201h.gghklgluob6 Book Nanomaterialsforfuelcellcataly 7694

Download as pdf or txt
Download as pdf or txt
You are on page 1of 583

Nanostructure Science and Technology

Series Editor: David J. Lockwood

Kenneth I. Ozoemena
Shaowei Chen Editors

Nanomaterials
for Fuel Cell
Catalysis
Nanostructure Science and Technology

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

More information about this series at http://www.springer.com/series/6331


Kenneth I. Ozoemena • Shaowei Chen
Editors

Nanomaterials for Fuel Cell


Catalysis
Editors
Kenneth I. Ozoemena Shaowei Chen
Council for Scientific and Department of Chemistry and Biochemistry
Industrial Research University of California
Pretoria, South Africa Santa Cruz, CA, USA

ISSN 1571-5744 ISSN 2197-7976 (electronic)


Nanostructure Science and Technology
ISBN 978-3-319-26249-9 ISBN 978-3-319-29930-3 (eBook)
DOI 10.1007/978-3-319-29930-3

Library of Congress Control Number: 2016942554

© Springer International Publishing Switzerland 2016


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or
dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt
from the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, express or implied, with respect to the material contained
herein or for any errors or omissions that may have been made.

Printed on acid-free paper

This Springer imprint is published by Springer Nature


The registered company is Springer International Publishing AG Switzerland
Preface

Ready availability of sufficient energy resources is critical in virtually every aspect


of our life. Whereas fossil fuels have remained our primary energy sources,
extensive efforts have been devoted to fuel cell research in the past few decades,
which represents a unique technology that will make substantial contributions to
our energy needs by converting the chemical energy stored in small (organic)
molecule fuels into electricity and, more importantly, exert minimal negative
impacts on the environment. In fuel cell electrochemistry, the reactions typically
involve the oxidation of fuel molecules at the anode and reduction of oxygen at the
cathode. Both reactions require appropriate catalysts such that a sufficiently high
current density can be generated for practical applications. Precious metals, in
particular, the platinum group metals, have been used extensively as the catalysts
of choice. Yet their high prices and limited reserves have severely hampered the
widespread commercialization of fuel cell technologies. Therefore, a significant
part of recent research efforts has been focused on the development of effective
electrocatalysts with reduced or even zero amounts (and hence costs) of precious
metals used that exhibit competitive or even improved electrocatalytic performance
as compared to state-of-the-art platinum-based catalysts. Toward this end, it is
imperative to understand the fundamental mechanisms involved such that an
unambiguous structure–activity correlation may be established, from which the
activity may then be further enhanced or even optimized. It should be recognized
that the reaction mechanisms in fuel cell electrochemistry are rather complicated
and not fully understood. Yet advances on the theoretical and experimental fronts
have yielded significant insights which offer important guidelines in the design and
engineering of fuel cell catalysts.
It is within this context that this book volume is conceived and developed. The
chapters are written by some of the leading experts in the field. The key goal is to
highlight recent progress in electrocatalysis at both fuel cell anode and cathode,
with a focus on the impacts of the design and engineering of electrode catalysts on
the catalytic performance. This is a critical first step toward the establishment of a
structure–activity correlation. A significant portion is devoted to oxygen reduction

v
vi Preface

reaction, as this has been recognized as a major bottleneck that largely determines
the overall fuel cell performance due to its sluggish electron-transfer kinetics and
complicated reaction pathways. Relevant research is typically centered around
three aspects. The first involves alloying and surface ligand engineering of
platinum-based electrocatalysts (the conventional catalysts) through the so-called
electronic and geometrical contributions such that the costs may be reduced and
concurrently the performance improved. The second entails the development of
non-platinum precious metal nanoparticle catalysts. The third is focused on cheap
transition-metal oxides or totally metal-free catalysts (e.g., doped carbons). In the
last two, while the costs of the catalysts may be substantially reduced as compared
to those of the platinum-based counterparts, their performances have mostly
remained subpar. Thus, how to further improve their activity is a leading challenge
in the field.
Similar issues have been found with the anode reactions. Conventionally, small
(organic) molecules, such as hydrogen, methanol, and formic acid, have been used
as potential fuels. More recently, C2 molecules such as ethanol and ethylene glycol
have also been attracting extensive interest because of their ready availability, low
toxicity, and high energy density. Yet the oxidation mechanisms involved are far
more complex than those of the C1 counterparts (namely, methanol and formic
acid), leading to reduced efficiency in the reaction. Therefore, a major thrust of
current research is to unravel the reaction mechanisms involved such that the
catalytic performance may be further improved and ultimately optimized.
Whereas these issues represent daunting challenges, one may also choose to
accept them as unique opportunities where breakthroughs will help advance fuel
cell technologies toward commercial applications. This is no doubt a
multidisciplinary endeavor, including materials science, (electro)chemistry, inter-
facial engineering, and so on. It is our hope that this book will offer a unique
glimpse of the state of the art of fuel cell electrocatalysis and therefore may serve as
a technical reference for researchers at all levels in the areas of nanoparticle
materials and fuel cell technologies.

Santa Cruz, CA, USA Shaowei Chen


Pretoria, South Africa Kenneth I. Ozoemena
Contents

1 Electrochemistry Fundamentals: Nanomaterials Evaluation


and Fuel Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Neil V. Rees
2 Recent Advances in the Use of Shape-Controlled Metal
Nanoparticles in Electrocatalysis . . . . . . . . . . . . . . . . . . . . . . . . . . 31
Francisco J. Vidal-Iglesias, José Solla-Gullón, and Juan M. Feliu
3 Pt-Containing Heterogeneous Nanomaterials for Methanol
Oxidation and Oxygen Reduction Reactions . . . . . . . . . . . . . . . . . . 93
Hui Liu, Feng Ye, and Jun Yang
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic
Nanocrystals for Fuel Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
Ruizhong Zhang and Wei Chen
5 Integrated Studies of Au@Pt and Ru@Pt Core-Shell
Nanoparticles by In Situ Electrochemical NMR,
ATR-SEIRAS, and SERS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
Dejun Chen, Dianne O. Atienza, and YuYe J. Tong
6 Recent Development of Platinum-Based Nanocatalysts
for Oxygen Reduction Electrocatalysis . . . . . . . . . . . . . . . . . . . . . . 253
David Raciti, Zhen Liu, Miaofang Chi, and Chao Wang
7 Enhanced Electrocatalytic Activity of Nanoparticle Catalysts
in Oxygen Reduction by Interfacial Engineering . . . . . . . . . . . . . . 281
Christopher P. Deming, Peiguang Hu, Ke Liu, and Shaowei Chen

vii
viii Contents

8 Primary Oxide Latent Storage and Spillover


for Reversible Electrocatalysis in Oxygen
and Hydrogen Electrode Reactions . . . . . . . . . . . . . . . . . . . . . . . . . 309
Milan M. Jaksic, Angeliki Siokou, Georgios D. Papakonstantinou,
and Jelena M. Jaksic
9 Metal-Organic Frameworks as Materials for Fuel Cell
Technologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367
Henrietta W. Langmi, Jianwei Ren, and Nicholas M. Musyoka
10 Sonoelectrochemical Production of Fuel Cell Nanomaterials . . . . . 409
Bruno G. Pollet and Petros M. Sakkas
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based
Nanocatalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 435
T.S. Almeida, N.E. Sahin, P. Olivi, T.W. Napporn,
G. Tremiliosi-Filho, A.R. de Andrade, and K.B. Kokoh
12 Direct Alcohol Fuel Cells: Nanostructured Materials
for the Electrooxidation of Alcohols in Alkaline Media . . . . . . . . . . 477
Hamish Andrew Miller, Francesco Vizza,
and Alessandro Lavacchi
13 Effects of Catalyst-Support Materials on the Performance
of Fuel Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 517
Paul M. Ejikeme, Katlego Makgopa, and Kenneth I. Ozoemena
14 Applications of Nanomaterials in Microbial Fuel Cells . . . . . . . . . . 551
R. Fogel and J.L. Limson

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 577
Chapter 1
Electrochemistry Fundamentals:
Nanomaterials Evaluation and Fuel Cells

Neil V. Rees

1.1 Introduction

The field of electrochemistry has enjoyed a renaissance in the last two decades as its
relevance to the ever-expanding realm of nanoscience has been appreciated. As new
nanomaterials have been discovered and designed by physical science, many of
their properties of interest have been associated with electron transfer, as solid state
devices or catalysts, most of which can be probed and investigated via electro-
chemical techniques. It is in the latter case that we focus this chapter, in particular to
fuel cell catalysis. Clearly the aim of a fuel cell is to achieve full oxidation of its fuel
to extract the maximum thermodynamic output, and we therefore require efficient
catalysts in order to achieve this at low to intermediate temperatures (roughly
333–413 K). Electrochemistry plays two roles in the case of fuel cells therefore:
first in characterising and investigating candidate catalysts, and second in under-
standing and optimising the reduction and oxidation (“redox”) processes occurring
in the fuel cell itself. This chapter will not attempt to cover the whole of this field:
there are many excellent texts on physical electrochemistry [1–4] and a growing
number of similarly authoritative works on the fundamental science of fuel cells
[5–8], but will provide a brief survey of the key concepts and illustrations from
recent literature that those of us working in the field of fuel cell science should be
aware.

N.V. Rees (*)


School of Chemical Engineering, University of Birmingham, Birmingham B15 2TT, UK
e-mail: [email protected]

© Springer International Publishing Switzerland 2016 1


K.I. Ozoemena, S. Chen (eds.), Nanomaterials for Fuel Cell Catalysis,
Nanostructure Science and Technology, DOI 10.1007/978-3-319-29930-3_1
2 N.V. Rees

1.2 Part I: Ex-Situ Electrochemistry

The aim of ex-situ testing of candidate materials for fuel cells is typically to
simplify conditions such that variables can be carefully controlled. In the majority
of cases the testing takes the form of solution-phase cyclic voltammetry of candi-
date electrocatalysts and so the degree of catalytic activity needs to be ascertained.
In order to understand how this can be performed correctly, avoiding many com-
mon pitfalls, we shall first review the underlying physical electrochemistry of the
cyclic voltammetric response.

1.2.1 The Electrode Interface

In general, any interface comprising of two different phases will develop a potential
difference due to different electrical potentials of the two phases. This is true also
for an electrode placed in a test solution. The existence of a potential difference
therefore leads to an electric field gradient within the solution close to the electrode;
where there are ions in the solution they migrate under the influence of this field
such that an electrical double layer is rapidly established, with the Gouy-Chapman-
Stern model most commonly used to describe it (see Fig. 1.1). Clearly as the
concentration of ions in solution increases, so the diffuse layer becomes more
compressed.
Changes to the electrode interface due to adsorption, etc, can be detected via the
flow of non-Faradaic currents. These are current flows required to maintain
electroneutrality, and are distinct from Faradaic currents which are associated
with electron transfer to/from electroactive species to effect reduction/oxidation.

1.2.2 Mass Transport

Transport of materials through a fluid can occur via diffusion along a concentration
gradient, migration along an electric field gradient, natural convection due to
thermal gradients, or forced convection (due to stirring, flow, etc. deliberately
imposed on the system).
Diffusion is typically described via Fick’s Laws

∂c
j ¼ D ð1:1Þ
∂x
2
∂c ∂ c
¼D 2 ð1:2Þ
∂t ∂x
1 Electrochemistry Fundamentals: Nanomaterials Evaluation and Fuel Cells 3

Fig. 1.1 The Gouy-


Chapman-Stern model of
the electrical double layer,
with adsorbed counterions
forming the outer
Helmholtz plane (OHP), a
surfeit of counterions in the
diffuse layer and then bulk
solution. Accordingly, the
potential drops from the
electrode potential (ϕm) to
the solution potential (ϕs)

Diffusion to an electrode is also described in terms of the degree of development of


the diffusion field around the electrode. As the electrode reaches a potential where
electron transfer occurs and so a concentration gradient is established, diffusion is
initially linear (or one-dimensional) and the diffusion field gradually develops such
that eventually it becomes convergent (see Fig. 1.2).
The time taken for the establishment of convergent diffusion, tconv, can be
estimated by the expression [4]

r2
tconv  ð1:3Þ
D

Migration effects are often undesired in ex-situ experiments, and so an excess of


inert (or supporting) electrolyte is commonly added to eliminate them. However, in
the absence of excess electrolyte (i.e. less than full support), then migration will
become significant and needs to be accounted for via the Nernst-Planck and Poisson
equations [9]
 
∂C zF  
¼D ∇ Cþ
2
C∇ ϕ þ ∇C∇ϕ
2
ð1:4Þ
∂t RT
4 N.V. Rees

Fig. 1.2 (a) Linear and (b) convergent diffusion fields to an electrode

ρ
∇2 ϕ ¼  ð1:5Þ
εs ε0

where z is the species charge, ϕ is the potential, εs is the dielectric constant of the
solvent medium, ε0 is the permittivity of free space, and ρ is the local charge
density, found from summing all local charges present
X
ρ ¼ F i zi Ci ð1:6Þ

Natural convection effects are notoriously unpredictable and so are generally


eliminated by thermostatitng of the electrochemical cell and conducting the exper-
iment over as short a time as practicable, where a timescale of 20–30 s is usually
held to be the maximum desired [10].
Forced convection is imposed deliberately on the system in order to increase
mass transport. As such the formed of convection is usually chosen such that the
hydrodynamics of the system are well-defined [11].

1.2.3 Electrode Kinetics

The simplest case is that of fully reversible behaviour (i.e. where the electron
transfer kinetics are effectively infinitely fast on the timescale of the mass trans-
port), where the Nernst equation will hold at all times at the electrode surface [1].
For quasi-reversible and irreversible systems, account must be made of the
non-Nernstian condition at the electrode surface and this is most commonly
achieved via the Butler-Volmer equations
αFη
kf ¼ k0 e RT ð1:7Þ
þβFη
kb ¼ k0 e RT ð1:8Þ

where α + β ¼ 1, and η ¼ E  E0.


These are routinely available in commercial simulation packages and are phe-
nomenological insofar as there is limited molecular insight gained from their
1 Electrochemistry Fundamentals: Nanomaterials Evaluation and Fuel Cells 5

results. More theoretically informative would be the use of Marcus-Hush theory


[12–18], which can be used in a predictive sense as well as modelling experimental
data. Here the reductive and oxidative rates constants (kf and kb respectively) can be
expressed as
n o
Zþ1 exp ΔG6¼ ðε Þ=k B T
red
kf ¼ Ared ðεÞ dε ð1:9Þ
1 þ expfeðε  EÞ=kB T g
1
n o
Zþ1 exp ΔG6¼ ðε Þ=k B T
red
kb ¼ Aox ðεÞ dε ð1:10Þ
1 þ expfþeðε  EÞ=kB T g
1

where e is the electronic charge, kB the Boltzmann constant, Ared/ox(ε) is the


pre-exponential factor that includes the influence of the electronic states on metallic
electrode and electroactive species and the density of states of the electrode. If A is
assumed to be independent of energy, ε, then the general form for the reductive and
oxidative rate constants can be written as
(  )
I red η* ; λ*
kf ¼ k0   ð1:11Þ
I red 0; λ*
(  )
I ox η* ; λ*
kb ¼ k0   ð1:12Þ
I ox 0; λ*

where η*,
 λ* aredimensionless overpotential and reorganisation energy given by
η ¼ RT E  E0f and λ* ¼ RT
* F Fλ
, respectively. I(η*, λ*) is an integral of the form
n o
Zþ1 exp ΔG6¼ ðχ Þ
  red
I red η; λ *
¼ dχ ð1:13Þ
1 þ expfχ g
1

Zþ1 
  exp ΔG6¼ ox ðχ Þ
I ox η; λ *
¼ dχ ð1:14Þ
1 þ expfþχ g
1

where χ ¼ RTF
ðε  EÞ.
It is most commonly encountered in its symmetric representation [19, 20], where
 2
λ* η* þ χ
ΔG6¼
sym, red ðχ Þ ¼ 1þ ð1:15Þ
4 λ*
 2
λ* η* þ χ
ΔG6¼
sym, ox ð χ Þ ¼ 1  ð1:16Þ
4 λ*
6 N.V. Rees

However, this can lead to poor fitting to experimental data in certain circumstances,
for example where the electron transfer is extremely irreversible, and so the
asymmetric representation is potentially more powerful [19, 20].

* 2  * (  * 2 )
λ η *
þ χ η þ χ η þ χ
ΔG6¼
asym, red ðχ Þ ¼ 1þ þ β* 1
4 λ* 4λ* λ*
2
β*
þ ð1:17Þ
16λ*
 2  * (  * 2 )
6¼ λ* η* þ χ * η þχ η þχ
ΔGasym, ox ðχ Þ ¼ 1 þβ 1
4 λ* 4λ* λ*
2
β*
þ ð1:18Þ
16λ*

1.2.4 The Cyclic Voltammogram

The cyclic voltammogram, commonly abbreviated to “CV”, shows the current


response to a triangular voltage ramp (see Fig. 1.3) and owes its shape to the
complex interplay of both electrode kinetics and mass transport effects. As such
it can take different forms depending on the exact details of these two factors, often
described as transient or steady-state.

1.2.4.1 Transient Cyclic Voltammetry

Figure 1.3a shows the transient form of the CV, which is obtained for linear
diffusion systems, most commonly “macro”electrodes (of characteristic dimension
>100 μm) or smaller electrodes at very short timescales (i.e. high voltage scan
rates) [21]. Here the characteristic shape is due to the interaction of kinetics with
mass transport: the current maximum is due to the competing factors of (i) an
exponentially increasing rate of electron transfer (see Butler-Volmer equations),
and (ii) rapid depletion of the electroactive material in the double layer caused by
relatively slow diffusion of fresh material from bulk solution. In a sufficiently large
volume of solution, these factors equilibrate at the diffusion limiting current (Ilim).
The transient CV is usually characterised in terms of the peak currents (anodic
and cathodic, Ip,a and Ip,c) and the separation of the anodic and cathodic peak
potentials (ΔEpp ¼ Ep,a  Ep,c). The reversibility of the redox couple is defined in
terms of the peak separation [1] (Fig. 1.4, Table 1.1):
1 Electrochemistry Fundamentals: Nanomaterials Evaluation and Fuel Cells 7

I/A (a) (b)

I/A

E/V E/V

Fig. 1.3 Typical cyclic voltammograms for (a) transient and (b) steady-state responses, due to
linear and convergent diffusion respectively

(Epa, Ipa)
I/A

(Epc, Ipc)

E/V

Fig. 1.4 Characteristics of a transient cyclic voltammogram


8 N.V. Rees

Table 1.1 Cyclic ΔEpp ¼ Epa  Epc (mV) Definition Other relations
voltammetric definitions Epa þEpc
60/n Reversible E0 ¼ 2
60/n < ΔEpp < 120/n Quasi-reversible –
>120/n Epa þEpc
Irreversible E0 ¼ 2

It should be noted here that the reversibility of a redox system is clearly a


function of the timescale of the experiment, with a redox couple typically
displaying reversible behaviour at slow voltage scan rates, and quasi-reversible or
even irreversible behaviour at higher voltage scan rates.
Analytical results for the peak currents are available for the reversible and
irreversible cases, due to Randles and Sèvčı́k [1]:
 
I p, rev ¼ 2:69  105 nACbulk D1=2 ν1=2 ð1:19Þ
 
I p, irrev ¼ 2:99  105 n3=2 α α
3=2
ACbulk D1=2 ν1=2 ð1:20Þ

where, n is the total number of electrons transferred, nα is the number of electrons


transferred in the rate limiting step, A the geometric electrode area, Cbulk the bulk
concentration of the electroactive species, D the diffusion coefficient, α the transfer
coefficient, and ν the voltage scan rate.

1.2.4.2 Steady-State Voltammetry

In Fig. 1.3b, a steady-state CV is shown: this is most commonly observed for


microelectrode voltammetry (for electrodes with characteristic dimension
<20 μm), where the diffusion field is fully convergent. The sigmoidal shape is
due to the rate of mass transport being sufficiently fast to ‘keep up’ with the
accelerating rate of electron transfer at least until the limiting current (Ilim) is
reached.
Since the reverse scan retraces the forward scan in the absence of hysteresis
effects, the steady-state voltammogram is often recorded as a linear sweep rather
than a triangular voltage ramp. The limiting current is given by [1]

I lim ¼ 4nFCbulk Dr ð1:21Þ

where n, Cbulk, and D have their usual meaning, F is the Faraday constant and r the
disk radius.

1.2.4.3 The Transition Between Transient and Steady-State

The position of the voltammetric “wave” can be seen to shift to higher


overpotentials as the voltammogram shifts from fully transient to fully steady-
state, that is, as the mass transport accelerates from linear to convergent
1 Electrochemistry Fundamentals: Nanomaterials Evaluation and Fuel Cells 9

Fig. 1.5 Simulation cyclic voltammograms showing transition from transient to steady state and
the subsequent potential shift in the voltammogram. All parameters are the same, except from
electrode radius which varies: 1 mm (black), 10 μm (magenta), 1 μm (red), 100 nm (green), and
10 nm (blue)

(Fig. 1.5). There is no change to the electrode kinetics in this case, the effect is
purely one of mass transport. A simple way to understand this is that the higher rates
of diffusion sweeps material to and from the electrode more rapidly, hence the
electroactive species spends less time in the near vicinity of the electrode and a
higher overpotential is required to drive the electron transfer at a fast enough rate
for the species to react in that shorter time period.

1.2.4.4 Adsorbed and Thin-Layer Voltammetry

The voltammetry described thus far has been implicitly concerned with a large
volume of solution containing a bulk concentration of electroactive species which is
not significantly depleted by the electrolysis occurring at the electrode.
Figure 1.6 shows a schematic of both adsorbed and thin-layer systems with their
characteristic voltammetric responses. In both cases the current response decays to
zero as the concentration of electroactive species is exhausted. Note that in the
adsorbed case, the anodic and cathodic peaks are ideally symmetrical about the
potential axis due to the absence of mass transport. For thin-layer voltammetry,
there is an asymmetry due to the limited mass transport (usually diffusion) occur-
ring in the thin-layer volume.
10 N.V. Rees

Fig. 1.6 Schematic voltammograms for (a) adsorbed species, and (b) thin-layer (finite volume)
solution

1.2.5 Hydrodynamic Voltammetry

There are a wide range of different hydrodynamic electrodes, where forced con-
vection is applied to achieve a well-defined fluid behaviour [11]. These commonly
differ in:
(i) size of electrode—macro vs micro
(ii) geometry—disk, band, tube, ring, ring-disk
(iii) method of convection—flow systems, rotating systems, impinging jet systems
(iv) symmetry—axisymmetric vs non axisymmetric
(v) uniformity of accessibility
(vi) flow regime—laminar vs fully turbulent
Alternative detailed reviews consider this as a topic [11, 22], but here we shall
focus on the rotating disk electrode (RDE) and rotating ring-disk electrode (RRDE)
as they are commonly used within the fuel cell community (Fig. 1.7).
The RDE (and RRDE) have been in use for many years, having been developed
prior to the advent of microelectrodes as a means to shorten the experimental
timescale and measure more rapid processes than could be achieved via
diffusion-only studies with macroelectrodes [23, 24].
The electrode is mounted axisymmetrically on an insulated rotating shaft, and
rotated at sufficient frequency to set up laminar motion in the solution.
In this case the high rates of convective mass transport ensure a steady-state
voltammetric response where the limiting current is given by the Levich equation
[23]:
1 Electrochemistry Fundamentals: Nanomaterials Evaluation and Fuel Cells 11

Fig. 1.7 Schematic


diagrams of (a) rotating
disk, and (b) rotating ring-
disk electrodes

I lim ¼ 0:62nFACbulk D2=3 υ1=6 ω1=2 ð1:22Þ

where υ is the kinematic viscosity of the solution and ω is the angular speed (rad s1).
Kinetic parameters can be extracted via Koutecky-Levich analysis, which
deconvolutes the total current into kinetic and limiting contributions [25, 26]:

I 1 ¼ I 1 1
lim þ I k ð1:23Þ

Plotting 1/I vs ω1/2 at a range of potentials yields intercepts of value 1/Ik. Since Ik is
a function of E, given by

I k ¼ nFACbulk kðEÞ ð1:24Þ

where k(E) is the potential-dependent electrochemical rate constant given by the


Butler-Volmer expression
 
αFη
kðEÞ ¼ k0 exp ð1:25Þ
RT

Then a further (Tafel) plot of ln Ik vs η provides α and k0. The Koutecky-Levich


method is widely used for multi-electron systems, most notably oxygen reduction
(ORR). However, recent work by Masa et al. suggests that for accurate results it is
necessary to take account of the surface roughness of the catalyst deposit, since the
12 N.V. Rees

ratio of actual to geometric surface area will determine how the apparent rate
constant, k0app, differs from the true rate constant, k0 [27].
The RRDE is usually used in a generator-collector type experiment which can be
illustrated by the (simplified) ORR reaction as follows.

k1
O2 ðgÞ þ 4H þ ðaqÞ þ 4e ! 2H2 OðlÞ
k2 k3
O2 ðgÞ þ 2H þ ðaqÞ þ 2e ! H 2 O2 ðadsÞ ! H 2 O2 ðaqÞ
k4
H 2 O2 ðadsÞ þ 2H þ ðaqÞ þ 2e ! 2H2 OðlÞ

The disk is subjected to a normal voltage scan, in this case to reduce oxygen, whilst
the ring potential (Ering) is held at a potential sufficient to oxidise the intermediate
(here, hydrogen peroxide). The disk current therefore provides information on the
rate of reduction of parent species, whilst the ring current provides information on
the amount of intermediate produced.

1.2.6 The Voltammetry of Nanoparticles

The vast majority of catalysts are nanomaterials (assumed spherical) and hence it is
important to identify those characteristics that define their voltammetry.
An isolated nanoparticle on a planar substrate behaves as a spherical electrode
with a hindered convergent diffusion field (see Fig. 1.8).
For which Bobbert et al. have derived the limiting current expression to be [28]:

I lim ¼ 4π ðln2ÞnFCbulk Dr np ð1:26Þ

where rnp is the radius of the nanoparticle.

Fig. 1.8 Schematic


diagram showing
convergent diffusion to a
(nano)sphere on a plane
1 Electrochemistry Fundamentals: Nanomaterials Evaluation and Fuel Cells 13

The voltammetry for such as system would appear to be relatively straightfor-


ward: the nanoscale size of the ‘electrode’ (i.e. particle) ensures that all diffusion is
convergent on all practical timescales (see Eq. 1.3), and the exceptionally high rates
of mass transport associated with this would be expected to cause a shift of the wave
to higher overpotentials.
However, it is not common to measure single particle electrochemistry (see
later), and instead nanocatalysts are usually studied and used in very large numbers.
The catalysts can be deposited directly onto a substrate electrode (usually glassy
carbon, GC) or be deposited upon a carbon support material during their fabrica-
tion, and the catalyst/support then deposited onto the substrate. In this case the
carbon support is usually carbon black (often the commercial Vulcan XC-70), but
studies are increasingly using nanotubes, graphenes, etc. There are subtle differ-
ences to these two cases, so we consider them in turn.

1.2.6.1 Nanoparticle-Substrate Systems

In this case the nanoparticles are deposited upon the substrate, often via spraying or
drop-casting and subsequent evaporation of solvent. The particles are essentially
randomly distributed across the surface, and typically are in such numbers that they
do not form a monolayer across the entire substrate electrode surface (i.e. coverage
<100 %), shown schematically in Fig. 1.9.
By considering the diffusion to each individual particle, it should be clear that
there are two extreme cases (and several intermediate ones) which may occur as a
result of the relative spacing of the particles from each other [29].
First, where the particles are very far apart (i.e. extremely low coverage) such
that each individual particle develops a convergent diffusion field independent of
every other particle. Each particle is diffusionally isolated, and so the voltammetry
of the ensemble (or N particles) will appear to be a sigmoidal (steady-state)
voltammogram, with limiting current given by

I lim ¼ 4Nπ ðln2ÞnFCbulk Drnp ð1:27Þ

Second, where the particles are very close together (i.e. high coverages) such that
each individual particle’s diffusion field interferes or overlaps with its neighbour’s,
then the overall effect is for the ensemble to behave as if the whole area covered is

Fig. 1.9 Randomly


deposited particles on a
substrate electrode
14 N.V. Rees

Fig. 1.10 The four cases of diffusion to spheres on a plane (particles on a substrate electrode)

subject to linear diffusion. The voltammetry will therefore appear to be of a


transient voltammogram corresponding to an active area equal to that of the
geometric area over which the particles are deposited (i.e. the area of the substrate
electrode). The usual equations relating to macroelectrode voltammetry
(i.e. Randles-Sèvčı́k) will apply.
These two extremes are often termed Case 1 and Case 4 behaviour respectively,
and as they names suggest, two intermediate cases have been identified
corresponding to increasing degrees of overlap of diffusion fields associated with
neighbouring nanoparticles [19] (Fig. 1.10).
When considering the likely Case in operation for a given system, it is worth-
while considering the diffusion length on the experimental timescale, t [30]:
pffiffiffiffiffiffiffiffi
hxi  2Dt ð1:28Þ

For the reason that even if the particles are small and relatively far apart, if the
timescale of the experiment is sufficiently long that ‹x› approaches half of the inter-
particle separation, then diffusion field interference will occur and Cases 2–4 will
apply.

1.2.6.2 Nanoparticle@Support-Substrate Systems

There are three main ways for the presence of carbon support particles to alter the
voltammetry. First, by increasing the resistance between the catalyst particle and
substrate: this is usually not a significant effect in the cases of highly conductive
1 Electrochemistry Fundamentals: Nanomaterials Evaluation and Fuel Cells 15

carbons (carbon black, nanotubes, etc.), but can become noticeable for less con-
ductive supports such as graphene oxide.
Second, the partial or full occlusion of the catalyst particle. This can occur as
part of the catalyst fabrication, as this commonly involves a chemical reduction of
the dissolved catalyst precursor salt in the presence of the carbon support in
suspension. Any degree of aggregation of the support particles will necessarily
occlude catalyst particles. In some cases, it has been reported that during fabrication
the catalyst particles may intercalate graphitic particles of carbon support. Alter-
natively, occlusion may occur during the deposition of the catalyst@support parti-
cles onto the substrate through multilayer formation. In these scenarios, the
voltammetric changes will be slight, with Cases I–III affected to a decreasing
degree and Case IV unaffected. In Case I, the voltammetry will remain steady-
state, except the limiting current will reflect the lower apparent radius of the
nanoparticles.
Third, the deposition of higher coverages, or strongly aggregated
catalyst@support particles will lead to the existence of voids within the
catalyst@support deposit. There will be extremely hindered diffusion to these
voids from bulk solution and a mixed diffusion regime will be established, where
either convergent or linear diffusion occurs to catalyst particles on the surface of the
deposit and thin-layer diffusion occurs to catalyst particles within voids in the body
of the layer [31–33] (Fig. 1.11).
The resulting voltammetry of such a mixed diffusion regime is therefore some
algebraic sum of linear (or convergent is applicable) diffusion and thin-layer
diffusion signals. Since the latter, by definition, occurs at a lower overpotential
than the former, there is a necessary shift in overpotential to lower values than a
fully linear diffusion response (Fig. 1.12).

1.2.7 Electrocatalysis

An electrocatalyst increases the rate of an electrochemical reaction by providing a


lower energy pathway across the reaction coordinate. Due to the potential depen-
dence of the electrochemical rate constant on overpotential (see Eqs. 1.7 and 1.8)

Fig. 1.11 A multi-layer deposit with illustrative voids


16 N.V. Rees

250
Semi-infinite diffusion
200 Thin layer diffusion

150

100
i / mA

50

-50

-100

-150

-200
-0.8 -0.6 -0.4 -0.2 -0.0 0.2 0.4 0.6 0.8
E/V
Fig. 1.12 Illustration of how mixed diffusion regime can cause a lowering of overpotential.
Reproduced with permission from [34], copyright 2008 Elsevier

this necessarily means that the overpotential required for the electrochemical
reaction is decreased and the voltammetric wave is shifted to lower overpotentials.
However, the converse is certainly not necessarily true.
There is a widespread naı̈vety that a candidate catalyst that causes a shift of the
voltammetric wave to lower overpotentials must therefore be catalytic. The pre-
ceding discussion should make it clear that this conclusion can only be reached if
the mass transport to the system has been fully characterised and understood. We
have seen that the size of the particle (i.e. electrode) can cause a shift in
overpotential through the increasing rates of mass transport via diffusion as particle
size decreases. Further, the arrangement of the deposited particles will affect the
voltammetric shape (and therefore position on the potential axis) depending on
whether Case I–IV behaviour is followed. Finally, and most subtle, the existence of
a mixed (linear-thin layer) diffusion regime is itself sufficient to cause a lowering of
the observed overpotential.
In any evaluation of nanoparticulate materials as possible catalysts, it is there-
fore imperative to ensure that the role of coverage, support particles, etc. are taken
into account, through careful control experiments and simulation. Ideally, any study
of candidate catalysts should measure the kinetic parameters for the reaction, as this
is the only truly unambiguous measure of catalytic activity.
1 Electrochemistry Fundamentals: Nanomaterials Evaluation and Fuel Cells 17

1.0

Current / mA 0.5

0.0

-0.5

-1.0

-1.5
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
E / V (RHE)

Fig. 1.13 Cyclic voltammogram of Pt film electrode in 0.5 M H2SO4 at 50 mV s1. Reproduced
with permission from [35], copyright 2007 Elsevier

1.2.8 Electrochemical Surface Area (ECSA) Measurements

Catalyst preparations are often characterised by their ECSA, usually quoted in


m2 g1, as a measure of their activity in comparison to metal loading. The most
commonly used method for determining the ECSA is to measure the CV of the
catalyst in acid solution, determining the charge passed during the adsorption/
desorption of protons, and relating this to an average charge per unit area for the
metal surface (Fig. 1.13).
This method is particularly used for Pt-based catalysts as the adsorptive charge
of 210 μC cm2 is well known from an average of the values for the (1 1 1), (1 1 0),
and (1 0 0) surfaces from single crystal studies [36].
However, the measurement of ECSA is often more approximate than acknowl-
edged: the use of surface physisorption as a methodology requires the sample and
solution to be scrupulously clean: this is often not achievable for some catalyst
testing. The appearance of the resulting CV clearly shows this effect via truncated
or poorly resolving adsorption/desorption peaks.
It may therefore be desirable to use additional methods to provide assurance on
the ECSA: this can be obligatory for some non-Pt catalysts where hydrogen
adsorption is not reliable (e.g. Pd surfaces where H atoms readily absorb). In
these cases, CO adsorption is often used, and even the deposition of metals such
as Cu [36–38].
18 N.V. Rees

a b

A B
± e¯

Electrode
Fig. 1.14 Schematic diagram showing: (a) the nanoparticle becoming a nanoelectrode on contact
with the substrate electrode. Note the latter is of a different material to the nanoparticle such that
the A/B redox couple is inactive at the substrate electrode and only occurs at the nanoparticle
whilst it is in electrical contact with the former. (b) For the contact period of milliseconds a
convergent diffusion field is almost instantly established leading to steady-state currents for the
conversion of A to B. Reproduced with permission from [46] copyright 2012, Elsevier

1.2.9 Processes at Single Nanoparticles

There has been considerable recent interest in the ability to measure kinetic
processes at individual nanoparticles [39, 40]. This has been pioneered by the
groups of Bard and Compton and cover reactions of the nanoparticles themselves
(oxidation/reduction) as well as reactions occurring at the particle surface of
adsorbed and solution species [41–46].
This is achieved via a conceptually simple, but often experimentally challeng-
ing, process of observing the current signal produced when collisions between free
nanoparticles and a substrate electrode (usually carbon) which is held at a suitable
potential. During the contact phase of the particle-electrode collision, the nanopar-
ticle becomes a spherical electrode on a plane (see Fig. 1.8, Eq. 1.26) and a fully
convergent diffusion field is set up within 1 μs of the contact which typically lasts
for 1–20 ms at 298 K in aqueous solution for particles of size ca 10 nm. It has been
shown that kinetic parameters can be extracted from these current signals, as in the
study be Kahk et al. into the proton reduction at gold and silver nanoparticles [46]
(Fig. 1.14).
In part of this study, Kahk et al. introduced 7 nm radius gold nanoparticles into a
solution of 10 mM perchloric acid and 0.6 M sodium perchlorate. A carbon
microelectrode was potentiostatted at a range of potentials and reductive current
spikes recorded, as illustrated in Fig. 1.15a. These spikes were analysed as a
function of potential in Fig. 1.15b, and the data points fitted to a steady-state
voltammogram with commercial modelling software (DigiSim™, BASI Inc.) to
extract the kinetic parameters k0 ¼ (7.0  5.0)  107 cm s1, and α ¼ 0.56  0.05.
To demonstrate the high rate of mass transport to the nanoparticle in contact with
the electrode, Kahk added acetic acid to the reaction, since the rate of dissociation
1 Electrochemistry Fundamentals: Nanomaterials Evaluation and Fuel Cells 19

a
31.5

-32.0

-32.5
I / nA

-33.0

-33.5

-34.0

-34.5

4.3 4.4 4.5 4.6 4.7 4.8


t/s

b 0.0
Average spike height / nA

-0.5

-1.0

-1.5

-2.0

-1.6 -1.4 -1.2 -1.0 -0.8 -0.6


E/V
Fig. 1.15 (a) AuNP impact spikes for the potential held at 1.40 V, and (b) a plot of average
spike height vs potential for AuNP impacts. Reproduced with permission from [46] copyright
2012, Elsevier

to release protons is known to be kd ¼ 1.3  106 s1. Repeat experiments were


conducted, and no increase in the apparent proton concentration was detected.
Simulation was re-performed to include the dissociation kinetics of acetic acid
and identical fits were obtained (Fig. 1.16).
20 N.V. Rees

0.0

-0.2
Average spike height / nA

-0.4

-0.6

-0.8

-1.0

-1.2

-1.4

-1.8 -1.6 -1.4 -1.2 -1.0


E/V
Fig. 1.16 Spike height vs potential plots for 7 nm radius AuNPs in solutions of 10 mM HClO4 and
0.6 M NaClO4 (circle) and also including 0.1 M acetic acid ( filled square). The solid line is a
computer simulation to include the dissociation of acetic acid and free perchloric acid protons.
Reproduced with permission from [46] copyright 2012, Elsevier

1.3 Part II: Examples

In this section we briefly discuss case studies with reference to the foregoing,
beginning with screening catalysts for a simple 1-electron hydrogen evolution
reaction (HER, or proton reduction) and progressing to a study involving the
4-electron oxygen reduction reaction.

1.3.1 Pt@TiO2 Clusters for HER

Blackmore et al. [47] have recently published a study into fabricating Pt@TiO2 core
shell particles as potential catalysts, and have characterised them via simple cyclic
voltammetry. Having made the NPs via a magnetron-sputtering gas condensation
cluster beam source, and size selected them via time-of-flight mass selection, the
particles were soft landed onto a glassy carbon (GC) target. This formed the active
surface of a working electrode, and after TEM imaging, basic cyclic voltammetry
was carried out (Fig. 1.17).
In this example, the coverage of the electrode by the nanoparticles was calcu-
lated to be (5  2) %. Therefore it is reasonable to assume that no multilayer
deposits of particles have occurred, especially in the light of the high resolution
1 Electrochemistry Fundamentals: Nanomaterials Evaluation and Fuel Cells 21

Fig. 1.17 STEM images of Pt-TiO2 clusters, showing single and multiple Pt cores (yellow) in
TiO2 shells (pink). Reproduced with permission from [47] copyright 2015, Royal Society of
Chemistry

imaging which shows well-dispersed particles. Hence, a simple shift in the


voltammetry towards lower overpotentials can be safely interpreted as an increase
in catalytic function: this is indeed what is observed in comparison to the control
signals obtained from bare GC, and GC modified with TiO2 particles of a similar
size (see Fig. 1.18).
Note that despite the small size of the particles the diffusion length is long
enough at this scan rate to give a “transient” shaped CV.

1.3.2 Evaluation of Surfactant-Stabilised Pt NPs for the ORR

In a study of the effects of surfactants as stabilisers in Pt catalyst fabrication Newton


et al. [48] compared commercial TKK Pt catalyst with Pt/C catalysts made by the
addition of nonylphenolethoxylate (NP9) and tetradecyltrimethylammonium bro-
mide (TTAB) as neutral and cationic surfactants respectively, to the platinum
precursor and carbon black. In characterising the three catalysts, they illustrated
the difficulty in obtaining reliable ECSA measurements using hydrogen adsorption
alone, and compared with CO adsorption measurements (see Fig. 1.19). From this
they were able to determine the following ECSA values (all in m2 g1): TKK—91,
Pt + NP9—22, and Pt + TTAB—1.7.
Due to the fact the catalyst particles were comprised of Pt deposited onto carbon
black (nanospheres), it is more appropriate to treat as if multilayer deposits on the
electrode are likely. The preferred method therefore is to use a hydrodynamic
method, here the RDE, in order to swamp any intra-layer diffusion effects with
forced convection to the surface of the deposit (Scheme 1.1).
22 N.V. Rees

20

-20

-40
I / µA

-60

-80

-100

-120

-140
-1.8 -1.6 -1.4 -1.2 -1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2
E / V (vs Ag/AgCl)
Fig. 1.18 Cyclic voltammograms recorded at 25 mV s1 for a solution of 2 mM HClO4 in 0.1 M
NaClO4 at the following surfaces: bare GC (dotted line); GC modified with TiO2 particles (dashed
line), and GC modified with Pt@TiO2 particles (solid line). Reproduced with permission from [47]
copyright 2015, Royal Society of Chemistry

In addition, the use of the RRDE enables quantitative data to be extracted


regarding the relative importance of the direct 4e pathway and indirect (2e, 2e)
pathway via hydrogen peroxide, for the catalysts under scrutiny (Fig. 1.20).
The number of electrons transferred, n, can be quantified from the disk and ring
currents using

4I D
n¼ ð1:29Þ
I D þ ðI R =N Þ

where ID is the modulus of the total disk current (i.e. the reduction of O2 and H2O2),
IR is the ring current (i.e. oxidation of H2O2), and N is the collection efficiency
determined to be 0.21 in this case (from prior calibration experiments).
This expression can therefore yield the number of electrons passed on each
catalyst—with n ¼ 4 the best possible result. This suggests that the Pt + NP9/C
catalysts is similar in performance to TKK (Figs. 1.21 and 1.22).
Koutecky-Levich analysis was then performed on the RDE data. Kinetic anal-
ysis can be taken further using the model of Hsueh and Chin [49], via plots of ID/IR
vs ω1/2 and IDL/(IDL  ID) vs ω1/2:
 
ID 1 2k1 2k3 ð1 þ k1 =k2 Þ 1=2
¼ 1þ þ ω ð1:30Þ
IR N k2 NZ2
1 Electrochemistry Fundamentals: Nanomaterials Evaluation and Fuel Cells 23

Fig. 1.19 Cyclic voltammograms recorded at 25 mV s1 of Pt/C catalysts of loading 20 μg Pt cm2
on 0.196 cm2 GC electrodes in N2-saturated 0.1 M HClO4 at 298 K. Each compared to a 0.196 cm2
Pt disk electrode. (a) Current vs potential, and (b) CO-stripping voltammogram. Reproduced with
permission from [48] copyright 2014, Royal Society of Chemistry

Scheme 1.1 A simplified ORR mechanism


24 N.V. Rees

Fig. 1.20 Results for oxygen reduction at an RRDE for Pt + TTAB/C, Pt + NP9/C, TKK Pt
catalyst and a Pt disk electrode. (a) Ring currents when held at +1.10 V, and (b) disk currents at
different rotation speeds measured at 298 K in O2-saturated 0.1 M HClO4 solution on a 0.196 cm2
substrate electrode. Reproduced with permission from [48] copyright 2014, Royal Society of
Chemistry

I DL ðk1 þ k2 Þ 1=2
¼1þ ω ð1:31Þ
I DL  I D Z1

where IDL is the disk limiting current, Z1 ¼ 0.62D(O2)2/3v1/6, and Z2 ¼ 0.62D


(H2O2)2/3v1/6, and the corresponding plots are (Figs. 1.23 and 1.24):
If the gradients of the plots in Figs. 1.20 and 1.21 are noted as G1 and G2
respectively, and the corresponding intercepts C1 and 1. Then rearranging Eqs. 1.30
and 1.31 enables the rate constants to be evaluated (Fig. 1.25)
 
C1 N  1
k1 ¼ G2 Z1 ð1:32Þ
C1 N þ 1
2Z 1 G2
k2 ¼ ð1:33Þ
C1 N þ 1
NZ 2 G1
k3 ¼ ð1:34Þ
C1 N þ 1
1 Electrochemistry Fundamentals: Nanomaterials Evaluation and Fuel Cells 25

Fig. 1.21 Calculated number of electrons and fraction of hydrogen peroxide produced as a
function of potential for each of the three catalysts and a Pt disk electrode. Reproduced with
permission from [48] copyright 2014, Royal Society of Chemistry

Fig. 1.22 Koutecky-Levich plots for potentials E ¼ 0.86, 0.87, 0.88, 0.89, 0.90 V. Reproduced by
permission from [48] from the Royal Society of Chemistry
26 N.V. Rees

Fig. 1.23 Plots of Eq. 1.30 using the data in Fig. 1.22. Reproduced with permission from
[48] copyright 2014, Royal Society of Chemistry

Fig. 1.24 Plots of Eq. 1.31 from the data in Fig. 1.22. Reproduced with permission from [48]
copyright 2014, Royal Society of Chemistry
1 Electrochemistry Fundamentals: Nanomaterials Evaluation and Fuel Cells 27

Fig. 1.25 Plot of k1/k2 according to above equations. Reproduced with permission from [48]
copyright 2014, Royal Society of Chemistry

1.4 Conclusions

The electrochemical evaluation of nanomaterials for use in fuel cells can be


straightforward, but it is important to consider the fundamentals of the system
before starting ex-situ testing. There are established as well as newly emerging
techniques available, but it is vital to fully understand the interplay of kinetics and
mass transport to the test material in order to make a reliable judgment on its
catalytic activity. Ideally, kinetic parameters should be determined at this stage,
enabling true comparison between candidate catalyst materials.
It should then be possible to move on to considering the role of the carbon
support, gas diffusion layer, etc in stages so that at each step towards single fuel cell
can be optimised and controlled analytically.

Acknowledgement N.V.R. is grateful to Mr. P.H. Robbs for assistance in the preparation of
figures for this article.

References

1. Bard AJ, Faulkner LR (2001) Electrochemical methods: from fundamentals to applications,


2nd edn. Wiley, New York
2. Girault HH (2004) Analytical and physical electrochemistry. EPFL Press, Lausanne
28 N.V. Rees

3. Gileadi E (2011) Physical electrochemistry: fundamentals, techniques and applications. Wiley,


New York
4. Compton RG, Banks CE (2010) Understanding voltammetry, 2nd edn. Imperial College Press,
London
5. Santos E, Schmickler W (2011) Catalysis in electrochemistry: from fundamental aspects to
strategies for fuel cell development. Wiley, Hoboken
6. Wieckowski A, Koper MTM (2009) Fuel cell catalysis: a surface science approach. Wiley,
Hoboken
7. Wieckowski A, Norskov JK (2010) Fuel cell science: theory, fundamentals and biocatalysis.
Wiley, Hoboken
8. Barbir F (2011) PEM fuel cells: theory and practice, 2nd edn. Academic, San Diego
9. Dickinson EJF, Limon-Petersen JG, Rees NV, Compton RG (2009) How much supporting
electrolyte is required to make a cyclic voltammetry experiment quantitatively “Diffusional”?
A theoretical and experimental investigation. J Phys Chem C 113:11157–11171
10. Albery WJ (1975) Electrode kinetics. Oxford University Press, Oxford
11. Brett CMA, Oliveira Brett AMCF (1986) Hydrodynamic electrodes. In: Bamford CH, Tipper
CFH, Compton RG (eds) Comprehensive chemical kinetics, vol 26. Elsevier, Amsterdam
12. Marcus RA (1956) On the theory of oxidation-reduction reactions involving electron transfer
I. J Chem Phys 24:966
13. Marcus RA (1964) Chemical and electrochemical electron transfer theory. Annu Rev Phys
Chem 15:155
14. Hush NS (1958) Adiabatic rate processes at electrodes: energy-charge relationships I. J Chem
Phys 28:962
15. Hush NS (1999) Electron transfer in retrospect and prospect 1: Adiabatic electrode processes. J
Electroanal Chem 470:170
16. Henstridge MC, Laborda E, Rees NV, Compton RG (2012) Marcus-Hush-Chidsey theory of
electron transfer applied to voltammetry: a review. Electrochim Acta 84:12–20
17. Weaver MJ (1987) Redox reactions at metal-solution interfaces. In: Compton RG
(ed) Comprehensive chemical kinetics, vol 27. Elsevier, Amsterdam
18. Fletcher S (2010) The theory of electron transfer. J Solid State Electrochem 14:705
19. Laborda A, Henstridge MC, Compton RG (2012) Asymmetric Marcus theory: application to
electrode kinetics. J Electroanal Chem 667:48–53
20. Marcus RA (1965) On the theory of electron transfer reactions VI. Unified treatment for
homogeneous and electrode reactions. J Chem Phys 43:679
21. Amatore C, Maisonhaute E, Simonneau G (2000) Ultrafast cyclic voltammetry: performing in
the few megavolts per second range without ohmic drop. Electrochem Commun 2:81
22. Rees NV, Compton RG (2008) Hydrodynamic microelectrode voltammetry. Russ J
Electrochem 44:368–389
23. Levich VG (1962) Physicochemical hydrodynamics. Prentice-Hall, Englewood Cliffs
24. Albery WJ, Hitchman ML (1971) Ring-disc electrodes. Oxford University Press, Oxford
25. Koutecky J, Levich VG (1958) The use of a rotating disk electrode in the studies of electro-
chemical kinetics and electrolytic processes. Zh Fiz Khim 32:1565–1575
26. Koutecky J, Levich VG (1957) An application of a rotating disk electrode o the studies of
kinetic and catalytic processes in electrochemistry. Dokl Akad Nauk SSSR 117:441–444
27. Masa J, Batchelor-McAuley C, Schuhmann W, Compton RG (2014) Koutecky–Levich anal-
ysis applied to nanoparticle modified rotating disk electrodes: electrocatalysis or misinterpre-
tation? Nano Res 7:71–78
28. Bobbert PA, Wind MM, Vlieger J (1987) Diffusion to a slowly growing truncated sphere on a
substrate. Physica A 141:58
29. Davies TJ, Compton RG (2005) The cyclic and linear sweep voltammetry of regular and
random arrays of microdisc electrodes: theory. J Electroanal Chem 58:63–82
30. Einstein A (1956) Investigations of the theory of Brownian movement. Dover, Englewood
Cliffs
1 Electrochemistry Fundamentals: Nanomaterials Evaluation and Fuel Cells 29

31. Sims MJ, Rees NV, Dickinson EJF, Compton RG (2010) Effects of thin-layer diffusion in the
electrochemical detection of nicotine on basal plane pyrolytic graphite (BPPG) electrodes
modified with layers of multi-walled carbon nanotubes (MWCNT-BPPG). Sens Actuators B
144:153–158
32. Henstridge MC, Compton RG (2012) Mass transport to micro- and nanoelectrodes and their
arrays: a review. Chem Rec 12:63–71
33. Bae JH, Han J-H, Han D, Chung TD (2013) Effects of adsorption and confinement on
nanoporous electrochemistry. Faraday Discuss 164:361–376
34. Streeter I, Wildgoose GG, Shao L, Compton RG (2008) Cyclic voltammetry on electrode
surfaces covered with porous layers: an analysis of electron transfer kinetics at single-walled
carbon nanotube modified electrodes. Sens Actuators B 133:462–466
35. Kunimatsu K, Senzaki T, Samjeske G, Tsushima M, Osawa M (2007) Hydrogen adsorption
and hydrogen evolution reaction on a polycrystalline Pt electrode studied by surface-enhanced
infrared absorption spectroscopy. Electrochim Acta 52:5715
36. Wieckowski A (1999) Interfacial electrochemistry: theory, experiment and applications.
Marcel Dekker, New York
37. Farias MJS, Herrero E, Feliu JM (2013) Site selectivity for CO adsorption and stripping on
stepped and kinked platinum surfaces in alkaline medium. J Phys Chem C 117:2903–2913
38. Fang L-L, Tao Q, Li M-F, Liao L-W, Chen D, Chen Y-X (2010) Determination of the real
surface area of palladium electrode. Chin J Chem Phys 23:543
39. Rees NV (2014) Electrochemical insight from nanoparticle collisions with electrodes. A mini-
review. Electrochem Commun 43:83
40. Rees NV, Zhou YG, Compton RG (2012) Making contact: charge transfer during particle-
electrode collisions. RSC Adv 2:379
41. Bard AJ, Xiao XY (2007) Observing single nanoparticle collisions at an ultramicroelectrode
by electrocatalytic amplification. J Am Chem Soc 129:9610
42. Kwon SJ, Fan FRF, Bard AJ (2010) Observing iridium oxide (IROx) single nanoparticle
collisions at ultramicroelectrodes. J Am Chem Soc 132:13165
43. Kwon SJ, Zhou H, Fan FRF, Vorobyev V, Zhanf B, Bard AJ (2011) Stochastic electrochem-
istry with electrocatalytic nanoparticles at inert ultramicroelectrodes—theory and experi-
ments. Phys Chem Chem Phys 13:5394
44. Zhou YG, Rees NV, Compton RG (2011) The electrochemical detection and characterization
of silver nanoparticles in aqueous solution. Angew Chem Int Ed 50:4219
45. Rees NV, Zhou YG, Compton RG (2011) The aggregation of silver nanoparticles in aqueous
solution investigated via anodic particle coulometry. ChemPhysChem 12:1645
46. Kahk JM, Rees NV, Pillay J, Tshikhudo R, Vilakazi S, Compton RG (2012) Electron transfer
kinetics at single nanoparticles. NanoToday 7:174
47. Blackmore CE, Rees NV, Palmer RE (2015) Modular construction of size-selected
multiplecore Pt-TiO2 nanoclusters for electro-catalysis. Phys Chem Chem Phys 17
(42):28005–28009. doi:10.1039/c5cp00285k
48. Newton JE, Preece JA, Rees NV, Horswell SL (2014) Nanoparticle catalysts for proton
exchange membrane fuel cells: can surfactant effects be beneficial for electrocatalysis? Phys
Chem Chem Phys 16:11435
49. Hsueh KL, Chin DT, Srinivasan S (1983) Electrode kinetics of oxygen reduction: a theoretical
and experimental analysis of the rotating ring-disc electrode method. J Electroanal Chem
Interfacial Electrochem 153:79–95
Chapter 2
Recent Advances in the Use
of Shape-Controlled Metal
Nanoparticles in Electrocatalysis

Francisco J. Vidal-Iglesias, José Solla-Gullón, and Juan M. Feliu

2.1 Introduction

The surface structure, that is, the particular arrangement of the atoms at the surface
is a very relevant parameter and recognized to strongly determine the
electrocatalytic properties of the material under study. In this regard, the extensive
and intensive knowledge acquired with the use of metal single crystals is not only of
outstanding importance for the establishment of the relationships between reactiv-
ity and surface structure but also should be used as a reference for those studies
dealing with the understanding and evaluation of the effect of the surface structure
on the electrocatalytic properties of metal nanoparticles [1–4]. However, this
knowledge should not be limited to foresee the expected results or tendencies but
also should be used to learn the methods, protocols and precautions that must be
satisfied to properly find such surface structure-reactivity correlations. In the quest
of these correlations, the use of shape-controlled metal nanoparticles is of outstand-
ing importance because the shape of a nanoparticle remarkably determines its
surface atomic arrangement and coordination. In fact, the shape of a nanoparticle
anticipates which surface facets will be present at its surface due to the intrinsic
correlation between the shape and the surface structure of a nanoparticle. As
described in previous contributions, for fcc metals, it is possible to build a stereo-
graphic triangle that correlates the crystalline surface Miller index and the shape of
the nanoparticle [5–7]. Thus, for instance, for a perfect Pt octahedron, all the
exposed facets would be {1 1 1}, while on a perfect Pt cube these facets would be
{1 0 0} domains. However, it is worth noting that these are ideal polyhedral crystals
containing perfect surface planes, which is, indeed, an unrealistic situation. There-
fore, the real surface of a nanoparticle will contain a determined number of defect,

F.J. Vidal-Iglesias • J. Solla-Gullón • J.M. Feliu (*)


Instituto de Electroquı́mica, Universidad de Alicante, Apdo. 99, Alicante E-03080, Spain
e-mail: [email protected]

© Springer International Publishing Switzerland 2016 31


K.I. Ozoemena, S. Chen (eds.), Nanomaterials for Fuel Cell Catalysis,
Nanostructure Science and Technology, DOI 10.1007/978-3-319-29930-3_2
32 F.J. Vidal-Iglesias et al.

corner, edge, step and kink sites and also ordered surface domains of different
dimensions. This complex scenario will particularly determine the resulting
electrocatalytic properties of the nanoparticle. Consequently, it is of capital rele-
vance to recall that, from an electrocatalytic point of view, the shape of a nanopar-
ticle is not the key point but its real surface structure. This fact clearly indicates that
having tools to characterize in great detail the particular surface structure of a
nanoparticle is of paramount importance to understand its intrinsic electrocatalytic
activity. This point will be discussed later in this chapter.
To the best of our knowledge, the first examples of shape-controlled metal
nanoparticles in electrocatalysis were reported by our group in 2004 [8, 9]. In
these contributions, we prepared cubic Pt nanoparticles using the pioneering meth-
odology described in 1996 by El-Sayed and co-workers [10] (this paper is consid-
ered to be the first example of synthesis of shape-controlled metal nanoparticles),
which consisted in a chemical reduction of K2PtCl4 by H2 in the presence of sodium
polyacrylate (NaPA). These cubic Pt nanoparticles were used to electrocatalyze
ammonia electrooxidation in alkaline solution and the resulting current density was
significantly higher than that found for spherical Pt nanoparticles. It is important to
stress that this reaction was selected because of its extreme sensitivity towards the
surface structure of single crystal platinum electrodes, taking place almost exclu-
sively on {1 0 0} sites. The enhanced activity was attributed to an increase in the
{1 0 0} terrace surface sites as deduced from the voltammetric profile obtained with
the cubic Pt nanoparticles in 0.5 M H2SO4. This experimental finding clearly
illustrated that controlling the surface structure of the nanoparticles was a rational
option to optimize their electrocatalytic activity. From these contributions, many
other cases have been reported in the literature and most of them appear summa-
rized and described in some relevant reviews about the use of these shaped metal
nanoparticles for electrocatalytic applications [11–22].
From the point of view of the synthesis of these nanoparticles, after the afore-
mentioned pioneering contribution by El-Sayed et al. [10], many different
approaches have been developed for the preparation of shape-controlled metal
nanoparticles. In this chapter, we are not going to go into details about this point
and the reader is directed to some of the most relevant reviews dealing with the
synthesis of such metal nanoparticles [20, 23–28]. Among many options, the use of
colloidal routes is, without any doubt, one of the most, if not the most, conventional
approach. This method is essentially based on the chemical reduction of a metallic
precursor in the presence of a capping agent. This capping agent is extremely
important in the process because it modifies the surface energies of the
nanoparticles during their nucleation and growth steps, allowing the thermody-
namic limitations to be overcome. As a result, metal nanoparticles with a particular
shape can be obtained in a controlled way. Obviously, the nature of the capping
agent is not the only parameter determining the resulting shape but also other
experimental parameters including temperature, electrolyte, reducing time, nature
of the reducing agent, and others, play an important role in defining the particular
shape finally obtained.
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 33

As previously stated, the presence of a capping agent is of capital importance for


the preparation of the shaped metal nanoparticles. However, if these nanoparticles
are going to be applied in an electrocatalytic application, the presence, even in
residual amounts, of such capping agents at the surface of the nanoparticles
represents, from our point of view, one of the most critical aspects and should be
evaluated with particular attention. In this sense, it is well-recognized from single-
crystal studies that clean surfaces are required for a correct understanding of the
relationships between surface structure and electrocatalytic surface reactivity. In
this sense, the flame annealing protocol developed by Clavilier in 1980 represented
the starting point that allowed not only ordered but also clean Pt single crystal
surfaces to be obtained and reproduced in a significant number of laboratories
worldwide [29]. This historical perspective points out the importance of working
with “clean” shaped nanoparticles. Consequently, once the shaped metal
nanoparticles are synthesized through appropriate colloidal routes, these must be
submitted to specific decontamination procedures in order to completely remove
the adsorbed species present at the surface of the nanoparticles. When this step is
not performed, the electrocatalytic activity of the nanoparticles is affected in a
non-controlled way which does not allow reproducible and comparable results to be
obtained. In this chapter, and due to its extreme importance, we will include a
particular section in which we will describe in detail some of the specific decon-
tamination protocols described in the literature and applied for shape-controlled
metal nanoparticles.

2.2 Surface Cleanness and Surface Site Determination

Once the importance of surface cleanness is understood, one critical question arises,
how should this surface cleanness be evaluated? In this regard, the detailed
knowledge gained with metal single crystals again gives us some interesting
options about how to proceed in this aspect. For instance, it is well established
that for Pt and Pd electrodes, the so-called hydrogen/anion adsorption-desorption
states are an extremely sensitive process for the evaluation of the level of cleanness
of the surface of such materials [30, 31]. Thus, when the different hydrogen/anion
adsorption-desorption states are well-defined (sharpness) and reversible (symme-
try), that is an indisputable proof of the correct cleanness of the surface. In addition,
these processes may also provide other relevant aspects such as (i) electroactive
surface area of the nanoparticles, the total adsorption or desorption charge being
directly proportional to the total amount of exposed surface atoms and
(ii) qualitative information about the nature and distribution of the different surface
sites present at the surface of the nanoparticles. This latter point is of great interest
because it allows the establishment of qualitative correlations between the shape
and the surface structure of the nanoparticles to be obtained. In fact, the resulting
voltammetric response given by a particular sample in a reference supporting
electrolyte can be taken as a fingerprint of its specific surface structure.
34 F.J. Vidal-Iglesias et al.

Consequently, from the analysis of this electrochemical process it is possible to


evaluate not only the level of cleanness of the samples but also their active surface
area and surface structure, which are basic parameters to be then correlated with
their electrocatalytic properties. Complementary site determination procedures
based on irreversibly adsorbed adatoms will be described below.
Similarly, the analysis of the oxide formation/reduction region can be also used
to extract information about the surface cleanness and also for the estimation of the
active surface area of Pt and Pd and also for Au electrodes. In addition, from the
voltammetric profile of the oxide region, it is also possible to evaluate the specific
surface structure of the sample. This analysis is based on previous contributions by
Hamelin [32, 33] dealing with Au single crystal electrodes and it has been applied
to characterize different shape-controlled Au nanoparticles [34, 35]. However, in
this case, and due to the high upper potential limit used, the surface stability of these
shaped Au nanoparticles could be compromised, thus resulting in a possible surface
restructuring. In fact, from the studies of the oxide formation on the gold single
crystal electrodes, it was established that gold oxidation/reduction cycles induces
the modification of the surface order of the electrodes [36, 37], and after several
oxidation/reduction cycles, the initially flat surface of the single crystal electrode
becomes rough. Consequently, and despite the changes induced by a single oxida-
tion cycle are smaller than those seen for Pt electrodes [38], this oxidation/reduction
treatment in the supporting electrolyte should be avoided.
Another interesting option for Au surfaces was described by El-Deab et al. [39–
41], following previous observations by Porter et al. with polyoriented Au elec-
trodes [42, 43]. They were able to estimate the fraction of the surface sites in
various gold nanostructures by using the reductive desorption of different thiol
compounds such as cysteine, mercaptoacetic acid or cystamine.
With a similar aim of determining the surface structure, other easy-to-do
approaches should be also highlighted. Among them, some underpotential deposi-
tion (UPD) processes are particularly relevant because these electrochemical reac-
tions are well-known to be very sensitive to the surface structure of the electrodes
[44]. In this sense, lead (Pb) UPD on Au [45–51] and copper (Cu) UPD on Pd [52–
54] and Pt [55, 56] surfaces have been reported to be very useful for the analysis of
their surface structure. In addition, these UPD processes can also be used to
calculate the real area of the electrode from the charge involved in the UPD reaction
and subsequently normalized with a specific charge density normalization (gener-
ally obtained from a polyoriented surface). These different approaches have been
very satisfactorily applied for different shape-controlled Au, Pt and Pd
nanoparticles. An additional advantage of all these measurements is that they are
performed in solution, i.e. the same environment in which the electrochemical
reactions will take place. For instance, Fig. 2.1 shows the Pb UPD in alkaline
solution of different shaped Au nanoparticles. As it is observed, the different peaks
are extremely sensitive towards the shape of the nanoparticle, and therefore towards
their surface structure. In particular, the results clearly show a preferential {1 0 0}
and {1 1 1} surface structure of the (C) cubic and (D) octahedral Au samples,
respectively. In addition, the spherical samples, and independently of their particle
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 35

Fig. 2.1 Pb UPD


voltammetric profiles of the
unsupported gold
nanoparticles in
Ar-saturated 0.1 M NaOH
containing 103 M Pb
(NO3)2: (a) spherical 5 nm
(b) spherical 30–40 nm (c)
cubic and (d) octahedral
nanoparticles. Scan rate:
50 mV s1

size, show a similar Pb UPD profile, which is characteristic of a polyoriented


surface structure.

2.2.1 Surface Active Sites Quantification

It is worth noting that these previous approaches are able to provide qualitative
information about the surface structure of the clean samples. However, it would be
much more relevant to have surface probes to quantitatively estimate the different
surface sites present at the surface of the nanoparticles. In this regard, to the best of
our knowledge, this has been exclusively achieved for shape-controlled Pt
nanoparticles using different approaches developed by our group [57]. The meth-
odology is based on a detailed analysis of the redox behaviour of some adatoms (Bi,
36 F.J. Vidal-Iglesias et al.

Te and Ge) spontaneously adsorbed at the surface of different Pt single crystals,


both ideally basal planes and series of vicinal stepped surfaces. Thus, whereas Bi
[58, 59] and Te [60] were sensitive to the presence of {1 1 1} terrace domains, Ge
[61] was sensitive to the {1 0 0} ordered domains. From the specific response of the
adatom for each Pt surface, calibration plots showing the charge density measured
under the adatom redox peak and the corresponding ({1 0 0} or {1 1 1}) terrace
density were obtained. The quantification of the {1 1 1} and {1 0 0} oriented
domains at the surface of the shape-controlled Pt nanoparticles was obtained
performing similar irreversibly adsorbed Bi and Ge experiments and then simply
using the normalization factors obtained in the calibration plots [57, 59]. This
methodology is currently used in different laboratories [62–65].
It is also interesting to note that from the information gained with these Bi and
Ge surface probes, we have also reported that other surface structure sensitive
reactions could be used as electroanalytical detection tools to monitor the presence
of some particular surface sites. Thus for instance, we showed that the desorption-
adsorption of hydroquinone-derived adlayers [66] and the ammonia electro-
oxidation [67] can be used to selectively quantify the {1 1 1} and {1 0 0} domains
at the surface of preferentially oriented Pt nanoparticles, respectively.
Additionally, from the knowledge of the amount of {1 1 1} ordered domains
obtained by using Bi or Te, it is also possible to determine the total amount of
{1 0 0} and {1 1 0} sites at the surface of the different Pt electrodes (both bulk
electrodes and nanoparticles) through pure deconvolution of the voltammograms
obtained in the so-called hydrogen region [57]. However, it is important to note
some differences between these two different approaches (Ge analysis vs
deconvolution). Thus while from deconvolution the total fraction of {1 0 0} surface
sites, that is, both step and terrace contributions can be estimated and even sepa-
rately evaluated, from Ge analysis only those {1 0 0} terrace domains with a terrace
width 4 Pt atoms (each Ge atom requires four platinum atoms to be adsorbed) can
be determined. In any case, as expected, for {1 0 0} terrace contributions, both
approaches give comparable measurements [57].
Finally, and before moving forward to the next section, it is also important to
recall that these previous approaches are convenient for pure metal nanoparticles
(Pt, Pd and Au). However, this kind of analysis becomes much more complicated
when Pt and/or Pd based alloys or even core-shell nanostructures are used. In this
regard, we would like to point out recent contributions by Stamenkovic et al. [68],
Shao et al. [69], and Strasser et al. [70] in which different protocols including
adsorption in the hydrogen region, CO stripping and Cu UPD charges are used,
discussed and compared for different Pt- and Pd-based alloy nanoparticles also
including shape-controlled alloy nanoparticles. From these contributions, one may
easily realize that a correct determination of the electrochemically active area on
these samples is not a trivial task and several reasonable considerations should be
taken into account. In addition, the scenario turns almost unexplored if detailed
information about the surface structure of these shape-controlled Pt- or Pd-based
alloy nanoparticles is needed. These are remaining questions that should be
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 37

addressed in the forthcoming years and that would allow a comparison of samples
prepared with different methodologies to be properly performed.

2.2.2 Surface Cleaning Methodologies

As previously stated, surface cleanness is a critical requirement. If the surface of the


nanoparticles is contaminated, it is not possible to reproduce and understand the
resulting electrocatalytic activity. Consequently, we have considered of interest to
specifically detail those protocols that have been proven to be effective for the
cleaning of shape-controlled metal nanoparticles prepared from different colloidal
routes, that is, in presence of different capping agents. However, it is worth noting
that only those that have unequivocally evaluated the surface cleanness of the
samples, using any of the surface cleanness proofs previously described, will be
incorporated in this section.
In our first contributions, cubic Pt nanoparticles were prepared in the presence of
NaPA [8, 9]. The removal of this NaPA was performed by simple addition of some
NaOH pellets to the colloidal suspension. After the addition of NaOH, the colour of
the solution changed from golden to black/grey and the Pt nanoparticles began to
slowly precipitate. After complete precipitation, the supernatant was discarded and
the nanoparticles were collected and transferred to a glass beaker where ultrapure
water was added and the nanoparticles redispersed. Once nanoparticles again
precipitated (no additional NaOH was required), the sample was again washed
with ultrapure water. This water washing was repeated at least two times after
which the sample was kept in water for the electrochemical measurements. The
cyclic voltammograms obtained in 0.5 M H2SO4 showed a very well-defined
voltammetric profile in the hydrogen region which clearly indicated the effective-
ness of the decontamination procedure. In subsequent contributions, we also dem-
onstrated that this decontamination protocol also resulted effective with
cuboctahedral and tetrahedral-octahedral [57, 71–73] Pt nanoparticles prepared
using a slightly modified route to that used for the cubic ones but also using
NaPA as capping agent. Figure 2.2 shows some representative voltammetric pro-
files of these clean shape-controlled Pt nanoparticles in different supporting elec-
trolytes (H2SO4, HClO4 and NaOH). It is straightforward to notice the different
current scales and deduce which test electrolyte is the most indicated to distinguish
the surface characteristics of the different samples.
Furthermore, we also reported that this decontamination procedure was appro-
priate to obtain clean cubic Pd nanoparticles prepared in presence of cetyltrimethy-
lammonium bromide (CTAB) [74–76]. It is worth noting that in all these cases, we
always performed a classical CO adsorption stripping step to maximise the surface
cleanness of the samples in the electrochemical environment. This CO adsorption-
stripping allows the removal of residual impurities on the surface of the
nanoparticles without altering their initial surface structure. This decontamination
38 F.J. Vidal-Iglesias et al.

0.5 M H2SO4 0.1 M HClO4 0.1 M NaOH


100 A 50 A 50 A
50
0 0 0
-50
-50 -50
-100
0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.0 0.2 0.4 0.6 0.8 1.0

100 B 50 B 50 B
50
0 0 0
j / µA cm-2

-50
-50 -50
-100
0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.0 0.2 0.4 0.6 0.8 1.0

100
C 50 C 50 C
50
0 0 0
-50
-50 -50
-100
0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.0 0.2 0.4 0.6 0.8 1.0

100 D 50 D 50 D
50
0 0 0
-50
-50 -50
-100
0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.0 0.2 0.4 0.6 0.8 1.0

E vs RHE / V
Fig. 2.2 Representative voltammetric profiles of (a) quasi-spherical, (b) cubic, (c) octahedral/
tetrahedral and (d) cuboctahedral Pt nanoparticles obtained in different supporting electrolytes at
50 mV s1. Reproduced with permission from [73], © 2012 American Chemical Society
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 39

protocol has been adapted in other labs and research groups in other contributions
[62, 77, 78].
In this sense, Coutanceau and co-workers [79, 80] also demonstrated that this
decontamination procedure was efficient for cubic Pt nanoparticles prepared in the
presence of tetradecyltrimethylammonium bromide (TTAB). The voltammetric
response of such cubic Pt nanoparticles was very similar to those cubic samples
prepared in presence of NaPA [57, 72]. Nevertheless, it is worth noting that this
protocol cannot be considered as a universal cleaning method. In fact, Tong and
co-workers used this decontamination procedure with shaped Pt nanoparticles
(cubic and octahedral/tetrahedral) [81–83] synthesized by one-step polyol-based
synthetic procedure, previously developed by Somorjai and coworkers [84], in
which polyvinylpyrrolidone (PVP) was used as capping agent. The voltammetric
responses of the samples showed insufficient definition and symmetry of the
adsorption/desorption peaks which was attributed to the presence of residual PVP
blocking the surface sites. Thermogravimetric analyses (TGA) were also performed
in order to determine the PVP content of the samples after the NaOH cleaning. The
PVP content was determined to be 16 and 11.3 wt% for the octahedral/tetrahedral
and cubic samples, respectively thus indicating that a significant amount of PVP
still remained at the surface of the nanoparticles. Similarly, Coutanceau and
co-workers were also unable to clean similar shaped Pt nanoparticles [79] to
those prepared by Tong and co-workers. However, more recently, Tong and
co-workers have developed a new protocol to obtain PVP-free cubic and octahe-
dral/tetrahedral Pt nanoparticles using an adapted liquid phase UV photo-oxidation
(UVPO) technique [65]. In brief, the PVP-capped Pt nanoparticles were initially
centrifuged with acetone to remove the excess ethylene glycol and PVP, until the
nanoparticles were clearly separated. The resulting Pt nanoparticles were then
dispersed in an ethanol-hexane mixture and again centrifuged. This ethanol-hexane
cleaning was repeated at least three times and finally the sample was dispersed in
ethanol. Furthermore, and based on previous contributions, a modified procedure of
liquid phase UV combined with oxygenated H2O2 was applied to remove the
residual PVP present at the surface of the nanoparticles. In detail, a small portion
of the ethanolic-suspension containing the nanoparticles was dispersed into about
2 mL of H2O with 200–250 mg NaOH to yield an alkaline environment. On the
other hand, a basic and O2-saturated H2O2 solution was prepared by adding
200–250 mg NaOH to a 10 mL H2O2 solution which was then bubbled for at
least 10 min with O2 gas. The basic solution containing the nanoparticles was UV
irradiated (254 nm) for 1 h while about 1 mL of O2-saturated H2O2 solution was
added to the solution every 10 min. After the UV exposure, the solution was finally
purified by repetitive centrifugation and precipitation process with ethanol. The
voltammetric responses of the shaped Pt nanoparticles in 0.5 M H2SO4, Fig. 2.3,
were rather similar to those previously reported with similar and clean cubic and
octahedral/tetrahedral Pt nanoparticles prepared by our group (see Fig. 2.2). In
addition, TGA exhibited no major weight loss due to the desorption of PVP, thus
suggesting its absence at the surface of the nanoparticles.
40 F.J. Vidal-Iglesias et al.

Fig. 2.3 (a) Voltammetric profiles of the Pt black (black), cubic (red) and octahedral/tetrahedral
(blue) Pt nanoparticles in 0.5 M H2SO4. Comparison of CV profiles of pristine (solid line) cubic
(b) and octahedral/tetrahedral (c) Pt nanoparticles with CVs after germanium (dotted line) and
bismuth (dashed line) adsorption in 0.5 M H2SO4. Reproduced with permission from [65], © 2014
Royal Society of Chemistry

However, this methodology was a modified protocol based on a previous


contribution by Koper and co-workers in which PVP-protected Pt nanoparticles
were cleaned using a H2O2/H2SO4 solution [85]. Interestingly, the effectiveness of
the methodology was also tested with a Pt(1 1 1) single crystal electrode which
demonstrated that the surface was decontaminated without a loss of its initial
crystalline surface structure. In addition, they suggested that the decontamination
of the surface was in fact a physical removal process of the PVP induced by the
oxygen bubbling produced during the decomposition of the H2O2. As far as we
know, this contribution was the first example of effective cleanness with
nanoparticles prepared in the presence of PVP.
More recently, Yang et al. have proposed a new procedure for removing
different capping agents commonly used for the synthesis of shape-controlled Pt
nanoparticles such as PVP and oleylamine/oleic acid, by a simple electrochemical
potential cycling in 0.5 M NaOH [86]. The effectiveness of this protocol was
proven by using Pt nanocubes prepared in the presence of such capping agents.
The obtained particles were washed by centrifugation with a saturated ethanol
solution of NaOH for those nanoparticles prepared in presence of PVP, or in
hexane/ethanol mixture for those ones prepared with oleylamine/oleic acid, and
finally redispersed in ethanol and hexane, respectively. The different nanoparticles
were deposited on a gold electrode and cycled between 0 and 1.0 V (vs RHE)
in 0.5 M NaOH with a scan rate of 0.5 V s1. After 100 cycles, the electrodes were
transferred to a second electrochemical cell containing 0.5 M H2SO4 where the
cleanness was evaluated. The voltammetric response obtained clearly suggested
that this electrochemical potential cycling in 0.5 M NaOH solution is an effective
way to remove both capping agents, PVP and oleylamine/oleic acid from the
surface of the nanoparticles. Unfortunately, this cleaning method has some practical
limitations, particularly if large amounts of nanoparticles are needed. In
this regard, we have recently developed a new decontamination procedure to obtain
a new collection of clean shape-controlled Pt nanoparticles prepared using
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 41

Fig. 2.4 Voltammetric profiles of Pd (a) nanospheres, (b) nanocubes and (c) nanooctahedra in
0.5 M H2SO4 at 5 mV s1. The insets present TEM images of the Pd nanoparticles. Reproduced
with permission from [89], © 2014 American Chemical Society

oleylamine/oleic acid as capping material/solvent [87]. The method is not based on


electrochemical cleaning but in a chemical one which allows important amounts of
Pt nanoparticles to be cleaned.
Coutanceau and co-workers [88, 89] have recently reported that clean shaped
controlled Pd nanoparticles prepared in the presence of PVP can be obtained using a
NaOH treatment similar to that unsuccessfully used in previous contributions
[79, 81–83]. From our point of view, the main differences of this treatment in
comparison with the previous ones seem to deal with the required NaOH concen-
tration and the addition of a particular amount of water. In more detail, the resulting
colloidal suspension containing the shaped Pd nanoparticles was diluted with H2O
after which NaOH pellets were added to have a 1 M NaOH solution (this NaOH
concentration is significantly higher than that previously used). The nanoparticles
were collected by precipitation. This protocol was repeated until clean CV
voltammetric profiles were obtained, Fig. 2.4.
Neergat and co-workers have recently proposed two new decontamination
procedures for the removal of adsorbed stabilisers and capping agents used for
the preparation of shape-controlled Pd nanoparticles [90, 91]. The shaped Pd
42 F.J. Vidal-Iglesias et al.

nanoparticles were prepared with PVP also in the presence of other additives such
as Br, Cl and citrate ions. After some preliminary cleaning steps for the removal
of the excess of chemicals used during the syntheses, the samples were then treated
with solutions containing tert-butylamine [91] or NaBH4 [90] for a particular time,
after which were collected by centrifugation, washed with different solvents (eth-
anol or water) and finally dried. The samples were electrochemically characterized
in 0.1 M HClO4 and the response of the treated samples was clearly much better in
comparison with the as-prepared ones. This effective surface cleaning also allowed
better ORR activities to be recorded as a consequence of the improved surface
cleanness.
Another interesting option that has been applied for shape-controlled Au
nanoparticles is based on the electrochemical deposition of a film of PbO2 in
alkaline solution during the Pb UPD experiments [45–47, 92]. This methodology
was proven to be very effective for the removal of different capping agents
(polyethylene glycol dodecyl ether (Brij®30) and CTAB) without altering the initial
surface structure of the samples. This fact was demonstrated by testing the protocol
with both single crystal and polyoriented Au electrodes.

2.2.3 Surface Cleanness Versus Surface Disordering

As previously stated, surface cleanness is a fundamental requirement to properly


correlate the surface structure and reactivity of shape-controlled metal
nanoparticles. However, this surface cleaning must be performed without
perturbing the intrinsic/initial surface structure of the nanoparticles. To illustrate
this situation, we will discuss the use of UV/ozone as decontamination protocol.
With the aim of evidencing the importance of an effective surface cleaning,
Somorjai and co-workers evaluated the use of UV/ozone and its effects on the
catalytic properties of some shape-controlled Pt nanoparticles [93–95]. Interest-
ingly, they observed a remarkably increase in the catalytic activity after the UV/
ozone procedure which was justified in terms of a more efficient cleanness of the
samples and consequently enhanced catalytic properties. Also, they demonstrated
by TEM measurements that both shape and size of the UV/ozone treated
nanoparticles remained unmodified in comparison with the untreated samples. It
should be stressed that electrocatalysis depends on the real number of clean atoms
of the catalyst in contact with the solution, that is, the more clean atoms, the higher
the activity. A similar UV/ozone treatment was also used by Kiwi-Minsker to clean
PVP-stabilized Pd nanocubes [96]. They also reported improved catalytic activities
and, based on high-resolution scanning electron microscopic images, they also
stated that the Pd nanocubes were morphologically stable. However, as previously
stated, the shape of the nanoparticles is not the key parameter determining the
resulting catalytic activity but their surface structure at atomic level and this is
important for particular reactions. In fact, our group demonstrated that this UV/
ozone treatment strongly perturbed the surface structure of the nanoparticles but, as
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 43

Fig. 2.5 Voltammetric profiles before (black) and after 30 (red), 60 (green), 90 (blue), 120 (cyan)
and 150 (magenta) minutes of UV/ozone exposure. (a) cubic nanoparticles, (b) tetrahedral–
octahedral nanoparticles. Test solution: 0.5 M H2SO4. Sweep rate 50 mV s1 Reproduced with
permission from [97], © 2011 Elsevier

previously shown by Somorjai and co-workers, without altering neither the size nor
the shape of the nanoparticles [97]. As a consequence of the perturbation of
arrangement of the atoms at the surface of the nanoparticles, in particular those
present as terrace domains, the electrocatalytic properties of the samples were
drastically modified as illustrated in Fig. 2.5. It should be pointed out that in all
cases the surfaces were equally clean, but after the ozone treatment, the surface
atoms were less ordered and some particular reactions only take place on ordered
domains.
44 F.J. Vidal-Iglesias et al.

2.2.4 Can the Residual Presence of Capping Agents Have


Any Beneficial Effect in Electrocatalysis?

Although this aspect is not going to be analysed in the present chapter, it is


important to point out the existence of a number of interesting contributions that
claim that a determined presence of certain species, including some capping agents,
which are usually regarded as poisoning species because block the active
electrocatalytic sites and slow reaction kinetics, may provide beneficial effects in
terms of electrocatalytic properties. These effects point out to a third-body role of
the surface residues, unfortunately uncontrolled and likely not stable. For those
readers interested in this topic we refer to [98–107].
Finally, and before moving forward to the next section, it is also worth noting
that shape-controlled metal nanoparticles can also be prepared in the absence of
capping agents. In this regard, the electrochemical approach developed by Sun and
co-workers is one of the most interesting methodologies [5, 108, 109]. Other
interesting approaches are surfactant-free solvothermal synthesis [110–112] or
solid-state chemistry methods [113].

2.3 Electrocatalytic Properties of Shape Controlled Metal


Nanoparticles

In this section, we will summarize some of the most relevant advances on the use of
these shaped metal nanoparticles in electrocatalysis and, in particular, towards
reactions of interest for low temperature fuel cells such as formic acid, methanol
and ethanol electrooxidations and oxygen reduction. However, only those contri-
butions that, from our point of view, fulfil the particular requirements of surface
cleanness and surface structure–electrochemical reactivity correlation will be
discussed.

2.3.1 Formic Acid Electrooxidation

Formic acid electrooxidation has been the subject of innumerable contributions, not
only because it is a model reaction for a two-electron-transfer reaction, but also for
its promising use as fuel in direct formic acid fuel cells (DFAFCs) [3, 114–116]. It
is widely accepted that the oxidation of formic acid takes place through a dual path
mechanism [117–119]: a direct pathway via an adsorbed active intermediate, the
nature and role of which are still under strong discussion at the fundamental level
and that readily yield to CO2, and a second pathway involving the formation of
adsorbed CO, which is considered to be a poisoning intermediate, that can be
further oxidized to CO2 at high overpotentials. On Pt surfaces, it is well recognized
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 45

that both reaction paths are structure sensitive, being the Pt(1 0 0) surface the most
active (in the absence of CO poisoning), but also the most poisoned [115]. The
remarkable structure sensitivity of this electrochemical process has been also
illustrated with shape-controlled Pt nanoparticles. Tian et al. prepared
tetrahexahedral (THH) Pt nanocrystals with enhanced electrocatalytic activity
towards formic acid electrooxidation using a novel electrochemical approach
[108]. The electrocatalytic activity of the THH Pt nanoparticles was always higher
than those observed for polycrystalline Pt nanospheres and commercial Pt/C cata-
lyst (E-TEK Co., Ltd), independently of the applied potential range. Interestingly,
the best enhancement factors (about four and three times higher than for the
nanospheres and the commercial catalyst, respectively) were observed at low
potential values of about 0.1 V vs SCE. This contribution was the first synthesis
of Pt nanoparticles enclosed with high-index facets and clearly evidenced the great
potential of the metal nanoparticles containing high-index facets for some
electrocatalytic applications. Some of them will be discussed in this chapter.
Taking into account the previous studies with Pt single crystals in which the Pt
(1 0 0) surface was shown to be the most active basal plane, we evaluated the
response of different shape-controlled Pt nanoparticles. Thus, spherical, cubic,
cubooctahedral and tetrahedral-octahedral Pt nanoparticles were prepared and
electrochemically characterized [72]. From the voltammetric profiles of the sam-
ples we deduced that the surface structure of these samples was, preferentially
polyoriented, (1 0 0), (1 0 0 )-(1 1 1) and (1 1 1), respectively, as expected from the
stereographic triangle correlation between shape of the nanoparticle and the surface
structure of an fcc metal. However, as previously discussed, the results also
evidenced that the nanoparticles did not contain a unique type of surface site but
that all surface sites were observed, despite having a clear preferential surface
structure. In terms of activity towards formic acid electrooxidation, the results
obtained perfectly agreed with those obtained with the model Pt surfaces and the
different samples showed a similar behaviour to that of their corresponding Pt
model surfaces. Thus, for instance, in similarity with a Pt(1 0 0) electrode, those Pt
particles having a preferential (1 0 0) orientation showed the best activity in the
negative going sweep of the cyclic voltammetric measurements but very low
activity during the chronoamperometric experiments as consequence of their fast
CO poisoning.
Using similar shape-controlled Pt nanoparticles, we re-evaluated their activity
towards formic acid electrooxidation but in this case using pulsed voltammetry
[120]. This technique is very useful to independently study both reaction paths and
allows the intrinsic activities and the poisoning rates to be determined. Figure 2.6
shows the activities of the (1 0 0) and (1 1 1) preferentially oriented Pt nanoparticles
obtained at t ¼ 0 (intrinsic activity) corresponding with the activity of the electrode
through the active intermediate reaction path in the absence of poison, and that
obtained for t ¼ 1 s after which the surface poisoning is already observed.
Huang et al. [121] prepared concave Pt nanocrystals having {4 1 1} high-index
facets by using amines as capping agents. The electrocatalytic properties of these
samples towards formic acid electrooxidation showed that the activity (normalized
46 F.J. Vidal-Iglesias et al.

