium diox-
ide (21 nm) were used as nanopowders. With PVDF as polymer binder, membranes
with nanosized pores were cast, the pore diameter mainly being below 2 nm. The
use of such a membrane consisting of PVDF, SiO2 and triflic acid in a DMFC led to
high power densities of up to 500 mW cm2 [38]. The membrane material had 10
20 times higher water permeation than NafionTM , and it was expected that this high
permeation leads to a high back-diffusion of water to the anode side, thus avoiding
the flooding of the cathode catalyst and GDL.
All these membranes still have to prove their practical applicability.

2.4 Nanostructured Membranes

A first attempt to use an array of nanochannels as an electrolyte is described in [39].


The effect of the enhancement of proton conductivity by overlapping electrical dou-
ble layers was intended to be used to fabricate an improved micro fuel cell. First ex-
periments with micro channels of depth between 50 nm and 50 m were investigated
using diluted aqueous HClO4 as electrolyte. Normalising the experimental results
showed an increase in apparent proton conductivity for the low values of channel
depth and being higher than the bulk proton conductivity. The effect was even ob-
served with channels of depth 12 m, which is one order of magnitude larger than
the thickness of the electrical double layer. A concise scientific explanation of the
observed results still is lacking.
During fuel cell operation the membranes are stressed mainly by mechanical in-
terference. Differences in the local gas and water distribution would lead to different
processes in chemical reactions, shrinkage or expansion. The mechanical stress can
induce changes in the distribution of the catalytic nanoparticles by e.g. agglomera-
tion at fissures.
Intensive efforts are made in the field of gas separation of industrial gas mixtures.
Membranes with a porosity of less than 1 nm allow the uncomplicated separation
of propylene/propane, benzene/cyclohexane and high-pressure CO2 /CH4 mixtures
Nanotechnology for Fuel Cells 165

[40]. The reaction takes place via a solution-diffusion mechanism, which requires a
well-defined modification of the membrane in the nanometer range. The application
of these types of membrane is also useful for fuel cells since they allow a molecular
transport of the fuel gas and water [41].

2.5 Ionic Liquids (ILs)

The use of ionic liquids in fuel cells is a further attempt to develop new types of
membranes. Because of the current development in the field of ionic liquids, this
material class is a promising alternative to the other attempts of polymer materials.
The general advantage of the ionic liquids is that the conductivity is independent of
the water content.
The main nanotechnological approach of the ionic liquids is the opportunity to
immobilise the catalysts by covering the nanoparticles with a liquid phase onto a
solid surface. Huang et al. immobilised Pd nanoparticles on molecular sieves by
ionic liquids [42]. The catalytic system was used for solvent-free hydrogenation.
The combination of nanoparticles, ionic liquids and solid surface showed excellent
synergistic effects to enhance the activity and durability of the catalyst. In fuel cell
systems, this approach can be used to enhance the catalytic oxidation of hydrogen
by improved immobilisation. Another approach is related to the functionalisation of
nafion membranes by replacement of water by ionic liquid used to realise a suffi-
cient proton conductivity of NafionTM membranes. The main advantage is that such
a system is independent of the water content formed during the fuel cell reaction.
A first approach was published [43] in which NafionTM was swollen in ionic liq-
uids instead of water. These membranes using 1-butyl,3-methyl imidazolium triflate
(BMITf) and BMI tetraflouroborate (BMIBF4 ) as ionic liquids show excellent con-
ductivity at elevated temperatures up to 200 C. BMITf-imbibed membrane samples
(Nafion, Dow) even show higher conductivity values in a temperature range from
40 to 180 C.
The first evaluation of ILs as electrolytes for fuel cells has just been done [25].
This appears to have been the first attempt to apply ILs under aprotic conditions.
Adequate hydrophobic properties of the ionic liquids are necessary to realize
the three-phase boundary necessary for the fuel cell operation. Furthermore, the
transport of water throughout the membrane will be hindered. The water produced
during the reaction will not diffuse through the membrane, and a sophisticated water
management is not necessary.
Ionic liquids are considered more and more as alternatives for conventional elec-
trolytes [44]. The reported ionic conductivity is sufficient enough, even though the
values of 100 mS cm1 are based on the IL itself; they do not include the target ions
such as protons and the primary charge carriers are still not known yet and are under
discussion.
The systems are intensively studied for proton transfer towards fuel cell applica-
tions under water-free conditions ([44] and references therein). These ILs are known
as Brnstedt acidbase ILs and require certain conditions in preparation.
166 A. Heinzel and U. Konig

One of the most interesting systems is the class of imidazoles. They are self-
dissociation compounds with high proton conductivity (>100 mS cm1 ) without any
acid doping. Further enhancement of the conductivity and also the thermal stability
of the system can be realised by the addition of acidic components [45]. This is due
to the proton transfer via the Grotthuss mechanism.
Two other types of ionic liquids are very promising candidates for conducting
polymers. They are ionic liquids based on choline chloride, which have already
shown superior properties in electrochemical processes (e.g. metal finishing) [46],
and single-ended or double-ended diallylammonium ionic liquids, which are protic
compounds with a high potential for excellent proton conductivity [47].
Furthermore, the solubility of the fuel gas oxygen and hydrogen has to be as
small as possible but not completely insoluble. In general, the experimental results
indicate a very small solubility [44,48]. Nevertheless, it is possible that the solubility
of the gases influences each other [49].
It is important to realise a low pH value to enhance the proton conductivity. This
can be done by an adequate anion such as PO4 3 . The challenge in preparing the ad-
equate ionic liquid is the realisation of the Grotthuss mechanism in the proton trans-
port. For example it is shown by Kerr et al. that the linkage of aliphatic chains to the
imidazole molecule results in a decrease in conductivity [50]. This is explained by
the hindering of the N-substituted imidiazole to participate in the Grotthuss mecha-
nism by a structural effect. Kerr et al. also indicate that due to possible volatility the
IL must be fixed to a matrix. Current models developed using Hole theory [51] will
be used to design the compounds with optimum conductivity and viscosity.
The requirements of an adequate liquid are as follows:
Broad electrochemical window
Thermal stability up to 300 C
Stable against hydrolysis
Hydrophobic
Free of halogen
Recyclable
Cheaper than NafionTM
For structuring, the IL has to be immobilised. This can be done using i.e. zeolitic
structures or molecular sieves. It is obvious that with increasing surface area of the
solid phase, the motion of the liquid and the proton transport will be hindered. From
polymerisation experiments it is known that the stiffening of polymers by cross-
linking can be compared with the polymersurface interaction. Electrode surfaces
and solids such as silica, carbon black or cathode powder also stiffen the polymer
[52]. This can be explained by different transport properties at the interfaces. As a
consequence it must be expected that at the surface of the added particles the ionic
liquid will behave in a different way than in the immobilised liquid phase.
Studies of the use of molecular sieve exhibit that Pd nanoparticles can be immo-
bilised by the combination of molecular sieves and ionic liquids. The results indicate
that the nanoparticles immobilised onto the molecular sieve by ionic liquids were
very active and were stable catalysts for the solvent-free hydrogenation of olefins
[42, 53, 54].
Nanotechnology for Fuel Cells 167

Another nanotechnological approach is the use of a metalorganic chemical va-


por deposition (MOCVD) process. This process will be used for producing highly
nanodispersed platinum particles on GDL allowing the decrease of overpotential re-
sistance of oxygen reduction reaction. The interest of the MOCVD process is linked
with vapour penetration inside the first ten m of the substrate (carbon layer for ex-
ample) allowing a 3D repartition of the catalytic element inside the carbon network.
MOCVD process brings then a rational approach of the platinum loading inside
the active layer in terms of platinum accessibility and electroactivity. This MOCVD
process will be developed on a range of catalysts in order to immobilise the catalytic
phase on electrode (carbon support) or on membrane support, and to select a highly
stable catalyst in contact with the ionic liquid [55].
Table 1 compares the relevant properties of various alternative materials.

Table 1 Comparison of conductivity data and values of glass transition temperature for various
membrane materials
Material Test conditions Conductivity Glass transition Remarks
T (C) RH (%) mS cm1 temperature RH = relative
T (C) humidity/water
content

Aqueous systems
FS-PEEK [56] 120 50 30 >140 = f (RH)
BPSH [56] 120 50 55 135 = f (RH)
F-PES [56] 120 50 20 140 = f (RH)
F-PES [57] 120 50 330 90 = f (RH)
HPA Nafion [56] 120 50 30 90 = f (RH)
Nafion 117 [57] 120 50 30 90 = f (RH)
Nafion 1100 [58] 120 50 30 = f (RH)
3M [56] perfluorinated 120 0 40
sulfonic acid
Dupont [56] 120 50 >150 = f (RH)
Ionic liquids
Nafion [26] 3-methyl 180 0 100 200
imidazolium BF4
PBI [58] (polybenzimidazol) 200 30 50 200 = f (RH)
EMImBF4 [59] 1-ethyl-3- 120 <5% 100 = f (RH)
methylimidazolium BF4
BMImBF4 [59] 1-butyl-3- 120 <5% 50 = f (RH)
methylimidazolium BF4
BMIm [29] 1-butyl-3- 200 <5% 80 300 = f (RH)
methylimidazolium
BF4/H3PO4
BMPBETI [59] 1-butyl-3- 120 <5% 20 = f (RH)
pyrazolium BETI
Gel-type membranes [60], 150 100 = f (RH)
various polyether
Silica gel stabilised [27] 230 3080 = f (RH)
imidazolium (IONOGEL)
168 A. Heinzel and U. Konig

3 Electrocatalysts in Polymer Electrolyte Membrane Fuel Cells


(PEMFC) and PAFC

The electrocatalysts for PAFC and PEMFC are quite similar; the increase in op-
eration temperature from 80 C being typical for the PEMFC to 200 C for PAFC
is not so significant that the use of expensive noble metals could be avoided. The
CO tolerance of the PAFC anode is much better with a tolerable level of approx-
imately 1 vol% of CO compared with 10 maximum 100 ppm in short transients
for the PEMFC. Thus, there were a lot of parallels in the catalyst development, and
PEMFC developers could make use of the insight gained by PAFC development
and vice versa. Therefore, the following chapter refers to the PEMFC but uses re-
sults being generated for PAFC.

3.1 General Properties

Possible fuels for PEMFC are besides pure hydrogen, mainly reformate and
methanol. Other alcohols and organic fuels may also be converted but with much
lower current densities. At the low operation temperature of the PEMFC, noble
metal catalysts still can not be replaced, though a lot of research is going on for
developing non-noble metal catalysts [61]. Because of the high cost of noble metals,
the loading of platinum, ruthenium or other noble metals is a focal point: the first
PEMFC prototypes operated with 24 mg cm2 of pure noble metal, but a reduction
to a tenth of this value was soon achieved by preparation of supported catalysts
noble metal nanoparticles on a carbon carrier material. In the same time, current
densities were improved, leading to remarkable lower cost in respect to power
per m2 . The manufacturing methods of supported precious metal catalysts made
significant progress. The first well-known preparation method was impregnation of
the carbon carrier material with a solution of a salt of the respective noble metal.
Drying, reduction and sintering led to an even distribution of nanoparticles on the
carbon surface. Commercial catalysts and catalyst-coated membranes were man-
ufactured. An improvement was achieved by introducing the sol/gel preparation
method to the manufacturing of fuel cell electrocatalysts, leading to even finer
particles [62].
Especially at elevated temperature small particles tend to agglomerate, therefore
a spatial separation is important. It was found [31] that the platinum surface area
correlates with the BET surface of the carbon carrier material. This is easily under-
standable as a better dispersion of the noble metal particles leading to higher elec-
trochemical activity. But as soon as the platinum particles reside in very small pores,
smaller than 40 nm, they do not contribute to the electrochemical reaction anymore
[63]. Thus, optimisation of the carrier material as well as of the composition and
dispersion and manufacturing of the catalysts still is going on.
Nanotechnology for Fuel Cells 169

Because of the importance of this topic, several research groups have investigated
the possibilities for further improvement [64].
In the electrocatalytical field the main challenge is the control of the particle size.
On the one hand the optimum size for e.g. oxygen reduction has to be controlled.
Different values of optimum size are reported ranging from 1.5 up to 50 nm ([65]
and references cited therein). It is supposed that this is related to the different an-
alytical range for particle characterisation and electrochemical measurements. The
former is carried out over a small region of the electrode of some hundreds of nm
whereas the latter is performed on a high surface area of cm scale. Therefore, mi-
croelectrodes have to be used to guarantee a comparable measuring range.
On the other hand the stability of the nanoparticles during the fuel cell reaction
is important. The Pt particles can corrode within the system and will be deposited
at different locations [66]. This effect will lead to an agglomeration of the particles
at fissures [67] (Fig. 8). Furthermore, it is well known that dispersion and surface
area of Pt particles will change during the application of high electrical fields. The
details are reported recently by e.g. Antolini [68].
Since the electrical properties of the nanostructures will govern the catalytic
properties, their change in the nanoregion is important. The quantum size effects tak-
ing place in this regime become more and more important even though the electronic
conducting materials in the quantum confinement regime have not been described
yet [69].

3.2 Relevant Reactions

3.2.1 The Oxygen Reduction Reaction

The reduction of oxygen is the main reason for the overvoltage occurring in a fuel
cell. Thus, lot of emphasis was put into improving the catalyst activity by different
means such as alloying the noble metals or accurate particle size control during the
preparation processes. Some Pt alloys with transition metals showed higher oxygen
reduction activity compared with pure Platinum [70]. Gas diffusion electrodes with
Pt/Co as catalyst are commercially available (E-TEK), showing slightly inferior be-
haviour compared with the platinum gas diffusion electrodes.

3.2.2 The Hydrogen Oxidation Reaction

The hydrogen oxidation is a much faster reaction, and is usually not limiting the
fuel cell performance. Platinum is the optimal catalyst, but nevertheless, loadings of
0.45 mg cm2 are state of the art of commercial available fuel cell electrodes. A fur-
ther reduction and thus a better use of the noble metal loading should be possible
and has been investigated.
The more important challenge is to maintain catalyst activity, when reformate
is used as fuel. Reforming of natural gas, gasoline or diesel leads to a gas mixture
170 A. Heinzel and U. Konig

Fig. 8 Agglomeration of Pt particles on a PtO2 /NafionTM layer after electrochemical treatment


[66]. a EDX mapping of a freshly prepared PtO2 /NafionTM layer, b Pt-EDX mapping of a
PtO2 /NafionTM layer after electrochemical cycling

containing CO2 and CO in addition to the hydrogen being generated. At low temper-
atures, CO is a catalyst poison [71] even in the low concentrations being achieved by
the methods for fine purification of reformate (selective catalytic oxidation or selec-
tive catalytic methanisation of CO). The removal of CO from Pt catalyst sites can
successfully be achieved by oxygen-containing species; most well known for this
purpose is Ru, which is present as RuOx under fuel cell operation conditions [72].
The Pt/Ru ratio and the degree of alloying have been varied to develop electrodes
with optimal CO tolerance. Tin and molybdenum are other possible candidates.
Nanotechnology for Fuel Cells 171

3.2.3 Oxidation of Methanol

The degree of catalyst utilisation is even lower in DMFC; high noble metal loadings
are required according to the state of the art. At the high loadings in the mg/cm2
range, which are stiff state of the art, a carrier material is not necessarily required;
the noble metal coating on the membrane usually has a sufficient thickness and
lateral electronic conductivity. For DMFC also, the reduction of this high amount of
noble metal in the electrodes is a focal point for development.
Various procedures have been investigated to prepare active DMFC anode cata-
lysts. The basis is platinum and ruthenium: platinum is the most electroactive metal,
and ruthenium at least partly covered with hydroxides is used for removal of CO by
transfer of oxygen. The simplest approach is physical mixing of Pt and Ru powders;
improved preparation methods include alloy formation, electrodeposition of Ru on
Pt, PtRu codeposition and adsorption of Ru on Pt. The main factors for catalyst ac-
tivity are alloy formation, surface structure and good dispersion of the comparable
small particles. For the increased noble metal loadings on carbon carrier materials
being usually necessary for good methanol oxidation performance, it is important
to avoid the initiation of agglomeration of the deposited nanoparticles [73].
The reduction of metal salts (PtCl2 , RuCl3 ) in solution by LiBH4 and their sub-
sequent dispersion and stabilisation in THF to form a colloidal solution were used
in [74] to synthesize electrocatalysts for DMFC anodes with a well-defined parti-
cle size of 1.7 nm. By adding the carbon carrier stirring and removing the solvent,
catalyst samples were prepared. With a loading of 3 mg cm2 of noble metal on
the anode side a relatively high activity was achieved (160 mW cm2 at 70 C, dry
oxygen and 2 molar methanol solution).
Sol/gel techniques to prepare modified Vulcan XC-72 electrodes by applying a
platinum sol and a silica sol to the graphite material have recently been reported to
lead to electrodes with a tenfold higher activity towards methanol oxidation [75]. In
a first step, Pt deposits on the carbon particles maintaining their initial size (2 nm);
in a second step, a layer of the catalyst is covered by silica sol, being then transferred
into a gel and finally to aerogel by supercritical processing. The samples show a
BET surface of 731 m2 g1 and a mass normalised current for methanol oxidation
of 60 mA mg1 .
Preliminary investigations have been performed on using C60 -fullerenes as car-
rier material [76]. Electrophoretically deposited C60 nanoclusters were deposited
onto electrically conducting glass sheets, and platinum particles were deposited
by electrical reduction of a solution of H2 PtCl6 . An increased activity of the so-
prepared electrodes compared with platinum was observed, but current density in
general was low due to the low platinum loadings (maximum 100 g cm2 ).
Another approach is the synthesis of Ru-covered Pt nanoparticles and their
fixation to anionic phosphodecatungstate PW12 O3 40 [77]. By this layer-by-layer
preparation method a network film can be realised. A carrier material was alter-
nately immersed into the Pt/Ru solution and into phosphodecatungstate solution.
First electrochemical measurements proved their activity in principle, but the noble
metal loading was too low (typically 0.08 mg cm2 ) for direct comparison with com-
mercial catalysts.
172 A. Heinzel and U. Konig

Because of this diversity in investigated preparation procedures future


improvement of catalytic activity and reduction of noble metal loading for DMFC
anode can be expected to be realised. As higher temperatures are extremely
favourable for DMFC operation, the long-term stability of catalyst nanoparticles
will also be a future issue.