Fig. 2.6 Intrinsic activities E / V vs RHE


(t ¼ 0 s) (black) and current 0.0 0.2 0.4 0.6 0.8 1.0
densities after t ¼ 1 s (red) 8
obtained from pulse
voltammetry transients for 7 A
cubic (a) and octahedral/ 6

-2
tetrahedral (b) Pt

j / mA·cm
nanoparticles in 0.5 M 5
H2SO4 þ 0.1 M HCOOH. 4
Reproduced with
permission from [120], 3
© 2010 American Chemical 2
Society
1
0

3.0 B
-2
j / mA·cm

2.5

2.0

1.5

1.0

0.5

0.0
0.0 0.2 0.4 0.6 0.8 1.0
E / V vs RHE

to the electrochemically active surface and measured at 0.61 V vs SCE) was two
and five times higher than that obtained with a commercial Pt black and Pt/C,
respectively. In addition, these concave Pt nanocrystals showed an excellent sta-
bility after the electrochemical experiments.
Li et al. [122] reported the preparation of trapezohedral (TPH) Pt nanocrystals
enclosed by {5 2 2} high-index facets using a similar electrochemical approach to
that developed by Tian et al. [108]. These nanocrystals were tested for different
reactions of interest including formic acid electrooxidation. The peak current
density (normalized to the electroactive surface area estimated from the so-called
hydrogen region) of formic acid electrooxidation at 0.60 V in the positive potential
scan was observed to be about three times higher than that obtained with commer-
cial Pt/C samples.
Zhang et al. [123] prepared Pt concave nanocubes with high-index {h k 0} facets
using glycine as surface modifier. These nanoparticles showed enhanced specific
activities (again normalized to the electroactive surface area estimated from the
so-called hydrogen region) towards formic acid electrooxidation in comparison to
commercial Pt black and Pt/C catalysts. Unfortunately, the samples were irradiated
with a UV lamp (185 and 254 nm) for 12 h in air in order to remove the capping
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 47

agents. This cleaning methodology was shown to strongly perturb the surface
structure of the samples [97].
Xia et al. [124] prepared highly concave Pt nanoframes with high-index facets
using an oleylamine-assisted solvothermal method. The resulting Pt nanoframes
exhibited a reasonable electrocatalytic activity and an improved electrochemical
stability for different reaction of interest including formic acid electrooxidation.
More recently, Korzeniewski et al. reported an interesting surfactant-free
solvothermal method for the preparation of <10 nm shape-controlled Pt
nanoparticles [111]. The shape of the different Pt nanoparticles (truncated octahe-
dral, cuboctahedral or cubic) was essentiality determined by the water volume
fraction (optimal 3 %) in N,N-dimethyformamide (DMF) and the reaction time.
Both the voltammetric profiles obtained in 0.5 M H2SO4 and the responses towards
formic acid electrooxidation were similar to those reported in our previous works
[57, 72].
The influence of the shape/surface structure on Pd nanoparticles towards formic
acid electrooxidation has been studied more extensively. Interestingly, on Pd
surfaces the dehydration step does not take place (CO is not formed) and conse-
quently the reaction directly proceeds through the direct path. Moreover, the onset
potential for the oxidation is ca. 200 mV lower than that observed for Pt [125–
127]. However, from single crystal studies, it was also shown that the reaction is
structure sensitive, being the Pd(1 0 0) the most active surface among the basal
planes, both in sulphuric and perchloric acid solutions [128, 129]. As a conse-
quence, the preparation of shape-controlled Pd nanoparticles and particularly cubic
nanoparticles in which the (1 0 0) orientation is maximised, has been a subject of
great interest. From our point of view, the first relevant contributions were reported
by Jin et al. [130] and by Zhang et al. [131]. Jin et al. prepared different Pd
nanoparticles including cubes, truncated cubes, cuboctahedra, truncated octahedra,
and octahedra. That is, Pd nanoparticles containing different fractions of {1 0 0}
and {1 1 1} surface domains. As expected from Pd single crystal studies, the activity
systematically improved for those nanoparticles having a higher fraction of {1 0 0}
domains, being the cubic sample the most active one, Fig. 2.7. Cu UPD was used in
this case for the determination of the electroactive surface area of the nanoparticles.
Similarly, Zhang et al. synthesized rhombic dodecahedral and cubic Pd
nanoparticles using a simple one-pot synthesis in the presence of CTAB as capping
agent and KI as additive [131]. The voltammetric profiles of these nanoparticles
both in H2SO4 and in NaOH solutions showed characteristic features corresponding
to preferentially oriented {1 1 0} and {1 0 0} Pd nanoparticles, respectively. The Pd
nanocubes again exhibited a higher electrocatalytic activity toward formic acid
electrooxidation.
Almost in parallel with these contributions, we also studied the electrocatalytic
properties of spherical and cubic Pd nanoparticles towards formic acid
electrooxidation both in H2SO4 and HClO4 solutions [76]. Interestingly, we dem-
onstrated that, to properly compare the intrinsic electrocatalytic properties of both
nanoparticles, it was required to minimize the amount of samples during the
measurements to avoid diffusional problems of reactants and products that result
48 F.J. Vidal-Iglesias et al.

0.65

Peak potential (V)


0.60

150 0.55

0.50
jmax (mA/cm2)

0.45

Cube Octahedron
100

50

0
Catalysts

Fig. 2.7 Maximum current densities for formic acid electrooxidation (normalized to the electro-
chemical surface area) for different shaped Pd nanoparticles in 0.1 M HClO4 þ 2 M HCOOH at
10 mV s1. The inset shows their corresponding peak potentials. Reproduced with permission
from [130], © 2012 Royal Society of Chemistry

in a lower apparent activity. Under these controlled conditions, we observed that the
activity of the Pd nanocubes was remarkably higher than that found for the
spherical nanoparticles. In addition, we reported that the activity found in HClO4
was significantly higher than that obtained in H2SO4 in the whole potential range
thus pointing out the competitive effect of the adsorbed anions towards formic acid
oxidation at these surfaces.
From these initial contributions, new interesting approaches have been reported.
Shao et al. prepared different concave Pt nanocrystals (nanocubes, 5-fold twinned
nanorods and right bipyramids) [132]. These concave nanocrystals showed
enhanced electrocatalytic activity (normalized to electrochemically active surface
area estimated from Cu UPD measurements). Kuo et al. studied the effect of lattice
strain on the catalytic properties of some shape-controlled Pd nanoparticles
[133]. They prepared core–shell AuPd nanocubes and nanooctahedra, and twinned
Pd icosahedra that showed lattice strain, detected by X-ray diffraction measure-
ments, due to the intrinsic twinning in the Pd icosahedra and the lattice mismatch
between the Au and Pd shells in the AuPd core–shell particles. The electrocatalytic
properties of these strained nanoparticles towards formic acid electrooxidation
were compared with those obtained with Pd nanocubes and nanooctahedra. The
strained nanoparticles showed enhanced activities due to the upshift of the d-band
center of Pd as consequence of the lattice expansion.
Xia and co-workers also prepared different shape-controlled Pd nanoparticles
including cubes, octahedra, icosahedra and right bipyramids which were used as
electrocatalysts towards formic acid electrooxidation [134–136]. Unfortunately,
and due to the different experimental conditions used in the electrocatalytic
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 49

experiments, it is not possible to establish a clear tendency. However, we would like


to extract a controversial point from one of these contributions. Shao et al. [136]
stated that the use of shape-controlled Pd nanoparticles did not produce any
advantage in comparison with conventional Pd nanoparticles. From our point of
view, this fact can be explained in terms of an insufficient cleanness of the
nanoparticles. In fact, the reported voltammetric profiles are remarkably different
to those obtained by Coutanceau et al. [88, 89] using the same synthetic protocol but
after a careful surface cleaning of the resulting Pd nanoparticles. This is a clear
example of how important the surface cleanness is for electrocatalytic
measurements.
Zhang et al. evaluated the electrocatalytic properties of different-shaped and
monodispersed (50 nm) palladium nanocrystals, including cubes, octahedra and
rhombic dodecahedra, towards formic acid electrooxidation both in HClO4 and
H2SO4 solutions [137]. The voltammetric responses, in the absence of formic acid,
of the different samples showed some of the characteristic features observed with
Pd single crystals, which warranties the level of cleanness of the nanoparticles. In
both supporting electrolytes, the cubic Pt nanoparticles showed the highest activity
as a result of their preferential {1 0 0} surface structure. In addition, the activities
reported in HClO4 were higher than those obtained in H2SO4 due to the
corresponding competitive adsorption of anions. These results were in full agree-
ment with previous observations reported by our group [76].
More recently, Zhang et al. [138] prepared cubic, cuboctahedral and octahedral
Pd nanoparticles after HCl oxidative etching, occurring at the corners of cubic Pd
nanoparticles. The cubic Pd nanoparticles again showed the highest activity
towards formic acid electrooxidation.
Tang et al. have recently reported the synthesis of length tuneable penta-twinned
Pd nanorods and their activity toward formic acid electrooxidation in comparison
with cubic Pd nanoparticles [139]. Before the electrochemical testing, the different
samples were subjected to an electrochemical potential cycling between 1.2 and
þ0.6 V vs Ag/AgCl at 500 mV s1 and for 100 cycles in a CO-saturated 0.1 M
NaClO4 þ 1 mM NaOH solution. Despite they claimed that the resulting CVs were
characteristic of clean surfaces and similar to those reported in previous contribu-
tions, we are not in agreement with this assumption and consider that neither the
surface cleanness nor surface structure preservation were demonstrated. These are,
from our point of view, critical aspects that must be clearly shown.
It is well-known that the incorporation of a second element both on Pt and Pd
surfaces is an interesting approach to improve the electrocatalytic properties of
these surfaces towards formic acid electrooxidation [116]. This improvement has
been explained in terms of third-body, electronic and bifunctional effects. Third-
body effects take mainly place when the surfaces are directly modified by the
inclusion of a second element that acts as a mere spectator, and only blocks surface
sites, diminishing the activity for an undesired reaction. However, electronic effects
occur when the modified surface also changes its electronic properties. In this latter
point, the so-called d-band center model [140] is widely accepted and explains the
possible modifications of the electronic properties and binding energies of a
50 F.J. Vidal-Iglesias et al.

transition metal surface, after the incorporation of a second metal, and their
correlations with the resulting catalytic activity. Finally, bifunctional effects appear
when the second metal contributes to the reaction providing a required group at
lower potentials, which facilitates the reaction mechanism.
For shape-controlled metal nanoparticles, two main strategies have been
employed (i) deposition of different adatoms at the surface of the shaped
nanoparticles and (ii) preparation of shape-controlled Pt or Pd alloy based
nanoparticles. In the following, we will describe some of the most relevant contri-
butions in these two approaches.
Our group has intensively worked on the use of adatoms to selectively decorate
the surface of different shape-controlled Pt nanoparticles towards formic acid
electrooxidation. Following previous results obtained with adatom decorated Pt
single crystals [115, 141], we have used Bi [142, 143], Pd [144, 145], Sb [146] and
more recently Tl [147] as surface modifiers. The reported results showed that
improved electrocatalytic activity can be obtained in all cases, although the
enhancement factor strongly depends on the surface structure of the substrate and
the nature of the adatom. In particular, the systems Pd/Ptcubic, Bi/Ptoctahedral,
Sb/Ptoctahedral and Tl/Ptcubic were the most convenient combinations. Nevertheless,
it is important to note that the maximum oxidation current values, as well as the
optimal potential region, are different in these systems. Thus, for instance, whereas
the system Pd/Ptcubic showed a clear shift of the oxidation process towards much
lower potential values (maximum currents at 0.3 V) but without a great enhance-
ment in terms of maximum oxidation currents, the system Bi/Ptoctahedral provided an
important benefit in relation to the maximum oxidation current but at similar
potential values (about 0.6 V) to that of the unmodified samples.
In collaboration with the Sun group, we also demonstrated that this strategy also
provided interesting results with high-index facet Pt nanoparticles [142]. In this
way, THH Pt nanocrystals prepared using the electrochemical approach developed
by Tian et al. [108] were decorated with Bi adatoms and their electrocatalytic
activity was evaluated for the formic acid electrooxidation. The results obtained
again led to an enormous enhanced activity in comparison with the bare THH Pt
nanocrystals. In voltammetric experiments an enhancement factor of about
20 (peak current density) was found for the highest Bi coverage (θBi¼0.9). In
addition, chronoamperometric measurements also showed a very significant
enhancement that varied from 65 to 1.5 depending on the electrode potential and
Bi coverage.
Later, Sun’s group published a new contribution in which Au-decorated THH Pt
nanocrystals were used towards formic acid electrooxidation [148]. The gold
decoration was performed through an initial sub-monolayer Cu UPD which was
then exchanged with Au adatoms by simple galvanic replacement. The presence of
Au adatoms clearly inhibited the formation of CO trough the dehydration step
although, in comparison with the Bi modification, did not provoke an enhanced
electrocatalytic activity. This is due to the fact that while in the case of Au
decoration, only a third-body effect is occurring, for Bi decoration, where the
peak current is more than five times higher than that of Au-modified THH Pt,
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 51

Fig. 2.8 (a) Positive scan of the cyclic voltammograms for 0.05 ML, 0.5 ML, 3.17 ML, 5 ML
samples, and commercial Pt/C catalyst obtained in 0.5 M HCOOH þ 0.1 M H2SO4 solution at
50 mV s1, (b) comparison of the mass activities measured at 0.32 V and poisoning ratios (current
density at 0.67 V divided by the current density at 0.32 V). Reproduced with permission from [35],
© 2013 American Chemical Society

both third-body effect and electronic (ligand) effect are contributing to the enhance-
ment activity.
Yang and Lee reported a very interesting approach dealing with the epitaxially
deposition of Pt on gold octahedral nanocrystals enclosed by {1 1 1} facets
[35]. The amount/coverage of Pt was finely controlled from fully covered multiple
overlayers (about five monolayers) to atomically dispersed submonolayer (0.05
monolayer). Very interestingly, the electrocatalytic activity of Pt modified gold
octahedra remarkably increased for decreased Pt coverages. For 0.05 monolayer of
Pt a spectacular mass activity of about 63 A mgPt1 was found. This activity was
about 120 and 170 times higher than those obtained with 5 ML and with a
commercial Pt/C catalyst, Fig. 2.8. Such enormous enhancement was attributed to
a third body (lack of ensembles) and bifunctional effects at the Pt-Au sites.
Several shape-controlled Pt and Pd alloy nanoparticles have been tested for the
electrooxidation of formic acid. Wang et al. synthesized PdPt alloy nanocubes with
tunable compositions [149] and their electrocatalytic activity towards formic acid
electrooxidation was compared with those obtained with Pd nanocubes and a
commercial Pd black. They observed that the activity of the PdPt nanocube alloys
was, independently of their composition, higher than those of the Pd nanocubes and
commercial Pd black. In particular, the Pd74.4Pt25.6 alloy sample showed the highest
formic acid oxidation activity which was about two times higher than that observed
for the Pd nanocubes and commercial Pd black.
Fang’s group has reported different contributions about the use of shape-
controlled Pt alloy nanoparticles as electrocatalysts towards formic acid
electrooxidation including Pt3Fe nanocubes [150], PtCu nanocubes of different
atomic composition (PtxCu100x (x ¼ 54–80 at.%)) [151, 152] and Pt-Cu
nanooctahedra [153, 154]. In addition, Fang et al. have recently published two
interesting papers in which the electrocatalytic activity-enhancement of different
shape-controlled Pt-based bimetallic nanocrystals, also including high-index noble
metal nanostructures, has been studied [155, 156].
52 F.J. Vidal-Iglesias et al.

Kang and Murray reported the synthesis and electrocatalytic properties of PtMn
nanocubes [157]. The electrocatalytic properties of the PtMn nanocubes (including
oxygen reduction and formic acid and methanol oxidation reactions) were evalu-
ated and compared with that obtained with spherical PtMn nanoparticles, with a
commercial E-TEK Pt catalysts and with a Pt black. For formic acid oxidation both
cubic and spherical PtMn nanoparticles were found to be less active (currents were
normalized to surface areas estimated from the charge of the hydrogen adsorption-
desorption region) than commercial ETEK Pt/C catalyst, although the cubic PtMn
nanoparticles showed higher activity than the spherical ones.
Yu et al. [158] synthesized polyhedral AuPd core-shell structures (60–80 nm)
having high-index facets (concave trisoctahedral (TOH) and hexoctahedral (HOH)
crystals both with {h k l} facets and THH crystals with {h k 0} facets)). Concave
TOH gold nanocrystals were used as seeds onto which Pd layers were heteroepi-
taxially grown. Their activity for formic acid electrooxidation was also evaluated
and compared with that obtained with cubic and octahedral AuPd core-shell
nanocrystals enclosed by low-index {1 0 0} and {1 1 1} facets. The maximum
current density of formic acid oxidation obtained in the positive going-scan
followed the order of octahedra < TOH < HOH < cubes ffi THH.
Zhang et al. [159] obtained HOH AuPd alloy nanoparticles with 48{hkl} facets
with a uniform size distribution (~55 nm). Interestingly, these samples showed an
activity towards formic acid electrooxidation about five times higher than that
obtained with a commercial Pd black catalyst, which was attributed to a synergy
effect of the high-index facets and the AuPd alloy composition.
Deng et al. [160] prepared THH PdPt alloy nanocrystals enclosed by {10, 3, 0}
high-index facets using the electrochemical approach [108]. At the optimal atomic
composition (Pd90Pt10) the THH nanocrystals exhibited a catalytic activity towards
formic acid electrooxidation in perchloric solution that was about three times higher
than that on THH Pd nanocrystals, and six times higher than on commercial Pd
black catalyst, Fig. 2.9. The enhanced activity was again related to the synergy
effect of high-index facets and particular electronic structure of the alloy formed.
Zhang et al. [161] prepared different shape-controlled PtPd bimetallic alloy
nanoparticles including PtPd cubes, bars, flowers, concave cubes and dendrites
through a one-pot solvothermal synthesis and using different capping agents,
surface modifiers and solvents. The electrocatalytic properties of all these PtPd
alloy nanoparticles toward formic acid oxidation were higher than those obtained
with a commercial Pt black and a Pt/C catalyst. Among the shape-controlled PtPd
bimetallic alloy nanoparticles, flowers and dendrites provided the best activities.
Nevertheless, the samples were not only irradiated with an UV lamp for 12 h to
remove the organic capping agents but also cycled (about 50 cycles) between 0.24
and 1.0 V vs SCE in 0.5 M H2SO4 before use. As previously discussed in this
chapter, these steps strongly perturb the surface structure at atomic level of the
samples.
More recently, Jia et al. [162] described the preparation of novel excavated
rhombic dodecahedral (ERD) PtCu3 alloy nanoparticles containing ultrathin
nanosheets of high-energy {1 1 0} facets. These ERD PtCu3 nanoparticles showed
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 53

a 80 b
jp jp
Pd0.94Pt0.06
60 Pd0.92Pt0.08
60 j@0v
j / mA cm-2

j / mA cm-2
Pd0.90Pt0.10
j@0v Pd0.88Pt0.14
40 Pd0.82Pt0.18 40
Pd THH
Pd black
20 20

0
0

06

08

10

14

18
k

0.

0.

0.

0.

0.
-0.2 0.0 0.2 0.4 0.6 0.8

ac

Pt

Pt

Pt

Pt

Pt
TH
bl

94

92

90

86

82
Pd

0.

0.

0.
Pd

0.

0.
E / V (SCE)

Pd

Pd

Pd

Pd

Pd
Fig. 2.9 Comparison of electrocatalytic activities of PdPt and Pd THH nanoparticles and commer-
cial Pd black towards formic acid oxidation. (a) Current–potential curves recorded at 50 mV s1 in
0.25 M HCOOH þ 0.25 M HClO4 (currents normalized to the electroactive surface area of the
samples). (b) Comparison of oxidation current density at the peak ( jP) and at 0 V ( j@0V).
Reproduced with permission from [160], © 2012 Royal Society of Chemistry

better electrocatalytic activity towards formic acid electrooxidation than a com-


mercial Pt black and octahedral PtCu3 alloy nanoparticles. In addition, CO stripping
experiments showed that CO is oxidised at a potential about 50 and 100 mV lower
than with the reference materials. As the authors stated, before the electrocatalytic
experiments the samples were electrochemically cleaned by continuous potential
cycling between 0.2 and þ0.9 V vs SCE (the number of cycles is not mentioned).
Thus, this potential cycling will induce the formation of a Pt rich surface. As the
authors also stated, this Pt enrichment is obviously much more evident after the
electrochemical stability test (1000 cycles between 0.2 and 0.8 V vs SCE)
although from SEM, TEM and EDX of the ERD PtCu3 nanoparticles, these seem
unaltered both in terms of shape and atomic composition. This fact again illustrates
the extremely sensitivity of the electrochemical measurements to changes at the
surface of the nanoparticles.
Very recently, Sneed et al. [163] reported the preparation of cubic Pd–Ni–Pt
core-sandwich-shell nanoparticles and their electrocatalytic properties toward
methanol and formic acid electrooxidation. Interestingly, the effect of the Ni
layer thickness and the particle size of the core-sandwich-shell on their catalytic
activities were also evaluated and compared with that obtained with PdPt
nanocubes. Despite the reported activity enhancement, the samples showed an
important restructuring, particularly in acidic solution. However, restructuring in
acidic conditions suggests a more complex catalytic behaviour from changes in
composition.
Finally, Zhang et al. have very recently reported the synthesis of PdCu bimetallic
tripods for formic acid oxidation [164]. The PdCu tripods of different atomic
composition (Pd87Cu13, Pd80Cu20 and Pd93Cu7 (prepared after selective etching
with HCl)) showed higher specific and mass activities than those found with
54 F.J. Vidal-Iglesias et al.

commercial Pd black. However, in terms of catalytic stability, that sample


containing a lower Cu content displayed higher stability during the electrochemical
essays.

2.3.2 Methanol Electrooxidation

Methanol oxidation has been extensively studied because of its great interest for
fuel cell applications, as using this fuel removes the problem of storing or gener-
ating hydrogen found in hydrogen/oxygen fuel cells. Methanol is the simplest
alcohol and its electrochemistry is also the easiest [165]. Platinum is the pure
metal with the highest activity towards this reaction in spite of its low tolerance
for CO, which is an intermediate species and poisons its surface. Due to the
sensitivity of this reaction towards the surface structure of platinum [166, 167]
and the need to have a second oxophilic metal to minimize the CO poisoning
drawback, many research manuscripts have focused during the last 2–3 years on the
use of shaped Pt-based bimetallic nanoparticles with the aim of testing them
towards this reaction. Other metals have also received attention and some results
will be also summarized in this section.
As previously stated, methanol oxidation was found to be sensitive to the Pt
surface structure. This dependence was reproduced with shape-controlled (cubic,
spherical, octahedral/tetrahedral and truncated octahedral) nanoparticles [72]. For
that shape effect comparison, we have to bear in mind that a cube is ideally enclosed
by 6 {1 0 0} faces or an octahedron by 8 {1 1 1} faces. Because of the high activity
found for Pt(1 0 0) single crystal, platinum nanocubes have been more recently
reported to enhance this reaction. Park’s group synthesized 4.5 nm cubes by a
thermal reduction process in the presence of PVP. Enhanced catalytic activities (2.4
times), in comparison with spherical nanoparticles (no preferential surface struc-
ture) were obtained in HClO4 solutions. The authors also concluded that the cubic
nanoparticles suffered lower accumulation of residues during the electrooxidation
due to a higher ratio in the peak current for the forward and backward scans. It is
worth noting that although this ratio is commonly used to gauge CO-tolerance and
catalytic activity of Pt-based electrocatalysts, a recent in situ enhanced IR study
suggests its inadequacy [168]. Similar sized nanocubes were also prepared by
Zheng’s group but in this case in the absence of any surfactant or capping agent
[169]. They were prepared in benzyl alcohol under a CO atmosphere and 1.2 times
higher current densities in comparison with spherical Pt nanoparticles were
recorded in H2SO4 medium. Capping agents are normally used in the synthesis of
shaped nanoparticles but they make the cleaning step a critical one. Concerning the
importance of the cleaning step, the same authors reported much lower oxidation
current densities for nanoparticles prepared with PVP and oleylamine, which had
been only partially cleaned [169]. Tong’s group also focused its attention towards
the importance of the residual presence of capping agents towards this reaction
[98]. In several papers it has been reported the remaining presence of PVP on the
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 55

nanoparticles [81, 98], which could have an effect on the catalytic activity, and
recently the same authors have reported the possibility to remove this capping agent
using an adapter liquid phase UV photo-oxidation technique [65]. Interestingly, the
authors claimed that the surface-bound residual PVP enhanced the activity for the
octahedral/tetrahedral nanoparticles while it was suppressed for cubic
nanoparticles. Wei’s group was also concerned about the importance of using
capping-free nanoparticles because of the great effect these molecules could
impose on the catalytic activity of the catalysts [170]. They prepared cubic
nanoparticles prepared in the absence of surfactant with an average size of around
3.5 nm. They used Pt(acac)2 which was reduced with ascorbic acid in DMF (N,
N-dimethylformamide). The higher activity obtained towards methanol oxidation
in KOH medium in comparison with the commercial Pt/C catalysts was ascribed
not only to the surface structure (large amount of {1 1 0} and {1 0 0} sites) but also
to the clean surface of the catalysts. Sun’s group prepared electrochemically THH
Pt nanocrystals, bound by well-defined high index crystal planes [171]. The activity
was not very high because these planes facilitate, for pure Pt, the dissociative
chemisorption of methanol, leading to poisoning by strongly adsorbed
CO. Nevertheless, the decoration of these nanoparticles with Ru adatoms exhibited
greatly superior catalytic currents and CO2 yields in the low potential range, when
compared with a commercial PtRu alloy nanoparticle catalyst, Fig. 2.10. In addi-
tion, the onset potential was shifted 100 mV towards more negative potentials.
Palladium is another pure metal which has been extensively studied as catalysts
towards this reaction regardless of its lower activity, due to the generation of COads

Fig. 2.10 (a) Comparison of chronoamperometric current transients recorded at 0.25 V in 1.0 M
CH3OH þ 0.1 M HClO4 solution for the various Ru-decorated THH Pt nanocrystals (solid line), Pt
black (dashed line) and PtRu/C (dashed-dotted line) electrodes. (b) Potential dependence of the
enhancement factor R (ratio of the chronoamperometric current density to the current density for the
relevant clean electrode). Reproduced with permission from [171], © 2012 American Chemical
Society
56 F.J. Vidal-Iglesias et al.

species. This reaction has also proven to be sensitive to the surface structure of
palladium, for which the shape has a great influence on the catalytic behavior.
Arriaga’s group prepared 10 nm cubic Pd nanoparticles in the presence of ascorbic
acid, sodium bromide and PVP and observed a 4-fold increase at low potentials in
the methanol oxidation in alkaline medium in comparison with commercial Pd
[172]. The authors were concerned about the site-blocking due to the PVP. They
confirmed by TGA and EDS analyses the difficulty to remove PVP and proposed an
electrochemical method, which consisted in cycling the samples in acidic and
alkaline media. Although this methodology can be valid to remove the PVP from
the surface of the nanoparticles, the high potentials reached during those continuous
cycles (oxidation/reduction) modify the surface structure of the nanoparticles,
removing a large amount of the inherent {1 0 0} sites of the cubic nanoparticles
present at the surface before this cleaning step.
Cubic nanoparticles connected in a chain-like structure were also prepared by
Opallo’s group and reported 11.5 times higher activity in comparison with the
commercial Pd/C catalyst in HClO4 solution [173]. The synthesis was performed by
reducing PdCl2 in aqueous solution by 5-hydroxytryptamine and the resulting cubes
had an average size of approximately 180 nm. Wang’s group also reported the
synthesis of cubic Pd nanoparticles dispersed on graphene oxide and observed
higher activity and stability than commercial Pd/C catalyst for this reaction in
H2SO4 medium. The cubes were prepared by reducing Na2PdCl4 with ascorbic
acid, without the presence of any surfactant or capping agent. In fact the authors
mentioned that it is the graphene oxide which acts as protecting and structure-
directing agent, ensuring that the nanoparticles have a clean surface, thus allowing a
high activity. The same authors published high activities using star-like and con-
cave Pd nanocrystallites [174], and highlighted the importance of the cleaning step
even indicating in the title of the manuscript the word “clean” as they do not use any
surfactant to obtain these shapes. The synthesis of the nanostars (approx. 250 nm)
was performed in aqueous medium adding ascorbic acid (reducing agent) and KBr
to a H2PdCl4 solution. For the concave nanocrystallites (approx. 50 nm), NaOH was
also added before the reducing agent. Those structures showed more than three
times higher activity than that of the commercial Pd/C in alkaline medium. Another
strategy to have palladium nanoparticles with high activity is to prepare three-
dimensional porous Pd nanoflowers with large electrochemically active surface
areas. Feng’s group prepared this type of structures with a high yield by a simple
one-pot wet-chemical method at room temperature using Good’s buffers, N-2-
hydroxyethylpiperazine-N9-2-ethanesulfonic acid, both as reducing and shape-
directing agent [175]. These structures have an average size of 12–68 nm and
show, without using any seed or surfactant, a preferential growth along the
{1 1 1} directions and exhibited excellent electrocatalytic activity and stability
towards methanol oxidation in alkaline medium.
Platinum and palladium are the two pure metals which have attracted most
attention towards this reaction and consequently PtPd bimetallic nanoparticles
have been also widely studied. It is interesting to highlight that the control over
the surface structure/shape of bimetallic systems is not so easy as in the case
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 57

nanoparticles of only one metal, and only in the last years it has been achieved.
Chen and coworkers prepared bimetallic PtPd nanocubes on graphene with a
surfactant-free method, in which DMF was used as a bi-functional solvent for the
reduction of both metal precursors and graphene oxide [176]. This catalyst, with a
mean size of 8 nm, showed a higher catalytic activity in H2SO4 and a better
tolerance to carbon monoxide poisoning than PtPd nanoparticles deposited on
carbon and pure Pt nanoparticles deposited on graphene. This was ascribed to the
synergetic effects of Pt and Pd and the enhanced electron transfer by graphene. Yin
et al. also prepared PtPd cubes, enriched in {1 0 0} sites, and PtPd nanotetrahedra,
enriched in {1 1 1} sites, through a hydrothermal route [177]. The mean particle
size was around 8.5 and 4.9 nm for the cubic and the tetrahedral nanoparticles,
respectively. In the synthesis procedure some small ions and PVP were used as
facet selective ({1 0 0} or {1 1 1}) agents and capping ligand, respectively. Those
ions were large amounts of Br and tiny amounts of I for the {1 0 0} facets and
Na2C2O4 and formaldehyde for the {1 1 1} facets. This study confirmed the facet-
dependent enhanced electrocatalytic activity and durability. The {1 0 0}-facet-
enclosed PtPd nanocubes demonstrated the highest activity, whereas the {1 1 1}-
facet-enclosed PtPd nanotetrahedra exhibited a better durability in HClO4 medium
and also a higher activity than pure platinum nanoparticles.
As in the case of the tetrahedral nanoparticles, octahedral nanoparticles are also
ideally enclosed by {1 1 1} facets. Park’s group synthesized octahedral PtPd alloy
nanoparticles and also reported an improved catalytic activity and stability towards
methanol oxidation [178]. The nanoparticles, with a mean size of 8.3 nm, were
prepared by means of a polyol process with glycerol as a reducing agent in an
aqueous solution. Interestingly, a comparison with pure platinum nanoparticles was
made and a stability test was performed on both catalysts, which consisted in
holding the catalysts at 0.45 V for 1 h in 0.1 M HClO4 þ 2 M CH3OH, Fig. 2.11.
In the case of the Pt/C, the current density at 0.4 V after the stability test seriously
decreases (reduction of 78.2 % from the initial value). In contrast, the PtPd alloy

Fig. 2.11 Cyclic voltammograms before and after stability test of (a) PtPd alloy/C and (b) Pt/C in
0.1 M HClO4 þ 2 M CH3OH at 50 mV s1 and at 25  C. The insets indicate CVs in 0.1 M HClO4
(blue line) and 0.1 M HClO4 þ 2 M CH3OH (black line) of the catalysts. Reproduced with
permission from [178], © 2011 Royal Society of Chemistry
58 F.J. Vidal-Iglesias et al.

still had quite high current densities (decrease of 36.5 %) at the end of the test. In
addition, the morphology of the catalysts was evaluated before and after the
stability test, as this parameter can affect the whole activity. Therefore, shape and
size distribution was observed and compared to the initial stage. In the case of the
PtPd alloy, the size and morphology of the catalysts still remained after the stability
test, thus resulting in much improved catalytic stability. On the other hand, the pure
platinum nanoparticles show massive agglomeration and it was observed an
increase in the average size from 3.44 to 4.85 nm, together with a wider size
distribution after the stability test, thus resulting in a deteriorated catalytic activity.
Yin et al. also prepared multiply twinned PtPd nanoicosahedra and reported
them to be highly active electrocatalysts for methanol oxidation [179]. In fact these
{1 1 1}-enclosed nanoicosahedra were compared to {1 1 1}-enclosed PtPd
nanotetrahedra as well as commercial Pt catalysts. The higher activity of the
icosahedra in comparison with the tetrahedra, being both {1 1 1} enriched, was
ascribed to the different surface structures between multiply twinned icosahedra
and single crystalline tetrahedra, as both of them had practically the same compo-
sition and were ideally enclosed by sole {1 1 1} facets. In addition, the larger
amount of multiply twinned defects on the icosahedrons, were also likely to play
an important role and contribute to the significantly enhanced electrocatalytic
activity. The icosahedra had a mean size of 11.2 nm and were prepared under
hydrothermal conditions by reducing K2PtCl4 and Na2PdCl4 with HCHO in the
presence of PVP as capping agent and Na2C2O4 as a facet-selective agent. In
addition, a high concentration of Hþ in solution would hamper the reducing
capability of HCHO molecules and thus slow down the reducing rate in synthesis.
In this experiment, a proper amount of HCl solution was needed for the preparation
of the PtPd nanoicosahedral nanoparticles. Without HCl added, single-crystalline
PtPd nanotetrahedra were obtained, but with an excess of HCl, the reducing rate
was too slow and only a small amount of nanocrystals with undesired broad size-
distributions were obtained.
Concave PtPd nanocubes have been also synthesized and their electrocatalytic
activity towards this reaction has been evaluated. Zhan et al. reported how to
prepare them in ethylene glycol by reducing H2PtCl6 and Na2PdCl4 with ascorbic
acid and KBr as reducing and capping agents, respectively [180]. PtPd regular
nanocubes were initially formed by co-reducing Na2PdCl4 and H2PtCl6 with
ascorbic acid due to the selective chemisorptions of Br ions on the {1 0 0} facets.
These regular nanocubes were then excavated from the side faces due to bromide-
induced galvanic replacement between [PtCl6]2 ions and Pd atoms. Meanwhile,
the overgrowth at corners and edges of the nanocubes arising from the co-reduction
of [PdCl4]2 and [PtCl6]2 ions eventually led to the PtPd alloy concave nanocubes.
It was found that the combination of the galvanic replacement induced by the
bromide and the co-reduction was responsible for the formation of these alloy
concave nanocubes. By modifying the composition, a great impact on the rate of
galvanic replacement was observed and thus the concave content of such alloy. The
highest catalytic activity was found for Pt60Pd40 concave nanocubes, which showed
much improved CO poisoning tolerance, in addition to a 4.6-fold increase in
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 59

specific activity, in comparison with a commercial Pt catalyst. This enhancement


was ascribed to the synergetic effects between Pt and Pd and to the unique surface
structure with high-index facets.
Au nanoparticles with different shapes have been also prepared and the struc-
ture/shape sensitivity towards this reaction has been confirmed. Zhang
et al. published a paper where cubic, rod-like, quasi spherical, concave cubic and
trisoctahedral nanoparticles were prepared and their catalytic activity towards
methanol oxidation in alkaline medium evaluated [181], being the latter two shapes
the most active as a result of the presence of high-index facets on their surface. The
different shapes were achieved by reducing HAuCl4 in the presence of trace
amounts of NaBH4, ascorbic acid and surfactants. The surfactants used were,
depending on the desired shape, CTAB or cetyltrimethylammonium chloride
(CTAC).
Cu is another pure metal, the activity of which towards methanol oxidation has
been evaluated using shape controlled nanoparticles. Cu is a noble metal with
relative low cost and is emerging as a promising alternative to other noble metals
such as Pt, Au or Pd. Pesika’s group prepared electrochemically platelike Cu
crystals enriched in {1 1 1} sites from a copper sulfate solution in the presence of
KBr and compared its catalytic activity to that of truncated octahedron-shaped
nanoparticles [182]. The former were more active towards this reaction in alkaline
medium than the latter. In order to understand this dependency towards the shape/
surface structure of the nanoparticles, the authors investigated this reaction with Cu
single crystals. They observed that the {1 1 1} and {1 1 0} crystal planes were more
active in comparison to the {1 0 0} plane and that was the reason to explain why the
{1 1 1} enriched plates were more active than the truncated octahedron-shaped
nanoparticles, for which the three crystal facets are exposed almost equally.
Cu has been also used with the aim of reducing the amount of more expensive
noble-metal based catalysts, such as Pt or PtPd, by preparing PtCu [183] or PtPdCu
[184] catalysts. Qi et al. prepared PtCu (Pt59Cu41) concave nanocubes enclosed by
high-index {5 1 1} facets by co-reducing Pt(acac)2 and Cu(acac)2 with oleylamine
in the presence of CTAB and trioctylphosphine oxide at 180  C. These concave
nanocubes exhibited substantially enhanced electrocatalytic properties for metha-
nol oxidation in HClO4 medium relative to commercial Pt/C in addition to a lower
amount of platinum needed [183]. In fact, the concave cubes, with a mean size of
15 nm, showed 3.0 and 4.7 times higher current densities than those of PtCu
nanoparticles and commercial Pt/C respectively. Chronoamperometric experiments
were performed and concave PtCu cubes had, over the entire time, higher current
densities than the other two tested catalysts. The authors ascribed this improvement
to the combination of Pt with Cu, which can cause a small compressive strain due to
their weak lattice constant, therefore reducing the binding strength of CO on the
catalyst. In addition, Cu helps in the oxidation of CO by generating hydroxyl
species from water. They also suggested that the surface structure plays an impor-
tant role in determining the catalytic activity, thus explaining the differences
between the PtCu nanoparticles and the PtCu concave nanocubes.
60 F.J. Vidal-Iglesias et al.

Platinum has been also alloyed with other inexpensive metals, especially 3d
transition metals in order not only to decrease costs but also to provide the catalysts
with different electronic configurations. These transition metals have the ability to
weaken the interaction between Pt and the unwanted PtOH intermediate, thus
enhancing the catalytic performance [157, 185]. In fact Fang’s group has reviewed
the effect of the shape control and the electrocatalytic activity of Pt-based bimetallic
nanocrystals, including high-index Pt-M nanostructures [156], in which special
attention is paid to the synthesis approach commonly used for their preparation.
Hu et al. prepared hollow PtNi nanocatalysts supported on graphene by a galvanic
displacement reaction with PtCl62 and Ni2þ as precursors [186]. First, Ni2þ ions
were reduced by NaBH4 to Ni nanosheets as well as GO to graphene. Then, the
platinum precursor was added and the hollow PtNi nanostructures were formed due
to the displacement between the Ni nanosheets and PtCl62 ions. The nanoparticles,
with a mean size of 30 nm and a pore size of 1 nm showed a high activity and good
poison tolerance. The authors ascribed this improvement to the synergistic role of Pt
and Ni in the catalysts in comparison to solid PtNi deposited on graphene and to
commercial Pt/C nanocatalysts. The Ni hydroxide species, which exist on the
surface [187], offer OH species to remove the intermediate CO that blocks active
platinum sites. In addition, the open pores in the shells could greatly enhance the
surface area and then a better use of the platinum in the catalyst, thus reducing the
costs. Furthermore, the mesoporous structured shell provided very large fraction of
edges and corner atoms [188, 189], endowing the hollow Pt–Ni–graphene
nanocatalysts with good electrocatalytic performance. Graphene also provided a
high electrical conductivity and good stability for keeping a stable catalyst structure.
Similar results were obtained by Sun’s group with hollow PtNi nanospheres
prepared, in this case, by chemical successive reduction method in the presence of
PVP [190]. The facet dependent electrocatalytic activity was also confirmed for
Pt3Co nanocrystals [191]. Nanoflowers and nanocubes, enriched in {1 1 1} and
{1 0 0} facets, respectively, were prepared and the former showed a higher activity
in HClO4 medium towards methanol oxidation than the latter (1.5-fold higher) and
Pt3Co nanoparticles (2.5-fold higher). In the synthesis, oleylamine and oleic acid
were used together with benzyl ether and the metal precursors. With a similar
synthesis methodology, but using 1-octadecene instead of benzyl ether, PtFe con-
cave nanocubes of 7.6 nm were prepared, showing an enhanced methanol oxidation
activity and CO tolerance, in comparison to PtFe nanocubes and commercial Pt
catalysts, in HClO4 medium. This enhancement was ascribed to the presence of
high-index atomic steps. With the same aim, reduce Pt content while reducing CO
poisoning, Zn was also used with Pt to prepare highly active catalysts.
Murray’s group prepared PtZn nanocrystals in oleylamine and oleic acid and
found the activity of the spherical nanoparticles higher than that of nanocubes
[192]. However, upon annealing, the PtZn alloy phase could be transformed into
the Pt3Zn intermetallic phase, which showed better performance than the alloy,
together with an excellent poisoning tolerance. The activities recorded in H2SO4
medium were very similar to those of commercial platinum while reducing cost and
enhancing poisoning tolerance.
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 61

2.3.3 Ethanol Electrooxidation

Ethanol electrooxidation [193] is a reaction of great interest as ethanol can be


used as fuel in fuel cells or in internal engines. Although methanol has been more
extensively studied in the fuel cell field, ethanol is attracting more and more
interest because it can be obtained from renewable sources and because its
toxicity and corrosive power are lower than those of methanol. An ideal complete
oxidation of ethanol to carbon dioxide releases 12 electrons. Nevertheless it is
quite difficult to break the C-C bond, and therefore a much lower number of
electrons per molecule is obtained. Thus, the main oxidation products in acid
media are C2 molecules, acetaldehyde (two electrons) and acetic acid (four
electrons), being the formation of CO2 quite low at ambient temperature. In
acid media, although pure Pt becomes poisoned during the reaction, platinum
based catalysts are the most active.
Our group studied the ethanol oxidation by cyclic voltammetry and FTIR
techniques using spherical, cubic and octahedral Pt nanoparticles and evaluated
the role of the surface structure/shape of the nanoparticles over their catalytic
activity. The behavior was that predicted from the results obtained for the single
crystal electrodes [194], and the facet geometries of the nanoparticles. Spherical
nanoparticles (4 nm) were prepared using a water in oil microemulsion (n-heptane/
Brij®30/water), while octahedral and cubic nanoparticles (8–10 nm) were prepared
with a colloidal method using NaPA as capping agent. For all the platinum samples,
both the incomplete oxidation towards acetaldehyde and acetic acid and the for-
mation of CO was observed, although the ratio of both paths was different. The
ethanol oxidation was performed in H2SO4 and HClO4 media, Fig. 2.12. Octahedral
nanoparticles, enriched in {1 1 1} sites, showed almost exclusively acetic acid
formation, therefore showing a low poisoning rate and being very similar the
positive and negative sweeps. Cubic nanoparticles are ideally enclosed by
6 {1 0 0} faces, and the Pt(1 0 0) is the most active platinum basal plane at splitting
the C-C bond. Thus, cubic nanoparticles showed the largest activity for the cleavage
of the C-C bond. Finally, spherical nanoparticles, which show no preferential
orientation, possess a high amount of low coordinated atoms on the surface
which is also effective to cleave the C-C bond, and CO was also obtained.
Park et al. prepared small Pt cubes (4.5 nm) by a thermal reduction process in the
presence of PVP and reported a higher electrocatalytic activity in comparison with
spherical nanoparticles (3.5 nm) [195]. They observed current densities 2.4 times
higher for the cubic nanoparticles, which also showed a much lower onset potential
(0.21 V) in HClO4. In addition to the higher current densities the authors also
concluded that the cubic nanoparticles exhibited a lowered poisoning rate in base of
the ratio between the current densities in the forward and backward scans, which is
much higher in the case of the cubic nanoparticles (0.84 vs 0.57) compared with that
of spherical nanoparticles.
62 F.J. Vidal-Iglesias et al.

E vs RHE/V E vs RHE/V
0.0 0.2 0.4 0.6 0.8 0.2 0.4 0.6 0.8 1.0
3.5 3.5
a b
3.0 3.0

2.5 2.5
j/mAcm-2

j/mAcm-2
2.0 2.0

1.5 1.5

1.0 1.0

0.5 0.5

0.0 0.0

3.0
c d 3.0

2.5 2.5

2.0 2.0
j/mAcm-2

j/mAcm-2
1.5 1.5

1.0 1.0

0.5 0.5

0.0 0.0
0.0 0.2 0.4 0.6 0.8 0.2 0.4 0.6 0.8 1.0
E vs RHE/V E vs RHE/V

Fig. 2.12 Ethanol oxidation (first cycle) on (a) (poly)Pt, (b) (1 0 0)Pt, (c) (1 0 0)-(1 1 1)Pt and (d)
(1 1 1)Pt nanoparticles in 0.5 M H2SO4 þ 0.2 M CH3CH2OH (black line) and 0.1 M HClO4 þ 0.2 M
CH3CH2OH (red line). Sweep rate: 50 mV s1. Reproduced with permission from [194], © 2013
Royal Society of Chemistry

Sun et al. also studied this reaction using Pt nanoflowers, which were synthe-
sized in deep eutectic solvents and characterized by SEM, TEM, XRD, XPS and
electrochemical tests [196]. The size and the shape of these nanoflowers could be
controlled and ranged from 45 to 250 nm approximately. Current densities could be
up to nearly three times higher in HClO4 medium. These nanoflowers possess sharp
single crystal petals and a high density of atomic steps to which the higher
electrocatalytic activity was ascribed to.
Probably due to the high activity observed with single crystal surfaces with low
Miller indices, several papers have been published with nanoparticles enclosed by
high index facets. In the case of platinum Sun et al. published different manuscripts
where they studied this reaction with such particles. They observed that concave
THH nanocrystals prepared in deep eutectic solvents bounded by {9 1 0} faces were
more active than platinum black in HClO4 medium [197]. By modifying the
synthesis conditions, the size and shape could be controlled. The electrocatalytic
activity was evaluated by cyclic voltammetry and chronoamperometry. The former
concluded that the THH were two times more active than Pt black and the latter,
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 63

a b 2.5
10
Concave Pt Concave Pt
Pt black 2.0 Pt black
8
Pt/C Pt/C
j / mAcm-2

j / mAcm-2
6 1.5

4 1.0

2 0.5

0 0.0
-0.2 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8
E / V vs SCE E / V vs SCE

Fig. 2.13 Voltammetric profiles for the electrooxidation of (a) formic acid and (b) ethanol of
concave Pt nanocrystals, commercial Pt black, and Pt/C (E-TEK). The formic acid oxidation was
recorded in 0.5 M H2SO4 þ 0.25 M HCOOH solution at a scan rate of 50 mV s1. The ethanol
oxidation was recorded in 0.1 M HClO4 þ 0.1 M CH3CH2OH solution at a scan rate of 50 mV s1.
Reproduced with permission from [121], © 2011 American Chemical Society

than they were more stable after 2 h at constant potential (0.45 V vs SCE). Also,
triambic icosahedral nanoparticles (160 nm, enclosed by {7 7 1} high-index facets)
were prepared electrochemically in deep eutectic solvents and showed higher
activity (2–4 times) and stability than a commercial Pt black catalyst also in
HClO4 medium [198]. The same group reported how Pt nanocubes (10 nm) could
be transformed into THH nanocrystals after a short square-wave potential treatment
(2 min) in H2SO4 with sizes from 6 to 20 nm [199]. With this methodology, low
index {1 0 0} facets were transformed into {3 1 0} high-index facets, and the
presence of this type of sites was the cause of a remarkable increase in activity
towards ethanol oxidation, in comparison with the original Pt nanocubes and
commercial platinum nanoparticles. Also, Zheng’s group reported an amine-
assisted synthesis of concave polyhedral Pt nanoparticles having {4 1 1} high-
index faces, which exhibited an enhanced electrocatalytic activity (2.3–5.6 times)
over commercial Pt catalysts towards ethanol oxidation reaction in HClO4 [121],
Fig. 2.13.
Zhang et al. also reported a higher activity with multipod and concave Pt
nanocrystals, mainly exposing {2 1 1} and {4 1 1} high-index facets, respectively
[200]. The authors also ascribed the higher activity (in HClO4) in comparison to Pt/C
and Pt nanocubes with low-index facets to the high density of surface
atomic steps.
Among the other pure metals, Pd and Rh have also been used to prepare shaped
nanoparticles, with which ethanol oxidation has been studied. Arriaga
et al. prepared cubic Pd nanoparticles with a very high yield by a chemical
reduction in aqueous medium employing ascorbic acid, PVP and sodium bromide
as reducing agent, surfactant and additive, respectively [172]. The nanoparticles
were prepared with (17.5 nm) and without (10.6 nm) Vulcan carbon as support and
64 F.J. Vidal-Iglesias et al.

the PVP was removed electrochemically. Pd nanocubes exhibited a much better


performance than commercial Pd nanoparticles towards ethanol oxidation,
reporting current densities around three times higher in voltammetric experiments
in KOH medium. Pd was also used by Xie et al. to prepare concave nanocubes with
high-index {7 3 0} facets and vicinal {h k 0} facets such as {3 1 0} [201]. These
were prepared by a simple reduction of an aqueous solution of H2PdCl4 with
ascorbic acid in the presence of CTAC and CTAB at room temperature. Depending
of the ratio CTAC:CTAB, the authors were able to have crystals with different
degrees of concavity. Those with the ratio 4:1 were the nanoparticles showing the
highest electrocatalytic activity in NaOH medium, showing from 4 to 6.5 times
higher catalytic activities than a commercial Pd black catalyst and a good stability.
Shaped pure Rh nanoparticles were also used to study this reaction. Sun’s group
prepared THH Rh nanocrystals enclosed by {8 3 0} high-index facets through an
electrochemical square wave potential method at room temperature in a H2SO4
solution [202]. Voltammetric experiments in NaOH gave current densities nearly
seven times higher for the THH nanoparticles than for Rh black. Also, for
chronoamperometric experiments, after 30 min, these high-index nanoparticles
showed higher current densities (1.8 times).
With these metals (Pt, Pd, Rh) and also Au, different bimetallic or trimetallic
nanoparticles have been synthesized and their behavior towards ethanol oxidation
has been evaluated. Chen et al. reported the synthesis of PtPd alloys with different
shapes (spherical, nanoflowers and nanodendrites) deposited in graphene in aque-
ous solution in the presence of ethanol [203]. For all of them, higher electrocatalytic
activities and better tolerance to reaction intermediate poisoning was observed in
comparison with Pt nanoflowers supported on graphene, Pd nanoparticles supported
on graphene and PtPd nanoparticles supported on carbon black, which was ascribed
to the synergetic effects of Pt and Pd and the enhanced electron transfer properties
of graphene.
PtRh nanocrystals are also very interesting from a catalytic point of view,
because several studies have confirmed that the addition of Rh to Pt can signifi-
cantly enhance the C-C bond cleavage due to a change of electronic properties
[204–206]. In this sense, Sun’s group reported the synthesis of PtRh nanocubes with
different atomic ratio supported on graphene using the modified polyol method with
Br for the shape-directing agents and in the absence of any capping agent
[207]. The most active catalyst towards ethanol oxidation was that with a Pt:Rh
atomic ratio ¼ 9:1, which showed superior specific activity and long-term stability
to commercial Pt/C catalyst in HClO4 medium, being current densities for that
alloy, twice that for Pt/C for both voltammetric and chronoamperometric experi-
ments. In addition FTIR spectroscopy was used to evaluate the C-C bond cleavage
in ethanol. It was found that alloyed Rh can promote the split of this bond, being the
catalyst with the ratio Pt:Rh ¼ 1:1 the best for breaking the C-C bond in ethanol.
The same group, supported on the advantage of the PtRh nanocrystals and that of
nanoparticles with high-index facets for this reaction, also prepared PtRh
nanocrystals with high-index facets: {8 3 0}-bound THH and {3 1 1}-bound trape-
zohedron (TPH) [109]. The nanoparticles were prepared electrochemically and the
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 65

Fig. 2.14 Cyclic


voltammograms for ethanol
oxidation on Pt–Rh THH,
Pt–Rh TPH, Pt THH, Pt
TPH and commercial Pt/C
catalysts at 50 mV s1 in
0.1 M ethanol þ 0.1 M
HClO4. Reproduced with
permission from [109],
© 2013 Royal Society of
Chemistry

shapes could be tuned from THH to TPH by altering the electrode potential. They
observed excellent electrocatalytic properties and a high ability to break the C-C
bond of ethanol in HClO4 medium of both nanoparticles enclosed by high-index
facets. The {3 11} facets of the trapezohedron nanocrystals have a higher step
density than {8 3 0} facets on THH nanoparticles, explaining the highest catalytic
activity of the former. That activity was about 6.3 times higher than that of a
commercial Pt catalyst and the onset potential was slightly shifted towards more
negative potentials, Fig. 2.14.
Au was also used, together with Pd, to synthesize highly active catalysts towards
ethanol electrooxidation. AuPd octapodal nanoparticles were prepared by selective
etching of {1 0 0} facets by in situ generated Br ions in aqueous medium. Their
electrocatalytic behavior was compared to that of flower-like AuPd alloy
nanoparticles with similar composition and size and that of a commercial Pd/C
catalyst [208]. Peak current densities were about two and four times higher, in
comparison with those of flower-like and Pd/C catalysts, respectively. Chronoam-
perometric experiments also revealed a superior electrochemical stability of the
octapodal nanoparticles. This enhancement was ascribed to the presence of a
number of active sites on their surfaces such as highly active facets, gaps between
pods, and some defect sites, which originate from their octapodal structures with
rough surface features. Moreover, it was also attributed to the high fraction of
{1 0 0} facets in comparison with the other catalysts. By using density functional
theory (DFT), Pd(1 0 0) was reported to be the best basal plane for the ethanol
dissociation [209]. The same bimetallic system was studied by Zheng’s group,
which reported the preparation of monodisperse Au–Pd alloy nanoparticles with
systematic shape evolution from rhombic dodecahedral (RD) to trisoctahedral
(TOH), and hexoctahedral (HOH) structures by varying the concentration of sur-
factant in the surfactant-mediated synthesis [210]. The catalytic activities toward
ethanol electrooxidation were in the order of HOH > RD > TOH, which follows the
order of their corresponding surface energies. Finally, Xu’s group also prepared
66 F.J. Vidal-Iglesias et al.

AuPd nanocrystals, but in this case it was Au@Pd concave nanocubes (21 nm)
enclosed by high-index facets which were also decorated with Pt on top [211]. The
authors normalized their activity to the mass or Pd or Pd þ Pt and obtained 7-9
higher current density values than for a Pd black catalyst in KOH medium.

2.3.4 Oxygen Reduction Reaction

The oxygen reduction reaction (ORR) is probably the most important electrochem-
ical reaction due to its multiple and relevant applications [212–218]. Consequently,
the ORR has been extensively and intensively studied through the years. In the
complete reduction of oxygen to water, there are four electrons exchanged, involv-
ing the double bond breaking and the formation of four O-H bonds. Thus, this
multi-electron transfer process necessarily requires the existence of several steps in
its reaction mechanism and therefore involves the possible existence of several
adsorption intermediates. However, and although knowledge of the oxygen reduc-
tion reaction process has advanced considerably, the exact mechanism remains not
fully understood [216, 217, 219, 220].

2.3.4.1 ORR in Shape-Controlled Pt Nanoparticles

Among pure metals, Pt is well-recognized to be the electrode with the highest


electrocatalytic activity for the 4 e pathway [217, 221]. In addition, from single
crystal studies, it is known that the ORR on Pt is a structure-sensitive
electrocatalytic reaction. In particular, and considering half-wave potentials, the
electrocatalytic activity decreases in the order of Pt(1 1 0) > Pt(1 1 1) > Pt(1 0 0) in
perchloric acid solution and of Pt(1 1 0) > Pt(1 0 0) > Pt(1 1 1) in sulphuric acid
solution [1, 222]. The different reactivity in both electrolytes points out the impor-
tant role that the competitive anion adsorption plays on the reaction. Moreover,
from studies with stepped Pt single-crystal surfaces, it was clearly observed that
these stepped surfaces showed, irrespectively of their step site symmetry, higher
ORR catalytic activity than those found with basal low-index crystal surfaces
[219, 223–225]. To the best our knowledge, the first use of shape-controlled Pt
nanoparticles towards ORR was reported by Inaba et al. [226]. They prepared cubic
Pt nanoparticles in presence of NaPA which showed a high activity for ORR in
H2SO4 solutions. In addition, potential cycling surface structure perturbation led to
a decrease on the ORR activity as a consequence of the structural change from the
initial {1 0 0} surface orientation to a polycrystalline orientation.
Later, we evaluated the ORR activity of different and clean shape-controlled Pt
nanoparticles both in H2SO4 and HClO4 solutions by using scanning electrochem-
ical microscopy (SECM) [227]. Under these conditions, we found a good agree-
ment between the reactivity of the nanoparticles and that previously observed with
single crystal electrodes. In particular, the activity for the ORR followed the trend
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 67

Current (A)
-6.50e-8 -7.69e-8 -8.88e-8 -1.01e-7 -1.13e-7

0 µm
a

0.7 V 1350 µm
0 µm 3000 µm

PtNPtetra PtNPhexa PtNPsphe PtNPcubic

750 µm
b

0.4 V 1350 µm
0 µm 3000 µm

-3.72e-7 -3.50e-7 -3.27e-7 -3.05e-7 -2.82e-7

Current (A)

Fig. 2.15 SECM TG/SC images displaying the reduction current collected for ORR in 0.1 M
HClO4 solution at a Pt NPs array. This array is formed by spots of four different types of shape-
controlled Pt NPs (Tetrahedral (tetra), hexahedral (hexa), spherical (sphe) and cubic (cubic) Pt
nanoparticles). Scan rate ¼ 125 μm/s. (a) Image of the first and second rows of the array held
constant at 0.7 V. (b) Image of the second row of the array held constant at 0.4 V. Reproduced with
permission from [227], © 2010 American Chemical Society

PtNPhexahedral > PtNPtetrahedral  PtNPspherical > PtNPcubic in 0.1 M HClO4


and the trend PtNPhexahedral > PtNPcubic > PtNPspherical > PtNPtetrahedral in
0.5 M H2SO4. Figure 2.15 shows a representative image of the different Pt
nanoparticles obtained at two different potentials (0.7 V (kinetic region) and
0.4 V (diffusion region)) in 0.1 M HClO4 solution. At 0.4 V, and due to the fact
that the current only depends on the diffusion rate of O2 toward the surface, which is
the same in all cases, the activity of the samples is essentially the same. However, at
0.7 V, the Pt nanoparticles showed different ORR activities, which are determined
by their specific surface structure in the electrolyte used.
Later, Yu et al. [228] published the synthesis of concave Pt nanocubes enclosed
by high-index facets such as {5 1 0}, {7 2 0}, and {8 3 0}. These Pt concave
68 F.J. Vidal-Iglesias et al.

Fig. 2.16 (a) Voltammetric profiles of different Pt catalysts in 0.5 M H2SO4 at 20 mV s1. (b)
Polarization curves for ORR in an O2 saturated 0.5 M H2SO4 solution (1600 rpm at 10 mV s1). (c)
Mass and specific activities (at 0.8 V vs. RHE). (d) Electrochemical durability test of concave Pt
nanoframes in 0.5 M H2SO4 at 20 mV s1. Reproduced with permission from [124], © 2013 Wiley
and Sons

nanocubes showed a specific activity towards ORR in HClO4 solutions (normalized


to the electroactive surface area) about 2–4 higher than that obtained with Pt
nanocubes, cuboctahedral (that is Pt nanoparticles bounded by {1 0 0} and {1 1 1}
low-index facets) and commercial Pt/C nanoparticles. These results show the low
activity of the Pt cubes and the high activity of the concave Pt nanocubes containing
high index surfaces in HClO4 solutions, and agrees with our previous observations
[227] and with that expected from Pt single crystal studies [222].
Xia et al. reported an interesting contribution dealing with the preparation of
highly concave Pt nanoframes with high-index {7 4 0} facets and their activity
towards ORR in H2SO4 solution [124]. They found that the concave Pt nanoframes
were not only more active (both mass and specific activities, Fig. 2.16b, c) than Pt
nanocubes and a commercial Pt/C catalyst (Johnson Matthey, JM 20 wt%) but also
much more stable under electrochemical conditions, Fig. 2.16d. Care should be
taken, however, with surface cleanness and with the fact that diffusion currents
should be the same in all samples for a proper comparison.
Fu et al. [229] again analyzed the activity toward ORR of Pt nanocubes (about
6 nm), prepared using a green hydrothermal method with polyallylamine hydro-
chloride (PAH) as capping agent and formaldehyde as a reducing agent. The Pt
nanocubes displayed an enhanced mass activity and specific activity both in H2SO4
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 69

Fig. 2.17 (a) CV curves for the Pt nanoparticles and commercial E-TEK Pt black in N2-saturated
0.5 M H2SO4 solutions at 50 mV s1. (b) Hydroxyl surface coverage (ΘOH) for the Pt nanoparticles
and E-TEK Pt black. (c) ORR polarization curves for the Pt nanoparticles and E-TEK Pt black in
O2-saturated 0.5 M H2SO4 solution at 5 mV s1 and rotation rate of 1600 rpm. (d) Specific activity
of the Pt nanoparticles and commercial E-TEK Pt black at a fixed potential of 0.85 and 0.9 V for
the ORR. Reproduced with permission from [229], © 2013 Royal Society of Chemistry

and HClO4 solutions in comparison to a commercial Pt black (E-TEK). Despite this


enhancement, it is worth noting that the Pt nanocubes were UV/Ozone treated
(185 and 254 nm in air for 4 h) and electrochemically cycled (0.2 to 1.2 V vs
RHE for 50 cycles) in order to remove the capping agent and other possible
contaminants. These treatments are known to induce important modifications on
their surface structure thus resulting in a more disordered {1 0 0} surface [97]. This
effect can be visualized by direct comparison of the characteristic response of the
{1 0 0} domains in H2SO4 solution of the Pt nanocubes used (Fig. 2.17a) with that
obtained in previous contributions, Fig. 2.1 [73]. In addition, as authors stated, from
FT-IR experiments they concluded that, owing to its excellent chemical stability,
the PAH molecules attached on the surface of the Pt nanocrystals cannot be
completely removed by UV/ozone treatment. This strong Pt-N bonding was respon-
sible of the negative shift (about 0.3 eV) of the Pt 4f binding energy of the Pt
nanocubes in comparison with the Pt black. Thus, the enhanced electrocatalytic
activity of the Pt nanocubes in HClO4 solution was related to this change of the
d-band center of Pt that is, to an electronic effect.
70 F.J. Vidal-Iglesias et al.

Zhou et al. [230] prepared icosahedral Pt nanoparticles (8.8 nm edge length)


using the hot injection-assisted GRAILS (gas reducing agent in liquid solution)
method [231, 232]. The specific activity (normalized to the ECSA obtained from
the charge in the hydrogen region) of the carbon supported Pt icosahedra obtained
in HClO4 was about 0.83 mA cm2Pt which was higher than that found for Pt
nanocubes (0.35 mA cm2Pt) and for a Pt/C reference material (0.2 mA cm2Pt).
This enhancement was attributed to the presence of {1 1 1} domains and twin planes
at the surface of the nanoparticles. The stability of the carbon supported Pd
icosahedral nanoparticles was evaluated after being cycled in O2 saturated solutions
for 2500 cycles. Unfortunately, this stability was exclusively evaluated (using
TEM) in terms of particle shape but not in terms of surface structure.
More recently, Devivaraprasad et al. [63] have reported a detailed electrochem-
ical study of different shape-controlled Pt nanoparticles prepared with NaPA and
TTAB as capping agents. The samples were electrochemically characterized using
most of the surface probes previously discussed in this chapter, including cyclic
voltammetry in different supporting electrolytes (H2SO4, HClO4 and NaOH), Cu
UPD and Bi and Ge analyses. After this complete electrochemical characterization,
the samples were tested towards ORR both the acidic and alkaline electrolytes.
They found that the ORR specific activity trends of the shape controlled Pt
nanoparticles in the three electrolytes was in good agreement with that expected
from single crystals contributions [222]. Thus, whereas the {1 0 0} oriented Pt
nanoparticles were more active in H2SO4, the {1 1 1} ones displayed better activ-
ities both in HClO4 and NaOH. Interestingly, the trends in peroxide formation were
also evaluated. Nevertheless, as authors also stated, the analysis of the H2O2
formation in this kind of nanomaterials is rather complicated and several aspects
should be considered including effect of the particle size, active site density, layer
thickness, reaction mechanism and “surface blocking” by adsorbed species.
Interestingly, Tripkovic et al. [233] reported a density functional theory (DFT)
study about the effect of the size and shape of Pt nanoparticles on the ORR activity
in a non-adsorbing electrolyte. Also, the ORR activity of these shapes at 0.9 V was
compared with that of the “equilibrium shape”, which was determined by calculat-
ing the changes in surface energies with potential for low-index Pt facets. The
results showed that the tetrahedral Pt nanoparticles provided the highest activities
and the cubic Pt nanoparticles the lowest ones, Fig. 2.18. These findings are in good
agreement with previous experimental observations on both shape-controlled Pt
nanoparticles and Pt single crystals in non-adsorbing electrolytes such as HClO4.
In this sense, Li et al. [234] have very recently discussed about the activity and
stability for ORR in acidic media of size and shape controlled Pt nanoparticles.
These aspects were analysed in direct relationship with the knowledge gained from
Pt single crystals. They found that under ORR relevant conditions (0.6–1.1 V, 4000
scans, 50 mV s1 and 60  C), the particle size and shape effects constantly evolved
due to the potential-induced modifications. In fact, they stated that for long-term
catalytic performance, the shape-controlled activity will disappear and the activity
will be determined by that coming from thermodynamically stable round-like
nanoparticles that is, the “equilibrium shape” described by Tripkovic et al. [233].
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 71

Fig. 2.18 Dependence of the specific activities and mass activities with the particle diameter and
specific surface area. Reproduced with permission from [233], © 2014 Springer

2.3.4.2 ORR in Shape-Controlled Pd Nanoparticles

Among the pure metals, Pd is one of the most active catalysts for ORR. From single
crystal studies, it was shown that the ORR on Pd is also a structure sensitive
reaction. Kondo et al reported that, in HClO4 solution, the ORR activity on Pd
single crystals increases in the order of Pd(1 1 0) < Pd(1 1 1) < Pd(1 0 0) (estimated
at þ0.90 V (RHE)) [235]. In addition, on Pd(S)[n(1 0 0)  (1 1 1)] and Pd(S)[n
(1 0 0)  (1 1 0)] single crystals, the activity increased linearly with the increase of
the terrace atom density, thus pointing out that the {1 0 0} domains contain the most
active surface sites towards ORR [236]. To the best our knowledge, the first
example of shape-controlled Pd nanoparticles toward ORR was reported by Xiao
et al. [237]. They found that the specific activity of Pd nanorods was about 10 times
higher than that of Pd spherical nanoparticles and comparable to that of bulk
Pt. This enhancement was attributed, in contrast with the Pd single crystal studies,
to the presence of {1 1 0} facets in which, according with their DFT calculations,
the interaction between an O adatom and a {1 1 0} site is exceptionally weak.
Almost in parallel, Shao et al. [238] and our group, in collaboration with the
Tammeveski group [74], reported the ORR activity of Pd nanocubes, that is
nanoparticles containing a preferential {1 0 0} surface structure. Shao et al. showed
72 F.J. Vidal-Iglesias et al.

Fig. 2.19 ORR behaviour


of Pd cubes and octahedra
supported on carbon black
and Pd/C annealed at
500  C for 2 h (Pd/C-HT) in
0.1 M HClO4 at 10 mV s1
and 1600 rpm. Inset:
Comparison of specific
activities at 0.9 V.
Reproduced with
permission from [238],
© 2011 Royal Society of
Chemistry

that the ORR specific activity of Pd nanocubes (about 6–7 nm) in HClO4 was about
one order of magnitude higher than that of Pd octahedral (about 5–6 nm edge length)
and a commercial cubooctahedral Pd/C catalyst (BASF, about 7 nm (after a thermal
treatment)) and comparable to that obtained with the state-of-the-art Pt/C catalysts
(TKK, 46.6 wt% about 3 nm), Fig. 2.19. The electroactive surface area of the Pd
samples was estimated by integrating the stripping charge of the Cu UPD process,
assuming 420, 490, and 460 μC cm2 for full Cu monolayer coverage on cubes,
octahedra, and cubooctahedra, respectively.
On the other hand, it was shown [74] that the ORR specific activity (estimated at
0.85 V vs RHE in H2SO4 solutions) of clean cubic Pd nanoparticles (about 27 nm)
was about three times higher than that obtained with spherical Pd nanoparticles
(about 3 nm) and a bulk Pd electrode. In this case, the surface area of the Pd
nanoparticles was determined, after oxygen reduction measurements, by charge
integration under the oxide reduction peak, assuming the value of 424 μC cm2 as
the charge density for the reduction of a monolayer of PdO [239]. In addition, in a
forthcoming contribution [75], we also studied the ORR in alkaline solution with
similar ~27 nm Pd nanocubes. The ORR specific activity at 0.95 V of the Pd
nanocubes was again about three times higher than that obtained with spherical
Pt nanoparticles and a bulk Pd electrode.
A similar contribution was published by Lee et al. [240]. They prepared Pd
nanocubes with different particle size (27, 48 and 63 nm) and evaluated their ORR
activity in alkaline solution. It was found that the ORR specific activity (measured
at 0.1 V vs Ag/AgCl) of the 48 nm Pd nanocubes was higher than that obtained
with the 27 and 63 nm Pd nanocubes. In addition, the activity of these Pd nanocubes
was, independently of their particle size, higher than that obtained with 9 nm Pd
nanoparticles. The electrochemical active surface area of the Pd nanocubes was
estimated from the integration of the PdO reduction charge.
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 73

Shao et al. [136] extended their previous observations about the ORR activity of
cubic and octahedral Pd nanoparticles (5–7 nm) in HClO4 [238], H2SO4 and
alkaline solutions. They observed that, in H2SO4, the ORR activity of Pd cubes
was again much higher (about 17 times) than that of Pd octahedra. This behaviour
was explained as due to the stronger adsorption of (bi)sulfate anion on the {1 1 1}
domains present at the surface of the Pd octahedra. Nevertheless, in alkaline
solution, the ORR was shown to be almost insensitive to the surface structure of
the different nanoparticles, in clear contradiction with previous findings. In this
regard, it is worth recalling the important differences in particle size which could
change the binding energy of the species involved in the reaction mechanism, in
particular, OHad species. More work would be required to understand these contra-
dictory results.
More recently, Shao [241] and Zhang et al. [242] have published relevant
reviews about the recent progress about the electrocatalysis of Pd-based
nanomaterials for ORR.

2.3.4.3 ORR in Shape-Controlled Au Nanoparticles

ORR has been also extensively evaluated on Au surfaces. From Au single crystals
studies, it is well-established that the ORR is sensitive to the surface structure and
particularly to the {1 0 0} bidimensional domains in alkaline solution [214, 243,
244]. In alkaline solution, the ORR on Au(1 0 0) is especially interesting because
not only shows an electrocatalytic activity similar to that of Pt but also due to the
fact that exchanges four electrons in the potential window between þ0.55 and
þ0.95 V vs RHE. However, the origin of this high catalytic activity is still under
discussion. As a practical consequence of this surface structure sensitivity, the ORR
in alkaline solution has been explored in different shape-controlled Au
nanoparticles. Nevertheless, and taking into account that we have very recently
reviewed the Au electrocatalysis for ORR, also including a particular section
devoted to the use of shape/surface structure-controlled Au nanoparticles, this
aspect will not be included/discussed in this chapter. For those readers interested
in this point we refer to [34, 214, 245].

2.3.4.4 ORR in Shape Controlled Pt and Pd Based Alloy Nanoparticles

A great number of contributions have been published on the use of Pt or Pd based


alloy nanoparticles to enhance ORR electrocatalysis with two main objectives:
(a) improving ORR reactivity and (b) decreasing costs. In this sense, the control
of the shape of these Pt- or Pd-based nanoparticles and consequently their surface
structure has been also deeply explored [215, 218, 241, 246, 247]. To the best of our
knowledge, the first attempt to produce shape controlled PtPd alloy nanoparticles
was published in 2009 by Lim et al. [248]. They reported the synthesis of PdPt
bimetallic nanodendrites which displayed about 2.5 times higher mass activity than
74 F.J. Vidal-Iglesias et al.

the state-of-the-art Pt/C catalyst. However, the use of these shape controlled Pt
alloy nanoparticles is marked by the publication made by Markovic’s group [249]
in which the specific activity of a single crystal Pt3Ni(1 1 1) surface was shown to be
about 18 mA cm2 at 0.9 V, that is, about 90 times higher than that of a commercial
Pt/C (about 0.2 mA cm2 Pt) in 0.1 M HClO4 solution. This specific activity is still
the highest ORR activity ever measured. This striking enhancement was attributed
to a low coverage of hydroxyl species induced by the specific electronic structure of
the Pt3Ni sample. The Pt3Ni(1 1 1) surface was found to form a Pt-skin layer with a
peculiar oscillatory near-surface compositional Pt and Ni profile across the 2–4
outermost layers of the {1 1 1} surface. This particular arrangement of Pt and Ni
atoms in the top three layers shifts the d-band center relative to the Fermi level of
the topmost layer of Pt atoms, thus affecting the adsorption coverage of hydroxyl
spectator species and consequently enhancing the ORR activity. Later, Gasteiger
and Markovic [250] summarized the most relevant approaches for different oxygen
reduction electrocatalysts and, based on the results obtained with Pt3Ni(1 1 1), they
estimated that, even with large 30 nm Pt3Ni octahedral nanoparticles having a Pt
skin layer, the mass activity should be about 10 times higher (about 1.6 A mg1 Pt)
than the commercial Pt/C. Since then, several contributions have been published in
the last years about the development of shape-controlled Pt-alloy nanoparticles.
However, and due to its fundamental and practical importance, in the following
section we will exclusively focus our attention on the development and ORR
activity of shape-controlled Pt-Ni nanomaterials.
The first example of shape-controlled Pt3Ni nanoparticles was reported by
Zhang et al. [251]. They prepared Pt3Ni nanoctahedra and nanocubes using a
wet-chemical approach previously reported [252]. They found that the ORR spe-
cific activity of the Pt3Ni nanoctahedra was about five times higher than that of
Pt3Ni nanocubes in good agreement with what should be expected from the results
obtained with Pt3Ni(1 1 1) single crystals, Fig. 2.20. Interestingly, in comparison
with a commercial Pt/C electrocatalysts, the specific and mass activities of Pt3Ni-
octahedra/C were about seven and four times higher, respectively, despite the three
times larger particle size of the Pt3Ni nanoctahedra.
Later, the Yang group reported some relevant contributions. They initially
showed that the ORR specific activity of truncated octahedral Pt3Ni nanoparticles
was about four times higher than that found for a reference Pt catalyst [253]. In
terms of ORR mass activity, a value of 0.53 A mg1 Pt was obtained for the Pt3Ni
sample which was about four times higher than that of the commercial Pt/C catalyst
(0.14 A mg1 Pt) and 1.8 times higher than that of carbon-supported octahedral
Pt3Ni catalysts previously used by Zhang et al. (0.3 A mg1 Pt) [251]. Subsequently,
they also reported the ORR activity of different truncated octahedral Pt3Ni
nanoparticles in alkaline solution [231, 254]. They found a specific activity of
about 0.76 mA cm2, measured at 0.9 V vs RHE, which was 4.5 times that of a
commercial Pt/C catalyst (0.17 mA cm2). A mass activity of 0.30 A mg1 Pt was
obtained which was about twice that of the commercial Pt/C catalyst. Interestingly,
they also reported a new approach for the preparation of shape-controlled Pt-M (M:
Co, Fe, Ni, Pd) alloy nanoparticles using a gas reducing agent, carbon monoxide in
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 75

Fig. 2.20 (a) ORR behaviour of Pt3Ni nanoctahedra, Pt3Ni nanocubes, and Pt nanocubes
supported on a rotating GC disk electrode in O2 saturated 0.1 M HClO4 solution at 20 mV s1
and at 295 K; rotation rate, 900 rpm. (b) Comparison of the ORR activities at 0.9 V vs RHE at
295 K. Reproduced with permission from [251], © 2010 American Chemical Society

this case, in liquid solution (GRAILS) method [231]. Using this methodology, they
prepared Pt3Ni nanocubes and nanooctahedra (about 10 nm) which were tested
toward ORR in HClO4 solution. The specific activities of the samples were
0.85 mA cm2 Pt for the nanocubes, and 1.26 mA cm2 Pt for the nanooctahedra,
which were remarkably higher than those typically obtained with a Pt/C catalyst
(0.20 mA cm2 Pt). In addition, an ORR mass activity of about 0.44 A mg1 Pt
was obtained for the octahedral Pt3Ni catalyst, three times higher than that of the
reference Pt/C (0.14 A/mgPt). Finally they reported an additional improvement by
using icosahedral Pt3Ni nanoparticles (about 13 nm) [255]. The ORR specific and
mass activities of the icosahedral samples were about 1.83 mA cm2 Pt and
0.62 A mg1 Pt, respectively. These values were higher than those previously
reported with the Pt3Ni nanooctahedra [231].
Carpenter et al. [112] reported a solvothermal method for the preparation of
shape controlled Pt based alloy nanoparticles in which DMF acted as both
solvent and reducing agent. They observed a maximum specific activity of about
3 mA cm2 with octahedral PtNi (Pt/Ni ratio about 1) nanoparticles, corresponding
with a 15 fold specific activity enhancement in comparison with the state-of-the art
Pt/C catalyst. However, in this remarkable improvement, the mass activity was only
about five times higher over Pt. Using a similar approach, Cui et al. [110] showed
that 9.5 nm octahedral PtNi nanoparticles displayed ORR specific and mass activ-
ities of 3.14 mA cm2 Pt and 1.45 A mg1 Pt at 0.9 V vs RHE in HClO4 solutions.
These values represented a tenfold specific and mass activity enhancement in
comparison with the state-of-art Pt/C catalyst. Such great enhancement was attrib-
uted not only to the {1 1 1} surface structure of the octahedral samples but also to
the particularly high (about 41 Pt at. %) near-surface Pt-to-Ni atomic surface
composition.
Later, Choi et al. [256] increased this previous enhancement using 9 nm PtNi
octahedra prepared in presence of oleylamine and oleic acid as capping agents
[252], introducing benzyl ether as solvent to decrease the surface coverage of the
76 F.J. Vidal-Iglesias et al.

Fig. 2.21 (a) CO stripping and subsequent voltammogram of octahedral Pt2.5Ni/C (HAc treat-
ment for 2 h) in a N2-saturated 0.1 M HClO4 solution at 50 mV s–1. (b) ORR behaviour of
octahedral Pt2.5Ni/C (HAc treatment for 2 h) and Pt/C in an O2-saturated 0.1 M HClO4 solution at
10 mV s–1, rotation speed 1600 rpm. (c) Mass activities and (d) specific activities (calculated from
the charges associated with HUPD, CO stripping, and CuUPD) of Pt/C, Pt3Ni/C, and octahedral
Pt2.5Ni/C at 0.9 and 0.93 V. Reproduced with permission from [256], © 2013 American Chemical
Society

surfactant on the PtNi samples, but without altering their shape/surface structure.
Additionally, the samples were treated with acetic acid (HAc) at 60  C in order to
maximise the surface cleanness as well as to modify the Pt/Ni near-surface atomic
composition. Under optimised conditions, a mass activity of 3.3 mA g1 Pt was
obtained at 0.9 V vs RHE, which is about 17 times higher than that of a Pt/C sample
and two times higher than that previously reported by Cui et al. [110]. In terms of
ORR specific activity, measured at þ0.93 V, this optimised octahedral PtNi sam-
ples showed an enhancement factor that varied from 64 to 38, as a function of the
methodology employed to determine the electroactive surface area (H UPD, CO
stripping or Cu UPD), Fig. 2.21. As previously discussed, the determination of the
real surface area of this kind of shape-controlled Pt alloy nanoparticles is a critical
point that is still under discussion [68–70]. More recently, the ORR activity of these
octahedral PtNi nanoparticles has been also evaluated as a function of the particle
size (6–12 nm) and the Pt/Ni atomic composition [257]. Octahedral PtNi
nanoparticles with a particle size of 9 nm and a Pt/Ni ratio of 2.4 displayed the
best ORR specific and mass activities.
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 77

Chen et al. [258] have very recently reported an impressive mass activity of
5.7 A mg1 Pt in HClO4 solution at 0.9 V vs RHE, using Pt3Ni nanoframes. These
Pt3Ni nanoframes were obtained after controlled erosion of rhombic dodecahedral
PtNi3 nanoparticles (about 20 nm) in nonpolar solvents. In addition, a subsequent
thermal treatment induced the formation of a smooth Pt-skin type structure. Inter-
estingly, not only the ORR activity was improved, but also its durability under ORR
operation conditions. The results showed that both the ORR activity and nanoframe
structure remained almost identical after the durability test. This stability was
attributed to the particular electronic properties of the Pt-skin surface of the
nanoframes.
Other interesting octahedral PtNi nanoparticles have been recently prepared and
evaluated towards ORR [113, 259–262], but without additional specific and mass
activity improvements than those previously discussed. Finally, and despite this
remarkable progress on both ORR specific and mass activities, it is important to
recall that, from a practical point of view, the stability is a critical issue and has been
the subject of some relevant contributions [258, 263–267].
Based on this previous discussion, interesting contributions are being published
dealing with the use of shape-controlled core-shell nanoparticles [246, 247, 268–
271] and doped shape-controlled nanoparticles for enhanced activity and stability
issues [272]. Unfortunately, this aspect will not be covered in this chapter but will
represent a step beyond in the understanding of the role of shape/surface structure
and atomic composition towards advanced ORR electrocatalysts with improved
activity and stability.

2.4 Conclusions

The use of shape-controlled metal nanoparticles in electrocatalysis has produced an


important impact of several reactions of interest, both in terms of enhanced activity,
selectivity and in some cases stability. Firstly, and due to the fact that most of these
materials are prepared in the presence of capping agents through colloidal routes,
we have discussed about the importance of surface cleanness and how this cleaning
can be achieved by using different approaches. Subsequently, we have summarized
some of the most relevant contributions on the use of these shaped clean metal
nanoparticles towards reactions of interest for low temperature fuel cells such as
formic acid, methanol and, ethanol electrooxidation and oxygen reduction. How-
ever it is worth noting that only those contributions fulfilling this surface cleaning
requirement have been considered in the present chapter. The results here reviewed
clearly point out the outstanding role played by the surface structure of these shaped
nanomaterials on their subsequent electrocatalytic properties.

Acknowledgments This work has been financially supported by the MINECO of Spain through
projects CTQ2013-44083-P and CTQ2013-48280-C3-3-R and Generalitat Valenciana through
project PROMETEOII/2014/013.
78 F.J. Vidal-Iglesias et al.