3.3 Optimise Carrier Material

In membrane fuel cells and PAFC as well, usually carbon is used as carrier for the
noble metal catalyst. It fulfils the most important general requirements of chemical
stability and electrical conductivity. One of the first and very commonly used carbon
carrier materials was Vulcan XC-72 (Cabot Corp.) with a BET surface area of about
250 m2 g1 . The effect of platinum loading on the electrochemical activity mainly
for oxygen reduction under PAFC conditions was thoroughly investigated [78]. For
PAFC it was reported [31] that these ungraphitised carbon materials are not stable
on the cathode side and graphitising led to a remarkable loss in surface area of the
carbon. These early investigations already showed that the carbon carrier material
significantly influences the activity of the deposited noble metal particles.
A recent systematic comparison of the suitability of various commercially avail-
able carbon carrier materials as electrode for fuel cells [79] using exactly the same
preparation method for platinum loading led to the result that surface roughness
has a superior positive influence on stability of nanoparticles. The carbon black
Printex XE2 showed to stabilise the precipitated PtOx nanoparticles during the elec-
trochemical reduction process, which was applied to the catalyst material. Four out
of 15 samples showed these advantageous properties. The investigation of carbon
carrier materials from the Sibunit family being prepared by pyrolysis of natural gas
followed by an activation process to achieve the desired surface area and pore vol-
ume was carried out [80]. Carbon materials with 1500 m2 g1 of high purity, high
electrical conductivity and consisting of uniform spherical particles could be pro-
duced. The carbon particles do not contain micro pores: pore diameter ranges from
3100 nm. The goal was to design the optimal carrier material for a DMFC anode.
The catalysts were characterised in a DMFC half cell. In this work it was shown that
the positive effect of a low loading of Pt/Ru catalyst dominates the positive influ-
ence of a high carbon surface area. The utilisation of the noble metal is the highest
with 10% catalyst low loading on low BET surface area carbon (70 m2 g1 ). The
result is explained by the negative influence of small pores <20 nm being present
in high surface area carbon carrier materials on electrochemical activity, either by
diffusion hindrance or by blocking of small pores for example by ionomer due to the
fabrication process of the membrane/electrode assembly. This example shows that
various factors influence the activity of a fuel cell electrode, the BET surface of the
carrier material, the pore volume and size, the distribution of noble metal particle
size as well as the manufacturing process for the MEA.
Nanotechnology for Fuel Cells 173

In new approaches, achieving higher surface areas for improved fuel cell
performance is the goal; carbon nanotubes and carbon aerogels are investigated
for the improvement of electrode properties.

3.3.1 Nanostructured Carrier Material

Another approach to synthesize carbon carrier material with defined structures is


the silica template method, which means carbonizing a polymer silica composite
and removing the silica [81]. Again, the mean pore size could be varied by this
preparation method the mean diameter being either 50 nm, 90 nm or a mixture of
both. Large pores and especially the micro-porous structure with a large surface area
led to a good metal dispersion and a high activity in a DMFC.
Recently, the development of carbon-free materials has been published [82]. The
organic material Perylen red serves as a conductive and stable support, forming a
surface structure with very dense whiskers. The activity of this new electrode is
higher than that of carbon supported electrode and shows even better electrochemi-
cal stability. The structure of this electrode material is shown in Fig. 9.

3.3.2 Nanocomposites

Another option is the replacement of carbon as electronic conducting carrier ma-


terial and the proton-conducting NafionTM in the electrode layer as well by a
conducting polymer, by which an improvement of the interfacial properties was ex-
pected [83]. The manufacture of a Pt/Rupolymer nanocomposite was successfully
carried out and first samples were tested in a DMFC. As electronically conduct-
ing polymers, poly(N-vinyl-carbazole) and poly(9-(4-vinyl-phenyl)-carbazole were

Fig. 9 Whisker-like carrier material


174 A. Heinzel and U. Konig

used. The electrochemical data were slightly lower than with carbon-supported cat-
alysts, but it is expected that an improvement of the electronic conductivity of the
polymer used would lead to even better data.

3.3.3 Nanotubes

Up to now the operating temperature is limited by the perfluorined membranes used


as electrolyte. They can not be used above temperatures T = 120 C, because they
will dry out and then they are not anymore ion-conductive, and on the other hand
they will become gas permeable, because they will be operated above their glass
transition temperature. By increasing the operating temperature the second main
problem of the PEMFC will be solved, i.e. the catalyst poisoning through carbon
monoxide (CO). In case of the DMFC carbon monoxide exists even more, because
it is produced as a by-product during methanol oxidation. Beginning at an operation
temperature of T > 150 C the carbon monoxide will be oxidised by thermal energy
to carbon dioxide and will leave the cell with the gas stream, so that the catalyst
stays active.
The application of carbon nanotubes (CNT) sensitised with metal clusters (Pt,
Ru) opens new possibilities to enhance the fuel cell process [84]. The nanoparticles
deposited on the CNT with a diameter of 2 nm agglomerate on the CNT surface,
forming clusters of about 2040 nm. It is not known yet if the particles will agglom-
erate during the fuel cell process.
XPS measurements show in comparison to conventional Pt/graphite electrodes
an enhanced surface concentration of OH, CO and CO2 Hgroups. Conductivity
measurements and cyclovoltametric investigations (CV) confirm their suitability as
fuel cell electrodes (high electronic conductivity, electrochemical stability against
potentials U > +1 V vs. NHE). The newly developed electrodes exhibit improved
exchange current density in H2 and also in DMFC operation by the factor 510.
It is assumed that the CNT surface groups make the substrate more hydrophilic and
therefore they speed up the transport of the protons from the reaction site. On the
other hand the, CNT surface groups act as a co-catalyst and make the kinetics of the
oxidation reaction more efficient.
One important issue is the use of these materials for hydrogen storage [85, 86].
High hydrogen adsorption capacity was reported for various carbon nanotubes. The
capacity can be significantly increased by doping with e.g. Li or K. Li-doped mate-
rial adsorbs up to 20 wt% and K-doped nanotubes up to 14 wt%, and this depends
on the moisture [87]. Even though the mechanism of hydrogen adsorption is not
clarified yet, the increase in catalytic adsorption is proven. For a review of recent
work, see [88] or [89].
The doping of hydrides with carbon shows an improvement of the hydrogen sorp-
tion kinetics, which confirms the catalytic role of the carbon structure [90].
Nanotubes show also very promising properties in the fuel cell reaction. Kim
et al. used gold nanostructures such as nanotubes or nanoparticles to oxidise CO
[91]. The process is based on the high catalytic activity of gold nanoparticles for CO
Nanotechnology for Fuel Cells 175

oxidation [92]. A reducible polyoxometalate (POM) such as H3 PMo12 O40 serves as


a strong oxidizing agent for CO and as an energy-storage agent for electrons and
protons. The reduced POM can be reoxidised in fuel cell systems that contain simple
carbon anodes.

3.3.4 Electrochemical Deposition of Catalysts

The deposition of catalysts onto the carbon takes place usually by mechanical mix-
ing. A number of catalysts do not have a sufficient contact to the carbon and can not
participate in the reaction. This problem can be overcome by depositing the catalysts
by an electrochemical reaction from adequate precursors. As a consequence, the
phase boundary between electrolyte and collector electrode system will be specif-
ically nanostructured [93, 94]. It was proven that the efficiency of the catalysts in-
creases by the factor 23 up to 100% [95]. The performance of these systems in
PEFC and DMFC systems is at least comparable to that of conventional systems.

3.4 Structure of the Interface

The most sophisticated task is the preparation method to realize a layer with suffi-
cient porosity, hydrophobicity and good access for the reactants to the catalyst parti-
cles. A mixture of the supported catalyst with liquid ionomer in a solvent is usually
the basis forming a so-called ink. The ink is used to coat the membrane by pasting,
screen printing, spraying or similar methods. With NafionTM ionomer solution and
low loadings of 0.120.16 mg Pt/cm2 thin electrodes <10 m were prepared and a
good electrochemical performance was achieved. Starting from this basic method,
many attempts to improve the structure and thus the performance of the catalyst
layer were undertaken. One example is the formation of a colloidal dispersion of
the perfuorosulfonate ionomer [96] FlemionTM in this case cross-linked to Pt-
loaded carbon particles. The performance of an electrode with a lower loading of
0.1 mg/cm2 was found to be optimal due to the use of an optimised carbon car-
rier material and improved preparation process leading to a finer dispersion of the
coagulated ionomer.

4 Bipolar Plates

The materials for bipolar plates have to fulfil several requirements. The most impor-
tant properties are the electrical and thermal conductivity. A review about the present
state is given in [97]. The resistance of a PEM fuel cell usually is dominated by the
membrane electrolyte, and thus the bipolar plate shall not significantly contribute
to this value. Besides the bulk resistance of the material the contact resistance be-
tween bipolar plate and GDL has to be considered. A second important aspect is
176 A. Heinzel and U. Konig

the chemical stability; hydrogen on the anode side as reducing agent and air on the
cathode side causing an oxidising environment combined with the presence of water
and elevated temperatures are the main operation conditions. In principle, corrosion-
resistant metals and carbon composite materials can be used. Metals are the first
choice for high power density applications, as their conductivity is superior to that
of all carbon composites. Graphite is corrosion resistant, but needs at least an im-
pregnation to show the required gas tightness. In addition to that possible impu-
rities of carbon have to be taken into account as catalyst poison. According to the
state of the knowledge, embossed or hydroformed metal plates, hot pressed or injec-
tion moulded carbon composite materials either with duroplastic or thermoplastic
binder materials will be the options.

4.1 Corrosion-Resistant Coatings for Metallic Bipolar Plates

The investigation of use of stainless steel as metallic bipolar plates started quite
late [98]. First results were promising but long-term measurement showed that all
the commonly used stainless steels show severe corrosion damages after several
thousands of hours of operation at elevated temperature. The influence of operation
temperature was shown to be crucial [99]. State of the art is a corrosion-resistant
coating with gold, but as well material cost as cost for the coating process should be
avoided. Physical vapour deposition and chemical vapour deposition have been con-
sidered to be feasible coating processes for metallic bipolar plates. Thus, research
with focus on metallic materials and cheaper coating processes was continued, but
published results are scarce.
It is well known that uncoated stainless steel bipolar plates exhibit high transition
impedances due to formation of an oxide layer by corrosion under fuel cell operation
conditions.
The development of electrical conducting coatings included oxides, carbides, ni-
trides and borides of the metals Cr, Ti, Mo, W, V or Fe [100]. The electrical re-
sistance (voltage drop at the single plates, the air and the fuel plate, respectively)
and the accumulation of metal cations in the MEA during cell operation have been
investigated and taken as a measure for the quality of the coating [101].

4.2 Carbon Composite Bipolar Plates

Thermoplastic carbon composite materials are a favourable material combination


for bipolar plates because they can be manufactured by the mass production pro-
cess of injection moulding [102]. Electrical conductivity of a carbon composite re-
quires a high content of carbon, usually a mixture of graphite and active coal. The
percolation limit of the graphite in the polymer binder has to be exceeded, leading
to direct contact between graphite particles. Additionally, a basic conductivity of
the polymer matrix by the smaller active carbon particles is achieved. The injection
Nanotechnology for Fuel Cells 177

Fig. 10 View of a section through an injection moulded bipolar plate

moulding process leads to a favourable orientation of graphite particles in the direc-


tion of the current flow and to a thin and gas-tight skin, and thus the most important
criteria for use in a fuel cell are fulfilled. A picture of a section through an injection
moulded bipolar plate is given in Fig. 10 [103].

5 Analytical Challenges

To improve the development of nanostructures and their application adequate ana-


lytical methods are required.
Ex-situ methods such as diffraction methods show details about the electronic
properties of nanostructures [104]. Catalysts made of electronically conducting
RuO2 are surrounded by hydrous proton-conducting regions [105], which is
necessary for the high activity of this material as a co-catalyst for CO-tolerant
Pt-RuOx fuel cell electrocatalysts.
Understanding the relationship between the nanoscale structure and electrochem-
ical properties in materials will also lead to the design of other active materials for
electrochemical power sources.
178 A. Heinzel and U. Konig

More complicated are in-situ methods to analyse the electrochemical behaviour


of fuel cell systems since the reactions analysed separately can differ significantly
from those taking place in a real fuel cell system [106]. It is well known that the
electrochemical kinetics are strongly dependent on the crystallographic orientation
of the metal surface [107, 108].
Especially the performance of the metallic catalysts is governed by the orien-
tation. Furthermore, the porosity of the system dominates the transport properties
within the electrolyte [109111].
Conventional electrochemical methods require the use of a reference electrode
to control the potential. Since the thickness of the membrane electrolyte is about
50200 m, the implementation of a Haber-Luggin capillary is complicated. To
avoid as much disturbance of the fuel cell process as possible the diameter of the
capillary has to be less than 1 mm [112]. Even though the disturbance of the water
management and the conductivity of the membrane can be minimised a change in
the distribution of the electrical potential has to be taken into account.
Sufficient results can be obtained from scanning electrochemical experiments.
By the use of the scanning electrochemical microscope (SECM) the electrochem-
ical properties of catalytic materials can be achieved with high spatial resolution
[113]. From the results specific preparation routes for new catalytic systems can be
derived [114].
Other methods like neutron imaging are very complicated. Nevertheless, neutron
imaging is a valuable in-situ method to explore the two-phase flow phenomena in
fuel cell systems with relatively mean interference with the cell [115]. A spatial
resolution in the 1-m scale can be reached [116]. It can be expected that in due
time the resolution reaches the nm scale and important information about the water
distribution during the fuel cell reaction can be obtained.
A direct monitoring of degradation processes of membranes by means of ESR
spectroscopy allows the development of better materials with adequate nanostruc-
tural properties [117].

6 Future Application Power Sources Based on Fuel Cells


for Nanotechnological Applications [118]

Fuel cells are especially promising for long-term application with constant energy
consumption. The number of miniaturised applications especially in the electronic
and medical region will increase even more in the next periods of time, such as
integrated chip-battery/fuel cell systems or active implants for biocontrol. To sup-
ply these applications with constant energy, the fuel cell has to be miniaturised as
well. Nanotechnology will provide the miniaturization of this development with
important aspects.
One important aspect is related to fabrication. The realization of miniaturised
devices is largely due to the use of advanced materials and/or fabrication methods,
however, there is a limit in the scaling of the total system because of the burden of
Nanotechnology for Fuel Cells 179

their auxiliary components. A successful design requires an adequate understanding


of how the performance is influenced by the design and manufacturing process
[119]. Even though theoretical calculations predict an increasing fuel cell power
density increase with decreasing channel size due to the reduced diffusion blockage
of ribs and increased convection in micro channels, the experimental observation
revealed an optimum channel size with maximum power density. Furthermore, the
particle perimeter dependency of Faradic impedance changes to particle area depen-
dency as the particle size increases. This is explained by the generation of cracks in
larger catalyst particles, which serve as triple-phase boundaries.
Miniature fuel cells with a peak power exceeding 40 mW/cm2 [120] have been
designed utilizing a clever flip-flop design and advanced etching and deposition
techniques. The characteristic feature is that the interconnection of electrodes from
two different cells is located on the same side of the membrane realizing a planar
design. Microfuel processors have also been designed for converting methanol into
hydrogen using 5-mm2 -sized reformers and combustors [121]. From these kinds of
micro fuel cell devices, opportunities arise in the field of nanoscale power sources
where the fuel can be delivered directly to the point where the energy is needed.
Micro fuel cell designs without polymeric membranes can overcome some PEM-
related issues such as fuel crossover, anode dry-out or cathode flooding. In these
membraneless laminar flow-based fuel cells (LF-FC) two or more liquid streams
merge into a single microfluidic channel. The stream flows over the anode and the
cathode electrodes placed on opposing side walls within the channel. The reaction
of fuel and oxidant takes place at the electrodes while the two liquid streams and
their liquidliquid interface provide the necessary ionic transport [122, 123].
The advantage of nanoscale fuel cell devices is that a direct electrochemical
reaction can be used for electricity production. The potential fuels might include
reagents in blood or water, and/or metal nanoparticles. One approach has been re-
ported for a biofuel cell operating in a glucose-containing buffer solution [124].
Carbon fibres are coated with catalysts for the selective oxidation of glucose (glu-
cose oxidase) and reduction of oxygen (bilirubin oxidase), which produce 50 nW
per mm of fibre.
The examples described earlier show the high importance of nanotechnology
for the fuel cell devices. Also the molecular application of nanodevices and the
necessary fuel support will finally lead to smart refuelling systems.

References

1. Nanoforum Energy Report, April 2004 (nanoforum.org)


2. VDI-Nachrichten, 2002 (VDI-nachrichten.com)
3. C.H. Steele, A. Heinzel, Nature 414, 2001, 345352
4. Handbook of Fuel Cells, Fundamentals, Technology and Applications, W. Vielstich,
A. Lamm, H. Gasteiger, Eds., Wiley, 2003
5. W. Grot, Chem. Ing. Tech. 47, 1975, MS 260/75
180 A. Heinzel and U. Konig

6. T.A. Zawodzinski, Jr., C. Derouin, S. Radzinski, R.J. Sherman, V.T. Smith, T.E. Springer,
S. Gottesfeld, J. Electrochem. Soc. 140, 1993, 10411047
7. M.S. Wilson, J.A. Valerio, S. Gottesfeld, Electrochim. Acta 40(3), 1995, 355363
8. E. Gulzow, T. Kaz, R. Reissner, H. Sandner, L. Schilling, M.v. Bradke, J. Power Sources 105,
2002, 261266
9. Handbook of Fuel Cells, Fundamentals, Technology and Applications, Vol. 4, W. Vielstich,
A. Lamm, H. Gasteiger, Eds., Wiley, 2003, pp. 930, 947948.
10. A. Muller, Proc. of Third European Solid Oxide Fuel Cell Forum, P. Stevens, Ed., Lucerne,
1998, p. 353
11. D. Herbstsritt, Proc. Fourth European Solid Oxide Fuel Cell Forum, Lucerne 2000, p. 697;
E. Ivers-Tiffee, A. Weber, D. Herbstritt, J. Eur. Ceram. Soc. 21, 2001, 1805
12. M.A. Laguna-Bercero, A. Larrea, R.I. Merino, J.I. Pena, V.M. Orera, J. Am. Ceram. Soc. 11,
2005, 3215
13. J.A. Kolde, B. Bahar, M.S. Wilson, T.A. Zawodzinski, S. Gottesfeld, In: S. Gottesfeld,
G. Halpert, A. Landgrebe, Eds., Proc. First Int. Symp. on Proton Conducting Membrane Fuel
Cells, The Electrochemical Society Proceedings, Advanced Composite Polymer Electrolyte
Fuel Cell Membranes, Vol. 95-23, 1995, 193201
14. B. Bahar, A.R. Hobson, J.A. Kolde, D. Zuckerbrod, US patent 5,547,551, 1996
15. N. Nakao, M. Yoshitake In: Handbook of Fuel Cells, Fundamentals, Technology and Appli-
cations, Vol. 3, Part 1, W. Vielstich, A. Lamm, H. Gasteiger, Eds., Wiley, 2003, pp. 412419
16. Creavis Gesellschaft fur Technologie und Innovation, WO 03/07543 (2008)
17. R. Savinell, E. Yeager, D. Tryk, J. Wainright, D. Weng, K. Lux, M. Litt, C. Rogers,
J. Electrochem. Soc. 141(4), 1994, L46
18. L. Xiao, H. Zhang, E. Scanlon, L.S. Ramanathan, E.-W. Choe, D. Rogers, T. Apple, and
B.C. Benicewicz, Chem. Mater., 17(21), 2005, 53285333; T. Schmidt, 208th Meeting of the
Electrochemical Soc., MA 2005-02, October 1621, 2005, Los Angeles, CA
19. A.E. Steck, C. Stone In: New Materials for Fuel Cells and Modern Battery Systems II,
O. Savadogo, Ed., Ecole Polytechnique, Montreal, 1997
20. F.N. Cornet, G. Geble, R. Mercier, M. Pineri, B. Silion, In: New Materials for Fuel Cells and
Modern Battery Systems II, O. Savadogo, P.R. Roberge, Eds., Ecole Polytechnique Montreal,
Montreal, 1997, p. 818
21. J.A. Kerres, J. Membr. Sci. 185, 2001, 327
22. K.D. Kreuer, J. Membr. Sci. 185, 2001, 2939
23. B. Ruffmann, B. Rohland, Chem. Ing. Tech. 2005, 539548
24. M. Yoshizawa, H. Ohno, Anhydrous proton transport system based on zwitterionic liquid
and HTFSI, Chem. Commun., 2004, 1828
25. Md.A.B. Susan, A. Noda, S. Mitsushima, M. Watanabe, Chem. Commun. 2003, 938939
26. M. Doyle, S.K. Choi, G. Proulx, High-temperature proton conducting membrane based
on perfluorinated ionomer membrane-ionic liquid composites, J. Electrochem. Soc. 147(1),
2000, 34
27. M.-A. Neouze, J. Le Bideau, F. Leroux, A. Vioux, A route to heat resistant solid membranes
with performances of liquid electrolytes, Chem. Commun., 2005, 10821084
28. Z. Li, H. Liu, A. Liu, P. He, J. Li, A room-temperature ionic-liquid-templated proton-
conducting gelationous electrolyte, J. Phys. Chem. B 2004, 108, 1751217518
29. C.A. Angell, W. Xu, J.-P. Belieres, M. Yoshizawa, Ionic liquids and ionic liquids acids with
high temperature stability for fuel cell and other high temperature applications, methods of
making and cell employling same, US Patent WO2004114445, 2004
30. D.J. Jones, J. Rozi`ere In: Handbook of Fuel Cells, Fundamentals, Technology and Applica-
tions, Vol. 3, Part 1, W. Vielstich, A. Lamm, H. Gasteiger, Eds., Wiley, 2003, pp. 447455
31. B. Bonnet, D.J. Jones, J. Rozi`ere, L. Tichicaya, G. Alberti, M. Casciola, L. Masinelli,
B. Bauer, A. Peraio, E. Ramunni, J. New Mater. Electrochem. Syst. 3, 2000, 87
32. F. Bauer, M. Willert-Porada, J. Power Sources 145, 2005, 101107
33. G. Alberti, M. casciola, M. Pica, G. Di Cesare, Ann. N.Y. Acad. Sci 984, 2003, 208225
34. I. Honma, Y. Takeda, J.M. Bae, Solid State Ionics 120, 1999, 255264
Nanotechnology for Fuel Cells 181