References

1. Wieckowski A (1999) Interfacial electrochemistry: theory, experiment, and applications.


Marcel Dekker, New York
2. Wieckowski A, Savinova ER, Vayenas CG (2003) Catalysis and electrocatalysis at nanopar-
ticle surfaces. CRC Press, New York
3. Koper MTM (2009) Fuel cell catalysis: a surface science approach. Electrocatalysis and
electrochemistry. Wiley, Hoboken
4. Climent V, Feliu JM (2011) Thirty years of platinum single crystal electrochemistry. J Solid
State Electrochem 15(7–8):1297–1315. doi:10.1007/s10008-011-1372-1
5. Zhou Z-Y, Tian N, Huang Z-Z, Chen D-J, Sun S-G (2009) Nanoparticle catalysts with high
energy surfaces and enhanced activity synthesized by electrochemical method. Faraday
Discuss 140:81–92
6. Proussevitch AA, Sahagian DL (2001) Recognition and separation of discrete objects within
complex 3D voxelized structures. Comput Geosci 27(4):441–454. doi:10.1016/s0098-3004
(00)00141-2
7. Tian N, Zhou ZY, Sun SG (2008) Platinum metal catalysts of high-index surfaces: from
single-crystal planes to electrochemically shape-controlled nanoparticles. J Phys Chem C 112
(50):19801–19817
8. Vidal-Iglesias FJ, Solla-Gullón J, Rodrı́guez P, Herrero E, Montiel V, Feliu JM, Aldaz A
(2004) Shape-dependent electrocatalysis: ammonia oxidation on platinum nanoparticles with
preferential (100) surfaces. Electrochem Commun 6(10):1080–1084. doi:10.1016/j.elecom.
2004.08.010
9. Solla-Gullón J, Vidal-Iglesias FJ, Rodrı́guez P, Herrero E, Feliu JM, Clavilier J, Aldaz A
(2004) In situ surface characterization of preferentially oriented platinum nanoparticles by
using electrochemical structure sensitive adsorption reactions. J Phys Chem B 108
(36):13573–13575. doi:10.1021/jp0471453
10. Ahmadi TS, Wang ZL, Green TC, Henglein A, El-Sayed MA (1996) Shape-controlled
synthesis of colloidal platinum nanoparticles. Science 272(5270):1924–1926
11. Koper MTM (2011) Structure sensitivity and nanoscale effects in electrocatalysis. Nanoscale
3(5):2054–2073
12. Solla-Gullon J, Vidal-Iglesias FJ, Feliu JM (2011) Shape dependent electrocatalysis. Annu
Rep Prog Chem, Sect C 107:263–297
13. Bing Y, Liu H, Zhang L, Ghosh D, Zhang J (2010) Nanostructured Pt-alloy electrocatalysts
for PEM fuel cell oxygen reduction reaction. Chem Soc Rev 39(6):2184–2202
14. Chen A, Holt-Hindle P (2010) Platinum-based nanostructured materials: synthesis, proper-
ties, and applications. Chem Rev 110(6):3767–3804
15. Peng Z, Yang H (2009) Designer platinum nanoparticles: control of shape, composition in
alloy, nanostructure and electrocatalytic property. Nano Today 4(2):143–164
16. Chen J, Lim B, Lee EP, Xia Y (2009) Shape-controlled synthesis of platinum nanocrystals for
catalytic and electrocatalytic applications. Nano Today 4(1):81–95
17. Vismadeb Mazumder YL, Sun S (2010) Recent development of active nanoparticle catalysts
for fuel cell reactions. Adv Funct Mater 20:1224–1234
18. Wu B, Zheng N (2013) Surface and interface control of noble metal nanocrystals for catalytic
and electrocatalytic applications. Nano Today 8(2):168–197. doi:10.1016/j.nantod.2013.02.
006
19. Kleijn SEF, Lai SCS, Koper MTM, Unwin PR (2014) Electrochemistry of nanoparticles.
Angew Chem Int Ed 53(14):3558–3586
20. You H, Yang S, Ding B, Yang H (2013) Synthesis of colloidal metal and metal alloy
nanoparticles for electrochemical energy applications. Chem Soc Rev 42(7):2880–2904
21. Sanchez-Sanchez CM, Solla-Gullon J, Montiel V (2013) Electrocatalysis at nanoparticles. In:
Electrochemistry. Nanosystems electrochemistry, vol 11. The Royal Society of Chemistry,
London, pp 34–70. doi:10.1039/9781849734820-00034
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 79

22. Dai Y, Wang Y, Liu B, Yang Y (2014) Metallic nanocatalysis: an accelerating seamless
integration with nanotechnology. Small 11(3):268–289. doi:10.1002/smll.201400847
23. Burda C, Chen X, Narayanan R, El-Sayed MA (2005) Chemistry and properties of
nanocrystals of different shapes. Chem Rev 105(4):1025–1102
24. Xia Y, Xiong Y, Lim B, Skrabalak SE (2009) Shape-controlled synthesis of metal
nanocrystals: simple chemistry meets complex physics? Angew Chem Int Ed 48(1):60–103
25. Tao AR, Habas S, Yang P (2008) Shape control of colloidal metal nanocrystals. Small 4
(3):310–325
26. Sau TK, Rogach AL (2010) Nonspherical noble metal nanoparticles: colloid-chemical
synthesis and morphology control. Adv Mater (Weinheim, Germany) 22(16):1781–1804
27. Leong GJ, Schulze MC, Strand MB, Maloney D, Frisco SL, Dinh HN, Pivovar B, Richards
RM (2014) Shape-directed platinum nanoparticle synthesis: nanoscale design of novel
catalysts. Appl Organomet Chem 28(1):1–17
28. Gu J, Zhang YW, Tao F (2012) Shape control of bimetallic nanocatalysts through well-
designed colloidal chemistry approaches. Chem Soc Rev 41(24):8050–8065
29. Clavilier J, Faure R, Guinet G, Durand R (1980) Preparation of monocrystalline. Pt micro-
electrodes and electrochemical study of the plane surfaces cut in the direction of the {111}
and {110} planes. J Electroanal Chem 107(1):205–209
30. Clavilier J (1999) Flame-annealing and cleaning technique. In: Wieckowski A (ed) Interfacial
electrochemistry. Marcel Dekker, New York, pp 231–248
31. Hara M, Linke U, Wandlowski T (2007) Preparation and electrochemical characterization of
palladium single crystal electrodes in 0.1 M H2SO4 and HClO4 Part I. Low-index phases.
Electrochim Acta 52(18):5733–5748
32. Hamelin A (1996) Cyclic voltammetry at gold single-crystal surfaces. 1. Behaviour at
low-index faces. J Electroanal Chem 407(1–2):1–11
33. Hamelin A (1996) Cyclic voltammetry at gold single-crystal surfaces. 2. Behaviour of high-
index faces. J Electroanal Chem 407(1-2):13–21
34. Ke F-S, Solomon B, Ding Y, Xu G-L, Sun S-G, Wang ZL, Zhou X-D (2014) Enhanced
electrocatalytic activity on gold nanocrystals enclosed by high-index facets for oxygen
reduction. Nano Energy 7:179–188. doi:10.1016/j.nanoen.2014.01.003
35. Yang S, Lee H (2013) Atomically dispersed platinum on gold nano-octahedra with high
catalytic activity on formic acid oxidation. ACS Catal 3(3):437–443. doi:10.1021/cs300809j
36. Nichols RJ, Magnussen OM, Hotlos J, Twomey T, Behm RJ, Kolb DM (1990) An in-situ
STM study of potential-induced changes in the surface topography of Au(100) electrodes. J
Electroanal Chem 290:21–31
37. Schneeweiss MA, Kolb DM (1997) Oxide formation on Au(111)—an in situ STM study.
Solid State Ion 94(1–4):171–179
38. Itaya K, Sugawara S, Sashikata K, Furuya N (1990) Insitu scanning tunneling microscopy of
platinum (111) surface with the observation of monatomic steps. J Vac Sci Technol A Vac
Surf Films 8(1):515–519
39. El-Deab MS (2009) On the preferential crystallographic orientation of Au nanoparticles:
effect of electrodeposition time. Electrochim Acta 54(14):3720–3725
40. El-Deab MS, Sotomura T, Ohsaka T (2005) Size and crystallographic orientation controls of
gold nanoparticles electrodeposited on GC electrodes. J Electrochem Soc 152(1):C1–C6
41. El-Deab MS, Arihara K, Ohsaka T (2004) Fabrication of Au(111)-like polycrystalline gold
electrodes and their applications to oxygen reduction. J Electrochem Soc 151(6):E213–E218
42. Walczak MM, Alves CA, Lamp BD, Porter MD (1995) Electrochemical and X-ray photo-
electron spectroscopic evidence for differences in the binding sites of alkanethiolate mono-
layers chemisorbed at gold. J Electroanal Chem 396(1–2):103–114
43. Zhong CJ, Zak J, Porter MD (1997) Voltammetric reductive desorption characteristics of
alkanethiolate monolayers at single crystal Au(111) and (110) electrode surfaces. J
Electroanal Chem 421(1–2):9–13
80 F.J. Vidal-Iglesias et al.

44. Herrero E, Buller LJ, Abruna HD (2001) Underpotential deposition at single crystal surfaces
of Au, Pt, Ag and other materials. Chem Rev 101(7):1897–1930. doi:10.1021/cr9600363
45. Hernández J, Solla-Gullón J, Herrero E, Aldaz A, Feliu JM (2007) Electrochemistry of shape-
controlled catalysts: oxygen reduction reaction on cubic gold nanoparticles. J Phys Chem C
111(38):14078–14083. doi:10.1021/jp0749726CCC
46. Hernández J, Solla-Gullón J, Herrero E, Feliu JM, Aldaz A (2009) In situ surface character-
ization and oxygen reduction reaction on shape-controlled gold nanoparticles. J Nanosci
Nanotechnol 9(4):2256–2273. doi:10.1166/jnn.2009.SE38
47. Hernández J, Solla-Gullón J, Herrero E (2004) Gold nanoparticles synthesized in a water-in-
oil microemulsion: electrochemical characterization and effect of the surface structure on the
oxygen reduction reaction. J Electroanal Chem 574(1):185–196. doi:10.1016/j.jelechem.
2003.10.039
48. Hamelin A (1979) Lead adsorption on gold single crystal stepped surfaces. J Electroanal
Chem 101(2):285–290
49. Hamelin A, Katayama A (1981) Lead underpotential deposition on gold single-crystal
surfaces: the (100) face and its vicinal faces. J Electroanal Chem 117(2):221–232
50. Hamelin A (1984) Underpotential deposition of lead on single crystal faces of gold. Part
I. The influence of crystallographic orientation of the substrate. J Electroanal Chem 165
(1–2):167–180
51. Hamelin A, Lipkowski J (1984) Underpotential deposition of lead on gold single crystal
faces. Part II. General discussion. J Electroanal Chem 171(1–2):317–330
52. Chierchie T, Mayer C (1988) Voltammetric study of the underpotential deposition of copper
on polycrystalline and single crystal palladium surfaces. Electrochim Acta 33(3):341–345
53. Cuesta A, Kibler LA, Kolb DM (1999) A method to prepare single crystal electrodes of
reactive metals: application to Pd(hkl). J Electroanal Chem 466(2):165–168
54. Fang LL, Tao Q, Li MF, Liao LW, Chen D, Chen YX (2010) Determination of the real
surface area of palladium electrode. Chin J Chem Phys 23(5):543–548
55. Francke R, Climent V, Baltruschat H, Feliu JM (2008) Electrochemical deposition of copper
on stepped platinum surfaces in the 01(1)over-bar zone vicinal to the (100) plane. J
Electroanal Chem 624(1–2):228–240
56. Danilov AI, Molodkina EB, Rudnev AV, Polukarov YM, Feliu JM (2005) Kinetics of copper
deposition on Pt(111) and Au(111) electrodes in solutions of different acidities. Electrochim
Acta 50(25–26):5032–5043. doi:10.1016/j.electacta.2005.02.078
57. Solla-Gullón J, Rodrı́guez P, Herrero E, Aldaz A, Feliu JM (2008) Surface characterization of
platinum electrodes. Phys Chem Chem Phys 10(10):1359–1373. doi:10.1039/b709809j
58. Rodrı́guez P, Solla-Gullón J, Vidal-Iglesias FJ, Herrero E, Aldaz A, Feliu JM (2005)
Determination of (111) ordered domains on platinum electrodes by irreversible adsorption
of bismuth. Anal Chem 77(16):5317–5323. doi:10.1021/ac050347q
59. Rodrı́guez P, Herrero E, Solla-Gullón J, Vidal-Iglesias FJ, Aldaz A, Feliu JM (2005) Specific
surface reactions for identification of platinum surface domains—surface characterization
and electrocatalytic tests. Electrochim Acta 50(21):4308–4317. doi:10.1016/j.electacta.2005.
02.087
60. Rodrı́guez P, Herrero E, Aldaz A, Feliu JM (2006) Tellurium adatoms as an in-situ surface
probe of (111) two-dimensional domains at platinum surfaces. Langmuir 22
(25):10329–10337. doi:10.1021/la060981e
61. Rodrı́guez P, Herrero E, Solla-Gullón J, Vidal-Iglesias FJ, Aldaz A, Feliu JM (2005)
Electrochemical characterization of irreversibly adsorbed germanium on platinum stepped
surfaces vicinal to Pt(100). Electrochim Acta 50(15):3111–3121. doi:10.1016/j.electacta.
2004.10.086
62. Brimaud S, Pronier S, Coutanceau C, Léger JM (2008) New findings on CO electrooxidation
at platinum nanoparticle surfaces. Electrochem Commun 10(11):1703–1707
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 81

63. Devivaraprasad R, Ramesh R, Naresh N, Kar T, Singh RK, Neergat M (2014) Oxygen
reduction reaction and peroxide generation on shape-controlled and polycrystalline platinum
nanoparticles in acidic and alkaline electrolytes. Langmuir 30(29):8995–9006
64. Bertin E, Garbarino S, Guay D (2014) Formic acid oxidation on Bi covered Pt
electrodeposited thin films: influence of the underlying structure. Electrochim Acta
134:486–495. doi:10.1016/j.electacta.2014.04.111
65. Levendorf AM, Chen D-J, Rom CL, Liu Y, Tong YJ (2014) Electrochemical and in situ
ATR-SEIRAS investigations of methanol and CO electro-oxidation on PVP-free cubic and
octahedral/tetrahedral Pt nanoparticles. RSC Adv 4(41):21284–21293. doi:10.1039/
c4ra00815d
66. Rodriguez-Lopez M, Solla-Gullon J, Herrero E, Tunon P, Feliu JM, Aldaz A, Carrasquillo A
(2010) Electrochemical reactivity of aromatic molecules at nanometer-sized surface
domains: from Pt(hkl) single crystal electrodes to preferentially oriented platinum
nanoparticles. J Am Chem Soc 132(7):2233–2242. doi:10.1021/Ja909082s
67. Martı́nez-Rodrı́guez RA, Vidal-Iglesias FJ, Solla-Gullón J, Cabrera CR, Feliu JM (2014)
Synthesis and electrocatalytic properties of H2SO4-induced (100) Pt nanoparticles prepared
in water-in-oil microemulsion. ChemPhysChem 15(10):1997–2001. doi:10.1002/cphc.
201400056
68. Van Der Vliet DF, Wang C, Li D, Paulikas AP, Greeley J, Rankin RB, Strmcnik D,
Tripkovic D, Markovic NM, Stamenkovic VR (2012) Unique electrochemical adsorption
properties of Pt-skin surfaces. Angew Chem Int Ed 51(13):3139–3142
69. Shao M, Odell JH, Choi SI, Xia Y (2013) Electrochemical surface area measurements of
platinum- and palladium-based nanoparticles. Electrochem Commun 31:46–48
70. Rudi S, Cui C, Gan L, Strasser P (2014) Comparative study of the electrocatalytically active
surface areas (ECSAs) of Pt alloy nanoparticles evaluated by Hupd and CO-stripping
voltammetry. Electrocatalysis 5(4):408–418
71. Solla-Gullón J, Vidal-Iglesias FJ, Herrero E, Feliu JM, Aldaz A (2006) CO monolayer
oxidation on semi-spherical and preferentially oriented (100) and (111) platinum
nanoparticles. Electrochem Commun 8(1):189–194. doi:10.1016/j.elecom.2005.11.008
72. Solla-Gullón J, Vidal-Iglesias FJ, López-Cudero A, Garnier E, Feliu JM, Aldaz A (2008)
Shape-dependent electrocatalysis: methanol and formic acid electrooxidation on preferen-
tially oriented Pt nanoparticles. Phys Chem Chem Phys 10(25):3689–3698. doi:10.1039/
b802703j
73. Vidal-Iglesias FJ, Aran-Ais RM, Solla-Gullon J, Herrero E, Feliu JM (2012) Electrochemical
characterization of shape-controlled Pt nanoparticles in different supporting electrolytes.
ACS Catal 2(5):901–910. doi:10.1021/Cs200681x
74. Erikson H, Sarapuu A, Tammeveski K, Solla-Gullón J, Feliu JM (2011) Enhanced
electrocatalytic activity of cubic Pd nanoparticles towards the oxygen reduction reaction in
acid media. Electrochem Commun 13(7):734–737
75. Erikson H, Sarapuu A, Alexeyeva N, Tammeveski K, Solla-Gullón J, Feliu JM (2012)
Electrochemical reduction of oxygen on palladium nanocubes in acid and alkaline solutions.
Electrochim Acta 59:329–335. doi:10.1016/j.electacta.2011.10.074
76. Vidal-Iglesias FJ, Aran-Ais RM, Solla-Gullon J, Garnier E, Herrero E, Aldaz A, Feliu JM
(2012) Shape-dependent electrocatalysis: formic acid electrooxidation on cubic Pd
nanoparticles. Phys Chem Chem Phys 14(29):10258–10265. doi:10.1039/c2cp40992e
77. Brimaud S, Jusys Z, Behm RJ (2014) Shape-selected nanocrystals for in situ spectro-
electrochemistry studies on structurally well defined surfaces under controlled electrolyte
transport: a combined in situ ATR-FTIR/online DEMS investigation of CO electrooxidation
on Pt. Beilstein J Nanotechnol 5:735–746. doi:10.3762/bjnano.5.86
78. Brimaud S, Jusys Z, Behm RJ (2011) Controlled surface structure for in situ ATR-FTIRS
studies using preferentially shaped Pt nanocrystals. Electrocatalysis 2(2):69–74. doi:10.1007/
s12678-011-0040-7
82 F.J. Vidal-Iglesias et al.

79. Coutanceau C, Urchaga P, Brimaud S, Baranton S (2012) Colloidal syntheses of shape- and
size-controlled Pt nanoparticles for electrocatalysis. Electrocatalysis 3(2):75–87. doi:10.
1007/s12678-012-0079-0
80. Urchaga P, Baranton S, Napporn TW, Coutanceau C (2010) Selective syntheses and electro-
chemical characterization of platinum nanocubes and nanotetrahedrons/octahedrons.
Electrocatalysis 1:3–6
81. Susut C, Tong Y (2011) Size-dependent methanol electro-oxidation activity of Pt
nanoparticles with different shapes. Electrocatalysis 2(2):75–81. doi:10.1007/s12678-011-
0041-6
82. Susut C, Chapman GB, Samjeské G, Osawa M, Tong Y (2008) An unexpected enhancement
in methanol electro-oxidation on an ensemble of Pt(111) nanofacets: a case of nanoscale
single crystal ensemble electrocatalysis. Phys Chem Chem Phys 10(25):3712–3721
83. Susut C, Nguyen TD, Chapman GB, Tong Y (2008) Shape and size stability of Pt
nanoparticles for MeOH electro-oxidation. Electrochim Acta 53(21):6135–6142
84. Song H, Kim F, Connor S, Somorjai GA, Yang P (2005) Pt nanocrystals: shape control and
Langmuir-Blodgett monolayer formation. J Phys Chem B 109(1):188–193
85. Monzó J, Koper MTM, Rodriguez P (2012) Removing polyvinylpyrrolidone from catalytic Pt
nanoparticles without modification of superficial order. ChemPhysChem 13(3):709–715
86. Yang H, Tang Y, Zou S (2014) Electrochemical removal of surfactants from Pt nanocubes.
Electrochem Commun 38:134–137. doi:10.1016/j.elecom.2013.11.019
87. Arán-Ais RM, Vidal-Iglesias FJ, Solla-Gullon J, Herrero E, Feliu JM (2014) Electrochemical
characterization of clean shape-controlled Pt nanoparticles prepared in presence of
oleylamine/oleic acid. Electroanalysis 27(4):945–956
88. Zalineeva A, Baranton S, Coutanceau C (2013) Bi-modified palladium nanocubes for glyc-
erol electrooxidation. Electrochem Commun 34:335–338
89. Zalineeva A, Baranton S, Coutanceau C, Jerkiewicz G (2015) Electrochemical behavior of
unsupported shaped palladium nanoparticles. Langmuir 31(5):1605–1609. doi:10.1021/
la5025229
90. Nalajala N, Gooty Saleha WF, Ladewig BP, Neergat M (2014) Sodium borohydride treat-
ment: a simple and effective process for the removal of stabilizer and capping agents from
shape-controlled palladium nanoparticles. Chem Commun 50(66):9365–9368
91. Naresh N, Wasim FGS, Ladewig BP, Neergat M (2013) Removal of surfactant and capping
agent from Pd nanocubes (Pd-NCs) using tert-butylamine: its effect on electrochemical
characteristics. J Mater Chem A 1(30):8553–8559
92. Hernández J, Solla-Gullón J, Herrero E, Aldaz A, Feliu JM (2005) Characterization of the
surface structure of gold nanoparticles and nanorods using structure sensitive reactions. J
Phys Chem B 109(26):12651–12654. doi:10.1021/jp0521609
93. Park JY, Aliaga C, Renzas JR, Lee H, Somorjai GA (2009) The role of organic capping layers
of platinum nanoparticles in catalytic activity of CO oxidation. Catal Lett 129(1-2):1–6
94. Aliaga C, Park JY, Yamada Y, Lee HS, Tsung CK, Yang P, Somorjai GA (2009) Sum
frequency generation and catalytic reaction studies of the removal of organic capping agents
from Pt nanoparticles by UV-Ozone treatment. J Phys Chem C 113(15):6150–6155
95. Krier JM, Michalak WD, Baker LR, An K, Komvopoulos K, Somorjai GA (2012) Sum
frequency generation vibrational spectroscopy of colloidal platinum nanoparticle catalysts:
disordering versus removal of organic capping. J Phys Chem C 116(33):17540–17546
96. Crespo-Quesada M, Andanson JM, Yarulin A, Lim B, Xia Y, Kiwi-Minsker L (2011)
UV-ozone cleaning of supported poly(vinylpyrrolidone)-stabilized palladium nanocubes:
effect of stabilizer removal on morphology and catalytic behavior. Langmuir 27
(12):7909–7916
97. Vidal-Iglesias FJ, Solla-Gullón J, Herrero E, Montiel V, Aldaz A, Feliu JM (2011) Evaluating
the ozone cleaning treatment in shape-controlled Pt nanoparticles: evidences of atomic surface
disordering. Electrochem Commun 13(5):502–505. doi:10.1016/j.elecom.2011.02.033
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 83

98. Levendorf A, Sun S-G, Tong Y (2014) In situ FT-IR investigation of methanol and CO
electrooxidation on cubic and octahedral/tetrahedral Pt nanoparticles having residual PVP.
Electrocatalysis 5(3):248–255. doi:10.1007/s12678-014-0186-1
99. Tong YJ (2012) Unconventional promoters of catalytic activity in electrocatalysis. Chem Soc
Rev 41(24):8195–8209
100. Susut C, Chen DJ, Sun SG, Tong YJ (2011) Capping polymer-enhanced electrocatalytic
activity on Pt nanoparticles: a combined electrochemical and in situ IR spectroelectro-
chemical study. Phys Chem Chem Phys 13(16):7467–7474
101. Shen J, Ziaei-Azad H, Semagina N (2014) Is it always necessary to remove a metal
nanoparticle stabilizer before catalysis? J Mol Catal A Chem 391(1):36–40
102. Chung YH, Chung DY, Jung N, Sung YE (2013) Tailoring the electronic structure of
nanoelectrocatalysts induced by a surface-capping organic molecule for the oxygen reduction
reaction. J Phys Chem Lett 4(8):1304–1309
103. Miyabayashi K, Nishihara H, Miyake M (2014) Platinum nanoparticles modified with
alkylamine derivatives as an active and stable catalyst for oxygen reduction reaction.
Langmuir 30(10):2936–2942
104. Niu Z, Li Y (2013) Removal and utilization of capping agents in nanocatalysis. Chem Mater
26(1):72–83. doi:10.1021/cm4022479
105. Lee H (2014) Utilization of shape-controlled nanoparticles as catalysts with enhanced
activity and selectivity. RSC Adv 4(77):41017–41027
106. Newton JE, Preece JA, Rees NV, Horswell SL (2014) Nanoparticle catalysts for proton
exchange membrane fuel cells: can surfactant effects be beneficial for electrocatalysis? Phys
Chem Chem Phys 16(23):11435–11446
107. Du L, Zhang S, Chen G, Yin G, Du C, Tan Q, Sun Y, Qu Y, Gao Y (2014) Polyelectrolyte
assisted synthesis and enhanced oxygen reduction activity of Pt nanocrystals with controlla-
ble shape and size. ACS Appl Mater Interfaces 6(16):14043–14049
108. Tian N, Zhou Z-Y, Sun S-G, Ding Y, Wang ZL (2007) Synthesis of tetrahexahedral platinum
nanocrystals with high-index facets and high electro-oxidation activity. Science 316
(5825):732–735. doi:10.1126/science.1140484
109. Tian N, Xiao J, Zhou Z-Y, Liu H-X, Deng Y-J, Huang L, Xu B-B, Sun S-G (2013) Pt-group
bimetallic nanocrystals with high-index facets as high performance electrocatalysts. Faraday
Discuss 162:77–89. doi:10.1039/c3fd20146e
110. Cui C, Gan L, Li HH, Yu SH, Heggen M, Strasser P (2012) Octahedral PtNi nanoparticle
catalysts: exceptional oxygen reduction activity by tuning the alloy particle surface compo-
sition. Nano Lett 12(11):5885–5889
111. Gumeci C, Marathe A, Behrens RL, Chaudhuri J, Korzeniewski C (2014) Solvothermal
synthesis and electrochemical characterization of shape-controlled Pt nanocrystals. J Phys
Chem C 118(26):14433–14440
112. Carpenter MK, Moylan TE, Kukreja RS, Atwan MH, Tessema MM (2012) Solvothermal
synthesis of platinum alloy nanoparticles for oxygen reduction electrocatalysis. J Am Chem
Soc 134(20):8535–8542
113. Zhang C, Hwang SY, Trout A, Peng Z (2014) Solid-state chemistry-enabled scalable
production of octahedral Pt-Ni alloy electrocatalyst for oxygen reduction reaction. J Am
Chem Soc 136(22):7805–7808
114. Shao M (2013) Electrocatalysis in fuel cells: a non- and low-platinum approach. Springer,
London
115. Feliu JM, Herrero E (2003) Formic acid oxidation. In: Vielstich W, Gasteiger H, Lamm A
(eds) Handbook of fuel cells, vol 2, Fundamentals, technology and applications. Wiley,
Chichester
116. Jiang K, Zhang HX, Zou S, Cai WB (2014) Electrocatalysis of formic acid on palladium and
platinum surfaces: from fundamental mechanisms to fuel cell applications. Phys Chem Chem
Phys 16(38):20360–20376
84 F.J. Vidal-Iglesias et al.

117. Capon A, Parsons R (1973) The oxidation of formic acid at noble metal electrodes: I. Review
of previous work. J Electroanal Chem 44(1):1–7
118. Capon A, Parsons R (1973) The oxidation of formic acid on noble metal electrodes: II. A
comparison of the behaviour of pure electrodes. J Electroanal Chem 44(2):239–254
119. Capon A, Parsons R (1973) The oxidation of formic acid at noble metal electrodes Part III.
Intermediates and mechanism on platinum electrodes. J Electroanal Chem 45(2):205–231
120. Grozovski V, Solla-Gullon J, Climent V, Herrero E, Feliu JM (2010) Formic acid oxidation
on shape-controlled Pt nanoparticles studied by pulsed voltammetry. J Phys Chem C 114
(32):13802–13812. doi:10.1021/jp104755b
121. Huang X, Zhao Z, Fan J, Tan Y, Zheng N (2011) Amine-assisted synthesis of concave
polyhedral platinum nanocrystals having {411} high-index facets. J Am Chem Soc 133
(13):4718–4721
122. Li Y, Jiang Y, Chen M, Liao H, Huang R, Zhou Z, Tian N, Chen S, Sun S (2012)
Electrochemically shape-controlled synthesis of trapezohedral platinum nanocrystals with
high electrocatalytic activity. Chem Commun 48(76):9531–9533. doi:10.1039/c2cc34322c
123. Zhang Z-C, Hui J-F, Liu Z-C, Zhang X, Zhuang J, Wang X (2012) Glycine-mediated
syntheses of Pt concave nanocubes with high-index {hk0} facets and their enhanced
electrocatalytic activities. Langmuir 28(42):14845–14848. doi:10.1021/la302973r
124. Xia BY, Wu HB, Wang X, Lou XW (2013) Highly concave platinum nanoframes with high-
index facets and enhanced electrocatalytic properties. Angew Chem Int Ed 52
(47):12337–12340. doi:10.1002/anie.201307518
125. Lu GQ, Crown A, Wieckowski A (1999) Formic acid decomposition on polycrystalline
platinum and palladized platinum electrodes. J Phys Chem B 103(44):9700–9711
126. Rice C, Ha S, Masel RI, Wieckowski A (2003) Catalysts for direct formic acid fuel cells. J
Power Sources 115(2):229–235. doi:10.1016/s0378-7753(03)00026-0
127. Rice C, Ha RI, Masel RI, Waszczuk P, Wieckowski A, Barnard T (2002) Direct formic acid
fuel cells. J Power Sources 111(1):83–89
128. Hoshi N, Kida K, Nakamura M, Nakada M, Osada K (2006) Structural effects of electro-
chemical oxidation of formic acid on single crystal electrodes of palladium. J Phys Chem B
110(25):12480–12484
129. Baldauf M, Kolb DM (1996) Formic acid oxidation on ultrathin Pd films on Au(hkl) and Pt
(hkl) electrodes. J Phys Chem 100(27):11375–11381
130. Jin M, Zhang H, Xie Z, Xia Y (2012) Palladium nanocrystals enclosed by {100} and {111}
facets in controlled proportions and their catalytic activities for formic acid oxidation. Energy
Environ Sci 5(4):6352–6357
131. Zhang H-X, Wang H, Re Y-S, Cai W-B (2012) Palladium nanocrystals bound by {110} or
{100} facets: from one pot synthesis to electrochemistry. Chem Commun 48(67):8362–8364.
doi:10.1039/c2cc33941b
132. Shao Z, Zhu W, Wang H, Yang Q, Yang S, Liu X, Wang G (2013) Controllable synthesis of
concave nanocubes, right bipyramids, and 5-fold twinned nanorods of palladium and their
enhanced electrocatalytic performance. J Phys Chem C 117(27):14289–14294
133. Kuo CH, Lamontagne LK, Brodsky CN, Chou LY, Zhuang J, Sneed BT, Sheehan MK, Tsung
CK (2013) The effect of lattice strain on the catalytic properties of Pd nanocrystals.
Chemsuschem 6(10):1993–2000
134. Xia X, Choi SI, Herron JA, Lu N, Scaranto J, Peng HC, Wang J, Mavrikakis M, Kim MJ, Xia
Y (2013) Facile synthesis of palladium right bipyramids and their use as seeds for overgrowth
and as catalysts for formic acid oxidation. J Am Chem Soc 135(42):15706–15709
135. LV T, Wang Y, Choi SI, Chi M, Tao J, Pan L, Huang CZ, Zhu Y, Xia Y (2013) Controlled
synthesis of nanosized palladium icosahedra and their catalytic activity towards formic-acid
oxidation. Chemsuschem 6(10):1923–1930
136. Shao M, Odell J, Humbert M, Yu T, Xia Y (2013) Electrocatalysis on shape-controlled
palladium nanocrystals: oxygen reduction reaction and formic acid oxidation. J Phys Chem C
117(8):4172–4180
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 85

137. Zhang X, Yin H, Wang J, Chang L, Gao Y, Liu W, Tang Z (2013) Shape-dependent
electrocatalytic activity of monodispersed palladium nanocrystals toward formic acid oxida-
tion. Nanoscale 5(18):8392–8397. doi:10.1039/c3nr03100d
138. Zhang J, Feng C, Deng Y, Liu L, Wu Y, Shen B, Zhong C, Hu W (2014) Shape-controlled
synthesis of palladium single-crystalline nanoparticles: the effect of HCl oxidative etching
and facet-dependent catalytic properties. Chem Mater 26(2):1213–1218. doi:10.1021/
cm403591g
139. Tang Y, Edelmann RE, Zou S (2014) Length tunable penta-twinned palladium nanorods:
seedless synthesis and electrooxidation of formic acid. Nanoscale 6(11):5630–5633. doi:10.
1039/c4nr00299g
140. Nørskov JK, Abild-Pedersen F, Studt F, Bligaard T (2011) Density functional theory in
surface chemistry and catalysis. Proc Natl Acad Sci 108(3):937–943. doi:10.1073/pnas.
1006652108
141. Boronat-Gonzalez A, Herrero E, Feliu JM (2014) Fundamental aspects of HCOOH oxidation
at platinum single crystal surfaces with basal orientations and modified by irreversibly
adsorbed adatoms. J Solid State Electrochem 18(5):1181–1193. doi:10.1007/s10008-013-
2209-x
142. Chen QS, Zhou ZY, Vidal-Iglesias FJ, Solla-Gullon J, Feliu JM, Sun SG (2011) Significantly
enhancing catalytic activity of tetrahexahedral Pt nanocrystals by Bi adatom decoration. J Am
Chem Soc 133(33):12930–12933. doi:10.1021/ja2042029
143. López-Cudero A, Vidal-Iglesias FJ, Solla-Gullón J, Herrero E, Aldaz A, Feliu JM (2009)
Formic acid electrooxidation on Bi-modified polyoriented and preferential (111) Pt
nanoparticles. Phys Chem Chem Phys 11(2):416–424. doi:10.1039/b814072c
144. Vidal-Iglesias FJ, Solla-Gullon J, Herrero E, Aldaz A, Feliu JM (2010) Pd adatom decorated
(100) preferentially oriented Pt nanoparticles for formic acid electrooxidation. Angew Chem
Int Ed 49(39):6998–7001. doi:10.1002/anie.201002501
145. Vidal-Iglesias FJ, Lopez-Cudero A, Solla-Gullon J, Aldaz A, Feliu JM (2012) Pd-modified
shape-controlled Pt nanoparticles towards formic acid electrooxidation. Electrocatalysis 3
(3–4):313–323. doi:10.1007/s12678-012-0094-1
146. Vidal-Iglesias FJ, López-Cudero A, Solla-Gullón J, Feliu JM (2013) Towards more active
and stable electrocatalysts for formic acid electrooxidation: antimony-decorated octahedral
platinum nanoparticles. Angew Chem Int Ed 52(3):964–967. doi:10.1002/anie.201207517
147. Buso-Rogero C, Perales-Rondon JV, Farias MJS, Vidal-Iglesias FJ, Solla-Gullon J,
Herrero E, Feliu JM (2014) Formic acid electrooxidation on thallium-decorated shape-
controlled platinum nanoparticles: an improvement in electrocatalytic activity. Phys Chem
Chem Phys 16(27):13616–13624. doi:10.1039/c4cp00304g
148. Liu HX, Tian N, Brandon MP, Pei J, Huangfu ZC, Zhan C, Zhou ZY, Hardacre C, Lin WF,
Sun SG (2012) Enhancing the activity and tuning the mechanism of formic acid oxidation at
tetrahexahedral Pt nanocrystals by Au decoration. Phys Chem Chem Phys 14
(47):16415–16423
149. Yuan Q, Zhou Z, Zhuang J, Wang X (2010) Pd-Pt random alloy nanocubes with tunable
compositions and their enhanced electrocatalytic activities. Chem Commun 46
(9):1491–1493. doi:10.1039/b922792j
150. Zhang J, Yang H, Yang K, Fang J, Zou S, Luo Z, Wang H, Bae IT, Jung DY (2010)
Monodisperse Pt3Fe nanocubes: synthesis, characterization, self-assembly, and
electrocatalytic activity. Adv Funct Mater 20(21):3727–3733
151. Yang H, Dai L, Xu D, Fang J, Zou S (2010) Electrooxidation of methanol and formic acid on
PtCu nanoparticles. Electrochim Acta 55(27):8000–8004. doi:10.1016/j.electacta.2010.03.026
152. Xu D, Bliznakov S, Liu Z, Fang J, Dimitrov N (2010) Composition-dependent
electrocatalytic activity of Pt-Cu nanocube catalysts for formic acid oxidation. Angew
Chem Int Ed 49(7):1282–1285. doi:10.1002/anie.200905248
86 F.J. Vidal-Iglesias et al.

153. Zhang J, Yang H, Martens B, Luo Z, Xu D, Wang Y, Zou S, Fang J (2012) Pt-Cu
nanoctahedra: synthesis and comparative study with nanocubes on their electrochemical
catalytic performance. Chem Sci 3(11):3302–3306
154. Bromberg L, Fayette M, Martens B, Luo ZP, Wang Y, Xu D, Zhang J, Fang J, Dimitrov N
(2013) Catalytic performance comparison of shape-dependent nanocrystals and oriented
ultrathin films of Pt4Cu alloy in the formic acid oxidation process. Electrocatalysis 4
(1):24–36
155. Quan Z, Wang Y, Fang J (2013) High-index faceted noble metal nanocrystals. Acc Chem Res
46(2):191–202
156. Porter NS, Wu H, Quan Z, Fang J (2013) Shape-control and electrocatalytic activity-
enhancement of Pt-based bimetallic nanocrystals. Acc Chem Res 46(8):1867–1877. doi:10.
1021/ar3002238
157. Kang Y, Murray CB (2010) Synthesis and electrocatalytic properties of cubic Mn–Pt
nanocrystals (nanocubes). J Am Chem Soc 132(22):7568–7569. doi:10.1021/ja100705j
158. Yu Y, Zhang Q, Liu B, Lee JY (2010) Synthesis of nanocrystals with variable high-index Pd
facets through the controlled heteroepitaxial growth of trisoctahedral Au templates. J Am
Chem Soc 132(51):18258–18265
159. Zhang L, Zhang J, Kuang Q, Xie S, Jiang Z, Xie Z, Zheng L (2011) Cu2 þ -assisted synthesis
of hexoctahedral Au-Pd alloy nanocrystals with high-index facets. J Am Chem Soc 133
(43):17114–17117
160. Deng YJ, Tian N, Zhou ZY, Huang R, Liu ZL, Xiao J, Sun SG (2012) Alloy tetrahexahedral
Pd-Pt catalysts: enhancing significantly the catalytic activity by synergy effect of high-index
facets and electronic structure. Chem Sci 3(4):1157–1161
161. Zhang ZC, Hui JF, Guo ZG, Yu QY, Xu B, Zhang X, Liu ZC, Xu CM, Gao JS, Wang X
(2012) Solvothermal synthesis of Pt-Pd alloys with selective shapes and their enhanced
electrocatalytic activities. Nanoscale 4(8):2633–2639
162. Jia Y, Jiang Y, Zhang J, Zhang L, Chen Q, Xie Z, Zheng L (2014) Unique excavated rhombic
dodecahedral PtCu3 alloy nanocrystals constructed with ultrathin nanosheets of high-energy
{110} facets. J Am Chem Soc 136(10):3748–3751. doi:10.1021/ja413209q
163. Sneed BT, Young AP, Jalalpoor D, Golden MC, Mao S, Jiang Y, Wang Y, Tsung C-K (2014)
Shaped Pd–Ni–Pt core-sandwich-shell nanoparticles: influence of Ni sandwich layers on
catalytic electrooxidations. ACS Nano 8(7):7239–7250. doi:10.1021/nn502259g
164. Zhang L, Choi S-I, Tao J, Peng H-C, Xie S, Zhu Y, Xie Z, Xia Y (2014) Pd-Cu bimetallic
tripods: a mechanistic understanding of the synthesis and their enhanced electrocatalytic
activity for formic acid oxidation. Adv Funct Mater 24(47):7520–7529. doi:10.1002/adfm.
201402350
165. Li M, Adzic R (2013) Low-platinum-content electrocatalysts for methanol and ethanol
electrooxidation. In: Shao M (ed) Electrocatalysis in fuel cells, vol 9, Lecture notes in energy.
Springer, London, pp 1–25. doi:10.1007/978-1-4471-4911-8_1
166. Xia XH, Iwasita T, Ge F, Vielstich W (1996) Structural effects and reactivity in methanol
oxidation on polycrystal line and single crystal platinum. Electrochim Acta 41(5):711–718
167. Lamy C, Leger JM, Clavilier J, Parsons R (1983) Structural effects in electrocatalysis—a
comparative-study of the oxidation of CO, HCOOH and CH3OH on single-crystal Pt elec-
trodes. J Electroanal Chem 150(1-2):71–77
168. Hofstead-Duffy AM, Chen D-J, Sun S-G, Tong YJ (2012) Origin of the current peak of
negative scan in the cyclic voltammetry of methanol electro-oxidation on Pt-based
electrocatalysts: a revisit to the current ratio criterion. J Mater Chem 22(11):5205–5208.
doi:10.1039/c2jm15426a
169. Chen G, Tan Y, Wu B, Fu G, Zheng N (2012) Carbon monoxide-controlled synthesis of
surface-clean Pt nanocubes with high electrocatalytic activity. Chem Commun 48
(22):2758–2760. doi:10.1039/c2cc17984a
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 87

170. Chen J, Mao J, Zhao J, Ren M, Wei M (2014) Surfactant-free platinum nanocubes with
greatly enhanced activity towards methanol/ethanol electrooxidation. RSC Adv 4
(55):28832–28835. doi:10.1039/c4ra03446e
171. Liu H-X, Tian N, Brandon MP, Zhou Z-Y, Lin J-L, Hardacre C, Lin W-F, Sun S-G (2012)
Tetrahexahedral Pt nanocrystal catalysts decorated with Ru adatoms and their enhanced
activity in methanol electrooxidation. ACS Catal 2(5):708–715. doi:10.1021/cs200686a
172. Arjona N, Guerra-Balcázar M, Ortiz-Frade L, Osorio-Monreal G, Álvarez-Contreras L,
Ledesma-Garcı́a J, Arriaga LG (2013) Electrocatalytic activity of well-defined and homoge-
neous cubic-shaped Pd nanoparticles. J Mater Chem A 1(48):15524–15529
173. Kannan P, Maiyalagan T, Opallo M (2013) One-pot synthesis of chain-like palladium
nanocubes and their enhanced electrocatalytic activity for fuel-cell applications. Nano
Energy 2(5):677–687. doi:10.1016/j.nanoen.2013.08.001
174. Qin Y-L, Zhang X-B, Wang J, Wang L-M (2012) Rapid and shape-controlled synthesis of
“clean” star-like and concave Pd nanocrystallites and their high performance toward meth-
anol oxidation. J Mater Chem 22(30):14861–14863. doi:10.1039/c2jm32682e
175. Wang A-J, Li F-F, Zheng J-N, Xi H-X, Meng Z-Y, Feng J-J (2013) Green synthesis of porous
flower-like palladium with high electrocatalytic activity towards methanol oxidation. RSC
Adv 3(26):10355–10362. doi:10.1039/c3ra40556g
176. Chen X, Cai Z, Chen X, Oyama M (2014) Synthesis of bimetallic PtPd nanocubes on
graphene with N,N-dimethylformamide and their direct use for methanol electrocatalytic
oxidation. Carbon 66:387–394. doi:10.1016/j.carbon.2013.09.014
177. Yin AX, Min XQ, Zhang YW, Yan CH (2011) Shape-selective synthesis and facet-dependent
enhanced electrocatalytic activity and durability of monodisperse sub-10 nm Pt-Pd tetrahe-
drons and cubes. J Am Chem Soc 133(11):3816–3819
178. Lee Y-W, Ko AR, Han S-B, Kim H-S, Park K-W (2011) Synthesis of octahedral Pt-Pd alloy
nanoparticles for improved catalytic activity and stability in methanol electrooxidation. Phys
Chem Chem Phys 13(13):5569–5572. doi:10.1039/c0cp02167a
179. Yin A-X, Min X-Q, Zhu W, Wu H-S, Zhang Y-W, Yan C-H (2012) Multiply twinned Pt-Pd
nanoicosahedrons as highly active electrocatalysts for methanol oxidation. Chem Commun
48(4):543–545. doi:10.1039/c1cc16482a
180. Zhan F, Bian T, Zhao W, Zhang H, Jin M, Yang D (2014) Facile synthesis of Pd-Pt alloy
concave nanocubes with high-index facets as electrocatalysts for methanol oxidation.
CrystEngComm 16(12):2411–2416. doi:10.1039/c3ce42362j
181. Zhang J, Xi C, Feng C, Xia H, Wang D, Tao X (2014) High yield seedless synthesis of high-
quality gold nanocrystals with various shapes. Langmuir 30(9):2480–2489. doi:10.1021/
la404602h
182. Venkatasubramanian R, He J, Johnson MW, Stern I, Kim DH, Pesika NS (2013) Additive-
mediated electrochemical synthesis of platelike copper crystals for methanol
electrooxidation. Langmuir 29(43):13135–13139. doi:10.1021/la4027078
183. Qi Y, Bian T, Choi S-I, Jiang Y, Jin C, Fu M, Zhang H, Yang D (2014) Kinetically controlled
synthesis of Pt-Cu alloy concave nanocubes with high-index facets for methanol electro-
oxidation. Chem Commun 50(5):560–562. doi:10.1039/c3cc48061e
184. Yin A-X, Min X-Q, Zhu W, Liu W-C, Zhang Y-W, Yan C-H (2012) Pt-Cu and Pt-Pd-Cu
concave nanocubes with high-index facets and superior electrocatalytic activity. Chem Eur J
18(3):777–782. doi:10.1002/chem.201102632
185. Choi S-I, Choi R, Han SW, Park JT (2011) Shape-controlled synthesis of Pt3Co nanocrystals
with high electrocatalytic activity toward oxygen reduction. Chem Eur J 17
(44):12280–12284. doi:10.1002/chem.201101138
186. Hu Y, Wu P, Zhang H, Cai C (2012) Synthesis of graphene-supported hollow Pt–Ni
nanocatalysts for highly active electrocatalysis toward the methanol oxidation reaction.
Electrochim Acta 85:314–321. doi:10.1016/j.electacta.2012.08.080
187. Hu Y, Wu P, Yin Y, Zhang H, Cai C (2012) Effects of structure, composition, and carbon
support properties on the electrocatalytic activity of Pt-Ni-graphene nanocatalysts for the
88 F.J. Vidal-Iglesias et al.

methanol oxidation. Appl Catal Environ 111–112:208–217. doi:10.1016/j.apcatb.2011.10.


001
188. Wang L, Nemoto Y, Yamauchi Y (2011) Direct synthesis of spatially-controlled Pt-on-Pd
bimetallic nanodendrites with superior electrocatalytic activity. J Am Chem Soc 133
(25):9674–9677. doi:10.1021/ja202655j
189. Wang L, Yamauchi Y (2011) Synthesis of mesoporous pt nanoparticles with uniform particle
size from aqueous surfactant solutions toward highly active electrocatalysts. Chem Eur J 17
(32):8810–8815. doi:10.1002/chem.201100386
190. Zhou X-W, Zhang R-H, Zhou Z-Y, Sun S-G (2011) Preparation of PtNi hollow nanospheres
for the electrocatalytic oxidation of methanol. J Power Sources 196(14):5844–5848. doi:10.
1016/j.jpowsour.2011.02.088
191. Luo X, Liu Y, Zhang H, Yang B (2012) Shape-selective synthesis and facet-dependent
electrocatalytic activity of CoPt3 nanocrystals. CrystEngComm 14(10):3359–3362. doi:10.
1039/c2ce25088h
192. Kang Y, Pyo JB, Ye X, Gordon TR, Murray CB (2012) Synthesis, shape control, and
methanol electro-oxidation properties of Pt-Zn alloy and Pt 3Zn intermetallic nanocrystals.
ACS Nano 6(6):5642–5647
193. Koper MTM, Lai SCS, Herrero E (2009) Mechanisms of the oxidation of carbon monoxide
and small organic molecules at metal electrodes. In: Koper MTM (ed) Fuel cell catalysis. A
surface science approach. Wiley, Hoboken, pp 159–208
194. Buso-Rogero C, Grozovski V, Vidal-Iglesias FJ, Solla-Gullon J, Herrero E, Feliu JM (2013)
Surface structure and anion effects in the oxidation of ethanol on platinum nanoparticles. J
Mater Chem A 1(24):7068–7076. doi:10.1039/c3ta10996h
195. Lee YW, Han SB, Kim DY, Park KW (2011) Monodispersed platinum nanocubes for
enhanced electrocatalytic properties in alcohol electrooxidation. Chem Commun 47
(22):6296–6298
196. Wei L, Fan YJ, Wang HH, Tian N, Zhou ZY, Sun SG (2012) Electrochemically shape-
controlled synthesis in deep eutectic solvents of Pt nanoflowers with enhanced activity for
ethanol oxidation. Electrochim Acta 76:468–474
197. Wei L, Fan YJ, Tian N, Zhou ZY, Zhao XQ, Mao BW, Sun SG (2012) Electrochemically
shape-controlled synthesis in deep eutectic solvents—a new route to prepare Pt nanocrystals
enclosed by high-index facets with high catalytic activity. J Phys Chem C 116(2):2040–2044
198. Wei L, Zhou ZY, Chen SP, Xu CD, Su D, Schuster ME, Sun SG (2013) Electrochemically
shape-controlled synthesis in deep eutectic solvents: triambic icosahedral platinum
nanocrystals with high-index facets and their enhanced catalytic activity. Chem Commun
49(95):11152–11154
199. Zhou ZY, Shang SJ, Tian N, Wu BH, Zheng NF, Xu BB, Chen C, Wang HH, Xiang DM, Sun
SG (2012) Shape transformation from Pt nanocubes to tetrahexahedra with size near 10 nm.
Electrochem Commun 22(1):61–64
200. Zhang L, Chen D, Jiang Z, Zhang J, Xie S, Kuang Q, Xie Z, Zheng L (2012) Facile syntheses
and enhanced electrocatalytic activities of Pt nanocrystals with {hkk} high-index surfaces.
Nano Res 5(3):181–189
201. Zhang J, Zhang L, Xie S, Kuang Q, Han X, Xie Z, Zheng L (2011) Synthesis of concave
palladium nanocubes with high-index surfaces and high electrocatalytic activities. Chem Eur
J 17(36):9915–9919
202. Yu N-F, Tian N, Zhou Z-Y, Huang L, Xiao J, Wen Y-H, Sun S-G (2014) Electrochemical
synthesis of tetrahexahedral rhodium nanocrystals with extraordinarily high surface energy
and high electrocatalytic activity. Angew Chem Int Ed 53(20):5097–5101. doi:10.1002/anie.
201310597
203. Chen X, Cai Z, Chen X, Oyama M (2014) Green synthesis of graphene-PtPd alloy
nanoparticles with high electrocatalytic performance for ethanol oxidation. J Mater Chem 2
(2):315–320. doi:10.1039/c3ta13155f
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 89

204. de Souza JPI, Queiroz SL, Bergamaski K, Gonzalez ER, Nart FC (2002) Electro-oxidation of
ethanol on Pt, Rh, and PtRh electrodes. A study using DEMS and in-situ FTIR techniques. J
Phys Chem B 106(38):9825–9830. doi:10.1021/jp014645c
205. Lima FHB, Gonzalez ER (2008) Ethanol electro-oxidation on carbon-supported Pt-Ru, Pt-Rh
and Pt-Ru-Rh nanoparticles. Electrochim Acta 53(6):2963–2971
206. Yuan Q, Zhou Z, Zhuang J, Wang X (2010) Seed displacement, epitaxial synthesis of Rh/Pt
bimetallic ultrathin nanowires for highly selective oxidizing ethanol to CO2. Chem Mater 22
(7):2395–2402
207. Rao L, Jiang Y-X, Zhang B-W, Cai Y-R, Sun S-G (2014) High activity of cubic PtRh alloys
supported on graphene towards ethanol electrooxidation. Phys Chem Chem Phys 16
(27):13662–13671. doi:10.1039/c3cp55059a
208. Hong JW, Lee YW, Kim M, Kang SW, Han SW (2011) One-pot synthesis and
electrocatalytic activity of octapodal Au-Pd nanoparticles. Chem Commun 47
(9):2553–2555. doi:10.1039/c0cc04856a
209. Wang ED, Xu JB, Zhao TS (2010) Density functional theory studies of the structure
sensitivity of ethanol oxidation on palladium surfaces. J Phys Chem C 114
(23):10489–10497. doi:10.1021/jp101244t
210. Zhang J, Hou C, Huang H, Zhang L, Jiang Z, Chen G, Jia Y, Kuang Q, Xie Z, Zheng L (2013)
Surfactant-concentration-dependent shape evolution of Au–Pd alloy nanocrystals from rhom-
bic dodecahedron to trisoctahedron and hexoctahedron. Small 9(4):538–544. doi:10.1002/
smll.201202013
211. Zhang G-R, Wu J, Xu B-Q (2012) Syntheses of sub-30 nm Au@Pd concave nanocubes and
Pt-on-(Au@Pd) trimetallic nanostructures as highly efficient catalysts for ethanol oxidation. J
Phys Chem C 116(39):20839–20847. doi:10.1021/jp304570c
212. Xing W, Yin G, Zhang J (2014) Rotating electrode methods and oxygen reduction
electrocatalysts. Elsevier Science, Amsterdam
213. Katsounaros I, Cherevko S, Zeradjanin AR, Mayrhofer KJJ (2014) Oxygen electrochemistry
as a cornerstone for sustainable energy conversion. Angew Chem Int Ed 53(1):102–121
214. Vidal-Iglesias F, Solla-Gullón J, Herrero E, Feliu J (2013) Au electrocatalysis for oxygen
reduction. In: Shao M (ed) Electrocatalysis in fuel cells, vol 9, Lecture notes in energy.
Springer, London, pp 483–512. doi:10.1007/978-1-4471-4911-8_16
215. Guo S, Zhang S, Sun S (2013) Tuning nanoparticle catalysis for the oxygen reduction
reaction. Angew Chem Int Ed 52(33):8526–8544
216. Lee J, Jeong B, Ocon JD (2013) Oxygen electrocatalysis in chemical energy conversion and
storage technologies. Curr Appl Phys 13(2):309–321
217. Gewirth AA, Thorum MS (2010) Electroreduction of dioxygen for fuel-cell applications:
materials and challenges. Inorg Chem 49(8):3557–3566. doi:10.1021/ic9022486
218. Wu J, Yang H (2013) Platinum-based oxygen reduction electrocatalysts. Acc Chem Res 46
(8):1848–1857. doi:10.1021/ar300359w
219. Gomez-Marin AM, Rizo R, Feliu JM (2014) Oxygen reduction reaction at Pt single crystals: a
critical overview. Catal Sci Technol 4(6):1685–1698. doi:10.1039/C3cy01049j
220. Gomez-Marin AM, Rizo R, Feliu JM (2013) Some reflections on the understanding of the
oxygen reduction reaction at Pt(111). Beilstein J Nanotechnol 4:956–967. doi:10.3762/
Bjnano.4.108
221. Adzic RR (1998) Recent advances in the kinetics of oxygen reduction. In: Lipkowski J, Ross
PN (eds) Electrocatalysis. Wiley-VCH, New York, pp 197–242
222. Markovic NM, Gasteiger HA, Ross PN (1995) Oxygen reduction on platinum low-index
single-crystal surfaces in sulfuric-acid-solution—rotating ring-Pt(Hkl) disk studies. J Phys
Chem 99(11):3411–3415
223. Kuzume A, Herrero E, Feliu JM (2007) Oxygen reduction on stepped platinum surfaces in
acidic media. J Electroanal Chem 599(2):333–343. doi:10.1016/j.jelechem.2006.05.006
224. Bandarenka AS, Hansen HA, Rossmeisl J, Stephens IEL (2014) Elucidating the activity of
stepped Pt single crystals for oxygen reduction. Phys Chem Chem Phys 16(27):13625–13629
90 F.J. Vidal-Iglesias et al.

225. Maciá MD, Campina JM, Herrero E, Feliu JM (2004) On the kinetics of oxygen reduction on
platinum stepped surfaces in acidic media. J Electroanal Chem 564(1-2):141–150. doi:10.
1016/j.jelechem.2003.09.035
226. Inaba M, Ando M, Hatanaka A, Nomoto A, Matsuzawa K, Tasaka A, Kinumoto T, Iriyama Y,
Ogumi Z (2006) Controlled growth and shape formation of platinum nanoparticles and their
electrochemical properties. Electrochim Acta 52(4):1632–1638
227. Sanchez-Sanchez CM, Solla-Gullon J, Vidal-Iglesias FJ, Aldaz A, Montiel V, Herrero E
(2010) Imaging structure sensitive catalysis on different shape-controlled platinum
nanoparticles. J Am Chem Soc 132(16):5622–5624. doi:10.1021/ja100922h
228. Yu T, Kim DY, Zhang H, Xia Y (2011) Platinum concave nanocubes with high-index facets
and their enhanced activity for oxygen reduction reaction. Angew Chem Int Ed 50
(12):2773–2777
229. Fu G, Wu K, Jiang X, Tao L, Chen Y, Lin J, Zhou Y, Wei S, Tang Y, Lu T, Xia X (2013)
Polyallylamine-directed green synthesis of platinum nanocubes. Shape and electronic effect
codependent enhanced electrocatalytic activity. Phys Chem Chem Phys 15(11):3793–3802
230. Zhou W, Wu J, Yang H (2013) Highly uniform platinum icosahedra made by hot injection-
assisted GRAILS method. Nano Lett 13(6):2870–2874
231. Wu J, Gross A, Yang H (2011) Shape and composition-controlled platinum alloy
nanocrystals using carbon monoxide as reducing agent. Nano Lett 11(2):798–802. doi:10.
1021/nl104094p
232. Kang Y, Ye X, Murray CB (2010) Size- and shape-selective synthesis of metal nanocrystals
and nanowires using CO as a reducing agent. Angew Chem Int Ed 49(35):6156–6159
233. Tripković V, Cerri I, Bligaard T, Rossmeisl J (2014) The influence of particle shape and size
on the activity of platinum nanoparticles for oxygen reduction reaction: a density functional
theory study. Catal Lett 144(3):380–388
234. Li D, Wang C, Strmcnik DS, Tripkovic DV, Sun X, Kang Y, Chi M, Snyder JD, van der
Vliet D, Tsai Y, Stamenkovic VR, Sun S, Markovic NM (2014) Functional links between Pt
single crystal morphology and nanoparticles with different size and shape: the oxygen
reduction reaction case. Energy Environ Sci 7:4061–4069. doi:10.1039/c4ee01564a
235. Kondo S, Nakamura M, Maki N, Hoshi N (2009) Active sites for the oxygen reduction
reaction on the low and high index planes of palladium. J Phys Chem C 113
(29):12625–12628
236. Hitotsuyanagi A, Kondo S, Nakamura M, Hoshi N (2011) Structural effects on the oxygen
reduction reaction on n(1 1 1)-(1 0 0) series of Pd. J Electroanal Chem 657(1–2):123–127
237. Xiao L, Zhuang L, Liu Y, Lu J, Abru~ na HD (2008) Activating Pd by morphology tailoring for
oxygen reduction. J Am Chem Soc 131(2):602–608. doi:10.1021/ja8063765
238. Shao M, Yu T, Odell JH, Jin M, Xia Y (2011) Structural dependence of oxygen reduction
reaction on palladium nanocrystals. Chem Commun 47(23):6566–6568
239. Grdeń M, Łukaszewski M, Jerkiewicz G, Czerwiński A (2008) Electrochemical behaviour of
palladium electrode: oxidation, electrodissolution and ionic adsorption. Electrochim Acta 53
(26):7583–7598
240. Lee C-L, Chiou H-P, Liu C-R (2012) Palladium nanocubes enclosed by (100) planes as
electrocatalyst for alkaline oxygen electroreduction. Int J Hydrogen Energy 37
(5):3993–3997. doi:10.1016/j.ijhydene.2011.11.118
241. Shao M (2013) Palladium-based electrocatalysts for oxygen reduction reaction. In: Shao M
(ed) Electrocatalysis in fuel cells, vol 9, Lecture notes in energy. Springer, London, pp
513–531. doi:10.1007/978-1-4471-4911-8_17
242. Zhang H, Jin M, Xiong Y, Lim B, Xia Y (2012) Shape-controlled synthesis of Pd
nanocrystals and their catalytic applications. Acc Chem Res 46(8):1783–1794. doi:10.
1021/ar300209w
243. Rodriguez P, Koper MTM (2014) Electrocatalysis on gold. Phys Chem Chem Phys 16
(27):13583–13594. doi:10.1039/c4cp00394b
2 Recent Advances in the Use of Shape-Controlled Metal Nanoparticles. . . 91

244. Mei D, He ZD, Zheng YL, Jiang DC, Chen Y-X (2014) Mechanistic and kinetic implications
on the ORR on a Au(100) electrode: pH, temperature and H-D kinetic isotope effects. Phys
Chem Chem Phys 16(27):13762–13773. doi:10.1039/c4cp00257a
245. Erikson H, Sarapuu A, Tammeveski K, Solla-Gullón J, Feliu JM (2014) Shape-dependent
electrocatalysis: oxygen reduction on carbon-supported gold nanoparticles. ChemElectroChem
1(8):1338–1347. doi:10.1002/celc.201402013
246. Gan L, Cui C, Rudi S, Strasser P (2014) Core-shell and nanoporous particle architectures and
their effect on the activity and stability of Pt ORR electrocatalysts. Top Catal 57
(1–4):236–244
247. Oezaslan M, Hasché F, Strasser P (2013) Pt-based core-shell catalyst architectures for oxygen
fuel cell electrodes. J Phys Chem Lett 4(19):3273–3291
248. Lim B, Jiang M, Camargo PHC, Cho EC, Tao J, Lu X, Zhu Y, Xia Y (2009) Pd-Pt bimetallic
nanodendrites with high activity for oxygen reduction. Science 324(5932):1302–1305
249. Stamenkovic VR, Fowler B, Mun BS, Wang GF, Ross PN, Lucas CA, Markovic NM (2007)
Improved oxygen reduction activity on Pt3Ni(111) via increased surface site availability.
Science 315(5811):493–497
250. Gasteiger HA, Markovic NM (2009) Just a dream or future reality? Science 324(5923):48–49
251. Zhang J, Yang H, Fang J, Zou S (2010) Synthesis and oxygen reduction activity of shape-
controlled Pt3Ni nanopolyhedra. Nano Lett 10(2):638–644
252. Zhang J, Fang J (2009) A general strategy for preparation of Pt 3d-transition metal (Co, Fe,
Ni) nanocubes. J Am Chem Soc 131(51):18543–18547
253. Wu J, Zhang J, Peng Z, Yang S, Wagner FT, Yang H (2010) Truncated octahedral Pt3Ni
oxygen reduction reaction electrocatalysts. J Am Chem Soc 132(14):4984–4985
254. Wu J, Yang H (2011) Synthesis and electrocatalytic oxygen reduction properties of truncated
octahedral Pt3Ni nanoparticles. Nano Res 4(1):72–82
255. Wu J, Qi L, You H, Gross A, Li J, Yang H (2012) Icosahedral platinum alloy nanocrystals
with enhanced electrocatalytic activities. J Am Chem Soc 134(29):11880–11883
256. Choi SI, Xie S, Shao M, Odell JH, Lu N, Peng HC, Protsailo L, Guerrero S, Park J, Xia X,
Wang J, Kim MJ, Xia Y (2013) Synthesis and characterization of 9 nm Pt-Ni octahedra with a
record high activity of 3.3 A/mgPt for the oxygen reduction reaction. Nano Lett 13
(7):3420–3425
257. Sang-Il Choi SX, Shao M, Lu N, Guerrero S, Odell JH, Park J, Wang J, Kim MJ, Xia Y (2014)
Controlling the size and composition of nanosized Pt–Ni octahedra to optimize their catalytic
activities toward the oxygen reduction reaction. Chemsuschem 7(5):1476–1483
258. Chen C, Kang Y, Huo Z, Zhu Z, Huang W, Xin HL, Snyder JD, Li D, Herron JA,
Mavrikakis M, Chi M, More KL, Li Y, Markovic NM, Somorjai GA, Yang P, Stamenkovic
VR (2014) Highly crystalline multimetallic nanoframes with three-dimensional
electrocatalytic surfaces. Science 343(6177):1339–1343. doi:10.1126/science.1249061
259. Sakamoto R, Omichi K, Furuta T, Ichikawa M (2014) Effect of high oxygen reduction reaction
activity of octahedral PtNi nanoparticle electrocatalysts on proton exchange membrane fuel cell
performance. J Power Sources 269:117–123. doi:10.1016/j.jpowsour.2014.07.011
260. Chou S-W, Lai Y-R, Yang YY, Tang C-Y, Hayashi M, Chen H-C, Chen H-L, Chou P-T
(2014) Uniform size and composition tuning of PtNi octahedra for systematic studies of
oxygen reduction reactions. J Catal 309:343–350. doi:10.1016/j.jcat.2013.09.008
261. Xu X, Zhang X, Sun H, Yang Y, Dai X, Gao J, Li X, Zhang P, Wang H-H, Yu N-F, Sun S-G
(2014) Synthesis of Pt–Ni alloy nanocrystals with high-index facets and enhanced
electrocatalytic properties. Angew Chem Int Ed 126:12730–12735. doi:10.1002/anie.
201406497
262. Huang X, Zhao Z, Chen Y, Zhu E, Li M, Duan X, Huang Y (2014) A rational design of
carbon-supported dispersive Pt-based octahedra as efficient oxygen reduction reaction cata-
lysts. Energy Environ Sci 7(9):2957–2962. doi:10.1039/c4ee01082e
263. Wu J, Yang H (2012) Study of the durability of faceted Pt3Ni oxygen–reduction
electrocatalysts. ChemCatChem 4(10):1572–1577. doi:10.1002/cctc.201200242
92 F.J. Vidal-Iglesias et al.

264. Cui C, Gan L, Heggen M, Rudi S, Strasser P (2013) Compositional segregation in shaped
Pt alloy nanoparticles and their structural behaviour during electrocatalysis. Nat Mater
12(8):765–771. doi:10.1038/nmat3668
265. Ahmadi M, Behafarid F, Cui C, Strasser P, Cuenya BR (2013) Long-range segregation
phenomena in shape-selected bimetallic nanoparticles: chemical state effects. ACS Nano
7(10):9195–9204. doi:10.1021/nn403793a
266. Tuaev X, Rudi S, Petkov V, Hoell A, Strasser P (2013) In situ study of atomic structure
transformations of Pt–Ni nanoparticle catalysts during electrochemical potential cycling.
ACS Nano 7(7):5666–5674. doi:10.1021/nn402406k
267. Cui C, Ahmadi M, Behafarid F, Gan L, Neumann M, Heggen M, Cuenya BR, Strasser P
(2013) Shape-selected bimetallic nanoparticle electrocatalysts: evolution of their atomic-
scale structure, chemical composition, and electrochemical reactivity under various chemical
environments. Faraday Discuss 162:91–112. doi:10.1039/c3fd20159g
268. Wu J, Shi M, Yin X, Yang H (2013) Enhanced stability of (111)-surface-dominant core–shell
nanoparticle catalysts towards the oxygen reduction reaction. Chemsuschem 6
(10):1888–1892. doi:10.1002/cssc.201300388
269. Choi S-I, Shao M, Lu N, Ruditskiy A, Peng H-C, Park J, Guerrero S, Wang J, Kim MJ, Xia Y
(2014) Synthesis and characterization of Pd@Pt–Ni core–shell octahedra with high activity
toward oxygen reduction. ACS Nano 8(10):10363–10371. doi:10.1021/nn5036894
270. Xie S, Choi S-I, Lu N, Roling LT, Herron JA, Zhang L, Park J, Wang J, Kim MJ, Xie Z,
Mavrikakis M, Xia Y (2014) Atomic layer-by-layer deposition of Pt on Pd nanocubes for
catalysts with enhanced activity and durability toward oxygen reduction. Nano Lett 14
(6):3570–3576. doi:10.1021/nl501205j
271. Shao M, He G, Peles A, Odell JH, Zeng J, Su D, Tao J, Yu T, Zhu Y, Xia Y (2013)
Manipulating the oxygen reduction activity of platinum shells with shape-controlled palla-
dium nanocrystal cores. Chem Commun 49(79):9030–9032. doi:10.1039/c3cc43276a
272. Li Y, Quan F, Chen L, Zhang W, Yu H, Chen C (2014) Synthesis of Fe-doped octahedral
Pt3Ni nanocrystals with high electro-catalytic activity and stability towards oxygen reduction
reaction. RSC Adv 4(4):1895–1899. doi:10.1039/c3ra46299d
Chapter 3
Pt-Containing Heterogeneous Nanomaterials
for Methanol Oxidation and Oxygen
Reduction Reactions

Hui Liu, Feng Ye, and Jun Yang

3.1 Introduction

The strong growing interest in using direct methanol fuel cells (DMFCs) as portable
and mobile power sources is rooted in their desirable features of relatively small
environmental footprint, compact system design, and higher volumetric energy
densities compared with existing technologies [1, 2]. DMFC relies on the oxidation
of methanol on a catalyst layer to form carbon dioxide (CO2). Water (H2O) is
consumed at the anode and is produced at the cathode. Hydrogen ions are
transported across the proton exchange membrane to the cathode where they
react with oxygen (O2) to produce water. Electrons are transported through an
external circuit from anode to cathode, providing power to external devices. The
overall reaction occurred in a DMFC can be described as
2CH3OH þ 3O2 ! 4H2O þ 2CO2.
It has been well recognized that the success of a DMFC technology depends
largely on the electrocatalysts. Currently, platinum (Pt)-based nanomaterials are the
most effective electrocatalysts to facilitate both the anodic (methanol oxidation
reaction, MOR) and cathodic (oxygen reduction reduction, ORR) reactions in
DMFCs [3–6]. However, they are susceptible to CO poisoning in the MOR and
their ORR activity is still limiting the fuel cell performance [7–11]. Other than the
classical approaches of increasing the Pt catalytic performance through particle size
reduction, particle shape control, and alloying with oxophilic metals [5, 6, 12–14],
there is increasing interest in combining structural engineering and the synergistic
effects of an adjuvant metal to enhance the Pt catalytic properties [15–17].

H. Liu • F. Ye • J. Yang (*)


State Key Laboratory of Multiphase Complex Systems, Institute of Process Engineering,
Chinese Academy of Sciences, Beijing 100190, China
e-mail: [email protected]

© Springer International Publishing Switzerland 2016 93


K.I. Ozoemena, S. Chen (eds.), Nanomaterials for Fuel Cell Catalysis,
Nanostructure Science and Technology, DOI 10.1007/978-3-319-29930-3_3
94 H. Liu et al.

Although a wide variety of recipes have been developed to synthesize, assemble


and package nanomaterials/nanostructures into forms more amenable to applica-
tions, controllable synthesis of nanomaterials with heterogeneous structure,
e.g. core-shell, hollow interiors, cage-bell structures, stellated/dendritic
morpholoies, dimeric, or composite construction, is much more complicated
[14, 18–24]. The synthesis of heterogeneous nanomaterials can be carried out in
the solid, gaseous, or solution state. For example, multi-metallic compounds can be
traditionally obtained using metallurgical techniques which melts two or more
kinds of bulk metals under proper conditions [25]. This solid state method requires
high-temperature heating and annealing for long periods of time, and it is difficult
to obtain nanocrystalline compounds with high surface areas or desired structures
which are urgently needed in various applied fields such as energy, environment,
and catalysis [26]. In contrast, the solution-based method is more powerful and
versatile and has been the preferred method by many researchers lately toward the
synthesis of heterogeneous nanomaterials [14, 15, 24]. In a solution-based synthetic
system, the nucleation and growth process of the nanomaterials can be easily
controlled by adjusting the reaction parameters including the concentration of
reactants, the mole ratio between precursors and surfactants, and the reaction
temperature and time. The synthesis, property, and application of Pt-based nano-
structured materials have been recently reviewed by Chen and Holt-Hindle
[14]. Therefore, in the following, after a brief update of the literature, we prefer
to devote this chapter to summarize the Pt-containing nanomaterials with hetero-
geneous structures prepared via solution-based approaches, their characterization
and potential applications as electrocatalysts for MOR and ORR so as to provide the
readers a systematic and coherent picture of the field. Most of these works have
only been carried out in the last several years, particularly by the authors in different
laboratories.

3.2 Pt-Based Heterogeneous Nanomaterials


with Core-Shell Constructions

Bi- and multi-metallic nanoparticles with a core–shell construction are a type of


conventional heterogeneous nanomaterials. However, they still have attracted tre-
mendous research interest in past few years [27–36]. The complex electron inter-
action between/among the two or more electron-rich elements, the lattice strain
created in these core–shell particles, as well as the heterometallic bonding interac-
tions, modify the surface electronic properties of the nanoparticles [37]. Therefore,
core–shell nanoparticles often exhibit improved catalytic properties compared to
their alloyed counterparts or to mixtures of monometallic nanoparticles [38, 39].
The seed-mediated growth reaction based on the successive reduction of two or
more metal precursor salts is one of the commonest methods to prepare core-shell
structured nanoparticles. Turkevich and Kim in the 1970s had tried to obtain
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 95

Fig. 3.1 (a) HAADF-STEM, (b) high-resolution HAADF-STEM, and (c) elemental mapping
images of the 5 nm/1 nm Pd/FePt NPs. The experiments were carried out on an aberration-
corrected JEOL 2200FS microscope. HAADF-STEM images were acquired with a convergence
angle of 27 mrad and an inner collection angle of 100 mrad. EDS analysis was carried out with an
electron beam size of 2 Å. Reproduced with permission from [27] Copyright American Chem-
ical Society

gold-layered Pd nanoparticles by this method [40]. The deposition of one metal on


preformed nanoparticles of another metal is convenient, provided that the deposi-
tion can be carried out selectively on the core particles.
Among the latest investigations into the synthesis and electrocatalytic properties
of Pt-containing multi-metallic core-shell nanoparticles, the work from Sun group
at Brown University is especially relevant. As a typical example, in 2010, Sun and
co-workers reported a unique approach for synthesizing core-shellstructured
Pd-FePt NPs with a 5 nm Pd core and a FePt shell whose thickness is tunable
from 1 to 3 nm [27]. They first prepared 5 nm Pd seed particles by reducing the Pd
(acac)2 in oleylamine and borane t-butylamine at 75  C, and then they coated the Pd
seeds with FePt shell in a mixture consisting of oleylamine, oleic acid and
1-octadecene. the core-shell structure of the Pd-FePt nanoparticles was confirmed
by transmission electron microscopy (TEM) and aberration-corrected high-angle
annular dark-field scanning TEM (HAADF-STEM), as shown in Fig. 3.1. The
authors demonstrated that the ORR catalysis of the core-shell Pd-FePt nanoparticles
is dependent on the thickness of the FePt shell and that the thin (1 nm or less) FePt
shell is both active and durable for ORR in 0.1 M HClO4 solution. The authors
proposed that the much-enhanced ORR stability of the 5–1 nm Pd-FePt
nanoparticles may arise from the thin FePt coating, as a thin metallic shell can
offer the desired thermodynamic stability for a metallic core/shell structure [41].
Subsequently, the Sun group replaced the Pd core with Au, and prepared
multimetallic Au-FePt3 nanoparticles with core-shell construction [42]. They
found that the multimetallic core-shell Au-FePt3 nanoparticles possess both the
96 H. Liu et al.

high catalytic activity of Pt-bimetallic alloys and the superior durability of the
tailored morphology and composition profile, with mass-activity enhancement of
more than one order of magnitude over Pt catalysts. Experimental observations
suggest that the Au core could enhance the ORR catalytic performance of FePt3
alloy shell through altering the surface segregation and hindering place-exchange
mechanisms. The obtained fundamental insight into the synergy among these
materials could enable the researchers to resolve, define, and utilize the exact role
of each constituent in a ternary system, which has guided the synthesis of tailored
nanoparticles possessing favorable coordination of surface active sites, distribution
of elements and amount of Pt in the system.
Pt nanoparticles with controlled morphologies, such as tetrahedral [43], cubic
[12, 44, 45], truncated octahedral [46] and high-index tetra hexahedrons [47], have
been demonstrated as highly active catalysts for various reactions. In particular,
icosahedral nanoparticles with a high density of compressed twins on the surfaces
are expected to be the most active catalysts [48]. The excellent activity of the
multiply twinned particles (MTPs) of Pt with (1 1 1) facet of icosahedral morphol-
ogy with a high density of twins and corners on their surface was also verified
computationally [49]. However, unlike palladium (Pd) [48], silver (Ag) [50] and
gold (Au) [51], MTPs of Pt were rarely formed [24].
In recent, Yang and co-workers reported for the first time the synthesis of
monodispersed core-shell AgPd-Pt MTPs with icosahedral morphology, and inves-
tigated their application in catalyzing ORR at room and moderate temperatures
[52]. In this method, oleylamine was used as solvent, reducing agent and stabilizer
for the formation of nanoparticles. The Ag component of the AgPd alloy inner core
was crucial for the construction of the multiply twinned structure of the core-shell
nanoparticles, while the Pd component was employed to reduce the tensile strain
effect of the Ag on the deposited Pt layers. The core-shell AgPd-Pt MTPs exhibited
superior catalytic activity toward ORR, as compared to the commercial Pt
nanoparticles. The enhancement was attributed to the higher surface fraction of
atoms on the (1 1 1) facets of icosahedral Pt MTPs.
In the first step, multiply twinned Ag nanoparticles with an average diameter of
ca. 7 nm and an icosahedral morphology were prepared in oleylamine at elevated
temperature. These Ag MTPs were then used as seeds for the formation of core-
shell nanoparticles. To prepare core-shell AgPd-Pt MTPs, the PdCl2 and Pt(acac)2
metal precursors were immediately added to Ag colloidal solution in oleylamine,
which was kept at 250  C for 1 h under a nitrogen flow and magnetic stirring. Upon
the addition of Pd(II) and Pt(II) salts, the replacement reaction between Ag
nanoparticles and Pd(II) ions occurred immediately to form AgPd alloy
nanoparticles, analogous to the formation of AgAu or AgPd alloy nanoparticles
by replacement reaction between Ag nanoparticles and Au or Pd metal ions in
aqueous solution [53]. The alloying process was realized by the rapid interdiffusion
of metal atoms as a result of the reduced dimension of the silver templates, the
elevated temperature, and the large number of interfacial vacancy defects generated
by the replacement reaction. The subsequent reduction of Pt(II) precursors resulted
in the homogeneous deposition of Pt on the preformed AgPd alloy nanoparticles.
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 97

Fig. 3.2 (a) TEM and (b) HRTEM images of icosahedral core-shell AgPd@Pt nanoparticles.
HRTEM images of AgPd@Pt MTP along a (c) two-fold, (d) three-fold, and (e) five-fold symmetry
axis. (f) HAADF-STEM image. (g) EDX spectrum of a single AgPd@Pt MTP boxed in (f). (h) Ag,
Pd and Pt elemental profiles along the red line across the MTP (inset of h). Reproduced with
permission from [52] Copyright American Chemical Society

The as-prepared colloidal solution was then cooled down to room temperature, and
the nanoparticles were precipitated, washed with methanol, and re-dispersed in a
non-polar organic solvent (e.g. toluene, hexane or chloroform). A typical TEM
image of the core-shell AgPd-Pt nanoparticles was shown in Fig. 3.2a. The
nanoparticles are nearly monodispersed, with an average diameter of ca. 15 nm.
The narrow size distribution of nanoparticles results in the formation of
98 H. Liu et al.

two-dimensional close-packed hexagonal arrays on the TEM grid. The core-shell


AgPd-Pt nanoparticles are predominantly multiply twinned with an icosahedral
morphology, as illustrated by the high-resolution TEM (HRTEM) image
(Fig. 3.2b). They were found in different orientations (along a two-fold symmetry
axis in Fig. 3.2c, along a three-fold symmetry axis in Fig. 3.2d, and along a five-fold
symmetry axis in Fig. 3.2e) [51, 54, 55]. The energy-dispersive X-ray (EDX)
spectroscopy analysis (Fig. 3.2g) of an arbitrary single particle boxed in the
HAADF-STEM image of Fig. 3.2f confirms that the particle is indeed composed
of Ag, Pd and Pt. Since the lattice mismatch between the AgPd alloy core and the Pt
shell was only ca. 1.1 % [52], the core-shell nanoparticles were difficult to deter-
mine accurately using standard TEM. However, the formation of core-shell AgPd-
Pt nanoparticles could be confirmed by the elemental profiles of these particles. As
shown in Fig. 3.2h, the Pt signal is noted across the entire particle (ca. 15 nm),
whereas the Ag and Pd signals is obtained only across the core (ca. 10 nm).
The addition of Pd(II) or Pt(II) salt precursor alone to icosahedral Ag seeds
would result in AgPd alloy nanoparticles of ca. 10 nm or core-shell Ag-Pt
nanoparticles of ca. 12 nm, respectively, which are also predominantly multiply
twinned with an icosahedral morphology [52]. The XRD patterns of Ag, AgPd
alloy, core-shell Ag-Pt, core-shell AgPd-Pt, and pure Pt nanoparticles prepared in
oleylamine in the absence of any seeds were presented in Fig. 3.3. As compared to
pure Pt nanoparticles, a shift to a lower angle is observed for the (1 1 1) diffraction
peak of core-shell Ag-Pt (Fig. 3.3c). This is because Ag has larger lattice parameter
(0.4090 nm) than that of Pt (0.3923 nm), and the Ag cores exert a tensile effect on
the Pt atoms deposited on their surface. In contrast, the (1 1 1) diffraction peak of
core-shell AgPd-Pt is very close to that of pure Pt, indicating that the presence of Pd
in alloy AgPd cores could reduce the tensile effect induced by Ag on the Pt layer.
Based on the calculation from Fig. 3.3b, the lattice parameter for the (1 1 1) facet of
AgPd alloy was 0.3967 nm, very similar to that of pure Pt.

Fig. 3.3 XRD patterns of


(a) Ag, (b) alloy Ag/Pd, (c)
core-shell Ag-Pt, (d) core-
shell Ag/Pd-Pt, and (e) Pt
nanoparticles. Reproduced
with permission from [52]
Copyright American
Chemical Society
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 99

Fig. 3.4 (a) TEM and (b) HRTEM images of Ag/Pd-Pt/C catalysts. (c) HRTEM image of a single
Ag/Pd-Pt icosahedral nanoparticle supported on XC-72 carbon. Reproduced with permission from
[52] Copyright American Chemical Society

For the electrochemical characterizations, the core-shell AgPd-Pt nanoparticles


were loaded on Vulcan XC-72 carbon supports. TEM and HRTEM images of the
AgPd-Pt/C nanoparticles show that the core-shell AgPd-Pt nanoparticles are well
dispersed on XC-72, maintaining their particle size and the multiply twinned
structure with icosahedral morphology (Fig. 3.4).
The catalytic activity of the core-shell AgPd-Pt MTPs with icosahedral mor-
phology towards ORR was evaluated by sweep voltammetry in O2-saturated
HClO4, and benchmarked against that of the commercial Pt/C catalysts with an
average diameter of ca. 3 nm. Cyclic voltammograms of commercial Pt/C and core-
shell AgPd-Pt/C in argon-purged 0.1 M HClO4 at room temperature were used to
obtain the electrochemically active surface areas (ECSAs) from the hydrogen
adsorption/desorption regions (0.0–0.3 V vs. RHE in Fig. 3.5a). In the cathodic
scan, the oxide (OHads) stripping peak (0.8 V) for core-shell AgPd-Pt/C is higher
than that for commercial Pt/C (0.78 V). The positive shift of ca. 0.02 V in the oxide
stripping peak for core-shell AgPd-Pt/C suggests a weaker binding of the OHads
species on the surface of this catalyst [56]. This could be attributed to a weaker
interaction with the adsorbed species on the dominant (1 1 1) facets [57]. Figure 3.5b
shows the ORR polarization curves of these two catalysts in oxygen-saturated
0.1 M HClO4 at room temperature and 60  C. As indicated, at room temperature,
the core-shell AgPd-Pt/C MTPs exhibit a more positive half-wave potential
100 H. Liu et al.

0
1.0 Pt/C (A) (B)

Current Density (mA cm-2)


Current Density (mA cm-2)

AgPd@Pt/C -1
0.5 Pt/C at RT
-2 Pt/C at 60°C
0.0 AgPd@Pt/C at RT
-3
AgPd@Pt/C at 60°C
-0.5
-4

-1.0 -5

-1.5 -6
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Potential (V vs. RHE) Potential (V vs. RHE)
0.99 1.0
(C) (D)

ik Mass Activity (mA mg-1 Pt)


Pt/C at RT
Potential (V vs. RHE )

0.96 0.8
Pt/C at 60°C
AgPd@Pt/C at RT
0.93 0.6
AgPd@Pt/C at 60°C

Pt/C at RT 0.4
0.90
Pt/C at 60°C
AgPd@Pt/C at RT 0.2
0.87
AgPd@Pt/C at 60°C
0.0
0.01 0.1 1 0.90 0.85
Specific Activity (mA cm-2 Pt) Potential (V vs. RHE)

Fig. 3.5 (a) Cyclic voltammograms of Pt/C and Ag/Pd-Pt/C in argon-purged 0.1 M HClO4; sweep
rate ¼ 20 mV/s, room temperature. (b) Linear sweep voltammograms in oxygen-saturated 0.1 M
HClO4 showing the positive-going scans; sweep rate ¼ 20 mV/s, room temperature or 60  C,
1600 rpm. (c) Tafel plot for ORR at high potential, normalized by Pt ECSA. (d) Kinetic mass
activities of ORR at 0.85 and 0.90 V RHE, normalized by Pt mass. Reproduced with permission
from [52] Copyright American Chemical Society
(0.88 V) than that of commercial Pt/C catalysts (0.86 V), indicating that they have a
higher ORR activity.
By normalizing with the Pt surface area, the specific activities of core-shell
AgPd-Pt/C and commercial Pt/C were obtained (Fig. 3.5c). At room temperature
and 0.9 V vs. RHE, a specific activity of 0.36 mA cm2 Pt was attained for core-
shell AgPd-Pt/C, which is >2 times higher than that for commercial Pt/C
(0.16 mA cm2 Pt). Kinetic mass activity calculated from Koutecký-Levı̀ch equa-
tion (Fig. 3.5d) also confirms that the core-shell AgPd-Pt/C (0.23 mA μg1 Pt) is
superior to commercial Pt/C (0.11 mA μg1 Pt). In addition, regardless of whether
Pt surface area or Pt mass is used as the basis for normalizing the measured current,
the specific activity (0.68 mA cm2 Pt) and mass activity (0.44 mA μg1 Pt) of
core-shell AgPd-Pt/C at 60  C and 0.90 V are higher than those of commercial Pt/C
(0.24 mA cm2 Pt and 0.17 mA μg1 Pt, respectively). The former values are also
close to the 2015 targets set by the U.S. Department of Energy (DoE) at 80  C and
0.9 V (0.72 mA cm2 Pt and 0.44 mA μg1 Pt, respectively). Even on the basis of
total mass of Pd and Pt, the core-shell AgPd-Pt/C catalyst shows greater mass
activities at room temperature and 60  C (0.15 mA μg1 Pt þ Pd and 0.28 mA μg1
Pt þ Pd, respectively) than those of the commercial Pt catalyst (0.11 mA μg1 Pt,
and 0.17 mA μg1 Pt, respectively).
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 101

It should be noted that the catalytic activities of AgPd alloy and core-shell Ag-Pt
nanoparticles towards ORR were found to be much lower than that of the core-shell
AgPd-Pt nanoparticles. It is well known that most Pd-based catalysts are lower in
intrinsic ORR activities than Pt [58, 59]. Therefore, it was not unexpected that the
AgPd alloy nanoparticles show a lower activity than that of core-shell Ag-Pt and
AgPd-Pt nanoparticles. For the core-shell Ag-Pt nanoparticles, the tensile effect of
Ag core on the Pt shell would result in an upward shift of the d-band center of Pt
[60]. As has been demonstrated [61], the surface of a Pt-based catalyst with a high
value in d-band center would bind adsorbents too strongly. Compared to that of the
core-shell Ag-Pt nanoparticles, the superior catalytic activity of core-shell AgPd-Pt
nanoparticles toward ORR is possibly because the latter has a more suitable d-band
center to balance the breaking of O–O bonds and the formation of O–H bonds, two
usual reactions in a common ORR process [9, 62, 63].
The synthesis of Pt-containing heterogeneous nanoparticles with core-shell
structure as efficient electrocatalysts is still a very hot topic in current scientific
communities. In 2012, Chen and co-workers prepared core-shell Ru-Pt
nanoparticles through a polyol reduction with a sequence controlling processes
[32]. The core-shell particles have a Pt shell with a precisely controlled thickness
around 1.5 atomic layers. The authors found that the activity of the core–shell Ru-Pt
nanoparticles in methanol oxidation reaction (MOR) is significantly dependent on
the crystalline structure of the core and the lattice strain at the core–shell interface.
Core-shell Ru-Pt nanoparticles delivered 6.1-fold peak MOR current density at
135 mV than Pt particles. The current density is improved by the compressive
lattice strain of the surface that is caused by the lattice mismatch between the Pt
shell and the Ru core. In addition, the Ru-core nanoparticles maintained a steady
current density of 0.11 mA cm2 at 500 mV in a half-cell system for 2 h, which is
100-fold higher than that of Pt particles. These results provide mechanistic infor-
mation for the development of fuel cell catalysts along with reduced Pt utilization
and programmable electrochemical performance.
Also in 2012, the Adzic group described a route to the development of novel
PtNiN core–shell catalysts with low Pt content shell and inexpensive NiN core,
which have high activity and stability for the ORR [33]. The PtNiN synthesis
involves nitriding Ni nanoparticles and simultaneously encapsulating them by
2–4 monolayer-thick Pt shell. The experimental data and the density functional
theory calculations indicate nitride has the bifunctional effect that facilitates for-
mation of the core––shell structures and improves the performance of the Pt shell
by inducing both geometric and electronic effects‘. More importantly, the synthesis
of inexpensive NiN cores opens up possibilities for designing of various transition
metal nitride based core–shell nanoparticles for a wide range of applications in
energy conversion processes.
To improve the catalytic activity of Pt-based nanomaterials for MOR, the Wang
and Yamauchi group prepared core-shell Au-Pt nanocolloids with nanostructured
dendritic Pt shells through chemically reducing both H2PtCl6 and HAuCl4 species
in the presence of a low-concentration surfactant (Pluronic F127) solution [64]. By
applying an ultrasonic treatment, the particle size of the core-shell Au-Pt
102 H. Liu et al.