35. Y.S. Kim, F. Wang, M. Hickner, T.A. Zawodzinski, J.E. McGrath, J. Membr. Sci. 212, 2003,
263
36. G. Alberti, M. Casciola, M. Pica, Guisi di Cesare, Ann. N Y Acad. Sci. 984, 2003, 2008
37. E. Peled, T. Duvdevani, A. Melman, Electrochem. Solid State Lett. 1(5), 1998, 210211
38. E. Peled, V. Livshits, M. Rakhmann, A. Aharon, T. Duvdevani, M. Pholosoph, T. Feiglin,
Electrochem. Solid State Lett. 7(12), 2004, 507510
39. S. Liu, Q. Pu, L. Gao, C. Korzeniewsky, C. Matzke, Nano Lett. 5(7), 2005, 1389
40. C. Staudt-Bickel, Crosslinked Polymeric Membranes for the Separation of Gaseous and Liq-
uid Mixtures, Soft Materials, Vol. 1, 3, 2003, 277
41. F. Pithan, C. Staudt-Bickel, Crosslinked copolyimide membranes for phenol recovery from
process water by pervaporation, ChemPhysChem, 4(9), 2003, 967
42. J. Huang, T. Jiang, H. Gao, B. Han, Z. Liu, W. Wu, Y. Chang, G. Zhao, Pd nanoparticles
immobilized on molecular sieves by ionic liquids: heterogeneous catalysts for solvent-free
hydrogenation, Angew. Chem., 116, 2004, S1421S1423
43. M. Doyle, S.K. Choi, G. Proulx, J. Electrochem Soc. 147, 2000, 3437
44. Electrochemical Aspects of Ionic Liquids, H. Ohno, Ed., Wiley, 2005
45. M. Yoshizawa*, Wu. Xu, C.A. Angell, J. Am. Chem. Soc., 125, 2003, 1341115419
46. A.P. Abbott,* G. Capper, D.L. Davies, R.K. Rasheed and V. Tambyrajah, Chem. Commun.,
2003, 7071
47. A.P. Abbott, G. Capper, D.L. Davies,* H.L. Munro, R.K. Rasheed and V. Tambyrajah, Chem.
Commun., 2001, 20102011
48. T. Hofener, M. Solinas, W. Leitner, Hydrogen solubility in the system ionic liquid/CO2/H2,
First Congress on Ionic Liquids, Salzburg, 2005
49. S.N.V.K. Aki, D.G. Hert, J.L. Anderson, J.F. Brennecke, Pure and mixed gas solubilities
CO2/IL mixtures, First Congress on Ionic Liquids, Salzburg, 2005
50. J. Kerr, X.-G. Sun, G. Liu, J. Xie, C. Reeder, New polymeric proton conductors for high
temperature applications, DOE Workshop, 2004
51. A.P. Abbott, ChemPhysChem 2004, 5, 1242
52. G. Tsagaropoulos, A. Eisenberg, Macromolecules, 28, 1995, 6067
53. G.S. Fonseca, A.P. Umpierre, P.F.P. Fichtner, S.R. Texeira, J. Dupont, On the use of imi-
dazolium ionic liquids for the formation and stabilization of Ir(0) and Rh(0) nanoparticles:
efficient catalysts for the hydrogenation of arenes. Chem. Eur. J., 9 (14), 2003, 32633269
54. C.W. Scheeren, G. Machado, J. Dupont, P.F.P. Fichtner, S.R. Texeira, Nanoscale Pt(0) parti-
cles prepared in imidazolium room temperature ionic liquids: synthesis from an organometal-
lic precursor, characterization and catalytic properties in hydrogenation reactions, Inorg.
Chem., 42(15), 2003, 47384742
55. M.A. Neouze, J. Le Bideau, A. Vioux, Versatile heat resistant solid electrolytes with perfor-
mances of liquid electrolytes, Prog. Solid State Chem., 24, 2005, 217222
56. P. Devlin, P. Devlin (US DoE), Collaborative Fuel Cell R&D, Second Implementation Liai-
son Committee Meeting, Gunzburg, Germany, March 13, 2004
57. D.D. Macdonald et al., Materials for high temperature PEM Fuel Cells, Workshop at Energy
Institute, Pennsylvania State University, 2003
58. H.A. Gasteiger, M.F. Mathias, Materials for high temperature PEM Fuel Cells, Workshop at
Energy Institute, Pennsylvania State University, 2003
59. T.L. Reitz, Project report AFLR-PR-WP-TR-20042130, US Air Force, November 2004
60. J.J. Xu, Project report 03-19, California Energy Commission, 2003
61. P. Bogdanoff, M. Hilgendorff, H. Schulenburg, M. Fieber-Erdmann, I. Dorbandt, H. Trib-
utsch, S. Fiechter, In: Technical Session: Fuel Cell Systems of the World Renewable Energy
Congress VII: Proceedings (7, 2002, Koln) D. Stolten [u.a.] Eds. Julich: Forschungszentrum
Julich GmbH, 2003 (Schriften des Forschungszentrums Julich Reihe Energietechnik; 26),
pp. 129132
62. P. Stonehart, Ber. Bunsenges. Phys. Chem. 94, 1990, 913921
63. M. Uchida, Y. Fukuoka, Y. Sudawara, N. Eda, A. Otah, A. Otah, J. Electrochem. Soc. 143,
1996, 245
182 A. Heinzel and U. Konig

64. DaimlerChrysler/Stab/, Bonnemann MPI, Stolten FZJ, Wokaun, PSI, 3M. Umicore (USA,
Japan? Fa. Tanaka Kikinzoku in Schweizer Arbeit und Grove)
65. K.-Y. Chan, J. Ding, J. Ren, S. Cheng, K.Y. Tsang, J. Mater. Chem. 14, 2004, 505 (review)
66. N. Fink, Untersuchung zur Kinetik der methanol-oxidation unter DMFC-Bedingungen, PhD
Thesis, Heinrich-Heine-Universitat Dusseldorf, 2005
67. N. Fink, M. Lopez, U. Konig, GDCh Monographien 23, 2002, 367
68. E. Antolini, J. Mater. Sci. 38, 2003, 2995
69. J. Schoonman, Solid State Ionics 157, 2003, 319
70. U.A. Paulus, A. Wokaun, G.G. Scherrer, T.J. Schmidt, V. Stamenkovic, V. Radmilovic,
N.M. Markovic, P.N. Ross, J. Phys. Chem. B, 106, 2002, 41814191
71. U. Stimming, H.F. Oetjen, V.M. Schmidt, F. Trila, J. Electrochem Soc. 143, 1996, 3838
72. T.A. Zawodzinski, C. Karuppaiah, F. Uribe, S. Gottesfeld, Proc. Electrochem. Soc. 97, 1997,
139
73. H. Bonnemann, R. Brinkmann, P. Britz, J. New Mater. Electrochem. Syst. 3, 2000, 199206
74. A. Lee, K. Park, J. Choi, B. Kwon, Y. Sung, J. Electrochem. Soc. 149(10), 2002, A1299
A1304
75. M.L. Anderson, R.M. Stroud, D. Rolison, Nano Lett. 2(3), 2002, 235240
76. K. Vinodgopal, M. Haria, D. Meisel, P. Kamat, Nano Lett. 4(3), 2004, 415418
77. M. Chojak, M. Mascetti, R. Wlodarcyk, R. Marassi, K. Karnicka, K. Mieznikowski,
P.J. Kulesza, J. Solid State Electrochem. 8, 2004, 854860
78. H.R. Kunz, G.A. Gruver, J. Electrochem. Soc. 122, 1975, 1279
79. M.T. Reetz, Chimia 58, 2004, 896
80. V. Rao, P.A. Simonov, E.R. Savinova, G.V. Plaksin, S.V. Cherpanova, G.N. Kryukova,
U. Stimming, J. Power Sources 145, 2005, 178187
81. P. Kim, H. Kim, J.B. Joo, W. Kim, I.K. Song, J. Yi, J. Power Sources 145, 2005, 139146
82. M.K. Debe, Handbook of Fuel Cells Fundamentals and Applications, Vol. 3: Fuel Cell
Technology and Applications, W. Vielstich, H.A. Gasteiger, A. Lamm, Ed., 2003, Wiley, pp.
576589
83. J.-H. Choi, K.-W. Park, H.-K. Lee, Y.-M. Kim, J.-S. Lee, Y.-E. Sung, Electrochim. Acta 48,
2003, 2781
84. Quintus, M, Composite Electrodes and Membranes for Polymer Electrolyte Membrane Fuel
Cells, PhD Thesis, University of Stuttgart, 2002, urn:nbn:de:bsz:93-opus-12074
85. A. Dillon, K.M. Jones, T.A. Bekkedahl, C.H. Kiang, D.S. Bethune, M.J. Heben, Nature 386,
1997, 377
86. A. Chambers, C. Park, R.T.K. Baker, N.M. Rodriguez, J. Phys. Chem 102, 1998, 4253
87. P. Chen, X. Wu, J. Lin, K.L. Tan, Science 285, 1999, 91
88. K.P. de Jong, J.W. Geus, Catal. Rev. Sci. Eng. 42, 2000, 481
89. A.C. Dillon, M.J. Heben, Appl. Phys. A 72, 2001, 133
90. Z. Dehouche, L. Lafi, N. Grimard, J. Goyette, R. Chahine, Nanotechnology 16, 2005, 402
91. W.B. Kim, T. Voitl, G.J. Rodriguez-Rivera, J.A. Dumesic, Science 305, 2004, 1280
92. M. Valden, X. Lai, D.W. Goodman, Science 281, 1998, 1647
93. M.-S. Loffler, B. Gro, H. Natter, R. Hempelmann, T. Krajewski, J. Divisek, Phys. Chem.
Chem. Phys. 3, 2001, 333
94. M.-S. Loffler, B. Gro, H. Natter, R. Hempelmann, T. Krajewski, J. Divisek, Scripta Mater.,
44, 2001, 2253
95. M.-S. Loffler, H. Natter, R. Hempelmann, K. Wippermann, Electrochim. Acta 48, 2003, 3047
96. M. Uchida, Y. Fukuoka, Y. Suawara, H. Ohara, A. Ohta, J. Electrochem. Soc. 145(11), 1998,
37083713
97. A. Hermann, T. Chaudhuri, P. Spagnol, Bipolar plates for PEM fuel cells: a review, Int.
J. Hydrogen Energy 30, 2005, 1297
98. R. Mallant, F. Koene, C. Verhoeve, A. Ruiter, Proc. Fuel Cell Seminar, 1994
99. O. Rau, Dissertation, University Duisburg, Duisburg, Germany, 1999
100. D. Repenning, R. Spah, W. Kaiser, J. Wind, WO01/78175
Nanotechnology for Fuel Cells 183

101. J. Wind, A. LaCroix, S. Braeuininger, P. Hedrich, C. Heller, M. Schudy In: Handbook of Fuel
Cells, Fundamentals, Technology and Applications, W. Vielstich, A. Lamm, H. Gasteiger,
Eds., Wiley, 2003, pp. 294307
102. A. Heinzel, F. Mahlendorf, O. Niemzig, C. Kreuz, J. Power Sources 131, 2004, 3540
103. T. Derieth, G. Bandlamundi, P. Beckhaus, A. Heinzel, C. Kreuz, F. Mahlendorf, J. New
Mater. Electrochem. Syst., 2008, 2129
104. R. Carlin, K. Swider-Lyons, The AMPTIAC Newsletter, 6(1), Spring 2002
105. K.E. Swider-Lyons, K.M. Bussmann, D.L. Griscom, C.T. Love, D.R. Rolison, W. Dmowski,
T. Egami, In: Solid State Ionic Devices II Ceramic Sensors, E.D. Wachsman, et al., Eds.,
Electrochemical Society Proceedings 2000-32, 2000, 48
106. Bernhard Heinrich ANDREAUS, Die Polymer Elektrolyt Brennstoffzelle Charakter-
isierung ausgewahlter Phanomene durch elektrochemische Impedanzspektroskopie, PhD-
Thesis, Ecole Polytechnique Federal Lausanne, 2002
107. U. Konig, B. Davepon, Microstructure of polycrystalline Ti and its microelectrochemical
properties by means of electron-backscattering-diffraction, Electrochim. Acta 47, 2001, 149
108. B. Davepon, J.W. Schultze, U. Konig, C. Rosenkranz, Crystallographic orientation of sin-
gle grains of polycristalline Ti and their influence on electrochemical processes, Surf. Coat.
Technol. 169170, 2003, 85
109. C. Fricke, U. Konig, J.W. Schultze, Untersuchung instationarer Prozesse der O2-Reduktion
an Platin, GDCh-Monographie 12, 1997, 163
110. A.A. El-Shafei, R. Hoyer, L.A. Kibler, D.M. Kolb, Methanol oxidation on Ru-modified pref-
erentially oriented Pt electrodes in acidic medium, J. Electrochem. Soc. 151(6), 2004, F141
111. P.N. Ross, Jr., Oxygen reduction reaction on smooth single crystal electrodes, In: Handbook
of Fuel Cells Fundamentals, Technology and Applications, Vol. 2, Part 5 (The Oxygen Re-
duction/Evolution Reaction), W. Vielstich, A. Lamm, and H.A. Gasteiger, Eds., Chichester,
UK, Wiley, 2003, pp. 465480
112. T. Hamelmann, A. Moehring, M. Pilaski, M.M. Lohrengel, Impedance spectroscopy, in micro
systems. In: P.L Bonora, Ed., Fifth Int. Symp. on Electrochem. Impedance Spectroscopy,
Marilleva, Italy, 2001, pp. 5556
113. K. Eckhard, O. Schluter, V. Hagen, B. Wehner, T. Erichsen, W. Schuhmann and M. Muhler,
Appl. Catal. A: Gen. Catal., 281, 2005, 115120
114. C. Liang, W. Xia, H. Soltani-Ahmadi, O. Schluter, R.A. Fischer and M. Muhler, Chem.
Commun., 2005, 282284
115. D. Kramer, E. Lehmann, G. Frei, P. Vontobel, A. Wokaun, G.G. Scherer, Nuclear Instrum.
Methods Phys. Res. A 542, 2005, 5260
116. N. Kardjilov, S.W. Lee, E. Lehmann, I.C. Lim, C.M. Sim, P. Vontobel, Nuclear Instrum.
Methods Phys. Res. A 542, 2005, 100105
117. A. Panchenko, H. Dilger, E. Moller, T. Sixt, and E. Roduner, J. Power Sources, 127, 2004,
325330
118. A.E. Curtright, P.J. Bouwman, R.C. Wartena, K.E. Swider-Lyons, Power sources for nan-
otechnology, Int. J. Nanotechnol., 1(12), 2004
119. S. Won Cha, R. OHayre, F.B. Prinz, Solid State Ionics 175, 2004, 789795
120. S.J. Lee*, A. Chang-Chien, S.W. Cha, R. OHayre, Y.I. Park, Y. Saito, F.B. Prinz, J. Power
Sources 112, 2002, 410418
121. J.D. Holladay, E.O. Jones, M. Phelps, J. Hu, J. Power Sources 108, 2002, 2127
122. P.J.A. Kenis, R.F. Ismagilov, G.M. Whitesides, Science 285, 1999, 83
123. E.R. Choban, J.S. Spendelow, L. Gancs, A. Wieckowski, P.J.A. Kenis, Electrochim. Acta
50(27), 2005, 5390
124. N. Mano, F. Mao, A. Heller, A miniature biofuel cell operating in a physiological buffer,
J. Am. Chem. Soc. 124(44), 2002, 12962
Vanadium Oxide Aerogels: Enhanced Energy
Storage in Nanostructured Materials

Winny Dong and Bruce Dunn

Abstract Aerogels constitute a unique class of nanostructured materials. They


possess a three-dimensional network of nanometer-sized solid particles within a
continuous mesoporous volume. The materials have low density and high surface
area and only recently have they begun to be explored as electrochemically ac-
tive materials. This chapter reviews the importance of aerogel nanoarchitecture
in achieving high-performance electrochemical properties. Results obtained with
vanadium oxide aerogels are highlighted as these materials exhibit a number of de-
sirable characteristics for secondary lithium batteries.

1 Introduction

Nanostructured materials have emerged as an increasingly important direction for


electrochemical energy storage [1]. This approach represents a significant depar-
ture from more traditional materials involving micron-sized powders. There are in-
dications that the use of nanostructured materials may lead to advantages in both
discharge rates and cycle life of lithium-ion batteries. The reaction of lithium with
simple 3d oxides, sulfides, fluorides, and nitrides leads to in situ formation of nanos-
tructured materials with excellent properties as anodes in lithium cells [2]. Research
on cathode materials has also led to interesting results; decreasing the crystal size
in anatase-type TiO2 extends the solid solution domain and leads to improved re-
versibility because of better accommodation of the structural changes that occur
upon lithium insertion/deinsertion [3].
Another route to nanostructured cathode materials is to tune the morphology
of the solid as occurs in the synthesis of aerogels. These high surface area ma-
terials have a large pore volume and exhibit lithium capacities much higher than

B. Dunn ()
Department of Materials Science and Engineering, University of California, Los Angeles, CA
90095, USA
e-mail: bdunn@ucla.edu

E.R. Leite (ed.), Nanostructured Materials for Electrochemical Energy Production 185
and Storage, Nanostructure Science and Technology, DOI 10.1007/978-0-387-49323-7 5,
c Springer Science+Business Media LLC 2009
186 W. Dong and B. Dunn

polycrystalline, nonporous V2 O5 . The large volume of free space allows for the
easy transport of ions and the large surface area provides an electrochemically
active surface that is not constrained by diffusion limitations. The high porosity
also enables the electrolyte to penetrate the entire cathode or anode structure. This
morphology is especially interesting to researchers working on batteries with high
rate capabilities the ability to deliver a large amount of capacity over a relatively
short period of time. Additionally, the large amount of surface area means that the
electrochemical properties are dominated by surface properties and surface defects
and not the bulk material. Since oxides have inherently defective surfaces, the de-
fect chemistry associated with high vacancy concentrations is likely to influence the
electrochemcial properties. One such surface effect is the large amount of double-
layer capacitance, in addition to the intercalation capabilities, observed for V2 O5
aerogels [4].
In this chapter, we focus on the importance of nanoarchitecture in determining
the electrochemical properties of transition metal oxide aerogels. After a brief intro-
duction to aerogels and their synthesis methods we give an overview of the elec-
trochemical characterization techniques employed when dealing with nanostruc-
tures. This chapter concludes with a case study on how aerogel nanostructures have
improved the electrochemical properties of vanadium pentoxide.

2 Materials Synthesis

2.1 Transition Metal Oxide Aerogels Through SolGel Synthesis

Transition metal compounds are important materials for electrochemistry due to


their ability to exist in various valence states. Several transition metal oxides (MoO3 ,
V2 O5 , MnO2 , etc.) have gained additional attention in the field of secondary lithium
batteries due to their layered structure. These layers can be propped open by in-
tercalated species such as solvated lithium and sodium ions, as well as larger
molecules [5]. The layered structure intercalates lithium while the mixed valence
transition metal centers allow for electron transfer.
Aerogels constitute a unique class of nanostructured materials. Aerogels pos-
sess a three-dimensional network of nanometer-sized solid particles surrounded by
a continuous macroporous (> 50 nm) and mesoporous (250 nm) volume as shown
in Fig. 1. Known for their low density and high surface area, aerogels provide both
molecular accessibility and rapid mass transport and have been part of the heteroge-
neous catalytic materials field for over 50 years [6]. Aerogels processed through
the solgel method have another interesting characteristic: they are often amor-
phous or nanocrystalline. The ability to prepare oxides in this metastable phase
provides another aspect where battery performance can be modified and potentially
enhanced [7].
Vanadium Oxide Aerogels: Enhanced Energy Storage in Nanostructured Materials 187

Fig. 1 Schematic of the


nanoscale solid network of
an aerogel interspersed with
both micro- and mesoporos-
ity (reprinted with permis-
sion from Royal Society of
Chemistry)

Although the main characteristic and advantage of aerogels is its high surface
area, the ability to produce metastable phases is also of importance. Metastable
phases are frequently present due to the low synthesis temperatures involved.
Solgel processing is a room temperature synthesis method in which liquid chem-
ical precursors react to form inorganic ceramic and glass materials [8]. This pro-
cess offers a different approach to the preparation of transition metal oxides which
generally are based on high-temperature heat treatment of powders. Because of
the lower preparation temperatures, unusual amorphous materials or ceramics with
metastable, or kinetically trapped, phases can be produced [9]. Amorphous metal
oxides of vanadium and manganese have demonstrated higher specific capacities
and reversible lithium-ion insertion capacity than do the crystalline forms.
Some of the other advantages of the solgel method include the ability to ob-
tain homogeneous multicomponent systems and better control of the entire synthe-
sis process. The resulting gels can be formed into fibers, films, or composites by
spinning, dip-coating, or impregnation. The synthesis of tailor-made materials is
possible through changing the composition of the precursors, the rate of gelation,
type of catalysts used, and the drying method. Aerogels are produced when the liq-
uid byproducts of the solgel process can be removed without collapsing the solid
network of the wet gel.
The solgel technique has been thoroughly described in review articles and books
[8, 10]. The field is largely based on silica and other insulating oxides such as alu-
mina. It is evident, however, that considerable attention is being given to the transi-
tion metal oxides [11]. Following is a brief discussion on the basic chemistry of the
solgel process.
The solgel process is based on the hydrolysis and condensation of molecular
precursors [11]. The most versatile precursors are the metal organic compounds,
metal alkoxides, M(OR)n . The transition metal is represented by M, n is the va-
lence of the metal, and R is an alkyl group such as methyl, ethyl, propyl, etc.
188 W. Dong and B. Dunn

Transition metal alkoxides are versatile precursors for solgel synthesis because
they are known for almost all transition metal elements [12]. With the notable ex-
ception of silicon, most transition metals are extremely reactive to water, which can
present problems such as the formation of precipitates. Therefore, most transition
metal alkoxides must be handled with great care, in a dry environment, and are often
stabilized by chemical modification to prevent precipitation upon hydrolysis.
Hydrolysis of the alkoxide occurs when water is added and a reactive MOH
hydroxo group is generated [11].

H2 O + MOR MOH + ROH (1)

This reaction follows a nucleophilic substitution mechanism [8]. The rate of reaction
will depend on the coordination unsaturation and oxidation state of the metal atom.
The higher the unsaturation and lower the oxidation state, the faster the substitution
reaction. Condensation occurs as bridging bonds between M and O are formed and
alcohol and/or water are generated.

MOH + MOR M O M + ROH (2)


MOH + MOH M O M + H2 O (3)

After a significant amount of hydrolysis and condensation has taken place, a three-
dimensional network of metal and oxygen forms within the sol (metaloxygen col-
loids suspended in a liquid) and the viscosity of the sol increases. As condensation
continues, the sol transforms into a nonfluid gel and an interconnected and fairly
rigid 3-D network extends throughout the entire sample container. The resulting wet
gel is an amorphous, porous metal oxide with water and alcohol in its mesoscopic
pores. Typically, the solid phase is between 5 and 10% of the total volume.
When the full coordination of the metal atom is not satisfied in the alkoxide,
bridging hydroxo groups can be formed in place of bridging oxygen.

M OH + M ROH M OH M + ROH (4)


M OH + M HOH M OH M + H2 O (5)

Since the coordination for most transition metals is not saturated in the alkoxide,
metal hydroxides can form very quickly, leading eventually to terminal instead of
bridging oxygen bonds, causing precipitates rather than a 3-D network. This reac-
tion must be slowed in order to form a transition metal oxide with bridging oxygen
bonds throughout the structure.
Reaction rates can be controlled through several methods. It has been found that
the rate of hydrolysis decreases with increasing size of the alkyl groups [8, 11].
The larger the alkyl group, the less electronegative the metal atom due to steric
hindrance (shielding of the metal atom by the alkyl group). Solvate formation is
another way of expanding the coordination and slowing down the hydrolysis rate.
Solvate formation, or dilution with a chemically inert solvent can lead to lower
association. Such is the case with forming vanadium pentoxide (V2 O5 ) gels. To
prevent precipitation (by slowing the hydrolysis and olation rates), a very small
Vanadium Oxide Aerogels: Enhanced Energy Storage in Nanostructured Materials 189

amount of water is added, diluted in copious amounts of acetone. Another method


is to chemically modify the metal alkoxides by using additives. Additives such as
solvents, acidic or basic catalysts, or stabilizing agents can react with the alkoxide,
giving rise to a less reactive molecular precursor [8, 10, 11].

2.2 Aerogels

Aerogels are highly porous nanostructured materials with low density. They were
first reported by Kistler in the early thirties [13]. As mentioned previously, these
materials possess a number of interesting and unique properties, one being their
high surface area. The amount of surface area differs from one metal oxide to an-
other and can depend on the synthesis process. However, in virtually every case the
surface area is at least 200 m2 /g and values of >1, 000 m2 /g have been observed.
This characteristic is in sharp contrast to metal oxides prepared through traditional
synthetic methods that have surface areas <10 m2 /g. Another unique property is the
extremely high porosity (80% to >99%) of the aerogel. This results in very low
thermal conductivity and heat capacity. When the metal oxide is an insulator and
the pore diameter is <100 nm, transparent aerogels are obtained. The combination
of high surface area and high porosity has led to aerogel applications in thermal in-
sulation, catalysts and catalyst supports, acoustics, and gas filters [1416]. Although
silica aerogels are the most widely studied system, aerogels can be made from a va-
riety of oxides, cellulose, and even rubber [13].
The synthesis of aerogels is based on the formation of a gel. This two-phase
material consists of solvent molecules trapped in a solid network. Once the wet
gels are formed there are two general methods for removing the solvents. One is
by exposing the wet gel to ambient atmosphere (under ambient pressure) and al-
lowing the solvents to evaporate. The rate of solvent evaporation can be controlled
by varying the temperature and the amount of surface area exposed to the ambient
atmosphere. Capillary forces [(6); Fig. 2] produced by the surface tension of the
solvents will cause some of the wet gel network to collapse, resulting in significant
overall shrinkage.
Capillary pressure = (4L cos )/d. (6)
In (2.6), L is the surface tension of the liquid, d is the pore diameter, and is the
contact angle at the liquidvapor interface. Water at 25 C leads to capillary forces of
around 1,500 atm. Not many systems can sustain this kind of force. After complete
solvent removal, the dry gel will have a porosity of approximately 50% and be
termed a xerogel.
To prevent capillary forces from developing, supercritical drying is used. The
final, supercritically dried gel (i.e., the aerogel) exhibits minimal shrinkage when
drying, preserving the highly porous nanostructure of the wet gel [17]. In the super-
critical drying process the wet gel is placed inside an autoclave. The temperature and
pressure inside the autoclave are raised until the solvent is in its supercritical phase.
190 W. Dong and B. Dunn

Fig. 2 Schematic of a pore of


diameter, d, in a wet gel

vapor

liquid
q

Supercritical drying
Liquid 300 m2 /g
1
Solid
Pressure

2
Ambigel
100 200 m2/g

Freeze-drying Gas
280 m2/g

Temperature
Fig. 3 Methods for synthesizing V2 O5 materials with aerogel-like morphology (reprinted with
permission from The Electrochemical Society)

Above the critical point (Fig. 3), the liquid no longer exists but is in a supercritical
vapor or gaseous phase. Therefore, a liquidvapor interface never develops and no
capillary forces are exerted on the solid network (6). The solvent, in its supercritical
phase, is then vented from the autoclave and the sample returns to ambient pressure
Vanadium Oxide Aerogels: Enhanced Energy Storage in Nanostructured Materials 191

and temperature. Removing the pore liquid without collapsing the microstructure
preserves both the solid and porous networks so that the resulting solid, an aerogel,
retains the high porosity, low density, and high surface area of the wet gel.
It should be appreciated that there are several routes for developing aerogel-like
morphologies for V2 O5 . In addition to supercritical drying, it is possible to use
ambient evaporation methods (ambigels) and freeze-drying techniques to achieve
mesoporous materials with high surface area [18]. These processes are distinctive
because they are based on different methods of solvent removal as shown in Fig. 3.
While the surface areas and densities are slightly different, the electrochemical prop-
erties are all very similar.

3 Characterization Techniques

3.1 Circumventing the Conductivity Problem

3.1.1 Sticky Carbon

The vanadium oxide, molybdenum oxide, and the ruthenium oxide/titanium oxide
systems have all shown that aerogels serve as amplifiers of surface-dominated ef-
fects [1921]. Despite the apparent benefit of high surface area, the electrochem-
ical properties of aerogels received relatively little attention until the mid-1990s.
This reluctance was mainly due to the incompatibility of traditional electrochemical
characterization techniques with high surface area materials. Since transition metal
oxide aerogels do not exhibit very good electronic conduction, it is necessary to
combine aerogels with other components (carbon conductor and binder) to form a
composite electrode structure (Fig. 4). This traditional electrode fabrication route
is inappropriate for aerogels. The aerogel particles are subsequently agglomerated;

Stainless Steel Stainless Steel


Mesh Mesh

Carbon Black Carbon Black Carbon Black


Particles Particles
Transition Metal
Oxide Powder Buried Transition Buried Transition
Metal Oxide Particles Metal Oxide Particles
Solvent

Fig. 4 Traditional preparation route for transition metal oxide electrodes. The aerogel morphology
is not retained
192 W. Dong and B. Dunn

Stainless Steel
Stainless Steel Mesh
Powders are pressed Mesh
onto the wax substrate
Sticky Carbon
Sticky Carbon

Transition Metal Oxide


Particles Transition Metal
Oxide Particles

Fig. 5 Fabrication of the sticky carbon electrode in which the aerogel particles are in contact with
both the electrolyte and the carbon current collector

their surface area is reduced, and their porous structure is altered. In this case, the
liquid electrolyte may not be able to penetrate the pores and access all of the par-
ticles. Thus, the results reported initially for V2 O5 aerogels did not represent the
electrochemical properties of the actual aerogel, but rather the behavior of an ag-
glomerated aerogel system. Frequently, the metal oxide/carbon/binder composite is
mixed in a dispersion solvent to ensure good mixing and minimize agglomeration.
The removal of this dispersion solvent can also cause partial to complete collapse
of the aerogel structure, making quantitative analysis of surface area and porosity
difficult if not impossible.
The sticky carbon electrode described by Long et al. [22] circumvents these lim-
itations (Fig. 5). By mixing acetylene black and wax, it is possible to prepare a con-
ductive electrode that effectively holds the finely dispersed, high surface area parti-
cles, exposes these particles to the electrolyte, and provides electrical contact. These
particles have not been altered and maintain the original aerogel structure and sur-
face area. Using the sticky carbon approach, Long et al. demonstrated that the aero-
gel structure indeed contributed to enhanced electrochemical properties. They were
able to measure the theoretical specific capacitance value for hydrated ruthenium
oxide gel (900 F/g). This was approximately 25% higher than the value measured
using a composite electrode structure. The sticky carbon method was successfully
applied to V2 O5 aerogels and the sweep voltammetry curves were significantly dif-
ferent from those samples prepared with traditional electrodes. As discussed later,
these results suggest a different mode of intercalation in aerogels.

3.1.2 Nanocomposites

Efforts to enhance the conductivity of transition metal oxide electrodes have


included the preparation of composites of the oxide with a conductive material,
such as carbon black. Traditional composite electrodes, however, are characterized
by aggregation of the carbon black particles [23]. These aggregates are typically
on the order of hundreds of nanometers in diameter and may occlude the oxide
aerogel surface. Work to enhance the conductivity of the transition metal oxide
Vanadium Oxide Aerogels: Enhanced Energy Storage in Nanostructured Materials 193

aerogels by replacing the carbon black conducting additive with single wall carbon
nanotubes (SWNTs) has shown great progress [4]. Aerogels for this experiment
are prepared as nanocomposites of vanadium oxide and SWNTs where both com-
ponents are incorporated during the solgel process [24]. The results show that
at high specific currents (2,800 mA/g), where kinetic limitations become a factor,
the V2 O5 /SWNT nanocomposite performed much better (310 mAh/g) than the
traditional V2 O5 /carbon black composite (150 mAh/g) [4].
SWNTs consist of bundles of nanotubes in which the characteristic bundle diam-
eter is in the order of 10 nm. Thus, the SWNTs can provide electronic conduction
in the electrode without blocking the electrolyte access to the aerogel material [24].
Results of the study by Sakamoto et al. show that intimate contact between the V2 O5
aerogel and SWNTs is established as several aerogel fibers are intertwined with sev-
eral SWNTs. This contact was observed between the two phases at the nanodimen-
sional level at multiple points along the ribbons. The macroscopic conductivity of
electrodes prepared with carbon nanotubes as the conductive network is substan-
tially higher (by more than four times) than that of electrodes containing the same
weight fraction of carbon black.

3.2 Deciphering Mechanisms of Charge Storage

3.2.1 EXAFS/XANES

Although it is often assumed that lithium intercalation in amorphous or nanocrys-


talline aerogels occurs between layers of the oxides and results in the reduction of
the transition metal, it is often difficult to correlate Li capacity with the oxidation
state or the local structure of the transition metal. The nanostructure of the aerogels,
along with their often amorphous nature, clearly plays an important role in deter-
mining electrochemical properties. For these reasons, it is highly desirable to be
able to study the local structure and valence of the transition metal in relationship
to Li intercalation. Mansour et al. have demonstrated that this can be accomplished
through X-ray absorption spectroscopy (XAS), especially through the study of ex-
tended X-ray absorption fine structure (EXAFS) and X-ray absorption near edge
structure (XANES) [25]. The XANES region of XAS provides information with
regard to the oxidation state and local symmetry of the absorbing atom. The EX-
AFS region contains information with regard to local structure parameters such as
coordination numbers, bond lengths, and disorder.
In an XAS experiment, photons of uniform energy impinge on a sample and
are absorbed. If the photon energy is high enough, core level electrons may be-
come excited and move into unoccupied states. The level of core electron exci-
tation is measured indirectly through electron decay, giving rise to fluorescence
light or Auger electrons. These experiments can be performed at synchrotron ra-
diation facilities where high-intensity, monochromatic X-ray sources are available.
XAS, with its element specificity and ability to monitor changes in the electronic
194 W. Dong and B. Dunn

and atomic structures without the requirement of long-range order, is uniquely


suited for investigating amorphous and nanophase materials [25]. The nondestruc-
tive nature of XAS also avails itself for monitoring the structural and electronic
changes in the oxide system during Li intercalation. XAS is not only applicable
before and after intercalation, but in situ experiments have also been successfully
demonstrated [26]. In addition to the V2 O5 aerogel studies reported in the follow-
ing paragraph, we previously used the XAS technique to show that nanocrystalline
MoO3 intercalates greater amounts of Li than either the crystalline or the amorphous
forms [20].
In situ XAS studies have established that the oxidation state of V in the V2 O5
aerogel is consistent with the amount of Li inserted (i.e., Lix V2 O5 ). Pentavalent
vanadium is reduced to tetravalent V in the intercalation range 0 < x < 2 and then
tetravalent V is reduced to trivalent V upon insertion of more Li (x > 2) [27]. Al-
though the local structure of the V2 O5 aerogel does not seem to be affected by the
preparation method [28], the long-range order (interplanar distances) and the extent
of reduction of the pentavalent V appear to be influenced by the synthesis parame-
ters and the total surface area of the V2 O5 [26].