Fig. 3.6 (a) Low-magnification and (b) high-magnification TEM images of Au@Pt nanocolloids
prepared from a typical solution with a Pt/Au molar ratio of 1.0. (c) High-magnification TEM
image of the Pt shell. (d) Selected-area electron diffraction (SAED) patterns taken from the Pt shell
region. Reproduced with permission from [64] Copyright American Chemical Society
nanocolloids is dramatically decreased and their size distribution becomes very
narrow. The difference in reduction potentials of the two soluble metal salts (Au
(III) and Pt(IV) species) plays a key role in the one-step synthesis of the core-shell
structure. Because of the different reduction potentials, the reduction of Au ions
preferentially occurs over a short time to form the Au seeds. It is followed by
overgrowth of Pt nanodendritic nanowires on the Au seeds, as confirmed by the
TEM images of the final products (Fig. 3.6). The thickness of the Pt shell on the Au
cores can be easily tuned by controlling the Pt/Au molar ratios in the starting
precursor solutions. Through the optimization of the Pt shell thicknesses, the
core-shell Au-Pt nanocolloids exhibit enhanced activity as an electrocatalyst for
MOR.
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 103

Subsequently, Kang et al. demonstrated a facile one-pot aqueous approach for


the controllable synthesis of trimetallic Au-PdPt core-shell nanoparticles with a
well-defined octahedral Au core and a highly crystalline dendritic PdPt alloy shell
[35]. The simultaneous reduction of multiple metal precursors with dual reducing
agents, namely, ascorbic acid and hydrazine, gave a fine control over the nucleation
and growth kinetics of nanoparticles, resulting in the formation of novel core-shell
Au-PdPt nanoparticles. The prepared core-shell particles display excellent catalytic
performance for methanol electrooxidation, which can be attributed to their opti-
mized binding strength toward adsorbate molecules due to the improved charge
transfer between core and shell of the particles.
Most recently, the Sun group reported a core-shell nanoparticle with dual
electrocatalytic activities for both methanol oxidation and oxygen reduction reac-
tions [36]. They firstly prepared Au seed particles of 5 nm in diameter in a mixture
of tetralin and oleylamine, and then coated these Au seeds with alloy CuPt shell in
1-octadecene at elevated temperature. The thickness of the alloy CuPt shell is
ca. 1.5 nm, and the final product was re-dispersed in hexane. The CuPt composi-
tions (Pt65Cu35, Pt50Cu50, and Pt36Cu64) in the alloy shell could be controlled by the
molar ratio of Cu and Pt metal precursors. The CuPt alloy effect and interactions
between core and shell make these core-shell Au-CuPt nanoparticles a promising
catalyst for both ORR and MOR in 0.1 M HClO4 solution. As shown in Fig. 3.7, the
specific (mass) reduction and oxidation activities of these core-shell particles reach

a b 10
Au/Cu35Pt65
Specific Activity (mA / cm2)

0.30
Before 5000 cycles Au/Cu50Pt50
After 5000 cycles Au/Cu64Pt36
Current (mA)

0.15
Pt
0.00 1

-0.15

-0.30
0.1
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.86 0.88 0.90 0.92 0.94 0.96
Potential V (vs. RHE) Potential V (vs. RHE)
c d
Mass Activity (mAmg-1pt)

2500
Specific Activity (mA / cm )
2

2000
3.0
2.5 1500
2.0
1000
1.5
1.0 500
0.5
0.0 0
Au/Cu35Pt65 Au/Cu50Pt50 Au/Cu64Pt36 Pt

Fig. 3.7 (a) CVs of core-shell Au-Cu35Pt65/C catalysts before and after 5000 potential cycles
between 0.6 and 1.0 V; (b) activity–potential plots of Au-CuxPt100x/C and Pt/C catalysts; (c)
specific and Pt mass activities of Au-CuxPt100x/C and Pt/C catalysts at 0.9 V. (d) TEM of
Au-Cu35Pt65/C catalysts before (left) and after (right) 5000 potential cycles. Reproduced with
permission from [36] Copyright American Chemical Society
104 H. Liu et al.

2.72 mA cm2 (1500 mA mg1 Pt) at 0.9 V and 0.755 mA cm2 (441 mA mg1 Pt)
at 0.8 V (vs RHE), respectively. Their studies show that the existence of an Au core
not only minimizes the Pt usage but also improves the stability of the core-shell
Au-CuPt catalyst for fuel cell reactions. The results suggest that the core-shell
design is indeed effective for optimizing nanoparticle catalysis. The same concept
may be extended to other multimetallic nanoparticle systems, making it possible to
tune nanoparticle catalysis for many different chemical reactions.
Although great successes have been achieved in preparation and electrochemical
application of core-shell structured nanomaterials, to date, as summarized by Yang
in a mini-review [28], the reports on the synthesis of well-defined multi-metallic
core-shell nanoparticles with sizes below ca. 10 nm are still relatively uncommon.
Besides the intrinsic challenges in the synthesis, one important reason perhaps lies
in the difficulty to characterize the core-shell structure in multi-metallic
nanoparticles in detail. In addition, in view of the profoundly effect of morphology
of nanoparticles on the catalytic activity, the performance of core-shell structured
nanomaterials for MOR and ORR would be dramatically increased if their overall
shape could also be finely controlled at this length scale.

3.3 Pt-Based Heterogeneous Nanomaterials with Hollow


Interiors or Cage-Bell Structures

Preparation of Pt-based nanomaterials with hollow interiors is the most common


practice to enhance their electrocatalytic property for MOR and ORR [65–72]. For
instance, Pt hollow nanospheres are found to be twice as active as solid Pt
nanoparticles of roughly the same size for methanol oxidation [65, 66]. The
increase in activity could be attributed mainly to the larger surface area of the
hollow structure, where the porous shell allows the internal surface of the catalyst to
be accessible to the reactants. The continuing research efforts in this area have also
given rise to the recent possibility of creating hybrids of core-shell and hollow
structures, or a new class of core-shell structures with a distinctive core-void-shell
configuration, which are called cage-bell, yolk-shell or rattle-type structures. With
the unique properties of a movable core, interior void spaces, and controlled
porosity and composition of the shell, cage-bell structured (CBS) nanomaterials
have great potential for a diverse range of applications, such as nanoreactors
[73–75], drug delivery systems [76–78], lithium-ion batteries [79–83],
photocatalysis [84], and photonics [85]. For example, polymeric hollow spheres
with a movable Au nanoparticle core have been synthesized, which allow the
optical sensing of chemicals diffused into the cavity [86]. Recently, Fan and
co-workers developed a photocatalytic approach using densely packed optically
active porpyrins to template the synthesis of well-defined hollow Pt nanostructures
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 105

which were excellent catalyst for the methanol oxidation reaction [87]. The authors
also demonstrated reusability of the template by an acid-base treatment. A general
strategy for the fabrication of hollow or CBS structures is template-assisted selec-
tive etching of core-shell particles. The core particle in this case is overlaid with a
single or double shell of a different material. The core or the inner shell is then
selectively removed by calcination or with a solvent.
Over the past 20 years, various methods based on sacrificial templates of
polymer and inorganic spheres [72, 88–93], liquid droplets [94], vesicles [95–97],
and microemulsion droplets [98–100], have been developed for the synthesis of
hollow and CBS nanomaterials, and were reviewed recently by the Lou and Jiang
groups [101, 102]. These preparative methods are however system specific, more
successful for metal oxides than for noble metals. Specific methods based on other
principles such as galvanic replacement [65, 68, 103–110], Kirkendall effect [111],
Ostwald ripening [112, 113], layer-by-layer assembly [91, 114], copper sulfide
leaching [115], and structure transformation [116], have also been used for the
synthesis of hollow or cage-bell nanomaterials. These methods are only applicable
to a small number of metals (mostly Au and Pt) and there are significant limitations.
For example, the core and the shell of cage-bell nanostructures prepared this way
are usually made of the same material, and there is no effective control of the shell
thickness and shell structure [117, 118]. In summary, a general approach to
rationally fabricate hollow and cage-bell structures for a sufficiently wide spectrum
of noble metals is still lacking and it poses significant challenges.
In this section, we focus on a facile generic approach reported by the Yang group
in 2012, which is based on the inside-out diffusion of Ag in core-shell structures, for
the fabrication of hollow or cage-bell noble metal nanomaterials [119]. In this
strategy, core-shell Ag-M nanoparticles or core-shell-shell MA-Ag-MB
nanoparticles are first prepared in an organic solvent, followed by the removal of
Ag from the core or the inner shell with bis(p-sulfonatophenyl)phenylphosphane
dihydrate (BSPP), which binds strongly with Ag(I)/Ag(0) to promote the inside-out
diffusion of Ag in the core-shell structure, which is typically completed in 24~48 h.
The hollow and CBS nanoparticles prepared as such have relatively lower densities,
which usually translate to a higher surface area than their solid counterparts, hence
a higher degree of metal utilization (for example, in catalysis) can be expected. In
particular, CBS Pt-Ru nanoparticles were found to exhibit outstanding methanol
tolerance for the cathode reaction of the direct methanol fuel cell (DMFC). This
may be attributed to the difference in the diffusion of methanol and oxygen through
the shell of the CBS nanoparticles. Hence a diffusion-limited design, rather than the
intrinsic properties of the catalytic metals, may be used to deliver the desired
catalyst selectivity for the mitigation of the methanol crossover problem in DMFCs.
106 H. Liu et al.

3.3.1 Inside-Out Diffusion of Ag in Ag-Containing Single or


Double Shell Core-Shell Metal Nanoparticles

The inside-out diffusion of Ag in Ag-containing single or double shell core-shell


metal nanoparticles is the basis for synthesis of noble metal nanomaterials with
hollow interiors or cage-bell structures [119]. Typically, core-shell Ag-Pt
nanoparticles were used to illustrate this interesting diffusion phenomenon. The
TEM images in Fig. 3.8a, b show the initial uniform core-shell Ag-Pt nanoparticles.
After storage in toluene for seven months at room temperature, Ag diffuses out
from the interior of the core-shell Ag-Pt nanoparticles and TEM shows a product
very different from the original nanoparticles (Fig. 3.8a, c): There is increasing

Fig. 3.8 Inside-out diffusion of Ag in core-shell Ag-Pt nanoparticles. (a) TEM and (b) HRTEM
images of original core-shell Ag-Pt nanoparticles; (c) TEM and (d) HRTEM images of core-shell
Ag-Pt nanoparticles after aging in toluene for 7 months at room temperature. Reproduced with
permission from [119] Copyright American Chemical Society
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 107

presence of isolated Ag nanoparticles and hollow Pt nanoparticles in a mixture of


nanoparticles. The HRTEM image of Fig. 3.8d reveals the structures of the isolated
Ag and hollow Pt nanoparticles: Two different sets of lattice fringes with separa-
tions of 0.236 and 0.228 nm corresponding well with the (1 1 1) planes of face-
centered cubic (fcc) Ag and Pt, are found in the solid and hollow nanoparticles,
respectively. The inside-out diffusion of Ag in core-shell Ag-Ru and Ag-Rh
nanoparticles were also observed and confirmed by TEM characterizations [119].
The inside-out diffusion of Ag is promoted by the structure of the Ag seeds. The
Ag seed particles synthesized in oleylamine usually have multiply twinned
decahedral structure [119]. The twinned structure has a strong influence on the
stability of the Ag nanoparticles and the overlaid Pt shell. The decahedral particles
are mosaic structures consisting of five single crystal tetrahedrons oriented radially
about a central axis so that all five tetrahedrons share a common edge in the center,
and each tetrahedron has two sides in contact with its neighbors. However, this
model is not a space filling one. Since the angle between two (1 1 1) planes of a
tetrahedron is theoretically 70.53 , five tetrahedrons with adjoining (1 1 1) planes
will leave a gap of 7.35 . To make up for this difference, significant lattice
distortions and defects must occur [24, 120]. These imperfections in the Ag seed
nanoparticles are disruptive to the epitaxial deposition of Pt atoms on the Ag seeds,
resulting in roughness and discontinuity in the Pt shell subsequently formed. The
twinned Ag nanoparticles are also inherently unstable and could slowly be etched
by dissolved O2 and Cl dissociated from the Pt precursor. The crystallographic
defects in the twinned particles might be the active sites for the oxidative dissolu-
tion of the nanoparticles. The Agþ ions released by the O2/Cl etching of twinned
Ag nanoparticles could be re-reduced by excess oleylamine in the solution to single
crystalline Ag nanoparticles. In principle, the transformation from twinned to single
crystalline particles is reversible since O2/Cl etching should also operate on single
crystalline Ag nanoparticles. However, as pointed out by Xia and co-workers [120],
single crystalline nanoparticles are less soluble and would continue to grow slowly
even in the presence of etchant at the expense of the twinned nanoparticles. Thus,
the inside-out diffusion of Ag in core-shell Ag-Pt nanoparticles can be rationalized
by the scheme shown in Fig. 3.9. Because of the concentration gradient between the
interior of the core-shell particles and the surrounding solution, the Agþ ions
generated from the O2/Cl etching of twinned Ag seeds diffuse out through the
discontinuous Pt shell, and are reduced by oleylamine to form isolated single
crystalline Ag nanoparticles in the colloidal solution. With the progress of time
the core-shell Ag-Pt nanoparticles would eventually disappear due to the etching
and outward diffusion of Ag to leave behind a physical mixture of hollow Pt
nanoparticles and single crystalline Ag nanoparticles. Oleylamine is an indispens-
able multifunctional agent in this case: It was simultaneously a particle stabilizing
agent and the reducing agent for the etched Ag ions. Without it the nucleation and
growth of isolated single crystal Ag nanoparticles from dissolved Ag twined
particles would not be possible.
108 H. Liu et al.

Fig. 3.9 Schematic for the


inside-out diffusion of Ag in
Etch by O2/Cl-
core-shell Ag-Pt
nanoparticles. Reproduced
Reduce by amine
with permission from [119]
Copyright American
Chemical Society

by amine
Reduce
Growth
Reduce by amine

3.3.2 Monometallic and Alloy Nanoparticles with Hollow


Interiors

The inside-out diffusion of Ag in core-shell nanoparticles with Ag residing in the


core region could be used to generate hollow metal nanoparticles. In this protocol,
the preparation of core-shell Ag-M nanoparticles is an important step preceding the
fabrication of hollow noble metal nanoparticles. In addition, the inside-out diffu-
sion process could also be accelerated by introducing a chemical reagent BSPP,
which binds strongly with Ag as Agþ-BSPP coordination complexes to drive the
outward diffusion of Ag in the core-shell nanoparticles [66, 121], enabling the
completion of the inside-out diffusion of Ag in a much shorter timeframe of
24~28 h (with agitation) after mixing the core-shell Ag-M organosol with an
aqueous BSPP solution. The BSPP-Agþ coordination compounds are water soluble
and their continuous formation left behind an organosol of hollow noble metal
nanoparticles, shown by TEM (Fig. 3.10a–d for Ru, Rh, Os, and Pt, respectively) as
an increase in the image contrast between the core and shell regions, and the
development of visible discontinuity in the metal shell. The removal of Ag was
confirmed by EDX analysis of the final products [119].
With a slight modification, the protocol can be used to synthesize hollow alloy
nanoparticles. In this case, co-reduction in the presence of preformed Ag seeds
involves two or more shell precursor metals in order to form an alloy shell on the Ag
core. The second step remains the same—removal of the Ag core by BSPP
treatment. TEM images of the hollow PtRu, PtRh, PtOs, PtRuOs, and PtRhPOs
nanoparticles are given in Fig. 3.10e, f, g, h and i, respectively. In comparison with
their core-shell templates, the particle size and morphology were virtually
unchanged by the BSPP treatment, suggesting that the removal of the Ag core
from the core-shell nanoparticles did not cause the collapse of the particle
geometry [119].
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 109

Fig. 3.10 TEM images of hollow structured Ru (a), Rh (b), Os (c), Pt (d), PtRu (d), PtRh (f), PtOs
(g), PtRuOs (h), and PtRhOs nanoparticles (i) prepared upon the inside-out diffusion of Ag in core-
shell nanostructures. Reproduced with permission from [119] Copyright American Chemical
Society

3.3.3 Noble Metal Nanoparticles with Cage-Bell Structures

The protocol can be further modified to synthesize noble metal nanoparticles with a
cage-bell structure, which refers to a movable core enclosed by a shell with nano-
channels. In principle, the core and shell components may be made of metals or
alloys. To achieve the synthesis of cage-bell structured (CBS) nanoparticles, the
formation of core-shell-shell MA-Ag-MB nanoparticles, in which the inner Ag shell
serves as the sacrificial component, is most critical. The Yang group chose Au and
Pt as the core component (MA) in their study, with Ru, Os, Ir, Pt and alloy PtRh as
the outermost shell component (MB). The seed nanoparticles (MA, Au and Pt) were
overlaid with Ag first, followed by the growth of another metal (MB) shell to form
110 H. Liu et al.

Fig. 3.11 TEM images of CBS Au-Os (a), Au-Ir (b), Au-Pt (c), Pt-Ru (d), Pt-Os (e), and Pt-Pt
nanoparticles (f) prepared upon the inside-out diffusion of Ag in core-shell nanostructures.
Reproduced with permission from [119] Copyright American Chemical Society

MA-Ag-MB nanoparticles with the requisite core-shell-shell structure. After BSPP


treatment, the inner Ag layer was removed from the MA-Ag-MB nanoparticles,
leaving behind Au-Os, Au-Ir, Au-Pt, Pt-Ru, Pt-Os, and Pt-Pt nanoparticles with the
cage bell structure, as shown by the TEM images in Fig. 3.11, which show the
preservation of the size and morphology of the core-shell-shell nanoparticles in the
CBS nanoparticles. The void space between the core and the outer shell regions,
formed upon the elimination of the Ag inner shell by BSPP, is discernible by the
strong brightness contrast in TEM images.
Most of the materials reported by the Yang group are valuable and important
catalytic metals (Pt, Rh, Ru, PtRh, PtRu, etc), thus one would expect the hollow and
CBS metal nanoparticles to yield significant applications for catalysis due to their
advantage of low density, which translated to a higher surface area and would be
beneficial to the catalytic reaction.
An unique application was raised by the CBS nanoparticles. The anode and
cathode catalysts of current DMFCs are commonly based on Pt. These catalysts are
not selective to methanol oxidation reaction (MOR) or oxygen reduction reaction
(ORR) and hence any methanol crossover from the anode to the cathode through the
proton exchange membrane (PEM) can be oxidized by the cathode catalyst. This
results in the creation of a mixed potential at the cathode which degrades the fuel
cell performance [122–124]. Yang and co-workers discovered that the CBS Pt-Ru
nanoparticles have none of the failing of common Pt-based catalysts, and are
effectively methanol tolerant in oxygen reduction [119]. As a proof, Fig. 3.12a
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 111

shows the polarization curves of ORR on CBS Pt-Ru nanoparticles in the presence
of different methanol concentrations. Even with the presence of methanol in
concentrations as high as 1.0 M in the electrolyte, the catalytic reduction of oxygen
on the CBS Pt-Ru nanoparticles was hardly affected, demonstrating the effective
inhibition of methanol oxidation on the CBS Pt-Ru nanoparticles. For comparison,
oxygen reduction on a commercial E-Tek Pt/C catalyst (3~5 nm nanoparticles on
carbon powder) with and without methanol was also measured (Fig. 3.12b). The
ORR polarization curve in this case was clearly affected even at a low methanol
concentration of 0.1 M: a valley was formed at the potential for methanol oxidation.
In the CBS Pt-Ru catalyst, the catalytically active metal, i.e. Pt, is located in the
core region shielded by a porous Ru shell. Reactants must diffuse through the
porous shell of the CBS nanoparticles to access the active core for catalysis to
occur. In this case the selectivity for ORR may be caused by the porous shell
permitting only the passage of small molecule reactants. The situation is depicted
by the scheme in Fig. 3.12c. Methanol and oxygen must diffuse into the CBS
nanoparticle interior through the porous Ru shell for MOR and ORR to occur.
However, a methanol molecule is larger than an oxygen molecule (the diameters of

Fig. 3.12 Inhibition of


methanol oxidation by CBS (a)
Pt-Ru nanoparticles: (a)
Polarization curves of ORR
on CBS Pt-Ru nanoparticles
in the presence of (black
circle) 0.0 M methanol, (red
circle) 0.1 M methanol,
(green circle) 0.5 M
methanol, and (blue circle)
1.0 M methanol; (b) (b)
polarization curves of ORR
on commercial Pt/C catalyst
in the presence of (black
circle) 0.0 M methanol and
(red circle) 0.1 M methanol;
(c) Schematic illustration of
the differential diffusion
and reaction of reactants in 0.0 0.2 0.4 0.6 0.8
CBS Pt-Ru nanoparticles. Potential (V vs Ag/AgCl)
Reproduced with
permission from [119] (c) Porous Ru shell

Copyright American
Chemical Society

Catalytic
Ptcore

Reactants Products
112 H. Liu et al.

methanol and oxygen molecules are 0.44 nm and 0.34 nm, respectively). Hence the
diffusion of O2 is faster than the diffusion of methanol in CBS Pt-Ru nanoparticles,
rendering the oxidation of methanol on CBS Pt-Ru a non-competitive event. By
tailoring the structures (e.g. the size of Pt core and the porosity of Ru shell) of the
CBS nanoparticles, one would expect the ORR catalytic activity and methanol-
tolerant property of CBS Pt-Ru nanoparticles could be further enhanced.

3.3.4 Carbon-Supported Noble Metal Nanoparticles


with Hollow Interiors

The protocol based on the inside-out diffusion of Ag in core-shell nanostructures


and BSPP treatment is very general for the fabrication of noble metal nanomaterials
with hollow interiors [119]. Unfortunately, the expensive nature of BSPP is a major
obstacle to the application of this protocol for large-scale production of hollow
structured metal nanomaterials. Therefore, a universal and cost-effective approach
to rationally fabricate hollow structures for a sufficiently wide spectrum of noble
metals is undoubtedly significant for given technological applications.
Aiming at finding a more cost-effective alternative to BSPP to produce hollow
structured metal nanomaterials, the Yang group reported an improved approach to
the fabrication of carbon-supported hollow structured noble metal nanoparticles,
labeled as hNMNPs/C [125]. Analogous to the protocols they developed previ-
ously, this strategy also starts with the synthesis of core-shell noble metal
nanoparticles with Ag residing in the core regions, which are then loaded on the
carbon supports and refluxed in acetic acid to remove the original organic surfac-
tant. The refluxing in acetic acid also increased the aqueous affinity of the core-shell
nanoparticles. The carbon-supported core-shell Ag-noble metal nanoparticles are
subsequently agitated in saturated Na2S or NaCl solution to eliminate Ag from the
core region, leading to the formation of noble metal nanoparticles with hollow
interiors.
This strategy is universal for the generation of carbon-supported monometallic
and alloy noble metal nanoparticles with hollow interiors. As displayed by
Fig. 3.13, the strong brightness contrasts in the TEM and HRTEM images of
hNMNPs/C including hPt/C (a1 and a2), hPtRh/C (b1 and b2), hPtRu/C (c1 and
c2), hPtIr/C (d1 and d2), hPtOs/C (e1 and e2), hPtRhRu/C (f1 and f2), and hPtIrOs/C
(g1 and g2) show that only carbon-supported nanoparticles with hollow interiors
were obtained in the final products after treatment with aqueous Na2S (Fig. 3.13a1,
b1, c1, d1, e1, f1, g1) or NaCl solution (Fig. 3.13a2, b2, c2, d2, e2, f2, g2). Analogous to
the results observed in BSPP treatment, these electron microscopy images also
show the preservation of the size and morphology of the core-shell nanoparticles in
the hollow structured nanoparticles. In comparison with the hollow noble metal
nanostructures formed through other selective etching process [72, 119, 126], the
present synthesis by Na2S or NaCl shows remarkable simplicity and universality,
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 113

Fig. 3.13 TEM images of hPt/C (a1, a2), hPtRh/C (b1, b2), hPtRu/C (c1, c2), hPtIr/C (d1, d2),
hPtOs/C (e1, e2), hPtRhRu/C (f1, f2), and hPtIrOs/C (g1, g2) obtained by Na2S (a1, b1, c1, d1, e1, f1,
g1) or NaCl treatment (a2, b2, c2, d2, e2, f2, g2). Reproduced with permission from [125] Copyright
American Chemical Society
114 H. Liu et al.

which might be used to design and generate other hollow nanostructures for wide
application explorations.
In particular, the electrochemical measurements demonstrate that the carbon-
supported Pt-containing hollow structured noble metal nanoparticles prepared by
NaCl treatment exhibit higher catalytic activity for MOR than that of their core-
shell counterparts due to the increase of surface area induced by the hollowing
process, whereas those obtained from Na2S treatment display very poor activity for
MOR at room temperature, suggesting that S2 anions have strong poisonousness
for the surface of Pt-containing metal shells, which induce serious blockage of the
surface areas of the catalysts [125].

3.3.5 Hollow Pt and Ru Assemblies via Electrostatic


Interaction

Although the mastery over the structure and composition of metal nanoparticles is
an effective way to improve the electrocatalytic activity on a mass basis, it is
generally accepted that Pt alone has rather poor activity for methanol oxidation.
At room and moderate temperatures, Pt catalysts could be readily poisoned by
carbon monoxide (CO), an intermediate product of methanol oxidation [7, 127,
128]. Without a second oxophilic metal, Pt itself would have to dissociate water to
react away CO intermediate at higher potentials.
Instead of alloying Pt with other transitional metals (e.g. Ru, Rh, Ni, or Os)
[129–133], the Yang group reported a facile, aqueous route to fabricate the assem-
blies of hollow Pt (hPt) and ultrafine Ru nanoparticles through electrostatic inter-
action, and their applications for catalyzing methanol oxidation, the anodic reaction
in DMFCs at room temperature [134]. In this approach, negatively charged hollow
Pt nanospheres with an average size of 12 nm and positively charged ultrafine Ru
nanoparticles with an average size of 0.9 nm were first prepared, respectively.
Subsequently, they were mixed together and assemblies of hPt and Ru are
formed upon the electrostatic attraction between the particles with opposite
charges. Figure 3.14 shows the TEM and HRTEM images of as-prepared hPt-Ru
assemblies at different Pt/Ru molar ratios. As indicated, the assembly with ultrafine
Ru nanoparticles does not lead to apparent change in particle size of hPt
nanospheres. However, the comparison among these microscope images clearly
illustrates that the assembly with ultrafine Ru nanoparticles using electrostatic
interaction results in the significant change in surface roughness of hPt particles,
and the degree of the roughness increased with increasing the molar ratio of Ru in
the hPt-Ru assemblies.
Compared with the commercial PtRu alloy nanoparticles, the electrochemical
measurements show that the hPt-Ru assemblies have superior catalytic activity
toward MOR, of which the assemblies with Pt/Ru molar ratio of 2/1 are the best
[134]. XPS analyses confirm that most of the Ru presents as Ru oxides in hPt-Ru
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 115

Fig. 3.14 TEM (a, c, e) and HRTEM images (b, d, f) of electrostatic interaction based hPt-Ru
assemblies at Pt/Ru molar ratio of 2:1 (a, b), 1:1 (c, d), and 1:2 (e, f), respectively. Reproduced
from [134] with permission from the Royal Society of Chemistry

assemblies, which would be favorable for methanol oxidation upon the investiga-
tions reported by Rolison and co-workers [135, 136]. They found the homogeneous
PtRu alloy catalysts are orders of magnitude less active than mixed-phase
electrocatalysts consisting of Pt metal and hyfrous ruthenium oxides. The high
activity of a mixed-phase electrocatalysts is derived from the presence of hydrous
ruthenium oxides, the electron and proton conducting materials, which are inti-
mately mixed with Pt. Their experimental observations were subsequently con-
firmed by many other researchers [137–143]. For example, Qiu and co-workers
deposited Pt on preformed hydrous Ru oxide and observed enhanced activity for
catalyzing MOR [140]. The comparison between alloy PtRu catalysts prepared
from co-reduction of the metal precursors and from the mixing of preformed metal
colloids was carried out by Lamy and co-workers [144]. They concluded that
alloyed catalysts had the lowest activity because of the migration of Ru to the
alloy surface under the operating conditions, thereby reducing the Pt surface area
available for methanol adsorption.
The less surface dilution effect may also account for the enhanced catalytic
activity of hPt-Ru assemblies toward methanol oxidation. It has been generally
accepted that methanol oxidation commences by methanol adsorption on multiple
Pt sites (3-fold methanol adsorption sites, donated as Pt3) [122]. The surface
conditions of commercial PtRu/C catalysts and hPt-Ru assemblies could be sche-
matically described by Fig. 3.15a, b, respectively. For the alloy PtRu/C catalysts,
the atomic mixing and uniform distribution of Pt and Ru on the surface may inhibit
116 H. Liu et al.

(a) (b)
1 1

3 2 3 2

Pt Ru

Fig. 3.15 Schematic illustration for the surface conditions of alloy PtRu catalysts (a) and hPt-Ru
assemblies (b). Inlets show the possible direction for methanol adsorption on the surface of
catalysts. Reproduced from [134] with permission from the Royal Society of Chemistry

the adsorption of methanol at the directions labeled as 1 and 3 (inlet of Fig. 3.15a),
and therefore lead to negative effect on the catalytic activity for methanol oxida-
tion. However, for the hPt-Ru assemblies, several closely neighboring Pt atoms
share a ultrafine Ru cluster on the surface, not only making use of the oxygen-
containing species brought by Ru cluster, but also allowing the adsorption of
methanol at all directions (inlet of Fig. 3.15b). Compared with the surface dilution
of Pt induced by alloying with Ru, the hPt-Ru assemblies can maintain good
catalytic activity toward methanol oxidation at suitable Pt/Ru molar ratios (for
example, Pt/Ru of 2/1), although the coherent interfaces between hPt and Ru
nanoparticles in the assemblies resulted in some blockage of the surface area of
the Pt domains.

3.4 Pt-Based Heterogeneous Nanomaterials with Stellated/


Dendritic Morphologies

Extensive research efforts have been devoted towards the development of bimetal-
lic nanomaterials with heterogeneous strtucture, also called hybrid structure or
particle-on-particle structure, in which one metal is distributed at single site or at
intervals on the surface of another metal particles [6, 15, 17, 145, 146]. These
heterogeneously nanostructured materials usually exhibit catalytic properties dif-
ferent from the each one of the constituent materials due to the synergistic effect
between the two component metals. For example, the Pt nanoparticles modified by
Au clusters were found to display excellent stability for the electrocatalysis of
oxygen reduction reaction (ORR) at room temperature [16]. The electronic
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 117

interaction between Au and Pt can account for the observed stabilization of


Pt. When clusters of the softer Au metal are placed on the surface of considerably
harder Pt, the surface alloying of Au with Pt, although unlikely, also would modify
the Pt electronic structure toward a lower Pt surface energy, or lower-lying Pt d-
band state, which is helpful to increase the oxidation potential of Pt electrocatalysts
against dissolution under potential cycling regimes.

3.4.1 Pt-Containing Heterogeneous Nanomaterials via the


Seed-Mediated Growth

Analogous to the synthesis of core-shell nanostructures, the general strategy for the
fabrication of heterogeneously bimetallic nanoparticles is also the seed-mediated
growth method, whereby previously formed seed particles of one metal serve as the
sites for nucleation and growth of another metal [17, 147–151]. The advantage of
seed-mediated growth is a number of physical/chemical parameters, including the
mismatch of lattice structures and constants, the correlation of surface and interface
energies, the difference in electronegativity between the two metals, the reducing
and stabilizing agent, and even the reaction kinetics, could be selected or tuned for
directing the heterogeneous nucleation and growth mode of the second metal on the
surface of seed particles in a wet-chemistry preparation [6, 150, 152, 153]. Typical
examples include the recent reported bimetallic Pt-on-Pd hetero-nanostructures or
nanodendrites with enhanced catalytic activity for ORR [154–157]. The synthesis
of such electrocatalysts starts with the preparation of truncated Pd cores through
wet chemistry. A dense array of Pt branches are formed subsequently onto the Pd
cores from the metal precursors via a facet-selective growth. These Pt-on-Pd
heterogeneous nanostructures show two and a half times more activity on the
basis of equivalent Pt mass for the ORR than the state-of-the-art Pt/C catalyst and
five times more activity than the first-generation supportless Pt-black catalyst
(Fig. 3.16).
Recently, using the seed-mediated growth method, Kim and co-workers reported
the synthesis of AuPt bimetallic hetero-nanostructures consisting of Au nanocrystal
cores with well-defined morphologies and dendritic shells of Pt [158]. They first
prepared Au seed particles with controlled morphologies (cube, rod, or octahedron)
in aqueous phase using ascorbic acid and cetyltrimethylammonium bromide
(CTAB) as reducing agent and stabilizer, respectively. Then the aqueous Pt pre-
cursors (K2PtCl4) was added, and the Pt atoms reduced from K2PtCl4 by ascorbic
acid deposit on the surface of Au seeds, forming Pt multibranches on Au nanocrys-
tal cores. The final AuPt bimetallic heterogeneous nanostructures would follow the
original morphologies of the Au seeds, as schematically illustrated by Fig. 3.17.
The authors found that the AuPt hetero-nanostructures exhibit higher
electrocatalytic activity and durability for ORR than those of the monometallic Pt
catalyst, indicating that the formation of heterostructures can provide higher active
118 H. Liu et al.

Fig. 3.16 (a) CV, (b) hydroxyl surface coverage (ΘOH), (c) ORR polarization curves, and (d)
specific kinetic current densities (ik) for carbonsupported Pt-on-Pd and Pt catalysts. Reproduced
with permission from [155] Copyright American Chemical Society

Fig. 3.17 Schematic illustration to show the synthesis of AuPt bimetallic hetero-nanostructures
consisting of Au nanocrystal cores with well-defined morphologies and dendritic shells of
Pt. Reproduced from [158] with permission from Wiley-VCH

catalytic surfaces. Interestingly, the ORR activities are highly dependent on the
shape of the cores. The usage of Au nanooctahedron core would result in the largest
improvement of the ORR activity. Their data thus reveal the importance of the core
structure for further enhancing the activity of core–shell type nanocatalysts.
To sufficiently increase the catalytic activity but reduce the amount of Pt metals,
Wang and Yamauchi reported a facile synthesis of Pt-Pd metallic nanocages, which
possess a hollow interior and porous dendritic shell [126]. They prepared dendritic
Pt-on-Pd nanoparticles as the starting materials using an one-pot approach they
developed before [157, 159]. Then in a typical synthesis of metallic nanocages, the
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 119

Fig. 3.18 TEM (a, b) and HAADF-STEM images (c, d) for dendritic Pt-on-Pd nanoparticles
before chemical etching (a, c) and dendritic nanocages after chemical etching (b, d). Reprinted
with permission from [126] Copyright American Chemical Society

starting Pt-on-Pd dendritic nanoparticles were mixed with an excess amount of


concentrated nitric acid, and then the mixture solution was stirred for 5 days at room
temperature, allowing selective dissolution of the Pd core. The product was finally
collected by consecutive washing/centrifugation cycles with water. As shown by
the TEM images in Fig. 3.18, the nanocages are strikingly uniform in shape. The
average particle size is ca. 42 nm, which is almost the same size as the starting Pt-
on-Pd dendritic nanoparticles before chemical etching. In addition, the resultant
dendritic nanocages have well-developed hollow interiors, which quite differed
from the starting sample with solid interiors. The exteriors of the dendritic
nanocages are well covered by Pt nanoarms with an average diameter of 2 nm.
The Pt nanoarms are highly branched in various directions, resulting in the porous
wall in each individual entity. Because these metallic nanocages have sufficient
120 H. Liu et al.

catalytic active sites at both their interior and exterior surfaces, they are highly
active catalysts for MOR in comparison with other Pt materials. The developed
facile synthesis can be scaled up easily and should be highly valuable for routinely
producing new metallic nanocages with large surface areas.

3.4.2 Pt Stellates with an Ag Core or a Hollow Interior via


an One-Pot Approach

After a careful review of the literature on recent designs of electrocatalysts for fuel
cell applications [6, 147–149, 151, 154–157], Yang and co-workers proposed that
the heterogeneous growth of Pt on a metal with different electronegativity may
combine the electronic effect in catalysis with the exposure of selective active
facets through nanoparticle morphology engineering to increase the diversity in the
control of Pt catalytic properties. They have demonstrated this concept with the
design and preparation of stellated Pt nanoparticles with a Ag core (SPtNPs-A)
[160]. The bimetallic stellated nanoparticles could be formed in one-pot by the
competing growth of Pt on the twin sites of in-situ formed decahedral Ag seeds, as
shown by Fig. 3.19. Although both Ag and Pt precursors were heated in one-pot, the
Ag precursor was preferentially reduced because the reduction of Ag-oleylamine
complexes is kinetically more facile. The fast kinetics results in the formation of
multiply twinned decahedral Ag nanoparticles, which seed the subsequent deposi-
tion of Pt. The twin structure of the Ag seeds, which accounts for many interesting
electronic, optical, and catalytic properties [50, 161, 162], are essential for the
development of the stellated Ag-Pt heterogeneous nanostructure. The anisotropic
growth of Pt on the surface of Ag seeds is due to the selective deposition of Pt at
high-energy twin boundary sites.
TEM was used to follow the particle growth process. The results are shown in
Fig. 3.20. The TEM (Fig. 3.20a) and high-resolution TEM (HRTEM) images
(Fig. 3.20b) show only multiply twinned decahedral Ag nanoparticles with an

Fig. 3.19 Schematic


illustration of the
mechanism for the
formation of stellated Ag-Pt
nanoparticles in the one-pot
synthesis. Reproduced from
Ag(I) g
[160] with permission from
MacMillan
Pt(II) Pt
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 121

a b c 300
C

250

200

Counts (a.u.)
150
Cu

100 Ag

50
O Cu Ag

0
50 nm 2 nm 0 5 10 15
Energy (keV)
20 25

d e f 300

Cu
250

C
200

Counts (a.u.)
150

O Ag
100

Pt Cu
50
Pt Pt Ag

0
50 nm 2 nm 0 5 10 15
Energy (keV)
20 25

g h i 300
Cu

250

200
Counts (a.u.)
150

100 C
Ag
Pt Cu
50 Pt
Pt Ag

0
50 nm 2 nm 0 5 10 15
Energy (keV)
20 25

j k l 300
Cu

250

200
Counts (a.u.)

150 Pt

100 Ag
Pt
Cu
50 C Pt
Ag

0
50 nm 2 nm 0 5 10 15
Energy (keV)
20 25

m n o 300
Cu
C
250

200
Counts (a.u.)

150

100
Pt
Cu
Pt
50 Ag
Pt
Ag
0
50 nm 2 nm 0 5 10 15
Energy (keV)
20 25

Fig. 3.20 Representative TEM images (a, d, g, j, m), HRTEM images (b, e, h, k, n) and
corresponding EDX spectra (c, f, i, l, o) of the stellated Ag-Pt nanoparticles made after the reaction
for 30 min (a, b, c), 60 min (d, e, f), 90 min (g, h, i), 120 min (j, k, l), and 180 min (m, n, o),
respectively. Reproduced from [160] with permission from MacMillan
average diameter of 12.2 nm after reaction for 30 min, although both Ag and Pt
precursors are introduced to oleylamine simultaneously. The absence of the Pt
component at this stage is confirmed by EDX analysis (Fig. 3.20c). The Pt signal
122 H. Liu et al.

is only detected by EDX of the nanoparticles recovered from the reaction mixture
after 60 min, as shown in Fig. 3.20f. There are also significant changes in the
morphology and the size of the particles in the TEM and HRTEM images
(Fig. 3.20d, e). With the deposition of Pt on the in-situ formed Ag nanoparticles,
the average diameter is increased to 16.6 nm. Core-shell nanoparticles with a
number of Pt humps on the shell are the majority product at this time. At 90 min,
the Pt humps developed into short branches on the surface of the core-shell Ag-Pt
nanoparticles (Fig. 3.20g, h), and the Pt signal in the EDX spectrum increased
accordingly (Fig. 3.20i). This stellating growth of Pt continued until the Pt precur-
sor in the reaction mixture is completely exhausted (120 min). The final product is
SPtNPs-A with an average size of 25.2 nm (Fig. 3.20j, k) which is morphologically
distinct from the spherical nanoparticles formed in the early stages of the reaction.
EDX analysis (Fig. 3.20l) of the nanoparticles measured the Ag/Pt atomic ratio to
be approximately 1.0/1.9, which is in good agreement with the Ag/Pt atomic ratio
used in the preparation (1/2). This is suggestion that all of the Ag and Pt precursors
had been completely reduced by oleylamine under the prevailing experimental
conditions. The completion of reaction could also be confirmed by extending the
reaction time to 180 min, where no further changes in the particle size, morphology,
and Pt/Ag atomic ratio were detected by TEM, HRTEM, and EDX (Fig. 3.20o),
respectively.
The twinned structure of the Ag seeds provides an expeditious means to modify
the internal structure of SPtNPs-A. As discussed in previous section, the multiply-
twinned Ag nanoparticles are inherently unstable; known to be slowly etched by
dissolved O2 and Cl dissociated from the Pt precursor (K2PtCl4), and influence
strongly on the continuity and compactness of the overlaid Pt shell [119, 120]. The
Agþ released from the O2/Cl etching of twinned Ag seeds diffuses out through the
discontinuous Pt shell due to the prevailing Agþ concentration gradient between the
core of the Ag-Pt nanoparticles and the surrounding solution. This diffusion has
been developed into a general protocol for the fabrication of noble metal
nanoparticles with a hollow or cage-bell structure [119]. A variation of this protocol
could be to completely remove the Ag core from SPtNPs-A, resulting in hollow Pt
nanoparticles with an intact stellation morphology (SPtNPs-H, or stellated Pt
nanoparticles with a hollow interior). After aging the mixture of SPtNPs-A colloi-
dal solution in toluene with aqueous solution of BSPP for 48 h at room temperature,
the disappearance of the Ag signal in the EDX analysis of SPtNPs-A evidences the
complete elimination of the Ag core from SPtNPs-A (Fig. 3.21d). The void space in
the core region formed by the removal of Ag by BSPP could be discerned by the
strong contrast between central and surface regions in the TEM and HRTEM
images of Fig. 3.21a–c. The particle size and the stellation morphology are prac-
tically unchanged by the BSPP treatment. Hence the BSPP treatment is highly
selective and did not cause the collapse of the SPtNPs-A structure.
Specifically, the electrochemical measurements demonstrate that the SPtNPs-A
are highly active for MOR due to presence of the Ag core which alters the Pt
activity through the Pt-Ag electronic coupling effects, while the SPtNPs-H lever-
ages on the abundance of atomic steps, edges, corner atoms, as well as dominant
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 123

a b

50 nm 2 nm
c d 300
Cu

250 C

200
Counts (a.u.)

150

100 O Pt
Cu

50 Pt
Pt

0
0 5 10 15 20 25
Energy (keV)

Fig. 3.21 TEM (a), HRTEM (b, c), and corresponding EDX spectrum (d) of stellated Pt
nanoparticles with a hollow interior prepared from the stellated Ag-Pt bimetallic nanoparticles.
Reproduced from [160] with permission from MacMillan

{1 1 1} facets on the stellating geometry to increase the ORR activity of the Pt


surface [160]. The stellated morphology offers a convenient platform to construct
heterogeneous nanoparticles with very varied catalytic properties quite easily.

3.4.3 Bimetallic Au-Noble Metal Nanodendrites

Considering the great potential of the hybrid nanostructures in many catalytic


applications, the production of heterogeneously bimetallic nanoparticles consisting
of a wide spectrum of noble metals would be undoubtedly important. Employing
Au as a substitute for Ag seeds, the Yang group developed a more generic approach,
which is also based on preferential nucleation and growth of noble metals on the
active twins of Au seed particles, for the fabrication of bimetallic nanodendrites
124 H. Liu et al.

[163]. Analogous to the synthesis of bimetallic Ag-Pt stellates, this strategy starts
with the preparation of the Au nanoparticles with multiple twins in oleylamine at
elevated temperature. These multi-twinned Au particles are served as seeds for the
subsequent nucleation and growth of other noble metals, resulting in the formation
of bimetallic Au-noble metal nanoparticles with dendritic morphologies.
This strategy is general enough for the generation of dendritic bimetallic
nanoparticles of Au and other noble metals. The synthesis follows the common
seed-mediated growth protocol besides the temperature, which is 165  C for Au-Pd
and Au-Pt, 320  C for Au-Ru and Au-Ir, respectively [163]. The TEM (Fig. 3.22a,
d, g, j) and HRTEM images (Fig. 3.22b, e, h, k) of bimetallic Au-Pt (Fig. 3.22a, b),
Au-Pd (Fig. 3.22d, e), Au-Ru (Fig. 3.22g, h), and Au-Ir systems (Fig. 3.22j, k) show
that only heterogeneous nanoparticles with dendritic morphologies were obtained
in the final products, which are very different from the starting spherical Au seed
particles. The presence of the corresponding noble metals in the bimetallic
nanodendrites (Au and Pt for bimetallic Au-Pt, Au and Pd for bimetallic Au-Pd,
Au and Ru for bimetallic Au-Ru, Au and Ir for bimetallic Au-Ir) could be confirmed
by the results of EDX analyses (Fig. 3.22c, f, i, l) of these four bimetallic
nanosystems. In comparison with the dendritic or particle-on-particle
nanostructures formed through either Stranski-Krastanov or Volmer-Weber mode
[72], the present seed-mediated growth synthesis shows remarkable universality. In
addition, the mechanistic understanding for the fabrication of dendritic bimetallic
nanoparticles from multiply twinned seeds might be used to design and generate
other heterogeneous nanostructures for wide application explorations.
As reported by Yang and co-workers, the heterogeneous Au-Pt bimetallic
nanodendrites display superior catalytic activity toward MOR for the presence of
unique dendritic structures, as compared with that of commercial Pt/C catalysts
[163]. Therefore their study offers a vivid example to demonstrate the tailoring of
the catalyst properties by means of a heterogeneous construction, and provides a
method with remarkable universality for creating heterogeneously structured
nanomaterials for given technological applications.

3.4.4 Oleylamine Synthesis of Dendritic Noble Metal


Nanomaterials

Among the different strategies for synthesis of metal nanoparticles with controlled
morphologies, the particle synthesis in oleylamine is of particular interest.
Oleylamine is an amine of the fatty acid (oleic acid), and is a common reagent in
the chemical synthesis of nanoparticles [164]. It can function both as a reducing
agent and as a capping ligand to stabilize the surface of the particles. In addition,
oleylamine has very high boiling point, and can be a suitable solvent for the reaction
mixture. This means amine-based nanoparticle synthesis can usually be conducted
in a single-phase reaction system. Compared with that in two-phase system, such as
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 125

(a) (b) Cu
(c)
C

Counts (a.u.)
Cu
Cu
Au
Pt
Pt
Au Pt
Au

50 nm 2 nm 0 4 8
Energy (keV)
12 16 20

(d) (e) C (f)


Cu

Counts (a.u.)
Au

Au
Cu Pd Cu
Au
Pd

50 nm 2 nm 0 5 10
Energy (keV)
15 20 25

(g) (h) C
Cu
(i)
Counts (a.u.)

Au
Ru
Cu
Cu
Au
Au Ru

50 nm 2 nm 0 5 10
Energy (keV)
15 20 25

(j) (k) Cu
(l)
C
Counts (a.u.)

Au
Cu
Ir
Cu Au Ir Au

50 nm 2 nm 0 3 6
Energy (keV)
9 12 15

Fig. 3.22 TEM (a, d, g, j), HRTEM (b, e, h, k), and corresponding EDX spectra (c, f, i, l) of the
dendritic Au-Pt (a, b, c), Au-Pd (d, e, f), Au-Ru (g, h, i), and Au-Ir bimetallic nanoparticles (j, k, l)
as-prepared by seed-mediated growth method. Reproduced from [163] with permission from the
Royal Society of Chemistry

the case of Brust-Schiffrin reaction [165], the particle synthesis in single-phase


could typically offer the ease of kinetic control. This poses significant advantages to
the understanding of nanoparticle growth kinetics.
A number of literatures have been reported for the synthesis of metal
nanoparticles in or using oleylamine, particularly the work from the Sun group
126 H. Liu et al.

[166–178]. For example, in 2009, they reported a simplified approach using


oleylamine as both reducing agent and stabilizer for the synthesis of Fe3O4
nanoparticles via the thermal decomposition of Fe(acac)3 [175]. By varying the
volume ratio of oleylamine in benzyl, the Fe3O4 nanoparticles from 7 to 10 nm,
which show reasonably large magnetization, could be easily produced. In an earlier
study, Huo and co-workers investigated the mechanism of gold nanoparticle syn-
thesis using oleylamine as both reducing agent and protecting ligand. They found
that amine ligands are oxidized to amides in metal precursor reduction and form a
protecting layer of hydrogen bond network on the metal nanoparticles [179].
By reducing the metal precursors in oleylamine at elevated temperature
(160  C), the Yang group recently found that a large variety of noble metal
nanodendrites (NMNDs) including monometallic (e.g. Pt, Pd, Ru, Rh, Ir) and
alloy (e.g. PtPd, PtRh, PtIr, PtPdRu, PtPdRh, PtRuOs, and PtRhOs) could be
synthesized via a facile one-pot approach [180]. No extra reducing agent and
stabilizers are needed to achieve these syntheses. After reaction, the dendritic
products could be purified by precipitation with methanol, centrifugation, washing
with methanol, and re-dispersed in toluene. The TEM images of these NMNDs
were shown in Fig. 3.23, which indicate that the NMNDs were very uniform in size.
By tracking the growth process of the NMNDs using TEM, Yang and
co-workers proposed a competing mechanism to interpret the one-pot synthesis of
NMNDs in oleylamine [180]. According to Huo and co-workers [179], amine
ligands are oxidized to amides during the metal precursor reduction, forming a
protecting layer of hydrogen bond network on the metal nanoparticles. Therefore,
the mechanism for the fabrication of NMNDs in oleylamine could be summarized
by the scheme in Fig. 3.24. At initial stage, the metal precursors are reduced into
metal atoms by oleylamine, which grow into metal nanoparticles, while
oleylamides are simultaneously generated from oleylamine as capping agents to
stabilize the nanoparticles. Then the particle aggregation would compete with the
oleylamide passivation, resulting in the formation of a large number of particle
aggregates in the solution (step (a) in Fig. 3.24). Finally, the particle aggregates
continuously grow with the expense of small nanoparticles in solution via a
ripening process, e.g. Ostwald ripening [181], to form larger and more stable
nanodendrites (step (b) and (c) in Fig. 3.24).
The competition between particle aggregation and oleylamide passivation at
initial stage of the synthesis is essential for the formation of NMNDs, which might
be affected by the type of metal precursor and reaction temperature since the ligand
in different metal precursors and the temperature would have significant influence
on the reduction kinetics of the noble metal ions. For examples, concerning the Pt
(acac)2 and K2PtCl4 as metal precursors, the acetylacetonate ion in Pt(acac)2 and Cl

ions in K2PtCl4 may have different effect on the reduction of Pt(II) ions, and
therefore affect the competition between the aggregation and oleylamide passiv-
ation of Pt particles, resulting in Pt with different morphologies. Analogously, since
the particle aggregation and oleylamide generation are varied at different temper-
ature. The temperature could influence the morphology of the final metal products
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 127

Fig. 3.23 TEM images of NMNDs including Pt, Pd, Ru, Rh, Ir, PtPd, PtRh, PtIr, PtPdRu, PtPdRh,
PtRuOs, and PtRhOs synthesized in oleylamine at temperature of 160  C. Reproduced from [180]
with permission from the Royal Society of Chemistry

through affecting the competition between the aggregation and oleylamide passiv-
ation of noble metal particles.
128 H. Liu et al.

(a)

(b)

(c)

Fig. 3.24 Schematic illustration to demonstrate the mechanism for the formation of NMNDs via
the reduction of metal precursors in oleylamine at elevated temperature: (a) Competition between
particle aggregation and oleylamide passivation results in the formation of particle aggregates; (b,
c) particle aggregates grow into nanodendrites via Ostwald ripening. Reproduced from [180] with
permission from the Royal Society of Chemistry

Yang and co-workers characterized the electrocatalytic activities of a number of


Pt-containing NMNDs (Pt, PtPd, PtIr, PtPdRu, and PtRuOs) toward the room
temperature MOR and benchmarked with that of commercial PtRu/C catalysts
(E-TEK, alloy PtRu nanoparticles with ca. 3.5 nm in diameter on Vulcan XC-72
carbon support) [180]. From the comparison of current densities, the Pt-containing
alloy NMNDs show higher catalytic activities than that of Pt nanodendrites. It is
worthy to note that the PtRuOs nanodendrites particles display the highest catalytic
activity for MOR, definitely supporting that the introduction of Os to the traditional
Pt-containing alloy nanomaterials is an effective way to enhance the activity of a
MOR catalyst [182, 183], although the effect of the dendritic structure on the
catalytic activity of the alloy nanomaterials cannot be ruled out. Os is supposed
to modify the adsorptive properties of alloy PtRu towards oxygen-containing
species at potentials slightly more negative than for Ru [184]. According to the
bi-functional catalysis [185, 186], a timely supply of oxygen-containing species
would be helpful to reduce the adsorption of CO-like species, which are the
intermediate products during MOR and are the catalyst poisons, on the Pt particles,
therefore PtRuOs nanoparticles are superior to pure Pt or other Pt-containing alloy
electrocatalysts.
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 129

3.5 Pt-Based Heterogeneous Nanomaterials with Dimetric


Structures

Bimetallic heterodimers, in which one metal is distributed at single site on the


surface of another metal particles, might exhibit unique and superior properties
different from the each one of the constituent materials due to the synergistic effect
between the two component metals. A number of strategies including localized
overgrowth [148, 150], high temperature reduction [187–189], plasmon-mediated
synthesis [146], manipulation of reaction kinetics [153], and selective transforma-
tion [190], have been developed for the synthesis of bimetallic heterodimers.
Recently, Puntes and coworkers reported the fabrication of Au-Pt heterodimers at
room temperature using oleylamine as reducing agent (Fig. 3.25), and found that
the Pt seeds play an important role for the nucleation and growth of Au region in the
heterodimer structures [191]. The Pt seeds markedly accelerate the reaction, serving
both as nucleation platforms and as an initial catalytic reducer of the Au ions in
solution, as supported by the initial shrinking of the Pt seeds.

Fig. 3.25 (a) Schematic for the heterodimer formation, (b) TEM images and (c) plot of diameter
vs. time for the heterodimer evolution. (Inset) TEM of the Pt seeds. All scale bars correspond to
20 nm. Reproduced from [191] with permission from the Royal Society of Chemistry
130 H. Liu et al.

Unfortunately, up to date, unlike the abundant studies on the semiconductor/


metal oxide-noble metal heterodimers, which will be discussed in next section, the
literatures reported on heterodimers consisting of two chemically distinct metals are
still very limited due to the lack of synthetic inaccessibility. The most commonly
used seed-mediated growth method usually results in the formation of bimetallic
nanoparticles with core-shell or dendritic aggregates [149, 151, 154–156, 192–
195]. Considering the further exploration of the properties and consequent appli-
cations of these heterogeneous nanostructures, the development of effective
approaches for the production of bimetallic heterodimers would be undoubtedly
important and pose significant challenges.
In their recent work, Yang and co-workers presented a facile and high-yield
approach to the synthesis of bimetallic heterodimers consisting of Ag and hollow
structured Pt nanoparticles, labeled as Ag-hollow Pt (Ag-hPt), which is based on an
unique diffusion phenomenon of Ag they observed in core-shell Ag-Pt
nanoparticles with Ag residing in the core region [196]. In their strategy, core-
shell Ag-Pt nanoparticles are first prepared in oleylamine and re-dispersed in an
organic medium (e.g. toluene) as starting templates. The core-shell Ag-Pt templates
are subsequently converted into Ag-hPt heterodimers via the inside-out migration
of Ag in core-shell Ag-Pt nanoparticles. A heating treatment at elevated tempera-
ture is employed to promote the inside-out diffusion of Ag from the core region of
the core-shell nanoparticles. The fabrication of bimetallic Ag-hPt heterodimers via
the inside-out migration of Ag in core-shell Ag-Pt nanoparticles can be rationalized
interpreted as following. As mentioned in previous sections, the twinned Ag seeds
used for the formation of starting core-shell Ag-Pt nanoparticles are inherently
unstable and could slowly be etched by dissolved O2 and Cl dissociated from the
Pt precursor to generate Agþ ions [119, 120]. These Agþ ions are then diffused out
through the discontinuous Pt shells, which are produced during the epitaxial
deposition of Pt atoms on the Ag seeds with a large number of imperfections on
their surfaces, due to the Agþ concentration gradient between the interior of core-
shell Ag-Pt nanoparticles and the surrounding solution. At suitable temperature
(e.g. 80  C), the Agþ ions diffused out from core-shell Ag-Pt nanoparticles are
re-reduced by excess reducing agent (e.g. oleylamine) in toluene to form single
crystalline Ag nanoparticles decorated on the outer surface of the Pt shell. The Ag
nanoparticles grew with the continued outward diffusion of Agþ ions until the Ag
core is completed depleted, leaving behind a colloidal solution of bimetallic Ag-hPt
with heterogeneous nanostructures. Finally, the Ag nanoparticles on the surface of
the Pt shell undergo a ripening process to form larger and more stable domains on
the surface of Pt shell.
After heating the core-shell Ag-Pt colloidal solution in toluene at 80  C for 72 h
under air, Ag is migrated to the surface of Pt shells from the interior of core-shell
Ag-Pt nanoparticles, resulting in the formation of bimetallic heterodimers
consisting of Ag solid particles and Pt nanoparticles with a hollow interior. The
TEM and STEM images (Fig. 3.26a, b) after heating treatment in toluene show that
only heterodimers with strong imaging contrast in different domains are observed
in the final products, which are clearly distinct from the starting core-shell
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 131

(a) (b)

20 nm 20 nm

Cu (c) Ag
Pt
(d)
Intensity (a.u.)

C
Counts (a.u.)

Pt

Ag Pt
Cu
Cu
Pt Ag

0 5 10 15 20 25 0 5 10 15 20 25 30 35 40 45
Energy (keV) Position (nm)

(e) (f)

5 nm 5 nm

Fig. 3.26 (a) TEM image, (b) STEM image, (c) STEM-EDX spectrum, (d) elemental profiles,
and (e, f) HRTEM images of bimetallic Ag-hPt heterodimers synthesized via the inside-out
migration of Ag in core-shell Ag-Pt nanoparticles at elevated temperature. Reproduced from
[196] with permission from the Royal Society of Chemistry
132 H. Liu et al.

nanoparticles. The void space between the core and the outer Pt shell regions,
formed upon the inside-out migration of the Ag cores by heating treatment, is
discernible by the strong brightness contrast in TEM and STEM images, while Ag
appears as lighter domains sharing the solid-state interfaces with the Pt domains. As
indicated by Fig. 3.26c, the EDX analysis of an arbitrary single dimer boxed in the
high-angle annular dark-field STEM image (Fig. 3.26b) confirms the presence of
Ag and Pt after heating treatment of core-shell Ag-Pt nanoparticles. A more direct
evidence for the formation of Ag-hPt heterodimers is provided by the line scanning
analysis of an arbitrary single dimer in the high-angle annular dark-field STEM
mode. As shown in Fig. 3.26d, the Pt signal is present at one side throughout the
heterodimer whereas the Ag signal is detected only at the other side. The interplanar
spacings of approximate 0.24 and 0.23 nm indicated in the HRTEM images of the
biemtallic Ag-hPt heterodimers (Fig. 3.26e, f) correspond to the (1 1 1) planes of
face-centered cubic (fcc) Ag and Pt, respectively.
To demonstrate the effect of the structure of the bimetallic Ag-Pt systems on the
catalytic properties, the core-shell Ag-Pt nanoparticles and bimetallic Ag-hPt
heterodimers were loaded on Vulcan carbon and examined for their electrocatalytic
activities towards the room-temperature ORR. Cyclic voltammograms of core-
shell Ag-Pt/C and bimetallic Ag-hPt/C in argon-purged 0.1 M HClO4 at room
temperature are used to obtain the electrochemical active surface areas (ECSAs)
from the hydrogen adsorption/desorption regions (0.2 to 0.1 V vs Ag/AgCl). As
shown in Fig. 3.27a, the ECSAs normalized by the mass of Pt for core-shell Ag-Pt
nanoparticles and bimetallic Ag-hPt heterodimers are 40.6 m2 gPt1 and 48.8 m2
gPt1, respectively. The inside-out migration of Ag from core-shell Ag-Pt
nanoparticles may lead to the increase of ECSAs by releasing the inner surface
of Pt shell, whereas the growth of Ag domains on the outer surface of Pt shell
would result in the decrease of ECSAs due to the contact interfaces between hollow
Pt and Ag in the heterodimers, which may induce some blockage of the surface
area of the Pt shells. These two effects might have offset each other, such that the
ECSA of bimetallic Ag-hPt heterodimers is only slightly higher than that of core-
shell Ag-Pt nanoparticles.
Figure 3.27b shows the ORR polarization curves in the potential range of
0.8–0 V for core-shell Ag-Pt nanoparticles and bimetallic Ag-hPt heterodimers in
oxygen-saturated 0.1 M HClO4 at room temperature. The half-wave potentials for
the core-shell Ag-Pt nanoparticles and bimetallic Ag-hPt heterodimers are 518 mV
and 606 mV, respectively. It has been noted that the half-wave potential for
bimetallic Ag-hPt heterodimers is much more positive than that of core-shell
Ag-Pt nanoparticles, indicating that the Pt shells in the Ag-hPt heterodimers have
higher catalytic activity for ORR in comparison with that of core-shell Ag-Pt
nanoparticles under the experimental conditions. Furthermore, chronoamperometry
of core-shell Ag-Pt nanoparticles and bimetallic Ag-hPt heterodimers at 0.6 V in
oxygen-saturated 0.1 M HClO4 solution is used to obtain some indications of the
long-term performance of the catalysts in ORR. As shown in Fig. 3.27c, the
“steady-state” specific activity of bimetallic Ag-hPt heterodimers is higher than
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 133

3
(a)
2 Ag@Pt
Ag-hPt
1

j (mA cm-2)
0

-1

-2

-3

-4
-0.2 0.0 0.2 0.4 0.6 0.8 1.0
E (V vs Ag/AgCl)

10
0 (b) (c)
0
-1

-2 Ag@Pt -10
j (mA cm-2)

-3 j (mA cm-2) -20


Ag-hPt
-4
-30 Ag@Pt
-5
Ag-hPt -40
-6

-7 -50

-8 -60
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 2000 4000 6000 8000 10000
E (V vs Ag/AgCl) E (V vs Ag/AgCl)

Fig. 3.27 (a) Cyclic voltammograms of core-shell Ag-Pt nanoparticles and bimetallic Ag-hPt
heterodimers in argon-purged HClO4 (0.1 M) at room temperature, scan rate: 50 mV s1; (b) ORR
polarization curves for core-shell Ag-Pt nanoparticles and bimetallic Ag-hPt heterodimers,
recorded at room temperature in an O2-saturated HClO4 solution (0.1 M) at a sweep rate of
20 mV s1 and a rotating speed of 1600 rpm; (c) chronoamperograms of core-shell Ag-Pt
nanoparticles and bimetallic Ag-hPt heterodimers at 0.6 V in an O2-saturated HClO4 solution
(0.1 M) at room temperature and a rotating speed of 1600 rpm. Reproduced from [196] with
permission from the Royal Society of Chemistry

that of core-shell Ag-Pt nanoparticles after 2.5 h, indicating the greater stability of
the bimetallic heterodimers.
Considering the electronic coupling between Ag and Pt are present both in core-
shell Ag-Pt nanoparticles and bimetallic Ag-hPt heterodimers although it is asym-
metric in the latter, the lower catalytic activity of the core-shell Ag-Pt nanoparticles
in comparison with that of the bimetallic Ag-hPt heterodimers could be attributed to
the tensile effect imposed on the Pt shell by the Ag core in core-shell structures.
Since Ag has larger lattice parameter (0.4090 nm) than that of Pt (0.3923 nm), the
Ag cores would exert a tensile effect on the Pt atoms deposited on their surfaces
[52]. In a common ORR process, i.e. the series 4 electron pathway, the reactions
involve both the breaking of an O–O bond and the formation of O–H bonds [9, 62,
63]. The most active Pt-based catalyst should have the d-band center with an
intermediate value since the optimal ORR catalyst needs to facilitate both bond-
breaking and bond-making steps without hindering one or the other [197, 198]. How-
ever, the tensile effect of Ag core on Pt shell, which leads to narrower d-band,
results in an up-shift of the d-band center of the Pt shell. As has been discussed in
134 H. Liu et al.

previous section (Sect. 3.2), the surface of a Pt-based catalyst with high value of
d-band center tends to bind adsorbents too strongly, thereby enhancing the kinetics
of dissociation reactions producing these adsorbents. In this case, the Pt shells may
suffer from hindered bond-making step, therefore are less active for ORR. On the
other hand, the bimetallic Ag-hPt heterodimers shows good catalytic activity
toward ORR possibly because they have a more suitable d-band center to balance
the breaking of O-O bonds and the formation of O-H bonds.
The purpose of Yang and co-workers is to provide an unique strategy (inside-out
migration of Ag in core-shell Ag-Pt nanoparticles) to fabricate bimetallic
Ag-hollow Pt nanoparticles with heterodimeric structures and to demonstrate the
physical/chemical properties of a bimetallic system might be tuned by tailoring its
structure (by relegating Ag to the core or on the shell region in bimetallic Ag-Pt
nanoparticles). Their study offers a vivid example to demonstrate the tuning of the
material properties by means of a structural tailoring. In addition, the mechanistic
understanding of the structural transformation from core-shell to heterogeneous
nanoparticles might be used to design and fabricate other heteronanostructures for a
given application.

3.6 Pt-Based Composite Nanomaterials

Besides the more conventional heterogeneous nanomaterials, e.g. core-shell, hol-


low, and bimetallic heterostructures, there has been increasing interest devoted
towards the development of semiconductor-metal nanocomposites that consist of
different classes of materials with solid-state interfaces [71, 199–225]. This type of
nanostructures combines materials with distinctly different physical and chemical
properties to yield a unique hybrid nanosystem with multifunctional capabilities,
and tunable or enhanced properties that may not be attainable otherwise. The early
studies of semiconductor-metal nanocomposites involved different metals (e.g. Au,
Ag, and Pt) deposited on or doped in TiO2 powders for photocatalytic applications
[226–232]. In these structures, the metal domain induces the charge equilibrium in
photoexcited TiO2 nanocrystals, to affect the energetics of the nanocomposites by
shifting the Fermi level to more negative potentials. The shift in Fermi level is
indicative of improved charge separation in TiO2-metal systems, and is effective
towards enhancing the efficiency of photocatalysis [233, 234].

3.6.1 Nanocomposites Consisting of Semiconductors


and Gold

The nanocomposites constructed by depositing gold on the surface of semiconduc-


tor nanocrystals are the topics being studied most sufficiently [71, 199–202, 205,
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 135

209, 235–242]. Within the last decade have wet-chemistry methods blossomed and
become a powerful approach toward the synthesis of semiconductor-Au
nanocomposites. The deposition of noble metals on the surface of semiconductor
nanocrystals is usually conducted in nonpolar organic solvent via seed-mediated
growth at room temperature. In detail, the semiconductor nanocrystals with high
quality are firstly prepared in organic solvent at elevated temperature, and simul-
taneously, the metal precursors are also transferred into nonpolar organic solvent
using established methods. The semiconductor organosol and the metal precursor
solution in organic solvent are then mixed together, followed by addition of weak
reducing agent (usually dodecylamine, DDA). The DDA reduces the metal pre-
cursors into metal atoms in the presence of semiconductor nanocrystals, which
nucleate and grow on the surface of semiconductor nanocrystals, resulting in
formation semiconductor-noble metal nanocomposites. Noble metal precursors
usually cannot dissolve in nonpolar organic solvents. However, for gold, there are
a number of approaches, e.g. Brust-Schiffrin method [165], ethanol-mediated
general phase transfer protocol [71], and dodecyldimethylammonium (DDAB)
facilitated strategy developed by Banin and coworkers [199], could be used to
transfer its precursors, e.g. commonly used HAuCl4, from aqueous phase to non-
polar organic solvents. In addition, the successful deposition of noble metals on the
surface of semiconductor nanocrystals has some specific requirements on the
reducing agent. For example, too strong reducing agent, e.g. NaBH4, usually
leads to the separated nucleation of noble metal atoms in the solution instead of
growth on semiconductor nanocrystals. Therefore, although the ethanol-mediated
general phase transfer protocol could make a wide variety of metal precursors
dissolve in nonpolar organic solvents [71], only the deposition of gold on the
surface of semiconductor nanocrystals gain most sufficient study over the past
decade.
Banin and co-workers are pioneers in nanocomposites consisting of semicon-
ductor and gold, and they made a breakthrough in 2004 [199]. They demonstrated a
solution synthesis for nanocomposites via the selective growth of gold tips on the
apexes of hexagonal-phase CdSe nanorods at room temperature (Fig. 3.28). The
authors prepared CdSe rods and tetrapods of different dimensions by high-
temperature pyrolysis of suitable precursors in a coordinating solvent containing
a mixture of trioctylphosphineoxide and phosphonic acids. They dissolved AuCl3 in
toluene with the addition of dodecyldimethylammonium bromide (DDAB) and
dodecylamine, which serves as reducing agent. Then they mixed this solution at
room temperature with a toluene solution of the colloidal-grown CdSe nanorods or
tetrapods. After the reaction, the composite products were precipitated by addition
of methanol, separated by centrifugation, and re-dispersed in toluene for further
characterizations. The novel nanostructures display modified optical properties due
to the strong coupling between the gold and semiconductor components. The gold
tips show increased conductivity, as well as selective chemical affinity for forming
self-assembled chains of rods. The architecture of these nanostructures is qualita-
tively similar to bifunctional molecules such as dithiols, which provide two-sided
chemical connectivity for self-assembly and for electrical devices, and contacting
136 H. Liu et al.

Fig. 3.28 TEM images showing controlled growth of Au onto the tips of CdSe quantum rods. (a)
Original rod sample, 29  4 nm. (b–d) Rod samples after Au treatment using gradually increased
AuCl3 concentrations; increased Au tip sizes are visible. Reproduced from [199] with permission
from the American Association for the Advancement of Science

points for colloidal nanorods and tetrapods. The Banin group later reported the
synthesis of asymmetric semiconductor–metal heterostructures whereby gold was
grown on one side of CdSe nanocrystalline rods and dots. Theoretical modeling and
experimental analysis showed that the one-sided nanocomposites were transformed
from the two-sided architectures through a ripening process [235].
Following the breakthrough made by Banin group, various strategies were
developed for the synthesis of semiconductor–Au nanocomposites by anisotropic
growth of gold on the surface of semiconductor nanocrystals through chemical
reduction, physical deposition or photochemistry. The structure displayed in
Fig. 3.29 for nanocomposites composed of PbS and gold was reported by Yang
and coworkers [238]. They chose PbS prepared in oleylamine as the target semi-
conductor nanocrystal because it has a typical cubic crystal structure and can easily
be produced in controlled shapes and sizes. They dissolved gold precursors into
toluene using Brust-Schiffrin method [165], and then mixed them together,
followed by addition of dodecylamine served as reducing agent, resulting in
formation of PbS-Au nanocomposites. High quality and monodisperse PbS-Au1,
PbS-Au4 nanostructures and PbS-Aun nanocubes could be obtained by controlling
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 137

Fig. 3.29 (a, b) TEM images of the ordered array of PbS-Au4 heterogeneous nanostructures; (c,
d) TEM images of a typical PbS-Au4 nanostructure; (e) HRTEM image of a typical PbS-Au4
nanostructure; (f) SAED pattern and (g) FFT pattern obtained from the ordered array of PbS-Au4
heterogeneous nanostructures. Reprinted with permission from [238]. Copyright American Chem-
ical Society

the molar ratio of PbS and Au precursors. Their method provides a reasonable
method to control the size, number, and distribution of gold nanocrystals on semi-
conductors, which is important to the design and fabrication of composite
nanomaterials for catalysis, optoelectronic devices, and chemical/biological sen-
sors. Owing to their narrow size distribution and intrinsic high symmetry, the
as-synthesized PbS-Au4 nanocomposites could easily self-assemble into hexagonal
arrays. Nonlinear optical measurements show that the PbS-Au4 nanostructure could
alter the nonlinear response time of PbS nanocrystals, thereby confirming the
applicability of metal-decorated semiconductor nanocrystals in the development
of saturable absorber devices.
In a recent study, Yang and co-workers systematically conducted the investiga-
tion into the deposition of Au/Ag on the surface of metal sulfide nanocrystals
[243]. Upon the characterizations using TEM, HRTEM, STEM, and EDX, they
found that both the reaction time and metal/semiconductor precursor ratio have
138 H. Liu et al.

apparent effect on the deposition of Au on the surface of semiconducting metal


sulfide nanocrystals. With the increase of reaction time or metal/semiconductor
precursor ratio, both the number and the size of the Au domains on each metal
sulfide nanocrystal are increased. The deposition of Au on the metal sulfide could
be fulfilled in 30 min, suggesting that the surface of semiconducting metal sulfide
nanocrystal has catalytic property for the reduction of gold. The deposition mode of
Au on metal sulfide nanocrystals might be related to the structural similarity of
these two different type of nanoparticles. With the increase of the structural
similarity, the number of the sites on the surface of metal sulfide nanocrystals for
Au deposition is increased. Specifically, for the PbS nanocrystals, Au nucleates and
grows at multiple sites on the surface of PbS nanocrystals, while Ag exchanged
with the Pb in PbS nanocrystals to form core-shell PbS@Ag2S nanostructures.

3.6.2 Nanocomposites Consisting of Semiconductors


and Other Noble Metals

Motivated by the unique property and great potentials in wide variety of applica-
tions, the syntheses of nanocomposites consisting of semiconductor and noble
metals other than gold have also received significant attention in recent years
[203, 204, 244]. Actually, early in 80s of last century, researchers had loaded
platinum (Pt) or Rhodium (Rh) on the surface of CdS or ZnS semiconductors for
producing hydrogen [245–247]. However, the overall size of the semiconductor
substrates in those reports have exceeded the scope of nanotechnology.
The Alivisatos group conducted a pioneer study on the photo-deposition of Pt on
the surface of CdS and CdSe/CdS nanocrystals [244]. Different from the early
aqueous-based deposition of Pt on the surface of semiconductors using photore-
duction approach, they performed the deposition reaction in organic phase, which
includes CdS nanorods, a organic soluble Pt precursor—(1,5-cyclooctadiene)
dimethylplatinum (II) ((CH3)2PtCOD), and excess of a tertiary amine, which is
used as hole-scavenging gent. An excitation wavelength of 458 nm was chosen by
the researchers to ensure that CdS is the only light-absorbing component, thus
preventing homogeneous nucleation of platinum. After irradiating the reaction
mixture for a while under an inert atmosphere, the solution turns from translucent
yellow to translucent brown, and the fluorescence of CdS is quenched. The TEM
images before and after irradiation show the formation of heterostructures
consisting of small nanoparticles positioned along the length of the CdS nanorods
(Fig. 3.30). The average diameters of the nanoparticles range from 1.5 to 2.7 nm,
depending on the reaction conditions. The results obtained by the Alivisatos group
is in contrast to CdSe-Au heterostructures synthesized by thermal methods, where
the gold deposition often occurs preferentially on the nanorod ends [199, 235]. The
authors also found that the photodeposition of Pt on CdS shows a strong depen-
dence on the nature of the amine used. Highest yields are observed with bulky
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 139

Fig. 3.30 Photodeposition of Pt on CdS nanorods. (a) TEM images of CdS nanorods before
irradiation and (b) of the photodeposition product. After exposure to light, Pt nanoparticles appear
along the length of the nanorods. Reproduced from [244] with permission from John Wiley and
Sons

tertiary amines. Comparison of octylamine, dioctylamine, and trioctylamine reveals


no photodeposition for the primary amine, very little for the secondary, and the
most for trioctylamine.
Subsequently, the Banin group reported the synthesis of rod-like CdSe-Pt
nanocomposites and their photocatalytic property toward methylene blue
[203]. They produced CdSe nanorods (70 nm  8 nm) using the previously
established method [248]. After synthesis the CdSe nanorods were transferred to
an aqueous solution by exchanging the alkylphosphine surface ligands with
mercapto-undecanoic acid (MUA) [249]. In brief, the CdSe nanorods are dissolved
in chloroform along with MUA and the solution is stirred until the MUA dissolved
completely. In a different vial KOH is dissolved in triple distilled water (TDW) to
give a pH 14 solution. The stock KOH solution is added to the nanorod/MUA
solution, followed by vigorous stirring until a white emulsion is formed. The
aqueous and organic phases are separated by centrifuging. After centrifugation
the water phase is clear and brown colored, indicating the presence of phase-
transferred CdSe nanorods, while the organic phase remains nearly colorless. The
aqueous nanorod solution contained excess MUA, which is removed by precipitat-
ing the rods with methanol and centrifuging. The nanorods were then re-dissolved
in TDW for next utilization.
The deposition of Pt on the surface of CdSe nanorods could be achieved through
a facile route. The platinum precursor (PtCl4) is dissolved in water and mixed
vigorously with the aqueous CdSe nanorod solution at room temperature for 2 days.
A dark brown/black precipitate is formed and collected by centrifugation to yield
the CdSe–Pt composite particles. No additional reducing agent is needed during the
deposition process. A comparison between TEM images of samples taken before
140 H. Liu et al.

and after the reaction with platinum shows that the mean length of the CdSe
nanorods is reduced—from 70 to 55 nm.
The above-mentioned method developed by the Banin group for growth of Pt on
the surface of CdSe nanorods is very complex and time consuming. More impor-
tantly, although the rod-like CdSe-Pt nanocomposites they prepared have superior
visible light photocatalytic activity for the reduction of methylene blue, the
nanocomposites are easily aggregated in aqueous phase, and this lead to a signif-
icant decrease of the surface area available for the photocatalytic reduction of
methylene blue. On the other hand, the authors apply phase transfer to avoid the
use of expensive organic metal precursors for producing semiconductor-noble
metal nanocomposites in aqueous phase, and hence offer a good strategy to deposit
noble metals on the surface of semiconductor nanocrystals.
Mokari and coworkers prepared CdS nanorods in organic solvent, and they did
not transfer them into aqueous phase. Instead, they directly deposited Pt on the
surface of CdS nanorods using platinum acetylacetonate (Pt(acac)2) as metal
precursor [204]. The method they used to synthesize CdS nanorods is very simple:
A mixture of CdO, trioctylphosphine oxide, and tetradecylphosphonic acid is
heated at 300  C for 15 min. The temperature is then raised to 320  C and a solution
of S dissolved in trioctylphosphine is injected before decreasing the temperature to
300  C. The reaction mixture is heated for 45 min, cooled, and then precipitated
with methanol and separated by centrifugation.
For the deposition of Pt metal, oleic acid, oleylamine, and 1,2-hexadecanediol
are heated in diphenyl ether at 80  C under vacuum for 30 min to remove traces of
water. The Pt acetylacetonate is added to a suspension of CdS rods in dichloroben-
zene and heated at 65  C for 10 min to promote dissolution of the Pt precursor. The
mixture of surfactants and diphenyl ether is purged with nitrogen and heated to
200  C before injecting the Pt precursor and semiconductor rods. After several
minutes, the reaction is removed from heat and quenched in a water bath. The
CdSe-Pt composite product is washed twice by precipitation in ethanol followed by
centrifugation, and then separated twice by centrifugation.
Analogous to the deposition of Au on the CdSe nanorods, as reported by Banin
and coworkes [199], upon the concentration of Pt precursor, Pt is also selectively
grown at one or two tips of the CdS nanorods since the reactivity of the nanorods is
higher at the tips than along the body of the rod due to the increased surface energy,
as shown by Fig. 3.31. The method developed by Mokari and coworkers could be
easily extended to deposit bimetallic PtNi or PtCo on the tips of CdS nanorods, and
these heterostructures might be of special interest for a variety of applications
including photocatalysis, water splitting, and magnetic applications.

3.6.3 Nanocomposites Consisting of Ag2S and Noble Metals

The first example to apply the semiconductor-noble metal nanocomposites for


catalyzing the reactions (MOR and ORR) in DMFCs is the work reported by
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 141

101
002
100

110
113
102
bulk Cds

30 40 50 60 70 80

Intensity (a.u.)
Cds rods

20 nm

Pt-Cds

Pt (111)

Pt (200)

Pt (220)
20 30 40 50 60 70 80
2θ (degrees)

20 nm
Pt

Intensity (a.u.)
Cu Pt

S Pt
Cd Cu Pt
Pt

2 4 6 8 10 12 14
20 nm
Energy (keV)

Fig. 3.31 Selective growth of Pt nanoparticles with different sizes on CdS nanorods, (a) CdS rods
(120 nm  4 nm), (b) CdS with small single Pt tips (4.3 nm), (c) CdS with larger double Pt tips
(5.7 nm), (d) XRD patterns of CdS rods and Pt-CdS hybrid structures with corresponding CdS and
Pt bulk patterns (stick patterns shown above and below, respectively), and (e) selected area EDS
spectrum of a single Pt tip, with inset HRTEM image of two Pt-CdS hybrids. Reproduced with
permission from [204] Copyright American Chemical Society

Yang and Ying in 2011. They demonstrated an aqueous route for the synthesis of
nanocomposites consisting of Ag2S and different noble metals [213]. By reducing
various noble metal precursors using citrate in aqueous phase in the presence of
preformed Ag2S nanocrystals, uniform semiconductor-noble metal heterogeneous
142 H. Liu et al.

nanostructures are obtained as the dominant product. In addition to binary


nanocomposites, ternary and quaternary hybrid systems are also achieved via the
successive deposition of different noble metals on the surface of Ag2S nanocrystals.
In particular, the Pt-containing nanocomposites are found to exhibit superior
catalytic activity for MOR, the key reaction in direct methanol fuel cell (DMFC),
due to the electronic coupling effect between the ultrafine Pt crystallites and the
semiconductor domains.
A large number of literatures have been devoted to the synthesis of Ag2S
nanocrystals in organic media [250–256]. For example, the Zhao group reported
an one-step route to control the organization of semiconductor Ag2S nanocrystals
[253]. In a typical synthetic process, they mixed aqueous AgNO3 solution and
toluene solution of dodecylthiol to form a yellowy micellar solution under
inHowever, the tensile effect of Ag ively stirring at room temperature. After
30 min, aqueous Na2S solution was dropped into the microemulsion for the
formation of Ag2S nanocrystals, which were usually self-assemble into hexagonal,
quasi-orthogonal, and chain-like arrays on the TEM grid. There are also a few of
studies reported on the synthesis of Ag2S nanocrystals in aqueous phase, either use
protein [257, 258], or polymeric membrane [259] to passivate the surface of Ag2S
nanocrystals for preventing them from aggregation in solution. However, their self-
assembly and poor quality in term of morphology, size, and size distribution enable
them not suitable substrates for the deposition of noble metals.
Alternatively, Yang and Ying developed a room-temperature method based on
the reaction between sodium sulfide (Na2S) and coordinating compounds formed by
Agþ ions and BSPP to derive aqueous dispersible Ag2S nanocrystals [213]. Typi-
cally, 600 mg of BSPP were added to 300 mL of aqueous AgNO3 solution (1 mM)
in a 1000-mL beaker. The mixture was stirred for 1 h to form BSPP-Agþ complexes
[66, 121], followed by the prompt addition of 10 mL of aqueous Na2S solution
(50 mM), which results in a series of color changes before finally arriving at a
brown Ag2S hydrosol. A TEM image of the as-prepared Ag2S nanocrystals is
shown in Fig. 3.32a. The nanocrystals are spherical, nearly monodispersed, and
have an average size of ca. 7.2 nm. The HRTEM image (Fig. 3.32b) illustrates the
lattice planes in these nanocrystals, showing an interplanar spacing of ca. 0.26 nm,
which corresponds to the (121) planes of monoclinic Ag2S.
The Ag2S nanocrystals are used as seeds for the formation of nanocomposites
with different metals. With sodium citrate as reducing agent, under the experimen-
tal conditions (aqueous phase and elevated temperature), noble metals nucleated
preferentially on the existing Ag2S nanocrystals, rather than homogeneously. The
TEM images of binary Ag2S-Au, Ag2S-Pt, Ag2S-Os, and Ag2S-Pd nanocomposites
are displayed in Fig. 3.33a, b, c, and d, respectively. The deposition of noble metals
on the Ag2S nanocrystals could be clearly identified via brightness contrast, and
confirmed by the EDX analysis of an arbitrary single particle with HAADF-
STEM [213].
The final morphology of the nanocomposites depended on whether the surface of
substrate particles allows for only a single nucleation site or multiple ones. Unlike
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 143

Fig. 3.32 (a) TEM and (b) HRTEM images of the as-prepared Ag2S nanocrystals. d ¼ 7.2 nm,
σ ¼ 0.6 nm, σ ¼ 8.4 %. σ and σ are the standard derivation and relative standard derivation,
respectively. Reproduced from [213] with permission from John Wiley and Sons

Fig. 3.33 TEM images of binary Ag2S-Au (a), Ag2S-Pt (b), Ag2S-Os (c), Ag2S-Pd (d), multiple
Ag2S-Au-Pt (e), Ag2S-Au-Os (f), Ag2S-Pt-Os (g), and Ag2S-Au-Pt-Os nanocomposites (h).
Reproduced from [213] with permission from John Wiley and Sons

the face-centered cubic (fcc) or hexagonal materials, monoclinic Ag2S has very
complicated crystal structures. It has many facets with different lattice spacing,
which provides favorable sites to match the lattice planes of various noble metals
for their epitaxial growth on the substrate seeds. The experimental results show that
only a single site on the Ag2S nanocrystal surface is suitable for the nucleation of
gold clusters, but multiple sites existed over the surface of Ag2S seed for Pt and Os
noble metals. An extreme case is observed in the Ag2S-Pd system (Fig. 3.33d).
Numerous sites are provided for the nucleation of Pd nanoparticles, which could
grow and eventually coalesce to form a continuous shell on each Ag2S nanocrystal.
144 H. Liu et al.