3.2.2 FTIR and Spectroelectrochemistry

Fourier transform infrared (FTIR) spectroscopy and spectroelectrochemistry are two


other techniques that have been used with great success at deciphering the charge
storage mechanism in aerogel electrodes. Pre- and postintercalation along with in
situ FTIR experiments have been used in the determination of valence states of the
transition metal and in characterizing local bonding [29].
For in situ data, the cation-insertion process can be monitored through the si-
multaneous current and optical response during the charging and discharging pro-
cess of the cathode. These spectroelectrochemical methods measure the voltammet-
ric response described by the differential capacitance ( Q/ E) and the optical re-
sponse by the change in absorbance ( A/ E) at a given wavelength. This technique
has been referred to as derivative cyclic voltabsorptometry [30, 31]. For an elec-
trochromic process that can be described by the BeerLambert law, the differential
absorbance A can be related to the electrochemical current i by

dA/dt = a z F i , (7)

where is the molar absorptivity, a is the electrode area, z is the number of electrons
produced or consumed, and F is the Faraday constant [32,33]. This method has been
used successfully to investigate cation-insertion reactions in solgel-derived nanos-
tructured vanadium oxide and manganese oxide [34, 35]. For V2 O5 , this technique
is useful because the optical absorbance of V2 O5 is a function of the degree of Li+
insertion and concomitant electron insertion into the oxide [36]. Monitoring the rel-
ative absorbances at specific wavelengths can provide insight into changes of the
local structures during lithium insertion and deinsertion.
Vanadium Oxide Aerogels: Enhanced Energy Storage in Nanostructured Materials 195

4 Case Study: Effects of Nanostructure on the Electrochemical


Properties of V2 O5

A number of nanostructured materials are being investigated as cathodes and anodes


for secondary lithium batteries. Among the aerogel systems that have been exam-
ined for battery applications include SnO2 -based anodes, MnO2 cathodes, MoO3
cathodes, and carbon aerogel anodes [37]. Many of these transition metal oxide gels
retain the electronic properties of the traditionally prepared crystalline counterparts
or show moderate improvements. However, the electrochemical results for vana-
dium oxide aerogels are distinctly different; this system clearly shows the ability of
nanostructured materials to exhibit electrochemical properties that far exceed those
of other amorphous V2 O5 gels that do not have the nanoarchitecture of aerogels [5].
Vanadium oxide belongs to the well-known family of mixed-valence compounds
[38]. The vanadium ions generally exhibit several valence states so that electron
transfer can occur. This particular property has made them the subject of intense
study for a variety of different applications, especially as battery electrodes. One
motivation for synthesizing V2 O5 gels is for the application of cathodes in sec-
ondary lithium batteries. Reversible electrochemical intercalation of Li+ ions into
V2 O5 gels (V2 O5 1.6H2 O) was first reported in 1983 [39]. Since then, considerable
study has been conducted on this subject [40]. Vanadium pentoxide gels can be ei-
ther amorphous or crystalline and they exhibit semiconducting behavior. These ma-
terials possess a characteristic ribbon-like morphology. Under ambient conditions,
the water content of V2 O5 nH2 O gels corresponds to n = 1.8 [41]. This water can
be removed reversibly upon heating, to a composition of V2 O5 0.5H2 O. Below
this water content, the dehydration becomes irreversible, leading to crystalline
V2 O5 . Amorphous V2 O5 has been reported to have specific capacities in excess
of 500 mAh/g with 4 Li/V2 O5 [42]. In addition to battery cathodes, thin films of
V2 O5 have demonstrated electrochromic behavior.
The specific capacity for lithium is greater with V2 O5 aerogels than it is with ei-
ther the corresponding xerogels or crystalline V2 O5 [43]. Comparisons to xerogels
are especially fascinating since the differences between aerogels and xerogels are
morphological (primarily porosity and surface area) and not chemical; the materials
differ only by the manner in which they are dried [6, 44]. Aerogel materials of-
fer considerable promise for battery applications as prior research has shown these
materials to possess a specific lithium capacity in excess of 300 mAh/g at a C/4
discharge rate [42].
The increase in specific capacity for the V2 O5 aerogels is not just related to mor-
phology. While aerogels have shorter diffusion distances and the mesoporous vol-
ume leads to greater accessibility of the ions, these are not the only considerations.
There is an inherent difference in properties due to the nanostructured nature of the
aerogel. In the following paragraphs, we will focus on how nanostructure enhances
the electrochemical properties of V2 O5 aerogels in order to shed some light on
how nanostructures can be used to enhance electrochemical performance for other
materials.
196 W. Dong and B. Dunn

4.1 Pseudocapacity and the Importance of Pore Architecture

Cyclic voltammetry studies with V2 O5 aerogel powders on sticky carbon have


yielded evidence of pseudocapacitance not observed in crystalline V2 O5 nor V2 O5
xerogels [4]. When the high surface area V2 O5 aerogels are completely accessi-
ble to the Li+ ions, evidence of pseudocapacitance is observed. This is the first
time such behavior has been observed for V2 O5 materials and occurs in combi-
nation with the lithium intercalation typically seen for V2 O5 . As shown in Fig. 6,
the Faradaic features are broad and capacitive and the intercalation peaks appear
superimposed on the capacitive response [19]. A specific capacitance of 2,300 F/g
(1, 480 F/cm2 ) and a lithium capacity of 650 mAh/g were reported. Besides the
pseudocapacitance of V2 O5 aerogels, another important observation of this study is
that the capacitance did not vary directly with surface area. Instead, there is a fairly
linear trend between specific capacitance and pore volume. This behavior highlights
the importance of pore architecture and of lithium ion accessibility to the high sur-
face area V2 O5 aerogel. Pore diameters less than 5 nm may not be accessible to the
solvated lithium ions and therefore do not contribute to the overall capacitance of
the sample.
The high surface area and short diffusion length in the solid also contribute to
the ability of V2 O5 aerogels to reversibly intercalate ions with larger radii and/or
valences greater than that of Li+ [45]. Tang et al. have shown that Na+ , K+ ,
Mg2+ , and Ba2+ can all be reversibly inserted into the V2 O5 aerogel structure. The
charge storage mechanism for these ions is most likely related to pseudocapacitive
behavior.

4.2 Effects of Surface Defects

In addition to the large surface areas of V2 O5 gels, surface defects have also shown
to have significant influences on the electrochemical capacity, electronic conductiv-
ity, electrode potential, and ionic transport in metal oxides. Aerogel surfaces tend
to be more defective than those of crystalline materials. The high surface defect
concentration contributes to the amplification of these electrochemical properties.
Work by Swider-Lyons et al. has demonstrated that various temperatureatmosphere
treatments can significantly affect the Li-ion capacities of micrometer-sized V2 O5
polycrystalline powders by controlling the nature of vacancies [46]. Anion vacan-
cies created by heating polycrystalline V2 O5 under Ar (low partial pressure of O2 )
lowered the Li-ion capacity relative to the as-received powder. Heating polycrys-
talline V2 O5 under O2 /H2 O atmosphere formed proton-stabilized cation vacancies
and increased the Li-ion capacity by >20%. Rhodes et al. have demonstrated that
controlled disorder can be created through various atmospheric treatments as a way
to improve electrochemical performance [36].
In the work by Rhodes et al., spectroelectrochemical methods were used to deter-
mine the nature of specific insertion sites within the V2 O5 gels and to monitor how
Vanadium Oxide Aerogels: Enhanced Energy Storage in Nanostructured Materials 197

Fig. 6 Voltammetric sweeps for V2 O5 aerogels using different electrodes. a Sticky-carbon elec-
trode, b traditional composite electrode. The arrows refer to the intercalation/deintercalation peaks
in b (reprinted with permission from the Electrochemical Society)

the distribution of these sites changes due to the different temperatureatmosphere


treatments. The changes in the relative absorbance of the disordered V2 O5 gels at
400 nm suggest that electron insertion occurs into stoichiometric V2 O5 states that
are locally defect-free, while changes in the relative absorbance at 800 nm suggest
that insertion occurs into the vanadiumoxygen structure in the vicinity of vacancies
and defects [36].
198 W. Dong and B. Dunn

5 Conclusions

Aerogels represent a novel nanostructured material for advanced energy storage. As


an earlier publication has indicated, there are both advantages and disadvantages for
using nanomaterials in energy storage [1]. This chapter has shown how vanadium
oxide aerogels represent a very attractive electrode material for secondary lithium
batteries. The aerogels exhibit significantly better electrochemical properties than
either crystalline, nonporous V2 O5 or xerogels of V2 O5 . The reason for this differ-
ence is that there are a number of features that become prominent in aerogels but
are absent with the other forms of V2 O5 . The interconnected porosity allows for
the easy transport of ions and the large surface area provides an electrochemically
active surface that is not constrained by diffusion limitations. The high porosity also
enables the electrolyte to penetrate the entire cathode structure; this accessibility en-
sures high utilization of the aerogel material. A unique surface effect observed with
the V2 O5 aerogels is the large amount of pseudocapacitance, which, in combination
with the excellent lithium intercalation behavior, leads to greater amounts of charge
storage for V2 O5 aerogels. The fact that the V2 O5 aerogels support high discharge
rates makes these materials attractive for electrical actuation systems in micro- and
nanoelectromechanical systems (MEMS and NEMS). As for disadvantages, there
are unresolved questions concerning the cyclability of V2 O5 aerogels. In particular,
structural changes over time are a concern. Finally, there is a trade-off to consider.
Aerogels offer excellent gravimetric capacity, but poor volumetric capacity. Assem-
bling aerogels into hierarchical electrode structures is one direction that offers the
prospect of overcoming this limitation.

Acknowledgments The authors are grateful for the support of their research by the Office of
Naval Research.

References

1. Arico AS, Bruce P, Scrosati B, Tarascon J-M, Van Schalkwijk W, Nat. Mater. 4, 366 (2005)
2. Poizot P, Laruelle S, Grugeon S, Dupont L, Tarascon J-M, Nature 407, 496 (2000)
3. Sudant G, Baudrin E, Larcher D, Tarascon J-M, J. Mater. Chem. 15, 1263 (2005)
4. Dong W, Sakamoto J, Dunn B, Sci. Technol. Adv. Mater., 4, 3 (2003)
5. Schollhorn R, Kuhlmann R, Besenhard JO, Mater. Res. Bull. 11, 83 (1976)
6. Pajonk GM, Catal. Today, 35, 319 (1997)
7. Xu JJ, Kinser AJ, Owens BB, Smyrl WH, Electrochem. Solid State Lett.1, 1 (1998)
8. Brinker CJ, Scherer GW, SolGel Science, Academic Press, New York, 1990
9. Mackenzie JD, J Non-Cryst. Solids, 73, 631 (1985)
10. Hench LL, West JK, Chem. Rev. 90, 33 (1990)
11. Livage J, Henry M, Sanchez C, Prog. Solid State Chem, 18, 341 (1988)
12. Bradley DC, Mehrotra RC, Gaur DP, Metal Alkoxides, Academic Press, London, 1978
13. Kistler SS, Nature 127, 741 (1931)
14. Fricke J, Aerogels, Springer, Berlin, 1986
15. Ayen RJ, Iacobucci PA, Rev. Chem. Eng. 5, 157 (1988)
16. Schneider M, Baiker A, Catal. Rev. Sci. Eng. 37, 515 (1995)
Vanadium Oxide Aerogels: Enhanced Energy Storage in Nanostructured Materials 199

17. Harreld JH, Dong W, Dunn B, Mater. Res. Bull. 33, 561 (1998)
18. Sudant G, Baudrin E, Dunn B, Tarascon J-M, J. Electrochem. Soc. 151, A666 (2004)
19. Dong W, Rolison DR, Dunn B, Electrochem. Solid State Lett. 3, 457 (2000)
20. Dong W, Mansour AN, Dunn B, Solid State Ionics, 144, 31 (2001)
21. Swider KE, Merzbacher CI, Hagans PL, Rolison DR, Chem. Mater. 9, 1248 (1997)
22. Long JW, Swider KE, Merzbacher CI, Rolison DR, Langmuir 15, 780 (1999)
23. Mandal S, Amarilla J, Ibanez J, Rojo JM, J. Electrochem. Soc. 148, A24 (2001)
24. Sakamoto JS, Dunn B, J. Electrochem. Soc. 149, A26 (2002)
25. Mansour AN, Dallek S, Smith PH, Baker WM, J. Electrochem. Soc., 149, A1589 (2002)
26. Mansour AN, Smith PH, Baker WM, Balasubramanian M, McBreen J. J. Electrochem. Soc.
150, A403 (2003)
27. Mansour AN, Smith PH, Baker WM, Balasubramanian M, McBreen J, Electrochim. Acta, 47,
3151 (2002)
28. Stallworth PE, Johnson FS, Greenbaum SG, Passerini S, Flowers J, Smyrl W, Fontanella J,
Solid State Ionics 146, 43 (2002)
29. Kumar PM, Badrinarayanan S, Sastry M, Thin Solid Films 358, 122 (2000)
30. Bancroft EE, Sidwell JS, Blount HN, Anal. Chem. 53, 1390 (1981)
31. Zamponi S, Czerwinski A, Marassi R, J. Electroanal. Chem. 266, 37 (1989)
32. Corodoba de Torresi SI, Electrochim. Acta, 40, 1101 (1995)
33. Chung CY, Wen CT, Gopalan A, Electrochim. Acta, 47, 423 (2001)
34. Long JW, Qadir LR, Stroud RM, Rolison DR, J. Phys. Chem. B, 105, 8712 (2001)
35. Long JW, Young AL, Rolison DR, J. Electrochem. Soc. 150, A1161 (2003)
36. Rhodes CP, Dong W, Long JW, Rolison DR, in Solid State Ionics III, ed. E.D. Wachsman,
et al. PV 2002-26 (Electrochemical Society, Pennington, NJ) pp.478489
37. Long, JW, Dunn B, Rolison DR, White HW, Chem. Rev. 104, 4463 (2004)
38. Robin MB, Day P, Adv. Inorg. Chem. Radiochem., 10, 247 (1967)
39. Araki B, Mailhe C, Baffier N, Livage J, Vedel J, Solid State Ionics 9, 439 (1983)
40. Livage J, Chem. Mater. 3, 578 (1991)
41. Aldebert P, Baffier N, Gharbi N, Livage J, Mater. Res. Bull. 16, 669 (1981)
42. Coustier F, Smyrl WHP, J. Electrochem. Soc., 145, L73 (1998)
43. West K, Zachau-Christiansen B, Jacobsen T, J. Power Sources, 43, 127 (1993)
44. Rolison DR, Dunn B, J. Mater. Chem. 11, 963 (2001)
45. Tang PE, Sakamoto JS, Baudrin E, Dunn B, J. Non-Cryst. Solids, 350, 67 (2004)
46. Swider-Lyons KE, Love CT, Rolison DR, Solid State Ionics 152153, 99 (2002)
Nanostructured Composites: Structure,
Properties, and Applications in Electrochemistry

Joop Schoonman, Sergey Zavyalov, and Alla Pivkina

Abstract Hybrid metal/metal oxidepoly-para-xylylene (PPX) nanocomposites


have attracted great interest, because of a broad spectrum of applications. A simple,
low-cost preparation technique has been developed and comprises a cold-wall vac-
uum co-deposition technique. This co-deposition technique has been applied to
synthesize nanocomposites, containing PPX and nanoparticles of Al, Sn, Zn, Ti and
their oxides. Important is the oxidation kinetics of the metal clusters to their oxides
in relation to the percolation threshold.
The structure, microstructure, and properties of these composite materials have
been studied by X-Ray diffraction, AFM, TEM, and electrical measurements.
Anticipated applications of the present composites are photanodes for hydrogen
production, active lithium-ion rechargeable battery materials, and chemical gas
sensors. An important aspect of these composite materials is the nature of the inter-
facial properties of the nanostructured particles and the PPX material.
The properties of selected composites will be discussed with focus on the nanos-
tructured TiO2 /PPX composites as an anode material for a rechargeable lithium-ion
battery.

1 Introduction

In the past decades there has been a world-wide growing interest in innovative
design, production, storage, and application of nanocomposite materials, i.e., com-
posite materials having spatially organized nanosized components, for advanced in-
novative devices for the conversion and storage for renewable energy sources. Such
composites open up new possibilities for tailoring the fundamental properties of the

J. Schoonman ()
Delft University of Technology, Delft Institute for Sustainable Energy, P.O. Box 5045, 2600 GA
Delft, The Netherlands
e-mail: J.Schoonman@tudelft.nl

E.R. Leite (ed.), Nanostructured Materials for Electrochemical Energy Production 201
and Storage, Nanostructure Science and Technology, DOI 10.1007/978-0-387-49323-7 6,
c Springer Science+Business Media LLC 2009
202 J. Schoonman et al.

materials to reinforce considerably the useful properties of their components, and for
creating unique new properties because, for example, of self-assembling processes
during the synthesis. Examples of such materials are metal (metal oxide)/polymer
nanocomposites, where nanoparticles reveal specific interparticle interactions and
interactions with the matrix in which they are dispersed [1, 2]. Nanostructured
anatase titanium dioxide has attracted widespread attention as a photo-electrode in
an advanced regenerative dye-sensitized solar cell, referred to as the Gratzel solar
cell [3]. It has been shown also that the nanostructured anatase material exhibits an
enhancement factor of about 3 106 compared to the mean lithium-ion intercalation
time of a dense layer of this Li-battery anode material [4].
Nanostructured materials comprising 3-d transition metal oxide nanoparticles
or alloys have been investigated extensively for their potential application as
anode materials in lithium-ion batteries. A serious drawback of such systems
is the substantial volume change of the active phase (up to 300%) during the
charge/discharge process, which leads to mechanical disintegration of the elec-
trode. The use of the polymeric matrix could stabilize the nanoparticles within the
nanocomposite.
A wide range of metal (metal oxide)/polymer nanocomposites has been syn-
thesized, using Al, Sn, Zn, Pd, and Ti as a metal source and poly-para-xylylene
(PPX) as a polymeric matrix. The properties of the nanocomposites were studied
by comparing the structure, morphology, electrical properties, oxidation kinetics,
and electrochemical parameters. Samples of Ti(TiO2 )/PPX nanocomposites were
investigated as anode material for rechargeable Li-ion batteries, and a conventional
two-electrode cell was employed. Addition of components with a high electronic
conductivity could increase the efficiency of the synthesized materials as anode for
rechargeable Li-ion batteries.

2 Experimental Aspects

2.1 Formation of Nanocomposites

The flux of metal atoms in vacuum (Pd, Sn, Al, Ti, Zn), evaporated from a bulk sam-
ple condenses onto a cooled substrate together with the monomer. The condensate
consists of nanoparticles of the metal and the monomer (Fig. 1). Upon heating the
substrate to ambient temperature the monomer polymerises to PPX. The structure
thus obtained is a porous matrix with dispersed nanoparticles in it. The properties
of these nanocomposites containing metal and/or metal-oxide nanoparticles in the
polymeric matrix are presented. Manipulation of the synthesis conditions, i.e., the
distance between the vapour source and the substrate, the tilt angle of the beam, and
the deposition time allowed for optimising the deposition regime. Measuring the
electrical resistance of the condensate and composite permitted the control of the
film formation in relation to the oxidation behaviour.
Nanostructured Composites: Structure, Properties, and Applications in Electrochemistry 203

Fig. 1 Schematics of the cold-wall vacuum co-deposition process

CH2 CH2
400K 930K
2CH2 CH2

CH2 CH2
Para-xylylene
Di-para-xylylene

2CH2 CH2 CH2 CH2 CH2 CH2

Bi-radical

CH2 CH2 CH2 CH2


n
Poly-para-xylylene

Fig. 2 Pathway for pyrolysis of di-para-xylylene to form PPX

2.2 PPX Vacuum Co-Deposition

The schematic of the para-xylylene monomer polymerization [5] is presented in


Fig. 2. The monomer beam was introduced from the source consisting of the zone
of evaporation of di-para-xylylene and its pyrolysis zone. Di-para-xylylene was
introduced into the evaporation zone, which then evaporated (without destruction) in
the temperature range 350400 K. Then the molecules of di-para-xylylene reached
the pyrolysis zone with a temperature of 930 K. Under these conditions the CC
bond shows destruction with almost 100% output of bi-radical. The monomer thus
obtained condenses onto the cooled substrate. With heating up to room temperature,
the condensed monomer polymerises into PPX as indicated in Fig. 2.
204 J. Schoonman et al.

Five types of substrates have been used in the experiments reported here, i.e., (1)
a polished quartz substrate of size 5 mm 5 mm and 1-mm thick with Pt contacts
for electrical measurements, (2) a polished NaCl single-crystalline substrate of the
same size and thickness for TEM analysis, and (3) a polished quartz substrate of
size 10 mm 20 mm and 2-mm thick for optical and AFM investigations. To obtain
identical samples both types of substrates were fixed close to each other onto a
cooled surface of the sample holder (4). Additionally, we used Al-foil and Cu-foil
as substrates for nanocomposites.

2.3 Methods of Nanocomposite Analysis

Oxidation kinetics during air exposure after vacuum synthesis was measured using
the data acquisition board L-1250 connected to a PC. The temperature coefficient of
the electrical resistance in vacuum (the slope of Rv (T )/Rv (293 K) vs. temperature)
was measured after cooling of the synthesized composite from 293 to 77 K.
The morphology of the nanocomposites was studied with Transmission Electron
Microscopy (TEM JEM-2000 EX-II at 200 kV). Samples for TEM were prepared
by standard procedures, including separation of the nanocomposite layer from the
NaCl substrate in water and the film deposition onto a Cu grid for further derailed
investigations. The metal content of the composites was calculated by atomic ab-
sorption analysis using a Perkin-Elmer 503 spectrometer.
The surface morphology, film thickness, lateral forces, and spreading resistance
were studied by AFM (P47-SPM-MDT, Russia, NT-MDT) with silicon cantilevers
having a tip radius less than 10 nm and 20 apex angle (NSC11, Estonia, Mikro-
masch) and conductive cantilevers (silicon coated with TiPt) having a tip radius of
40 nm and 30 apex angle (CSC21, Estonia, Mikromasch). Topographic and spread-
ing resistance AFM images were obtained in tapping and contact modes in air.
The optical spectra of the films were recorded with a spectrophotometer
(Shimadzu UV-3100) in the wavelength range 2002,000 nm.
For electrochemical measurements, a conventional two-electrode cell was
employed using 1.4-cm diameter electrodes. The deposited thin film was the work-
ing electrode and metallic lithium was used as the reference and counter electrode.
The electrolyte consists of 1 MLiClO4 in a 1:2 molar ratio of ethylene carbonate
to propylene carbonate solvent. A porous polyethylene sheet was used as an elec-
trolyte separator. The cells were sealed in a coin cell casing (Hohsen) in an Ar-filled
glove box. Specific capacity and cycling measurements were performed at room
temperature using a Maccor battery test system. The cells were cycled between 0.08
and 2.6 V vs. metallic lithium at a constant current in the range of 0.10.0005 mA.
The typical initial open-circuit voltage for these cells was about 2.9 V.
Nanostructured Composites: Structure, Properties, and Applications in Electrochemistry 205

3 Results and Discussion

3.1 Pd/PPX Nanocomposites

Samples of nanoporous composites of a metal and the polymer PPX have been
synthesized in the form of thin films. The AFM study reveals the metal nanopar-
ticles to have a size of 730 nm. Within the composite the polymer forms more
or less spherical globules with a maximum size of about 200 nm (Fig. 3). The in-
terfacial regions between neighbouring polymeric spherulites contain nanoparticles
of inorganic filler. Depending on the metal content in as-prepared samples, three
different types of surface morphology are distinguished, i.e., isolated nanoparticles
localized on the polymer spherulite surfaces, formation and growth of nanoparticle
chains, and spreading of interconnected chains.
The study of the thin film deposition process shows that the film structure is
strongly dependent on the crystal and energetic properties of the surfaces in con-
tact. As a rule, the epitaxial contact between two phases is the most energetically
favourable, but because of the high deposition rate of the thin films an amorphous
structure is formed instead of an epitaxial layer. With increase of the film thickness
or at higher temperatures, the amorphous phase is converted into a crystalline film.

3.2 Sn (SnO2 )/PPX Nanocomposites

Freshly synthesized tin-containing samples exhibit a grey colour with a metallic


lustre. On contacting with the ambient air during 2 min the samples became trans-
parent for composites with Sn concentrations below the percolation threshold of
10 vol%, while the samples with a tin content beyond 10 vol% do not change their
colour during several months (Fig. 4). Experimental data [6] reveal that for the metal

Fig. 3 AFM image of Pd/PPX nanocomposite: a Phase-contrast image: dark regions are poly-
mer spherulites and light spots are Pd nanoparticles located at the boundaries between polymer
globules, b Cross-section A-A. Profile maximums correspond to Pd particles embedded into the
boundary surfaces between the polymeric globules
206 J. Schoonman et al.

Fig. 4 Sn-containing samples of nanocomposites after air exposure: a sample 1 (8 vol% of Sn), b
sample 3 (12 vol% of Sn)

Table 1 Properties of Sn(SnO2 )/PPX nanocomposites


Sample no. Sn vol% Resistivity in vacuum Morphology

1 8 6 MOhm Isolated particles, localization on


the polymer spherulite surface
2 10 880 Ohm Formation and growth of
3 12 45 Ohm nanoparticle chains
4 16 13 Ohm Chains exhibiting percolation

content in as-prepared samples below or at the percolation threshold (samples 1 and


2, Table 1) the inorganic particles are isolated and the interparticle distance varies
from 5 to 20 nm. Slightly above the percolation threshold (sample 3) the particles
form continuous filaments of varying diameter, but the maximal diameter never ex-
ceeds that of the single metal nanoparticle. Beyond the percolation threshold (sam-
ple 4), the nanoparticles form aggregates located on the boundaries between the
polymer spherulites. Hence, the metalpolymer and metal oxidepolymer nanocom-
posites are the components, wherein the inorganic particles form structured sub-
systems with respect to the polymeric matrix. In increasing the metal content, the
nanoparticles localize along the borders of polymeric spherulites accompanied with
the formation of conducting chain structures. Further increase in the metal content
gives rise to aggregation of nanoparticles and their coalescence.
Analysis of the Sn(SnO2 )/PPX composites reveals that for the metal content in
as-prepared samples below or at the percolation threshold the inorganic particles
Nanostructured Composites: Structure, Properties, and Applications in Electrochemistry 207

Fig. 5 TEM image of sample 1 SnO2 /PPX nanocomposite (8 vol% Sn)

(SnO2 ) are isolated and the interparticle distance varies from 5 to 100 nm (Fig. 5).
Slightly above the percolation threshold the metal particles (Sn) form continuous
filaments of varying diameter, but the maximal diameter never exceeds that of the
single metal nanoparticle. Beyond the percolation threshold, the nanoparticles form
aggregates located on the boundaries between the polymer globules.

3.3 Al(Al2 O3 )/PPX Nanocomposites

TEM analysis of the nanocomposite with an Al content beyond the percolation


threshold reveals spherical pure metal nanoparticles with a mean diameter of about
10 nm (Fig. 6a), while below the percolation threshold the composite contains ag-
glomerates of rhombohedral Al2 O3 (corundum) with a mean size of 55 nm (Fig. 6c).
A sample with a metal content just at the percolation threshold contains metal
nanoparticles of 10 nm and alumina aggregates of 28 nm in diameter (Fig. 6b). The
inorganic phase is homogeneously dispersed within the polymeric matrix in all of
the investigated samples. It has been shown that the nanocomposite structure de-
termines the oxidation behaviour of Al nanoparticles within the polymeric matrix
under air exposure.
Freshly synthesized Al-containing samples (Table 2) exhibit normally a dark
colour with a metallic lustre in vacuum. On contacting the composite with the am-
bient air, sample 7 became transparent, whereas samples 5 and 6 do not change
their colour. Substantial differences in oxidation behaviour exist between the inves-
tigated samples. Figure 7ac shows the resistivity change of freshly synthesized
samples 57, if exposed to air at 1 atm. According to the above TEM results,
nanoparticles in sample 5 are Al crystallites, which are clearly reflected in the minor
increase of the electrical resistance (Fig. 7a) during air exposure (Rmax = 5.5%).
However, the electrical resistivity of sample 7 increases dramatically during several
208 J. Schoonman et al.

Fig. 6 TEM images of Al(Al2 O3 )/PPX nanocomposites with different contents of the inorganic
phase: a sample 5 (12 vol%); b sample 6 (10 vol%), c sample 7 (8 vol%)

Table 2 Parameters of synthesized Al(Al2 O3 )/PPX nanocomposites


Sample number Crystal phase dm (Al) (nm) dm (Al2 O3 ) (nm)

5 Al 10
6 Al + Al2 O3 6 28
7 Al2 O3 55
Nanostructured Composites: Structure, Properties, and Applications in Electrochemistry 209

Fig. 7 Kinetics of oxidation during air exposure at room temperature of Al(Al2 O3 )/PPX nanocom-
posites with different contents of the inorganic phase: a sample 5 (12 vol%), b sample 6 (10 vol%),
c sample 7 (8 vol%)

seconds (Fig. 7c). In fact, nanoparticles in this sample comprise pure alumina dielec-
tric material. The high resistivity of the nanocomposite is caused by the high resis-
tivity of the alumina particles and large distance between them. TEM micrographs
of sample 6 revealed alumina and Al crystallites within the polymeric matrix. The
dramatic increase in resistivity is followed by a sharp decrease after 84 s of air ex-
posure (Fig. 7b).
210 J. Schoonman et al.

3.4 Ti (TiO2 )/PPX Nanocomposites

A series of samples of nanocomposites of Ti and PPX with different Ti content


has been synthesized (Table 3). AFM analysis shows that the inorganic phase com-
prises nanoparticles of 1020 nm in diameter, which are homogeneously distributed
between the polymer globules (Fig. 8).
XRD studies show that synthesized composites do not contain any crystal phase,
just an amorphous phase. Optical absorption measurements prove that synthesized
nanocomposites are containing TiO2 and Ti phases. For comparative analysis the
pure Ti containing thin film was deposited onto the cold substrate (77 K) and onto
the substrate at room temperature. The same result was obtained: XRD analysis
shows that the synthesized films only contain the amorphous phase. Kinetics of the
electrical resistance increase with the air exposure of Ti/PPX nanocomposites (after
synthesis under vacuum) is similar to that of the Al/PPX ones. For a metal content
below the percolation threshold the metal particles became insulator within several
seconds, whereas for the samples beyond the threshold the observed resistance in-
crease is per cents within several hours. DTA analysis revealed that the heating of
amorphous TiO2 nanoparticles up to a temperature of 480 C leads to a phase trans-
formation to anatase, whereas heating up to 580 C results in the anatase transfor-
mation to the rutile structure.

Table 3 Freshly synthesized Ti(TiO2 )/PPX nanocomposites: inorganic phase content


Sample Inorganic phase Inorganic phase
content (vol%)

8 TiO2 7
9 TiO2 9
10 TiO2 10
11 Ti 11
12 Ti 14

Fig. 8 AFM topography images: a sample 12 (14 vol% Ti/PPX), b sample 8 (8 vol% TiO2 /PPX)
Nanostructured Composites: Structure, Properties, and Applications in Electrochemistry 211

Fig. 9 UVvisible spectra of synthesized titanium/polymer nanocomposites: Samples 8, 9, and 10


contain amorphous particles of TiO2 , and samples 11 and 12 contain amorphous Ti

The optical absorption spectra of five different samples are shown in Fig. 9. The
polymeric matrix does not absorb in the range 200600 nm as shown by Zavyalov
et al. [6]. According to data [7], the strongest absorption band of the nanocomposite
samples 810 at 360 nm is assigned to TiO2 particles. The optical spectra of samples
11 and 12 show the non-selective absorption over the entire wavelength span of 220
620 nm, which is typical for free electrons in the Ti metal in the composite films.
Two types of inorganic filler are stabilized by the polymeric matrix, i.e., amorphous
TiO2 in samples 8, 9, and 10, and amorphous Ti in samples 11 and 12 (Table 3).
Simultaneously with topography acquisition under scanning of TiO2 /PPX
nanocomposites, one can see some other characteristics of the investigated sam-
ples. Superposition of topography, lateral forces, and spreading resistance images
allows to understand that the high-conductivity points are localized in between the
polymeric globules (Fig. 10).
Figure 11 shows the electrical resistivity vs. time history for sample 8 with the
metal content below the percolation threshold and for sample 11 above this thresh-
old. The resistivity of sample 8 increases fast, i.e., for the 20 s of oxidation, resis-
tance (R) grows three orders of magnitude, whereas the resistivity of sample 11
increases much slower, i.e., 1.5 times for 20 min. Kinetics of the curve for sample 8
(Fig. 11a) could be approximated by the dependency ln(R)1/ ln(t), and for sample
11 the R(t) dependency reveals R0.5 ln(t) (Fig. 11b).
According to literature data [8, 9] for the early stages of the low-temperature
metal oxidation, when the oxide layer thickness is below 23 nm, the oxide layer
growth depends on the oxygen and metal ion diffusion and on the electron diffusion
towards the reaction surface. The chemical potencial field is formed by the ad-
sorbed oxygen on the oxide surface and by the induced oxygen activity at the metal
metal oxide boundary. With the metal content growth, the conditions of diffusion via
212 J. Schoonman et al.

Fig. 10 AFM of sample 12 (14 vol% TiO2 /PPX): superposition of topography and spreading
resistance images. Scan sizes are a 250 nm 250 nm, b 550 nm 550 nm

a 1600 b 120

110
1200
R, Ohm

R, Ohm

100
800
90
400
80

0 70
0 5 10 15 20 25 0 5 10 15 20 25
t, min t, min

Fig. 11 Kinetics of oxidation of Ti/PPX nanocomposites: a sample 8, experimental data follow to


the function ln(R)1/ ln(t), b sample 11, experimental data follow to the function R0.5 ln(t)
Nanostructured Composites: Structure, Properties, and Applications in Electrochemistry 213

the oxide layer are changing because of the change in the charge transfer mechanism
within the ensemble of metal nanoparticles. As a result, the logarithmic oxidation
law transforms to the inverse logarithmic one.

3.5 Electrical Resistance in Vacuum

The mechanism of the charge transfer processes within an ensemble of ultra-fine


metal particles (in a composite of non-conducting particles) depends on the ratio of
the particles conductivity and the conductivity of the barrier regions between them
and also on the ratio of the particle size and the interparticle distance. Figure 12
presents the temperature coefficient (the slope of R(T )/R(20 C) vs. reciprocal tem-
perature) of the electrical resistivity of as-prepared composites with metal contents
ranging from 7 to 12 vol%. These samples were deposited on quartz substrates with
Pt contacts. The composites with a lower metal content show a semiconductor-like
negative temperature coefficient. This indicates a loss of metal-like contacts be-
tween metal particles. If the metal particle density increases, the temperature coef-
ficient increases. This is typical for porous films comprising islands of conducting
material. Previously, we reported about a sign changing of the temperature coeffi-
cient of metal/polymer nanocomposites once the metal content is close to the per-
colation threshold. The composites with a metal content of 12 vol% reveal a pos-
itive temperature coefficient, indicating an electrical conductivity determined by a
continuous network with metal-like contacts between the metallic nanoparticles. In
contrast, when the metal concentration is 8 vol%, the temperature coefficient be-
comes semiconductor-like. This indicates a loss of metal-like contacts between par-
ticles of the metal phase. This is to be expected from percolation behaviour of the

0,01 Sn
Temperature - co-efficient

0 Cu
2 7 12
0,01 Al

0,02 Ti

0,03 Pd

0,04

0,05
Metal content, Vol.%

Fig. 12 Temperature coefficient of the resistivity of synthesized metal(metal oxide)polymer


nanocomposites vs. metal content
214 J. Schoonman et al.

composites on the metal filler content. The percolation threshold can be determined
by the variation in the temperature dependence of the electrical resistance, which
is for the present case 10 vol% of metal. The synthesized nanocomposites demon-
strate two types of electrical conductivity, i.e., the electrical conductivity in vacuum
is limited by (a) the non-conducting polymer layer, and (b) the conductivity of the
metal nanoparticles.

3.6 Electrochemical Characterization

Figure 13 shows the cell potential as a function of the specific capacity for pure PPX,
10 vol% TiO2 /PPX matrix (sample 10), and 14 vol% TiO2 /PPX matrix (sample 12)
electrodes. From this figure, the performance of the PPX and 10 vol% TiO2 /PPX
matrix electrodes is the same. This indicates that the PPX polymer has some re-
versible capacity and that at 10 vol% TiO2 the active component is still the PPX
polymer matrix. By comparing these potential curves with the potential curve of the
14 vol% TiO2 /PPX film, the curve differs in that the potential for the 14 vol% TiO2
is higher on discharge and lower on charge. Since the reduction and oxidation po-
tential of TiO2 lies at a flat potential of approximately 1.8 V, TiO2 seems to be active
in the 14 vol% TiO2 film. This observation is seen despite the fact that this sample is
thicker than the other films and the current at which this cell was tested was greater
than that of the 10 vol% TiO2 film. As a note, the currents are directly comparable,
since the electrode areas for all cells were the same.
Since the reduction and oxidation potential of TiO2 is flat vs. Li metal at 1.8 V, the
14 vol% TiO2 film was cycled at lower currents in order to observe the intercalation
of TiO2 . In Fig. 14 this potential is shown in the flat part of the potential curve. Al-
though the flat potential is slightly off of what has been reported, the cell resistance
may account for this difference. The reversibility of the 14 vol% TiO2 film seems

Fig. 13 Percolation of TiO2 vol% in the PPX matrix


Nanostructured Composites: Structure, Properties, and Applications in Electrochemistry 215

Fig. 14 Reversibility of 14 vol% TiO2 /PPX nanocomposites at different current densities

Fig. 15 Specific capacity of sample 12 (14 vol%TiO2 /PPX) vs. cycle life

to be good even though the capacity is low for current rates up to 0.28 A/cm2 .
However, for the very low rate of 0.03 A/cm2 , the efficiency decreases as shown
in Fig. 15.