The species-dependent features of noble metal deposition on Ag2S nanocrystal


are further employed to derive multiple semiconductor-metal nanocomposites. The
complex semiconductor-noble metal nanocomposites including Ag2S-Au-Pt, Ag2S-
Au-Os, Ag2S-Pt-Os, and Ag2S-Au-Pt-Os, have been prepared by successive reduc-
tion of noble metal precursors using citrate in the presence of preformed Ag2S
nanocrystals, as characterized by TEM images in Fig. 3.33e, f, g, h. The growth of
more than one type of metal on each Ag2S nanocrystal demonstrates that the
nucleation sites on the surface of Ag2S nanocrystal are specific to different metals,
providing for the possibility of creating complex nanocomposites between Ag2S
and multiple metals via a simple and flexible route.
The Pt-containing Ag2S-noble metal nanocomposites are attractive for catalytic
application due to their relative small Pt domain size in the nanocomposite systems.
The smaller Pt crystallite size usually translates to a higher surface area, which
would be beneficial to the catalytic reaction. The electrochemically active surface
area (ECSA) of Pt in Pt-containing Ag2S-noble metal nanocomposites could be
determined using cyclic voltammetry. Although the contact interfaces between Pt
and Ag2S in the nanocomposites result in some blockage of the surface area of the
Pt domains, the average ECSA of the Pt in Pt-containing Ag2S-metal
nanocomposites is 84.5 m2 g1, which is 16 % higher than that of the commercial
Pt/C (72.9 m2 g1) due to the smaller Pt domain size of the former (ca. 1 nm) [213].
The other important feature of the nanocomposites is the electronic coupling
between the metal and semiconductor domains. The XPS analyses of the
Pt-containing nanocomposites demonstrate that electrons are transferred to Pt
from other domains of the nanocomposites due to the intra-particle charge transfer
(see Fig. 3.34 for the energy level diagram). Analogous charge transfer has been
observed in the Au@PbS system, whereby the electrons transfer from PbS shell to
the inner Au core resulted in the n-type to p-type change in hydrazine-treated PbS
[260]. The electron transfer from Ag2S to Pt could also be described with the
generation of a hole in the Ag2S domain. In the presence of Au domain (work
function ¼ 5.1 eV) [261], the alignment of energy levels in Au and Ag2S would be
favorable for electron transfer from Au to Ag2S to fill the hole generated by the

Fig. 3.34 Energy level


diagram for Ag2S-noble
metal nanocomposites
predicts intraparticle charge
transfer among different
domains. Reproduced from -3.63 eV Ec
[213] with the permission
from Wiley-VCH
Au -5.32 eV Ev
-5.10 eV Ag2S Pt
-5.65 eV
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 145

electron transfer to Pt domains, further promoting the electron transfer from Ag2S
to Pt in order for the Fermi levels to match at the interface. The electron transfer
from other domains to Pt in Pt-containing nanocomposites increases the electron
density around the Pt sites, causing the weakening of CO chemisorption and hence
the increase their electrochemical activity for MOR, a key reaction in direct
methanol fuel cell [213].
On the other hand, the Pt-containing Ag2S-noble metal nanocomposites exhibit
poor activity for ORR, another key reaction in DMFC [9, 11, 262, 263]. This is not
surprising based on a report by Watanabe and co-workers [264], which discussed
alloying of Fe, Ni or Co with Pt to enhance the electro-catalytic activity for oxygen
reduction. Fe, Ni or Co has more 5d vacancies than Pt and could withdraw electrons
from the latter. This withdrawing effect induced an increase in 5d vacancies in Pt,
increasing 2π electron donation from O2 to the Pt surface, and resulting in enhanced
O2 adsorption to favor oxygen reduction. However, for the Pt-containing Ag2S-
noble metal nanocomposites, the electron donation from the semiconductor to the
Pt domains would decrease the 5d vacancies in Pt. Thus, the adsorption of O2 on the
nanocomposites would be too weak for the O2 dissociation reaction [62, 63, 197,
198]. This would account for the poor activity of the Pt-containing nanocomposites
towards oxygen reduction.

3.6.4 Cadmium Selenide-Platinum Nanocomposites


with a Core-Shell Construction

The role of surface strain in catalysis, particularly as a means of tuning the catalytic
activity, has attracted great interest in recent years [265–267]. When a metal is
deposited on a substrate with different lattice parameters, a compressive or tensile
strain in the surface of the metal layer usually occurs to fulfill the requirement of
epitaxial growth, and often affects the overlap of the electron orbitals between the
metal atoms and therefore changes the electronic properties of the surface and its
reactivity, which will influence the bond strength of an adsorbate [268]. Therefore,
the application of compressive strain or tensile stain to a surface can be an effective
means to influence the surface reactivity [269]. Strain and the associated shift of the
d-band can be brought about by growing the desired metal on other materials with a
different lattice constant. The over-layer may thereby be strained or compressed
depending on the lattice mismatch between the two materials. The lateral strain has
been studied in a number of pseudomorphic metal monolayers formed on electrode
substrates [268, 270–273], or core-shell nanoparticles constructed by different
metals [274–276]. However, the investigations of lateral strain in semiconductor-
metal nanocomposites, a type of nanostructure that combine materials with dis-
tinctly different physical and chemical properties, are scarcely reported up to date.
Analogous to the deposition of Pd on Ag2S nanocrystals [213], the Yang group
found that in aqueous media, numerous sites on CdSe nanocrystals are also
146 H. Liu et al.

(a) (b) (c)


250
Cd
Se
200 Pt

Intensity (a.u.)
150

100

50

0
7 14 21 28 35
Position (nm)
(d) (e) (f)

Fig. 3.35 Core-shell CdSe-Pt nanocomposites synthesized in aqueous phase using 10 nm CdSe
cores: (a) TEM image, (b) STEM image, (c) element profile, and (d) HRTEM image of core-shell
CdSe-Pt nanocomposites at CdSe/Pt molar ratio of 1/1; (e, f) HRTEM images of core-shell
CdSe@Pt nanocomposites at CdSe/Pt molar ratio of 2/1 and 1/2, respectively. Reproduced from
[214] with permission from the Royal Society of Chemistry

provided for the nucleation of Pt nanoparticles, which then grow and eventually
coalesce to form a continuous shell on each CdSe nanocrystal, as displayed by the
TEM, HRTEM and STEM in Fig. 3.35 [214]. The thickness of the Pt shell could be
controlled by varying the CdSe/Pt molar ratio in the synthesis. Figures 3.35d, e, f
illustrate the core-shell CdSe@Pt nanocomposites synthesized at CdSe/Pt molar
ratios of 1/1, 2/1, and 1/2, respectively. The thickness of Pt shell could be varied as
shown by comparing those HRTEM images.
The XRD analyses of the core-shell CdSe@Pt nanocomposites at different
CdSe/Pt molar ratios suggest that the interplanar spacing of Pt is compressed to
match the lattice plane of CdSe for its epitaxial growth on the CdSe substrates. The
compression of Pt lattice spacing can also be verified by the HRTEM images in
Fig. 3.35. The digital analyses of images show that the spacing of Pt(1 1 1)
decreases from 0.2244 nm to 0.2148 nm with the decrease of Pt molar ratio in the
core-shell nanocomposites. The compressive strain effect of the CdSe core on the
deposited Pt shell might be a welcome feature for the core-shell composite
nanomaterials in a given catalytic reaction.
In a specific CdSe/Pt molar ratio, e.g. 1/1, the core-shell CdSe@Pt
nanocomposites display superior catalytic activity toward oxygen reduction reac-
tion (ORR) [214]. ORR is the key reaction at the cathode of direct methanol fuel
cell (DMFC), and the low activity of the cathode catalysts at room temperature is
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 147

one of the significant challenges need to be resolved before the commercialization


of DMFC [62, 155, 277]. In order to overcome this problem, it is necessary to
maximize the activity of Pt-based catalysts by engineering their structure, mor-
phology and/or composition.
The observed enhanced catalytic activity for ORR of core-shell CdSe@Pt
nanocomposites could be resulted from the reasonably compressive strain effect
of the CdSe core on the Pt shell. As discussed in Sect. 3.5, the common ORR
process involves both the breaking of an O–O bond and the formation of O–H bonds
[9, 62, 63]. The most active Pt-based catalyst should have the d-band center with an
intermediate value since the optimal ORR catalyst needs to facilitate both bond-
breaking and bond-making steps without hindering one or the other [197, 198]. The
surface of a Pt-based catalyst with high value of d-band center tends to bind
adsorbents more strongly, thereby enhancing the kinetics of dissociation reactions
producing these adsorbents, whereas a surface with lower d-band center value tends
to bind adsorbents more weakly and facilitates the formation of bonds amongst
them. The compressive stains of CdSe core on Pt shell leading to broader d-band
result in a down-shift of the d-band center of the Pt shell. At an appropriate CdSe/Pt
molar ratio (1/1) in core-shell nanocomposites, the downward shift of the d-band
center sufficiently balances the bond-breaking and bond-making steps of the ORR
process, thus offering optimum catalytic activity.
It is worthy to note that the core-shell CdSe@Pt nanocomposites also exhibit
excellent catalytic activity toward MOR at CdSe/Pt molar ratio of 1/2 [214]. The
enhanced MOR catalytic activity of the core-shell CdSe@Pt nanocomposites could
also be attributed to the reasonably compressive strain effect of the CdSe core on
the deposited Pt shell, which results in the appropriate shift of the d-band center of
Pt shell and improvement of the adsorption of methanol on the surface of Pt
catalysts. Interestingly, the CdSe/Pt molar ratio for core-shell CdSe@Pt
nanocomposites offering highest activity for MOR is different from that for ORR.
This is a welcome feature for these core-shell nanocomposites to be used in DMFC.
Comparing with the DMFC using Pt/C catalysts both at anode and cathode, the
inhibition of methanol crossover in certain level could be expected when the core-
shell nanocomposites with optimal CdSe/Pt molar ratio for ORR are used at
cathode. This is important for the commercialization of DMFC since crossover of
methanol from the anode to the cathode through the polymer electrolyte membrane
(PEM) interferes with oxygen reduction at the cathode, resulting in the creation of a
mixed potential and drastically decrease in DMFC performance [122–124].

3.6.5 Nanocomposites of Silver Sulfide and Noble Metals


with Controlled Nanostructure

The interest in nanostructured composites is the possibility of integrating materials


with vastly different physical and chemical properties into a hybrid nanosystem
148 H. Liu et al.

with multiplexing capability and tunable physical and chemical properties that may
not be obtainable otherwise [278, 279]. The interactions between the nanoscale
metal and semiconductor components, if synergistic, can lead to significant
improvements of application performance. On the other hand, noble metal
nanoparticles with controlled internal structures have been often used to modify
the noble metal properties in a diverse range of applications, such as catalysis [280–
282], nanoreactors [75], and drug delivery systems [76–78]. For example, Hyeon
and co-workers reported good catalytic activity for Pd nanoparticles with a hollow
interior in Suzuki cross-coupling reactions. The hollow Pd nanoparticle catalyst
could be reused seven times without the loss of catalytic activity [280].
Hence, the integration of semiconductor and noble metal nanoparticles into a
nanosystem with controlled architecture may lead to further improvements of the
application performance of noble metal-semiconductor hybrid materials. Upon the
inside-out diffusion of Ag in core-shell nanoparticles [119], the Yang group
developed a general approach to the fabrication of silver sulfide and noble metal
nanoparticles with an overall hollow or cage-bell structure [283]. In their strategy,
core-shell nanoparticles with Ag residing in the core or inner shell region are first
prepared as starting templates. Sulfur is then used to promote the inside-out
diffusion of Ag from the core or the inner shell region of the core-shell
nanoparticles, and to convert the diffused Ag into Ag2S. The overall result is the
conversion of the Ag-containing core-shell nanoparticles into silver sulfide-noble
metal nanoparticles with hollow or cage-bell structures.
The inside-out diffusion of Ag in core-shell nanoparticles has been developed
into a general protocol for the fabrication of noble metal nanoparticles with a
hollow or cage-bell structure [119]. The protocol begins with the synthesis of
core-shell Ag-M or core-shell-shell MA-Ag-MB nanoparticles in an organic solvent.
Ag is then removed from the core or from the inner shell by BSPP, which binds
strongly with Agþ ions to promote the inside-out diffusion process and to allow the
complete removal of Ag in 24~48 h, leaving behind an organosol of hollow or cage-
bell structured metal nanomaterials.
Subsequently, Yang and co-workers attempted to find a more cost-effective
alternative to produce hollow/cage-bell structured metal nanomaterials, and they
employed elemental sulfur (S) to replace expensive BSPP to remove the Ag
component from the core-shell particles. However, they experimentally found
that the inside-out diffusion of Ag in core-shell nanoparticles could be further
developed as a general protocol to fabricate nanocomposites of Ag2S and metal
nanoparticles with a hollow or cage-bell structure [283]. As illustrated by the
scheme in Fig. 3.36, after mixing core-shell nanoparticles and sulfur in toluene,
the Agþ ions released by the O2/Cl etching of the twinned Ag seeds diffuse out
through the discontinuous metal shell, and react with sulfur to form Ag2S
nanocrystals decorated on the outer shell of the metal nanoparticles (Fig. 3.36a).
The Ag2S nanocrystals grow with the continued outward diffusion of Agþ ions until
the Ag core is completely removed, leaving behind a colloidal solution of
nanocomposites consisting of Ag2S and metal nanoparticles with a hollow interior
(Fig. 3.36b). Finally, the Ag2S nanocrystals on the surface of the metal shell
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 149

Fig. 3.36 A schematic


illustration showing the
S
formation of
nanocomposites consisting (a)
of Ag2S and metal
nanoparticles with a hollow
interior based on the inside-
out diffusion of Agþ in Ag

(b)
S
Ag-containing core-shell Ag2S
nanoparticles. Reproduced
from [283] with permission Metal
from the Royal Society of
Chemistry
Ripening Process
(c)

underwent a ripening process, e.g. Ostwald ripening or electrochemical ripening


[181, 235], to form larger and more stable domains on the surface of the metal shell
(Fig. 3.36c).
With a slight modification, the protocol above could be used to prepare
nanocomposites of Ag2S and noble metal nanoparticles with a cage-bell structure.
As a typical example, the core-shell-shell Au-Ag-Pt nanoparticles are prepared first,
and after treatment with elemental sulfur, the internal Ag shell is removed from the
core-shell-shell Au-Ag-Pt nanoparticles by the aforementioned inside-out diffusion
of Agþ ions. Reaction of Agþ ions with sulfur results in the formation of Ag2S
patches on the surface of the now cage-bell Au-Pt nanoparticles, finalizing the
fabrication of nanocomposites of Ag2S and Au-Pt nanoparticles with a cage-bell
structure. The morphologies of the starting core-shell-shell Au-Ag-Pt nanoparticles
and final nanocomposite were shown by the TEM images in Fig. 3.37a, b. The cage-
bell structure of the metallic domains is revealed by the appearance of void space
between the core and the outer shell regions. The void space is formed by the
removal of the Ag inner shell by elemental sulfur. The Ag2S patches on the surface
of cage-bell metal domains display a stronger contrast than the void space, as can be
seen in the TEM image in Fig. 3.37b. The HRTEM image in Fig. 3.37c of a single
composite Ag2S-cage-bell Au-Pt nanoparticle shows three different sets of lattice
fringes with separations of 0.26 nm in the light patches, 0.24 nm in the core, and
0.23 nm in the shell regions. These fringes correspond well with the (121) planes of
monoclinic Ag2S, and (1 1 1) planes of face-centered cubic Au and Pt, respectively.
The TEM characterizations confirm that the size and morphology of the remaining
metallic domains in the nanocomposites are virtually unchanged after treatment
with sulfur, suggesting that the diffusion of the Ag from the inner shell of the core-
shell-shell nanoparticles does not cause the collapse of the particle geometry.
The strategy for the fabrication of nanocomposites of Ag2S and hollow struc-
tured noble metal nanoparticles could be further tuned to produce ternary
nanocomposites consisting of Ag2S, Au, and Pt nanoparticles with hollow interiors,
150 H. Liu et al.

(a) (c)

20 nm 5 nm
1.0
(b) (d)
0.8
Absorbance / a.u.

0.6

0.4

0.2

Core-shell-shell Au-Ag-Pt
Ag2S-cage-bell Au-Pt
20 nm 0.0
300 400 500 600 700
Wavelength / nm

Fig. 3.37 TEM images of core-shell-shell Au-Ag-Pt (a), Ag2S-cage-bell Au-Pt (b); HRTEM
images of nanocomposites of Ag2S-cage-bell Au-Pt (c); UV-visible spectra of core-shell-shell
Au-Ag-Pt nanoparticles before and after treatment with element sulfur (d). Reproduced from [283]
with permission from the Royal Society of Chemistry

labeled as Ag2S-Au-hPt [284]. In brief, the gold precursors (HAuCl4) are firstly
transferred from aqueous to toluene using the ethanol-mediated transfer protocol
[71], and then mixed with Ag2S-hollow Pt (Ag2S-hPt) nanocomposite organosol in
toluene. Ternary Ag2S-Au-hPt nanocomposites are found as the dominant product
after aging the mixture of Ag2S-hPt hetero-dimers and HAuCl4 in toluene for 2 h, as
indicated by the TEM and HRTEM images in Fig. 3.38. Isolated Au nanoparticles
are not observed, indicating that Au nucleates preferentially on the existing Ag2S-
hPt heterodimers under the experimental conditions. In most cases, Au is deposited
only at a single site on the Ag2S domain in each Ag2S-hPt hetero-dimers. The
average diameter of the deposited gold patches is ca. 8.3 nm, which could be
discernible by the strong brightness contrast in TEM and HRTEM images.
The direct evidences for the formation of ternary Ag2S-Au-hPt nanocomposites
are provided by the line scanning analysis and elemental mapping of an arbitrary
single composite nanoparticle (Fig. 3.38c) in the high-angle annular dark-field
STEM mode. As shown in Fig. 3.38d for the line-scanning analysis, the Au and
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 151

(a) (b) (e)

Ag
(f)
20 nm 2 nm

(c)
S
(g)

Au
(h)
50
(d) S
40
Ag
Intensity (a.u.)

Pt
30
Au
Pt
20 (i)
10

-10
0 10 20 30 40 50
Position (nm)

Fig. 3.38 TEM image (a), HRTEM image (b), STEM image (c), line-scan analysis (d), and
elemental mapping (e–i) of a single particle (c) of the ternary Ag2S-Au-hPt nanocomposites
as-prepared by depositing Au on the Ag2S-hPt heterodimers in toluene at room temperature.
Reproduced from [284] with the permission from the Nature Publishing Group

Pt signals are present at left- and right-hand side, respectively, whereas the Ag and
S signals are only concentrated at the core region. The line-scanning analysis is well
in accord with the nanoscale mapping results (Fig. 3.38e–i), which also manifest
that the Au and Pt components are respectively distributed on the two sides of the
ternary nanocomposites. Further, HRTEM image revealed that the crystal planes of
Au are not parallel to those of Ag2S in each heterodimerNafion membrane nano-
particle (Fig. 3.38b), indicating that the growth of Au on the surface of Ag2S
domain takes place in different orientations.
Analogous to the Ag2S-Au-Pt ternary nanocomposites discussed previously
[213], the ternary Ag2S-Au-hPt nanocomposites also exhibit enhanced
electrocatalytic property toward MOR due to the strong electronic coupling
between Pt and other domains in the hybrid particles [284]. The study by Yang
152 H. Liu et al.

and co-workers offers a vivid example to exhibit the enhancement of the material
properties by means of a structural tailoring, and the concept might be used toward
the design and synthesis of other hetero-nanostructures for catalytic applications
other than methanol oxidation.

3.6.6 Pt-Based Selective Catalysts for a Propotype


of Membraneless DMFCs

In current DMFCs, the anode and cathode electrocatalysts are commonly based on
Pt [3–6, 285, 286]. These catalysts are not selective to MOR at anode or ORR at
cathode, and hence any methanol crossover from the anode to the cathode through
the proton exchange membrane can be oxidized by the cathode catalyst. This results
in the creation of a mixed potential at the cathode, which degrades the fuel cell
performance [8, 287, 288]. Although a number of efforts have been devoted toward
the modification on the proton exchange membrane to overcome this key obstacle
for the commercialization of DMFCs, it is generally thought that the commonly
used Nafion membrane has an unacceptably high rate of methanol crossover
[289–295]. In this sense, synthesis of electrocatalysts with high selectivity for
MOR and ORR represents an alternative for solving this problem in DMFCs.
Upon the deep understanding of the mechanisms of the electrocatalytic reac-
tions, the Yang group reported the design and fabrication of Pt-based nanomaterials
with enhanced catalytic activity and high selectivity towards DMFC reactions by
sufficiently making use of the structural uniqueness and electronic coupling effects
among the different domains of the electrocatalysts so that the DMFCs can be
operated well without or with little dependence on the proton exchange membrane
[296]. In particular, the ternary Au@Ag2S-Pt nanocomposites display superior
MOR selectivity due to the electronic coupling effect among different domains of
the nanocomposites (Fig. 3.39a), while the cage-bell structured Pt-Ru nanoparticles
exhibit excellent methanol tolerance for ORR at the cathode because of the differ-
ential diffusion of methanol and oxygen in the porous Ru shell of the cage-bell
nanoparticles (Fig. 3.39b). The good catalytic selectivity of these Pt-based
nanomaterials via structural construction enables a DMFC to be built without a
proton exchange membrane between the fuel electrode and the oxygen electrode
(Fig. 3.39c).
The performance of the membraneless DMFC has been evaluated in term of the
open circuit voltage (OCV), the current-voltage (I–V ), and the current-power (I–P)
characteristics under ambient conditions [296]. The membraneless DMFC with the
selective MOR catalyst at the anode and ORR catalyst at the cathode could maintain
an open-circuit voltage of ca. 0.38 V for more than 120 min. It is of great
importance that reasonable power with the maximum value of ca.15 μW is obtained
without a separate membrane. The OCV of the membraneless DMFC is still rather
low, at only ca. 32 % of the theoretical cell voltage of DMFC (ca. 1.18 V)
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 153

(a) Pt
Energy level diagram

Pt
Pt
Au
-3.63 eV Ev
Pt -5.10 eV
Pt Au -5.32 eV Ec
-5.65 eV
Pt
Ag2S Pt
Pt

(b) Porous Ru shell

Methanol

Catalytic
Ptcore

Oxygen
Products
Reactants

Methanol molecule Oxygen molecule

(c)

O2 O2

Electrolyte

Au@Ag2S-Pt (Anode) CBS Pt-Ru(Cathode)

Fig. 3.39 Schematic illustration of the construction of ternary Au@Ag2S-Pt nanocomposites and
the energy level alignment in Ag2S-Au-Pt nanocomposites (a), schematic illustration of the
differential diffusion and reaction of methanol and oxygen in CBS Pt-Ru nanoparticles (b), and
schematic illustration for the membraneless DMFC (c). Reproduced from [296] with the permis-
sion from the Nature Publishing Group
154 H. Liu et al.

[297]. The performance of the prototype could be limited by the use of dissolved
oxygen as the source of oxygen. Despite the limited performance, the prototype
does demonstrate the viability of using selective MOR and ORR catalysts to
construct a DMFC without the proton exchange membrane. Since the exemption
of a proton exchange membrane not only reduces the cost and volume of the
DMFC, it also provides more flexibility and miniaturizability of the cell design.
Its implementation is also simpler than membraneless fuel cells that depend on
non-mixing laminar flows [298, 299] or 3D anode [300–302]. The performance of
the membraneless fuel cell may be improved by using the on-chip fuel cell
developed by the Osaka group, which can directly make use of the oxygen from
the air [303].

3.7 Conclusion Remarks and Perspectives

The efforts of many leading research groups have led to a rich variety of heteroge-
neous nanomaterials, and their accumulation creates great opportunities also a
tremendous challenge to apply these materials in various areas, e.g. energy conver-
sion and environmental remediation. In addition to structural advantages, the
welcome features for a heterogeneous nanoparticle include the electronic coupling
effect and lateral strain effect among different domains of the heterogeneous
materials. Future research challenges in heterogeneous nanomaterials may include:
(1) develop general strategies to prepare heterogeneous nanomaterials with the
control on the size and morphology of each domain in the heterogeneous particles;
(2) to understand the synergetic effect and interface boundary sites in heteroge-
neous nanosystems, which are of importance to enhance the properties,
e.g. catalysis, of the heterogeneous nanomaterials. The mechanistic understanding
of the underlying chemistry for the heterogeneous systems might be valuable for the
development of more metal-based heterogeneous nanomaterials with interesting
architectures and tailored functionalities; (3) explore the catalytic activity of the
heterogeneous nanomaterials for energy conversion. The as-synthesized noble
metal-based heterogeneous nanomaterials may possess superior properties for
high efficient energy conversion due to its ultrafine size, high stability, ideal
morphology, and welcome energy level alignment, some of which are difficult or
impossible to achieve by commercial catalysts. The application of these
nanomaterials toward DMFC and photocatalysis would be hot topics accordingly.
The specific goals include the evaluation of the heterogeneous nanomaterials as
catalysts for energy conversion, the devices designed with improved performance,
efficiency, durability and reduced costs, and the realization of the commercializa-
tion of the catalysts as a long-term objective; (4) Explore other scientific-related
issues. Many interesting scientific findings might be derived from the synthesis and
characterization of the heterogeneous nanomaterials, which would not only satisfy
everlasting human curiosity, but also promise new advances in nanoscience and
nanotechnology.
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 155

Acknowledgements Financial support from the 100 Talents Program of the Chinese Academy of
Sciences, National Natural Science Foundation of China (Grant Nos: 21173226, 21376247, and
21476246), and State Key Laboratory of Multiphase Complex Systems, Institute of Process
Engineering, Chinese Academy of Sciences (MPCS-2012-A-11) is gratefully acknowledged.

References

1. Antolini E (2003) Formation of carbon-supported PtM alloys for low temperature fuel cells: a
review. Mater Chem Phys 78:563–573
2. Liu H, Song C, Zhang L, Zhang J, Wang H, Wilkinson DP (2006) A review of anode catalysis
in the direct methanol fuel cell. J Power Sources 155:95–110
3. Steele BCH, Heinzel A (2001) Materials for fuel-cell technologies. Nature 414:345–352
4. Perry ML, Fuller TF (2002) A historical perspective of fuel cell technology in the 20th
century. J Electrochem Soc 149:S59–S67
5. Chen J, Lim B, Lee EP, Xia Y (2009) Shape controlled synthesis of platinum nanocrystals for
catalytic and electrocatalytic applications. Nano Today 4:81–95
6. Peng Z, Yang H (2009) Designer platinum nanoparticles: control of shape, composition in
alloy, nanostructure and electrocatalytic property. Nano Today 4:143–164
7. Parsons R, VanderNoot T (1988) The oxidation of small organic molecules: a survey of
recent fuel cell related research. J Electroanal Chem 257:9–45
8. Antolini E, Lopes T, Gonzalez ER (2008) An overview of platinum-based catalysts as
methanol-resistant oxygen reduction materials for direct methanol fuel cells. J Alloys
Compd 461:253–262
9. Marković NM, Ross PN (2002) Surface science studies of model fuel cells electrocatalysts.
Surf Sci Rep 45:117–229
10. de Brujin FA, Dam VAT, Janssen GJM (2008) Review: durability and degradation issues of
PEM fuel cell components. Fuel Cells 8:3–22
11. Chen W, Kim J, Sun S, Chen S (2008) Electrocatalytic reduction of oxygen by FePt alloy
nanoparticles. J Phys Chem C 112:3891–3898
12. Wang C, Daimon H, Onodera T, Koda T, Sun S (2008) A general approach to the size- and
shape-controlled synthesis of platinum nanoparticles and their catalytic reduction of oxygen.
Angew Chem Int Ed 47:3588–3591
13. Kang Y, Murray CB (2010) Synthesis and electrocatalytic properties of cubic Mn-Pt
nanocrystals (nanocubes). J Am Chem Soc 132:7568–7569
14. Chen A, Holt-Hindle P (2010) Platinum-based nanostructured materials: synthesis, proper-
ties, and applications. Chem Rev 110:3767–3804
15. Cozzoli PD, Pellegrino T, Manna L (2006) Synthesis, properties and perspectives of hybrid
nanocrystal structures. Chem Soc Rev 35:1195–1208
16. Zhang J, Sasaki K, Adzic RR (2007) Stabilization of platinum oxygen-reduction
electrocatalysts using gold clusters. Science 315:220–222
17. Zhang L-F, Zhong S-L, Xu A-W (2013) Highly branched concave Au/Pd bimetallic
nanocrystals with superior electrocatalytic activity and highly efficient SERS enhancement.
Angew Chem Int Ed 52:645–649
18. Murray CB, Kagan CR, Bawendi MG (2000) Synthesis and characterization of monodisperse
nanocrystals and close-packed nanocrystal assemblies. Annu Rev Mater Sci 30:545–610
19. Daniel M-C, Astruc D (2004) Gold nanoparticles: assembly, supramolecular chemistry,
quantum-size-related properties, and applications toward biology, catalysis, and nanotech-
nology. Chem Rev 104:293–346
20. Bruda C, Chen X, Narayanan R, El-Sayed MA (2005) Chemistry and properties of
nanocrystals of different shapes. Chem Rev 105:1025–1102
156 H. Liu et al.

21. Ferrando R, Jellinek J, Johnston RL (2008) Nanoalloys: from theory to applications of alloy
clusters and nanoparticles. Chem Rev 108:845–910
22. Laurent S, Forge D, Port M, Roch A, Robic C, Elst LV, Muller RN (2008) Magnetic iron
oxide nanoparticles: synthesis, stabilization, vectorization, physicochemical characteriza-
tions, and biological applications. Chem Rev 108:2064–2110
23. Murray RW (2008) Nanoelectrochemistry: metal nanoparticles, nanoelectrodes, and
nanopores. Chem Rev 108:2688–2720
24. Xia Y, Xiong Y, Lim B, Skrabalak SE (2009) Shape-controlled synthesis of metal
nanocrystals: simple chemistry meets complex physics? Angew Chem Int Ed 48:60–103
25. Novet T, Johnson DC (1991) New synthetic approach to extended solids: selective synthesis
of iron suicides via the amorphous state. J Am Chem Soc 113:3398–3403
26. Suryanarayana C (2001) Mechanical alloying and milling. Prog Mater Sci 46:1–184
27. Mazumder V, Chi M, More KL, Sun S (2010) Core/shell Pd/FePt nanoparticles as an active
and durable catalyst for the oxygen reduction reaction. J Am Chem Soc 132:7848–7849
28. Yang H (2011) Platinum-based electrocatalysts with core–shell nanostructures. Angew Chem
Int Ed 50:2674–2676
29. Wang A, Peng Q, Li Y (2011) Rod-shaped Au–Pd core–shell nanostructures. Chem Mater
23:3217–3222
30. Weber MJ, Mackus AJM, Verheijen MA, van der Marel C, Kessels WMM (2012) Supported
core/shell bimetallic nanoparticles synthesis by atomic layer deposition. Chem Mater
24:2973–2977
31. Cardinal MF, Mongin D, Crut A, Maioli P, Rodrı́guez-González B, Pérez-Juste J, Liz-Marzán
LM, Fatti ND, Vallée F (2012) Acoustic vibrations in bimetallic Au@Pd core–shell
nanorods. J Phys Chem Lett 3:613–619
32. Chen T-Y, Luo T-JM, Yang Y-W, Wei Y-C, Wang K-W, Lin T-L, Wen T-C, Lee CH (2012)
Core dominated surface activity of core–shell nanocatalysts on methanol electrooxidation. J
Phys Chem C 116:16969–16978
33. Kuttiyiel KA, Sasaki K, Choi Y, Su D, Liu P, Adzic RR (2012) Nitride stabilized PtNi core–
shell nanocatalyst for high oxygen reduction activity. Nano Lett 12:6266–6271
34. Zhang Q, Lee I, Joo JB, Zaera F, Yin Y (2013) Core–shell nanostructured catalysts. Acc
Chem Res 46:1816–1824
35. Kang SW, Lee YW, Park Y, Choi B-S, Hong JW, Park K-H, Han SW (2013) One-pot
synthesis of trimetallic Au@PdPt core–shell nanoparticles with high catalytic performance.
ACS Nano 7:7945–7955
36. Sun X, Li D, Ding Y, Zhu W, Guo S, Wang ZL, Sun S (2014) Core/shell Au/CuPt
nanoparticles and their dual electrocatalysis for both reduction and oxidation reactions. J
Am Chem Soc 136:5745–5749
37. Tedsree K, Li T, Jones S, Chan CWA, Yu KMK, Bagot PAJ, Marquis EA, Smith GDW,
Tsang SCE (2011) Hydrogen production from formic acid decomposition at room tempera-
ture using a Ag–Pd core–shell nanocatalyst. Nat Nanotechnol 6:302–307
38. Alayoglu S, Nilekar AU, Mavrikakis M, Eichhorn B (2008) Ru–Pt core–shell nanoparticles
for preferential oxidation of carbon monoxide in hydrogen. Nat Mater 7:333–338
39. Serpell CJ, Cookson J, Ozkaya D, Beer PD (2011) Core@shell bimetallic nanoparticle
synthesis via anion coordination. Nat Chem 3:478–483
40. Turkevich J, Kim G (1970) Palladium: preparation and catalytic properties of particles of
uniform size. Science 169:873–879
41. Paz-Borbon LO, Mortimer-Jones TV, Johnston RL, Posada-Amarillas A, Barcaro G,
Fortunelli A (2007) Structures and energetics of 98 atom Pd–Pt nanoalloys: potential stability
of the Leary tetrahedron for bimetallic nanoparticles. Phys Chem Chem Phys 9:5202–5208
42. Wang C, van der Vliet D, More KL, Zaluzec NJ, Peng S, Sun S, Daimon H, Wang G,
Greeley J, Pearson J, Paulikas AP, Karapetrov G, Strmcnik D, Markovic NM, Stamenkovic
VR (2011) Multimetallic Au/FePt3 nanoparticles as highly durable electrocatalyst. Nano Lett
11:919–926
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 157

43. Narayanan R, El-Sayed MA (2004) Shape-dependent catalytic activity of platinum


nanoparticles in colloidal solution. Nano Lett 4:1343–1348
44. Habas SE, Lee H, Radmilovic V, Somorjai GA, Yang P (2007) Shaping binary metal
nanocrystals through epitaxial seeded growth. Nat Mater 6:692–697
45. Wang C, Daimon H, Lee Y, Kim J, Sun S (2007) Synthesis of monodisperse Pt nanocubes and
their enhanced catalysis for oxygen reduction. J Am Chem Soc 129:6974–6975
46. Narayanan R, El-Sayed MA (2004) Changing catalytic activity during colloidal platinum
nanocatalysis due to shape changes: Electron-transfer reaction. J Am Chem Soc
126:7194–7195
47. Tian N, Zhou Z, Sun S, Ding Y, Wang ZL (2007) Synthesis of tetrahexahedral platinum
nanocrystals with high-index facets and high electro-oxidation activity. Science 316:732–735
48. Li C, Sato R, Kanehara M, Zeng HB, Bando Y, Teranishi T (2009) Controllable polyol
synthesis of uniform palladium icosahedra: effect of twinned structure on deformation of
crystalline lattices. Angew Chem Int Ed 48:6883–6887
49. Wang JX, Inada H, Wu L, Zhu Y, Choi Y, Liu P, Zhou WP, Adzic RR (2009) Oxygen
reduction on well-defined core-shell nanocatalysts: particle size, facet, and Pt shell thickness
effects. J Am Chem Soc 131:17298–17302
50. Tang Y, Ouyang M (2007) Tailoring properties and functionalities of metal nanoparticles
through crystallinity engineering. Nat Mater 6:754–759
51. Zhang Q, Xie J, Yang J, Lee JY (2009) Monodisperse icosahedral Ag, Au, and Pd
nanoparticles: size control strategy and superlattice formation. ACS Nano 3:139–148
52. Yang J, Yang J, Ying JY (2012) Morphology and lateral strain control of Pt nanoparticles via
core–shell construction using alloy AgPd core toward oxygen reduction reaction. ACS Nano
6:9373–9382
53. Zhang Q, Lee JY, Yang J, Boothroyd C, Zhang JX (2007) Size and composition tunable Ag–
Au alloy nanoparticles by replacement reactions. Nanotechnology 18:245605
54. Reyes-Gasga J, Tehuacanero-Nunez S, Montejano-Carrizales JM, Gao XX, Jose-Yacaman M
(2007) Analysis of the contrast in icosahedral gold nanoparticles. Top Catal 46:23–30
55. Ling T, Xie L, Zhu J, Yu HM, Ye HQ, Yu R, Cheng Z, Liu L, Yang GW, Cheng ZD, Wang
YJ, Ma XL (2009) Icosahedral face-centered cubic Fe nanoparticles: facile synthesis and
characterization with aberration-corrected TEM. Nano Lett 9:1572–1576
56. Gasteiger HA, Kocha SS, Sompalli B, Wagner FT (2005) Activity benchmarks and require-
ments for Pt, Pt-alloy, and non-Pt oxygen reduction catalysts for PEMFCs. Appl Catal B
56:9–35
57. Markovic N, Gasteiger H, Ross PN (1997) Kinetics of oxygen reduction on Pt(hkl) elec-
trodes: implications for the crystallite size effect with supported Pt electrocatalysts. J
Electrochem Soc 144:1591–1597
58. Li HQ, Xin Q, Li WZ, Zhou ZH, Jiang LH, Yang SH, Sun GQ (2004) An improved
palladium-based DMFCs cathode catalyst. Chem Commun 23:2776–2777
59. Hayre RO, Cha SW, Colella W, Prinz FB (2006) Fuel cell fundamentals. Wiley, New York
60. Bligaard T, Norskov JK (2007) Ligand effects in heterogeneous catalysis and electrochem-
istry. Electrochim Acta 52:5512–5516
61. Xu Y, Ruba AV, Mavrikakis M (2004) Adsorption and dissociation of O2 on Pt-Co and Pt-Fe
alloys. J Am Chem Soc 126:4717–4725
62. Stamenkovic VR, Mun BS, Arenz M, Mayrhofer KJJ, Lucas CA, Wang GF, Ross PN,
Markovic NM (2007) Trends in electrocatalysis on extended and nanoscale Pt-bimetallic
alloy surfaces. Nat Mater 6:241–247
63. Stamenkovic VR, Fowler B, Mun BS, Wang GF, Ross PN, Lucas CA, Markovic NM (2007)
Improved oxygen reduction activity on Pt3Ni(111) via increased surface site availability.
Science 315:493–497
64. Ataee-Esfahani H, Wang L, Nemoto Y, Yamauchi Y (2010) Synthesis of bimetallic Au@Pt
nanoparticles with Au core and nanostructured Pt shell toward highly active electrocatalysts.
Chem Mater 22:6310–6318
158 H. Liu et al.

65. Liang H, Zhang H, Hu J, Guo Y, Wan L, Bai C (2004) Pt hollow nanospheres: facile synthesis
and enhanced electrocatalysts. Angew Chem Int Ed 43:1540–1543
66. Yang J, Lee JY, Too H-P, Valiyaveettil S (2006) A bis( p-sulfonatophenyl)phenyl-phosphine-
based synthesis of hollow Pt nanospheres. J Phys Chem B 110:125–129
67. Chen G, Xia D, Nie Z, Wang Z, Wang L, Zhang L, Zhang J (2007) Facile synthesis of Co–Pt
hollow sphere electrocatalyst. Chem Mater 19:1840–1844
68. Chen HM, Liu RS, Lo MY, Chang SC, Tsai LD, Peng YM, Lee JF (2008) Hollow platinum
spheres with nano-channels: synthesis and enhanced catalysis for oxygen reduction. J Phys
Chem C 112:7522–7526
69. Guo S, Dong S, Wang E (2008) A general method for the rapid synthesis of hollow metallic or
bimetallic nanoelectrocatalysts with urchinlike morphology. Chem Eur J 14:4689–4695
70. Xu C, Wang L, Wang R, Wang K, Zhang Y, Tian F, Ding Y (2009) Nanotubular mesoporous
bimetallic nanostructures with enhanced electrocatalytic performance. Adv Mater
21:2165–2169
71. Yang J, Sargent EH, Kelley SO, Ying JY (2009) A general phase-transfer protocol for metal
ions and its application in nanocrystal synthesis. Nat Mater 8:683–689
72. Peng Z, Wu J, Yang H (2010) Synthesis and oxygen reduction electrocatalytic property of
platinum hollow and platinum-on-silver nanoparticles. Chem Mater 22:1098–1106
73. Lee J, Park JC, Song H (2008) A nanoreactor framework of a Au@SiO2 yolk/shell structure
for catalytic reduction of p-nitrophenol. Adv Mater 20:1523–1528
74. Gough DV, Wolosiuk A, Braun PV (2009) Mesoporous ZnS nanorattles: programmed size
selected access to encapsulated enzymes. Nano Lett 9:1994–1998
75. Liu J, Qiao SZ, Hartono SB, Lu GQ (2010) Monodisperse yolk-shell nanoparticles with a
hierarchical porous structure for delivery vehicles and nanoreactors. Angew Chem Int Ed
49:4981–4985
76. Gao J, Liang G, Zhang B, Kuang Y, Zhang X, Xu B (2007) FePt@CoS2 yolk-shell
nanocrystals as a potent agent to kill hela cells. J Am Chem Soc 129:1428–1433
77. Zhao W, Chen H, Li Y, Li L, Lang M, Shi J (2008) Uniform rattle-type hollow magnetic
mesoporous spheres as drug delivery carries and their sustained-release property. Adv Funct
Mater 18:2780–2788
78. Zhu Y, Ikoma T, Hanagata N, Kaskel S (2010) Rattle-type Fe3O4@SiO2 hollow mesoporous
spheres as carrier for drug delivery. Small 6:471–478
79. Zhang W-M, Hu J-S, Guo Y-G, Zheng S-F, Zhong L-S, Song W-G, Wan L-J (2008)
Tin-nanoparticles encapsulated in elastic hollow carbon spheres for high-performance
anode material in lithium-ion batteries. Adv Mater 20:1160–1165
80. Lou XW, Li CM, Archer LA (2009) Designed synthesis of coaxial SnO2@carbon hollow
nanospheres for highly reversible lithium storage. Adv Mater 21:2536–2539
81. Chen JS, Li CM, Zhou WW, Yan QY, Archer LA, Lou XW (2009) One-pot formation of
SnO2 hollow nanospheres and α-Fe2O3@SnO2 nanorattles with large void space and their
lithium storage properties. Nanoscale 1:280–285
82. Chem JS, Luan D, Li CM, Boey FYC, Qiao S, Lou XW (2010) TiO2 and SnO2@TiO2 hollow
spheres assembled from anatase TiO2 nanosheets with enhanced lithium storage properties.
Chem Commun 46:8252–8254
83. Chen JS, Wang Z, Dong XC, Chen P, Lou XW (2011) Graphene-wrapped TiO2 hollow
structures with enhanced lithium storage capabilities. Nanoscale 3:2158–2161
84. Li H, Bian Z, Zhu J, Zhang D, Li G, Huo Y, Li H, Lu Y (2007) Mesoporous titania spheres
with tunable chamber structure and enhanced photocatalytic activity. J Am Chem Soc
129:8406–8407
85. Benabid F, Couny F, Knight JC, Birks TA, Russell PJ (2005) Compact, stable and efficient
all-fibre gas cells using hollow-core phoyonic crystal fibres. Nature 434:488–491
86. Kamata K, Lu Y, Xia Y (2003) Synthesis and characterization of monodispersed core-shell
spherical colloids with movable cores. J Am Chem Soc 125:2384–2385
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 159

87. Bai F, Sun Z, Wu H, Haddad RE, Xiao X, Fan H (2011) Templated photocatalytic synthesis
of well-defined platinum hollow nanostructures with enhanced catalytic performance for
methanol oxidation. Nano Lett 11:3759–3762
88. Caruso F, Caruso RA, M€ ohwald H (1998) Nanoengineering of inorganic and hybrid hollow
spheres by colloidal templating. Science 282:1111–1114
89. Caruso F, Shi X, Caruso RA, Susha A (2001) Hollow titania spheres from layered precursor
deposition on sacrificial colloidal core particles. Adv Mater 13:740–744
90. Yang Z, Niu Z, Lu Y, Hu Z, Han CC (2003) Templated synthesis of inorganic hollow spheres
with a tunable cavity size onto core-shell gel particles. Angew Chem Int Ed 42:1943–1945
91. Lou XW, Yuan C, Archer LA (2007) Shell-by-shell synthesis of tin oxide hollow colloids
with nanoarchitectured walls: cavity size tuning and functionalization. Small 3:261–265
92. Wang Z, Luan D, Boey FYC, Lou XW (2011) Fast formation of SnO2 nanoboxes with
enhanced lithium storage capability. J Am Chem Soc 133:4738–4741
93. Lai X, Li J, Korgel BA, Dong Z, Li Z, Su F, Du J, Wang D (2011) General synthesis and
gas-sensing properties of multiple-shell metal oxide hollow microspheres. Angew Chem Int
Ed 50:2738–2741
94. Schmidt HT, Ostafin AE (2002) Liposome directed growth of calcium phosphate nanoshells.
Adv Mater 14:532–535
95. Hubert DHW, Jung M, German AL (2000) Vesicle templating. Adv Mater 12:1291–1294
96. Gao X, Zhang J, Zhang L (2002) Hollow sphere selenium nanoparticles: their in-vitro anti
hydroxyl radical effect. Adv Mater 14:290–293
97. Nakashima T, Kimizuka N (2003) Interfacial synthesis of hollow TiO2 microspheres in ionic
liquids. J Am Chem Soc 125:6386–6387
98. Qi L, Li J, Ma J (2002) Biomimetic morphogenesis of calcium carbonate in mixed solutions if
surfactants and double-hydrophilic block copolymers. Adv Mater 14:300–303
99. Zhang D, Qi L, Ma J, Cheng H (2002) Synthesis of submicrometer-sized hollow silver
spheres in mixed polymer-surfactant solutions. Adv Mater 14:1499–1502
100. Wong MS, Cha JN, Choi K-S, Deming TJ, Stucky GD (2002) Assembly of nanoparticles into
hollow spheres using block copolypeptides. Nano Lett 2:583–587
101. Lou XW, Archer LA, Yang Z (2008) Hollow micro-/nanostructures: synthesis and applica-
tions. Adv Mater 20:3987–4019
102. Zhao Y, Jiang L (2009) Hollow micro/nanomaterials with multilevel interior structures. Adv
Mater 21:3621–3638
103. Sun Y, Mayers BT, Xia Y (2002) Template-engaged replacement reaction: a one-step
approach to the large-scale synthesis of metal nanostructures with hollow interiors. Nano
Lett 2:481–485
104. Sun Y, Xia Y (2004) Mechanistic study on the replacement reaction between silver
nanostructures and chloroauric acid in aqueous medium. J Am Chem Soc 126:3892–3901
105. Chen J, Wiley B, McLellan J, Xiong Y, Li Z-Y, Xia Y (2005) Optical properties of Pd-Ag and
Pt-Ag nanoboxes synthesized via galvanic replacement reactions. Nano Lett 5:2058–2062
106. Chen M, Gao L (2006) Synthesis and characterization of Ag nanoshells by a facile sacrificial
template route through in situ replacement reaction. Inorg Chem 45:5145–5149
107. Skrabalak SE, Chen J, Sun Y, Lu X, Au L, Cobley CM, Xia Y (2008) Gold nanocages:
synthesis, properties, and applications. Acc Chem Res 41:1587–1595
108. Zhang H, Jin M, Wang J, Li W, Camargo PHC, Kim MJ, Yang D, Xie Z, Xia Y (2011)
Synthesis of Pd-Pt bimetallic nanocrystals with a concave structure through a bromide-
induced galvanic replacement reaction. J Am Chem Soc 133:6078–6089
109. Mohl M, Dobo D, Kukovecz A, Konya Z, Kordas K, Wei J, Vajtai R, Ajayan PM (2011)
Formation of CuPd and CuPt bimetallic nanotubes by galvanic replacement reaction. J Phys
Chem C 115:9403–9409
110. González E, Arbiol J, Puntes VF (2011) Carving at the nanoscale: sequential galvanic
exchange and Kirkendall growth at room temperature. Science 334:1377–1380
160 H. Liu et al.

111. Yin Y, Rioux RM, Erdonmez CK, Hughes S, Somorjai GA, Alivisatos PA (2004) Formation
of hollow nanocrystals through the nanoscale Kirkendall effect. Science 304:711–714
112. Zeng HC (2006) Synthetic architecture of interior space for inorganic nanostructures. J Mater
Chem 16:649–662
113. Zeng HC (2007) Ostwald ripening: a synthetic approach for hollow nanomaterials. Curr
Nanosci 3:177–181
114. Caruso F (2000) Hollow capsule processing through colloidal templating and self-assembly.
Chem Eur J 6:413–419
115. Macdonald JE, Sadan MB, Houben L, Popov I, Banin U (2010) Hybrid nanoscale inorganic
cages. Nat Mater 9:810–815
116. Chen C, Kang Y, Huo Z, Zhu Z, Huang W, Xin HL, Snyder JD, Li D, Herron JA,
Mavrikakis M, Chi M, More KL, Li Y, Markovic NM, Somorjai GA, Yang P, Stamenkovic
VR (2014) Highly crystalline multimetallic nanoframes with three-dimensional
electrocatalytic surfaces. Science 343:1339–1343
117. Wu X-J, Xu D (2009) Formation of yolk/SiO2 shell structure using surfactant mixtures as
template. J Am Chem Soc 131:2774–2775
118. Li G, Shi Q, Yuan SJ, Neoh KG, Kang ET, Yang X (2010) Alternating silica/polymer
multilayer hybrid microspheres templates for double-shelled polymer and inorganic hollow
microstructures. Chem Mater 22:1309–1317
119. Liu H, Qu J, Chen Y, Li J, Ye F, Lee JY, Yang J (2012) Hollow and cage-bell structured
nanomaterials of noble metals. J Am Chem Soc 134:11602–11610
120. Wiley B, Herricks T, Sun Y, Xia Y (2004) Polyol synthesis of silver nanoparticles: use of
chloride and oxygen to promote the formation of single-crystal, truncated cubes and tetrahe-
drons. Nano Lett 4:1733–1739
121. Tan Y-N, Yang J, Lee JY, Wang DIC (2007) Mechanistic study on the Bis( p-
sulfonatophenyl)-phenylphosphine synthesis of monometallic Pt hollow nanoboxes using
Ag*-Pt core-shell nanocubes as sacrificial templates. J Phys Chem C 111:14084–14090
122. Gasteiger HA, Markovic N, Ross PN, Cairns EJ (1993) Methanol electrooxidation on well-
characterized platinum-ruthenium bulk alloys. J Phys Chem 97:12020–12029
123. Wang JM, Brankovic SR, Zhu Y, Hanson JC, Adzic RR (2003) Kinetic characterization of
PtRu fuel cell anode catalysts made by spontaneous Pt deposition on Ru nanoparticles. J
Electrochem Soc 150:A1108–A1117
124. Antolini E, Salgado JRC, Gonzalez ER (2005) Carbon supported Pt75M25 (M ¼ Co, Ni) alloys
as anode and cathode electrocatalysts for direct methanol fuel cells. J Electroanal Chem
580:145–154
125. Liu H, Ye F, Yang J (2014) A universal and cost-effective approach to the synthesis of
carbon-supported noble metal nanoparticles with hollow interiors. Ind Eng Chem Res
53:5925–5931
126. Wang L, Yamauchi Y (2013) Metallic nanocages: synthesis of bimetallic Pt–Pd hollow
nanoparticles with dendritic shells by selective chemical etching. J Am Chem Soc
135:16762–16765
127. Bagotzky VS, Vassiliev YB, Khazova OA (1977) Generalized scheme of chemisorption,
electrooxidation and electroreduction of simple organic-compounds on platinum group
metals. J Electroanal Chem 81:229–238
128. Goodenough JB, Hamnett A, Kennedy BJ, Manoharan R, Weeks SA (1990) Porous carbon
anodes for the direct methanol fuel-cell. 1. The role of the reduction method for carbon
supported platinum electrodes. Electrochim Acta 35:199–207
129. Jusys Z, Schmidt TJ, Dubau L, Lasch K, Jorissen L, Garche J, Behm RJ (2002) Activity of
PtRuMeox (Me ¼ W, Mo or V) catalysts towards methanol oxidation and their characteriza-
tion. J Power Sources 105:297–304
130. Park KW, Choi JH, Ahn KS, Sung YE (2004) PtRu alloy and PtRu-WO3 nanocomposite
electrodes for methanol electrooxidation fabricated by a sputtering deposition method. J Phys
Chem B 108:5989–5994
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 161

131. Park KW, Sung YE, Han S, Yan Y, Hyeon T (2004) Origin of the enhanced catalytic activity
of carbon nanocoil-supported PtRu alloy electrocatalysts. J Phys Chem B 108:939–944
132. Park KW, Choi JH, Lee SA, Pak C, Chang H, Sung YE (2004) PtRuRhNi nanoparticle
electrocatalyst for methanol electrooxidation in direct methanol fuel cell. J Catal
224:236–242
133. Huang J, Yang H, Huang Q, Tang Y, Lu T, Akins DL (2004) Methanol oxidation on carbon-
supported Pt-Os bimetallic nanoparticle electrocatalysts. J Electrochem Soc 151:A1810–
A1815
134. Ye F, Yang J, Hu W, Liu H, Liao S, Zeng J, Yang J (2012) Electrostatic interaction based
hollow Pt and Ru assemblies toward methanol oxidation. RSC Adv 2:7479–7486
135. Rolison DR, Hagans PL, Swider KE, Long JW (1999) Role of hydrous ruthenium oxide in Pt–
Ru direct methanol fuel cell anode electrocatalysts: the importance of mixed electron/proton
conductivity. Langmuir 15:774–779
136. Long JW, Stroud RM, Swider KE, Rolison DB (2000) How to make electrocatalysts more
active for direct methanol oxidation—avoid PtRu bimetallic alloys! J Phys Chem B
104:9772–9776
137. Villullas HM, Mattos-Costa FI, Bulhőes LOS (2004) Electrochemical oxidation of methanol
on Pt nanoparticles dispersed on RuO2. J Phys Chem B 108:12898–12903
138. Chen Z, Qiu X, Lu B, Zhang S, Zhu W, Chen L (2005) Synthesis of hydrous ruthenium oxide
supported platinum catalysts for direct methanol fuel cells. Electrochem Commun 7:593–596
139. Scheiba F, Scholz M, Cao L, Schafranek R, Roth C, Cremers C, Qiu X, Stimming U, Fuess H
(2006) On the suitability of hydrous ruthenium oxide supports to enhance intrinsic proton
conductivity in DMFC anodes. Fuel Cells 6:439–446
140. Cao L, Scheiba F, Roth C, Schweiger F, Cremers C, Stimming U, Fuess H, Chen L, Zhu W,
Qiu X (2006) Novel nanocomposite Pt/RuO2  xH2O/carbon nanotube catalysts for direct
methanol fuel cells. Angew Chem Int Ed 45:5315–5319
141. Profeti LPR, Simőes FC, Olivi P, Kokoh KB, Coutanceau C, Léger J-M, Lamy C (2006)
Application of Pt þ RuO2 catalysts prepared by thermal decomposition of polymeric pre-
cursors to DMFC. J Power Sources 158:1195–1201
142. Huang S-Y, Chang C-M, Wang K-W, Yeh C-T (2007) Promotion of platinum-ruthenium
catalyst for electrooxidation of methanol by crystalline ruthenium dioxide. ChemPhysChem
8:1774–1777
143. Peng F, Zhou C, Wang H, Yu H, Liang J, Yang J (2009) The role of RuO2 in the
electrocatalytic oxidation of methanol for direct methanol fuel cell. Catal Commun
10:533–537
144. Dubau L, Hahn F, Coutanceau C, Léger J-M, Lamy C (2003) On the structure effects of
bimetallic PtRu electrocatalysts towards methanol oxidation. J Electroanal Chem
554–555:407–415
145. Beard KD, Borrelli D, Cramer AM, Blom D, Van Zee JW, Monnier JR (2009) Preparation
and structural analysis of carbon-supported Co core/Pt shell electrocatalysts using electroless
deposition methods. ACS Nano 3:2841–2853
146. Langille MR, Zhang J, Mirkin CA (2011) Plasmon-mediated synthesis of heterometallic
nanorods and icosahedra. Angew Chem Int Ed 50:3543–3547
147. Zhou S, Mcllwrath K, Jackson G, Eichhorn B (2006) Enhanced CO tolerance for hydrogen
activation in Au-Pt dendritic heteroaggregate nanostructures. J Am Chem Soc
128:1780–1781
148. Lee H, Habas SE, Somorjai GA, Yang P (2008) Localized Pd overgrowth on cubic Pt
nanocrystals for enhanced electrocatalytic oxidation of formic acid. J Am Chem Soc
130:5406–5407
149. Peng Z, Yang H (2009) PtAu bimetallic heteronanostructures made by post-synthesis mod-
ification of Pt-on-Au nanoparticles. Nano Res 2:406–415
150. Lim B, Kobayashi H, Yu T, Wang J, Kim MJ, Li Z-Y, Rycenga M, Xia Y (2010) Synthesis of
Pd-Au bimetallic nanocrystals via controlled overgrowth. J Am Chem Soc 132:2506–2507
162 H. Liu et al.

151. Wang F, Li C, Sun L-D, Wu H, Ming T, Wang J, Yu JC, Yan C-H (2011) Heteroepitaxial
growth of high-index-faceted palladium nanoshells and their catalytic performance. J Am
Chem Soc 133:1106–1111
152. Fan F-R, Liu D-Y, Wu Y-F, Duan S, Xie Z-X, Jiang Z-Y, Tian Z-Q (2008) Epitaxial growth
of heterogeneous metal nanocrystals: From gold nano-octahedra to palladium and silver
nanocubes. J Am Chem Soc 130:6949–6951
153. Zeng J, Zhu C, Tao J, Jin M, Zhang H, Li Z-Y, Zhu Y, Xia Y (2012) Controlling the
nucleation and growth of silver on palladium nanocubes by manipulating the reaction
kinetics. Angew Chem Int Ed 51:2354–2358
154. Lim B, Jiang M, Camargo PHC, Cho EC, Tao J, Lu X, Zhu Y, Xia Y (2009) Pd-Pt bimetallic
nanodendrites with high activity for oxygen reduction. Science 324:1302–1305
155. Peng Z, Yang H (2009) Synthesis and oxygen reduction electrocatalytic property of Pt-on-Pd
bimetallic heteronanostructures. J Am Chem Soc 131:7542–7543
156. Guo S, Dong S, Wang E (2010) Three-dimensional Pt-on-Pd bimetallic nanodendrites
supported on graphene nanosheet: Facile synthesis and used as an advanced nanoelectro-
catalyst for methanol oxidation. ACS Nano 4:547–555
157. Wang L, Nemoto Y, Yamauchi Y (2011) Direct synthesis of spatially-controlled Pt-on-Pd
bimetallic nanodendrites with superior electrocatalytic activity. J Am Chem Soc
133:9674–9677
158. Kim Y, Hong JW, Lee YW, Kim M, Kim D, Yun WS, Han SW (2010) Synthesis of AuPt
heteronanostructures with enhanced electrocatalytic activity toward oxygen reduction.
Angew Chem Int Ed 49:10197–10201
159. Wang L, Yamauchi Y (2009) Block copolymer mediated synthesis of dendritic platinum
nanoparticles. J Am Chem Soc 131:9152–9153
160. Liu H, Ye F, Yao Q, Cao H, Xie J, Lee JY, Yang J (2014) Stellated Ag-Pt bimetallic
nanoparticles: an effective platform for catalytic activity tuning. Sci Rep 4:3969
161. Wang ZL (2000) Transmission electron microscopy of shape-controlled nanocrystals and
their assemblies. J Phys Chem B 104:1153–1175
162. Brodersen SH, Grønbjerg U, Hvolbæk B, Schiøtz J (2011) Understanding the catalytic
activity of gold nanoparticles through multi-scale simulations. J Catal 284:34–41
163. Feng Y, Liu H, Yang J (2014) Bimetallic nanodendrites via selective overgrowth of noble
metals on multiply twinned Au seeds. J Mater Chem A 2:6130–6137
164. Mourdikoudis S, Liz-Marzán LM (2013) Oleylamine in nanoparticle synthesis. Chem Mater
25:1465–1476
165. Brust M, Walker M, Bethell D, Schiffrin DJ, Whyman R (1994) Synthesis of thoil-derivatised
gold nanoparticle in a two-phase Liquid-Liqiud synthem. J Chem Soc Chem Commun
7:801–802
166. Hiramatsu H, Osterloh F (2004) A simple large-scale synthesis of nearly monodisperse gold
and silver nanoparticles with adjustable sizes and with exchangeable surfactants. Chem Mater
16:2509–2511
167. Aslam M, Fu L, Su M, Vijayamohanan K, Dravid VP (2004) Novel one-step synthesis of
amine-stabilized aqueous colloidal gold nanoparticles. J Mater Chem 14:1795–1797
168. Yu H, Chen M, Rice PM, Wang SX, White RL, Sun S (2005) Dumbbell-like bifunctional Au–
Fe3O4 nanoparticles. Nano Lett 5:379–382
169. Wang C, Hou Y, Kim J, Sun S (2007) A general strategy for synthesizing FePt nanowires and
nanorods. Angew Chem Int Ed 46:6333–6335
170. Peng S, Sun S (2007) Synthesis and characterization of monodisperse hollow Fe3O4
nanoparticles. Angew Chem Int Ed 46:4155–4158
171. Chaubey GS, Barcena C, Poudyal N, Rong C, Gao J, Sun S, Liu JP (2007) Synthesis and
stabilization of FeCo nanoparticles. J Am Chem Soc 129:7214–7215
172. Kim J, Rong C, Lee Y, Liu JP, Sun S (2008) From core/shell structured FePt/Fe3O4/MgO to
ferromagnetic FePt nanoparticles. Chem Mater 20:7242–7245
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 163

173. Zeng H, Sun S (2008) Syntheses, properties and potential applications of multicomponent
magnetic nanoparticles. Adv Funct Mater 18:391–400
174. Lu X, Tuan H-Y, Korgel BA, Xia Y (2008) Facile synthesis of gold nanoparticles with narrow
size distribution by using AuCl or AuBr as the precursor. Chem Eur J 14:1584–1591
175. Xu Z, Shen C, Hou Y, Gao H, Sun S (2009) Oleylamine as both reducing agent and stabilizer
in a facile synthesis of magnetite nanoparticles. Chem Mater 21:1778–1780
176. Kim J, Rong C, Liu P, Sun S (2009) Dispersible ferromagnetic FePt nanoparticles. Adv Mater
21:906–909
177. Wang C, Peng S, Lacroix L-M, Sun S (2009) Synthesis of high magnetic moment CoFe
nanoparticles via interfacial diffusion in core/shell structured Co/Fe nanoparticles. Nano Res
2:380–385
178. Lacroix L-M, Huls NF, Ho D, Sun X, Cheng K, Sun S (2011) Stable single-crystalline body
centered cubic Fe nanoparticles. Nano Lett 11:1641–1645
179. Liu X, Atwater M, Wang J, Dai Q, Zou J, Brennan JP, Huo Q (2007) A study on gold
nanoparticle synthesis using oleylamine as both reducing agent and protecting ligand. J
Nanosci Nanotechnol 7:3126–3133
180. Feng Y, Ma X, Han L, Peng Z, Yang J (2014) A universal approach to the synthesis of
nanodendrites of noble metals. Nanoscale 6:6173–6179
181. Ostwald WF (1897) Studien uber die bildung und umwandlung fester korper. Z Phys Chem
22:289–330
182. Liu RX, Iddir H, Fan QB, Hou GY, Bo AL, Ley KL, Smotkin ES, Sung YE, Kim H, Thomas
S (2000) Potential-dependent infrared absorption spectroscopy of adsorbed CO and X-ray
photoelectron spectroscopy of arc-melted single-phase Pt, PtRu, PtOs, PtRuOs, and Ru
electrodes. J Phys Chem B 104:3518–3531
183. Jiang J, Kucernak A (2009) Electrodeposition of highly alloyed quaternary PtPdRuOs
catalyst with highly ordered nanostructure. Electrochem Commun 11:1005–1008
184. Ley KL, Liu RX, Pu C, Fan QB, Leyarovska N, Segre C, Smotkin ES (1997) Methanol
oxidation on single-phase Pt-Ru-Os ternary alloys. J Electrochem Soc 144:1543–1548
185. Watanabe M, Motoo S (1975) Electrocatalysis by ad-atoms. 2. Enhancement of oxidation of
methanol on platinum by ruthenium ad-atoms. J Electroanal Chem 60:267–273
186. Watanabe M, Motoo S (1975) Electrocatalysis by ad-atoms. 3. Enhancement of oxidation of
carbon-monoxide on platinum by ruthenium ad-atoms. J Electroanal Chem 60:275–283
187. Pellegrino T, Fiore A, Carlino E, Giannini C, Cozzoli PD, Ciccarella G, Respaud M,
Palmirotta L, Cingolani R, Manna L (2006) Heterodimers based on CoPt3-Au nanocrystals
with tunable domain size. J Am Chem Soc 128:6690–6698
188. Choi J-S, Choi HJ, Jung DC, Lee J-H, Cheon J (2008) Nanoparticle assisted magnetic
resonance imaging of the early reversible stages of amyloid β self-assembly. Chem Commun
19:2197–2199
189. Wang C, Tian W, Ding Y, Ma Y-Q, Wang ZL, Markovic NM, Stamenkovic VR, Daimon H,
Sun S (2010) Rational synthesis of heterostructured nanoparticles with morphology control. J
Am Chem Soc 132:6524–6529
190. Bai L, Kuang Y, Luo J, Evans DG, Sun X (2012) Ligand-manipulated selective transforma-
tions of Au-Ni bimetallic heteronanostructures in an organic medium. Chem Commun
48:6963–6965
191. Lim SI, Varon M, Ojea-Jiménez I, Arbiol J, Puntes V (2011) Pt nanocrystal evolution in the
presence of Au(III)-salts at room temperature: spontaneous formation of AuPt heterodimers.
J Mater Chem 21:11518–11523
192. Yang J, Lee JY, Chen LX, Too H-P (2005) A phase-transfer identification of core-shell
structures in Ag-Pt nanoparticles. J Phys Chem B 109:5468–5472
193. Luo J, Wang L, Mott D, Njoki PN, Lin Y, He T, Xu Z, Wanjala BN, Lim IS, Zhong C (2008)
Core/shell nanoparticles as electrocatalysts for fuel cell reactions. Adv Mater 20:4342–4347
194. Zhou W-P, Yang X, Vukmirovic MB, Koel BE, Jiao J, Peng G, Mavrikakis M, Adzic RR
(2009) Improving electrocatalysts for O2 reduction by fine-tuning the Pt–support interaction:
164 H. Liu et al.

Pt monolayer on the surfaces of a Pd3Fe(111) single-crystal alloy. J Am Chem Soc


131:12755–12762
195. Yang J, Chen X, Yang X, Ying JY (2012) Stabilization and compressive strain effect of AuCu
core on Pt shell for oxygen reduction reaction. Energy Environ Sci 5:8976–8981
196. Liu H, Yang J (2014) Bimetallic Ag–hollow Pt heterodimers via insideout migration of Ag in
core–shell Ag–Pt nanoparticles at elevated temperature. J Mater Chem A 2:7075–7081
197. Zhang J, Vukmirovic MB, Sasaki K, Nilekar AU, Mavrikakis M, Adzic RR (2005) Mixed-
metal Pt monolayer electrocatalysts for enhanced oxygen reduction kinetics. J Am Chem Soc
127:12480–12481
198. Zhang J, Vukmirovic MB, Xu Y, Mavrikakis M, Adzic RR (2005) Controlling the catalytic
activity of platinum-monolayer electrocatalysts for oxygen reduction with different sub-
strates. Angew Chem Int Ed 44:2132–2135
199. Mokari T, Rothenberg E, Popov I, Costi R, Banin U (2004) Selective growth of metal tips
onto semiconductor quantum rods and tetrapods. Science 304:1787–1790
200. Shi WL, Zeng H, Sahoo Y, Ohulchanskyy TY, Ding Y, Wang ZL, Swihart M, Prasad PN
(2006) A general approach to binary and ternary hybrid nanocrystals. Nano Lett 6:875–881
201. Yang J, Levina L, Sargent EH, Kelley SO (2006) Heterogeneous deposition of noble metals
on semiconductor nanoparticles in organic or aqueous solvents. J Mater Chem 16:4025–4028
202. Costi R, Saunders AE, Elmalem E, Salant A, Banin U (2008) Visible light-induced charge
retention and photocatalysis with hybrid CdSe-Au nanodumbbells. Nano Lett 8:637–641
203. Elmalem E, Saunders AE, Costi R, Salant A, Banin U (2008) Growth of photocatalytic CdSe-
Pt nanorods and nanonets. Adv Mater 20:4312–4317
204. Habas SE, Yang P, Mokari T (2008) Selective growth of metal and binary metal tips on CdS
nanorods. J Am Chem Soc 130:3294–3295
205. Yang J, Ying JY (2009) Room-temperature synthesis of nanocrystalline Ag2S and its
nanocomposites with gold. Chem Commun 22:3187–3189
206. Yang J, Ying JY (2010) Diffusion of gold from the inner core to the surface of Ag2S
nanocrystals. J Am Chem Soc 132:2114–2115
207. Zhang J, Tang Y, Lee K, Ouyang M (2010) Tailoring light–matter–spin interactions in
colloidal hetero-nanostructures. Nature 466:91–95
208. Zhang J, Tang Y, Lee K, Ouyang M (2010) Nonepitaxial growth of hybrid core-shell
nanostructures with large lattice mismatches. Science 327:1634–1638
209. Zhao N, Li L, Huang T, Qi L (2010) Controlled synthesis of PbS-Au nanostar-nanoparticle
heterodimers and cap-like Au nanoparticles. Nanoscale 2:2418–2423
210. Carbone L, Cozzoli PD (2010) Colloidal heterostructured nanocrystals: Synthesis and growth
mechanisms. Nano Today 5:449–493
211. Li M, Yu X-F, Liang S, Peng X-N, Yang Z-J, Wang Y-L, Wang Q-Q (2011) Synthesis of
Au-CdS core-shell hetero-nanorods with efficient exciton-plasmon interactions. Adv Funct
Mater 21:1788–1794
212. Qu J, Liu H, Wei Y, Wu X, Yue R, Chen Y, Yang J (2011) Coalescence of Ag2S and Au
nanocrystals at room temperature. J Mater Chem 21:11750–11753
213. Yang J, Ying JY (2011) Nanocomposites of Ag2S and noble metals. Angew Chem Int Ed
50:4637–4643
214. Yang J, Chen X, Ye F, Wang C, Zheng Y, Yang J (2011) Core-shell CdSe@Pt
nanocomposites with superior electrocatalytic activity enhanced by lateral strain effect. J
Mater Chem 21:9088–9094
215. Hu W, Liu H, Ye F, Ding Y, Yang J (2012) A facile solution route for the synthesis of PbSe–
Au nanocomposites with different morphologies. CrystEngComm 14:7049–7054
216. Haldar KK, Sinha G, Lahtinen J, Patra A (2012) Hybrid colloidal Au-CdSe pentapod
heterostructures synthesis and their photocatalytic properties. ACS Appl Mater Interfaces
4:6266–6272
217. Ding X, Zou Y, Jiang J (2012) Au-Cu2S heterodimer formation via oxidation of AuCu alloy
nanoparticles and in situ formed copper thiolate. J Mater Chem 22:23169–23174
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 165

218. Motl NE, Bondi JF, Schaak RE (2012) Synthesis of colloidal Au-Cu2S heterodimers via
chemically triggered phase segregation of AuCu nanoparticles. Chem Mater 24:1552–1554
219. Vinokurov K, Macdonald JE, Banin U (2012) Structures and mechanisms in the growth of
hybrid Ru–Cu2S nanoparticles: from cages to nanonets. Chem Mater 24:1822–1827
220. Wang D, Li X, Li H, Li L, Hong X, Peng Q, Li Y (2013) Semiconductor–noble metal hybrid
nanomaterials with controlled structures. J Mater Chem A 1:1587–1590
221. Ding X, Zou Y, Ye F, Yang J, Jiang J (2013) Pt–CuS heterodimers by sulfidation of CuPt
alloy nanoparticles and their selective catalytic activity toward methanol oxidation. J Mater
Chem A 1:11880–11886
222. Sitt A, Hadar I, Banin U (2013) Band-gap engineering, optoelectronic properties and appli-
cations of colloidal heterostructured semiconductor nanorods. Nano Today 8:494–513
223. Zhao Q, Ji M, Qian H, Dai B, Weng L, Gui J, Zhang J, Ouyang M, Zhu H (2014) Controlling
structural symmetry of a hybrid nanostructure and its effect on efficient photocatalytic
hydrogen evolution. Adv Mater 26:1387–1392
224. Vinokurov K, Bekenstein Y, Gutkin V, Popov I, Millo O, Banin U (2014) Rhodium growth
on Cu2S nanocrystals yielding hybrid nanoscale inorganic cages and their synergistic prop-
erties. CrystEngComm 16:9506–9512
225. Banin U, Ben-Shahar Y, Vinokurov K (2014) Hybrid semiconductor–metal nanoparticles:
from architecture to function. Chem Mater 26:97–110
226. Kraeutler B, Bard AJ (1978) Heterogeneous photocatalytic synthesis of methane from acetic
acid-new Kolbe reaction pathway. J Am Chem Soc 100:4317–4318
227. Baba R, Nakabayashi S, Fujishima A, Kenichi H (1985) Investigation of the mechanism of
hydrogen ecolution during photocatalytic water decomposition on metal-loaded semiconduc-
tor powders. J Phys Chem 89:1902–1905
228. Pastoriza-Santos I, Koktysh DS, Mamedov AA, Giersig M, Kotov NA, Liz-Marzan LM
(2000) One-pot synthesis of Ag@TiO2 core-shell nanoparticles and their layer-by-layer
assembly. Langmuir 16:2731–2735
229. Chandrasekharan N, Kamat PV, Hu JQ, Jones G (2000) Improving the photoelectrochemical
performance of nanostructured TiO2 films by adsorption of gold nanoparticles. J Phys Chem
B 104:10851–10857
230. Subramanian V, Wolf E, Kamat PV (2001) Semiconductor-metal composite nanostructures.
To what extent do metal nanoparticles improve the photocatalytic activity of TiO2 films? J
Phys Chem B 105:11439–11446
231. Kamat PV (2002) Photophysical, photochemical and photocatalytic aspects of metal
nanoparticles. J Phys Chem B 106:7729–7744
232. Hirakawa T, Kamat PV (2005) Charge separation and catalytic activity of Ag@TiO2 core-
shell composite clusters under UV-irradiation. J Am Chem Soc 127:3928–3934
233. Jakob M, Levanon H, Kamat PV (2003) Charge distribution between UV-irradiated TiO2 and
gold nanoparticles: determination of shift in the Fermi level. Nano Lett 3:353–358
234. Subramanian V, Wolf E, Kamat PV (2004) Catalysis with TiO2/gold nanocomposites. Effect
of metal particle size on the Fermi level equilibration. J Am Chem Soc 126:4943–4950
235. Mokari T, Sztrum CG, Salant A, Rabani E, Banin U (2005) Formation of asymmetric
one-side metal-tipped semiconductor nanocrystal dots and rods. Nat Mater 4:855–863
236. Saunders AE, Popov I, Banin U (2006) Synthesis of hybrid CdS-Au colloidal nanostructures.
J Phys Chem B 110:25421–25429
237. Mokari T, Aharoni A, Popov I, Banin U (2006) Diffusion of gold into InAsnanocrystals.
Angew Chem Int Ed 45:8001–8005
238. Yang J, Elim HI, Zhang Q, Lee JY, Ji W (2006) Rational synthesis, self-assembly, and optical
properties of PbS-Au heterogeneous nanostructures via preferential deposition. J Am Chem
Soc 128:11921–11926
239. Talapin DV, Yu H, Shevchenko EV, Lobo A, Murray CB (2007) Synthesis of colloidal PbSe/
PbS core-shell nanowires and PbS/Au nanowire-nanocrystal heterostructures. J Phys Chem C
111:14049–14054
166 H. Liu et al.

240. Menagen G, Mocatta D, Salant A, Popov I, Dorfs D, Banin U (2008) Selective gold growth on
CdSe seeded CdS nanorods. Chem Mater 20:6900–6902
241. Huang SS, Huang JM, Yang JA, Peng JJ, Zhang QB, Peng F, Wang HJ, Yu H (2010)
Chemical synthesis, structure characterization, and optical properties of hollow PbSx-solid
Au heterodimer nanostructures. Chem Eur J 16:5920–5926
242. Yang J, Lee JY, Ying JY (2011) Phase transfer and its applications in nanotechnology. Chem
Soc Rev 40:1672–1679
243. Liu H, Ye F, Ma X, Cao H, Yang J (2013) The growth of metal sulfide–Au/Ag
nanocomposites in a nonpolar organic solvent. CrystEngComm 15:7740–7747
244. Dukovic G, Merkle MG, Nelson JH, Hughes SM, Alivisatos AP (2008) Photodeposition of Pt
on colloidal CdS and CdSe/CdS semiconductor nanostructures. Adv Mater 20:4306–4311
245. Harbour JR, Wolkow R, Hair ML (1981) Effect of platinization on the photoproperties of
CdS pigments in dispersion. Determination by H2 evolution, O2 uptake, and electron spin
resonance spectroscopy. J Phys Chem 85:4026–4029
246. Reber J-F, Rusek M (1986) Photochemical hydrogen production with platinized suspensions
of cadmium sulfide and cadmium zinc sulfide modified by silver sulfide. J Phys Chem
90:824–834
247. Frank AJ, Goren Z, Willner I (1986) Photohydrogenation of acetylene and ethylene by Pt and
Rh supported on CdS semiconductor particles. J Chem Soc Chem Commun 15:1029–1030
248. Peng XG, Manna L, Yang WD, Wickham J, Scher E, Kadavanich A, Alivisatos AP (2000)
Shape control of CdSe nanocrystals. Nature 404:59–61
249. Salant A, Amitay-Sadovsky E, Banin U (2006) Direct self-assembly of gold-tipped CdSe
nanorods. J Am Chem Soc 128:10006–10007
250. Wang X, Zhuang J, Peng Q, Li Y (2005) A general strategy for nanocrystal synthesis. Nature
437:121–124
251. Shi H, Fu X, Zhou X, Hu Z (2006) Preparation of organic fluids with high loading concen-
tration of Ag2S nanoparticles using the extractant Cyanex301. J Mater Chem 16:2097–2101
252. Motte L, Pileni MP (1998) Influence of length of alkyl chains used to passivate silver sulfide
nanoparticles on two- and three-dimensional self-organization. J Phys Chem B
102:4104–4109
253. Gao F, Lu Q, Zhao D (2003) Controllable assembly of ordered semiconductor Ag2S
nanostructures. Nano Lett 3:85–88
254. Liu J, Raveendran P, Shervani Z, Ikushima Y (2004) Synthesis of Ag2S quantum dots in
water-in-CO2 microemulsions. Chem Commun 2582–2583Please provide the volume num-
ber for Ref. [254].This journal has no volume number.
255. Wang X, Liu W, Hao J, Fu X, Xu B (2005) A simple large-scale synthesis of well-defined
silver sulfide semiconductor nanoparticles with adjustable sizes. Chem Lett 34:1664–1665
256. Gao F, Lu Q, Komarneni S (2005) Interface reaction for the self-assembly of silver
nanocrystals under microwave-assisted solvothermal conditions. Chem Mater 17:856–860
257. Brelle MC, Zhang JZ, Nguyen L, Mehra RK (1999) Synthesis and ultrafast study of cysteine-
and glutathione-capped Ag2S semiconductor colloidal nanoparticles. J Phys Chem A
103:10194–10201
258. Yang L, Xing R, Shen Q, Jiang K, Ye F, Wang J, Ren Q (2006) Fabrication of protein-
conjugated silver sulfide nanorods in the bovine serum albumin solution. J Phys Chem B
110:10534–10539
259. Mo X, Krebs MP, Yu SM (2006) Directed synthesis and assembly of nanoparticles using
purple membrane. Small 2:526–529
260. Lee JS, Shevchenko EV, Talapin DV (2008) Au-PbS core-shell nanocrystals: plasmonic
absorption enhancement and electrical doping via intra-particle charge transfer. J Am Chem
Soc 130:9673–9675
261. Michaelson HB (1977) The work function of the elements and its periodicity. J Appl Phys
48:4729–4733
262. Yeager E (1984) Electrocatalysts for O2 reduction. Electrochim Acta 29:1527–1537
3 Pt-Containing Heterogeneous Nanomaterials for Methanol Oxidation. . . 167

263. Aricò AS, Srinivasan S, Antonucci V (2001) DMFCs: from fundamental aspects to technol-
ogy development. Fuel Cells 1:133–161
264. Toda I, Igarashi H, Uchida H, Watanabe M (1999) Enhancement of the electroreduction of
oxygen on Pt alloys with Fe Ni and Co. J Electrochem Soc 146:3750–3756
265. Xu C, Goodman DW (1996) Adsorption and reaction of formic acid on a pseudomorphic
palladium monolayer on Mo(110). J Phys Chem 100:245–252
266. Baldauf M, Kolb DM (1996) Formic acid oxidation on ultrathin Pd films on Au(hkl) and Pt
(hkl) electrodes. J Phys Chem 100:11375–11381
267. Kibler LA, Kolb DM (2003) Physical electrochemistry: recent developments. Z Phys Chem
217:1265–1280
268. Kibler LA, El-Aziz AM, Hoyer R, Kolb DM (2005) Tuning reaction rates by lateral strain in a
palladium monolayer. Angew Chem Int Ed 44:2080–2084
269. Kumar S, Zou S (2007) Electrooxidation of carbon monoxide and methanol on platinum-
overlayer-coated gold nanoparticles: effects of film thickness. Langmuir 23:7365–7371
270. Naohara H, Ye S, Uosaki K (2000) Electrocatalytic reactivity for oxygen reduction at
epitaxially grown Pd thin layers of various thickness on Au(111) and Au(100). Electrochim
Acta 45:3305–3309
271. Naohara H, Ye S, Uosaki K (2001) Thickness dependent electrochemical reactivity of
epitaxially palladium thin layers on Au(111) and Au(110) surfaces. J Electroanal Chem
500:435–445
272. El-Aziz AM, Kibler LA (2002) Influence of steps on the electrochemical oxidation of CO
adlayers on Pd(111) and on Pd films electrodeposited onto Au(111). J Electroanal Chem
534:107–114
273. Kibler LA, El-Aziz AM, Kolb DM (2003) Electrochemical behaviour of pseudomorphicov-
erlayers: Pd on Au(111). J Mol Catal A Chem 199:57–63
274. Yang JH, Cheng CH, Zhou W, Lee JY, Liu Z (2010) Methanol-tolerant heterogeneous
PdCo@PdPt/C electrocatalyst for the oxygen reduction reaction. Fuel Cells 10:907–913
275. Yang JH, Lee JY, Zhang Q, Zhou W, Liu Z (2008) Carbon-supported pseudo-core-shell Pd-Pt
nanoparticles for ORR with and without methanol. J Electrochem Soc 155:B776–B781
276. Yang JH, Zhou W, Cheng CH, Lee JY, Liu Z (2010) Pt-decorated PdFe nanoparticles as
methanol-tolerant oxygen reduction electrocatalyst. ACS Appl Mater Interfaces 2:119–126
277. Adzic RR, Zhang J, Sasaki K, Vukmirovic MB, Shao M, Wang JX, Nilekar AU,
Mavrikakis M, Valerio JA, Uribe F (2007) Platinum monolayer fuel cell electrocatalysis.
Top Catal 46:249–262
278. Shemesh Y, Macdonald JE, Menagen G, Banin U (2011) Synthesis and photocatalytic
properties of a family of CdS-PdX hybrid nanoparticles. Angew Chem Int Ed 123:1217–1221
279. Shaviv E, Schubert O, Alves-Santos M, Goldoni G, Felice RD, Vallée F, Fatti ND, Banin U,
S€onnichsen C (2011) Absorption properties of metal-semiconductor hybrid nanoparticles.
ACS Nano 5:4712–4719
280. Kim SW, Kim M, Lee WY, Hyeon T (2002) Fabrication of hollow palladium spheres and
their successful application to the recyclable heterogeneous catalyst for Suzuki coupling
reactions. J Am Chem Soc 124:7642–7643
281. Li Y, Zhou P, Dai Z, HuZ SP, Bao J (2006) A facile synthesis of PdCo bimetallic hollow
nanospheres and their application to Sonogashira reaction in aqueous media. New J Chem
30:832–837
282. Cheng F, Ma H, Li Y, Chen J (2007) Ni1-xPtx (x ¼ 0-0.12) hollow spheres as catalysts for
hydrogen generation from ammonia borane. Inorg Chem 46:788–794
283. Liu H, Ye F, Cao H, Ji G, Lee JY, Yang J (2013) A core–shell templated approach to the
nanocomposites of silver sulfide and noble metal nanoparticles with hollow/cage-bell struc-
tures. Nanoscale 5:6901–6907
284. Feng Y, Liu H, Wang P, Ye F, Tan Q, Yang J (2014) Enhancing the electrocatalytic property
of hollow structured platinum nanoparticles for methanol oxidation through a hybrid con-
struction. Sci Rep 4:6204
168 H. Liu et al.