4 Concluding Remarks

Thin-film metal (metal oxide)/polymer nanocomposites with different inorganic


phase contents were obtained by using the cold-wall vacuum co-deposition tech-
nique. A range of metals was shown to be applicable to form nanocomposite thin
films with PPX, i.e., Al, Ti, Pd, and Sn. AFM studies show the metal nanoparticles to
have a size of 750 nm. Within the composite the polymer forms more or less spher-
ical globules with a maximum size of about 200 nm. The interfacial regions between
neighbouring polymeric spherulites contain nanoparticles of the inorganic filler.
The chemical composition, surface morphology, electrical conductivity, opti-
cal absorption, and Li-ion conductivity have been investigated. It is found that in
216 J. Schoonman et al.

a relatively narrow concentration range of the inorganic phase, i.e., from 7 till
14 vol% within the polymeric matrix, the properties of these thin-film nanocom-
posites critically depend on content. Below the percolation threshold, the inorganic
phase consists of nanoparticles of the metal oxide (Al2 O3 , TiO2 , SnO2 ) and beyond
this threshold, of nanoparticles of the metal (Al, Ti, Sn). The different mechanisms
of charge transfer within the composites are found to be responsible for the differ-
ent oxidation behaviours of the metal nanoparticles. We have found experimentally
that the composites with a metal content beyond the percolation threshold show a
positive temperature coefficient indicating a conductivity determined by a continu-
ous network of metal-like contacts between the nanoparticles. In contrast, when the
metal concentration is below the percolation threshold, the temperature coefficient
is negative indicating a conductivity determined by a network of semiconductor-like
contacts. This indicates a loss of metal-like contacts between particles of the metal
phase. The percolation threshold can be determined by the variation of the temper-
ature dependence of the electrical resistance, which is found to be about 10 vol% of
metal. The synthesized nanocomposites demonstrate two types of electrical conduc-
tivity, i.e., the electrical conductivity in vacuum is limited by (a) the non-conducting
polymer layer, and (b) the conductivity of the metal nanoparticles.
It has been shown that the nanocomposite structure determines the oxidation be-
haviour of Al, Sn, and Ti nanoparticles within the polymeric matrix under air expo-
sure. For the metal content below the percolation threshold the metal particles be-
came insulators within several seconds, whereas for the samples above the threshold
the observed resistance growth is several per cent and full oxidation takes several
hours.
The synthesized thin films of nanostructured TiO2 /PPX composites exhibit a high
specific capacity and good cycling performance as anode in Li-ion batteries.

Acknowledgements We are grateful to Dr. Radmir Gaynutdinov (Shubnikov Institute of Crys-


tallography, Russian Academy of Science) for the atomic force microscopy characterization, to
Dr. Dan Simon (Delft University of Technology, The Netherlands) for electrochemical character-
ization, and to The Netherlands Organization for Scientific Research for financial support of the
project (NWO, grant no. 047.011.2003.004).

References

1. M.C. Roco, R.S. Williams, and A.P. Alivisatos, Nanotechnology Research Directions: IWGH
Workshop Report. Vision for Nanotechnology R&D in the Next Decade, Dordrecht, Kluwer
Academic, 2000
2. E.L. Nagaev, Small Metal Particles, Advances of Physical Science (in Russian), 1992 162(9)
49124
3. B. ORegan, M. Gratzel, Nature, 1991 353 737
4. J. Schoonman, Nanostructured Materials in Solid State Ionics, Solid State Ionics, 2000 135
519
5. M. Szwarc, Polym. Eng. Sci., 1976 16 473
6. S. Zavyalov, A. Timofeev, A. Pivkina, and J. Schoonman, Metal-polymer nanocomposites:
formation and properties near the percolation threshold in Nanostructured Materials: Selected
Nanostructured Composites: Structure, Properties, and Applications in Electrochemistry 217

Synthesis Methods, Properties and Applications, editors P. Knauth and J. Schoonman, Boston,
Kluwer Academic, 2002, 92117
7. S. Zavyalov, A. Pivkina, and J. Schoonman, Formation and characterization of metalpolymer
nanostructured composites, Solid State Ionics 2002 147 415419
8. A.T. Fromhold and E.L. Cook, Kinetics of oxide films growth on metal crystals: electron
tunneling and ionic diffusion, Phys. Rev., 1967 158 610612
9. N. Cabrera and N.F. Mott, Theory of oxidation of metals, Rep. Prog. Phys., 19481949 12
163184
Index

A Anisotropic growth, of nanocrystals, 55, 56


Aerogel-like morphology, synthesizing V2 O5 Atoms
materials with, 190, 191 nanoparticles surfaces relating to, 34, 35
Aerogel nanoarchitecture, vi quantum size effects relating to, 3839
Aerogels, 189191. See also Vanadium oxide
aerogels: enhanced energy storage, in B
nanostructured materials Band diagram, of DSC, 4
amorphous nature of, 193 Basic quantum mechanics, 38
carbon aerogel anodes, 195 Batteries, vi, 33, 34, 83
characteristics and advantage of, 185, 187 fuel cells v, 95
electrochemical properties of, 191 Li ion, 59, 93, 96100
as highly porous nanostructured materials, rechargeable, 91
with low density, 186, 189 solid-state, 91
synthesis of, 189 Beer-Lambert law, 194
transition metal oxide aerogels, through Bipolar plates, 175177
sol-gel synthesis, 186189 carbon composite, 176177
AFC. See Alkaline fuel cell membrane fuel cell relating to, 154
AFM studies, 201, 204, 205, 210, 212, 215 metallic, corrosion-resistant coatings for,
Alkaline fuel cell (AFC), 157 176
Al(Al2 O3 )/PPX nanocomposites, 207210, Bottom-up method
211 of synthesis, 56
Alternative energy conversion and storage of wet chemical nanocrystal synthesis, 56
devices, efficiency of, 81 Building blocks, 53, 60
Alternative energy devices, v, vi Bulk electrochemistry, 82, 83
nanoscale materials and electrochemistry Bulk electrolytes, ionic electromigration in,
needed for, 82 108
Alternative, for conventional electrolytes, ionic Bulk junctions, 1
liquids as, 165
Alternative membranes, 159162 C
composite, 161162 Calorimetric measurements, 46
dendrimer PTFE copolymers, 161 Capacitive current, 113
ionic liquid mixtures, 162 Carbon. See Sticky carbon; SWNTs
polymer blends, high-end polymers relating Carbon aerogel anodes, 195
to, 161 Carbon black, 192, 193
Amorphous nature, of aerogels, 193 Carbon composite bipolar plates, 176177
Analytical challenges, of nanotechnology, Carrier material. See Nanostructured carrier
177178 material; Optimise carrier material

219
220 Index

Catalysts. See also Electrocatalysts, in PEMFC Coulomb interaction, 41, 42


and PAFC CPE. See Constant phase element
electrochemical deposition of, 175 Critical nucleus, surface energy and,
fuel cell, 151152 correlation between, 5051
Catalytic studies, of nanoparticles surfaces, Crystallization
3637 phase stability and transformation, of
Cation-insertion process, 194 nanoparticles, relating to, 45, 47, 48
Cells. See also Fuel cells; Solar cells process of, metal oxide nanocrystals, 6266
electrochemical, 81, 94
electrolytic, 94 D
galvanic, 94, 100 Debye temperature, 83
Change storage, deciphering mechanisms of, Deciphering mechanisms, of change storage
193194 EXAFS/XANES, 193194
Characterization techniques, for enhanced FTIR, spectroelectrochemistry and, 194
energy storage, in nanostructured Decomposition method, of synthesis, 57,
materials 5859
circumventing conductivity problem, Dendrimer PTFE copolymers, 161
191193, 197 Density, low, aerogels as highly porous
deciphering mechanisms, of change storage, nanostructured materials with, 186, 189
193194 Device efficiency, problems of, v, 33
Charge carrier collection, 1417 Devices. See also Storage devices
Charge transfer energy, alternative, v, vi, 82
challenges of, 8687 energy conversion, fuel cells as, 152
in nanostructured electrodes, 137138 solid state p-n junction, 1
polarization and, at porous interface, Diffusion-convection, transport by, 110
134135 Diffusion layer
two-step, with adsorbed intermediate, of finite thickness, 122
124126 of infinite thickness, 120
Charge transfer resistance Rt , 120 Diffusion, transport by, 109
Chronoamperometry, 117119 Di-para-xylylene, 203
Circumventing conductivity problem Direct methanol fuel cells (DMFC), 156, 159,
nanocomposites, 192193 171172, 173, 175
sticky carbon, 191192, 197 Distributed features and dispersion, of porous
Colloidal stability, of nanoparticles surfaces, electrode geometry, 135137
3738, 84 DMFC. See Direct methanol fuel cells
Composite membranes, 161162, 163164 Double-layer structure, 111
Composites. See also Carbon composite DSC. See Dye-sensitized solar cell
bipolar plates; Nanocomposites; DSSC. See Dye-sensitized solar cells
Nanostructured composites: struc- Dye, criteria for long-term stability of, 23
ture, properties, applications, in Dye-sensitized solar cell (DSC), 7, 27, 28, 29
electrochemistry advantage of, 3
TiO2 /PPX, 201, 202 band diagram of, 4
Ti(TiO2) /PPX, 210212, 214216 conversion efficiency of, 3
Conductivity problem, circumvention of, long-term testing of, 4
191193, 197 mesoscopic film relating to, 2
Constant phase element (CPE), 136, 137 molecule used by, 3, 5
Controlled hydrolysis, 61, 63 operation of, 14, 18
Conventional electrolytes, ionic liquids as performance of, 21
alternative for, 165 solar light harvesting relating to, 3, 9, 10
Conversion efficiencies, 3, 20. See also IPCEs solid-state, 22
Conversion, of photons, 1014 stability of, 2226
Corrosion-resistant coatings, for metallic criteria for, 23
bipolar plates, 176 kinetic measurements, of solar cells, 24
Coulombic repulsion, 57 recent experimental results on, 2426
Index 221

structure of, 2 nanomaterials, and nanostructures,


Dye-sensitized solar cells (DSSC), 90, 102, introduction to, 8182
103, 128, 138 structure, properties, and applications in,
201217
E Electrodes. See also Porous electrode
Ecole Polytechnique Federale de Lausanne, 2 geometry; Porous electrodes
EDLC. See Electrochemical double-layer definition of, 88
capacitors electroactive, nanomaterials and nanostruc-
Electrical resistance, in vacuum, 213214 tured films as, 8791
Electrical resistivity, time history v, 211, 212 nanostructured, 89
Electric current, light-induced charge sepa- porous effect relating to, 90
ration and conversion of photons to, reactions of, investigating techniques for,
1014 113126
Electric double layer at interfaces, 111113 Electrolyte membranes, general properties of,
Electroactive electrodes, nanomaterials and 159
nanostructured films as, 8790 Electrolytes
Electroactive materials, 88, 90 bulk, ionic electromigration in, 108
Electrocatalysts conventional, ionic liquids as alternative for,
of low- and medium-temperature fuel cells, 165
152 nanomaterials as, 9192
in PEMFC and PAFC Electrolytic cell, 94
general properties of, 168169 Electromigration, ionic, in bulk electrolytes,
interface structure, 175 108
optimise carrier material relating to, Electrons. See also IPCEs; Scanning electron
172175 microscope views
relevant reactions of, 169174
kinetics of, 87
Electrochemical cells, 94
nanoscale materials and electrochemistry
operation of, 81
relating to, 85, 86
Electrochemical characterization, 214215
Electron transfer reaction, at interfaces,
Electrochemical deposition, of catalysts, 175
kinetics of, 106108
Electrochemical double-layer capacitors
Electron transfers, 86
(EDLC), 104, 135
Electrochemical generation, of electricity, 87 Electrostatic stabilization, metal oxide
nanocrystals relating to, 57, 6162
Electrochemical impedance, 119124
Electrochemical measurements, 204 Energy
Electrochemical process, mass transport enhanced storage of, 185200
phenomena involved in, 108111 Fermi, 15, 40, 103
Electrochemical properties free, 93
of V2 O5 , 195198 levels of, quantum size effects relating to,
of aerogels, 191 3840
Electrochemistry. See also Nanoscale storage devices, of sol-gel technology used
materials, electrochemistry and for, vi
basic principles of, 81 surface, critical nucleus and, correlation
bulk, 82, 83 between, 5051
concepts of, 104126 Energy conversion and storage devices,
electrode reactions, investigating alternative, efficiency of, 81
techniques for, 113126 Energy conversion and storage devices,
fundamental, 105113 operation principles of, 9497
definition of, 82 electrochemical double-layer capacitors,
energy conversion and storage in, 9394 104
interfacial, 8283 fuel cells, 100102
nanomaterials, and nanostructures, 81150 lithium ion batteries, 59, 93, 96, 97100,
future prospects for, 138 202
introduction to, 8182 photoelectrochemical solar cells, 102103
222 Index

Energy conversion and storage, in electro- low- and medium-temperature, electrocata-


chemistry, 9394 lysts of, 152
Energy conversion devices, fuel cells as, 152 low-temperature, 100
Energy devices, alternative, v, vi, 82 membrane, 153156
Enhanced energy storage, in nanostructured potential of, 152
materials, 185200 power sources based on, 178179
Enhanced red and near IR responses, by light technology and nanotechnology of, 152158
containment, 9 technology of, nanostructured materials
EQE. See External quantum efficiency impact on, v
Equations Future prospects, for solar cells, 28
evolution, steady-state solution of, 125
Laplace-Youngs, 47, 49 G
linearized nonsteady-state, solution of, 125 G24 Innovations, 26, 28
Maxwells, 42 Galvanic cell, 94, 100
Ostwald-Freundlich, 52 Gas diffusion layer (GDL), 154, 175
phase stability and transformation, of GDL. See Gas diffusion layer
nanoparticles relating to, 4449 Geometry
Schrodingers, 38, 41 lithium ion batteries relating to, 98
Equilibrium, 37 porous electrode, 127138
Evolution equation, steady-state solution of, Growth, nucleation and, 5056, 65
125
EXAFS. See X-ray absorption fine structure H
Experimental aspects, of nanostructured Heteropoly acids (HPA), 163164
composites High-end polymers, polymer blends relating
nanocomposite analysis, methods of, 204 to, 161
nanocomposites, formation of, 202, 203 Highest occupied molecular orbital (HOMO),
PPX vacuum co-deposition, 203204 10, 11, 102
Experimental results, on DSC stability, 2426 HOMO. See Highest occupied molecular
External quantum efficiency (EQE), 10 orbital
HPA. See Heteropoly acids
F Hybrid metal/metal oxide-PPX, 201
Fermi energy, 15, 40, 103 Hydrogen oxidation reduction, 169170
Finite thickness, diffusion layer of, 121 Hydrolysis
First large-scale field tests and commercial of alkoxide, 188
developments, of solar cells, 2628 controlled, 61, 63
Fluorine-doped tin dioxide (FTO), 18, 20 reaction rates of, 188
Fossil fuels, 33
Fourier transform infrared (FTIR) spec- I
troscopy, spectroelectrochemistry and, Imidazoles, 166
194 Impedance, 114, 115
Free energy, 93 electrochemical, 119124
FTIR. See Fourier transform infrared Warburg, 120, 121, 123
spectroscopy, spectroelectrochemistry Incident photon-to-electron conversion
and efficiencies (IPCEs), 3, 10, 17, 19
FTO. See Fluorine-doped tin dioxide Infinite thickness, diffusion layer of, 120
Fuel cell catalysts, drawbacks of, 151152 Interface
Fuel cells, 93, 96. See also AFC; DMFC; electric double layer at, 111113
MCFC; Nanotechnology, for fuel cells; electron transfer reaction at, 106108
PAFC; PEMFC; SOFC; SPEFE porous, polarization and charge transfer at,
batteries v, 96 133135
as efficient energy conversion devices, 152 Interface structure, 175
energy conversion and storage devices Interfacial electrochemistry, 8283
relating, 102 Ionic electromigration, in bulk electrolytes,
high-temperature, 100 108
Index 223

Ionic liquids, 165167 Maxwells equations, 42


as alternative, for conventional electrolytes, MCFC. See Molten carbonate fuel cell
165 Mean particle radius, 45
evaluation of, 165 Melting temperature, nanoparticles surfaces
mixtures, 162 relating to, 35
requirements of, 166 Membrane fuel cell, components of. See also
types of, 165166 PEMFC
use of, 165 bipolar plate, 154
Ionic transport. See Nanoscale electronic, ionic electrode, 154
transport and GDL, 154
Ions, easy transport of, 186 polymer membrane, 154155
IPCEs. See Incident photon-to-electron Membranes. See also Nanostructures
conversion efficiencies composite, 161162, 163164
Mesoporosity, 187
K Mesoporous TiO2 film, 8, 138
Kinetic measurements, of solar cells, 24 Mesoscopic film
Kinetics DSC relating to, 2
of electrons, 86 semiconductor, light harvesting by sensitizer
of electron transfer reaction, at interfaces, monolayer absorbed on, 69
106108 Mesoscopic solar cell variants, development
of oxidation, of Ti/PPX nanocomposites, of, 1
212 Mesoscopic TiO2 layer, 6, 8, 19, 47
Metallic bipolar plates, corrosion-resistant
L
coatings for, 176
Lambert Beers law, 7
Metalorganic chemical vapor deposition
Laplace-Youngs equation, 47, 49
(MOCVD) process, 167
Light containment, enhanced red and near IR
Metal oxide electrodes, transition, 191
responses by, 9
Metal oxide nanocrystals, 5967
Light-induced charge separation and con-
controlled hydrolysis relating to, 61, 63
version of photons, to electric current,
914 crystallization process of, 6266
Li ion batteries. See Lithium ion batteries electrostatic stabilization relating to, 57,
Linearized nonsteady-state equation, solution 6162
of, 125 sol-gel process relating to, 61
Linear sweep voltammetry, 117 synthesis of, 6061, 6466
Lithium, 195 Metal (metal oxide)/polymer nanocomposites,
Lithium ion batteries, 59, 93, 96100, 202 202
geometry relating to, 98 Methanol, oxidation of, 171172
reactions of, 97 Micrometric-scale materials, 83
Lowest unoccupied molecular orbital (LUMO), Microporosity, 187
10, 12, 102 MIE theory, 42
LUMO. See Lowest unoccupied molecular Mixed-valence compounds, 195
orbital MnO2 cathodes, 195
MOCVD process. See Metalorganic chemical
M vapor deposition process
Macrohomogenous concept, 131133 Molecule, DSCs use of, 3, 5
Mass transport phenomena, involved in Molten carbonate fuel cell (MCFC), 158
electrochemical process, 108111 Molybdenum oxide, 191
Materials synthesis MoO2 cathodes, 195
aerogels, 189191
transition metal oxide aerogels, through N
sol-gel synthesis, 186189 N3 dye, 2, 7, 10, 24
vanadium oxide aerogels: enhanced energy NafionTM , 152, 159, 160, 163, 165, 166, 170,
storage, in nanostructured materials, 173, 175
relating to, 186191 Nanoarchitecture, vi, 186
224 Index