285. Wang X, Waje M, Yan Y (2004) Methanol resistant cathodic catalyst for direct methanol fuel
cells. J Electrochem Soc 151:A2183–A2188
286. Xia BY, Wu HB, Yan Y, Low XW, Wang X (2013) Ultrathin and ultralong single-crystal
platinum nanowire assemblies with highly stable electrocatalytic activity. J Am Chem Soc
135:9480–9485
287. Liu F, Lu G, Wang C-Y (2006) Low crossover of methanol and water through thin mem-
branes in direct methanol fuel cells. J Electrochem Soc 153:A543–A555
288. Du CY, Zhao TS, Yang WW (2007) Effect of methanol crossover on the cathode behavior of
a DMFC: a half-cell investigation. Electrochim Acta 52:5266–5271
289. Jia N, Lefebvre MC, Halfyard J, Qi Z, Pickup PG (2000) Modification of Nafion proton
exchange membranes to reduce methanol crossover in PEM fuel cells. Electrochem Solid-
State Lett 3:529–531
290. Gurau B, Smotkin ES (2002) Methanol crossover in direct methanol fuel cells: a link between
power and energy density. J Power Sources 112:339–352
291. Kim Y-M, Park K-W, Choi J-H, Park I-S, Sung Y-E (2003) A Pd-impregnated
nanocomposite Nafion membrane for use in high-concentration methanol fuel in DMFC.
Electrochem Commun 5:571–574
292. Sahu AK, Pitchumani S, Sridhar P, Shukla AK (2009) Nafion and modified-Nafion mem-
branes for polymer electrolyte fuel cells: An overview. Bull Mater Sci 32:285–294
293. Zhang H, Huang H, Shen PK (2012) Methanol-blocking Nafion composite membranes
fabricated by layer-by-layer self-assembly for direct methanol fuel cells. Int J Hydrogen
Energy 37:6875–6879
294. Zhang Y, Cai W, Si F, Ge J, Liang L, Liu C, Xing W (2012) A modified Nafion membrane
with extremely low methanol permeability via surface coating of sulfonated organic silica.
Chem Commun 48:2870–2872
295. Beauger C, Lainé G, Burr A, Taguet A, Otazaghine B, Rigacci A (2013) Nafion®-sepiolite
composite membranes for improved proton exchange membrane fuel cell performance. J
Membr Sci 130:167–179
296. Feng Y, Yang J, Liu H, Ye F, Yang J (2014) Selective electrocatalysts toward a prototype of
the membraneless direct methanol fuel cell. Sci Rep 4:3813
297. Cameron DS, Hards GA, Harrison B, Potter RJ (1987) Direct methanol fuel cells. Platinum
Metals Rev 31:173–181
298. Ferrigno R, Stroock AD, Clark TD, Mayer M, Whitesides GM (2002) Membraneless
vanadium redox fuel cell using laminar flow. J Am Chem Soc 124:12930–12931
299. Jayashree RS, Egas D, Spendelow JS, Natarajan D, Markoski LJ, Kenis PJA (2006)
Air-breathing laminar flow-based direct methanol fuel cell with alkaline electrolyte.
Electrochem. Solid State Lett 9:A252–A256
300. Lam A, Wilkinson DP, Zhang J (2008) Novel approach to membraneless direct methanol fuel
cells using advanced 3D anodes. Electrochim Acta 53:6890–6898
301. Lam A, Wilkinson DP, Zhang J (2009) Control of variable power conditions for a
membraneless direct methanol fuel cell. J Power Sources 194:991–996
302. Lam A, Dara MS, Wilkinson DP, Fatih K (2012) Aerobic and anaerobic operation of an
active membraneless direct methanol fuel cell. Electrochem Commun 17:22–25
303. Tominaka S, Ohta S, Obata H, Momma T, Osaka T (2008) On-chip fuel cell: micro direct
methanol fuel cell of an air-breathing, membraneless, and monolithic design. J Am Chem Soc
130:10456–10457
Chapter 4
Synthesis and Electrocatalysis of Pt-Pd
Bimetallic Nanocrystals for Fuel Cells

Ruizhong Zhang and Wei Chen

4.1 Introduction

Over the past few decades, research on novel environment-friendly energy conver-
sion devices and their potential applications has attracted extensive interest due to
the depletion of fossil fuel reserves and thus the growing demand for efficient but
low-cost renewable energy [1, 2]. Fuel cells, such as proton-exchange membrane
fuel cells (PEMFCs), direct methanol fuel cells (DMFCs), and direct formic acid
fuel cells (DFAFCs), have been considered as the most promising power sources
because of high power density, high energy-conversion efficiency, and zero or low
emission of pollutants [3–6]. For fuel cells, platinum (Pt) has received unremitting
interest as an electrocatalyst because of the highest catalytic activity among the
studied catalysts for electro-oxidation of small organic fuels on the anode and
oxygen reduction on the cathode [7–10]. However, several issues arise from
using pure Pt as fuel cell catalysts. First, Pt surfaces are easily self-poisoned by
the strong adsorption of CO intermediates originated from small organic fuel
oxidation, leading to a severe decrease in the catalytic performance [11, 12]. Second,
using Pt as cathodic catalysts in DMFCs, methanol molecules crossover from anode
to cathode may lower the ORR performance because of the mixed potentials formed
from the simultaneous methanol oxidation and oxygen reduction [13–15]. Third,
the limited reserve in nature and the resulting sky-rocketing price of Pt has become

R. Zhang
State Key Laboratory of Electroanalytical Chemistry, Changchun Institute of Applied
Chemistry, Chinese Academy of Sciences, Changchun, Jilin 130022, China
University of Chinese Academy of Sciences, Beijing 100039, China
W. Chen (*)
State Key Laboratory of Electroanalytical Chemistry, Changchun Institute of Applied
Chemistry, Chinese Academy of Sciences, Changchun, Jilin 130022, China
e-mail: [email protected]

© Springer International Publishing Switzerland 2016 169


K.I. Ozoemena, S. Chen (eds.), Nanomaterials for Fuel Cell Catalysis,
Nanostructure Science and Technology, DOI 10.1007/978-3-319-29930-3_4
170 R. Zhang and W. Chen

one of the major barriers for the wide commercialization of fuel cells [16]. To
improve the catalytic activities and lower the costs of catalysts in fuel cells, much
effort has been devoted to combining Pt with another metal that is less constrained
in terms of reserve and economic dependence, to form bimetallic nanocatalysts. As
compared to monometallic Pt, bimetallic Pt-based nanocrystals are expected to
display not only a combination of the properties associated with two distinct metals,
but also new and unexplored properties because of a possible bi-functional mech-
anism or the so-called ligand effect [17–19].
Among the metals around Pt in the periodic table, palladium (Pd) is probably the
best candidate to generate bimetallic electrocatalysts with Pt due to the following
obvious advantages. First, Pd and Pt share the same face-centered cubic (fcc)
structure and almost identical lattice constants (with a mismatch of only 0.77%),
both of which are beneficial to the formation of Pt-Pd bimetallic nanocrystals with
single crystallinity. Second, the combination of Pt with Pd has a crucial impact on
the electronic structure of Pt, resulting in superior electrocatalytic activities for a
specific reaction owing to the formation of Pt-Pd bonds [20, 21]. According to the
studies by Nørskov and co-workers [22–25], the strain and electronic coupling
presented in a catalyst play a key role in determining the catalytic properties
because both of the two parameters can result in shifts of the d-band center
(calculated with respect to the Fermi level), which is directly related to the
adsorption energies of reactants on a catalyst as well as their activation barriers.
Of them, a compressive strain tends to down-shift the energy of the d-band center,
causing adsorbates to bind less strongly to the catalyst, while a tensile strain has the
opposite effect. On the other hand, the electronic coupling can result in shifts for the
d-band center due to a change in the density of states near the Fermi level. As for
Pt-Pd bimetallic catalysts, a small compressive strain resulting from their weak
lattice contract can cause a downshift of the d-band center, thus lowering the
binding strength for the adsorbed intermediates. As a consequence, the changes in
the d-band properties of Pt caused by its combination with Pd, and the small
mismatch in the lattice constants between Pt and Pd, makes Pt-Pd bimetallic
nanocrystals attractive catalysts for various reactions. Taking ORR as an example,
the rate-limiting step at high potential is the desorption of intermediates (O, OH,
etc.) produced during the reaction. The weakened binding strength of the interme-
diates on a Pt-Pd bimetallic catalyst can activate the Pt sites required for the
adsorption of O2 and then the splitting of the O-O bond, therefore accelerating
the kinetics of oxygen reduction [26]. In addition, a strong interaction between Pt
and Pd due to the formation of Pt-Pd bonds can also change the electronic structure
of Pt, causing the enhanced amount of O2 adsorbed on the surface of Pt and thus
improving ORR catalytic activity.
Moreover, it has been identified that the dissolution of Pt is a major reason for
the degradation of catalysts in proton-exchange membrane (PEM) fuel cells due to
the presence of dissolved molecular oxygen, highly aggressive condition in terms of
acidic pH, and the highly positive potential for device operation [27]. Therefore,
compared to other metals such as Ag, Cu, Co or Ni, the use of Pd may also help
minimize the corrosion and loss of catalysts in an acidic environment such as the
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 171

medium of a PEM fuel cell [28]. Especially, the introduction of Pd can prevent the
electrocatalysts from degradation to a certain extent by up-shifting the dissolution
potential of Pt and thus assure long-term stability [29]. From these perspectives, it
has been actively explored Pt-Pd bimetallic catalysts for a variety of reactions with
enhanced performance.
In addition to elemental compositions, the size, shape and structure of Pt-Pd
bimetallic system can also be finely manipulated to further enhance their catalytic
performance. Since electrocatalytic reactions are very sensitive to the exposed
crystal facets and the proportion of atoms located at the facets, edges or corners
of catalysts, recent years tremendous efforts have been devoted to the syntheses of
Pt-Pd bimetallic nanocrystals with well-defined shapes in high yields and purity by
tuning various experimental parameters [30]. By using Pd nanocrystals as the seeds
for overgrowth or sacrificial templates for galvanic replacement, Pt-Pd alloys with
different structures/shapes including alloys, core-shell, dendrities, alternating
multi-shells, and atomic monolayer, can be easily synthesized by different strate-
gies. These preparation methods include co-chemical reduction and its combination
with galvanic replacement to generate alloy nanocrystals, galvanic replacement
between Pd nanocrystals and a Pt salt precursor for generating dendritic
nanostructures, seed-mediated overgrowth for generating core-shell, multi-shell,
and dendritic nanostructures, and a combination of electrochemical deposition and
galvanic replacement for generating Pt monolayer on Pd nanocrystals, etc.
Besides the catalytic activity and costs, the stability and lifetime of an
electrocatalyst are also critical issues for its practical applications in fuel cells.
When Pt-Pd bimetallic nanocrystals are used as fuel cell catalysts, the catalyst
support also plays an important role in determining the catalytic properties. For a
catalyst support, it should have a high surface area for catalyst dispersion, a strong
affinity to immobilize the catalyst particles, a high electrical conductivity to
accelerate electron transfer in redox reactions, and a high electrochemical stability
in acidic or alkaline electrolytes to ensure a stable structure. Up to now, various
carbon materials including traditional carbon materials (e.g., Vulcan XC-72R
carbon black) and nanocarbon materials (e.g., carbon nanotubes, graphene and
ordered mesoporous carbon) have been used for the dispersion of catalyst particles.
However, earlier studies have shown that except for the dissolution and aggregation
of metal nanoparticles, the severe corrosion and oxidation of carbon support
materials in the harsh operating environment could also lead to quick degradation
of the electrocatalytic performance [31, 32]. Therefore, in recent years great efforts
have also been devoted to addressing the challenges of catalyst supports via
developing non-carbon support materials such as metal oxides, electronically
conductive polymer, nitrides and carbides [33].
Herein, we will first summarize recent progress in the development of experi-
mental techniques for the preparation of unsupported/supported Pt-Pd bimetallic
nanocrystals with unique structures, and then focus on their applications in fuel
cells as anode and cathode catalysts.
172 R. Zhang and W. Chen

4.2 Synthesis of Pt-Pd Bimetallic Nanocrystals

It is well-known that the catalytic activities of nanoparticle electrocatalysts are


strongly dependent on their composition, shapes, size, exposed surface planes and
the interactions between nanocrystals and catalyst supports. In the past few
decades, various methods have been applied in the synthesis of Pt-Pd bimetallic
nanocrystals with structures in the form of alloy, core-shell, dendrities, multi-shells,
and supported monolayer. These well-defined Pt-Pd nanostructures can be gener-
ally prepared through co-chemical reduction, galvanic replacement, seed-mediated
growth and electrochemical deposition synthetic routes. Typically, co-chemical
reduction synthesis of Pt-Pd bimetallic nanocrystals refers to the reduction of
Pt and Pd precursors in the presence of capping agents and/or stabilizers. Pt-Pd
nanoalloys can also be obtained through in situ oxidation and dissociation of Pd
nanocrystals by galvanic replacement with a Pt salt precursor. For a seed-mediated
growth approach, the well-defined shapes of Pt or Pd as seeds for epitaxial growth
of a Pd or Pt-shell. It is also feasible to control the shell by using electrochemical
deposition methods. In this section, we summarize various leading synthetic tech-
niques and the formation mechanisms for the preparation of Pt-Pd nanocrystals.

4.2.1 Co-chemical Reduction Method

In the past decades, co-chemical reduction has proved to be a straightforward


strategy for the facile synthesis of bimetallic nanocrystals. With this technique, it
is feasible to achieve simultaneous reduction of both Pt and Pd salt precursors in the
presence of a capping agent and/or a stabilizer due to their similar electrochemical
potentials of 0.74 V (versus a reversible hydrogen electrode) for PtCl62−/Pt and
0.62 V for PdCl42−/Pd. The obtained Pt-Pd bimetallic nanocrystals are interesting
for various electrocatalytic reactions due to the co-existence of Pt and Pd atoms on
the nanocrystal surfaces [34]. To date, various reducing agents have been used to
co-reduce Pd and Pt salt precursors for the synthesis of Pt-Pd alloy nanocrystals,
including sodium borohydride, alcohol, formic acid, formaldehyde, hydrazine,
ascorbic acid, etc. By using sodium borohydride (NaBH4) as a reductant, Crooks
and coworkers [35] prepared Pt-Pd bimetallic nanoparticles via the co-reduction of
K2PtCl4 and K2PdCl4 in a poly(amidoamine) dendrimer (G6-OH) aqueous solution.
Two steps are involved in the process: (i) loading Pt and Pd ions into the dentrimers
by fully complexing with the interior amines of dendrimers, and (ii) co-reducing the
complexed Pd and Pt ions by NaBH4. Interestingly, the synthesized bimetallic
Pt-Pd nanoparticles by this process have almost the same diameter (~1.8 nm)
containing an average of 180 atoms but with seven different Pt/Pd ratios. Here,
the dendrimer template plays a vital role in controlling the size of Pt-Pd
nanoparticles (smaller than 3 nm) and the composition variation (molar ratios
adjustable from 1:5 to 5:1). TEM and single-particle energy-dispersive
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 173

spectroscopy (EDS) indicated that the calculated particle diameter and Pt/Pd ratios
in the Pt and Pd precursors are very consistent with the measured ones. Therefore
this dendrimer-templating is a unique method for preparing nanoparticles having
particular Pt/Pd ratios, uniform size and composition. In another work,
one-dimensional ultrathin Pt-Pd alloy nanowires were synthesized in a cetyltri-
methylammonium (CTAB)-assisted water-chloroform micelles system [36]. In this
synthesis, a mixed aqueous solution of Pt and Pd salts was mixed with a chloroform
solution of CTAB, followed by the addition of a NaBH4 aqueous solution. In the
mixed solution, the reduction of precursor ions and metal growth occurred in the
swollen wormlike micelle networks of chloroform droplets with the CTAB mole-
cules. Because of the strict limitation of the wormlike micelle networks, the
obtained Pt-Pd nanowires showed an average size of 2.5 nm with a narrow diameter
distribution, and both Pt and Pd can be co-reduced to form alloy with an atomic
ratio of about 1:1, as confirmed by EDX measurements.
When Pt-Pd bimetallic materials are used as fuel cell catalysts, they are usually
dispersed on a catalyst support. Among the used catalyst supports, Vulcan XC-72
carbon black is the most popular carbon support for immobilizing and stabilizing
Pt-Pd nanoparticles [37–44], while carbon nanotubes [45–48], conducting polymer
composite matrix [49], and tungsten carbide [50] have also been explored as
potential supporting materials. For instance, Zhang et al. [51] synthesized a series
of PdxPt1−x nanoparticles dispersed on carbon black by reducing the mixture of Pd
(II) and Pt(II) precursors by NaBH4 in the presence of Vulcan XC-72 carbon. In the
synthesis, EDTA was also used as a chelator for Pd and Pt ions to ensure the
co-reduction of Pd(II) and Pt(II) species. The highly dispersed PdxPt1−x
nanoparticles on carbon black exhibited composition-dependent catalytic activity
for formic acid electro-oxidation and the Pd0.9Pt0.1 nanoparticles with a mean size
of 3.2 nm showed the best performance among the series.
In addition to strong reducing agents, Pt-Pd bimetallic nanostructures have also
been prepared by a polyol process, in which mild reductants such as ethylene glycol
[52–68], methanol [69], glycerol [70, 71], and 1,4-butanediol [53], have been used.
It is well-known that Pt nanocrystals with different exposed surfaces have different
electronic structures and surface atomic arrangements, and the appropriate crystal
phase and/or composition can greatly enhance reaction kinetics due to the mini-
mized surface energy and total excess free energy. Therefore, much work has been
done to prepare shape-controlled metallic nanostructures with desired exposed
crystal facets. For instance, Lee and co-workers [72] reported a glycerol reduction
method to synthesize Pt-Pd nanoparticles exhibiting dominantly exposed (111)
facets in octahedral shape with complete alloy formation between Pt and Pd
metallic phases. During the fast reduction process, the thermodynamically mini-
mized crystalline surface energy and thus the formation of crystal facets with a low
surface energy in the Pt-Pd structure can improve the electrocatalytic performance
of the obtained carbon supported Pt-Pd nanoparticles composite toward methanol
oxidation. By using a hollow-core mesoporous shell (HCMS) carbon as support,
Berker et al. [65] reported a rapid method to synthesize Pt-Pd/HCMS composite by
the co-reduction of H2PtCl6 and PdCl2 using ethylene glycol as a reducing agent
174 R. Zhang and W. Chen

under microwave irradiation. It was proposed that the HCMS carbon facilitated the
diffusion of hydrogen and oxygen as well as the water transport within fuel cells,
leading to significantly improved fuel cell performance. Formic acid has also been
used to synthesize Pt-Pd bimetallic nanocrystals [73–75]. Guo et al. [76] demon-
strated a simple procedure to synthesize Pt/Pd hybrid supported on polyaniline
(PANI) nanofibers with high conductivity and surface-to-volume ratios. In this
method, the PANI nanofibers were first synthesized by a wet-chemical approach
and then the as-obtained 1D PANI nanofibers were added into a mixture containing
H2PtCl6 and H2PdCl4. HCOOH was finally injected into the above mixture to
reduce the precursors at room temperature. The SEM and TEM images shown in
Fig. 4.1a, b indicate that a large number of PANI nanofibers with a diameter of 60–
100 nm have been obtained. Compared with the smooth surface of pristine PANI, a
rougher surface of the as-prepared PANI/Pt and PANI/PtPd indicates that Pt
(Fig. 4.1c, d) and PtPd nanoparticles (Fig. 4.1e, f) have been successfully deposited
on the surface of PANI fibers. Moreover, the corresponding magnified images
further revealed that the small PtPd nanoparticles formed a network and thus
leading to many nanoporous structures on the surface of hybrid nanofibers, which
favor a high electrocatalytic activity.
Most recently, by using graphene as support, Wang and co-workers [77] used the
formic acid method to prepare graphene-supported 1D PtPd nanorods (G-PtPd
NRs) by co-reducing H2PtCl6 and Pd(NO3)2 in the presence of HCOOH. In
comparison with the carbon-supported PtPd NRs and graphene-supported Pt NRs,
the G-PtPd NRs showed a larger diameter of about 4.4 nm and a longer length of
about 35 nm measured from the corresponding TEM images. Extended X-ray
absorption fine structure (EXAFS) studies confirmed the formation of G-PtPd
alloy with a Pt-rich inner core and a Pd-rich outer shell. XRD patterns indicated
that the growth of (111) and (220) planes of G-PtPd NRs was promoted for the
alloying of Pt and Pd on graphene support. In addition, the results of X-ray
absorption near edge spectroscopy (XANES) also showed that using graphene as
a support and alloying with Pd synergistically modified the d-band of Pt, and the
total number of unoccupied d-states for G-PtPd was reduced to as low as 0.295. All
these results suggested that the G-PtPd had lower unfilled d-states and more d-band
electrons were transferred from Pd to Pt, resulting in enhanced ORR performance.
Meanwhile, a simple microemulsion method was also developed to construct
Pt-Pd nanoparticles [78, 79]. With this technique, the preparation of metal
nanoparticles is realized by mixing two different micro-emulsions carrying the
specific reactants (metallic salts and reducing agent) dissolved in aqueous phase.
Microemulsion method has been accepted to be a unique method for the production
of metal particles with a very narrow size distribution. For example, Escudero and
coworkers [80] prepared alloyed Pt-Pd nanoparticles by reducing H2PtCl6 and
PdCl2 with hydrazine in a water-in-oil micro-emulsion of water/Berol 050/iso-
octane. The obtained Pt-Pd nanoparticles were smaller than 5 nm and exhibited
potential application in fuel cells.
In recent years, microwave-assisted technique has also been applied to the
synthesis of metal nanoparticles [81]. Compared with the traditional chemical
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 175

Fig. 4.1 (a) SEM and (b) TEM images of PANI nanofibers. (c, d) TEM images of PANI
nanofiber-supported supra-high density Pt NPs. (e, f) TEM images of PANI nanofiber-supported
Pt/Pd NPs at different magnifications. Reprinted with permission from [76]. Copyright 2009
Wiley-VCH

methods, the microwave heating provides homogeneous reaction conditions and a


fast kinetic process. Zhang and coworkers [82] reported a simple method to
synthesize three-dimensional Pd@Pt core-shell nanostructures with a high yield
via the co-reduction of K2PtCl4 and PdCl2 in the presence of CTAB by using
ascorbic acid as a reducing agent under microwave irradiation. It was shown that
the long alkyls of CTAB served as a shape-control agent to tailor the structure and
improve the dispersion of nanoparticles in the synthesis. TEM and SEM
176 R. Zhang and W. Chen

characterizations showed that the morphology of the synthesized Pd@Pt


nanostructures can be easily controlled by changing the molar ratio between Pt
and Pd precursors. And the Pd@Pt morphology was changed from cubic to spher-
ical shapes by decreasing the Pd/Pt molar ratio (3:1 to 1:3). The authors also studied
the shape effect on the eletrocatalytic activity for MOR and ORR, and among the
different Pd@Pt nanostructures, the sample at a Pd/Pt molar ratio of 1:3 exhibited
the best catalytic activity.
Interfacial self-assembly was found to be a simple and effective strategy for the
synthesis of noble metal nanomaterials, and the liquid-liquid, liquid-air and oil-
water-air interfaces have showed promise as flexible 2D platforms for nanoparticles
assembly [83–87]. Wu et al. [88] prepared macroscopic free-standing Pd/Pt
nanomembranes (Pd/Pt-FNMs) with well-defined monolayer structures as large as
several square centimeters by one-step self-assembly at a water-air interface. In the
synthesis, aqueous solutions of PdCl2, K2PtCl6 and sodium citrate (Na3CA) were
mixed in a beaker. Subsequently, the reducing agent NH2OH.HCl was quickly
added into the above solution. After boiling for 15 min and then cooling down to
room temperature, a monolayer of macroscopic Pd/Pt-FNMs floated at the water-air
interface and covered the solution surface of the whole beaker. The composition of
the Pd/Pt-FNMs (Pd53Pt47-FNMs, Pd33Pt67-FNMs, and Pd15Pt85-FNMs) was easily
tuned by adding a different amount of precursors and confirmed by ICP-AES. From
the TEM images, the obtained Pd/Pt-FNMs exhibited similar well-defined mor-
phologies, and Fig. 4.2a–c shows the details of the Pd33Pt67-FNMs. It can be seen
that the Pd/Pt nanoparticles have a uniform size with an average diameter of
9.25 nm. Although no obvious grain boundaries between Pd and Pt are observed
from the high-resolution TEM (HRTEM) measurements (Fig. 4.2d), the bimetallic
nature of the FNMs can be clearly seen from the high-angle annular dark-field
(HDDAF-STEM) image and the nanometer-scale TEM elemental mapping images
(Fig. 4.2e). In the mechanistic study, the authors found that when increasing the
dosage of Na3CA, Pd/Pt nanomembranes failed to form at the water-air interface.
However, with decreasing dosage of Na3CA, highly aggregated nanostructures
were produced. Therefore, the presence of an appropriate dosage of Na3CA was
believed to be the key to achieve high-quality Pd/Pt-FNMs.
In another study, Sun and coworkers [89] reported an oil-phase method for the
synthesis of polyhedral Pd-Pt alloy nanocrystals with a controlled size (3.5–6.5 nm)
and composition (Pd88Pt12 to Pd34Pt66). This synthesis involves co-reduction of Pd
(acac)2 and Pt(acac)2 with morpholine borane (MB) in oleylamine (OAm) at
different temperatures. The TEM and HAADF-STEM images confirmed the for-
mation of single-crystal nanoparticles in high yield with a uniform size. Both Pd
and Pt are uniformly distributed throughout each particle from EDX mapping
measurements. The linear dependence with a nearly unity slope implies that the
two precursors were co-reduced at the same rate in generating the bimetallic
nanocrystals. The formation of a single complex between the Pt and Pd precursors
with the capping agent facilitates the simultaneous reduction in an elevated
oil-phase. In a typical synthesis, the composition of the Pt-Pd alloy nanocrystals
could be adjusted by controlling the amounts of Pd(acac)2 and Pt(acac)2 added into
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 177

Fig. 4.2 (a–c) TEM images, (d) HRTEM image, and (e) HAADF-STEM image and the
corresponding nanometer-scale TEM elemental mapping (inset) of the Pd33Pt67-FNMs. Reprinted
with permission from [88]. Copyright 2012 Wiley-VCH
178 R. Zhang and W. Chen

the reaction solution. More importantly, at a constant composition, the size of the
Pt-Pd nanocrystals could be easily tuned by varying the temperature at which MB
was injected. When MB was injected at 40, 60 and 80 ˚C, 6.5, 5 and 4.5 nm Pd67Pt33
nanoparticles can be formed, respectively. The as-synthesized Pt/Pd nanoparticles
exhibited highly composition-dependent catalytic activities and high stability for
methanol oxidation reaction in acid media.
Although great success has been achieved in the synthesis of Pt-Pd alloy
nanoparticles via the afore-mentioned techniques, the products from these synthe-
ses were largely restricted to small nanoparticles (typically less than 5 nm in size)
with poorly defined crystallinity and morphology. By heating an aqueous solution
containing Na2PdCl4, K2PtCl4 and poly (vinyl pyrrolidone) (PVP) at 80 ˚C for 18 h,
Xia and coworkers [90] demonstrated a co-reduction approach for the synthesis of
Pt-Pd alloy nanocrystals with well-defined shapes and twinned structures. In this
process, the commercially available PVP was employed as a weak reducing agent
to manipulate the reduction kinetics owing to its hydroxy (OH)-end groups. It was
found that the slow reduction rate associated with the weak reducing power of PVP
is the prerequisite for the formation of Pt-Pd alloy nanocrystals with twinned
structures. From the TEM characterizations in Fig. 4.3a, b, the formed Pt-Pd
nanocrystals are mainly star-shaped decahedra with an average size of 40 nm and
triangular nanoplates with lateral dimensions of 30–50 nm, as well as a small
fraction of other shapes such as octahedra. The compositional line profiles of Pd
and Pt on a star-shaped decahedron shown in Fig. 4.3c indicate the formation of
Pt-Pd alloy. The HRTEM image (Fig. 4.3d) confirms the presence of fivefold twins
from the center of a star-shaped decahedron. In this work, the PVP-mediated slow
reduction rate could help retain the particles at small sizes for a long period of time
before nucleation. During the period, the small particles easily coalesced into larger
particles to reduce surface-to-volume ratio, leading to the formation of twinned
structures. When the reaction was conducted using a relatively high-rate reducing
agent of ethylene glycol, Pt-Pd nanocrystals with a truncated, octahedral shape
were produced, and this fast reduction process favored the formation of Pt-Pd
nanocrystals with a single-crystal structure.
Hydrothermal method is another useful and frequently used technique for the
preparation of Pt-Pd nanocrystals with the advantages of simplicity, free of tem-
plates, and easy shape-control of metal nanocrystals [91–93]. For instance, Yan and
coworkers [94] demonstrated a shape-selective synthesis of Pt-Pd nanotetrahedrons
(NTs) and nanocubes (NCs) with less than 10 nm in size via a one-step hydrother-
mal process by using small ions as efficient facet-selective agents (Fig. 4.4a). With
a combination of Na2C2O4 and formaldehyde as the (111) facet-selective agent and
reducing agent, single-crystalline Pt-Pd NTs enclosed by four (111) facets with a
shape selectivity of ~70% and an average size of 4.9 nm were produced (Fig. 4.4b).
The selective adsorption of C2O42− species on the (111) facets was found to be a
critical factor in directing the formation of Pt-Pd NTs. In comparison, Pt-Pd NCs
(~8.5 nm) with a shape selectivity of about 88% were produced in the presence of
both large amount of Br− and tiny amount of I− anions owing to their selective
capping for the (100) facet (Fig. 4.4c). Furthermore, reduction rate dependent on
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 179

a b

50 nm 10 nm

c d
300 Pd
Pt

200
Counts

100

0 10 20 30 40
1 nm
Distance / nm

Fig. 4.3 (a, b) TEM images, (c) EDX line-scan, and (d) HRTEM image of Pt-Pd alloy
nanocrystals synthesized by co-reduction of Na2PdCl4 and K2PtCl4 with PVP in an aqueous
solution. The inset in (b), (c) and (d) correspond to schematic illustration, the HAADF-STEM
image, and the Fourier transform pattern of the star-shaped decahedron, respectively. Reprinted
with permission from [90]. Copyright 2009 Wiley-VCH

the type of reductant also played a vital role in determining the shape of the Pt-Pd
products. A fast enough reducing rate by using a certain amount of formaldehyde
could improve the shape- and size-uniformity of the obtained Pt-Pd NTs, while a
slow reduction rate (with PVP instead of formaldehyde as the reducing agent) was
found to be beneficial to the formation of regularly shaped Pt-Pd NCs. The
electrocatalytic studies showed that the obtained alloy nanocrystals exhibit
enhanced and facet-dependent catalytic activity and stability for methanol
electrooxidation in the order of NCs>NTs>commercial Pt/C.
Besides the intrinsic catalytic properties of catalysts, the support used also plays
important roles in determining their catalytic performance. Due to the high elec-
tronic conductivity and large surface area, graphene has been recently used as a
support material for Pt-Pd nanocrystals dispersion. Our group developed a facile
hydrothermal method for the one-pot fabrication of reduced graphene oxide (rGO)-
supported Pt-Pd alloy nanocubes (PtPd/rGO) [95]. In a typical procedure, Pd(acac)2
and Pt(acac)2 were mixed with PVP, and NaI in DMF solution of graphene oxide.
180 R. Zhang and W. Chen

Fig. 4.4 (a) Schematic illustration of shape-controlled synthesis of Pt-Pd alloy nanocrystals with
tetrahedral and cubic shapes. (b, c) TEM images of Pt-Pd tetrahedrons and cubes, respectively.
The insets in (b) and (c) show the percentages of the alloy tetrahedrons and cubes, respectively.
Reprinted with permission from [94]. Copyright 2011 American Chemical Society

After ultrasonic treatment, the mixed solution was then transferred to a Teflon-lined
stainless steel autoclave and heated at 150 ˚C for 8 h. In this process, two key steps
were included: (1) the reduction of graphene oxide (GO) and the nucleation of
nanocrystals attached onto the surface of rGO, and (2) the gradual nuclei growth
into cubic shapes under the protection of PVP. From HRTEM images and elemental
mapping shown in Fig. 4.5a–g, single-crystalline Pt-Pd nanocubes with shape
selectivity of 82% and an average size of 8.5 nm were uniformly distributed on
the rGO surface. More recently, using GO as both a support material and a
structure- and/or morphology-directing agent, rGO-supported PtPd concave
nanocubes have also been successfully synthesized through a simple hydrothermal
process [96]. In sharp contrast, only cubic PtPd alloy nanocrystals were obtained in
the absence of GO. The as-prepared PtPd concave nanocubes exhibited enhanced
electrocatalytic activity and durability toward methanol oxidation owing to the
exposed high-index facets of {730} and the strong interaction between the catalysts
and graphene support.
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 181

Fig. 4.5 (a–c) High-resolution TEM micrographs of the PtPd alloy nanocubes supported on rGO
at different magnifications. The inset in (c) shows the FFT pattern of an individual PtPd nano-
crystal. (d) The high-angle annular dark-field (HAADF)-STEM image of PtPd/rGO and the
corresponding elemental mapping of (e) Pt, (f) Pd, and (g) the overlay. Reprinted with permission
from [95]. Copyright 2013 American Chemical Society

4.2.2 Galvanic Replacement and Its Combination


with Chemical Reduction

Galvanic replacement is a simple and popular route for controllably constructing


various types of bimetallic nanocrystals. Compared with the co-chemical reduction
method, galvanic replacement is based on an etching process without using haz-
ardous reducing agents. Therefore, this technique is considered to be a green
method for nanocrystals preparation. A galvanic replacement reaction is driven
by the different electrochemical potentials between a sacrificial metal template and
another metal ion in a solution phase. Typically, this process involves oxidation and
dissolution of the template accompanied by reduction of another metal ions and
deposition of the resultant atoms on the surface of template. Moreover, the size and
morphology of the final product can be easily manipulated by using sacrificial
templates with different size and shape and/or by controlling the extent of replace-
ment reaction.
In recent years, galvanic replacement has been applied to the synthesis of
supported- and unsupported-PtPd bimetallic nanocrystals [97–100]. By using Ag
nanowires as sacrificial templates, Chen et al. [101] synthesized Pt-Pd nanotubes
(PtPd NTs) with an average diameter of 45 nm, wall thickness of 7 nm and length of
10 μm. In this synthesis, Ag nanowires were first prepared via a polyol method, and
the PtPd NTs were subsequently obtained by a galvanic replacement reaction
182 R. Zhang and W. Chen

between Ag nanowires and the mixture of Pt(CH3COO)2 and Pd(CH3COO)2 in an


aqueous solution. The as-prepared supportless-PtPd NTs exhibited enhanced cata-
lytic activity and much improved durability for ORR compared to Pt NTs and the
commercial Pt/C, which can be ascribed to their unique dimensions
(i.e. micrometer-sized length) and anisotropic morphologies.
In theory, galvanic replacement reaction can occur between any pair of metals
with appropriate difference in their redox potentials. In spite of the large difference
in redox potential between PtCl62−/Pt and PdCl42−/Pd, it was found that no obvious
galvanic replacement reaction occurs between Pd nanocrystals and PtCl62− ions
unless under special reaction conditions. Xia and co-workers [102, 103] reported
that the presence of Br− ions could promote the initiation of galvanic replacement
between PtCl62− ions and Pd nanocrystals. Interestingly, the Br−-induced galvanic
replacement exhibited a preferential selectivity towards the (100) facets of Pd
nanocrystals, resulting in the formation of Pt-Pd bimetallic nanocrystals with a
concave structure due to the simultaneous dissolution of Pd atoms from the (100)
facets and deposition of Pt atoms on the (111) facets. Figure 4.6 shows the typical
TEM images of Pt-Pd nanocrystals synthesized via galvanic replacement reaction
with different reaction times using Pd nanocubes as templates. As illustrated by the
schematic drawings in the insets, the Pt-Pd nanocrystals evolved from Pd truncated
cubes to Pt-Pd concave cubes and Pt-Pd octapods. The Pt-Pd nanocrystals can also
be manipulated in terms of both morphology and size by using Pd templates with
different sizes and shapes [102]. More interestingly, when a reducing agent of citric
acid (CA) was introduced into the above system, Pt-Pd bimetallic nanocages rather
than concave nanocrystals were generated [103]. In the process, it was suggested
that two important stages were included: (1) Br−-induced selective galvanic
replacement reaction between (100) facets of Pd nanocubes and PtCl42−, resulting
in the formation of Pt-Pd concave nanocubes; (2) co-reduction of Pd2+ ions evolved
from the galvanic replacement, together with PtCl42− ions remained in the solution
by CA into atoms and the subsequent deposition of these atoms on the side faces of
the concave nanocubes. Thus, Pt-Pd alloy nanocages with cubic morphology could
be easily obtained via a combination of Pd dissolution (related to galvanic replace-
ment) and the following Pt-Pd overgrowth (due to co-reduction).
In a recent study [104], our group also investigated the galvanic replacement
between Pd nanowires and PtCl62− in aqueous solution. PtPd alloy nanorods with a
porous structure were successfully synthesized through this bromide-induced gal-
vanic replacement route. The obtained PtPd nanorods showed much larger electro-
chemically accessible surface area compared with the Pd nanowires and
commercial Pt/C, making them promising for application in fuel cells as cathode
catalysts.
Despite great success achieved in the synthesis of Pt-Pd nanostructures via the
afore-mentioned strategies, a key procedure for the synthesis of Pd templates is
required. Zheng and co-workers [105] reported a one-pot fabrication of hollow
Pt-Pd nanocubes by using a mixed precursors of Pd(acac)2 and Pt(acac)2, with a
PVP and NaI solution in DMF. In this synthesis, because of the strong coordination
of I− ions to Pd2+ ions, the addition of I− ions into the mixture of Pd(acac)2 and
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 183

Fig. 4.6 TEM images of Pt-Pd nanocrystals in the form of nanocubes, concave nanocubes, and
octapods that were formed through bromide-induced galvanic replacement at various reaction
times: (a) 0.5, (b) 4, (c) 9, and (d) 20 h. The yellow, blue and red balls represent Pd atoms, Pt
atoms, and Br− ions, respectively. Reprinted with permission from [102]. Copyright 2011 Amer-
ican Chemical Society

Pt(acac)2 can generate the new dominating precursors of PdI4− and Pt(acac)2. In the
DMF solution, PdI4− is more favorably reduced to form Pd nanocubes, and the
galvanic replacement between temporal Pd nanocubes and Pt2+ species occurred
subsequently to produce hollow Pt/Pd nanocubes. In addition to the use of iodide
ions as the shape controller, the authors also demonstrated that acetylacetonate
precursors can alter the reduction kinetics of metal cations and thus control the
one-pot synthesis of Pt-Pd hollow nanocubes. Compared to solid Pt-Pd nanocubes,
the hollow Pt-Pd nanocubes with increased accessible surface area exhibited
improved catalytic activity towards formic acid electrooxidation.
184 R. Zhang and W. Chen

4.2.3 Seed-Mediated Growth

Morphological control of nanocrystals has become increasingly important, as many


of their physical and chemical properties are highly shape-dependent. As a conve-
nient and versatile synthesis method, seed-mediated growth is probably the most
powerful route for the synthesis of bimetallic heterostructures with controlled
morphology. In a seed-mediated growth process, a pre-synthesized seed of one
metal is significant to serve as initial sites for the nucleation and growth of a second
metal. During nucleation, the second metal can follow two different pathways by
homogeneous and heterogeneous nucleation. Because the activation energy for
nucleation on a pre-synthesized seed is prominently lower than that in a homoge-
neous nucleation process, the heterogeneous nucleation is always thermodynami-
cally more favorable than homogeneous nucleation as long as the seed-mediated
growth proceeds at a relatively slow rate under mild conditions, such as using weak
reducing agent and at low temperature.
For seed-mediated growth, there are three major growth modes corresponding to
the products with three distinct structures: (i) Frank-vander Merwe mode (or layer-
by-layer growth) for core-shell nanocrystals, (ii) Volmer-Weber mode (or island
growth) for hybrid structures, and (iii) Stranski-Krastanow mode (or island-on-
wetting-layer growth) for branched structures [106]. According to the previously
demonstrated results [107], the nucleation and growth mode of the second metal
over the seed metal are mainly manipulated by various physical parameters,
including lattice match, difference in bond dissociation energy and electronegativ-
ity between these two metals. As for Pt and Pd, due to the negligible lattice
mismatch (only 0.77%) and the lack of galvanic replacement unless under the
modified conditions (the presence of Br−, I− ions, etc.), the growth mode of Pd
atoms on preformed Pt metal seeds (or Pt atoms on Pd seeds) is mainly determined
by the bond dissociation energy.
Controllable synthesis of Pt-Pd bimetallic nanocrystals has been attracting
increasing attention due to their novel catalytic properties which are distinctly
different from those of their monometallic counterparts. Particularly, recent
advances reveal that Pt-Pd bimetallic nanoelectrocatalysts with a core-shell struc-
ture have been recognized as a promising alternative to commercial catalysts for
effectively improving the catalytic activity and durability for fuel cell application.
To this end, a lot of studies have been performed to synthesize high-efficiency
bimetallic Pt-Pd nanoelectrocatalysts with well-defined core-shell structures
[108–115]. By manipulating the reaction kinetics, the growth and nucleation of
Pt-Pd core-shell nanocrystals can be directed to a layer-by-layer epitaxial mode,
leading to the formation of Pt-Pd core-shell nanocrystals. Yang and coworkers
[116] developed a facile method for the synthesis of Pt-Pd core-shell cubes,
cuboctahedra and octahedra through epitaxial growth of Pd on Pt cubic seeds.
This process proceeded through reducing K2PdCl4 by ascorbic acid with Pt
nanocubes (9.5 nm in edge length) as seeds and tetradecyltrimethylammonium
bromide (TTAB) as a surfactant agent in an aqueous solution. Interestingly, in
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 185

Fig. 4.7 SEM (the first column), TEM (the second column), HAADF-STEM (the third column)
images of Pt-Pd core–shell nanocubes (a, b, c), cuboctahedra (e, f, g) and octahedral (i, j, k), and
the modeled orientation (the fourth column) of the core and shell of Pt-Pd nanocubes (d), Pt-Pd
cuboctahedra (h) and octahedral (l), respectively. Reprinted with permission from [116]. Copyright
2007 Nature Publishing Group

this synthesis, the authors found that by using different controlled facets of Pt cubic
nanocrystals as nucleation centers for the overgrowth of Pd metal, shape-controlled
Pt-Pd heterostructures can be obtained. The epitaxial growth of cubic Pd shells on
cubic Pt seeds along the (100) and (111) directions resulted in the formation of
Pt-Pd core-shell cubes. However, cuboctahedrally and octahedrally shaped Pd
shells were formed upon addition of increasing amount of NO2 which can alter
the growth rates along the (100) and (111) directions to produce Pt-Pd core-shell
cuboctahedron and octahedron. Both scanning electron microscopy (SEM)
(Fig. 4.7a, e, i) and the corresponding transmission electron microscopy (TEM)
(Fig. 4.7b, f, j) images clearly show the overall cubic, cuboctahedral and octahedral
morphologies and monodispersity of the Pt-Pd products. Moreover, from the
HAADF-STEM images, the Pt cores and the shaped Pd shells are discernible,
demonstrating the layer-by-layer epitaxial growth on Pt nanocubes. As illustrated
by the schematic drawings in Fig. 4.7d, h, l, the products evolved from Pt nanocubes
to Pt-Pd cubes, cuboctahedra and octahedra along the (100) and (111) directions
with different growth rates. This method and concept could also be used to
synthesize other metal nanostructures such as FePt, CoPt, with desirable
morphology.
186 R. Zhang and W. Chen

In another study, Wang and coworkers [117] prepared Pt-Pd petal-like nanotubes
via a wet-chemical strategy, in which the Pt nanotubes with petal-like surface was
first synthesized using ultrathin Te nanowires as sacrificial templates and an
effective epitaxial growth was further employed to deposit thin Pd nanoshells on
the novel Pt nanotubes. It was found that the thickness of the Pd nanoshells can be
easily controlled through the synthetic parameters (the amount of added Pd precur-
sor, etc.). The obtained one-dimensional bimetallic Pt-Pd nanotubes with small
diameter and nanometer-sized wall thickness demonstrated promising application
in fuel cells as effective electrocatalysts.
Since Pt is extremely rare and expensive, it has been shown that deposition of Pt
on Pd single-crystal surface can reduce the cost of materials while enhance their
catalytic activity. By reducing H2PtCl6 with citric acid (CA) in the presence of Pd
nanoplates as seeds and PVP as a stabilizing agent in an aqueous solution, Xia and
coworkers synthesized Pd-Pt core-shell nanoplates with hexagonal and triangular
shapes through layer-by-layer epitaxial growth of Pt on Pd nanoplates [118]. When
AA was used instead of CA in this process, Pt-Pd nanodendrities rather than Pt-Pd
core-shell structures were produced, indicating that the slow reduction rate associ-
ated with the weak reducing ability of CA played a vital role in achieving the
epitaxial growth of Pt shells on Pd nanoplates. Furthermore, Pd-Pt core-shell
structures with different shapes, such as regular octahedra, truncated octahedra
and cubes, could also be obtained from the epitaxial growth of Pt on well-defined
Pd nanocrystals [119]. The epitaxial growth of Pt shells on regular and truncated
octahedra of Pd at slow reducing resulted in the formation of Pd-Pt core-shell
octahedra. However, an incomplete octahedral Pt shell was formed when the Pd
cube was used as a seed.
Most recently, by using rGO as a support, Bai et al. [120] developed a unique
synthetic approach to prepare core-shell-like Pt-Pd-rGO stack structures. Two
important steps (Fig. 4.8a) were suggested to be involved in the synthesis: (i) In
situ growth of Pd nanocubes on rGO sheets via the co-reduction of K2PdCl4 and GO
nanosheets by using ascorbic acid as a reducing agent, and (ii) Pt shells were coated
onto the Pd nanocrystals by reducing H2PtCl6 in DMF. As shown in Fig. 4.8b, c,
cubic nanocrystals with an average shell thickness of about 2 nm are dispersed on
rGO sheets. HRTEM images in Fig. 4.8d show that the Pt shell is a single crystal
enclosed by (100) facets, forming a perfect interface with the Pd nanocrystal.
STEM and EDX mapping studies further confirm that Pt and Pd are enriched in
the shells and cores, respectively. Importantly, the thickness of the Pt shell in the
Pd-Pt-rGO stack structure can be controlled by simply changing the ratio of rGO-Pd
to Pt precursors. Moreover, Pt can be selectively deposited on Pd nanocubes rather
than on rGO sheets. In the synthesis of Pd-Pt-rGO structure, the potential difference
of rGO and Pd causes the electrons accumulation on Pd surfaces, and then Pt can be
preferentially reduced on the Pd surfaces with a relatively high electron density.
Meanwhile, the more negative potential of rGO (0.38 V vs. SHE) than Pd (0.62 V
vs. SHE) could provide a steady electron supply to prevent the oxidation of Pd in
the redox reactions. All these results clearly demonstrate that rGO played a key role
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 187

Fig. 4.8 (a) Schematic illustration of the synthesis of Pt-Pd-rGO structures. (b, c) TEM images of
the Pt-Pd-rGO structure at different magnifications. (d) HRTEM image of a Pt-Pd nanocrystal
supported on rGO. (e) STEM image and (f–h) EDS mapping profiles of a single Pt-Pd nanocrystal
on rGO: (f) Pd (green), (g) Pt (red), and (h) Pt-Pd-STEM overlay. Reprinted with permission from
[120]. Copyright 2014 Wiley-VCH

in manipulating the reaction kinetics to generate layer-by-layer epitaxial growth of


Pt on Pd nanocrystals.
The layer-by-layer growth strategy was also extended to prepare Pt-Pd multi-
shelled nanocrystals with alternating shells of Pt and Pd. Xia and coworkers [121]
demonstrated a facile method for the heteroepitaxial growth of Pt-Pd nanocrystals
with multi-shelled structures by sequentially adding Pt and Pd salt precursors into
an aqueous solution containing Pt or Pd seeds with CA as both capping and
reducing agents. Figure 4.9a shows a schematic of this synthesis starting from a
188 R. Zhang and W. Chen

a
K2PtCl4
Citric acid Na2PdCl4 K2PtCl4

90°C 90°C 90°C

b c

20 nm 20 nm

d e

20 nm 20 nm

Fig. 4.9 (a) Schematic illustration of layer-by-layer epitaxial growth of Pt-Pd multi-shelled
nanocrystals on a Pd cuboctahedral seed. (b) TEM image of Pd cuboctahedra of 9 nm in size
that serve as seeds for the overgrowth steps. (c–e) TEM images of Pd@Pt (c), Pd@Pt@Pd (d), and
Pd@Pt@Pd@Pt (e) nanocrystals prepared by reducing K2PtCl4 with citric acid (CA) as a reducing
agent in the presence of cuboctahedral seeds of Pd. The insets show TEM images of individual
nanocrystals at a higher magnification. Reprinted with permission from [121]. Copy right 2011
American Chemical Society

cuboctahedral Pd seed. The TEM images in Fig. 4.9b–e clearly shows a set of
products obtained at different stages of this heteroepitaxial growth process, with
the shape evolving from cuboctahedra to octahedra. The morphological transition
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 189

from cuboctahedra to octahedra derived from the preferential adsorption of CA on


(111) facets of Pt, resulting in a faster growth rate along the Pt(100) direction than
the (111) facets. The thickness of Pd and Pt shells could be independently manip-
ulated by varying the amounts of Pd and Pt salt precursors added in the reaction
solution. The core-shell nanocrystals with Pt and Pd shells can be repeated more
times to generate larger and more complex Pd-Pt multi-shell nanocrystals. In
addition to the use of Pd cuboctahedra as seeds for the alternating deposition of
Pt and Pd shells, the authors found that Pd octahedra/plates and Pt cubes can also be
employed as seeds to prepare Pt-Pd multi-shelled nanocrystals with other different
shapes. For instance, starting from seeds of cubic Pt nanocrystals, Pt-Pd multi-shell
nanocrystals composing of alternating Pd and Pt shells were also obtained, except
for a morphological transition from cubes to octahedra owing to extensive over-
growth. All of these studies clearly demonstrated that the use of CA as both
reducing and capping agents is the key to the successful synthesis of the multi-
shelled nanocrystals. As a relatively weak reducing agent, CA can produce Pt and
Pd atoms at the right place to ensure layer-by-layer epitaxial growth of both metals.
The synthetic technique presented here can be used to prepare multi-shelled
nanostructures with other compositions and morphologies for various applications.
Except for the afore-mentioned layer-by-layer growth mode, the island-on-
wetting-layer mode is another preferred process for the growth of Pd on Pt seeds
due to the following order in bond dissociation energy: EPt-Pt (307 kJ/mol)>EPt-Pd
(191 kJ/mol)>EPd-Pd (136 kJ/mol), resulting in the formation of Pt-Pd branched
structures. Among various Pt-Pd heteronanostructures, Pt-on-Pd nanodendrites
with highly branched shapes have received great interest because of their unique
properties originating from the electronic coupling between the metals and a wide
variety of promising applications in catalysis. Xia and coworkers [122] developed a
facile, seed-mediated approach to the synthesis of Pt-Pd nanodendrites consisting of
a dense array of Pt branches on a Pd nanocrystal core. In the synthesis, Pd truncated
octahedra with an average size of 9 nm were used as seeds to direct the dendritic
growth of Pt upon the reduction of K2PtCl4 by L-ascorbic acid (AA) in an aqueous
solution containing PVP. It was proposed that the high initial supersaturation of Pt
atoms associated with fast reduction by AA was probably responsible for the
branched growth of Pt. In the reaction system, once Pt has nucleated on the surface
of a Pd nanocrystal upon fast reduction by AA, the Pt nuclei can serve as catalytic
sites for further reduction of the Pt precursor and create favorable sites for atomic
addition. Growth preferentially occurs on the Pt nuclei, and deposition proceeds
along the developing Pt branches. And the spatially separated Pt branches structures
could be generated due to the multiple nucleation sites provided by truncated
octahedral Pd seeds. The TEM images of the product showed that numerous Pt
branches have grown from each Pd core, resulting in the formation of Pt-Pd
nanodendrites. HRTEM characterizations clearly revealed the continuous lattice
fringes from the Pd core to the Pt branches, further indicating the Pt branches were
epitaxially nucleated and grown on the Pd seeds. The authors also found that the
generated product has a high surface area and particularly active facets, providing a
promising application in fuel cells. Furthermore, by using differently shaped Pd
190 R. Zhang and W. Chen

seeds and mediated growth mechanism, other Pt-Pd dendritic nanostructures have
also been generated [123, 124]. For example, Wang and coworkers [125] synthe-
sized ultralong Pt-Pd bimetallic nanowires with a 100% yield by employing Pd
nanowires as seeds to direct the dendritic growth of Pt upon the reduction of
K2PtCl4 by AA in aqueous solution. Interestingly, the as-prepared Pt-Pd nanowires
have the cores of Pd nanowires and shells of dendritic Pt, and the small single-
crystal Pt nanobranches interweave with each other, resulting in nanopores on the
surface of Pd nanowires. Due to the unique nanostructure, the synthesized Pt-Pd
nanowires exhibited a high surface area and enhanced electrocatalytic activity
towards methanol oxidation reaction.
In another study, Yang et al. [126] demonstrated an oil-phase synthetic approach
for the synthesis of Pt-Pd branched nanostructures by reducing Pt(acac)2 in a
mixture of diphenyl ether and oleylamine with 5 nm Pd nanoparticles as seeds
under an argon atmosphere at 180 ˚C. As shown in Fig. 4.10a, the branched arms of
Pt with an average diameter of 3 nm are distributed evenly on the surface of Pd

Fig. 4.10 Representative (a) TEM, (b) HRTEM (c) HAADF-STEM image and (d, e) EDX
mapping of Pd-Pt bimetallic nanodendrites synthesized by reducing Pt(acac)2 in a mixture of
diphenyl ether and oleyamine in the presence of preformed Pd nanoparticles. Reprinted with
permission from [126]. Copyright 2009 American Chemical Society
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 191

nanoparticles, generating Pt-Pd nanodendrite structures. From the HRTEM image


(Fig. 4.10b), the synthesized nanoparticles show well-defined crystalline fringes
and the Pt branches grew along the (111) crystal planes on the Pd seed. From the
representative STEM and the corresponding EDX maps (Fig. 4.10c–e), Pt is
dispersed throughout the entire particle, including the branches, whereas Pd could
only be detected in the core region, indicating the formation of Pt-Pd dendritic
structures.
Graphene nanosheets have been studied extensively owing to its unique elec-
tronic, thermal, mechanical properties arising from its strictly 2D structure. Partic-
ularly, its unique structure and the resulting properties endow it to be a promising
2D supporting material to load metal nanoparticles for application in fuel cells. By
using graphene sheets as support, Wang and coworkers [127] constructed high-
quality 3D Pt-on-Pd bimetallic nanodendrites supported on PVP-functionalized
graphene nanosheets. In the synthetic process, the PVP-functionalized graphene
was first obtained under the reduction of hydrazine. Pd/graphene seeds were then
synthesized using HCOOH as reducing agent at room temperature. Finally, Pt-on-
Pd nanodendrites supported on graphene sheets were produced by using graphene
sheets-supported Pd nanoparticles as seeds to direct the dendritic growth of Pt upon
the reduction of K2PtCl4 by ascorbic acid in an aqueous solution. The TEM results
indicated that the Pt-on-Pd bimetallic nanodendrites with an average size of 15 nm
were dispersed on graphene sheets, in which Pt branches with an average diameter
of about 3–5 nm were distributed evenly on the surface of a Pd nanoparticle.
Importantly, the number of Pt branches could be easily controlled through simply
manipulating the reaction parameters. For example, lower concentrations of
graphene can lead to relatively lower amounts of Pd nanoparticles adsorbed on
the surface of graphene nanosheets, thus resulting in Pt-on-Pd bimetallic
nanodentrites with more Pt branches supported on the graphene sheets. More
importantly, because the small single-crystal Pt nanobranches with porous structure
and good dispersion were directly grown onto the surface of graphene nanosheets,
the obtained hybrids exhibited an enlarged electrochemical surface area as high as
81.6 m2/g. All these unique structural features together with the synergetic effects
of the Pt-Pd and the enhanced electron transfer stemming from graphene support
are highly favorable for the application of the graphene nanosheets-supported 3D
Pt-on-Pd bimetallic nanodendrites in fuel cells, with a much higher catalytic
activity than conventional E-TEK Pt/C electrocatalysts for methanol electro-
oxidation.
In spite of the significant achievement in the synthesis of Pt-on-Pd nanodendritic
structures via the aforementioned seed-mediated growth strategy, the involved
processes of the syntheses were strongly dependent on the use of faceted Pd
seeds to direct the subsequent growth of Pt branches. Without the well-defined Pd
seeds, both the particle size and shape of the Pt-on-Pd nanodendrites are usually
uncontrollable. Recently, without using any pre-synthesized Pd seeds, organic
solvent and high temperature, Wang et al. [128] proposed a simple approach for
one-step direct synthesis of Pt-on-Pd nanodendrites in aqueous solution at room
temperature. In the synthesis, a block copolymer (Pluronic P123) was employed to
192 R. Zhang and W. Chen

mediate the reduction of K2PtCl4 and Na2PdCl4 by using AA as a reducing agent for
30 min at room temperature. It was found that the as-prepared product consists of
well-dispersed nanodendrites with Pd interior and dendritic Pt exterior, in which the
Pt nanoarms have widths of 3 nm branching in various directions. The preferential
reduction of the Pd precursor by AA caused by the different reduction kinetics of Pd
and Pt complex with AA was found to a key factor in directing the formation of Pt-
on-Pd nanodendrites. The formed Pd nanoparticles can serve as in situ seeds for the
subsequent deposition of Pt. Moreover, the use of Pluronic P123 and the selected
concentration (0.87 mM) also played critical roles in the synthesis of Pt-on-Pd
nanodendrites. In comparison, aggregated Pt-Pd nanoclusters were obtained with an
increase/decrease of the Pluronic P123 concentration or replacing Pluronic P123
with Pluronic F68. A surfactant of PVP was found to be unfavorable for the
formation of Pt-on-Pd nanodendrites. Most importantly, by a simple control of
the mole ratios of the Pt and Pd precursors in the reaction solution, Pt-on-Pd
nanodendrites with a designed Pt and Pd content could be obtained. In comparison
with the two-step seed-mediated methods, this rational block copolymer-mediated
synthesis could trigger the facile creation of novel bimetallic heterostructures with
designed compositions and desired properties.
In addition to seed-mediated growth involving direct reduction of a second metal
onto pre-formed seeds, Beer and co-workers [129] recently reported a novel anion
coordination route to control the formation of bimetallic core-shell nanoparticles
for any two noble-metals including Pt and Pd. This method uses ligand-based
supramolecular forces to ensure surface segregation of the shell metal onto the
pre-formed core before its reduction. And four different types of bimetallic core-
shell nanoparticles (Au-Pd, Pd-Au, Pt-Pd and Pd-Pt) with an average size of less
than 5 nm have been synthesized by using this new protocol. The success of this
synthesis was based on the ability to anchor metal ions to the pre-formed seeds
through amides-chlorometallates anion coordination by hydrogen bonding before
reduction occurred. This work not only provides novel core-shell nanoparticles with
small size (<5 nm), but also offers an impetus for the exploitation of supramolec-
ular interactions in the design and synthesis of structured nanoparticles with
controlled composition.

4.2.4 Electrochemical Deposition and Its Combination


with Galvanic Replacement

In recent years, to improve the mass activity of shaped nanocrystals, much work has
focused on the crystallographic orientation of metal atoms at the surface of
nanocrystals [30, 130]. Owing to the simplicity and no need of templates, electro-
chemical deposition is a useful approach to prepare Pt-Pd bimetallic nanocrystals
with decorated surface by Pt or Pd adatoms and consequently with enhanced
electrochemical properties [131–144]. For example, Feliu and coworkers [145]
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 193

have successfully electrodeposited Pd adatom on cubic Pt nanoparticles as an anode


electrocatalyst toward formic acid oxidation in fuel cells. The authors proposed that
the amounts of Pd on the Pt surfaces can be easily monitored in situ by observing
the voltammetric changes during the deposition process. Compared with the
Pd-modified quasispherical Pt nanoparticle, the Pd adatoms-decorated Pt(100)
nanoparticles exhibited enhanced catalytic activity for formic acid electrooxidation.
In the Pd adatoms-decorated Pt(100) nanoparticles, the high fraction of Pt(100)
sites can decrease surface poising, lower onset potential and thus greatly improve
the kinetics of formic acid oxidation over Pt surface.
In addition, the nanostructure of Pd nanocrystals covered by a Pt monolayer shell
has also attracted much attention because in such structure all Pt atoms located on
the surface of Pd core can be sufficiently utilized and thus improved Pt mass activity
can be achieved. DFT calculations demonstrated that the strong interaction between
the Pt monolayer and the base metal also plays a vital role in determining the
structure and properties of the Pt monolayer. In this case, the electrocatalytic
properties of the Pt monolayer can be easily manipulated by changing the base
metal. For example, the electrocatalytic activity of a Pt monolayer on different base
metals for ORR exhibited a volcano-type dependence [146]. Pd substrate was
thought not only to apply a compressive strain upon the Pt monolayer but also to
impart a so-called “ligand-effect”, leading to lowered energy of the weighted center
of the Pt d-band. Therefore, the Pd(111) decorated with Pt monolayer is located at
the top of the volcano curve with the highest ORR activity. The extremely low Pt
loading together with a perfect catalytic activity make the Pt-monolayer decorated
Pd nanocrystals very attractive for practical applications in fuel cells.
Recently, much work has been done to prepare Pt monolayer-decorated Pd
nanocrystals with improved electrocatalytic properties [147–150]. For example,
Adzic and coworkers [151] reported a simple approach for the synthesis of a Pt
monolayer on a Pd base metal by first electrochemically depositing a monolayer of
Cu atoms on Pd cores through Cu underpotential deposition (UPD), followed by
the controlled displacement of these adatoms with Pt via galvanic exchange.
Figure 4.11a schematically illustrates the key steps for the growth of a Pt
monolayer on a Pd core. The Z-contrast image from HAADF-STEM (Fig. 4.11b)
demonstrates the formation of a bright shell on the relatively darker nanoparticle
core. The EDX line-scan analysis (Fig. 4.11c) shows that the Pt trace has two peaks
in two sides, while the Pd trace has one peak in the center, further confirming the
formation of a Pt monolayer on the Pd core. In addition, the thickness of the Pt shell
could be gradually increased with repeated UPD and galvanic replacement
processes.
Although outstanding ORR activities have been achieved with the afore-
mentioned 0D core-shell catalysts, 1D nanostructures are characterized by their
uniquely anisotropic nature, which imparts advantageous structural and electronic
factors in the catalytic reduction of oxygen. In particular, Koenigsmann et al. [152]
designed a Pd nanowire core-Pt monolayer shell structure with enhanced
electrocatalytic activity and durability by successfully combining the uniquely
advantageous core-shell motif with the electrocatalytic advantages of ultrathin 1D
nanostuctures. In their synthesis, the ultrathin Pd nanowires (~2 nm) were first
194 R. Zhang and W. Chen

Fig. 4.11 (a) A schematic of the two major steps involved in the synthesis of a Pt monolayer on a
Pd core through a combination of electrochemical deposition and galvanic replacement.
(b) HAADF-STEM image and (c) the EDX line-scanning profile showing the formation of a
Pt monolayer on a Pd nanoparticle. Reprinted with permission from [151]. Copyright 2010
Wiley-VCH

synthesized, and the deposition of Pt onto the surface of the treated Pd nanowires
was achieved by Cu UPD followed by galvanic displacement of the Cu adatoms
with Pt2+. Importantly, the use of ultrathin nanowires with a diameter below 5 nm in
this work could maximize the surface area-to-volume ratio, achieving higher mass
activity compared with conventional commercial Pt nanoparticles, and core-shell
nanoparticles. Moreover, in the obtained Pd nanowire core-Pt monolayer shell
structure, a contraction of the core nanowire surface would be highly advantageous
for ORR electrocatalysis, because this would enhance the strain-induced contrac-
tion of the Pt monolayer and therefore improve the inherent ORR activity. In
addition, by varying the shape of the Pd cores (Pd cubes, octahedra or rods), the
morphology-tailored core-shell nanoparticles could also be obtained by the same
way [150, 153, 154].
In addition to the methods listed above, the synthesis of Pt-Pd (or supported
Pt-Pd) bimetallic nanocrystals can also be achieved by other techniques, such as
thermal treatment [155–160], plasma sputtering [161, 162], electroless deposition
[163, 164], and supercritical CO2 deposition [165, 166].
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 195

4.3 Applications of Unsupported/Supported Pt-Pd


Bimetallic Nanocrystals as Electrocatalysts in Fuel
Cells

4.3.1 Alcohol Oxidation Reaction

Among the various fuel cell technologies, direct methanol fuel cells (DMFCs) have
attracted special interests because of their ability to utilize methanol as a liquid fuel,
which can be easily and safely stored as well as transport comparing with hydrogen-
based counterparts [167, 168]. Moreover, methanol can be directly prepared either
from natural gas or renewable biomass and thus ideally meet the future power needs
with a high energy efficiency of 600 Wh/kg [169]. More importantly, DMFCs have
the unique advantage of operation at near ambient temperatures between 40 and
80 ˚C, which is much lower than the normal operating temperatures for solid oxide
fuel cells (800–1000 ˚C). Therefore, DMFCs technology represents a potentially
effective fuel cell candidate for future applications in the automotive, portable
power generating, and electronics industries [170]. However, it is significant to
note that large portions of the cost of DMFCs can be attributed to the high loading
of expensive electrocatalysts that are at the heart of the device itself. Meanwhile,
the slow anode reaction and methanol crossover reaction at the cathode are also the
vital limitation to the widespread application of DMFCs. Hence, lowering the costs
and improving efficiency of the catalysts have become two critical technological
issues towards the development of practical and inexpensive DMFCs.
The process of methanol oxidation reaction (MOR) at the anode in DMFCs
includes the methanol adsorption and the subsequent dissociation into adsorbed
intermediates [171]. According to the dual-pathway mechanism, CO is a poisoning
intermediate species, which can largely reduce the catalytic activity of catalysts,
especially Pt-based catalysts. To eliminate the CO poisoning to catalysts, oxygen-
containing surface species (e.g, OH) formed on adjacent catalyst sites are usually
needed to remove CO adsorbed on the catalyst surface (COad). Thus, to catalyze the
methanol oxidation efficiently, catalysts with multiple active sites are required for
the adsorption of methanol and formation of OH species. Therefore, a significant
amount of work has been dedicated to the synthesis of Pt-based nanostructures
combining with another metal so as to improve CO-tolerance. Among these, Pt-Pd
bimetallic nanocrystals represent an active and durable catalysts for methanol
oxidation.
Effectively controlling the size and shape of Pt-Pd bimetallic nanocrystals can
provide a great opportunity to improve their catalytic properties and increase their
mass or specific activity [57, 100, 110]. Huang and coworkers [115] employed a
seed-mediated growth method to synthesize Pt-Pd core-shell nanocrystals with
different Pd shell thicknesses. The catalytic activities of the different sizes
(i.e. Pd shell thickness) of Pt-Pd nanocrystals at Pd/Pt ratios from 1/6 to 2/3 for
MOR in alkaline media were compared in CV measurements. It was found that the
Pt-Pd core-shell nanocrystals with a Pd/Pt ratio of 1/3 (near monolayer Pd shell)
196 R. Zhang and W. Chen

yield the highest current density and the most negative potential for the oxidation
peak in the forward sweep due to the highest tolerance of the sample to CO
poisoning. Such results demonstrate the importance of Pd shell thickness of the
core-shell nanocrystals in the manipulation of the catalytic performance for fuel cell
applications.
To enhance the mass activity of Pt, the core-shell type of nanocrystal catalysts
with a Pt shell have also been developed to remarkably reduce Pt loading.
Yamauchi and coworkers [128] synthesized Pt-on-Pd nanodendrites and studied
their catalytic properties for methanol oxidation reaction in acid condition. The
electrochemical measurements showed that the forward peak current density of
MOR on Pt-on-Pd (0.49 A/mgPt) is much higher than that of Pt nanodendrites
(0.21 A/mgPt) and Pt black (0.11 A/mgPt), and at any oxidation current density, the
corresponding oxidation potentials on the Pt-on-Pd nanodendrites are obviously
lower than those on the Pt dendrites and Pt black, indicating that Pt-on-Pd
nanodendrites has the highest activity for methanol oxidation. It was found that
the formation of the inserted pseudo-Pd-Pt alloy heterointerface plays a critical role
in reducing the electronic binding energy in Pt and facilitating the C-H cleavage
reaction in methanol decomposition. Furthermore, the various atomic steps exposed
on the Pt branch surface can act as highly active sites for the methanol oxidation
reaction. Therefore, superior catalytic activity was exhibited through this open
dendritic structure with the designed Pt and Pd ratios. In addition, nanostructures
with high aspect ratios such as nanowires, nanotubes and nanorods can provide
improved mass transport and higher electrochemical active surface area than those
of low aspect-ratio nanoparticles [172]. For example, Guo et al. [125]. investigated
the activity of the ultralong Pt-Pd bimetallic nanowires for methanol oxidation
reaction. By comparing with other catalysts, the Pt-Pd nanowires exhibited much
larger electrochemical surface area (ECSA) (90.7 m2/g) than those of E-TEK Pt/C
catalyst (53.5 m2/g) and mesoporous Pt with giant mesocages (74 m2/g). Compared
to the commercial E-TEK catalyst, a significant enhancement of the peak current
and a negative shift of the onset potential of methanol oxidation were achieved on
the Pt-Pd nanowires. Moreover, it should be noted that the Pt-Pd nanowires also
exhibited much higher mass activity and stability than some state-of-art Pt-based
nanomaterials.
It has been well-documented that the catalytic activities of bimetallic
nanocrystals are strongly dependent on their compositions [139, 173]. Sun and
co-workers [89] studied polyhedral Pt-Pd bimetallic nanoparticles with the compo-
sitions ranging from Pd88Pt12 to Pd34Pt66 as anode catalysts for methanol oxidation
reaction. As shown in Fig. 4.12a, as compared to pure Pd nanoparticles, both peak
potentials and current densities of MOR change with the Pt content in the Pt-Pd
nanoparticles. The plots of peak potentials and peak current densities versus Pt
atomic % in the nanoparticles (Fig. 4.12b) showed a composition-dependent MOR
activity and the catalyst having 40–60% Pt exhibited the maximum activity plateau.
Cai and coworkers [82] synthesized three-dimensional Pt-Pd core-shell
nanostructures by using a one-step microwave heating method. It was found that
the Pt-Pd nanostructure with a Pd/Pt molar ratio of 1:3 exhibited the highest
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 197

Fig. 4.12 (a) J-V curves of MOR on Pd, BASF Pt and Pd-Pt catalysts with different compositions
in 0.1 M HClO4 and 0.1 M methanol. (b) Methanol oxidation peak current density and peak
potential vs the Pt atomic % in the Pt-Pd nanoparticles. Reprinted with permission from
[89]. Copyright 2011 American Chemical Society

electrocatalytic activity toward methanol oxidation as compared to Pt, Pd and other


Pt-Pd nanostructures. Moreover, the Pt-Pd nanostructure with a Pd/Pt molar ratio of
1:3 also exhibited high stability due to its enhanced tolerance to intermediate
species during the methanol oxidation. Jin and coworkers [88] investigated the
catalytic activity of the assembled free-standing Pt-Pd nanomembranes (PdPt-
FNMs) for methanol oxidation in alkaline condition. By using the hydrogen
adsorption-desorption method, the PdPt-FNMs have much higher ECSA values
(Pd53Pt47–FNMs (48.01 m2/g), Pd33Pt67–FNMs (52.04 m2/g), Pd15Pt85–FNMs
(50.55 m2/g)) than commercial Pt black (22.94 m2/g), indicating the enhanced
active sites on PdPt–FNMs for the electrooxidation of methanol. By comparing
the CVs of methanol oxidation, the Pd33Pt67–FNMs exhibited the best catalytic
performance with the most negative onset potential and the largest current density
among the PdPt-FNMs with different compositions and commercial Pt black.
Moreover, the PtPd-FNMs showed higher stability than commercial Pt black after
CV cycling tests, and the Pd33Pt67–FNMs exhibited the highest stability with the
least loss of the electrocatalytic activity among the studied PdPt-FNMs. The
enhanced MOR activity of all the above-mentioned Pt-Pd bimetallic nanocrystals
can be explained by the bifunctional methanol oxidation mechanism. In the Pt-Pd
nanocrystals, Pd is mainly responsible for the water dehydrogenation to form
Pd-OH, while Pt catalyzes the methanol dehydrogenation to form Pt-CO. The
reaction between Pd-OH and Pt-CO produces CO2 to regain the active metal
surfaces. However, without Pd, the water dehydrogenation on Pt occurs at a higher
potential, making the oxidation process sluggish. The activity also decreases with
the excessive Pd due to the lack of Pt for methanol dehydrogenation.
In the synthesis, metal nanoparticles are usually formed with certain facets to
minimize surface energy and the total free energy. On different crystal surfaces of
Pt-Pd materials, the electronic structures and atomic arrangements are quite differ-
ent, and the appropriate crystal phase is able to greatly enhance reaction kinetics
198 R. Zhang and W. Chen

Fig. 4.13 (a) Stable CV curves obtained from the Pt-Pd NCs, NTs, and Pt/C in 0.1 M HClO4 and
1 M CH3OH at a sweep rate of 50 mV/s. (b) CV curves obtained from different electrocatalysts
after 4000 additional cycles. Reprinted with permission from [94]. Copyright 2011 American
Chemical Society

[113]. Yin et al. [94] have successfully prepared the mono-disperse single-crystal-
line sub-10 nm Pt-Pd nanotetrahedra (NTs) and nanocubes (NCs). The single-
crystalline Pt-Pd NTs are enclosed by four (111) facets, while the Pt-Pd NCs are
enclosed by (100) facets. As shown in Fig. 4.13a, both the Pt-Pd NCs and NTs
exhibited high catalytic activities towards methanol electrooxidation in acid elec-
trolyte. As compared the peak voltage (Ef ¼0.85 V) and the peak current density
(Jf ¼0.51 mA/cm2) for Pt/C, more negative Ef and much larger Jf in a forward scan
were obtained with 0.85 V and 1.49 mA/cm2 for Pt-Pd NCs, and with 0.84 V and
1.12 mA/cm2 for Pt-Pd NTs, respectively. Moreover, the ratio of the current density
values in two sequential forward (positive) and backward (negative) sweeps (Jf/Jb)
is considered to be an important indicator of the catalyst tolerance to poisoning
species. The different Jf/Jb values for Pt-Pd NCs (2.5) and NTs (1.4) confirmed that
different reaction pathways might be adopted on the (100) or (111) surfaces.
Meanwhile, due to the more durable nature of the (111) facet of Pt-based
nanocrystals, the Pt-Pd NTs remained higher activities than both Pt-Pd NCs and
Pt/C after repeating the CV sweeps for over 4000 cycles (Fig. 4.13b). The different
electrocatalytic performances of the Pt-Pd NCs and NTs demonstrate the facet-
sensitive nature of methanol oxidation on Pt-Pd nanocrystals.
In another study, Lee et al. [72] synthesized octahedral Pt-Pd nanoparticles with
exposed (111) facets as anode catalysts for methanol oxidation. The octahedral
Pt-Pd alloy with dominantly exposed (111) facets exhibited an enhanced catalytic
performance with lower onset and peak potentials, and higher current density as
compared to polycrystalline Pt/C for MOR. Moreover, a higher ratio of the forward
to the reverse anodic peak current density than that of commercial Pt/C indicates
less accumulation of residues on the Pt-Pd octahedron during the forward anodic
scan. In addition, the remained size and morphology of the well-defined Pt-Pd alloy
octahedron compared to the massive agglomeration of Pt/C after the stability test
further confirmed the superior electrocatalytic activity and stability of the Pt-Pd
bimetallic nanocrystals for MOR. All these studies indicate that shape-controlled
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 199

synthesis of facet-sensitive multi-metal nanocrystals can open up new opportunities


for Pt-Pd nanocrystals with potential applications.
To further improve the catalytic activity of the Pt-Pd bimetallic nanocrystals and
lower the usage of noble metals, loading of catalysts on the surface of suitable
supporting materials is highly desirable [43, 54, 81, 96, 174, 175]. Wang and
coworkers [127] investigated the catalytic activity of three-dimensional Pt-on-Pd
bimetallic nanodendrites supported on graphene nanosheets (TP-BNGN). By com-
paring with the carbon nanofiber- or CNT-supported Pt nanoparticles, CNT/Pt
composite and commercial E-TEK Pt/C catalysts, the TP-BNGN exhibited much
improved catalytic activity for MOR with much higher mass current density, more
negative onset oxidation potential and more efficient removal of the poisoning
species on the catalyst surface. Recently, we studied the catalytic activities of
rGO-supported Pt-Pd nanocubes (PtPd/rGO) for methanol oxidation in acid
medium [95]. As shown in Fig. 4.14, compared with the unsupported PtPd alloy
nanocubes and commercial Pt/C catalysts, the PtPd/rGO showed enhanced

20 250
a Unsupported PtPd nanocubes b
PtPd/RGO 200
J (mA/mgmetal)
J (mA/mgmetal)

10 Pt/C
150
0
100 -0.2 0.0 0.2 0.4 0.6 0.8 1.0
E (V vs Ag/AgCl)
-10
50

-20 0

-0.2 0.0 0.2 0.4 0.6 0.8 1.0 -0.2 0.0 0.2 0.4 0.6 0.8 1.0

E (V vs Ag/AgCl) E (V vs Ag/AgCl) Specific activity (mA/cm )


Mass activity (mA/mgPt)

250 300
c d
Unsupported PtPd nanocubes 2.5
PtPd/RGO 250 Mass activity
J (mA/mgmetal)

200
PtPd/RGO Specific activity
200 2.0
150
150 1.5
100
100 1.0
50
50
0.5
0 0
0 200 400 600 800 1000
PtPd/RGO PtPd cubes Pt/C
Time (s)

Fig. 4.14 CVs of the unsupported and rGO-supported (PtPd/rGO) PtPd alloy nanocubes, and the
commercial Pt/C catalysts in (a) 0.1 M HClO4 solution, and (b) 0.1 M HClO4 +1.0 M CH3OH
solution. (c) Chronoamperometric curves of methanol oxidation at 0.62 V in 0.1 M HClO4 +1.0 M
CH3OH solution after the CO stripping treatment. Potential scan rate 50 mV/s. All currents were
normalized to the total mass of noble metals (Pt and Pd). (d) Comparison of mass and specific
activities of the three catalysts for methanol oxidation. Reprinted with permission from [95]. Copy-
right 2013 American Chemical Society
200 R. Zhang and W. Chen

electrocatalytic performance with increased ECSA, lower negative onset potential


and higher current density for methanol oxidation. Furthermore, the PtPd/rGO
exhibited high durability during the methanol oxidation. All these studies indicate
that graphene sheets can be used as excellent catalyst support for enhancing the
catalytic performance and improving the stability of Pt-Pd bimetallic catalysts.
With a high surface area, good electrical conductivity, light weight, high chem-
ical stability and inherent size and hollow geometry, carbon nanotubes (CNTs) are
one of the most attractive nanomaterials used as an efficient catalyst support in fuel
cells. Wai and coworkers [166] studied the catalytic activity of Pt-Pd nanoparticles
supported on multi-walled carbon nanotubes (MWCNT) for methanol electro-
oxidation. From the CVs of methanol oxidation, compared to MWCNT-supported
monometallic nanoparticles, the Pt-Pd/MWCNTs exhibited much lower onset
potential and higher ratio of the forward scan peak current over the backward
scan peak current, indicating the Pt-Pd/MWCNTs catalysts are more efficient at
lowering the adsorbed carbon monoxide and thus show an enhanced catalytic
activity for MOR.
In addition, conducting polymer nanostructures have recently received special
attention as supporting nanomaterials owing to their highly π-conjugated polymeric
chains, unique electrical properties, and controlled morphologies. Recently, Wang
and coworkers [76] reported a 1D nanoelectrocatalyst based on polyaniline (PANI)
nanofiber-supported Pt-Pd hybrid nanoparticles for methanol oxidation reaction.
From the CV measurements, the current density of MOR on the PANI nanofiber-
supported Pt-Pd nanoparticles catalyst was 2.99 times higher than that of E-TEK
Pt/C catalyst for MOR. And the hybrid catalyst exhibited a higher ratio (1.98) of the
forward oxidation current peak to the reverse current peak compared to commercial
E-TEK Pt/C (1.08), revealing the more effective removal of the poisoning species
on the surface of PANI nanofiber-supported Pt-Pd nanoparticles. The enhanced
catalytic activity of PANI nanofiber-supported Pt-Pd can be ascribed to the high
conductivity of PANI support in acidic conditions and the supra-high density
distribution of small Pt-Pd nanoparticles on the surface of PANI nanofibers. In
another work, highly dispersed Pt-Pd nanoparticles supported on polypyrrole (Ppy)
film were fabricated by electrochemical deposition method [137]. Because of the
uniform dispersion of Pt-Pd nanoparticles in the Ppy film, the synergistic effect of
the Pt and Pd and the effective reduction of CO species, the Ppy-supported Pt-Pd
nanoparticles showed remarkably improved electrocatalytic performance for meth-
anol oxidation reaction. Galal et al. [140] synthesized Pt-Pd nanoparticles
supported on poly(3-methylthiophene) (PMT) films via electro-deposition method.
The voltammetric measurements indicated a higher activity of PMT-supported
Pt-Pd nanoparticles than other types of deposited substrates such as Pt and glassy
carbon, for methanol oxidation. Moreover, it was also found that the thickness of
PTM in the Pt-Pd/PTM hybrid has an obvious effect on the catalytic performance,
and the best activity was obtained at the relatively thinner films.
Direct ethanol fuel cells (DEFCs) have also attracted enormous attention due
to the higher theoretical energy density but less toxicity of ethanol as compared to
methanol. Moreover, ethanol can be synthesized directly from biomass, whereas
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 201

methanol is most commonly generated from natural gas, an abundant but


non-renewable source [176]. For ethanol oxidation on Pt-based electrocatalysts,
several reaction mechanisms have been proposed with complex processes
[177, 178]. According to the mechanisms, it is widely accepted that the electro-
oxidation of ethanol can only occur with a multifunctional electrocatalyst. Platinum
itself is known to be rapidly poisoned on its surface by strongly adsorbed species
coming from the dissociative adsorption of the ethanol molecule. So other metals
modified with Pt can activate water at low potentials to produce oxygenated species
(e.g. adsorbed OH), and these OH species are necessary to oxidize the species from
the dissociation of ethanol to carbon dioxide. Moreover, unlike methanol oxidation,
the oxidation of ethanol is more difficult with the necessity to break the C-C bond to
obtain its complete oxidation. Except for the formation of adsorbed CO, acetalde-
hyde, acetic acid, or CO2 can be formed through different reaction pathways during
ethanol electro-oxidation reaction. In the past years, Pt-Pd bimetallic nanocrystals
have been recognized as a promising alternative to commercial catalysts for
effectively improving the activity and durability for ethanol electro-oxidation
[64, 142, 155]. For example, Guo et al. [117] synthesized Pt-Pd bimetallic
nanotubes with petal-like surface and further studied their catalytic activity for
ethanol oxidation in alkaline media. Figure 4.15a compares the CVs of Pt-Pd
bimetallic nanotubes with petal-like surfaces (PBNPSs, curve a), Pt nanotubes
with petal-like surfaces (PNPSs, curve b) and E-TEK Pd/C (curve c) electrodes in
a N2-saturated 0.5 M H2SO4 solution. It can be seen that the reduction peak
potential of Pd(OH)2 shifts positively on the PBPNSs (0.48 V vs. Ag/AgCl) as
compared with E-TEK Pd/C (0.38 V vs. Ag/AgCl), indicating the reduced
oxophilicity and weakened chemisorpotion with oxygen-containing species on
PBNPSs catalyst. Furthermore, from the CVs of ethanol oxidation shown in
Fig. 4.15b (specific activity), c (mass activity), the PBNPSs (curve a) exhibit
much higher specific and mass activities, which are about 3.7 and 3.1 times larger
than PNPSs (curve b), demonstrating the importance of thin Pd shells for enhancing
ethanol electro-oxidation. Meanwhile, from Fig. 4.15b, c, the PBNPSs also show
higher specific (3.05 times) and mass (3.7 times) activities than the E-TEK Pd/C
catalyst. In addition, after 1800 s continuous operating (Fig. 4.15d), the specific
activities of PBNPSs (curve a) are much higher than those of the PNPSs (curve b)
and commercial Pd/C (curve c), suggesting that the PBNPSs catalyst has great
stability in ethanol oxidation reaction.
Recently, Lee and co-workers [144] prepared bimetallic Pt-Pd alloy
nanoparticles on a Nafion-graphene film by a simple electrochemical deposition
method. Due to the synergistic effects of the PtPd nanoparticles and the enhanced
electron transfer originating from graphene, the hybrid exhibited efficient catalytic
activity and stability toward ethanol electro-oxidation in alkaline media. Datta
et al. [49] successfully incorporated Pt-Pd bimetallic nanocrystallites into poly-
vinyl carbazole (PNVC) cross-linked with V2O5 (PNVC-V2O5) and investigated its
electrocatalytic activity for ethanol oxidation in alkaline condition. As compared to
the Pt/C, PtPd/C and Pt/PNVC-V2O5, the PtPd/PNVC-V2O5 catalyst exhibited
enhanced electrocatalytic performance with increased electrochemical active
202 R. Zhang and W. Chen

60
a a
8 b a
b b
30 c c
6

j / mA cm-2
j/A mgmetal-1

0
4
-30
2
-60

0
-90

-0.3 0.0 0.3 0.6 0.9 1.2 1.5 -0.9 -0.6 -0.3 0.0 0.3
E / V vs. Ag/AgCl E / V vs. Ag/AgCl
5
c a d a
0.9 b 4 b
c c
j / mA cm-2
0.6 3

2
0.3

1
0.0
0

-0.9 -0.6 -0.3 0.0 0.3 0 300 600 900 1200 1500 1800
E / V vs. Ag/Agcl Time / s

Fig. 4.15 (a) CVs of the PBNPSs (curve a), PNPSs (curve b) and E-TEK Pd/C (curve c)
electrodes in a 0.5 M H2SO4 solution; CVs ((b): specific activity; (c): mass activity) of the
PBNPSs (curve a), PNPSs (curve b) and E-TEK Pd/C (curve c) electrodes in a 0.5 M H2SO4 +
1 M ethanol solution at the scan rate of 50 mV/s; (d) Current density-time curves (specific activity)
of the PBNPSs (curve a), PNPSs (curve b) and E-TEK Pd/C (curve c) electrodes in a 0.5 M NaOH
+1 M ethanol solution at −0.2 V. Reprinted with permission from [117]. Copyright 2010 Royal
Society of Chemistry

surface area, lower onset potential, and larger current density of ethanol oxidation.
This study indicates that PNVC-V2O5 polymer-metal composite can serve as a
promising catalyst supporting material for fuel cells.

4.3.2 Formic Acid Oxidation Reaction (FAOR)

Formic acid is a liquid at room temperature, and dilute formic acid is recognized as
a safe fuel used in low-temperature fuel cells. As compared to traditional hydrogen
fuel, aqueous solutions of formic acid exhibit the advantages of ease of transpor-
tation, storage and handling. Moreover, direct formic acid fuel cells (DFAFCs)
have a smaller crossover of formic acid than that of methanol fuels in DMFCs,
which simultaneously improves fuel utilization and enables the use of thinner
membranes. In addition, DFAFCs has a theoretical open circuit voltage (OCV) of
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 203

about 1.48 or 1.45 V, which is higher than hydrogen-oxygen (1.23 V) and


methanol-oxygen (1.18 V) fuel cells [179–181]. For formic acid oxidation, a
so-called dual pathway mechanism has been commonly accepted: (i) direct path-
way (dehydrogenation pathway), which involves the removal of two
hydrogen atoms from formic acid molecule to directly produce CO2 (HCOOH→
CO2 +2H+ +2e−), and (ii) indirect pathway (dehydration pathway), which involves
the dehydration of formic acid to yield a poisonous intermediate CO species
(COads) and then further oxidation to CO2 (HCOOH→COads +H2O→CO2 +2H+
+2e−). On the basis of such a mechanism, to enhance fuel cell efficiency and avoid
poisoning, the materials on which the direct dehydrogenation pathway predomi-
nantly occurs or have high CO tolerance are ideal anode catalysts for formic acid
oxidation. Pt-Pd bimetallic catalysts are considered as a substitute of platinum in
fuel cells due to their attractive performance and better CO tolerance during formic
acid oxidation than that of Pt catalysts [69, 135, 182].
By controlling the crystallographic orientation of the surface atoms, Pt-Pd
bimetallic nanocrystals with enhanced electrochemical properties can be designed
[124, 173]. For instance, by using highly faceted cubic Pt as seeds to direct the
epitaxial overgrowth of Pd, Yang and coworkers [116] prepared shape-controlled
Pt-Pd heterostructures and studied their catalytic activities for formic acid electro-
oxidation. From the CVs of formic acid oxidation shown in Fig. 4.16, (100)-facets
terminated cubic Pt-Pd core-shell particles (Fig. 4.16a) exhibit a much higher peak
current that is five times of that from the (111)-facets terminated Pt-Pd octahedra
(Fig. 4.16c), whereas the octahedral Pt-Pd shows a lower peak potential (0.15 V)
than the cubic Pt-Pd (0.36 V). Moreover, the Pt-Pd cubes show a sharper peak at
0.47 V in the negative scan than the octahedra, and the Pt-Pd cuboctahedra show
intermediate catalytic properties between the cubes and octahedra (Fig. 4.16b). The
surface-dependent properties can be ascribed to the different arrangement of atoms
at the surface of different shaped Pt-Pd nanocrystals, and thus the corresponding
surface energy affects adsorption, surface diffusion, intermediate formation, chem-
ical rearrangements and finally desorption of the products in the formic acid
oxidation.
In another work, Pt-Pd binary nanoparticles were fabricated by decorating Pd on
the well-defined Pt(100) single crystal surface via a seed-mediated growth process
[123]. Because the presence of adsorbed Pd on Pt(100) can minimize CO formation
at low potential, and thus decrease the self-poisoning and lower the oxidation
potential, the prepared Pt-Pd nanoparticles showed remarkably enhanced
electrocatalytic performance for formic acid oxidation. Feliu and coworkers [145]
prepared Pd adatom-modified (100) preferentially oriented cubic Pt nanoparticles
by an electrochemical deposition approach using a reductant- and surfactant-free
process. In electrochemical measurements, the formic acid oxidation current was
strongly inhibited in the positive scan on the bare cubic Pt nanoparticles owing to
the almost complete poisoning of (100) surface sites by CO. After the incorporation
of Pd, the formic acid oxidation was enhanced. However, due to the lower fraction
of Pt(100) sites in the quasispherical Pt nanoparticles, the current density obtained
from the cubic Pt nanoparticles is higher than that from quasispherical Pt
204 R. Zhang and W. Chen

Fig. 4.16 Electrocatalytic activities of the Pt-Pd binary metal nanocrystals for formic acid
oxidation. (a–c) Cyclic voltammograms for cubes (a), cuboctahedra (b) and octahedra (c) in
0.1 M H2SO4 and 0.2 M formic acid at a sweep rate of 50 mV/s. Reprinted with permission
from [116]. Copyright 2007 Nature Publishing Group

nanoparticles. Moreover, the Pd adatom-decorated cubic Pt nanoparticles exhibited


high stability during the formic acid oxidation. Such results indicate that modifi-
cation of shape-controlled Pt by surface adatoms is an efficient strategy to enhance
their electrocatalytic activities.
It was also found that the catalytic activities of Pt-Pd bimetallic nanocrystals
towards formic acid electro-oxidation depend on their compositions
[67, 183]. Zheng and coworkers [111] reported that Pt can be alloyed into the Pd
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 205

Fig. 4.17 (a) Cyclic voltammograms recorded at 50 mV/s for PdxPt1−x/C coated on GC electrodes
in 0.5 M H2SO4 solution containing 0.5 M formic acid. (b) Variation of Ep and ip with Pd atomic
fraction (x). Reprinted with permission from [51]. Copyright 2010 American Chemical Society

nanospheres and the Pt-Pd nanoparticles with a molar ratio of Pd/Pt¼3:1 exhibited
the highest activity for electrocatalytic oxidation of formic acid. Cai and coworkers
[51] synthesized a series of PdxPt1−x (x¼0.5–1) nanoparticles dispersed on carbon
black support and studied their catalytic activities for formic acid oxidation.
Figure 4.17a compares the CVs of formic acid oxidation on different compositions
of PdxPt1−x/C catalysts in 0.5 M H2SO4 solution containing 0.5 M formic acid (FA).
Obviously, the peak current density (ip) of FA oxidation is x-dependent, and the
Pd0.9Pt0.1/C shows the highest ip which is nearly twice as high as that of Pd/C and
206 R. Zhang and W. Chen

six times of that of commercial Pt/C. The larger gap of negatively shifted peak
potentials (Fig. 4.17b) between Pd/C and Pd0.9Pt0.1/C but smaller difference
between PdxPt1−x/C catalysts demonstrate the synergistic effect of Pd and Pt sites
in the FA electro-oxidation. Moreover, after 1000 s stability test, the oxidation
current density on Pd0.9Pt0.1/C is the highest among the PdxPt1−x/C catalysts, which
is 2.4- and 3.6-fold larger than those of commercial Pd/C and Pt/C, respectively.
Carbon nanotubes have also exhibited their potential applications as catalyst
supports in fuel cells [71]. Chen and coworkers [99] studied the catalytic properties
of hollow Pt-Pd nanospheres dispersed on carbon nanotubes. The electrochemical
measurements showed that the prepared hybrids have enhanced activity with a high
peak current and steady-state current toward formic acid oxidation as compared to
commercial Pt/C and E-TEK PtRu/C. In another study, PtxPdy alloy nanoparticles
supported on carbon nanotubes were also studied as anode catalysts for FAOR
[59]. It was found that the catalytic activity and stability of the CNT-supported
catalysts increase with more Pd content in the catalysts. In addition, titanium [92]
and conductive polymer (polythiophene) [138] were also employed as support
materials to disperse Pt-Pd bimetallic nanocrystal with enhanced electrocatalytic
activities for formic acid oxidation.

4.3.3 Oxygen Reduction Reaction (ORR)

Oxygen reduction reaction (ORR) on cathode plays a crucial role in manipulating


the performance of a fuel cell. Whereas much progress has been achieved on
cathodic catalysts in the past few decades, the inherently slow reaction kinetics of
ORR and the large overpotential still prevent the fuel cells from being scaled-up for
commercial applications. The ORR has complicated reaction pathways
and involves a multi-electron transfer process in which O2 is converted into H2O
or OH−, depending on the solution used in the electrochemical studies. In an acidic
solution, O2 can be reduced in an efficient 4e− process and converted into H2O: O2 +
4e− →2H2O. Meanwhile, O2 may also undergo a less efficiently partial 2e− reduc-
tion to form hydrogen peroxide, H2O2, followed by another 2e− reduction to convert
H2O2 into H2O: O2 +2H+ + 2e− →H2O2; H2O2 +2H+ + 2e− →2H2O. In alkaline
solution, O2 can be reduced by a 4e− process to form hydroxide, OH−: O2 +2H2O+
4e− →4OH−, or by two 2e− processes to form HO2− and then OH−: O2 +H2O+2e−
→HO2− +OH−; HO2− +H2O+2e− →3OH−. In addition, under common ORR con-
ditions, oxygen may be converted into different intermediates, such as oxygenated
(O), hydroxyl (OH) and superhydroxyl (OOH) species, which are very difficult to
be detected experimentally [184].
The kinetic parameters of ORR can be determined using the rotating disk
electrode (RDE) technique. The kinetic current density and the electron transfer
numbers can be derived from the following equations: [14]
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 207

1 1 1 1 1
¼ þ ¼ þ 1 ð4:1Þ
j jK jL jK B =2
2= 1=
B ¼ 0:2nFðD0 Þ 3 6 C0 ð4:2Þ

jK ¼ nFkC0 ð4:3Þ

where j is the measured current density, jK and jL are the kinetic and diffusion
limitation current densities, ω is the electrode rotating rate, n represents the number
of electrons transferred per oxygen molecule, F is the Faraday constant (F¼
96,485 C/mol), D0 is the diffusion coefficient of O2, ν is the kinetic viscosity of
the electrolyte, and C0 is the bulk concentration of O2 dissolved in the electrolyte.
The constant 0.2 is used when the rotation speed is expressed in rpm. According to
Eqs. 4.1–4.3, the number of electron transfer (n) and kinetic current density ( jK) can
be obtained from the slope and intercept of the Koutecky-Levich plots, respec-
tively. In addition, rotating ring-disk electrode (RRDE) can also be used to evaluate
the ORR performance of a catalyst. Electron transfer number (n) and the H2O2
percentage (H2O2%) generated at the disk electrode can be calculated from the ring
and disk current as the following equations (iD and iR represent the disk and ring
current, and N is the RRDE collection efficiency) [185]:

4iD
n¼ ð4:4Þ
iD þ iR =N

200iR =N
H2 O2 % ¼ ð4:5Þ
iD þ iR =N

At present, because of the outstanding catalytic properties, Pt has been the most
widely used electrocatalyst to speed up oxygen reduction at cathode. However,
despite extensive research efforts, the sluggish oxygen reduction kinetics at the
cathode, high cost, low stability and limited supply of platinum still largely restrict
the wide spread commercialization of fuel cells. Moreover, the state-of-art com-
mercial catalysts are usually prepared by dispersing Pt nanoparticles (2–5 nm) on
carbon supports, whose electrochemical surface areas and catalytic efficiency
degrade over time owing to the corrosion of the carbon supports and the dissolution,
aggregation of Pt nanoparticles during the catalysis processes. This has posed
tremendous challenges in the development of Pt-based catalysts for practical fuel-
cell applications. The combination of Pd with Pt can downshift the d-band center of
Pt catalyst, leading to a lower degree of adsorption of oxygenated spectator species
(such as OH−) and an increase of the number of active sites accessible to oxygen,
Pt-Pd bimetallic nanocrystals have been accepted as one type of the most active and
effective cathodic catalysts for oxygen reduction reaction.
The enhancement of the Pt mass or specific activity for ORR can be achieved by
controlling the composition of the alloyed Pt-Pd nanomaterials [38–40, 56, 60, 73].
Cooks and coworkers [35] prepared Pt-Pd bimetallic nanoparticles (~1.8 nm)
containing an average of 180 atoms with seven different Pt:Pd ratios
208 R. Zhang and W. Chen

(G6-OHPtnPd180−nDENs, n¼180, 150, 120, 90, 60, 30 and 0) by a dendrimer


templating method. From the results of CV and rotating disk voltammetry mea-
surements shown in Fig. 4.18, as compared to the equivalent monometallic Pt DENs
catalyst, the maximum relative mass activity enhancement of 240% was obtained
from the bimetallic PtPd DENs containing 17–33% Pd. And the G6-OH (Pt150Pd30)
DENs exhibited the most active ORR performance with the most positive onset
potential, the highest kinetic current density and an efficient 4e− oxygen reduction
process. Huang and coworkers studied the ORR activity of the Pt-Pd bimetallic
nanocrystals with different compositions [115]. It was found that the bimetallic
Pt-Pd nanocrystals have higher ORR activity than that of commercially available Pt
black and Pd black in acid solution, and the bimetallic nanocrystals with Pd/Pt ratio
of 1/3 gave the highest ORR activity. Lee et al. [70] reported that the Pt-Pd
nanocatalysts deposited on Vulcan XC-72R carbon support with a molar ratio of
1:3 exhibited much enhanced catalytic activity toward oxygen reduction reaction in
comparison with commercial Pt/C. By using carbon nanotubes as supports for
electrocatalyst, Ghosh et al. [55] investigated the catalytic properties of multiwalled

Fig. 4.18 Plots of (a)


kinetic current density and
(b) kinetic current density
normalized to the fraction
amount of Pt contained in
each nanoparticle
composition at different
potentials. Error bars
represent the span of data
for two independently
executed experiments (only
one experiment was carried
out for G6-OH (Pt60Pd120).
Reprinted with permission
from [35]. Copyright 2007
American Chemical Society
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 209

carbon nanotubes (MCNTs)-supported Pt-Pd nanoparticles with different compo-


sitions. The electrochemical measurements showed that the Pt46Pd54 nanoparticles
supported on MCNTs have an enhanced electrocatalytic activity toward ORR with
specific and mass activities of 378 μA/cm2Pt-Pd and 64 μA/μgPt-Pd at 0.8 V, as
compared to 150 μA/cm2Pt-Pd and 26.5 μA/μgPt-Pd for Pt64Pd36, 33 μA/cm2Pt-Pd and
4.5 μA/μgPt-Pd for Pt28Pd72, and 142 μA/cm2Pt-Pd and 45.7 μA/μgPt-Pd for commer-
cial Pt black. Moreover, the Pt46Pd54/MCNTs exhibited high stability and tolerance
to methanol during the oxygen reduction. The synergistic effect between Pt and Pd
and the strong MCNT-catalyst interaction was believed to enhance the overall
catalytic performance of the catalysts.
In addition to composition, the catalytic activities of Pt-Pd bimetallic
nanocrystals toward ORR are also dependent on their shapes. Pt-Pd core-shell
nanocrystals are a class of electrocatalysts for ORR with an enhanced performance.
For example, Iwasawa and coworkers [114] investigated the catalytic activities of
three types of bimetallic Pt-Pd nanoparticles, Pt/C, Pd (core)-Pt (shell), and Pt
(core)-Pd (shell) for ORR. By comparing the CVs, the Pt/C, PdPt alloy/C, and Pd
(core)-Pt (shell)/C catalysts with Pt-enriched surfaces showed much higher ORR
specific activities than the Pd/C and Pt (core)-Pd (shell)/C catalysts with
Pd-enriched surfaces. And the Pt surface-enriched bimetallic catalysts with a
core-shell structure showed higher Pt-based mass activity than the Pt monometal
catalyst. Moreover, after 1000 cycles of CV test, the loss of ECSA is less than 20%
for Pd (core)-Pt (shell)/C and PdPt alloy/C with Pt-enriched surfaces, while drop as
large as 53.8% and 41.6% for Pd/C and Pt (core)-Pd (shell)/C with Pd-enriched
surfaces, suggesting that Pt surface-enriched catalysts have great stability in ORR.
Pt monolayer deposited on Pd provides another promising core-shell structure to
enhance the ORR activity through the strong interaction between these two metals,
as well as the reduction of Pt loading down to a minimum level [110, 146, 148, 150,
186, 187]. Adzic and coworkers [153] reported a synthesis of well-defined Pd core-
Pt monolayer shell nanocatalysts for ORR. It was found that the optimized Pt
monolayer on commercial Pd nanoparticles exhibited outstanding mass and area-
specific ORR activities of 0.96 A/mgPt and 0.5 mA/cm2, respectively. The signif-
icant enhancement in activity could be attributed to the uniquely advantageous
core-shell motif. In the structure, the Pd core imparts a compressive strain and a
so-called “ligand-effect” upon the Pt monolayer, which can effectively lower the
energy of the Pt d-band and thus directly weaken its interaction with OHads group,
thereby yielding improved ORR activity. Sasaki et al. [151] investigated the
catalytic performance of a Pd core-protected Pt monolayer shell catalyst for
ORR. Compared to commercial Pt/Ketjen black and Pt/C, the Pd nanoparticles
covered by a Pt monolayer (Pt ML/Pd/C) exhibited much higher electrocatalytic
durability. The mass activity of the Pt ML/Pd/C decreased by 37% after
100,000 cycles of stability tests, while the mass activity of Pt/C lost almost 70%
after 60,000 potential cycles and the Pt/Ketjen black carbon experienced more than
40% loss after only 10,000 cycles. Surprisingly, due to the effect of potential
cycling on the particle size distribution, the Pt mass activity of PtML/Pd/C
210 R. Zhang and W. Chen

Fig. 4.19 Cyclic voltammograms (a) obtained for the ozone- and acid-treated Pt-PdNW/C core-
shell nanowires, after Pt monolayer deposition, loaded separately onto a GCE in a 0.1 M HClO4
solution at 20 mV/s. (b) The polarization curves for the treated Pt-Pd/C core-shell nanowire
composites were both obtained on a glassy carbon rotating disk electrode. Curves (anodic sweep
direction) were collected using a rotation rate of 1600 rpm in a 0.1 M HClO4 solution at 20 ˚C.
Kinetic current vs potential plots (inset) of treated Pt-Pd/C composites and commercial carbon
supported Pt nanoparticles are also presented. Reprinted with permission from [152]. Copyright
2011 American Chemical society

increased from the initial 3-fold to 5-fold enhancement over that of Pt/C after
60,000 cycles.
In another work, Adzic and coworkers [152] reported the deposition of a
monolayer quantity of Pt onto the surface of the treated Pd nanowires/C
(Pd NW/C) composites by Cu UPD followed by galvanic replacement of the Cu
adatoms with Pt2+, and their electrocatalytic activity for ORR in acid condition. As
shown in Fig. 4.19a, b, the ultrathin Pt monolayer shell-Pd nanowire core catalyst
(Pt-Pd NW/C) displayed dramatically enhanced ORR activity compared to the Pd
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 211

NW/C and commercial Pt nanoparticles. Moreover, the ozone-treated Pt-Pd NW/C


exhibited much higher area-specific (0.77 mA/cm2) and mass-specific
(1.83 A/mgPt) activities than those of acid-treated Pt-Pd nanowires (0.70 mA/cm2
and 1.47 A/mgPt). Such results indicate that the ozone treatment for Pd NW before
Cu UPD can lead to the exceptional enhancement in the catalytic activity of Pt-Pd
nanowires. Most importantly, after the accelerated durability half-cell test, the
ozone-treated Pt-Pd nanowires still maintained excellent electrochemical durability
and the area-specific activity increased by 1.5-fold.
Pt-Pd nanodendrites are another novel structure imparting its enhanced ORR
activity by maximizing the expression of certain highly active facets and improving
surface area. Xia and coworkers [122] synthesized Pt-Pd bimetallic nanodendrites
consisting of a dense array of Pt branches on a Pd core, and investigated their
catalytic activity for ORR in acid conditions. The calculated ECSA of the Pt-Pd
nanodendrites (57.1 m2/gPt) is much larger than that of commercial Pt black catalyst
(19.1 m2/gPt), but is lower than that of Pt/C catalyst (74 m2/gPt). It was found that
the mass activity from the Pt-Pd nanodendrites (0.241 mA/μgPt) is 2.5- and 5-fold
larger, respectively, than those of state-of-art Pt/C and the first-generation
supportless Pt-black catalyst at room temperature. Moreover, at 60 ˚C, the Pt
mass activity of the Pt-Pd nanodendrites (0.433 mA/μgPt) was still greater than
that of the Pt/C catalyst (0.204 mA/μgPt) and the Pt black (0.078 mA/μgPt).
In addition to mass activity, Pt-Pd nanodendrites exhibited a specific activity of
3.1–3.4-fold that of the Pt/C catalyst and 1.7–2.0-fold that of the Pt black depending
on the temperature, indicating a superior capability of the Pt-Pd nanodendrites in
accelerating ORR. The high ORR activity of the Pt-Pd nanodendrites can be
attributed to the high and accessible surface area of the dendritic morphology
and the preferential exposure of (111) facets along with some (110) and high-
index (311) facets on the Pt branches that are particularly active towards ORR. In
another work, Yang and coworkers [126] reported the synthesis of Pt-on-Pd
heteronanostructure supported on carbon black. The as-synthesized nanostructure
showed much larger electrochemically active surface area as compared to E-TEK
Pt/C (20 wt%Pt) catalyst (Fig. 4.20a). From Fig. 4.20b, it can be seen that the
incorporation of Pd also greatly altered the ability to absorb hydroxyl species
(OHads, E>0.6 V) compared to E-TEK Pt/C (20 wt%Pt) catalyst, and the Pt-on-
Pd catalyst exhibited much faster hydroxyl desorption ability with positively shifted
onset and peak potentials in the backward sweep. As shown in Fig. 4.20c, d, the
Pt-on-Pd catalyst shows a more positive onset potential and higher activity than the
commercially pure Pt. Moreover, after 30,000 cyclic voltammetry cycling, the
Pt-on-Pd catalyst lost only 12% of the initial ECSA and showed a small degradation
of 9 mV in the half-wave potential, while a loss of 39% of the initial ECSA and a
large decrease of 35 mV in the half-wave potential for Pt/C catalyst, suggesting the
enhanced stability of Pt-on-Pd/C catalyst for ORR.
In addition, the ORR catalytic activities of Pt-Pd bimetallic nanostructures with
other specific size and morphologies such as hollow urchin-like [97], bowl-like
[100], and rods [77], were also investigated. Yan et al. [101] found that Pt-Pd
nanotubes exhibited much enhanced ORR mass/specific activity, and improved
212 R. Zhang and W. Chen

Fig. 4.20 (a) CVs, (b) hydroxyl surface coverage (ΘOH), (c) ORR polarization curves, and (d)
specific kinetic current densities (ik) for carbon-supported Pt-on-Pd and Pt catalysts. Reprinted
with permission from [126]. Copyright 2009 American Chemical Society

stability in acid electrolyte as compared to commercial Pt/C and Pt black catalysts.


As reported by Xia and coworkers in another work [102], Pt-Pd concave cubes with
different weight percentages of Pt can be prepared by using bromide-induced
galvanic replacement method. It was found that the ORR activities of the Pt-Pd
concave nanocubes on the basis of Pt mass monotonically increase with decreasing
Pt weight percentage. And the Pt-Pd concave nanocubes with a Pt weight percent-
age of 3.4 showed the highest mass activity for ORR, which is 4 times higher than
that of the Pt/C catalyst in terms of equivalent Pt mass. In a recent work from our
group [104], Pt-Pd porous nanorods (PtPd PNRs) were successfully synthesized
through a bromide-induced galvanic replacement reaction between Pd nanowires
(Pd NWs) and K2PtCl6. The unique porous Pt-Pd nanorods showed greatly
enhanced ORR activity with the highest limiting current densities and the most
positive onset potential compared to those obtained from the original Pd NWs and
commercial Pt/C catalyst. In addition to the improved catalytic activity, the PtPd
PNRs also showed excellent durability in ORR (Fig. 4.21a–d) with only 5.88% loss
of the initial ECSA after the accelerated durability tests (1000 potential cycles),
whereas 21.6% and 40.4% of their original ECSA were lost for the Pd NWs and
commercial Pt/C catalyst, respectively. The enhanced catalytic activity of PtPd
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 213

800 300
a Pd NWs b Pt/C
200
400
100

I (µA)
I (µA)

0 0

-400 -100
1st -200 1 st
-800 1000th -300 1000 th
-1200 -400
-0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4

E (V vs Ag/AgCl) E (V vs Ag/AgCl)
4
200 c PtPd PNRs d

ECSA (cm 2)
100 3
I (µA)

0
-100 2

-200
1
-300 1st
1000th
-400 0
-0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 PtPd Pd Pt/C
E (V vs Ag/AgCl) PNRs NWs

Fig. 4.21 Cyclic voltammograms recorded on Pd NWs (a), Pt/C (b) and PtPd PNRs (c) before
(black curves) and after 1000 cycles (red curves) of the accelerated durability test, in 0.1 M HClO4
solution with a potential scan rate of 0.1 V/s. (d) Comparison of the electrochemical surface area
(ECSA) before and after 1000 cycles of the accelerated durability tests on the PtPd PNRs, Pd NWs,
and commercial Pt/C catalyst, respectively. Reprinted with permission from [104]. Copyright
2013 Elsevier

PNRs toward ORR with efficient 4e− pathway can be ascribed to its high specific
surface area, unique one-dimensional morphology and the alloy structure.

4.4 Conclusion

Over the past several years, Pt-Pd bimetallic nanocrystals have received great
attention owing to their attractive properties and extensive applications in fuel
cells. In this chapter, the recent development related to the synthesis of Pt-Pd
bimetallic nanocrystals with different structures, including Pt-Pd alloys, core-
shell, multi-shells, dendrites and supported monolayer has been extensively
reviewed. Furthermore, a number of strategies based on co-chemical reduction,
galvanic displacement, seed-mediated growth and the combination of these
methods used to control the formation of Pt-Pd bimetallic nanocrystals with a
specific structure have been discussed. Finally, the application of Pt-Pd
214 R. Zhang and W. Chen

(or supported Pt-Pd) bimetallic nanocrystals in fuel cells as catalysts for fuels
oxidation at anode and oxygen reduction reaction at cathode was summarized.
Despite substantial progress that has been made in designing Pt-Pd bimetallic
electrocatalysts and understanding their electrocatalytic mechanism, the develop-
ment of new methodologies, the rational control on size, shape, structure and
composition of bimetallic nanocrystals, in-depth understanding of the structure-
catalytic property relationship, and the exploration of more extensive applications
of bimetallic nanocrystals in fuel cells still need to be continuously studied.

Acknowledgements This work was supported by the National Natural Science Foundation of
China with the Grant Numbers of 21275136, 21043013 and the Natural Science Foundation of
Jilin Province, China (No. 201215090).

References

1. Potocnik J (2007) Renewable energy sources and the realities of setting an energy agenda.
Science 315:810–811
2. Armand M, Tarascon JM (2008) Building better batteries. Nature 451:652–657
3. Steele BCH, Heinzel A (2001) Materials for fuel-cell technologies. Nature 414:345–352
4. Russell AE, Rose A (2004) X-ray absorption spectroscopy of low temperature fuel cell
catalysts. Chem Rev 104:4613–4635
5. Lamy C, Lima A, LeRhun V, Delime F, Coutanceau C, Leger JM (2002) Recent advances in
the development of direct alcohol fuel cells (DAFC). J Power Sources 105:283–296
6. Vigier F, Rousseau S, Coutanceau C, Leger JM, Lamy C (2006) Electrocatalysis for the direct
alcohol fuel cell. Top Catal 40:111–121
7. Tian N, Zhou ZY, Sun SG, Ding Y, Wang ZL (2007) Synthesis of tetrahexahedral platinum
nanocrystals with high-index facets and high electro-oxidation activity. Science 316:732–735
8. Chen W, Kim J, Sun SH, Chen SW (2006) Electro-oxidation of formic acid catalyzed by FePt
nanoparticles. Phys Chem Chem Phys 8:2779–2786
9. Chen W, Kim JM, Sun SH, Chen SW (2007) Composition effects of FePt alloy nanoparticles
on the electro-oxidation of formic acid. Langmuir 23:11303–11310
10. Yang HZ, Zhang J, Sun K, Zou SZ, Fang JY (2010) Enhancing by weakening:
electrooxidation of methanol on Pt3Co and Pt nanocubes. Angew Chem Int Ed 49:6848–6851
11. Chen YX, Miki A, Ye S, Sakai H, Osawa M (2003) Formate, an active intermediate for direct
oxidation of methanol on Pt electrode. J Am Chem Soc 125:3680–3681
12. Tripkovic AV, Popovic KD, Grgur BN, Blizanac B, Ross PN, Markovic NM (2002) Meth-
anol electrooxidation on supported Pt and PtRu catalysts in acid and alkaline solutions.
Electrochim Acta 47:3707–3714
13. Qiao Y, Li CM (2011) Nanostructured catalysts in fuel cells. J Mater Chem 21:4027–4036
14. Zhang RZ, Chen W (2013) Non-precious Ir-V bimetallic nanoclusters assembled on reduced
graphene nanosheets as catalysts for the oxygen reduction reaction. J Mater Chem A
1:11457–11464
15. Liu MM, Zhang RZ, Chen W (2014) Graphene-supported nanoelectrocatalysts for fuel cells:
synthesis, properties, and applications. Chem Rev 114:5117–5160
16. Zhang H, Jin MS, Xia YN (2012) Enhancing the catalytic and electrocatalytic properties of
Pt-based catalysts by forming bimetallic nanocrystals with Pd. Chem Soc Rev 41:8035–8049
17. Gasteiger HA, Markovic N, Ross PN, Cairns EJ (1994) Temperature-dependent methanol
electrooxidation on well-characterized Pt-Ru alloys. J Electrochem Soc 141:1795–1803
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 215

18. Dinh HN, Ren XM, Garzon FH, Zelenay P, Gottesfeld S (2000) Electrocatalysis in direct
methanol fuel cells: in-situ probing of PtRu anode catalyst surfaces. J Electroanal Chem
491:222–233
19. Wang DS, Li YD (2011) Bimetallic nanocrystals: liquid-phase synthesis and catalytic
applications. Adv Mater 23:1044–1060
20. Calvo SR, Balbuena PB (2007) Theoretical analysis of reactivity on Pt(111) and Pt-Pd(111)
alloys. Surf Sci 601:4786–4792
21. Calvo SR, Balbuena PB (2007) Density functional theory analysis of reactivity of PtxPdy
alloy clusters. Surf Sci 601:165–171
22. Norskov JK, Bligaard T, Rossmeisl J, Christensen CH (2009) Towards the computational
design of solid catalysts. Nat Chem 1:37–46
23. Norskov JK, Abild-Pedersen F (2009) Catalysis bond control in surface reactions. Nature
461:1223–1225
24. Kitchin JR, Norskov JK, Barteau MA, Chen JG (2004) Role of strain and ligand effects in the
modification of the electronic and chemical properties of bimetallic surfaces. Phys Rev Lett
93:156801–156804
25. Hammer B, Norskov JK (2000) Theoretical surface science and catalysis-calculations and
concepts. Adv Catal 45:71–129
26. Norskov JK, Rossmeisl J, Logadottir A, Lindqvist L, Kitchin JR, Bligaard T, Jonsson H
(2004) Origin of the overpotential for oxygen reduction at a fuel-cell cathode. J Phys Chem B
108:17886–17892
27. de Bruijn FA, Dam VAT, Janssen GJM (2008) Durability and degradation issues of PEM fuel
cell components. Fuel Cells 8:3–22
28. Stamenkovic VR, Fowler B, Mun BS, Wang GF, Ross PN, Lucas CA, Markovic NM (2007)
Improved oxygen reduction activity on Pt3Ni(111) via increased surface site availability.
Science 315:493–497
29. Ramirez-Caballero GE, Hirunsit P, Balbuena PB (2010) Shell-anchor-core structures for
enhanced stability and catalytic oxygen reduction activity. J Chem Phys 133:134705–134713
30. Xia YN, Xiong YJ, Lim B, Skrabalak SE (2009) Shape-controlled synthesis of metal
nanocrystals: simple chemistry meets complex physics? Angew Chem Int Ed 48:60–103
31. Shao YY, Yin GP, Gao YZ (2007) Understanding and approaches for the durability issues of
Pt-based catalysts for PEM fuel cell. J Power Sources 171:558–566
32. Ferreira PJ, La O’ GJ, Shao-Horn Y, Morgan D, Makharia R, Kocha S, Gasteiger HA (2005)
Instability of Pt/C electrocatalysts in proton exchange membrane fuel cells—a mechanistic
investigation. J Electrochem Soc 152:A2256–A2271
33. Wang YJ, Wilkinson DP, Zhang JJ (2011) Noncarbon support materials for polymer elec-
trolyte membrane fuel cell electrocatalysts. Chem Rev 111:7625–7651
34. Greeley J, Stephens IEL, Bondarenko AS, Johansson TP, Hansen HA, Jaramillo TF,
Rossmeisl J, Chorkendorff I, Norskov JK (2009) Alloys of platinum and early transition
metals as oxygen reduction electrocatalysts. Nat Chem 1:552–556
35. Ye HC, Crooks RM (2007) Effect of elemental composition of PtPd bimetallic nanoparticles
containing an average of 180 atoms on the kinetics of the electrochemical oxygen reduction
reaction. J Am Chem Soc 129:3627–3633
36. Yang SC, Hong F, Wang LQ, Guo SW, Song XP, Ding BJ, Yang ZM (2010) Ultrathin
Pt-based alloy nanowire networks: synthesized by CTAB assistant two-phase water-chloro-
form micelles. J Phys Chem C 114:203–207
37. Basu D, Basu S (2012) Performance studies of Pd-Pt and Pt-Pd-Au catalyst for electro-
oxidation of glucose in direct glucose fuel cell. Int J Hydrogen Energy 37:4678–4684
38. Thanasilp S, Hunsom M (2011) Preparation of a high-performance Pt-Pd/C-electrocatalyst-
coated membrane for ORR in PEM fuel cells via a combined process of impregnation and
seeding: effect of electrocatalyst loading on carbon support. Electrochim Acta 56:1164–1171
216 R. Zhang and W. Chen

39. Thanasilp S, Hunsom M (2011) Effect of Pt: Pd atomic ratio in Pt-Pd/C electrocatalyst-coated
membrane on the electrocatalytic activity of ORR in PEM fuel cells. Renew Energy 36:1795–
1801
40. Tang YF, Zhang HM, Zhong HX, Xu T, Jin H (2011) Carbon-supported Pd-Pt cathode
electrocatalysts for proton exchange membrane fuel cells. J Power Sources 196:3523–3529
41. Maghsodi A, Hoseini MRM, Mobarakeh MD, Kheirmand M, Samiee L, Shoghi F, Kameli M
(2011) Exploration of bimetallic Pt-Pd/C nanoparticles as an electrocatalyst for oxygen
reduction reaction. Appl Surf Sci 257:6353–6357
42. Li XW, Zhu Y, Zou ZQ, Zhao MY, Li ZL, Zhou Q, Akins DL, Yang H (2009) Simple
complexing-reduction synthesis of Pd-Pt/C alloy electrocatalysts for the oxygen reduction
reaction. J Electrochem Soc 156:B1107–B1111
43. Kim IT, Lee HK, Shim J (2008) Synthesis and characterization of Pt-Pd catalysts for
methanol oxidation and oxygen reduction. J Nanosci Nanotechnol 8:5302–5305
44. Joo JB, Kim YJ, Kim W, Kim ND, Kim P, Kim Y, Yi J (2008) Methanol-tolerant PdPt/C
alloy catalyst for oxygen electro-reduction reaction. Korean J Chem Eng 25:770–774
45. Golikand AN, Asgari M, Lohrasbi E (2011) Study of oxygen reduction reaction kinetics on
multi-walled carbon nano-tubes supported Pt-Pd catalysts under various conditions. Int J
Hydrogen Energy 36:13317–13324
46. Golikand AN, Lohrasbi E, Asgari M (2010) Enhancing the durability of multi-walled carbon
nanotube supported by Pt and Pt-Pd nanoparticles in gas diffusion electrodes. Int J Hydrogen
Energy 35:9233–9240
47. Morales-Acosta D, Arriaga LG, Alvarez-Contreras L, Luna SF, Varela FJR (2009) Evalua-
tion of Pt40Pd60/MWCNT electrocatalyst as ethylene glycol-tolerant oxygen reduction cath-
odes. Electrochem Commun 11:1414–1417
48. Golikand AN, Lohrasbi E, Maragheh MG, Asgari M (2009) Carbon nano-tube supported
Pt-Pd as methanol-resistant oxygen reduction electrocatalyts for enhancing catalytic activity
in DMFCs. J Appl Electrochem 39:2421–2431
49. Datta J, Dutta A, Biswas M (2012) Enhancement of functional properties of PtPd nano
catalyst in metal-polymer composite matrix: application in direct ethanol fuel cell.
Electrochem Commun 20:56–59
50. Wu M, Shen PK, Wei ZD, Song SQ, Nie M (2007) High activity PtPd-WC/C electrocatalyst
for hydrogen evolution reaction. J Power Sources 166:310–316
51. Zhang HX, Wang C, Wang JY, Zhai JJ, Cai WB (2010) Carbon-Supported Pd-Pt nanoalloy
with low Pt content and superior catalysis for formic acid electro-oxidation. J Phys Chem C
114:6446–6451
52. Godinez-Garcia A, Gervasio DF (2014) Pd-Pt nanostructures on carbon nanofibers as an
oxygen reduction electrocatalyst. RSC Adv 4:42009–42013
53. Tao F, Grass ME, Zhang YW, Butcher DR, Renzas JR, Liu Z, Chung JY, Mun BS,
Salmeron M, Somorjai GA (2008) Reaction-driven restructuring of Rh-Pd and Pt-Pd core-
shell nanoparticles. Science 322:932–934
54. Arikan T, Kannan AM, Kadirgan F (2013) Binary Pt-Pd and ternary Pt-Pd-Ru nanoelectro-
catalysts for direct methanol fuel cells. Int J Hydrogen Energy 38:2900–2907
55. Ghosh S, Sahu RK, Raj CR (2012) Pt-Pd alloy nanoparticle-decorated carbon nanotubes: a
durable and methanol tolerant oxygen reduction electrocatalyst. Nanotechnology 23:385602–
285610
56. Zhao JA, Manthiram A (2011) Preleached Pd-Pt-Ni and binary Pd-Pt electrocatalysts for
oxygen reduction reaction in proton exchange membrane fuel cells. Appl Catal B Environ
101:660–668
57. Long NV, Hien TD, Asaka T, Ohtaki M, Nogami M (2011) Synthesis and characterization of
Pt-Pd alloy and core-shell bimetallic nanoparticles for direct methanol fuel cells (DMFCs):
enhanced electrocatalytic properties of well-shaped core-shell morphologies and
nanostructures. Int J Hydrogen Energy 36:8478–8491
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 217

58. Zhou ZM, Shao ZG, Qin XP, Chen XG, Wei ZD, Yi BL (2010) Durability study of Pt-Pd/C as
PEMFC cathode catalyst. Int J Hydrogen Energy 35:1719–1726
59. Winjobi O, Zhang ZY, Liang CH, Li WZ (2010) Carbon nanotube supported platinum-
palladium nanoparticles for formic acid oxidation. Electrochim Acta 55:4217–4221
60. He W, Liu JY, Qiao YJ, Zou ZQ, Zhang XG, Akins DL, Yang H (2010) Simple preparation of
Pd-Pt nanoalloy catalysts for methanol-tolerant oxygen reduction. J Power Sources
195:1046–1050
61. He W, Chen M, Zou ZQ, Li ZL, Zhang XG, Jin SA, You DJ, Pak C, Yang H (2010) Oxygen
reduction on Pd3Pt1 bimetallic nanoparticles highly loaded on different carbon supports. Appl
Catal B Environ 97:347–353
62. Wang WM, Huang QH, Liu JY, Zou ZQ, Zhao MY, Vogel W, Yang H (2009) Surface and
structure characteristics of carbon-supported Pd3Pt1 bimetallic nanoparticles for methanol-
tolerant oxygen reduction reaction. J Catal 266:156–163
63. Kadirgan F, Kannan AM, Atilan T, Beyhan S, Ozenler SS, Suzer S, Yorur A (2009) Carbon
supported nano-sized Pt-Pd and Pt-Co electrocatalysts for proton exchange membrane fuel
cells. Int J Hydrogen Energy 34:9450–9460
64. Kadirgan F, Beyhan S, Atilan T (2009) Preparation and characterization of nano-sized Pt-Pd/
C catalysts and comparison of their electro-activity toward methanol and ethanol oxidation.
Int J Hydrogen Energy 34:4312–4320
65. Ficicilar B, Bayrakceken A, Eroglu I (2009) Effect of Pd loading in Pd-Pt bimetallic catalysts
doped into hollow core mesoporous shell carbon on performance of proton exchange mem-
brane fuel cells. J Power Sources 193:17–23
66. Li HQ, Sun GQ, Li N, Sun SG, Su DS, Xin Q (2007) Design and preparation of highly active
Pt-Pd/C catalyst for the oxygen reduction reaction. J Phys Chem C 111:5605–5617
67. Blair S, Lycke D, Iordache CA (2006) Palladium-platinum alloy anode catalysts for direct
formic acid fuel cells. ECS Trans 3:1325–1332
68. Zhou WJ, Zhou ZH, Song SQ, Li WZ, Sun GQ, Tsiakaras P, Xin Q (2003) Pt based anode
catalysts for direct ethanol fuel cells. Appl Catal B Environ 46:273–285
69. Li X, Hsing I (2006) Electrooxidation of formic acid on carbon supported PtxPd1-x (x¼0-1)
nanocatalysts. Electrochim Acta 51:3477–3483
70. Lee YW, Ko AR, Kim DY, Han SB, Park KW (2012) Octahedral Pt-Pd alloy catalysts with
enhanced oxygen reduction activity and stability in proton exchange membrane fuel cells.
RSC Adv 2:1119–1125
71. Selvaraj V, Grace AN, Alagar M (2009) Electrocatalytic oxidation of formic acid and
formaldehyde on nanoparticle decorated single walled carbon nanotubes. J Colloid Interface
Sci 333:254–262
72. Lee YW, Ko AR, Han SB, Kim HS, Park KW (2011) Synthesis of octahedral Pt-Pd alloy
nanoparticles for improved catalytic activity and stability in methanol electrooxidation. Phys
Chem Chem Phys 13:5569–5572
73. Nishanth KG, Sridhar P, Pitchumani S, Shukla AK (2011) A DMFC with methanol-tolerant-
carbon-supported-Pt-Pd-alloy cathode. J Electrochem Soc 158:B871–B876
74. Antolini E, Zignani SC, Santos SF, Gonzalez ER (2011) Palladium-based electrodes: a way to
reduce platinum content in polymer electrolyte membrane fuel cells. Electrochim Acta
56:2299–2305
75. Lopes T, Antolini E, Gonzalez ER (2008) Carbon supported Pt-Pd alloy as an ethanol tolerant
oxygen reduction electrocatalyst for direct ethanol fuel cells. Int J Hydrogen Energy
33:5563–5570
76. Guo SJ, Dong SJ, Wang EK (2009) Polyaniline/Pt hybrid nanofibers: high-efficiency
nanoelectrocatalysts for electrochemical devices. Small 5:1869–1876
77. Chen HS, Liang YT, Chen TY, Tseng YC, Liu CW, Chung SR, Hsieh CT, Lee CE, Wang KW
(2014) Graphene-supported Pt and PtPd nanorods with enhanced electrocatalytic perfor-
mance for the oxygen reduction reaction. Chem Commun 50:11165–11168
218 R. Zhang and W. Chen

78. Solla-Gullon J, Rodes A, Montiel V, Aldaz A, Clavilier J (2003) Electrochemical character-


isation of platinum-palladium nanoparticles prepared in a water-in-oil microemulsion. J
Electroanal Chem 554:273–284
79. Chen CH, Hwang BJ, Wang GR, Sarma LS, Tang MT, Liu DG, Lee JF (2005) Nucleation and
growth mechanism of Pd/Pt bimetallic clusters in sodium bis(2-ethylhexyl)sulfosuceinate
(AOT) reverse micelles as studied by in situ X-ray absorption spectroscopy. J Phys Chem B
109:21566–21575
80. Escudero MJ, Hontanon E, Schwartz S, Boutonnet M, Daza L (2002) Development and
performance characterisation of new electrocatalysts for PEMFC. J Power Sources 106:206–
214
81. Zhang H, Xu XQ, Gu P, Li CY, Wu P, Cai CX (2011) Microwave-assisted synthesis of
graphene-supported Pd1Pt3 nanostructures and their electrocatalytic activity for methanol
oxidation. Electrochim Acta 56:7064–7070
82. Zhang H, Yin YJ, Hu YJ, Li CY, Wu P, Wei SH, Cai CX (2010) Pd@Pt core-shell
nanostructures with controllable composition synthesized by a microwave method and their
enhanced electrocatalytic activity toward oxygen reduction and methanol oxidation. J Phys
Chem C 114:11861–11867
83. Xia H, Wang D (2008) Fabrication of macroscopic freestanding films of metallic nanoparticle
monolayers by interfacial self-assembly. Adv Mater 20:4253–4256
84. Liang HW, Cao XA, Zhou F, Cui CH, Zhang WJ, Yu SH (2011) A free-standing Pt-nanowire
membrane as a highly stable electrocatalyst for the oxygen reduction reaction. Adv Mater
23:1467–1471
85. Jin YD, Dong SJ (2002) Diffusion-limited, aggregation-based, mesoscopic assembly of
roughened core shell bimetallic nanoparticles into fractal networks at the air-water interface.
Angew Chem Int Ed 41:1040–1044
86. Shi HY, Hu B, Yu XC, Zhao RL, Ren XF, Liu SL, Liu JW, Feng M, Xu AW, Yu SH (2010)
Ordering of disordered nanowires: spontaneous formation of highly aligned, ultralong Ag
nanowire films at oil-water-air interface. Adv Funct Mater 20:958–964
87. Kang YJ, Ye XC, Chen J, Cai Y, Diaz RE, Adzic RR, Stach EA, Murray CB (2013) Design of
Pt-Pd binary superlattices exploiting shape effects and synergistic effects for oxygen reduc-
tion reactions. J Am Chem Soc 135:42–45
88. Wu HX, Li HJ, Zhai YJ, Xu XL, Jin YD (2012) Facile synthesis of free-standing Pd-based
nanomembranes with enhanced catalytic performance for methanol/ethanol oxidation. Adv
Mater 24:1594–1597
89. Liu Y, Chi MF, Mazumder V, More KL, Soled S, Henao JD, Sun SH (2011) Composition-
controlled synthesis of bimetallic PdPt nanoparticles and their electro-oxidation of methanol.
Chem Mater 23:4199–4203
90. Lim B, Wang JG, Camargo PHC, Cobley CM, Kim MJ, Xia YN (2009) Twin-induced growth
of palladium-platinum alloy nanocrystals. Angew Chem Int Ed 48:6304–6308
91. Papageorgopoulos DC, Keijzer M, Veldhuis JBJ, de Bruijn FA (2002) CO tolerance of
Pd-rich platinum palladium carbon-supported electrocatalysts-Proton exchange membrane
fuel cell applications. J Electrochem Soc 149:A1400–A1404
92. Yi QF, Huang W, Liu XP, Xu GR, Zhou ZH, Chen AC (2008) Electroactivity of titanium-
supported nanoporous Pd-Pt catalysts towards formic acid oxidation. J Electroanal Chem
619:197–205
93. Huang D, Yuan Q, Wang HH, Zhou ZY (2014) Facile synthesis of PdPt nanoalloys with
sub-2.0 nm islands as robust electrocatalysts for methanol oxidation. Chem Commun
50:13551–13554
94. Yin AX, Min XQ, Zhang YW, Yan CH (2011) Shape-selective synthesis and facet-dependent
enhanced electrocatalytic activity and durability of monodisperse sub-10 nm Pt-Pd tetrahe-
drons and cubes. J Am Chem Soc 133:3816–3819
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 219

95. Lu YZ, Jiang YY, Wu HB, Chen W (2013) Nano-PtPd cubes on graphene exhibit enhanced
activity and durability in methanol electrooxidation after CO stripping-cleaning. J Phys Chem
C 117:2926–2938
96. Lu YZ, Jiang YY, Chen W (2014) Graphene nanosheet-tailored PtPd concave nanocubes with
enhanced electrocatalytic activity and durability for methanol oxidation. Nanoscale 6:3309–
3315
97. Guo SJ, Dong SJ, Wang EK (2008) A general method for the rapid synthesis of hollow
metallic or bimetallic nanoelectrocatalysts with urchinlike morphology. Chem Eur J
14:4689–4695
98. Yang JH, Lee JY, Zhang QB, Zhou WJ, Liu ZL (2008) Carbon-supported pseudo-core-shell
Pd-Pt nanoparticles for ORR with and without methanol. J Electrochem Soc 155:B776–B781
99. Liu B, Li HY, Die L, Zhang XH, Fan Z, Chen JH (2009) Carbon nanotubes supported PtPd
hollow nanospheres for formic acid electrooxidation. J Power Sources 186:62–66
100. Zhou XM, Fan LZ (2010) Pt/Pd alloy nanoparticles composed of bimetallic nanobowls
Synthesis, characterization and electrocatalytic activities. Electrochim Acta 55:8111–8115
101. Chen ZW, Waje M, Li WZ, Yan YS (2007) Supportless Pt and PtPd nanotubes as
electrocatalysts for oxygen-reduction reactions. Angew Chem Int Ed 46:4060–4063
102. Zhang H, Jin MS, Wang JG, Li WY, Camargo PHC, Kim MJ, Yang DR, Xie ZX, Xia YA
(2011) Synthesis of Pd-Pt bimetallic nanocrystals with a concave structure through a
bromide-induced galvanic replacement reaction. J Am Chem Soc 133:6078–6089
103. Zhang H, Jin MS, Liu HY, Wang JG, Kim MJ, Yang DR, Xie ZX, Liu JY, Xia YN (2011)
Facile synthesis of Pd-Pt alloy nanocages and their enhanced performance for preferential
oxidation of CO in excess hydrogen. ACS Nano 5:8212–8222
104. Lu YZ, Jiang YY, Chen W (2013) PtPd porous nanorods with enhanced electrocatalytic
activity and durability for oxygen reduction reaction. Nano Energy 2:836–844
105. Huang XQ, Zhang HH, Guo CY, Zhou ZY, Zheng NF (2009) Simplifying the creation of
hollow metallic nanostructures: One-pot synthesis of hollow palladium/platinum single-
crystalline nanocubes. Angew Chem Int Ed 48:4808–4812
106. Bauer E, Vandermerwe JH (1986) Structure and growth of crystalline superlattices-from
monolayer to superlattice. Phys Rev B 33:3657–3671
107. Fan FR, Liu DY, Wu YF, Duan S, Xie ZX, Jiang ZY, Tian ZQ (2008) Epitaxial growth of
heterogeneous metal nanocrystals: From gold nano-octahedra to palladium and silver
nanocubes. J Am Chem Soc 130:6949–6951
108. Beard KD, Van Zee JW, Monnier JR (2009) Preparation of carbon-supported Pt-Pd
electrocatalysts with improved physical properties using electroless deposition methods.
Appl Catal B Environ 88:185–193
109. Long NV, Asaka T, Matsubara T, Nogami M (2011) Shape-controlled synthesis of Pt-Pd core
shell nanoparticles exhibiting polyhedral morphologies by modified polyol method. Acta
Mater 59:2901–2907
110. Long NV, Ohtaki M, Hien TD, Randy J, Nogami M (2011) A comparative study of Pt and
Pt-Pd core-shell nanocatalysts. Electrochim Acta 56:9133–9143
111. Zhang L, Zhang JW, Jiang ZY, Xie SF, Jin MS, Han XG, Kuang Q, Xie ZX, Zheng LS (2011)
Facile syntheses and electrocatalytic properties of porous Pd and its alloy nanospheres. J
Mater Chem 21:9620–9625
112. Jung DH, Bae SJ, Kim SJ, Nahm KS, Kim P (2011) Effect of the Pt precursor on the
morphology and catalytic performance of Pt-impregnated on Pd/C for the oxygen reduction
reaction in polymer electrolyte fuel cells. Int J Hydrogen Energy 36:9115–9122
113. Nguyen VL, Ohtaki M, Matsubara T, Cao MT, Nogami M (2012) New experimental
evidences of Pt-Pd bimetallic nanoparticles with core-shell configuration and highly fine-
ordered nanostructures by high-resolution electron transmission microscopy. J Phys Chem C
116:12265–12274
114. Liu L, Samjeske G, Nagamatsu S, Sekizawa O, Nagasawa K, Takao S, Imaizumi Y,
Yamamoto T, Uruga T, Iwasawa Y (2012) Enhanced oxygen reduction reaction activity
220 R. Zhang and W. Chen

and characterization of Pt-Pd/C bimetallic fuel cell catalysts with Pt-enriched surfaces in acid
media. J Phys Chem C 116:23453–23464
115. Li YJ, Wang ZW, Chiu CY, Ruan LY, Yang WB, Yang Y, Palmer RE, Huang Y (2012)
Synthesis of bimetallic Pt-Pd core-shell nanocrystals and their high electrocatalytic activity
modulated by Pd shell thickness. Nanoscale 4:845–851
116. Habas SE, Lee H, Radmilovic V, Somorjai GA, Yang P (2007) Shaping binary metal
nanocrystals through epitaxial seeded growth. Nat Mater 6:692–697
117. Guo SJ, Dong SJ, Wang EK (2010) Pt/Pd bimetallic nanotubes with petal-like surfaces for
enhanced catalytic activity and stability towards ethanol electrooxidation. Energy Environ
Sci 3:1307–1310
118. Lim BW, Lu XM, Jiang MJ, Camargo PHC, Cho EC, Lee EP, Xia YN (2008) Facile synthesis
of highly faceted multioctahedral Pt nanocrystals through controlled overgrowth. Nano Lett
8:4043–4047
119. Jiang MJ, Lim B, Tao J, Camargo PHC, Ma C, Zhu YM, Xia YN (2010) Epitaxial overgrowth
of platinum on palladium nanocrystals. Nanoscale 2:2406–2411
120. Bai S, Wang C, Deng M, Gong M, Bai Y, Jiang J, Xiong YJ (2014) Surface polarization
matters: enhancing the hydrogen-evolution reaction by shrinking Pt Shells in Pt–Pd–
Graphene stack structures. Angew Chem Int Ed 126:12316–12320
121. Zhang H, Jin MS, Wang JG, Kim MJ, Yang DR, Xia YN (2011) Nanocrystals composed of
alternating shells of Pd and Pt can be obtained by sequentially adding different precursors. J
Am Chem Soc 133:10422–10425
122. Lim B, Jiang MJ, Camargo PHC, Cho EC, Tao J, Lu XM, Zhu YM, Xia YN (2009) Pd-Pt
bimetallic nanodendrites with high activity for oxygen reduction. Science 324:1302–1305
123. Lee HJ, Habas SE, Somorjai GA, Yang PD (2008) Localized Pd overgrowth on cubic Pt
nanocrystals for enhanced electrocatalytic oxidation of formic acid. J Am Chem Soc
130:5406–5407
124. Lim B, Jiang MJ, Yu T, Camargo PHC, Xia YN (2010) Nucleation and growth mechanisms
for Pd-Pt bimetallic nanodendrites and their electrocatalytic properties. Nano Res 3:69–80
125. Guo SJ, Dong SJ, Wang EK (2010) Ultralong Pt-on-Pd bimetallic nanowires with nanoporous
surface: nanodendritic structure for enhanced electrocatalytic activity. Chem Commun
46:1869–1871
126. Peng ZM, Yang H (2009) Synthesis and oxygen reduction electrocatalytic property of Pt-on-
Pd bimetallic heteronanostructures. J Am Chem Soc 131:7542–7543
127. Guo SJ, Dong SJ, Wang EK (2010) Three-dimensional Pt-on-Pd bimetallic nanodendrites
supported on graphene nanosheet: facile synthesis and used as an advanced nanoelectro-
catalyst for methanol oxidation. ACS Nano 4:547–555
128. Wang L, Nemoto Y, Yamauchi Y (2011) Direct synthesis of spatially-controlled Pt-on-Pd
bimetallic nanodendrites with superior electrocatalytic activity. J Am Chem Soc 133:9674–
9677
129. Serpell CJ, Cookson J, Ozkaya D, Beer PD (2011) Core@shell bimetallic nanoparticle
synthesis via anion coordination. Nat Chem 3:478–483
130. Lee H, Habas SE, Kweskin S, Butcher D, Somorjai GA, Yang PD (2006) Morphological
control of catalytically active platinum nanocrystals. Angew Chem Int Ed 45:7824–7828
131. Climent V, Markovic NM, Ross PN (2000) Kinetics of oxygen reduction on an epitaxial film
of palladium on Pt(111). J Phys Chem B 104:3116–3120
132. Arenz M, Schmidt TJ, Wandelt K, Ross PN, Markovic NM (2003) The oxygen reduction
reaction on thin palladium films supported on a Pt(111) electrode. J Phys Chem B 107:9813–
9819
133. Babu PK, Kim HS, Chung JH, Oldfield E, Wieckowski A (2004) Bonding and motional
aspects of CO adsorbed on the surface of Pt nanoparticles decorated with Pd. J Phys Chem B
108:20228–20232
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 221

134. Zhao MC, Rice C, Masel RI, Waszczuk P, Wieckowski A (2004) Kinetic study of electro-
oxidation of formic acid on spontaneously-deposited Pt/Pd nanoparticles-CO tolerant fuel
cell chemistry. J Electrochem Soc 151:A131–A136
135. Jayashree RS, Spendelow JS, Yeom J, Rastogi C, Shannon MA, Kenis PJA (2005) Charac-
terization and application of electrodeposited Pt, Pt/Pd, and Pd catalyst structures for direct
formic acid micro fuel cells. Electrochim Acta 50:4674–4682
136. Grace AN, Pandian K (2006) Pt, Pt-Pd and Pt-Pd/Ru nanoparticles entrapped polyaniline
electrodes—a potent electrocatalyst towards the oxidation of glycerol. Electrochem Commun
8:1340–1348
137. Selvaraj V, Alagar M, Hamerton I (2006) Electrocatalytic properties of monometallic and
bimetallic nanoparticles-incorporated polypyrrole films for electro-oxidation of methanol. J
Power Sources 160:940–948
138. Selvaraj V, Alagar M, Hamerton I (2007) Nanocatalysts impregnated polythiophene elec-
trodes for the electrooxidation of formic acid. Appl Catal B Environ 73:172–179
139. Xu Y, Lin X (2007) Facile fabrication and electrocatalytic activity of Pt0.9 Pd0.1alloy film
catalysts. J Power Sources 170:13–19
140. Galal A, Atta NF, Darwish SA, Ali SM (2008) Electrodeposited metals at conducting
polymer electrodes. II: Study of the oxidation of methanol at poly(3-methylthiophene)
modified with Pt-PdCo-catalyst. Top Catal 47:73–83
141. Zhou ZL, Kang TF, Zhang Y, Cheng SY (2009) Electrochemical sensor for formaldehyde
based on Pt-Pd nanoparticles and a Nafion-modified glassy carbon electrode. Microchim Acta
164:133–138
142. Mahapatra SS, Dutta A, Datta J (2010) Temperature effect on the electrode kinetics of
ethanol oxidation on Pd modified Pt electrodes and the estimation of intermediates formed
in alkali medium. Electrochim Acta 55:9097–9104
143. Mahapatra SS, Dutta A, Datta J (2011) Temperature dependence on methanol oxidation and
product formation on Pt and Pd modified Pt electrodes in alkaline medium. Int J Hydrogen
Energy 36:14873–14883
144. Yang X, Yang QD, Xu J, Lee CS (2012) Bimetallic PtPd nanoparticles on Nafion-graphene
film as catalyst for ethanol electro-oxidation. J Mater Chem 22:8057–8062
145. Vidal-Iglesias FJ, Solla-Gullon J, Herrero E, Aldaz A, Feliu JM (2010) Pd adatom decorated
(100) preferentially oriented Pt nanoparticles for formic acid electrooxidation. Angew Chem
Int Ed 49:6998–7001
146. Zhang JL, Vukmirovic MB, Xu Y, Mavrikakis M, Adzic RR (2005) Controlling the catalytic
activity of platinum-monolayer electrocatalysts for oxygen reduction with different sub-
strates. Angew Chem Int Ed 44:2132–2135
147. Zhang J, Mo Y, Vukmirovic M, Klie R, Sasaki K, Adzic R (2004) Platinum monolayer
electrocatalysts for O2 reduction: Pt monolayer on Pd (111) and on carbon-supported Pd
nanoparticles. J Phys Chem B 108:10955–10964
148. Adzic RR, Zhang J, Shao M, Sasaki K, Vukmirovic M, Uribe FA (2006) Platinum and mixed
platinum-metal monolayer fuel cell electrocatalysts: design, activity and long-term perfor-
mance stability. ECS Trans 3:31–36
149. Vukmirovic MB, Zhang J, Sasaki K, Nilekar AU, Uribe F, Mavrikakis M, Adzic RR (2007)
Platinum monolayer electrocatalysts for oxygen reduction. Electrochim Acta 52:2257–2263
150. Bliznakov ST, Vukmirovic MB, Yang L, Sutter EA, Adzic RR (2012) Pt Monolayer on
electrodeposited Pd nanostructures: advanced cathode catalysts for PEM fuel cells. J
Electrochem Soc 159:F501–F506
151. Sasaki K, Naohara H, Cai Y, Choi YM, Liu P, Vukmirovic MB, Wang JX, Adzic RR (2010)
Core-protected platinum monolayer shell high-stability electrocatalysts for fuel-cell cath-
odes. Angew Chem Int Ed 49:8602–8607
152. Koenigsmann C, Santulli AC, Gong KP, Vukmirovic MB, Zhou WP, Sutter E, Wong SS,
Adzic RR (2011) Enhanced electrocatalytic performance of processed, ultrathin, supported
222 R. Zhang and W. Chen

Pd-Pt core-shell nanowire catalysts for the oxygen reduction reaction. J Am Chem Soc
133:9783–9795
153. Wang JX, Inada H, Wu LJ, Zhu YM, Choi YM, Liu P, Zhou WP, Adzic RR (2009) Oxygen
reduction on well-defined core-shell nanocatalysts: particle size, facet, and Pt shell thickness
effects. J Am Chem Soc 131:17298–17302
154. Shao MH, He GN, Peles A, Odell JH, Zeng J, Su D, Tao J, Yu T, Zhu YM, Xia YN (2013)
Manipulating the oxygen reduction activity of platinum shells with shape-controlled palla-
dium nanocrystal cores. Chem Commun 49:9030–9032
155. Seweryn J, Lewera A (2012) Electrooxidation of ethanol on carbon-supported Pt-Pd
nanoparticles. J Power Sources 205:264–271
156. Ignaszak A, Song C, Zhu W, Wang YJ, Zhang J, Bauer A, Baker R, Neburchilov V, Ye S,
Campbell S (2012) Carbon–Nb0.07Ti 0.93O2 composite supported Pt–Pd electrocatalysts for
PEM fuel cell oxygen reduction reaction. Electrochim Acta 75:220–228
157. Lopes PP, Ticianelli EA, Varela H (2011) Potential oscillations in a proton exchange
membrane fuel cell with a Pd-Pt/C anode. J Power Sources 196:84–89
158. Kim KH, Yu JK, Lee HS, Choi JH, Noh SY, Yoon SK, Lee CS, Hwang TS, Rhee YW (2007)
Preparation of Pt-Pd catalysts for direct formic acid fuel cell and their characteristics. Korean
J Chem Eng 24:518–521
159. Parinyaswan A, Pongstabodee S, Luengnaruemitchai A (2006) Catalytic performances of
Pt-Pd/CeO2 catalysts for selective CO oxidation. Int J Hydrogen Energy 31:1942–1949
160. Koutsopoulos S, Johannessen T, Eriksen KM, Fehrmann R (2006) Titania-supported Pt and
Pt-Pd nanoparticle catalysts for the oxidation of sulfur dioxide. J Catal 238:206–213
161. Mougenot M, Caillard A, Brault P, Baranton S, Coutanceau C (2011) High performance
plasma sputtered PdPt fuel cell electrodes with ultra low loading. Int J Hydrogen Energy
36:8429–8434
162. Schmidt TJ, Markovic NM, Stamenkovic V, Ross PN, Attard GA, Watson DJ (2002) Surface
characterization and electrochemical behavior of well-defined Pt-Pd{111} single-crystal
surfaces: a comparative study using Pt{111} and palladium-modified Pt{111} electrodes.
Langmuir 18:6969–6975
163. Ohashi M, Beard KD, Ma SG, Blom DA, St-Pierre J, Van Zee JW, Monnier JR (2010)
Electrochemical and structural characterization of carbon-supported Pt-Pd bimetallic
electrocatalysts prepared by electroless deposition. Electrochim Acta 55:7376–7384
164. Wang C, Peng B, Xie HN, Zhang HX, Shi FF, Cai WB (2009) Facile fabrication of Pt, Pd and
Pt-Pd alloy films on Si with tunable infrared internal reflection absorption and synergetic
electrocatalysis. J Phys Chem C 113:13841–13846
165. Cangul B, Zhang LC, Aindow M, Erkey C (2009) Preparation of carbon black supported Pd,
Pt and Pd-Pt nanoparticles using supercritical CO2 deposition. J Supercrit Fluids 50:82–90
166. Yen CH, Shimizu K, Lin YY, Bailey F, Cheng IF, Wai CM (2007) Chemical fluid deposition
of Pt-based bimetallic nanoparticles on multiwalled carbon nanotubes for direct methanol
fuel cell application. Energy Fuel 21:2268–2271
167. Arico AS, Srinivasan S, Antonucci V (2001) DMFCs: from fundamental aspects to technol-
ogy development. Fuel Cells 1:133–161
168. Choi WC, Jeon MK, Kim YJ, Woo SI, Hong WH (2004) Development of enhanced materials
for direct-methanol fuel cell by combinatorial method and nanoscience. Catal Today 93–
95:517–522
169. Koenigsmann C, Wong SS (2011) One-dimensional noble metal electrocatalysts: a promising
structural paradigm for direct methanol fuel cells. Energy Environ Sci 4:1161–1176
170. Shukla AK, Ravikumar MK, Gandhi KS (1998) Direct methanol fuel cells for vehicular
applications. J Solid State Electrochem 2:117–122
171. Leger JM (2001) Mechanistic aspects of methanol oxidation on platinum-based
electrocatalysts. J Appl Electrochem 31:767–771
172. Zhang XY, Lu W, Da JY, Wang HT, Zhao DY, Webley PA (2009) Porous platinum nanowire
arrays for direct ethanol fuel cell applications. Chem Commun 2:195–197
4 Synthesis and Electrocatalysis of Pt-Pd Bimetallic Nanocrystals for Fuel Cells 223

173. Arenz M, Stamenkovic V, Ross PN, Markovic NM (2004) Surface (electro-)chemistry on Pt


(111) modified by a pseudomorphic Pd monolayer. Surf Sci 573:57–66
174. Qian W, Hao R, Zhou J, Eastman M, Manhat BA, Sun Q, Goforth AM, Jiao J (2013)
Exfoliated graphene-supported Pt and Pt-based alloys as electrocatalysts for direct methanol
fuel cells. Carbon 52:595–604
175. Feng JX, Zhang QL, Wang AJ, Wei J, Chen JR, Feng JJ (2014) Caffeine-assisted facile
synthesis of platinum@palladium core-shell nanoparticles supported on reduced graphene
oxide with enhanced electrocatalytic activity for methanol oxidation. Electrochim Acta
142:343–350
176. Bianchini C, Shen PK (2009) Palladium-based electrocatalysts for alcohol oxidation in half
cells and in direct alcohol fuel cells. Chem Rev 109:4183–4206
177. Lamy C, Rousseau S, Belgsir EM, Coutanceau C, Leger JM (2004) Recent progress in the
direct ethanol fuel cell: development of new platinum-tin electrocatalysts. Electrochim Acta
49:3901–3908
178. Simoes FC, dos Anjos DM, Vigier F, Leger JM, Hahn F, Coutanceau C, Gonzalez ER,
Tremiliosi G, de Andrade AR, Olivi P, Kokoh KB (2007) Electroactivity of tin modified
platinum electrodes for ethanol electrooxidation. J Power Sources 167:1–10
179. Yu XW, Pickup PG (2008) Recent advances in direct formic acid fuel cells (DFAFC). J
Power Sources 182:124–132
180. Rees NV, Compton RG (2011) Sustainable energy: a review of formic acid electrochemical
fuel cells. J Solid State Electrochem 15:2095–2100
181. Rees NV, Compton RG (2012) Sustainable energy: a review of formic acid electrochemical
fuel cells. J Solid State Electrochem 16:419–419
182. Grigoriev SA, Lyutikova EK, Martemianov S, Fateev VN (2007) On the possibility of
replacement of Pt by Pd in a hydrogen electrode of PEM fuel cells. Int J Hydrogen Energy
32:4438–4442
183. Baranova EA, Miles N, Mercier PH, Le Page Y, Patarachao B (2010) Formic acid electro-
oxidation on carbon supported PdxPt1− x(0≥x≥1) nanoparticles synthesized via modified
polyol method. Electrochim Acta 55:8182–8188
184. Guo SJ, Zhang S, Sun SH (2013) Tuning nanoparticle catalysis for the oxygen reduction
reaction. Angew Chem Int Ed 52:8526–8544
185. Jiang YY, Lu YZ, Lv XY, Han DX, Zhang QX, Niu L, Chen W (2013) Enhanced catalytic
performance of Pt-free iron phthalocyanine by graphene support for efficient oxygen reduc-
tion reaction. ACS Catal 3:1263–1271
186. Bliznakov ST, Vukmirovic MB, Yang L, Sutter EA, Adzic RR (2011) Pt monolayer on
electrodeposited Pd nanostructures-advanced cathode catalysts for PEM fuel cells. ECS
Trans 41:1055–1066
187. Sasaki K, Wang JX, Naohara H, Marinkovic N, More K, Inada H, Adzic RR (2010) Recent
advances in platinum monolayer electrocatalysts for oxygen reduction reaction: Scale-up
synthesis, structure and activity of Pt shells on Pd cores. Electrochim Acta 55:2645–2652
Chapter 5
Integrated Studies of Au@Pt and Ru@Pt
Core-Shell Nanoparticles by In Situ
Electrochemical NMR, ATR-SEIRAS,
and SERS

Dejun Chen, Dianne O. Atienza, and YuYe J. Tong

5.1 Introduction

Nanoscale noble metal materials, mostly in the form of nanoparticles (NPs), have
been subjected to intensive research in the context of their applications to fuel cell
electrocatalysis [1] traditionally and also to lithium-air batteries [2, 3] recently.
Among others, despite decades-long continuous efforts in discovering viable
replacements for the expensive and earthly-scant platinum (Pt), Pt-based
electrocatalysts [4, 5] are still the materials of choice in terms of having high
activity for CO oxidation reaction (COR) [6, 7], methanol oxidation reaction
(MOR) [8], ethanol oxidation reaction (EtOR) [9, 10], formic acid oxidation
reaction (FAOR) [11, 12], hydrogen evolution/oxidation reaction (HER/HOR)
[13, 14], oxygen reduction/evolution reaction (ORR/OER) [15–17], and CO2
reduction reaction (CO2R) [18, 19], just to name a few widely studied reactions.
Benefited from the fertile knowledge gained from the decades-long intensive
investigations of model electrocatalysts, such as single crystal surfaces, fundamen-
tal research on discovering, understanding and utilizing novel electronic, geomet-
ric, and/or bifunctional properties of Pt-based NPs [20–22] for the aforementioned
reactions continues to flourish and push the boundary of research in an environment
as close to that of practical applications as possible. Constant advance of the
sophistication of many in situ spectroelectrochemical techniques has been a critical
part of the latter. In this Chapter, we will discuss integrated studies of Au@Pt and
Ru@Pt NPs by in situ electrochemical (EC) nuclear magnetic resonance (NMR),
attenuated-total-reflection surface-enhanced IR reflection adsorption spectroscopy
(ATR-SEIRAS), and surface-enhanced Raman scattering spectroscopy (SERS)

D. Chen • D.O. Atienza • Y.Y.J. Tong (*)


Department of Chemistry, Georgetown University, 37th and O Streets, NW,
Washington, DC 20057, USA
e-mail: [email protected]

© Springer International Publishing Switzerland 2016 225


K.I. Ozoemena, S. Chen (eds.), Nanomaterials for Fuel Cell Catalysis,
Nanostructure Science and Technology, DOI 10.1007/978-3-319-29930-3_5
226 D. Chen et al.

spectroscopy, with a focus on how an organic integration of these techniques can


enhance the investigative power brought to bear by their complementary nature.
That is, an organically integrated whole is better than the sum of its parts.
The strength of NMR is in its unique ability to probe the electronic and structural
properties of Pt-based NPs [23]. Particularly, 195Pt NMR can distinguish, at least
semi-quantitatively, the Pt atoms on the surface of from those inside the NPs and
the Pt atoms in metallic from non-metallic (e.g. oxidized) state [24, 25]. Moreover,
using adsorbed 13CO as a molecular probe, 13C NMR can offer complementary
electronic and diffusional information of the surface [26–28]. However, NMR
continues to suffer from its intrinsic low mass detection sensitivity so it is generally
not a method of choice for detecting reaction intermediates during the reaction. On
the other hand, SEIRAS and SERS are more suited for the latter [29, 30].
Since its inception [31], EC-SEIRAS has been applied extensively to study
electrocatalysis mainly on conductive metal thin-films (~50 to 100 nm thick) used
as working electrodes [32]. Single metallic films, such as Au [32–34], Pt [35–43],
Pd [44, 45], Cu [46], Ag [47], and bimetallic PtRu [43, 48], PtPd [44] alloy film as
well as additional ultrathin-film (bi)metallic layers deposited on Au substrate film
[49, 50], have been the subjects of studies of COR [34, 41], MOR [40, 42, 43, 48],
EtOR [36], FAOR [39, 44, 45, 50], HER/HOR [35, 38], ORR [51, 52] and CO2R
[37]. Numerous surface species that are important to the reactions under investiga-
tion, such as adsorbed CO (COads) in linearly/bridge/multiple bonding configura-
tions [37, 43, 45, 50], formate [39, 40, 45, 53], adsorbed anion [33, 45] and
interfacial water [32, 54, 55], have been identified and studied [30]. On the other
hand, while SERS may not be as versatile and sensitive as SEIRAS, EC-SERS does
offer some complementary advantages, such as observing species that only have
symmetric vibrational modes, for instance O2 [56], accessing to lower
wavenumber range (100~1000 cm1) in which the vibrations of metal-adsorbate,
such as Pt-CO [57] or Pt-O [58], locate and not worrying about strong absorption by
water. With the enhancement offered by roughened bulk surfaces [29, 59] or
deposited NPs [60–62], EC-SERS has become increasingly used to study catalytic
reactions, such as COR [57, 60, 63, 64], MOR [57, 65], FAOR [62], ORR [56] and
H2O2 electro-reduction reaction [66, 67], on the respective electrocatalyst surface.
Despite their complementary nature, EC-NMR, -SEIRAS or -SERS has been
largely used alone for various reasons. But an integrated approach of combining
them together to investigate the same system is expected to provide a fuller
mechanistic picture of the system under investigation. With this in mind, and as
already briefly mentioned above, we will focus here on two samples, Ru@Pt and
Au@Pt NPs, which have been studied in various degree of detail by in situ
EC-NMR, -SEIRAS, and -SERS. This chapter will be structured as follows: we
will first present briefly the physical aspects of the samples (Sect. 5.2), then move to
NMR (Sect. 5.3), SEIRAS (Sect. 5.4), and SERS (Sect. 5.5) sequentially, which
will be followed by an integrated discussion (Sect. 5.6) of the results presented in
the previous sections. We will conclude the chapter with some summative
comments.
5 Integrated Studies of Au@Pt and Ru@Pt Core-Shell Nanoparticles. . . 227

5.2 Ru@Pt and Au@Pt NPs Samples

Ru@Pt NPs. The synthesis was adopted from the wet homogeneous reaction-like
procedure developed by Du et al. [68]. Briefly, the ethylene glycol (EG) was used as
both reaction medium and reducing agent of Johnson-Matthey (J-M) Ru Black
(~3 nm) and H2PtCl66H2O. The first step is the complete reduction of Ru NPs at
140  C for 2 h with vigorous stirring. This step is critical for the success of
controlling the surface Pt deposition where it has been determined experimentally
that Ru readily oxidizes to RuOx in the atmosphere [69]. The cleaned Ru NPs were
then collected and rinsed with ethanol and re-dispersed ultrasonically into a fresh
EG solution. The calculated volume of H2PtCl66H2O solution according to the
experimentally determined linear relationship [68] gave the desired 40 % Pt
packing density (PD). This mixture was ultrasonicated and heated at 120  C for
4 h to reduce Pt(IV) to Pt(0) on the surface of cleaned Ru NPs. This synthetic
method in a lower temperature homogeneous-reaction like media is advantageous
in preventing sintering effects. A representative TEM image is shown in Fig. 5.1a
from which the average particle size was determined to be 3.2 nm.
Au@Pt NPs. Au colloid NPs (3.5 nm) were first prepared from HAuCl43H2O in
aqueous solution using sodium citrate and sodium borohydride as reductant and
then supported spontaneously on Vulcan XC-72R that gave a metal loading of 30 wt
%. The Pt shell was deposited by reducing Pt4+ onto the surface of dispersed
carbon-supported Au NPs by using ascorbic acid with target Pt coverage of
100 % (atomic ratio of Pt/Au ¼ 1). The detail of the synthesis can be found in the
original paper [70]. A representative TEM image is shown in Fig. 5.1b that gives an
average particle size of 4.6 nm.

Fig. 5.1 TEM images of Ru@Pt (a) and Au@Pt (b). The average particle size is 3.2 nm for the
former and 4.6 nm for the latter. (b) is adapted with permission from [70] (© 2007 Elsevier)
228 D. Chen et al.

5.3 In Situ EC-NMR Study of Ru@Pt and Au@Pt NPs [71]

For EC-NMR measurements, the NMR samples (~50–80 mg) were treated at room
temperature by an EC cleaning process in a three-electrode setup with the NPs as
working electrode by holding the potential at 50 mV (vs. Ag/AgCl). For CO
adsorption from dissociation of 13C-labeled methanol (13C-MeOH) onto the
pre-clean NPs surface was done with the same cell but in 0.1 M HClO4 + 0.5 M
MeOH solution by holding the potential at 250 mV. The EC setup was under a
continuous stream of ultrapure Argon gas during the entire period of the above
treatments. After the EC cleaning or CO adsorption, the NMR sample cell filled
with the supporting electrolyte (0.1 M HClO4 or 0.1 M HClO4 + 0.5 M 13C-MeOH)
was sealed immediately with a one-to-one fitted ground-glass stopper, which was
then inserted into the NMR probe and lowered down to the cryostat that was
pre-cooled at 80 K. After the NMR measurements, the cell was immediately
reconnected back to the EC cleaning setup. The open-circuit potential was
re-checked to make sure that its stability was within a 20 mV difference before
and after the NMR measurements. The 13C NMR spectra of the adsorbed 13CO and
the corresponding inversion-recovery spin-lattice T1 measurements were obtained
at various temperature ranges of 80–120 K at a “home-assembled” spectrometer
equipped with an active-shielded 9.395 T widebore superconducting magnet, an
Oxford Spectrostat-CF cryostat (Oxford Instrument, U.K.), an AMT (Lancaster,
PA) 1 kW power amplifier, a Tecmag (Houston, TX) Apollo data acquisition
system, and a home-built single channel solenoid probe with a coil of 5 mm inner
diameter and 28 mm length. Both the 13C and 195Pt spectra were acquired using the
conventional “π/2-τ-π-τ-echo” Hahn spin-echo sequence with the π/2 pulse length
and τ of 3 μs and 25 μs respectively but for the 195Pt spectra the frequency was
manually changed between ca. 82–87 MHz with a 0.047 G/kHz increment between
each resonance position at 80 K because of the extremely broadness of the spectra.
The slow beats of the 195Pt signal at a given frequency (see Fig. 5.2e, f) are obtained
by varying the τ in the spin-echo sequence and the corresponding T1 measurements
were done with the conventional saturation-recovery method.
Figure 5.2a, b shows the point-by-point 195Pt NMR spectra of the as-received
(red) and EC-cleaned (blue) samples of Au@Pt (with Pt packing density (PD) ¼
3.8, A) and of Ru@Pt (with Pt PD ¼ 0.4, D) core-shell NPs in comparison with that
of the pure Pt/C NPs [73] (black dashed lines and open circles). The vertical dashed
line indicates the surface peak position (1.1000 G/kHz,) of the EC-cleaned pure
2.5 nm Pt/C NPs [73]. Although the frozen electrolyte at 80 K prevented active
potentiostatic control, the tightly vacuum-sealed sample cell ensured that the
sample surface potential would not change during the NMR measurements. This
was verified by measuring open circuit potentials before and after each NMR
measurement that always indicated that a constant potential was indeed achieved.
The as-received Au@Pt NPs had a shelf life of more than 48 months [70],
therefore, the surface was heavily oxidized, as indicated by the peak of Pt oxide
at 1.089 G/kHz (Fig. 5.2a). Notice that distinguishing a Pt surface covered by
5 Integrated Studies of Au@Pt and Ru@Pt Core-Shell Nanoparticles. . . 229

Fig. 5.2 Point-by-point 195Pt NMR spectra (a, b), spin–lattice relaxation curves (c) and (d), spin-
spin relaxation (J-coupling modulation or slow-beat) curves (e, f) for the Au@Pt (a, c, e) and
Ru@Pt (b, d, f) NPs, respectively. The calculated Pt PD (packing density) was 3.8 for the Au@Pt
and 0.4 for the Ru@Pt NPs. According to the straight line in Fig. 3 of [72], the expected Pt surface
fraction would be 1.3, i.e., a fully Pt covered Au core. The % of the total deposited Pt on the surface
would be ~1.3/3.8 ¼ 0.34, in an excellent agreement with the value of 33 % determined by 195Pt
NMR in a. Adapted from [71] with permission (© 2012 American Chemical Society)

adsorbed oxygen from that by Pt oxide is straightforward. After the EC cleaning,


the Pt oxide peak disappeared and a clean-surface peak appeared at 1.0975 G/kHz.
The latter amounts to a 2278-ppm positive shift with respect to the clean-surface
peak position (1.1000 G/kHz) of the pure Pt NPs, manifesting a strong electronic
effect of Au on Pt. The significant portion of the 195Pt atoms in the Au@Pt NPs
resonating at higher fields (above 1.1000 G/kHz) indicates that many deposited Pt
atoms went inside the Au NPs. Indeed, using a Gaussian de-convolution of the
surface peak (the dashed blue curve in Fig. 5.2a), we found that only 33 % of all Pt
deposited atoms were on the surface of the Au@Pt NPs, which is in agreement with
expected atomic exchange between deposited Pt atoms and the underlying Au
atoms [72]. Notice that such a ~33 % value is actually predicted by the straight
line in Fig. 3 of [72] (see caption of Fig. 5.2).
The as-received Ru@Pt NPs were freshly synthesized with a shelf life of only a
few hours [68]. Under the protection of ethylene glycol, little surface oxidation was
expected in such a short time. Indeed, no significant surface oxidation was observed
in the corresponding 195Pt NMR spectrum in Fig. 5.2b (red). The clean-surface peak
(blue line) now appeared at 1.1053 G/kHz, corresponding to a 4795-ppm negative
shift with respect to the clean-surface peak position (1.1000 G/kHz) of the pure Pt
NPs, revealing also a strong electronic effect of Ru on Pt as well, which is opposite
to that of Au. A Gaussian de-convolution of the surface peak gave a surface atom
fraction of 74 %, indicating a dominant Ru core Pt shell structure.
230 D. Chen et al.

The nuclear spin-lattice (T1) and spin-spin (T2) relaxation measurements are
presented in Fig. 5.2c, e for the EC-cleaned Au@Pt and in 2d and 2f for the
EC-cleaned Ru@Pt, respectively, at the spectral positions listed in the respective
figures. Detailed discussion of the relaxation results is out of scope here due to the
limited space, but it suffices to demonstrate the most important difference here.
Both the T1 (Fig. 5.2d) and T2 (Fig. 5.2f) at different spectral positions on the
Ru@Pt fall into the same respective relaxation curves, which is a manifestation of
the same chemical environment, i.e., the majority of the Pt atoms were on the
surface. Particularly, a J-coupling modulation analysis [25] of the data in Fig. 5.2f
by fitting to the J-coupling induced slow beats equation [74] S(τ)/S0 ¼ exp(–2τ/T2)
(P0 + exp(–(τ/T2J)2)[P1cos(Jτ) + P2cos2(Jτ)]) with J for J-coupling constant, 1/T2J
for spread in J, and P0 + P1 + P2 ¼ 1 (solid line) gives a Pt atomic fraction of 0.37.
Since for a full pseudo-morphic Pt monolayer on a Ru(0 0 0 1) surface, the Pt atomic
fraction among all the possible next nearest neighbors is 6/9 ¼ 0.67, a local Pt
atomic fraction of 0.37 is thus consistent with a Pt PD of 0.4 of mono-atomic Pt
islands. In contrast, both the T1 (Fig. 5.2c) and T2 (Fig. 5.2e) at different spectral
positions on the Au@Pt were different, which indicates a wider distribution of Pt
atoms on surface and inside the NPs as well.
From the T1 value, the known spectral position and temperature at which the T1
is measured, one can use the two-band model to calculate the s- and d-like Ef–
LDOSs, denoted here as Ds(Ef) and Dd(Ef) respectively [75]. These values for the
surface Pt atoms in Au@Pt and Ru@Pt samples are collected in Table 5.1. As can
be clearly seen, Au and Ru cores exert different (opposing) electronic influence on
Pt shell: the former lowers more the d-like while the latter lowers more the s-like
Ef–LDOSs.
Also collected (in the parentheses) in Table 5.1 are the average surface values
calculated by using the OpenMX DFT package [76] on Pt13@Pt42 (Pt55),
Ru13@Pt42, and Au13@Pt42 cubo-octahedral clusters. Considering the very sim-
plified cluster models, it is remarkable that the experiments and theoretical calcu-
lations have achieved a good agreement in numerical values and in trend for both
Ds(Ef) and Dd(Ef).
How changes in the surface Pt s- and d-like Ef-LDOS induced by the Au and
Ru cores influence metal-adsorbate bonding was probed and corroborated by 13C
NMR of 13CO adsorbed onto the Au@Pt and Ru@Pt NPs via dissociative adsorp-
tion of 13CH3OH (MeOH). The ability to probe the surface electronic properties by

Table 5.1 Ef-LDOS values deduced from 195


Pt and 13
C NMR data (1 Ry ¼ 13.6 eV and
mol ¼ molecule)
Ds(Ef)/Ry1 Dd(Ef)/Ry1 D5σ(Ef)/Ry1 D2π*(Ef)/Ry1
Samples atom1 atom1 mol1 mol1
Ru@Pt 3.7 (2.9) 13.3 (12.9) 0.7 4.9
Au@Pt 4.3 (3.0) 10.1 (12.7) 0.6 5.8
Pt 5.1 (3.4) 13.5 (15.2) 0.7 6.7
5 Integrated Studies of Au@Pt and Ru@Pt Core-Shell Nanoparticles. . . 231

Fig. 5.3 13C NMR spectra (9.39 T, 80 K) of chemisorbed CO on (a) Au@Pt and (b) Ru@Pt NPs.
Electro-deposition of CO was performed by holding the potential at 0.35 V vs. Ag/AgCl (3 M)
while immersed in 0.5 M D2ClO4 with 13CH3OH [13]. CH3OH at 49 ppm was also used as an
internal shift reference. (c) Temperature dependent T1 relaxation measurements obtained at
333 ppm for CO-Pt/Ru@Pt and CO-Pt/Au@Pt and at 219 ppm for CO-Ru/Ru@Pt. Adapted
from [71] with permission (© 2012 American Chemical Society). Inset in (c): Linear correlation
between the D2π*(Ef) and Ds(Ef)

using surface-adsorbed molecules, such as adsorbed 13CO, is among the unique


investigative strengths of in situ EC-NMR. Figure 5.3 shows the 13C NMR spectra
of 13CO on the Au@Pt (a) and Ru@Pt (b) NPs, with the assigned peaks indicated by
the arrows. For the Au@Pt sample, despite the dominant peaks from the remaining
MeOH in the electrolyte and the carbon support, the broad and weak—yet still
discernable—13C NMR peak at 333 ppm was assigned to the adsorbed CO, an
assignment corroborated by the temperature dependent T1 measurements (vide
infra). For the Ru@Pt sample, the peaks at 300 ppm and 219 ppm can be
assigned to CO on Pt and on Ru sites, respectively, based on the literature values
[27, 77]. The difference in the signal strength of the adsorbed CO on the Au@Pt and
Ru@Pt NPs may reflect the difference in dissociative adsorption of MeOH on the
respective surfaces.
The results of temperature dependent T1 measurements are presented in
Fig. 5.3c. For the CO on Pt, the T1’s were measured at 333 ppm for both Au@Pt
(T1T ¼ 83 s K) and Ru@Pt (T1T ¼ 99 s K) samples to facilitate comparison,
although the peak position for the latter was at  300 ppm. The pass-through-origin
straight lines are the hallmark of the Korringa relaxation behavior [78], which
indicates that the adsorbed CO molecules on three different sites all acquired
metallic characteristics through surface bonding. For CO on Ru sites, the shift
(219 ppm) and Korringa constant T1T ¼ 138 s K (where T is the absolute temper-
ature at which T1 is measured) are very close to those observed on pure Ru [77] and
on Ru deposited on Pt NPs [27], which lends strong support to our peak assignment
that is also consistent with the expected exposure of Ru core for a Pt PD of 40 %. On
232 D. Chen et al.

the other hand, the Korringa constants measured at the same spectral position
(333 ppm) were different for CO on the surface Pt atoms of the Au@Pt and on
those of the Ru@Pt NPs. The former showed a faster relaxation (i.e., smaller
Korringa constant T1T ), and thus a stronger metal-adsorbate bonding. In fact for
CO on Pt, we can also deduce the values of the respective 5σ- and 2π*-like Ef–
LDOS, i.e., D5σ(Ef) and D2π*(Ef), at 13C by using a different two-band model
developed for adsorbed CO on Pt [26].
As can be seen from the data in Table 5.1, the D5σ(Ef) was essentially constant,
while the D2π*(Ef) decreased proportionally (inset in Fig. 5.3c) with the Ds(Ef) for
CO on Pt from pure Pt to Au@Pt to Ru@Pt. This suggests that the electronic
alterations on metal-CO bonding, i.e., changes in metal to CO 2π* back-donation
[79], and therefore in bonding strength, were mainly caused by the variations in the
Ds(Ef) of the surface Pt atoms. On the other hand, it has been observed that the
ability to dissociatively adsorb MeOH that leads to adsorbed CO follows the order
Pt  Ru@Pt > Au@Pt, which shares the same order in the respective Dd(Ef) as
shown in Table 5.1. These results suggest that the d-like electrons are probably
responsible for agostic interaction that activates the three methyl protons in MeOH
during its dissociative adsorption.