Nanocomposites, vi, 173174, 192193 Nanoscale hydrophilic inorganic materials,


Al(Al2 O3 )/PPX, 207210, 211 152
analysis methods of, 204 Nanoscale inorganic or organic semiconduc-
formation of, 202, 203 tors, 1
metal (metal oxide)/polymer, 202 Nanoscale materials, design and properties of,
Sn (SnO2 )/PPX, 205207 3334
Ti/PPX, 212 Nanoscale materials, electrochemistry and
Nanocrystalline metal oxide semiconductors, alternative energy devices dependent on, 82
5960 charge transfer challenges relating to, 8687
Nanocrystalline solar cells, band diagram and electrons relating to, 85, 86
operational principle of, 45 size effects of, 8286
Nanocrystalline TiO2 layers, 7, 18, 59 Nanoscale materials, recent applications of:
Nanocrystals. See also Metal oxide nanocrys- solar cells, 132
tals; Transition metal nanocrystals, charge carrier collection, 1417
synthesis of DSC
anisotropic growth of, 55, 56 photovoltaic performance of, 18
SnO2 , 43, 44, 47, 54, 59, 63, 64, 65 stability of, 2226
Nanomaterials dye, criteria for long-term stability of, 23
electrochemistry, nanostructures and, enhanced red and near IR responses, by light
81150 containment, 9
as electrolytes, 9192 first large-scale field tests and commercial
nanostructured films and, as electroactive developments of, 2628
electrodes, 8791 future prospects for, 28
Nanometric-scale materials, physicochemical introduction to, 14
properties of, 84 kinetic measurements of, 24
Nanoparticles, v light-induced charge separation and
assembly and properties of, 3380 conversion of photons to electric current,
introduction to, 3334 914
metal oxide nanocrystals, 5967 mesoscopic semiconductor film, light
nanoparticles surfaces, 3438 harvesting by sensitizer monolayer
nucleation, growth and, 5056 absorbed on, 69
phase stability and transformation, 4449 nanocrystalline solar cells, band diagram
quantum size effects, 3844 and operational principle of, 45
synthetic methods, 4966 nanostructure, importance of, 56
transition metal nanocrystals, synthesis of, new sensitizers and redox systems,
5659 development of, 2122
for improved membrane properties - open circuit photovoltage, increase in, 21
composite membranes, 163164 overall conversion efficiency under global
production of, synthesization methods for, AM 1.5 standard reporting condition, 20
50 photocurrent action spectra, 1920
SnO2 , 54 quantum dot sensitizers, 1718
sodium, melting of, 36 solid-state DSCs, 22
Nanoparticles surfaces Nanoscience, nanotechnology and, 81
atoms relating to, 34, 35 Nanostructured anatase titanium dioxide, 202
catalytic studies of, 3637 Nanostructured carrier material, 173
characteristics of, 3435 Nanostructured cathode materials, 185
colloidal stability of, 3738, 84 Nanostructured composites: structure, prop-
melting temperature relating to, 35 erties, applications, in electrochemistry,
size of, 3435 201217
Nanoscale architectures, strategies for, 81 experimental aspects of, 202204
Nanoscale effect, examples of, 8384 introduction to, 201202
Nanoscale electronic, ionic transport and results and discussion about, 205215
across interfaces, 92 Nanostructured electrodes, 89
along interfaces, 92 charge transfer in, 137138
Index 225

Nanostructured films, nanomaterials and, as nanostructures, 159167


electroactive electrodes, 8791 relevance of, 151152
Nanostructured materials, 202 nanoscience and, 81
aerogels as, 186, 189 Nanotubes, 174175
alternative energy devices improved by, v, vi Network junctions, interpenetrating, 2, 5
enhanced energy storage in, 185200 New sensitizers and redox systems,
fuel cell technology impacted by, v development of, 2122
new technologies developed with, v Nonsteady-state techniques, 113, 114
Nanostructured membranes, 164165 Nucleation, growth and, 5056, 65
Nanostructure, electrochemical properties of
V2 O5 impacted by, 195198 O
pseudocapacity and importance of pore 1-D nanostructures, 99
architecture, 196 Open circuit photovoltage, increase in, 21
surface defects, effects of, 196197 Optimise carrier material, 172175
Nanostructures, 56, 8889 electrochemical deposition, of catalysts, 175
0-D, 99 nanocomposites, vi, 173174
1-D, 99 nanostructured, 173
2-D, 99 nanotubes, 174175
3-D, 99 Ostwald-Freundlich equation, 52
alternative membranes, 159162 Ostwald ripening model, 52, 53, 56
charge carrier collection, 1417 Overall conversion efficiency, under global
electrochemistry, nanomaterials, and, AM 1.5 standard reporting condition, 20
81150 Oxidation
electrolyte membranes, general properties of methanol, 171172
of, 159 of Ti/PPX nanocomposites, kinetics of, 212
enhanced red and near IR responses, by light Oxidation-reduction reactions, 86
containment, 9 Oxygen reduction reaction, 169
for fuel cells, 159167
ionic liquids, 165167 P
light-induced charge separation and PAFC. See Phosphoric acid fuel cell
conversion of photons, to electric Particle radius, 53
current, 914 PBM. See Purpose-built materials
mesoscopic semiconductor film, light Pd/PPX nanocomposites, 205
harvesting by sensitizer monolayer PEMFC. See Polymer-electrolyte-membrane
absorbed on, 69 fuel cell
nanoparticles, for improved membrane Percolation, of TiO2 vol%, in PPX matrix, 214,
properties - composite membranes, 215
163164 Perkin-Elmer 503 spectrometer, 204
nanostructured membranes, 164165 Phase stability and transformation, of
quantum dot sensitizers, 1718 nanoparticles, 4449
Nanostructuring, of electroactive materials, 90 crystallization relating to, 45, 47, 48
Nanotechnological applications, power sources equations relating to, 4549
based on fuel cells for, 178179 model for, 4748
Nanotechnology Phosphoric acid fuel cell (PAFC), 158,
basic principles of, v 168175
definition of, 83, 84 Photocurrent action spectra, 1920
emerging field of, 85 Photoelectrochemical solar cells, 102103
for fuel cells, 152158 Photovoltage, open circuit, increase in, 21
analytical challenges of, 177178 Photovoltaic cells, 1
bipolar plates, 175177 Photovoltaic performance, of DSC, 18
electrocatalysts, in PEMFC and PAFC, open circuit photovoltage, increase in, 21
168175 overall conversion efficiency, under global
future application of, 178179 AM 1.5 standard reporting condition, 20
introduction to, 151158 photocurrent action spectra, 1920
226 Index

Physicochemical properties, of nanometric- R


scale materials, 84 Reaction rates, of hydrolysis, 188
Polarization and charge transfer, at porous Reactions. See also Kinetics
interface, 134135 of electrocatalysts, in PEMFC and PAFC,
Polymer blends, high-end polymers relating to, 169174
161 of electrodes, investigating techniques for,
Polymer-electrolyte-membrane fuel cell 113126
(PEMFC), 101, 153, 155, 158, 159 electron transfer, at interfaces, 106108
electrocatalysts in, 168175 of lithium ion batteries, 97
Polymer membrane, 154155 oxidation-reduction, 86
Polyol process, of synthesis, 58 oxygen reduction, 86, 169
Poly-para-xylylene (PPX), 201, 202, 203, 210, precipitation, 50
214, 215 Rechargeable batteries, 91
Pore architecture, pseudocapacity and Ruthenium oxide/titanium oxide, 191
importance of, 196
Pore size distributions (PSD), 136 S
Porous effect, electrodes relating to, 90 Salt reduction method, of synthesis, 5758, 59
Porous electrode geometry, 127138 Scanning electron microscope views, 8
distributed features and dispersion, 135137 Schrodingers equation, 38, 41
macrohomogenous concept of, 131133 Semiconductor film, mesoscopic, 69
nanostructured electrodes, charge transfer Semiconductor quantum dots (QDs), 17
in, 137138 Semiconductors
porous interface, polarization and charge nanocrystalline metal oxide, 5960
transfer at, 134, 135 nanoscale inorganic or organic, 1
solid and electrolyte phases, transport in, quantum size effects relating to, 40
133 Sensitizer. See DSC
transmission line description of, 128131 Silica, 152
Porous electrodes Single wall carbon nanotubes (SWNTs), 193
theory of, 127128 SnO2 -based anodes, 195
transmission line description of, 128131 SnO2 colloidal suspensions, 55
Porous interface, polarization and charge SnO2 nanocrystals, 43, 44, 47, 54, 59, 63, 64,
transfer at, 134, 135 65
Power sources based on fuel cells, for SnO2 nanoparticle, 54
nanotechnological applications, 178179 SnO2 nanoribbon, 54
PPX. See Poly-para-xylylene Sn (SnO2 )/PPX nanocomposites, 205207
PPX matrix, percolation, of TiO2 vol%, in, analysis of, 206207
214, 215 properties of, 206
PPX vacuum co-deposition, 203204 Sodium nanoparticles, melting of, 36
Precipitation reactions, 50 SOFC. See Solid-oxide fuel cells
PSD. See Pore size distributions Solar cells. See also DSC; Mesoscopic solar
Pseudocapacity and importance of pore cell variants, development of; Nanoscale
architecture, 196 materials, recent applications of: solar
Purpose-built materials (PBM), 56 cells
photoelectrochemical, 102103
Solar light harvesting, DSC relating to, 3, 9, 10
Q Sol-gel process, metal oxide nanocrystals
QDs. See Semiconductor quantum dots relating to, 61
Quantum dots, 84 Sol-gel synthesis, transition metal oxide
Quantum dot sensitizers, 1718 aerogels through, 186189
Quantum size effects, 3844 Sol-gel technology, energy storage devices
atoms relating to, 3839 use of, vi
energy levels relating to, 3840 Solid and electrolyte phases, transport in, 133
semiconductors relating to, 40 Solid-oxide fuel cells (SOFC), 101, 158
Index 227

Solid-polymer-electrolyte fuel cell (SPEFE), TiO2 particles, 8


101 TiO2 polymorphs, 44, 45
Solid-state batteries, 91 TiO2 /PPX composites, 201, 202
Solid-state DSCs, 22 TiO2 surface, 10, 11, 12, 21
Solid state p-n junction devices, 1 TiO2 vol%, in PPX matrix, percolation of, 214,
Solution, of linearized nonsteady-state 215
equation, 125 Ti(TiO2) /PPX composites, 211213, 214, 216
Space-charge region, 92 Ti/PPX nanocomposites, 210, 212
Spectroelectrochemistry, FTIR and, 194 Toyota dream house, 27
Spectrometer, Perkin-Elmer 503, 204 Transition metal alkoxides, 188
Spectroscopy. See FTIR; XAS Transition metal nanocrystals, synthesis of
SPEFE. See Solid-polymer-electrolyte fuel cell bottom-up methods of, 56
Stability. See also Phase stability and decomposition method of, 57, 5859
transformation, of nanoparticles electrostatic stabilization relating to, 57
colloidal, of nanoparticles surfaces, 3738, polyol process relating to, 58
84 salt reduction method of, 5758, 59
of DSC, 2226 Transition metal oxide aerogels, through
Steady-state solution, of evolution equation, sol-gel synthesis, 186189
124 Transition metal oxide electrodes, 191
Sticky carbon, 191192, 197 Transmission Electron Microscopy (TEM),
V2 O5 relating to, 192 201, 204, 207, 208, 209
Storage devices Transmission line description, of porous
energy conversion and electrodes, 128131
alternative, efficiency of, 81 Transport. See also Mass transport phenomena,
fuel cells relating to, 102103 involved in electrochemical process
energy conversion and, operation principles by diffusion, 108
of, 59, 92, 93103, 202 by diffusion-convection, 108
of sol-gel technology, vi across interfaces, 92
Surface and transformation enthalpies, 46 along interfaces, 92
Surface defects, effects of, 196197 of ions, 186
Surface energy, critical nucleus and, correlation in solid and electrolyte phases, 133
between, 5051 2-D nanostructures, 89, 90
Surfaces. See Nanoparticles surfaces; TiO2 Two-step charge transfer, with adsorbed
SWNTs. See Single wall carbon nanotubes intermediate, 124126
Synthesis. See also Materials synthesis
of metal oxide nanocrystals, 6061, 6466
of tailor-made materials, 187 U
of transition metal nanocrystals, 5659 UV-visible spectra, of synthesized ti-
Synthesization methods, for nanoparticle tanium/polymer nanocomposites,
production, 50 211
Synthetic methods, for nanoparticles, 4966
metal oxide nanocrystals, 5966 V
nucleation, growth and, 5056 V2 O5 , 186, 194
aerogel-like morphology for, 190, 191
T electrochemical properties of, 195198
Tailor-made materials, synthesis of, 187 sticky carbon relating to, 192
Technology and nanotechnology, of fuel cells, V2 O5 /SWNT, 193
152158 Vanadium oxide aerogels: enhanced energy
TEM. See Transmission Electron Microscopy storage, in nanostructured materials,
Testing, long-term, of DSC, 4 185200
3-D nanostructures, 99, 139 characterization techniques for, 191194,
Ti(IV) ions, 10, 11, 12 197
Time history, electrical resistivity v, 211, 212 introduction to, 185186
TiO2 conduction band, 13, 24 materials synthesis relating to, 186191
228 Index

Vanadium oxide aerogels (Continued) X


nanostructures impact, on electrochemical XANES. See X-ray absorption near edge
properties of V2 O5 , 195198 structure
Vanadium pentoxide, 186 XAS. See X-ray absorption spectroscopy
X-ray absorption fine structure (EXAFS),
W 193194
Warburg diffusion resistance, 15 X-ray absorption near edge structure
Warburg impedance, 120, 121, 123 (XANES), 193194
Warburg straight line, 124, 137, 138 X-ray absorption spectroscopy (XAS), 193,
Wet chemical nanocrystal synthesis, bottom-up 194
methods of, 56 0-D nanostructures, 95
Color Plates

Fig. 1.1 Structure of the N3 dye cis RuL2 (SCN)2 (L = 2,2-bipyridyl-4,4 dicarbo-xylate)
Conducting
glass TiO2 Dye Electrolyte Cathode

Injection
S*
0.5
Maximum
Voltage
0 h
E vs
NHE Red Mediator Ox
(V) 0.5
Diffusion

1.0
S /S+
e- e-

Fig. 1.2 Energy band diagram of the DSC. Light absorption by the dye (S) produces an excited
state (S ) that injects an electron into the conduction band of a wide band gap semiconducting
oxide, such as TiO2 . The electrons diffuse across the oxide to the transparent current collector
made of conducting glass. From there they pass through the external circuit performing electrical
work and re-enter the cell through the back contact (cathode) by reducing a redox mediator (ox).
The reduced form of the mediator (red) regenerates the sensitizer closing the cyclic conversion of
light to electricity

Fig. 1.4 Uptake of N3 dye by a nanocrystalline TiO2 film, which is immersed in the dye solution.
The resulting deeply red-colored film is the photoactive part of the DSC
Fig. 1.7 Scanning electron micrograph showing anatase crystals of ca. 400 nm size, employed as
light scattering centres to enhance the red response of the DSC (courtesy of Dr. Tsuguo Koyanagi,
Catalysts & Chemicals Ind. Co. Ltd., Japan)

Side view Top view

Fig. 1.8 Side and top view of the RuL2 (NCS)2 (N3) sensitizer anchored to the (101) TiO2 anatase
surface through coordinative binding of two carboxyl groups to surface titanium ions. The green
and red spheres present titanium and oxygen, respectively. Note that the left carboxylate group
straddles two Ti(IV) surface ions from adjacent surface titanium rows while the right one forms an
ester bond. The structure shown represents the lowest energy configuration derived from molecular
dynamics calculations and the area occupied by one adsorbed N3 molecule being 1.64 nm2
Dye Dye Dye Dye Dye Dye Dye
Dye Dye Dye Dye Dye
Dye Dye Dye Dye
RS rtrans rtrans rtrans rtrans
Dye

Dye
Dye
rct rct rct Dye
Dye
RFTO/EL cch cch Dye
cch RCE
Dye Dye

CFTO/EL Electrolyte Zd

Dye Dye
CCE
Dye Dye Dye Dye Dye Dye Dye Dye Dye
Dye Dye
Dye Dye
Dye Dye

Dye

Dye
Dye
Dye Dye
Dye Dye Dye
Dye Dye Dye Dye Dye Dye Dye
Dye Dye Dye Dye

Fig. 1.12 Equivalent electric circuit diagram of a solar cell based on a nanocrystaline semiconduc-
tor film in contact with an electrolyte. Two transmission lines are used to model the motion of the
conduction band electron motion through a network of mesoscopic semiconductor particles and the
charge compensating flow of redox electrolyte. The electrical equivalent circuit treats each particle
as a resistive element. Interfacial electron transfer from the conduction band of the nanoparticle to
the triodide is modelled by a charge transfer resistance rct connected in parallel with their chemical
capacitance cch . The latter is defined as the electric charge (measured in Coulomb) that is required
to move the Fermi level of the of the semiconductor nanoparticles by 1 eV. Zd is the Warburg dif-
fusion resistance describing the motion of triiodide ions through the porous network to the counter
electrode while RCE and CCE are the charge transfer resistance for the reduction of triodiodide and
the double-layer capacitance of the counter electrode, respectively The red dots present cations
from the electrolyte
e
+ *
3 S /S


1 2
Oxidation
e
Potential
5

+ 4 Red/Ox
TiO2
S+/S Couple

Fig. 1.13 Photo-induced processes occurring during photovoltaic energy conversion at the surface
of the nanocrystalline titania films. 1: sensitizer (S) excitation by light, 2: radiative and nonradiative
deactivation of the sensitizer, 3: electron injection in the conduction band followed by electron
trapping and diffusion to the particle surface, 4: recapture of the conduction band electron by
the oxidized sensitizer (S+ ), 5: recombination of the conduction band electrons with the oxidized
form of the redox couple regenerating the sensitizer and transporting the positive charge to the
counterelectrode. Grey spheres: titania nanoparticles, red dots: sensitizer, green and blue dots:
oxidized and reduced form of the redox couple

Fig. 1.14 Cross-sectional view of the embodiment of DSC used in the laboratory for photovoltaic
performance measurements
COOH
O

HOOC
N
N N
Ru
N N
C N
S
C
S O

Scheme 1.1 K-19 sensitizer with an extended p-system in one of its ligands

Light

Conduction
Band
I
S S* e
e e e e

1
2 I2 S+

Semiconducting
e Membrane

Electrical Work

Fig. 1.17 The two coupled redox cycles involved in the generation of electricity from light in a
dye-sensitized solar cell
Fig. 1.19 Production of DSC prototypes by Aisin Seiki in Japan. Note the monolithic design of the
PV modules and the use of carbon as interconnect and counter electrode material. The red dye is
related to N-719 while the black dye has the structure RuL (NCS)3 where L = 2, 2 , 2 -terpyridyl-
4,4, 4 tricarboxylic acid

The Toyota Dream House

DSC
made by
AISIN -SEIKI

Fig. 1.20 The Toyota Dream House featuring DSC panels made by Aisin Seiki. For details see
web announcement http://www.toyota.co.jp/jp/news/04/Dec/nt04 1204.html
Fig. 1.21 First commercial flexible lightweight cell produced by G24 Innovation on a large scale
for us as telephone chargers

CHARGE

Power
supply

e
Load

O
DISCHARGE
Co

Li
ELECTROLYTE

Carbon

CATHODE ANODE
Li1-xCoO2 Graphite

Fig. 3.3 Schematic of a rechargeable lithium battery in discharge/charge mode. The lithium ion is
intercalated in the anode during charging and in the cathodes during discharging. The layered host
lattice in the cathode and anode is also illustrated, considering a cathode composed of a lithium
cobalt host and an anode composed of a crystalline structure of hexagonal graphite
a Load b Load

H2 O2 H2 O2

H2 H2 O H2 O O2

ANODE CATHODE ANODE CATHODE


diffusion path diffusion path

H O O2- H
+
e

Fig. 3.5 Schematic cross section of the simplified planar anodeelectrodecathode structure of
two typical fuel cells: a polymer-electrolyte membrane fuel cell and b solid oxide fuel cell

Вам также может понравиться