5.4 In Situ EC ATR-SEIRAS of Ru@Pt and Au@Pt NPs


[80]

Figure 5.4 shows the schematic of the EC IR cell that was used for in situ EC
ATR-SEIRAS measurements of the Ru@Pt and Au@Pt samples. A gold film was
first electrolessly deposited onto the reflecting plane of a Si attenuation total
reflection (ATR) prism of a triangular shape, which was polished with successively

Fig. 5.4 Schematic of the in situ EC ATR-SEIRAS cell used for ATR-SEIRAS measurements.
WE, RE and CE stand for working, reference and counter electrode respectively. A ~50-nm Au
film was first deposited onto the pre-polished Si prism onto which the Ru@Pt or Au@Pt NPs were
then drop-casted and dried with nitrogen flow
5 Integrated Studies of Au@Pt and Ru@Pt Core-Shell Nanoparticles. . . 233

finer grade alumina slurries down to 0.3 μm and cleaned by sonication in the Milli-
Q water. The detailed deposition procedure can be found elsewhere [81].
For the in situ EC-SEIRAS measurements, Ru@Pt (1.05 mg) and Au@Pt/C
(0.18 mg) were dispersed onto the as-prepared gold film. The catalytic adlayer of
the core-shell NPs was then dried in a gentle nitrogen flow at room temperature. No
Nafion® solution was used in order to avoid any optical interference. The working
electrode surface (Au film plus NPs) was cleaned by repetitive potential cycling
until stable and reproducible cyclic voltammograms (CV) were achieved in 0.1 M
HClO4 at 50 mV s1 between 0.3 V and 0.7 V for the Ru@Pt, and 0.25 V and
1.0 V for the Au@Pt/C NPs, respectively. Notice that the Ru@Pt had an expected Pt
PD or coverage of 40 % while that for the Au@Pt was 100 %.
Figure 5.5a, b present the SEIRAS spectra of two sequential COR potential scans
after gaseous CO adsorption on the Ru@Pt (a-1 and a-2) and the Au@Pt (b-1 and
b-2) NPs. The reference spectra were taken at 0.7 V for the former and 1.0 V for the
latter at which all adsorbed CO had been fully oxidized according to the
corresponding CO stripping voltammograms. While the same IR band of linearly-
bound CO-Pt was observed on the Au@Pt NPs for the two sequential COR runs
(Fig. 5.5b, d), there were changes for the CO spectra on the Ru@Pt NPs. In (a-1),
the spectra show a series of a single, potential-dependent CO IR band that varied
between 1998 and 2047 cm1, which can be assigned to a linearly-bonded COL
[43, 82]. The peak position of this single band at 0.15 V (spectrum (a) in the inset
of Fig. 5.5c) is 2005 cm1, a value in agreement with those of CO adsorbed on the
Ru sites of Ru-decorated Pt(1 1 1) surfaces (from 2001 to 2008 cm1) [82–84] but
between those of COL on pure Ru (~1995 cm1) [43, 82] and on PtRu alloy
(~2044 cm1) [43, 85] under the similar conditions. For CO absorbed on the
Ru-modified Pt(1 1 1) surfaces, two distinct CO stretching bands corresponding to
Pt-COL and Ru-COL respectively were observed; but on the Ru@Pt of the first COR
in this study and on PtRu alloys [43, 85, 86] in general, only one band was observed.
On the other hand, the SEIRAS spectra obtained between 0.3 to 0.3 V of the
CO during the second COR on the Ru@Pt can be reasonably de-convoluted into
three peaks whose values at 0.15 V are 1948 cm1, 2000 cm1 and 2031 cm1
respectively (spectrum b in the inset of Fig. 5.5c). Based on the available literature
values and assignments of the C-O stretching frequencies of adsorbed CO on Ru
[43], Ru-decorated Pt [82, 83, 87], and PtRu alloy [43, 88, 89] surfaces, we assigned
the three peaks to COL bonded to Ru-coordinated-to-Ru (or Ru sites distal to the Pt
islands–Ru-like sites, 1948 cm1), Ru-coordinated-to-Pt (or Ru sites proximal to
the Pt islands–PtRu sites, 2000 cm1), and Pt-islands-on-Ru-core (or Pt adlayer
sites that are most close to Pt sites in PtRu alloys–Pt-like sites, 2031 cm1) sites on
the surface of the Ru@Pt NPs respectively. The middle-frequency sites are most
likely around the peripheries of the Pt islands formed on the surface of Ru core [90].
Because Ru is highly oxophilic, the exposure of the Ru@Pt NPs to air after the
synthesis can easily lead to an enrichment of Ru species to the surface [91]. More-
over, neither repetitive CVs between 0.3 V and 0.7 nor holding electrode potential
at 0.3 V could readily modify such a Ru enrichment on the surface of the Ru@Pt
NPs. Consequently, only one C-O stretching band (Fig. 5a-1) similar to that of
234 D. Chen et al.

Fig. 5.5 The potential dependent EC-SEIRAS spectra of the pre-adsorbed CO during the first
(a-1/b-1) and the second (a-2/b-2) COR on the Ru@Pt/Au@Pt/C NPs respectively in 0.1 M
HClO4. The spectra taken at 0.7 V for the Ru@Pt and 1.0 V for the Au@Pt/C were used as the
corresponding reference spectra. The integrated areas of the spectra in (a) and (b) are plotted in (c)
and (d) respectively. The inset in (c) compares fine spectral features of (a-1) and (a-2) taken at
0.15 V. The spectra in (a-2) can be de-convoluted into three bands: 2031-cm1 (green stars),
2000-cm1 (blue triangles) and 1948-cm1 (pink squares) bands, which should be compared with
those in (a-1) where only one band (2005 cm1, black circles) was observed. In contrast, only one
same IR band that varied between 2025 and 2070 cm1 was observed in (b-1, black triangles in d)
and (b-2, red triangles in d). Adapted from [80] with permission (© 2011 American Chemical
Society)

COL-Ru of Ru-modified Pt(1 1 1) surface [82] was observed, suggesting a dominant


CO-Ru species on the surface.
On the other hand, the first COR acted like a CO annealing process and was able
to reduce (at least partially) the surface Ru oxide (RuOx or RuOxHy) to metallic Ru
and to bring the segregated Pt sites back to the surface of the Ru@Pt NPs so that CO
could adsorb onto different surface sites that led to the appearance of the three
different C-O stretching peaks as observed in the inset of Fig. 5.5c. The potential
dependent integrated IR band intensity of each de-convoluted band is presented in
Fig. 5.5c. What is intriguing is that the integrated intensity of each band behaves
5 Integrated Studies of Au@Pt and Ru@Pt Core-Shell Nanoparticles. . . 235

differently. The 2031-cm1 band (the Pt-islands-on-Ru-core sites, green stars)


shows the most negative onset potential of the COR, i.e., ~0.1 V, at which the
CO IR intensity starts decreasing. The 2000-cm1 band (the Ru sites proximal to
the Pt islands, blue solid triangles) shows little intensity changes from 0.3 to
0.0 V, a noticeable intensity increase from 0.0 to 0.25 V, and then a rapid intensity
decease beyond 0.25 V. The latter can be considered the COR onset potential of
these sites. Lastly, the CO IR intensity of the 1948-cm1 band (the Ru sites distal to
the Pt islands) decreases slowly until ~0.15 V beyond which a shaper decrease
starts. Thus, 0.15 V can also be considered as its COR onset potential.
The above observations led us to conclude that, in terms of the COR, the
Pt-islands-on-Ru-core sites were the most active; the proximal Ru sites were the
least active, and the distal Ru sites were in between. Although these differences in
the reaction rate of the COR may reflect differences in reaction barrier and/or CO
diffusion rate at these sites, their relatively lower onset potentials for the COR
should be contrasted to that of the first COR whose value is about 0.35 V (Fig. 5.5c,
black circles). This indicates that the more segregated surface structures of the
Ru@Pt NPs generated by the CO annealing, i.e., Pt islands on metallic Ru core, is
much more active than the (largely ruthenated) surface structure before the CO
annealing of which surface Ru oxides dominated, an observed was further con-
firmed by our recent study of the chemistry related to the activation of commercial
PtRu alloy electrocatalysts [92].
Figure 5.5b-1 and b-2 depict the potential dependent EC-SEIRAS spectra for the
two sequential CORs on the Au@Pt/C NPs. As briefly mentioned above and in
contrast to the cases of the Ru@Pt NPs presented in Fig. 5.5a-1 and a-2, no obvious
differences were observed in terms of the SEIRAS of the pre-adsorbed CO between
the two sequential CORs, as can been seen in Fig. 5.5b, d: only one CO IR
stretching band varying from 2025 to 2070 cm1 was observed over the entire
potential range, which can be assigned to a linearly bonded COL-Pt. This infers
rather constant surface structure and properties of the Au@Pt/C, an indication of
stability.
What is puzzling but also intriguing is the rather precipitous decrease in the IR
intensity below 0.0 V where adsorbed CO is supposed to be stable. Coincidently,
the potential dependence of the vibrational frequency in this region also behaved
oddly (not shown here but can be consulted in the original paper [80]). Since the
linearly increasing current during CO stripping positive-going potential scan was
largely negative in this potential regime [80], we speculated that an ongoing
protonation-like process converted adsorbed CO into a different species whose
exact identity is still unknown. However, the species-conversion hypothesis is
consistent with the observation of the non-monotonic frequency variation observed
in the same potential regime mentioned above [93] and with the observation of
co-adsorbed hydrogen on Pt electrodes by visible-infrared SFG [94] and by
SEIRAS [95–98]. Moreover, if this large decrease in IR intensity were caused by
a much earlier COR, it would be hard to reconcile with that no enhanced MOR was
observed on the Au@Pt NPs as will be discussed below.
236 D. Chen et al.

Fig. 5.6 The in situ SEIRAS spectra acquired in 0.5 M MeOH + 0.1 M HClO4 during positive
potential scan of MOR on the Au@Pt (a and a0 ) and Ru@Pt (b) NPs, with the spectra taken at
0.25 and 0.3 V as the respective references. The spectra in (a0 ) are the water bending bands for
the Au@Pt/C NPs obtained using the spectrum recorded at 0.1 V as the reference while the inset in
(b) amplifies the very weak CO spectra. (c) and (d) show the corresponding potential dependent
integrated IR intensity of the adsorbed COL (blue) and the surface water bending δ(HOH) (red).
The solid lines are for eye-guiding purpose. Adapted from [80] with permission (© 2011 American
Chemical Society)

The MOR on the Ru@Pt and Au@Pt/C NPs after the two sequential CORs
discussed above was followed by the in situ stair-step potential dependent
EC-SEIRAS in 0.5 M MeOH + 0.1 M HClO4. On the Au@Pt/C NP surface a
weak Pt-COL band around 2040 cm1 was observed as shown in Fig. 5.6a whose
corresponding integrated intensity is shown in Fig. 5.6c (solid blue circles). Notice
that while the COR of the pre-adsorbed gaseous CO (Fig. 5.5b) started with a fully
CO-covered surface, the MOR started with a CO free surface. Therefore, the
increase in IR intensity of the COL during the MOR before 0.45 V is a manifestation
of the accumulation of the surface-bound poisonous CO generated from the disso-
ciative adsorption of MeOH. Notice that the CO intensity started decreasing at
0.45 V, which is consistent with the main-current peak potential of the CO
stripping [80].
5 Integrated Studies of Au@Pt and Ru@Pt Core-Shell Nanoparticles. . . 237

During the potential dependent MOR, a somewhat bipolar water bending vibra-
tion δ(HOH) band was observed at ~1620 cm1 with the negative going signal
appearing at the lower frequency side. If the spectra were re-referenced by the
spectrum obtained at 0.1 V, the original bipolar spectral shape largely disappeared
(Fig. 5.6a0 ). This is because 0.1 V is around the potential of zero charge, PZC [99],
at which IR bending vibration of the absorbed water is expected to be weakest
[32]. The now dominantly mono-polar spectra in Fig. 5.6a0 reveals a more clear
trend: The peak intensity decreased first gradually and the bending frequency
moved from 1608 cm1 at 0.2 V to 1600 cm1 just below 0.1 V; it then started
increasing continuously with a constant bending frequency at 1622 cm1 above
0.1 V (red triangles in Fig. 5.6c). Such a behavior of the δ(HOH) is very similar to
that observed on Pt film [95], which was interpreted as due to changes in the
hydrogen-bonding-associated average orientation of water molecules with
oxygen-up or down on the negative or positive charged Pt surface separated by
the PZC.
Below the PZC, the Au@Pt/C NP surface was negatively charged so it attracted
the hydrogen end of the water molecules. As the electrode potential became less
negative, the attraction to hydrogen became weaker, so less hydrogen-down water
molecules would bond to the surface and led to weaker IR intensity of the water-
bending band. As the electrode potential moved further positively beyond the PZC,
the Au@Pt/C NP surface became more and more positively charged, leading to
attracting more and more water molecules via the oxygen-end of the water mole-
cules. Consequently, the IR intensity of the water-bending band became stronger.
Although the electro-oxidation of MeOH and adsorbed CO must have consumed
surface water, the continuous increase in water adsorption above 0.1 V indicates
that the MOR on the Au@Pt/C NPs was overall slow. The oxidation of the surface
above 0.75 V led to the observed level-off and then decrease in the IR intensity.
The band of the adsorbed CO generated during the MOR on the Au@Pt/C NPs
was much weaker than the pre-adsorbed gaseous CO for the COR. This is consistent
with the EC results that showed the degree of the suppression of hydrogen adsorp-
tion was much lower for the former than that for the latter. Yet, an even weaker
linearly bound CO band at about 1977 cm1 (at 0.05 V) was observed for the
MOR on the Ru@Pt NPs, whose integrated intensity, although about three times
smaller than that of the CO generated during the MOR on Au@Pt NPs, followed the
same increase-then-decrease pattern with the peak at 0.05 V (Fig. 5.6d). It is
highly likely that this CO band corresponds to the poisonous CO generated during
the MOR on the Pt-islands-on-Ru-core sites, for the Ru sites are essentially inactive
for MOR. Moreover, this onset is almost identical with that of COL on Pt islands of
Ru@Pt for second COR (green stars in Fig. 5.5c). The very low CO coverage was
probably responsible for the significant red shift in stretching frequency from the
full-coverage value of ~2030 cm1 to the low-coverage value of 1977 cm1.
Notice that the MOR generated CO-band peak potential of 0.05 V for the
Ru@Pt NPs was 0.5 V more negative than that (0.45 V) for the Au@Pt/C NPs. This
observation implies a much less CO poisoning and faster MOR on the Ru@Pt than
on the Au@Pt/C NPs. Indeed, the amplitude of the corresponding surface water
238 D. Chen et al.

bending vibration δ(HOH) band at 1641 cm1 (Fig. 5.6b) did not show an increase
after 0.1 V but instead reached a constant level (Fig. 5.6d), signaling that the MOR
on the Ru@Pt NPs was much faster than on the Au@Pt/C NPs such that it led to a
higher rate of surface water consumption, therefore a steady state of surface water
adsorption.
As activated surface water is a necessary surface reactant for both COR and
MOR [100, 101], its in situ IR spectroscopic investigation has proven informative
and revealing [7, 41, 102]. The IR vibrational frequency, band shape, and intensity
of the surface water at the EC interface are very sensitive to several parameters that
include the type of metal surface, the strength and direction of interfacial electric
field, and the co-adsorbed species. In addition to the bending mode δ(HOH)
discussed above, three distinguishable v(O-H) stretching bands of interfacial
water have been identified and reported: [32, 95, 103, 104] the non-hydrogen
bonded water monomer at ~3660 cm1, the strongly hydrogen-bonded ice-like
water at ~3040 cm1, and the disordered weakly hydrogen-bonded water at
~3400 cm1.
For the MOR, two water bands were observed (Fig. 5.7a): the weakly hydrogen-
bonded at 3380 cm1 and strongly hydrogen-bonded at 3075 cm1. Their potential
dependent, integrated IR intensity obtained by peak de-convolution are presented in
Fig. 5.7b. The initial intensity decrease in both bands can be again rationalized by
the reduced attraction to the hydrogen of the water molecule as the surface became

Fig. 5.7 (a) The potential dependent EC-SEIRAS spectra of the ν(O-H) bands observed on the
Ru@Pt NPs during MOR and (b) the corresponding integrated IR intensity of the 3380-cm1 (pink
stars) and 3075-cm1 (blue diamonds) bands as indicated by the vertical red lines in (a). Also
presented in (b) is the current transient recorded during the stair-step IR measurements (black
curve with spikes). Adapted from [80] with permission (© 2011 American Chemical Society)
5 Integrated Studies of Au@Pt and Ru@Pt Core-Shell Nanoparticles. . . 239

less negatively charged. The second large decrease in the integrated intensity for the
weakly hydrogen-bonded water at 3380 cm1 that peaked at ~0.45 V as compared
to the flat amplitude of the strongly hydrogen-bonded water band at 3075 cm1 over
the same potential region strongly suggests that the former type of water was at least
more actively involved in the MOR and therefore was consumed more. This
activity correlation is further supported by the simultaneously recorded current
transients over the potential-step changes (black curve in Fig. 5.7b). The clear
overlap of the MOR current peak with the IR intensity dip of the weakly hydrogen-
bonded water is the strongest experimental evidence showing that the weakly
hydrogen-bonded water was the active water species for the MOR on the Ru@Pt
NPs.
The in situ EC-SEIRAS spectra of the v(O–H) band of the interfacial water on
the Au@Pt/C NPs during the COR in 0.1 M HClO4 and MOR in 0.1 M
HClO4 + 0.5 M MeOH are presented in Fig. 5.8a, b. Very similar behaviors were
observed for both cases, consisting of a broad band at 3245 cm1 with a shoulder at
~3400 cm1 over the entire potential range studied. Additionally, a small yet visible
and also narrower peak at ~3580 cm1 was observed for potentials more positive
than ~0.3 V for the COR and than ~0.1 V for the MOR, respectively. The
potential dependent integrated band intensities obtained via de-convolution were
plotted in Fig. 5.8c, d (the blue circles for the COR and red triangles for the MOR)
for the 3435-cm1 (weakly hydrogen-bonded) and 3245-cm1 (strongly hydrogen-
bonded) water bands, respectively.

Fig. 5.8 Potential dependent EC-SEIRAS spectra of ν(O-H) stretch vibration for the Au@Pt/C
NPs during the COR (a) and MOR (b). The integrated IR intensity of the 3245-cm1 (c) and 3580-
cm1 (d) bands as a function of potential for the COR (blue circle) and MOR (red triangle)
obtained by spectral de-convolution. Adapted from [80] with permission (© 2011 American
Chemical Society)
240 D. Chen et al.

Among the three different types of water observed, only the weakly hydrogen-
bonded water in both the COR and MOR (Fig. 5.8d) show a variation pattern in the
potential-dependent, integrated IR band intensity that has three different potential
regimes: 0.25 to 0.1 V, 0.1–0.45 V, and 0.45–1.0 V, which closely match those of
IR band of the adsorbed CO (Fig. 5.5d). Interestingly, two distinct increasing rates
of the IR intensity were observed in the latter two potential regimes: 0.28, 0.12 a.u./
V below and 1.01, 0.79 a.u./V above ~0.45 V (see the black straight lines in
Fig. 5.8d) for MOR and COR, respectively, of which the electrode potential of
0.45 V coincides with the peak potential of the COR and the poisonous CO
generated during the MOR (Fig. 5.6c). Thus, the much bigger increasing rate of
the IR intensity above 0.45 V is most likely due to the oxidation of adsorbed COL in
COR and MOR that rendered more surface sites available for water adsorption.
Such a pattern match strongly suggests that the weakly hydrogen-bonded water
species most likely occupied the surface sites freed by the CO oxidation and might
also be the source of oxygen-containing species for the COR and MOR as suggested
by the observations made on the Ru@Pt NPs (Fig. 5.7b), although more direct
evidence supporting this assignment is still needed. Moreover, since no v(OH) band
corresponding to adsorbed Pt-OHads (~3700 cm1) [103] was observed during both
the COR and MOR and on both samples, the dissociative adsorption of water on Pt
(Pt-OH2 ! Pt-OHads + H+ + e) was probably the rate-determining step for both the
COR and MOR [7].

5.5 In Situ SERS of Pt and Ru@Pt

Figure 5.9 shows the Schematic of our in situ EC-SERS cell. The cell body is made
of Teflon with a quartz window through which the laser is shined onto the surface of
working electrode, which can be a rough Au film electrode, a roughened Pt disk
electrode, or NPs deposited on the substrate electrode. For the current study,
commercial Pt black or the synthesized Ru@Pt NPs were drop-casted onto a normal
Pt disk electrode from which Raman signal is too weak to be observed. The surface
enhancement came from the NPs themselves but it was time-consuming to locate a
“hot” spot. For the measurements, 0.1 M HClO4 or 0.1 M HClO4 + 0.5 M MeOH
were used for COR and MOR respectively. The EC-SERS spectra were obtained
using a confocal Raman microscope system (Renishaw RM1000) equipped with a
deep depletion CCD peltier cooled down to 70  C. The microscope attachment is
based on an Olympus BH2-UMA system and uses a 50 objective. A holographic
notch filter was used to filter the excitation line, and 1200 g mm1 selective
holographic grating was employed depending on the spectral resolution required.
The excitation wavelength was 785 nm from a Renishaw diode laser with a
maximum power of 100 mW. For each spectrum, exposure time was 120 s; each
potential step was held for 253 s during the in situ EC SERS measurements.
Figure 5.10 compares the in situ EC SERS spectra of the adsorbed CO at 0.25 and
0.0 V acquired during COR and MOR on Pt black and Ru@Pt respectively. It shows
5 Integrated Studies of Au@Pt and Ru@Pt Core-Shell Nanoparticles. . . 241

Fig. 5.9 The schematic of the in situ EC SERS cell with a photo of the working electrode encased
in the cell body as seen from the top. Adapted from [93] with permission (© 2014 Royal Society of
Chemistry)

Fig. 5.10 In-situ EC SERS


spectra of the bonding of
Pt–COads during the COR
(a–d) and MOR (e, f) at
0.25 and 0.0 V on the Pt
black and the Ru@Pt NPs

that the gaseous adsorption generated Pt–CO bands on Pt and Ru@Pt at 0.0 V
appeared at 485 cm1 and 490.5 cm1, respectively, with a difference of ~+5 cm1.
On the other hand, the v(C–O) on Pt and Ru@Pt at the same potential presents
bands at 2052 cm1 and 2006 cm1, respectively, with a difference of ~46 cm1.
In other words, in contrast to the large red shift (~46 cm1) for C-O stretching on
pure Pt vs on Ru@Pt, the corresponding difference in the Pt–CO vibrational frequency
shows a small but still significant blue shift (~+5 cm1). Such opposing shifts in νPt–CO
vs νC–O indicates that the addition of Ru makes the Pt sites more electron-accepting
or having higher Lewis acidity [105]. For MeOH generated CO on the Ru@Pt,
the band intensity was much weaker, indicating improved CO tolerance. But it also
showed the largest negative Stark tuning slope: 32 cm1/V (the other two were
242 D. Chen et al.

much smaller though still negative). That the SERS can access νPt–CO vibration, which
is directly related to metal–CO bonding and inaccessible to IR, offers insight that is
complementary to the intra-molecular vibration that can be measured by IR.

5.6 Discussion

In the above sections, we have presented in situ EC 195Pt and 13CO NMR, SEIRAS
and SERS data obtained on the same Ru@Pt and Au@Pt NPs, shedding light on
different aspects of their respective electrocatalytic behaviors. For instance, NMR
probes the electronic properties of the metal NPs, IR exams the intra-molecular
vibrations and by which identifies reaction intermediates, and Raman accesses to
the metal–adsorbate vibrations. Altogether, they can provide a fuller picture and
deeper understanding of the system under investigation.
NMR’s ability to provide quantitative information on electronic partition at the
Fermi level (Table 5.1) can be of good use to connect with some recent theoretical
development. For instance, recent quantum calculations on oxide-supported Pt
model systems [106] have (re-) unearthed a potentially important and also more
chemically intuitive surface bonding descriptor–the electronic partition, i.e., 6sp vs
5d electrons, at Pt that has been less (if not at all) investigated in electrocatalysis.
The calculations predict that the Pt–H bond would be stronger at a Pt site with
higher 6sp partition, while the Pt–O bond would be stronger at a Pt site with higher
5d partition, with Pt–CH3 and Pt–CH2 in between. The relatively recent application
of the Crystal Orbital Hamiltonian Population (COHP) formalism [107] within the
extended H€ uckel molecular orbital (MO) theory to chemisorption of CO, hydrogen,
methyl, and ethyl to metal surfaces by Hoffman and co-workers [108] also highlight
the revealing power of such a chemistry-based electronic-orbital-specific (EOS)
formalism. For instance, the COHP analysis revealed that the metal sp orbitals
actually contribute significantly (much more than previously believed) to the
overall CO-Ni and CO-Pt chemisorption [109–111]. Also, based on the COHP
analysis, strong agostic interactions between the C–H bonds of methyl and ethyl
groups and the Pt sites of high symmetries were proposed for C–H bond activation
[108]. This may find useful application in further delineating the mechanism of the
formic acid oxidation reaction (FAOR) and MOR on metal (Pt) surfaces, where C–
H bond-breaking is a necessary reaction step. This line of reasoning may also be
useful in designing suitable catalysts for partial oxidation of methane.
Indeed, the aforementioned EOS description has long been embodied in the
organometallic analogy of surface bonding advocated by Somorjai [112–114] and
in the frontier-orbital formalism by Hoffmann [115, 116], which has been very
successful in rationalizing many reaction mechanisms at solid/gas interfaces of
heterogeneous catalysis. It is somewhat surprising that it has not found wide-spread
use in electrocatalysis lately, which might have to do with the fact that the powerful
valence-electron-orbital-probing ultraviolet photoelectron spectroscopy is not
applicable to a solid/liquid interface, so a direct experimental connection with the
5 Integrated Studies of Au@Pt and Ru@Pt Core-Shell Nanoparticles. . . 243

chemically intuitive EOS description is thus lost. Nonetheless, we strongly believe


that an EOS description as briefly mentioned above can complement the highly
successful d-band center theory by offering more chemical specificity in terms of
surface-bonding-involved electronic orbitals by which the chemistry taking place at
an EC interface may be better nuanced. Moreover, FAOR, MOR, and oxygen
reduction reaction (ORR) at an electrode surface, all involve elementary reaction
steps that will necessarily have Pt–H, Pt–O, Pt-CH3, and/or Pt–CH2 bond formation
and C–H, O–H, and/or O¼O bond breakings, either sequentially or simultaneously.
They can thus serve as a natural and fertile testing ground for the EOS description.
Therefore, the EOS description can help parse more insightfully the chemistry of
M@Pt core-shell NPs for fuel cell electrocatalysis as a function of core element M
at an electronic and molecular level.
Figure 5.11 illustrates some EOS information obtained by the 195Pt and 13CO
NMR: the pure Pt has the highest s-like electrons at the Fermi level, Ru@Pt lowest,
with Au@Pt in between. Now if higher availability of s-like electrons leads to
stronger Pt-H bonding, the ability to abstract hydrogen would be strongest on Pt,
weakest on Ru@Pt, with Au@Pt in between. For MOR, stronger hydrogen
abstracting ability would lead to more generation of CO [117], which would
rationalize why only very low amount of CO was observed on the Ru@Pt
(Fig. 5.6b). At the same time, the availability of d-like electrons is not reduced at
all as compared to pure Pt so it would retain its ability to bind oxygen-containing
species, which would facilitate the formation of methoxy therefore enhance the
direct reaction pathway [117]. In other words, the higher MOR activity on Ru@Pt

Fig. 5.11 Correlation between the D2π*(Ef) and Ds(Ef) deduced by using the two-band model from
the 13C NMR Data of COads on the Pt, Ru@Pt and Au@Pt NPs as well as 195Pt NMR of them.
Adapted from [71] with permission (© 2012 American Chemical Society)
244 D. Chen et al.

might arise from its ability to enhance the direct reaction pathway for MOR. By the
same token, the substantially reduced d-like electrons at the Fermi level for the
Au@Pt would account for its inferior ability to do MOR because of its worsened
ability to bond oxygen-containing species.
As to the blue shift of the Pt–CO vibration from the pure Pt to Ru@Pt observed
by in situ EC-SERS, one could rationalize it by the reduction in s-like electrons at
the Fermi level as this correlates to a lower D2π*(Ef) (Fig. 5.11), an indication of
stronger Pt–CO bonding [105]. Consequently, νPt–CO became higher on the
Ru@Pt NPs.

5.7 Summary and Future Outlooks

We have shown in this chapter that integrating in situ EC NMR, SEIRAS and SERS
studies of the Ru@Pt and Au@Pt NPs has enabled us to achieve a better mecha-
nistic understanding of the two systems. In situ 195Pt and 13C NMR were able to
provide some quantitative EOS information based on which some SEIRAS and
SERS observations can be rationalized. The more relevant information one can
garner, the better one can understand the catalyst’s performance through which
better performing electrocatalysts can be designed and developed. For instance, for
Ru@Pt NPs, the SEIRAS data have suggested that segregated Pt and Ru ensembles
seem to help enhance the MOR reactivity. Also, lowering the s-like but increasing
the d-like electrons’ availability may guide MOR through the direct reaction
pathway without generating poisonous CO. Such specific information can be fed
to the next round of designing and developing better MOR electrocatalysts.
In terms of developing further in situ techniques, there are several promising
ones on the horizon that will become mature in the next decade. The first is the
synchrotron light source and free-electron laser [118] based in situ X-ray spectro-
scopic methods [119]. As the synchrotron light source becomes more intense, the
sensitivity of the technique will improve as well. It is also expected that the spatial
resolution of X-ray based microscope and imaging will also improve substantially
over the next decade. The second is the simple and straightforward high-resolution
in situ EC NMR [120] that can explore all the analytical power of routine NMR.
The third is nitrogen vacancy based scanning NMR microscope [121–123]. The key
features of these new developments are chemical specificity, sensitivity, and spatial
resolution.

Acknowledgments The authors gratefully acknowledge the financial supports provided by DOE
(DE-FG02-07ER15895), NSF (CHE-1413429) and ARO (66191-CH).
5 Integrated Studies of Au@Pt and Ru@Pt Core-Shell Nanoparticles. . . 245

References

1. Markovic NM, Ross PN (2000) New electrocatalysts for fuel cells from model surfaces to
commercial catalysts. CATTECH 4:110–126
2. Steele BCH, Heinzel A (2001) Materials for fuel-cell technologies. Nature 414
(6861):345–352
3. Girishkumar G, McCloskey B, Luntz AC, Swanson S, Wilcke W (2010) Lithium–air battery:
promise and challenges. J Phys Chem Lett 1(14):2193–2203. doi:10.1021/jz1005384
4. Shao Y, Park S, Xiao J, Zhang J-G, Wang Y, Liu J (2012) Electrocatalysts for nonaqueous
lithium–air batteries: status, challenges, and perspective. ACS Catal 2(5):844–857. doi:10.
1021/cs300036v
5. Zhang S, Yuan X-Z, Hin JNC, Wang H, Friedrich KA, Schulze M (2009) A review of
platinum-based catalyst layer degradation in proton exchange membrane fuel cells. J Power
Sources 194(2):588–600. http://dx.doi.org/10.1016/j.jpowsour.2009.06.073
6. Sealy C (2008) The problem with platinum. Mater Today 11(12):65–68
7. Arenz M, Mayrhofer KJJ, Stamenkovic V, Blizanac BB, Tomoyuki T, Ross PN, Markovic
NM (2005) The effect of the particle size on the kinetics of CO electrooxidation on high
surface area Pt catalysts. J Am Chem Soc 127(18):6819–6829. doi:10.1021/ja043602h
8. Lebedeva NP, Koper MTM, Feliu JM, van Santen RA (2002) Mechanism and kinetics of the
electrochemical CO adlayer oxidation on Pt(111). J Electroanal Chem 524–525:242–251.
http://dx.doi.org/10.1016/S0022-0728(02)00669-1
9. Iwasita T (2002) Electrocatalysis of methanol oxidation. Electrochim Acta 47:3663–3674
10. Wang H, Jusys Z, Behm RJ (2006) Ethanol electro-oxidation on carbon-supported Pt, PtRu
and Pt3Sn catalysts: a quantitative DEMS study. J Power Sources 154(2):351–359. http://dx.
doi.org/10.1016/j.jpowsour.2005.10.034
11. Xia XH, Liess HD, Iwasita T (1997) Early stages in the oxidation of ethanol at low index
single crystal platinum electrodes. J Electroanal Chem 437(1–2):233–240. http://dx.doi.org/
10.1016/S0022-0728(97)00404-X
12. Kang Y, Qi L, Li M, Diaz RE, Su D, Adzic RR, Stach E, Li J, Murray CB (2012) Highly
active Pt3Pb and core–shell Pt3Pb–Pt electrocatalysts for formic acid oxidation. ACS Nano 6
(3):2818–2825. doi:10.1021/nn3003373
13. Chen DJ, Zhou ZY, Wang Q, Xiang DM, Tian N, Sun SG (2010) A non-intermetallic PtPb/C
catalyst of hollow structure with high activity and stability for electrooxidation of formic
acid. Chem Commun 46(24):4252–4254. doi:10.1039/c002964e
14. Schmidt TJ, Ross Jr PN, Markovic NM (2002) Temperature dependent surface electrochem-
istry on Pt single crystals in alkaline electrolytes: Part 2. The hydrogen evolution/oxidation
reaction. J Electroanal Chem 524–525:252–260. http://dx.doi.org/10.1016/S0022-0728(02)
00683-6
15. Esposito DV, Hunt ST, Stottlemyer AL, Dobson KD, McCandless BE, Birkmire RW, Chen
JG (2010) Low-cost hydrogen-evolution catalysts based on monolayer platinum on tungsten
monocarbide substrates. Angew Chem Int Ed 49(51):9859–9862. doi:10.1002/anie.
201004718
16. Lim B, Jiang M, Camargo PHC, Cho EC, Tao J, Lu X, Zhu Y, Xia Y (2009) Pd-Pt bimetallic
nanodendrites with high activity for oxygen reduction. Science 324(5932):1302–1305.
doi:10.1126/science.1170377
17. Zhang J, Sasaki K, Sutter E, Adzic RR (2007) Stabilization of platinum oxygen-reduction
electrocatalysts using gold clusters. Science 315(5809):220–222
18. Lu YC, Xu ZC, Gasteiger HA, Chen S, Hamad-Schifferli K, Shao-Horn Y (2010) Platinum-
gold nanoparticles: a highly active bifunctional electrocatalyst for rechargeable lithium-air
batteries. J Am Chem Soc 132(35):12170–12171. doi:10.1021/ja1036572
19. Kerbach I, Climent V, Feliu JM (2011) Reduction of CO2 on bismuth modified Pt(110)
single-crystal surfaces. Effect of bismuth and poisoning intermediates on the rate of hydrogen
246 D. Chen et al.

evolution. Electrochim Acta 56(12):4451–4456. http://dx.doi.org/10.1016/j.electacta.2011.


02.027
20. Qu J, Zhang X, Wang Y, Xie C (2005) Electrochemical reduction of CO2 on RuO2/TiO2
nanotubes composite modified Pt electrode. Electrochim Acta 50(16–17):3576–3580. http://
dx.doi.org/10.1016/j.electacta.2004.11.061
21. Climent V, Garcia-Araez N, Feliu JM (2009) Clues for the molecular-level understanding of
electrocatalysis on single-crystal platinum surfaces modified by p-block adatoms. Fuel cell
catalysis a surface science approach. Wiley, Hoboken
22. Liu P, Nørskov JK (2001) Ligand and ensemble effects in adsorption on alloy surfaces. Phys
Chem Chem Phys 3(17):3814–3818. doi:10.1039/b103525h
23. Tong YY, Wieckowski A, Oldfield E (2002) NMR of electrocatalysts. J Phys Chem B 106
(10):2434–2446. doi:10.1021/jp0129939
24. Du B, Danberry AL, Park I-S, Sung Y-E, Tong Y (2008) Spatially resolved 195Pt NMR of
carbon-supported PtRu electrocatalysts: local electronic properties, elemental composition,
and catalytic activity. J Chem Phys 128(5):052311. doi:10.1063/1.2830952
25. Tan F, Du B, Danberry AL, Park I-S, Sung Y-E, Tong Y (2008) A comparative in situ 195Pt
electrochemical-NMR investigation of PtRu nanoparticles supported on diverse carbon
nanomaterials. Faraday Discuss 140:139–153. doi:10.1039/b803073a
26. Tong RC, Wieckowski A, Oldfield E (2000) A detailed NMR-based model for CO on Pt
catalysts in an electrochemical environment: shifts, relaxation, back-bonding, and the fermi-
level local density of states. J Am Chem Soc 122(6):1123–1129. doi:10.1021/ja9922274
27. Tong KHS, Babu PK, Waszczuk P, Wieckowski A, Oldfield E (2002) An NMR investigation
of CO tolerance in a Pt/Ru fuel cell catalyst. J Am Chem Soc 124(3):468–473. doi:10.1021/
ja011729q
28. Kobayashi T, Babu PK, Gancs L, Chung JH, Oldfield E, Wieckowski A (2005) An NMR
determination of CO diffusion on platinum electrocatalysts. J Am Chem Soc 127
(41):14164–14165. doi:10.1021/ja0550475
29. Tian ZQ, Ren B (2004) Adsorption and reaction at electrochemical interfaces as probed by
surface-enhanced Raman spectroscopy. Annu Rev Phys Chem 55:197–229. doi:10.1146/
annurev.physchem.54.011002.103833
30. Osawa M (2006) Diffraction and spectroscopic methods in electrochemistry: in-situ surface-
enhanced infrared spectroscopy of the electrode/solution interface, vol 9, Advances in
electrochemical science and engineering. Wiley-VCH, New York
31. Osawa M, Ataka K, Yoshii K, Yotsuyanagi T (1993) Surface-enhanced infrared ATR
spectroscopy for in situ studies of electrode/electrolyte interfaces. J Electron Spectrosc
Relat Phenom 64:371–379. uuid: CE0FC3F0-7DA6-4A00-8375-39A57E55959A
32. Ataka K, Yotsuyanagi T, Osawa M (1996) Potential-dependent reorientation of water mol-
ecules at an electrode/electrolyte interface studied by surface-enhanced infrared absorption
spectroscopy. J Phys Chem 100(25):10664–10672. doi:10.1021/jp953636z
33. Garcia-Araez N, Rodriguez P, Bakker HJ, Koper MTM (2012) Effect of the surface structure
of gold electrodes on the coadsorption of water and anions. J Phys Chem C 116
(7):4786–4792. doi:10.1021/jp211782v
34. Sun SG, Cai WB, Wan LJ, Osawa M (1999) Infrared absorption enhancement for CO
adsorbed on Au films in perchloric acid solutions and effects of surface structure studied
by cyclic voltammetry, scanning tunneling microscopy, and surface-enhanced IR spectros-
copy. J Phys Chem B 103(13):2460–2466
35. Yoshida M, Yamakata A, Takanabe K, Kubota J, Osawa M, Domen K (2009) ATR-SEIRAS
investigation of the fermi level of Pt cocatalyst on a GaN photocatalyst for hydrogen
evolution under irradiation. J Am Chem Soc 131(37):13218–13219. doi:10.1021/ja904991p
36. Shao MH, Adzic RR (2005) Electrooxidation of ethanol on a Pt electrode in acid solutions: in
situ ATR-SEIRAS study. Electrochim Acta 50(12):2415–2422. doi:10.1016/j.electacta.2004.
10.063
5 Integrated Studies of Au@Pt and Ru@Pt Core-Shell Nanoparticles. . . 247

37. Smolinka T, Heinen M, Chen YX, Jusys Z, Lehnert W, Behm RJ (2005) CO2 reduction on Pt
electrocatalysts and its impact on H-2 oxidation in CO2 containing fuel cell feed gas—a
combined in situ infrared spectroscopy, mass spectrometry and fuel cell performance study.
Electrochim Acta 50(25–26):5189–5199. doi:10.1016/j.electacta.2005.02.082
38. Kunimatsu K, Uchida H, Osawa M, Watanabe M (2006) In situ infrared spectroscopic and
electrochemical study of hydrogen electro-oxidation on Pt electrode in sulfuric acid. J
Electroanal Chem 587(2):299–307. doi:10.1016/j.jelechem.2005.11.026
39. Osawa M, K-i K, Samjeské G, Uchida T, Ikeshoji T, Cuesta A, Gutiérrez C (2011) The role of
bridge-bonded adsorbed formate in the electrocatalytic oxidation of formic acid on platinum.
Angew Chem Int Ed 50(5):1159–1163. doi:10.1002/anie.201004782
40. Chen YX, Miki A, Ye S, Sakai H, Osawa M (2003) Formate, an active intermediate for direct
oxidation of methanol on Pt electrode. J Am Chem Soc 125(13):3680–3681. doi:10.1021/
ja029044t
41. Samjeske G, Komatsu K, Osawa M (2009) Dynamics of CO oxidation on a polycrystalline
platinum electrode: a time-resolved infrared study. J Phys Chem C 113(23):10222–10228.
doi:10.1021/jp900582c
42. Kunimatsu K, Hanawa H, Uchida H, Watanabe M (2009) Role of adsorbed species in
methanol oxidation on Pt studied by ATR-FTIRAS combined with linear potential sweep
voltammetry. J Electroanal Chem 632(1–2):109–119
43. Yajima T, Uchida H, Watanabe M (2004) In-situ ATR-FTIR spectroscopic study of electro-
oxidation of methanol and adsorbed CO at PtRu alloy. J Phys Chem B 108(8):2654–2659.
doi:10.1021/jp037215q
44. Wang C, Peng B, Xie H-N, Zhang H-X, Shi F-F, Cai W-B (2009) Facile fabrication of Pt, Pd
and Pt–Pd alloy films on Si with tunable infrared internal reflection absorption and synergetic
electrocatalysis. J Phys Chem C 113(31):13841–13846. doi:10.1021/jp9034562
45. Miyake H, Okada T, Samjeske G, Osawa M (2008) Formic acid electrooxidation on Pd in
acidic solutions studied by surface-enhanced infrared absorption spectroscopy. Phys Chem
Chem Phys 10(25):3662–3669. doi:10.1039/b805955a
46. Wang HF, Yan YG, Hu SJ, Cai WB, Xu QH, Osawa M (2007) Seeded growth fabrication of
Cu-on-Si electrodes for in situ ATR-SEIRAS applications. Electrochim Acta 52
(19):5950–5957. doi:10.1016/j.electacta.2007.03.042
47. Delgado JM, Orts JM, Rodes A (2007) A comparison between chemical and sputtering
methods for preparing thin-film silver electrodes for in situ ATR-SEIRAS studies.
Electrochim Acta 52(14):4605–4613. doi:10.1016/j.electacta.2006.12.045
48. Yajima T, Wakabayashi N, Uchida H, Watanabe M (2003) Adsorbed water for the electro-
oxidation of methanol at Pt-Ru alloy. Chem Commun 7:828–829. doi:10.1039/b212197b
49. Yan LQ-X, Huo S-J, Ma M, Cai W-B, Osawa M (2005) Ubiquitous strategy for probing ATR
surface-enhanced infrared absorption at platinum group metal-electrolyte interfaces. J Phys
Chem B 109(16):7900–7906. doi:10.1021/jp044085s
50. Wang J-Y, Zhang H-X, Jiang K, Cai W-B (2011) From HCOOH to CO at Pd electrodes: a
surface-enhanced infrared spectroscopy study. J Am Chem Soc 133:14876–14879. doi:10.
1021/ja205747j
51. Vassilev P, Koper MTM (2007) Electrochemical reduction of oxygen on gold surfaces: a
density functional theory study of intermediates and reaction paths. J Phys Chem C 111
(6):2607–2613. doi:10.1021/jp064515+
52. Kunimatsu K, Yoda T, Tryk DA, Uchida H, Watanabe M (2010) In situ ATR-FTIR study of
oxygen reduction at the Pt/Nafion interface. Phys Chem Chem Phys 12(3):621–629. doi:10.
1039/B917306D
53. Cuesta A, Cabello G, Hartl FW, Escudero-Escribano M, Vaz-Domı́nguez C, Kibler LA,
Osawa M, Gutiérrez C (2013) Electrooxidation of formic acid on gold: an ATR-SEIRAS
study of the role of adsorbed formate. Catal Today 202:79–86. http://dx.doi.org/10.1016/j.
cattod.2012.04.022
248 D. Chen et al.

54. Shiroishi H, Ayato Y, Kunimatsu K, Okada T (2005) Study of adsorbed water on Pt during
methanol oxidation by ATR-SEIRAS (surface-enhanced infrared absorption spectroscopy). J
Electroanal Chem 581(1):132–138. doi:10.1016/j.jelechem.2005.04.027
55. Futamata M, Luo LQ (2007) Adsorbed water and CO on Pt electrode modified with Ru. J
Power Sources 164(2):532–537. doi:10.1016/j.jpowsour.2006.10.079
56. Li X, Gewirth AA (2005) Oxygen electroreduction through a superoxide intermediate on
Bi-modified Au surfaces. J Am Chem Soc 127(14):5252–5260. doi:10.1021/ja043170a
57. Yang H, Yang Y, Zou S (2007) In situ surface-enhanced raman spectroscopic studies of CO
adsorption and methanol oxidation on Ru-modified Pt surfaces. J Phys Chem C 111
(51):19058–19065. doi:10.1021/jp075929l
58. Xu B, Park I-S, Li Y, Chen D-J, Tong YJ (2011) An in situ SERS investigation of the
chemical states of sulfur species adsorbed onto Pt from different sulfur sources. J Electroanal
Chem 662:52–56. doi:10.1016/j.jelechem.2011.02.031
59. Tian ZQ, Ren B, Wu DY (2002) Surface-enhanced Raman scattering: from noble to transition
metals and from rough surfaces to ordered nanostructures. J Phys Chem B 106
(37):9463–9483. doi:10.1021/jp0257449
60. Gómez R, Pérez JM, Solla-Gullón J, Montiel V, Aldaz A (2004) In situ surface enhanced
raman spectroscopy on electrodes with platinum and palladium nanoparticle ensembles. J
Phys Chem B 108(28):9943–9949. doi:10.1021/jp038030m
61. Gómez R, Solla-Gullón J, Pérez JM, Aldaz A (2005) Surface-enhanced raman spectroscopy
study of ethylene adsorbed on a Pt electrode decorated with Pt nanoparticles.
ChemPhysChem 6(10):2017–2021. doi:10.1002/cphc.200500168
62. Solla-Gullón J, Gómez R, Aldaz A, Pérez JM (2008) A combination of SERS and electro-
chemistry in Pt nanoparticle electrocatalysis: promotion of formic acid oxidation by
ethylidyne. Eletrochem Commun 10(2):319–322. http://dx.doi.org/10.1016/j.elecom.2007.
12.010
63. Pu Zhang JC, Chen Y-X, Tang Z-Q, Dong C, Yang JL, Wu D-Y, Ren B, Tian Z-Q (2010)
Potential-dependent chemisorption of carbon monoxide at a gold core-platinum shell nano-
particle electrode: a combined study by electrochemical in situ surface-enhanced raman
spectroscopy and density functional theory. J Phys Chem C 114:403–411
64. Park I-S, Chen D-J, Atienza DO, Tong YYJ (2013) Enhanced CO monolayer electro-
oxidation reaction on sulfide-adsorbed Pt nanoparticles: a combined electrochemical and in
situ ATR-SEIRAS spectroscopic study. Catal Today 202:175–182. http://dx.doi.org/10.1016/
j.cattod.2012.05.045
65. Park I-S, Atienza DO, Hofstead-Duffy AM, Chen D, Tong YJ (2011) Mechanistic insights on
sulfide-adsorption enhanced activity of methanol electro-oxidation on Pt nanoparticles. ACS
Catal 2(1):168–174. doi:10.1021/cs200546f
66. Li X, Gewirth AA (2003) Peroxide electroreduction on bi-modified au surfaces: vibrational
spectroscopy and density functional calculations. J Am Chem Soc 125(23):7086–7099.
doi:10.1021/ja034125q
67. Li X, Heryadi D, Gewirth AA (2005) Electroreduction activity of hydrogen peroxide on Pt
and Au electrodes. Langmuir 21(20):9251–9259. doi:10.1021/la0508745
68. Du B, Rabb SA, Zangmeister C, Tong Y (2009) A volcano curve: optimizing methanol
electro-oxidation on Pt-decorated Ru nanoparticles. Phys Chem Chem Phys 11
(37):8231–8239. uuid:F4A2892D-47E4-478B-825C-8EB0584C7F0F
69. Le Rhun V, Garnier E, Pronier S, Alonso-Vante N (2000) Electrocatalysis on nanoscale
ruthenium-based material manufactured by carbonyl decomposition. Electrochem Commun
2(7):475–479
70. Park IS, Lee KS, Jung DS, Park HY, Sung Y-E (2007) Electrocatalytic activity of carbon-
supported Pt-Au nanoparticles for methanol electro-oxidation. Electrochim Acta
52:5599–5605. uuid:84E52FA9-41B4-4917-93F1-E2222D793214
5 Integrated Studies of Au@Pt and Ru@Pt Core-Shell Nanoparticles. . . 249

71. Atienza DO, Allison TC, Tong YJ (2012) Spatially resolved electronic alterations as seen by
in situ 195Pt and 13CO NMR in Ru@Pt and Au@Pt core–shell nanoparticles. J Phys Chem C
116(50):26480–26486. doi:10.1021/jp310313k
72. Du B, Zaluzhna O, Tong YJ (2011) Electrocatalytic properties of Au@Pt nanoparticles:
effects of Pt shell packing density and Au core size. Phys Chem Chem Phys 13
(24):11568–11574
73. Tong RC, Godbout N, Wieckowski A, Oldfield E (1999) Correlation between the knight shift
of chemisorbed CO and the fermi level local density of states at clean platinum catalyst
surfaces. J Am Chem Soc 121(13):2996–3003. doi:10.1021/ja9830492
74. Stokes HT, Rhodes HE, Wang PK, Slichter CP, Sinfelt JH (1982) NMR of platinum catalysts.
III. Microscopic variation of the Knight shifts. Phys Rev B 26(7):3575–3581.
uuid:572642F8-60BF-49D7-83A6-B7AE0A58B890
75. Bucher J, van der Klink J (1988) Electronic properties of small supported Pt particles: NMR
study of 195Pt hyperfine parameters. Phys Rev B Condens Matter 38(16):11038–11047.
doi:10.1103/PhysRevB.38.11038
76. Adekunle AS, Ozoemena KI (2008) Electron transfer behaviour of single-walled carbon
nanotubes electro-decorated with nickel and nickel oxide layers. Electrochim Acta 53
(19):5774–5782
77. Wang PK, Ansermet JP, Rudaz SL, Wang Z, Shore S, Slichter CP, Sinfelt JH (1986) NMR
studies of simple molecules on metal surfaces. Science 234(4772):35–41. doi:10.1126/sci
ence.234.4772.35
78. Korringa J (1950) Nuclear magnetic relaxation and resonance line shift in metals. Physica
XVI(7–8):601–610
79. Blyholder G (1964) Molecular orbital view of chemisorbed carbon monoxide. J Phys Chem
68(10):2772–2777. doi:10.1021/j100792a006
80. Chen D-J, Hofstead-Duffy AM, Park I-S, Atienza DO, Susut C, Sun S-G, Tong YJ (2011)
Identification of the most active sites and surface water species: a comparative study of CO
and methanol oxidation reactions on core–shell M@Pt (M ¼ Ru, Au) nanoparticles by in situ
IR spectroscopy. J Phys Chem C 115(17):8735–8743. doi:10.1021/jp200557m
81. Chen D-J, Xu B, Sun S-G, Tong YJ (2012) Electroless deposition of ultrathin Au film for
surface enhanced in situ spectroelectrochemistry and reaction-driven surface reconstruction
for oxygen reduction reaction. Catal Today 182:46–53. doi:10.1016/j.cattod.2011.08.052
82. Lin WF, Zei MS, Eiswirth M, Ertl G, Iwasita T, Vielstich W (1999) Electrocatalytic Activity
of Ru-Modified Pt(111) Electrodes toward CO Oxidation. J Phys Chem B 103
(33):6968–6977. doi:10.1021/jp9910901
83. Lu GQ, White JO, Wieckowski A (2004) Vibrational analysis of chemisorbed CO on the Pt
(111)/Ru bimetallic electrode. Surf Sci 564(1–3):131–140
84. Spendelow JS, Babu PK, Wieckowski A (2005) Electrocatalytic oxidation of carbon mon-
oxide and methanol on platinum surfaces decorated with ruthenium. Curr Opin Solid State
Mater Sci 9(1–2):37–48
85. Watanabe M, Sato T, Kunimatsu K, Uchida H (2008) Temperature dependence of
co-adsorption of carbon monoxide and water on highly dispersed Pt/C and PtRu/C electrodes
studied by in-situ ATR-FTIRAS. Electrochim Acta 53(23):6928–6937. doi:10.1016/j.
electacta.2008.02.023
86. Ianniello R, Schmidt VM, Stimming U, Stumper J, Wallau A (1994) CO adsorption and
oxidation on Pt and Pt Ru alloys: dependence on substrate composition. Electrochim Acta 39
(11–12):1863–1869
87. Friedrich KA, Geyzers KP, Dickinson AJ, Stimming U (2002) Fundamental aspects in
electrocatalysis: from the reactivity of single-crystals to fuel cell electrocatalysts. J
Electroanal Chem 524–525:261–272
88. Zheng MS, Sun SG, Chen SP (2001) Abnormal infrared effects and electrocatalytic proper-
ties of nanometer scale thin film of PtRu alloys for CO adsorption and oxidation. J Appl
Electrochem 31(7):749–757
250 D. Chen et al.

89. Lin WF, Iwasita T, Vielstich W (1999) Catalysis of CO electrooxidation at Pt, Ru, and PtRu
alloy. An in situ FTIR study. J Phys Chem B 103(16):3250–3257
90. Brankovic SR, Wang JX, Adzic RR (2001) Pt submonolayers on Ru nanoparticles—a novel
low Pt loading, high CO tolerance fuel cell electrocatalyst. Electrochem Solid State Lett 4
(12):A217–a220
91. Vogel W, Le Rhun V, Garnier E, Alonso-Vante N (2001) Ru clusters synthesized chemically
from dissolved carbonyl: in situ study of a novel electrocatalyst in the gas phase and in
electrochemical environment. J Phys Chem B 105(22):5238–5243. doi:10.1021/jp0100654
92. Chen DJ, Tong YYJ (2015) In situ Raman spectroscopic measurement of near-surface proton
concentration changes during electrochemical reactions. Chem Commun 51(26):5683–5686.
doi:10.1039/C5CC00427F
93. Zhu Y, Uchida H, Watanabe M (1999) Oxidation of carbon monoxide at a platinum film
electrode studied by Fourier transform infrared spectroscopy with attenuated total reflection
technique. Langmuir 15(25):8757–8764. doi:10.1021/la990835r
94. Peremans A, Tadjeddine A (1994) Vibrational spectroscopy of electrochemically deposited
hydrogen on platinum. Phys Rev Lett 73(22):3010
95. Osawa M, Tsushima M, Mogami H, Samjeske G, Yamakata A (2008) Structure of water at
the electrified platinum-water interface: a study by surface-enhanced infrared absorption
spectroscopy. J Phys Chem C 112(11):4248–4256. doi:10.1021/jp710386g
96. Kunimatsu K, Senzaki T, Samjesk G, Tsushima M, Osawa M (2007) Hydrogen adsorption
and hydrogen evolution reaction on a polycrystalline Pt electrode studied by surface-
enhanced infrared absorption spectroscopy. Electrochim Acta 52(18):5715–5724
97. Kunimatsu K, Uchida H, Osawa M, Watanabe M (2006) In situ infrared spectroscopic and
electrochemical study of hydrogen electro-oxidation on Pt electrode in sulfuric acid (vol
587, p 299, 2006). J Electroanal Chem 596(2):169–169. doi:10.1016/j.jelechem.2006.07.015
98. Futamata M, Luo L, Nishihara C (2005) ATR-SEIR study of anions and water adsorbed on
platinum electrode. Surf Sci 590(2–3):196–211
99. Climent V, Gomez R, Feliu JM (1999) Effect of increasing amount of steps on the potential of
zero total charge of Pt(111) electrodes. Electrochim Acta 45(4–5):629–637
100. Bergelin M, Herrero E, Feliu JM, Wasberg M (1999) Oxidation of CO adlayers on Pt(111) at
low potentials: an impinging jet study in H2SO4 electrolyte with mathematical modeling of
the current transients. J Electroanal Chem 467(1–2):74–84
101. Vidal-Iglesias FJ, Solla-Gullon J, Campina JM, Herrero E, Aldaz A, Feliu JM (2009) CO
monolayer oxidation on stepped Pt(S) [(n-1)(100)*(110)] surfaces. Electrochim Acta 54
(19):4459–4466
102. Roth C, Benker N, Buhrmester T, Mazurek M, Loster M, Fuess H, Koningsberger DC,
Ramaker DE (2005) Determination of O[H] and CO coverage and adsorption sites on PtRu
electrodes in an operating PEM fuel cell. J Am Chem Soc 127(42):14607–14615. doi:10.
1021/ja050139f
103. Coker DF, Miller RE, Watts RO (1985) The infrared predissociation spectra of water clusters.
J Chem Phys 82(8):3554–3562
104. Richmond GL (2002) Molecular bonding and interactions at aqueous surfaces as probed by
vibrational sum frequency spectroscopy. Chem Rev 102(8):2693–2724. doi:10.1021/
cr0006876
105. Wasileski SA, Koper MTM, Weaver MJ (2001) Field-dependent chemisorption of carbon
monoxide on platinum-group (111) surfaces: relationships between binding energetics,
geometries, and vibrational properties as assessed by density functional theory. J Phys
Chem B 105(17):3518–3530. doi:10.1021/jp003263o
106. Oudenhuijzen MK, van Bokhoven JA, Ramaker DE, Koningsberger DC (2004) Theoretical
study on Pt particle adsorbate bonding: influence of support ionicity and implications for
catalysis. J Phys Chem B 108(52):20247–20254. uuid:FEE96E5D-8FA4-4172-95A9-
A6C6AC66EE05
5 Integrated Studies of Au@Pt and Ru@Pt Core-Shell Nanoparticles. . . 251

107. Dronskowski R, Bloechl PE (1993) Crystal orbital Hamilton populations (COHP): energy-
resolved visualization of chemical bonding in solids based on density-functional calculations.
J Phys Chem 97(33):8617–8624. doi:10.1021/j100135a014
108. Papoian G, Norskov JK, Hoffmann R (2000) A comparative theoretical study of the hydro-
gen, methyl, and ethyl chemisorption on the Pt(111) surface. J Am Chem Soc 122
(17):4129–4144. doi:10.1021/ja993483j
109. Glassey WV, Hoffmann R (2001) A comparative study of the p (22)-CO/M (111), M ¼ Pt,
Cu, Al chemisorption systems. J Phys Chem B 105(16):3245–3260. doi:10.1021/jp003922x
110. Glassey WV, Hoffmann R (2001) A molecular orbital study of surface–adsorbate interactions
during the oxidation of CO on the Pt(111) surface. Surf Sci 475(1–3):47–60. doi:10.1016/
S0039-6028(00)01062-1
111. Glassey WV, Papoian GA, Hoffmann R (1999) Total energy partitioning within a
one-electron formalism: a Hamilton population study of surface–CO interaction in the c
(2  2)-CO/ Ni(100) chemisorption system. J Chem Phys 111(3):893–910. doi:10.1063/1.
479200
112. Somorjai GA, Aliaga C (2010) Molecular studies of model surfaces of metals from single
crystals to nanoparticles under catalytic reaction conditions. Evolution from prenatal and
postmortem studies of catalysts. Langmuir 26(21):16190–16203. doi:10.1021/la101884s
113. Somorjai GA, Contreras AM, Montano M, Rioux RM (2006) Clusters, surfaces, and catal-
ysis. Proc Natl Acad Sci U S A 103:10577–10583
114. Somorjai GA (1994) Introduction to surface chemistry and catalysis. Wiley, New York
115. Hoffmann R (1993) A chemical and theoretical approach to bonding at surfaces. J Phys
Condens Matter 5:A1–A16
116. Hoffmann R (1971) Interaction of orbitals through space and through bonds. Acc Chem Res 4
(1):1–9
117. Housmans THM, Wonders AH, Koper MTM (2006) Structure sensitivity of methanol
electrooxidation pathways on platinum: an on-line electrochemical mass spectrometry
study. J Phys Chem B 110(20):10021–10031. doi:10.1021/jp055949s
118. Huang Z, Kim K-J (2007) Review of X-ray free-electron laser theory. Phys Rev ST Accel
Beams 10(3):034801–034826. doi:10.1103/PhysRevSTAB.10.034801
119. Shearing P, Wu Y, Harris SJ, Brandon N (2011) In situ X-ray spectroscopy and imaging of
battery materials. Interface 20:43–47. uuid:27465B6F-DD87-45E1-8587-699AFDB658FF
120. Huang L, Sorte EG, Sun SG, Tong YYJ (2015) A straightforward implementation of in situ
solution electrochemical 13C NMR spectroscopy for studying reactions on commercial
electrocatalysts: ethanol oxidation. Chem Commun 51(38):1–3. doi:10.1039/C5CC00862J
121. DeVience SJ, Pham LM, Lovchinsky I, Sushkov AO, Bar-Gill N, Belthangady C, Casola F,
Corbett M, Zhang H, Lukin M, Park H, Yacoby A, Walsworth RL (2015) Nanoscale NMR
spectroscopy and imaging of multiple nuclear species. Nat Nanotechnol 10(2):129–134.
doi:10.1038/nnano.2014.313
122. Häberle T, Schmid-Lorch D, Reinhard F, Wrachtrup J (2015) Nanoscale nuclear magnetic
imaging with chemical contrast. Nat Nanotechnol 10(2):125–128. doi:10.1038/nnano.2014.
299
123. Rugar D, Mamin HJ, Sherwood MH, Kim M, Rettner CT, Ohno K, Awschalom DD (2015)
Proton magnetic resonance imaging using a nitrogen-vacancy spin sensor. Nat Nanotechnol
10(2):120–124. doi:10.1038/nnano.2014.288
Chapter 6
Recent Development of Platinum-Based
Nanocatalysts for Oxygen Reduction
Electrocatalysis

David Raciti, Zhen Liu, Miaofang Chi, and Chao Wang

6.1 Introduction

Fed by hydrogen and oxygen gases, a fuel cell generates electricity from the
hydrogen oxidation reaction at the anode and the ORR at the cathode. As it has
water as the only product, fuel cell is considered to be environmentally friendly.
With hydrogen produced from water splitting with solar electricity and oxygen
from the air, it is also a renewable solution for chemical-electrical energy conver-
sion. This has led to a broad range of applications of fuel cells, including stationary
power systems, mobile electronics and vehicles [1]. However, the need for platinum
(Pt) as catalysts for the electrochemical reactions, especially the ORR at the
cathode, has limited large-scale applications of fuel cells [2]. The sluggish kinetics
of ORR gives high overpotentials at the cathode, typically several hundred milli-
volts on Pt catalysts, and thereby large amounts of Pt are needed for practical
applications. In order to lower the system cost, substantial improvement of the
catalyst’s performance is demanded for reduced use of Pt. It is expected that a five-
fold enhancement of the ORR activity versus state-of-the-art Pt/C catalysts is
required to enable the commercial implementation of fuel cells in
transportation [3].
Various approaches have been developed to improve the performance of Pt
electrocatalysts for the ORR, including optimization of Pt particle size and shape,
alloying Pt with 3d transition metals and development of composite nanostructures.

D. Raciti • Z. Liu • C. Wang (*)


Department of Chemical and Biomolecular Engineering, Johns Hopkins University,
Baltimore, MD 21212, USA
e-mail: [email protected]
M. Chi
Center for Nanophase Materials Sciences, Oak Ridge National Laboratory,
Oak Ridge, TN, USA

© Springer International Publishing Switzerland 2016 253


K.I. Ozoemena, S. Chen (eds.), Nanomaterials for Fuel Cell Catalysis,
Nanostructure Science and Technology, DOI 10.1007/978-3-319-29930-3_6
254 D. Raciti et al.

Throughout these approaches the catalyst surface is maintained to be Pt, as Pt is


likely the only element that is both active and stable under PEMFC conditions.
Therefore, the design of advanced electrocatalysts beyond Pt usually goes through
the manipulation of subsurface nanoscale architectures to induce modifications to
the surface geometry and/or electronic structures, and thus alter the surface adsorp-
tion and catalytic properties of Pt.
Here we provide a focused review of the recent advancement of Pt-based
electrocatalysts for the ORR. We will discuss the various efforts on the develop-
ment of Pt and Pt-based alloy and composite nanostructures for electrocatalytic
applications, especially how to tailor surface structures towards enhanced catalytic
activity and stability for the ORR. Due to the presence of various protocols for ORR
measurements, we will focus our discussion on relative activities, typically
expressed in terms of improvement factors versus Pt bench marker. Though our
emphasis will be placed on high-surface-area nanocatalysts, we will also try to link
the structure-property relationships of nanocatalysts to those established on
extended-surface model catalysts. Possible approaches towards further improve-
ment of the Pt-based nanocatalysts will also be discussed.

6.2 Single-Component Pt Catalysts

Albeit the focus of research efforts is placed on alloy and composite nanostructures,
most of the catalysts employed in present fuel cells are still single-component Pt
particles supported on high-surface-area carbon. The Pt/C catalysts have been
extensively studied for more than half a century, with the ORR catalytic activity
found to be correlated to the particle size and shape [4]. The presence of extensive
studies in history, however, has not impeded the continual effort on this simple
system, where new insights and nanostructures with substantial catalytic improve-
ment have been developed.

6.2.1 Particle Size Effect

One purpose of using nanoparticulated materials as catalysts is to harvest the large


surface/volume ratio and number of surface sites. For many reactions, it is also
found that the catalytic activity of smaller particles, in terms of turnover frequency
(TOF), is higher than larger particles or extended surfaces due to the reduced
coordination of surface atoms [5–9]. These two trends ask for decreasing particle
size in catalytic improvement. The scenery of ORR on Pt, however, is different. It
has long been realized that the catalytic activity of Pt for the ORR is dependent on
particle size, with smaller particles giving lower specific activity (current normal-
ized by the real surface area of the catalyst) than larger ones. It is believed that the
size-dependent ORR activity originates from the varying ratios of ordered surface
6 Recent Development of Platinum-Based Nanocatalysts for Oxygen Reduction. . . 255

facets, i.e., (1 1 1) and (1 0 0) for cubo-octahedral particles, and edge and corner
sites on the surface as the particle size changes [4].
However, debates still exist in the literature. For example, Watanabe
et al. argued that the specific activity of ORR on Pt does not depend on the particle
size but rather on the interparticle distance [11]. In their studies of Pt clusters for the
ORR Yamamoto et al. found the Pt clusters of sizes smaller than 2 nm have higher
specific activities than commercial Pt/C catalysts (2.5 nm) [12]. On the other side, a
recent study by Shao et al. [10] confirmed the size-dependent ORR catalytic activity
with Pt/C catalysts prepared by layer-by-layer growth on the same seeds with
underpotential deposition methods, which can be considered to be with reasonable
control over the interparticle distance. They observed an increase of specific
activity with particle size and the maximum mass activity (current normalized by
the mass of Pt) at an intermediate particle size (Fig. 6.1a), consistent with the
previous trends established by Kinoshita [4] except that mass activity achieves the
maximum at ~2.2 nm (versus ~3 nm in Kinoshita’s report). Moreover, they calcu-
lated the oxygen binding energies for the particles of various sizes and was able to
clearly assign the increase in ORR specific activity with particle size to the increase
of Pt–O binding energy (Fig. 6.1b, due to the diminishing occupation of edge sites
on larger particles) [10], as the rate of ORR on Pt is limited by the desorption of OH.
It is noticed by this review that size-independent ORR activities have also been
reported recently. Arenz et al. did not observe a clear difference in ORR activity
among commercial Pt/C catalysts of 1–5 nm, though substantially different ORR
activities were found between unsupported Pt black (~30 nm) and the 1–5 nm Pt/C
catalysts [13]. Shao-Horn et al. revealed size-independent ORR activities on Pt
catalysts below 5 nm, which was ascribed to the similarity of surface compositions
and surface electronic structures of Pt particles below 5 nm as well as comparable
OHad coverage at the potential at which ORR was measured, though the catalysts
with different particle sizes in this study were produced by thermal treatment which
had caused increase of size dispersion [14].
From above one can see that the particle size effect, one of the most studied
topics of relevance to the ORR, still remains elusive for Pt electrocatalysts. While
the different experimental protocols with varying potential windows, catalyst
loading, electrolytes, etc. may all contribute to the controversial observations, the
materials that have been studied, the various Pt/C catalysts, could be of particular
importance. Commercial catalysts prepared by impregnation methods usually pos-
sess relatively wide distributions of particle size and shape, which may not be
considered as appropriate starting materials for fundamental studies. Development
of better defined catalysts, with monodisperse (but variable) and uniform (of the
same shape) particles by controlled synthesis is expected to resolve these issues and
help uncover the intrinsic particle size effect of Pt electrocatalysts for the ORR.
While the studies on particle size effect have been focused on the ORR activity,
it is necessary to point out that the size of Pt particles also plays an important role in
the stability of ORR electrocatalysts. In their studies of Pt/C catalysts of different
sizes Shao-Horn et al. also observed increasing catalyst stability upon potential
cycling as the particle size increases (Fig. 6.2) [14]. It is suggested that the
256 D. Raciti et al.

Fig. 6.1 (a) Size dependence of specific activity (blue diamond) and mass activity (red square) of
Pt/C for oxygen reduction reaction at 0.93 V; (b) averaged oxygen binding energy calculated as a
function of particle size for cubo-octahedrons (Reprinted with permission from [10]. Copyright
2011 ACS)

degradation of Pt/C catalysts in PEMFCs can undergo different routes, including Pt


dissolution and migration, Ostwald ripening and particle detachment from carbon
support [15]. Larger Pt particles are more stable as their surface atoms possess
lower chemical potential for dissolution or movement [16] which could be due to
the higher average coordination and presence of fewer defects. Therefore, a balance
between the both size-dependent activity and stability is desired for the design of
6 Recent Development of Platinum-Based Nanocatalysts for Oxygen Reduction. . . 257

Fig. 6.2 Specific electrochemical surface area (ESA) of Pt NPs as a function of the number of
potential cycles for Pt-46 % (2.0  0.6 nm, black squares), Pt-46 %-900  C-1 m (3.1  0.7 nm, red
circles), and Pt-46 %-900  C-2 h (4.7  1.2 nm, blue triangles) (Reprinted with permission from
[14]. Copyright 2012 ECS)

ORR electrocatalysts and the optimal particle size could be different from that
established by simply considering the maximal mass activity.

6.2.2 Particle Shape Effect

The ORR is a structure-sensitive reaction on Pt catalysts. Studies of single-crystal


Pt extended surfaces have revealed that the rate of ORR follows the trend (1 1 0) >
(1 1 1) > (1 0 0) [17, 18]. It is thus natural to pursue shape control over Pt catalysts
for enhanced electrocatalytic activity. However, except for a few reports [19–21],
the studies of shape-dependent ORR catalysis are very limited for Pt nanocatalysts
(in sharp contrast to the case of Pt-based alloy catalysts to be discussed in the
following). The challenges are mainly present in the difficulty of selectively
growing Pt nanocrystals enclosed by (1 1 0) or (1 1 1) facets, whereas synthesis of
Pt nanocubes with (1 0 0) surface has been found to be relatively easy and reported
by many groups [22–25]. In this sense, substantial effort may still be needed in
elucidating the particle shape effect in the ORR and correlate the electrocatalytic
performance of Pt nanoctalysts to their bulk counterparts.
Besides controlling the surface structure by directly growing Pt nanocrystals
with shape control, many other studies have focused on the development of Pt
electrocatalysts of alternated morphologies beyond simple NPs for the ORR
(Table 6.1). These include the syntheses of single-crystalline Pt nanowires on
carbon spheres (Fig. 6.3a) [26], porous Pt nanotubes by galvanic displacement of
258 D. Raciti et al.

Table 6.1 Comparison of ORR activities of Pt catalysts of different morphologies


Specific Specific surface
surface area activity at 0.9 V
Ref Pt nanostructure (cm2 mg1) vs. RHE (A/g) Measuring conditions
a Nanowires 436 120 In a single cell PEMFC at
80  C
b Porous 704 88 Argon saturated 0.1 M
nanotubes HClO4 electrolyte
c Hollow 600 1150 0.1 M HClO4 at 80  C,
nanoparticles 0.9 V in 10 mV/s potential
sweeps
d Mesoporous 560 170 N2-purged 0.1M HClO4
double gyroid (aq), 23  C, sweep rate
networks 50 mV/s

Pd nanowires (Fig. 6.3b) [27], hollow Pt NPs by galvanic displacement of Ni NPs


(Fig. 6.3c) [28], and mesoporous double gyroid Pt networks (Fig. 6.3d) [29]. High
mass activity has generally been achieved in these studies by manipulation of the
nanoscale architectures to increase the exposure of surface sites, whereas the
catalyst stability is typically superior than simply reducing the particle size due to
the formation of more continuous surface. While these nanostructures possess
unique properties and can potentially lead to unconventional advanced
electrocatalysts, their scaling up and incorporation into membrane electrode assem-
blies must also be considered which may reduce their potential for practical
applications.

6.3 Pt-Based Alloy Catalysts

Bimetallic alloys of Pt and 3d transition metals such as Cr, Mn, Fe, Ni, Co and Cu
have been extensively studied as electrocatalysts for the ORR [18, 30–35]. In these
alloys, the 3d transition metals are not directly involved in the ORR, but rather
enhance the catalytic activity of surface Pt atoms through electronic and geometric
modifications. Such modifications can lead to weaker Pt–O binding and reduced
adsorption of oxygenated spectator species, e.g., OH, on the surface, provide more
active sites accessible for molecular oxygen and thus improve the catalytic activity
[18, 31–33].
Although having promising catalytic activity, challenges still exist for the
employment of Pt-alloy catalysts for the ORR. The main concern is that the 3d
metals have much lower oxidation potential than Pt and they may be leached out
under the running conditions in PEMFCs [36]. Therefore, the architecture of
surface structures becomes critical for the design of Pt-alloy catalysts, which do
not only determine the catalytic activity, but also long-term durability of the
catalysts.
6 Recent Development of Platinum-Based Nanocatalysts for Oxygen Reduction. . . 259

Fig. 6.3 (a) SEM micrographs of the as-synthesized Pt nanowire/C nanostructure, with 60 wt% Pt
on carbon (inset: HRTEM micrograph of a branched Pt nanowire grown on carbon support);
(Reprinted with permission from [26]. Copyright 2008 Wiley) (b) HRTEM image of Pt nanotubes
(Reprinted with permission from [27]. Copyright 2010 Wiley); (c) STEM images of Pt hollow
nanoparticles (inset: High-resolution STEM images of Pt hollow particles) (Reprinted with
permission from [28]. Copyright 2011 ACS); (d) TEM image of the Pt mesoporous double gyroid
networks (Reprinted with permission from [29]. Copyright 2012 ACS)

Conventionally Pt-alloy catalysts are prepared by impregnation of 3d transition


metals into Pt/C catalysts [37, 38]. This type of synthesis usually involves high-
temperature thermal treatment to improve alloy homogeneity. The induced particle
sintering and agglomeration cause loss in surface area and control over particle size
and size distribution, making it challenging for the systematical study of the
Pt-alloy catalysts. Recently progress has been made on the synthesis of Pt alloy
nanoparticles (NPs) based on organic solution synthesis [39–48]. Monodisperse and
homogeneous colloidal NPs of the Pt-bimetallic alloys have been made by coupling
the reduction of Pt salts with the decomposition of organometallic precursors of 3d
transition metals [42–45] or by subsequent addition of the metal salt precursors
[44, 46–48], to address the different reduction and growth rates between Pt and 3d
metals. Pretreatments for surfactant removal and surface cleaning are usually
applied [45, 49–52] either directly on the as-synthesized NPs or after their
260 D. Raciti et al.

Fig. 6.4 Size dependent specific activity and specific surface area for the Pt3Co/C catalysts of
various particle sizes (Reprinted with permission from [45]. Copyright 2011 ACS)

deposition onto high-surface-area carbon, before the application as catalysts. By


separation of the nanoparticle growth and particle deposition onto support, control
over particle size, shape and composition has been achieved by the colloidal
approach, which has enabled systematic studies of the Pt-alloy catalysts for
the ORR.

6.3.1 Particle Size Effect

Based on monodisperse and homogeneous Pt3Co NPs obtained from organic


solution synthesis, the effect of particle size on electrocatalytic activity has been
investigated for Pt-bimetallic catalysts for the ORR (Fig. 6.4) [45]. It was found that
while the specific surface area (surface area normalized by the mass of Pt) decreases
as the particle size increases, the specific activity for the ORR increases due to less
oxophilic properties of the surface for large particles [3, 4, 53]. NPs of an interme-
diate size, ~4.5 nm, were found to give the maximum in mass activity, with an
improvement factor of about 3 versus Pt/C of similar particle size. This volcano-
type dependence is similar to the size-dependence activity of Pt/C (Fig. 6.1), where
the difference in the particle size maximizing mass activity implies the dissimilar
surface structures between these two systems (see the following section).
6 Recent Development of Platinum-Based Nanocatalysts for Oxygen Reduction. . . 261

6.3.2 Particle Shape Effect

Previous studies of Pt3Ni(h k l) single crystals have shown that the ORR activity is
an order of magnitude higher on Pt3Ni(1 1 1) than on Pt(1 1 1) and other low-index
Pt surfaces (Fig. 6.5) [18]. This has inspired significant efforts on the synthesis of Pt
alloy nanocatalysts of octahedral shape aiming at achieving the same level of
catalytic enhancement [54–56]. Pt3Ni NPs of octahedral shape with (1 1 1) faceted
surfaces were found to be ~4 and ~6 times more active than Pt3Ni nanocubes and Pt
nanocubes with (1 0 0) surfaces of comparable sizes, respectively (Fig. 6.6)
[54]. However, in another study Yang et al. only observed an improvement factor
of ~1.5 for the particle shape evolved from cube to octahedron [55] and ~2.2 to
icosahedron [57] for Pt3Ni. Such a divergence indicates that any trivial difference in
synthetic and catalyst preparation methods and/or particle sizes could lead to
substantially different catalytic performance, pointing to the lack of fundamental
understanding for the surface structure-property relationship of alloy nanocatalysts.
Although the single-crystal work [18] is usually cited for the dependence of
ORR activity on surface structure, it is necessary to point out that the nanoscale
alloy catalysts could possess different surface structures from the well-defined
extended surfaces and several factors might be present to impede the
achievement of similar level of catalytic improvement. First and most importantly,
the ORR activity of Pt3Ni(1 1 1) shown in Fig. 6.5 was recorded on single crystals
with a Pt-skin surface. This structure is formed by surface segregation via

Fig. 6.5 Catalytic activity trends of single-crystal Pt and Pt3Ni (h k l) surfaces for the ORR in
correlation to the d-band center positions (Reprinted with permission from [18]. Copyright
207 AAAS)
262 D. Raciti et al.

Fig. 6.6 (a) TEM and (b) HRTEM images of Pt3Ni nanoctahedra; (c) Shape dependent
electrocatalytic activities of Pt3Ni nanocatalysts for the ORR (Reprinted with permission from
[60]. Copyright 2010 ACS)

high-temperature (~1000 K) annealing in ultrahigh vacuum, which is unlikely


present in as-synthesized Pt alloy nanocatalysts. The same alloy catalysts but
with different surface structure could behave drastically differently in the ORR,
as will be discussed in the following (see Sect. 6.3.4). In principle, annealing could
bring about similar structuring processes in alloy nanocatalysts, but it has been
shown that high-temperature annealing also induces particle aggregation and
agglomeration, making it challenging to determine the effect of surface structure
in alloy nanocatalysts [58]. Secondly, alloy nanoparticles, even those claimed to be
in octahedral shape, have truncated corners and the surface comprises a certain
portion of other facets such as (1 0 0) and (1 1 0), which have much lower catalytic
activity than (1 1 1) surface of Pt alloys (see Fig. 6.5). An additional but not the last
factor is the presence of more defects on alloy nanocatalysts than on single-crystal
surfaces. Surfaces of nanoparticles could be enriched with steps, vacancies and
6 Recent Development of Platinum-Based Nanocatalysts for Oxygen Reduction. . . 263

Fig. 6.7 (a) TEM and (b) HRTEM images of Pt2.5Ni octahedral nanoparticles; (c, d) Comparison
of the ORR activities of Pt2.5Ni/C nanocatalysts with commercial cubo-octahedral Pt/C and Pt3Ni/
C catalysts (Reprinted with permission from [56]. Copyright 2013 ACS)

adatoms owing to the fast kinetics of nanocrystal nucleation and growth in typical
wet-chemical methods (as labelled by arrows in Fig. 6.7b). Similar to the corner and
edge sites discussed in the particle size effect, such defect sites have smaller
coordination number, higher oxophilicity and lower ORR activity than their coun-
terparts on ordered extended surfaces (which is more likely the case on single
crystals).
Besides the intrinsic structure difference, the method of catalyst preparation,
especially the way of cleaning and removing organic surfactants, could also affect
the catalytic performance of nanomaterials derived from solution synthesis [59]. A
recent study of octahedral Pt-Ni alloy nanocatalysts shows more than one order of
magnitude of enhancement in mass activity compared to commercial Pt (Fig. 6.7)
[56], with the more substantial enhancement than previous studies [54, 55] ascribed
to the adoption of acetic acid washing for surface cleaning. Although surface
defects are still visible in their HRTEM images, this approach is believed to
produce clean octahedral particles with better preserved (1 1 1) facets than the
other methods available in the literature [56].
264 D. Raciti et al.

6.3.3 Composition Effect and Dealloying

Despite their enhanced catalytic activity, concerns are present about the stability of
Pt-alloy catalysts for implementation into PEMFCs. Owing to their low oxidation
potentials, the 3d transition metals exposed on the surface of Pt-alloy catalysts
dissolve almost instantaneously in acidic electrochemical environments, as dem-
onstrated by low energy ion scattering (LEIS) studies on extended Pt-alloy surfaces
(Fig. 6.8) [61]. It has also been found in Pt-alloy nanocatalysts that a Pt-rich shell is
formed as a result of the depletion of non-precious metals in the near-surface
regions (also called dealloying) [48, 62–67]. Systematic studies based on PtxNi1x
(0 < x < 1) alloy NPs from organic solution synthesis show that the extent of
depletion, or the thickness of the formed Pt-rich shell, is dependent on the initial
alloy composition of the catalysts (Fig. 6.9) [48]. For the same particle size of
~5 nm, a maximum ratio of ~27 % was preserved for Pt0.5Ni0.5 among various
compositions, for which the Pt-rich shell thickness found to be 0.7~1 nm,
corresponding to about two atomic layers. Moreover, the dependence of ORR
catalytic activity on the composition for the PtxNi1x catalysts was found to follow
the same trend for the ratio of Ni left after surface depletion, with Pt0.5Ni0.5 gave the
highest activity among the investigated series [48]. Other studies on the Pt-alloy
catalysts of different compositions also revealed the dependence of ORR activity on
the initial alloy composition, albeit the formation of different nanostructures after
dealloying depending on the particle size and initial element distribution [67–70].
Strasser et al. studied the core/shell structures formed by dealloying of PtCu
alloy catalysts with greater details [71–74]. It was argued that the formed Pt-rich
shell is under compressive strain due to the smaller atomic size of Cu in comparison
to Pt. The composition-dependent ORR catalytic activity was thus ascribed to the
varying surface strain depending on the shell thickness, which is further determined
by the initial alloy compositions (Fig. 6.10) [71].
Intensity [arb. units]

0.2 0.4 0.6 0.8


0.2 0.4 0.6 0.8
E/E0 E/E0

Fig. 6.8 Schematic and LEIS studies for the formation of Pt skeleton after exposure of Pt-alloy
surfaces to the electrochemical environment (Reprinted with permission from [61]. Copyright
2006 ACS)
6 Recent Development of Platinum-Based Nanocatalysts for Oxygen Reduction. . . 265

Fig. 6.9 (a) Composition


line profiles for the (a)
PtxNi1x/C nanocatalysts
after electrochemical
depletion of surface Ni; (b)
summary of the specific
activities for the ORR at
0.9 V and improvement
factors versus Pt/C of he
PtxNi1x/C catalysts
(Reprinted with permission
from [48]. Copyright 2011
Wiley)

(b) 5

Improvement Factor (vs. Pt/C)


Specific Activity (mA/cm2)

3
3

2
Pt3Ni PtNi PtNi2 PtNi3

After surface depletion, the preserved 3d metal in the particle core may still not
be stable under the running conditions of PEMFCs. Two mechanisms can lead to
further degradation of Pt-bimetallic catalysts on the course of potential cycling,
even after pretreatment by acid. Redox reactions induced by potential cycling can
cause the dissolution of surface Pt atoms, exposing the non-precious atoms in the
subsurface [68, 75]. The embedded 3d metal atoms in the subsurface can also
diffuse to the surface and dissolve under the oxidative and acidic conditions in
the ORR, driven by their higher oxophilicity than Pt. The latter effect was demon-
strated by potential cycling of acid treated Pt3Co catalysts in alkaline electrolytes,
where the diffusion of 3d metal atoms onto surface was visualized by the appear-
ance and growth of Co2+/Co3+ redox peaks (Fig. 6.11) [36]. The observed fast
266 D. Raciti et al.

a b 0.12

0.10
Pt shell
0.08
1

ORR activity (eV)


2 0.06 Pt25Cu75
Pt50Cu50
0.04
Pt-Cu core Pt75Cu25
Pt25Cu75
0.02
Pt50Cu50
acore 0.00
Pt75Cu25
Pt
−0.02
ashell
−4 −3 −2 −1 0
Strain (ashell−aPt)/aPt (%)

Fig. 6.10 (a) A simple structural two-phase core/shell model for the dealloyed PtCu NPs; (b)
Experimentally observed (solid triangles) and predicted (black dashed line) relationships between
ORR activity and lattice strain for the dealloyed PtCu catalysts Reprinted by permission from
Macmillan Publishers Ltd. [Nature Chemistry] [71] Copyright 2010

Fig. 6.11 CV series (50 mV/s scan rate at room temperature) recorded for an acid leached Pt3Co
catalyst in Ar saturated 0.1 M KOH electrolyte. The gradual segregation of Co atoms to the surface
due to potential cycling was indicated by appearance of Co2+/Co3+ peaks (labelled by arrows)
(Reprinted with permission from [36]. Copyright 2009 ACS)

degradation of Pt-alloy catalysts under potential cycling is likely associated with


the disordered surface structure of dealloyed catalysts. The Pt-skeleton shell formed
after 3d metal leaching is usually enriched by defects, such as undercoordinated
6 Recent Development of Platinum-Based Nanocatalysts for Oxygen Reduction. . . 267

atoms and vacancies (Fig. 6.8) [68, 76]. The surface atoms with coordination
numbers smaller than their counterparts on ordered surfaces are less stable and
can be easier to dissolve [77]. The vacancies present in the Pt-skeleton shell could
provide the pathway for diffusion of the 3d metal atoms from the core to the surface.
Therefore, the Pt-skeleton surfaces formed by either acid pretreatment or in situ
electrochemical leaching may not be able to protect the nonprecious metal atoms
that remain in the particles.

6.3.4 Alloy Catalysts with Pt-Skin Surface

As discussed above (in particle size and shape effects), ORR electrocatalysis is
structure-sensitive and the absence of a Pt-skin surface on octahedral nanocatalysts
has impeded the achievement of the same level of catalytic enhancement as
demonstrated on single-crystal extended surfaces. Meanwhile, from the prior sec-
tion it is noticed that the continual loss of 3d metals causes the degradation of
Pt-bimetallic catalysts towards Pt-like catalytic performance. Because of these
undesirable factors, improvement of only 2–3 times in mass activity was observed
in Pt ally nanocatalysts [63, 64, 68, 78–80], which is substantially lower than that
expected based on fundamental studies of extended surfaces (Fig. 6.5). Moreover,
the dissolved 3d metal ions from alloy catalysts were found to be detrimental to the
cell performance of MEA, e.g., causing accumulation of protons at the cathode
[1]. Therefore, it is of utmost importance to protect the non-precious component for
the employment of Pt-alloy catalysts in PEMFCs.
While using large particles can generally preserve more non-precious metal
contents than smaller ones, considering that a thicker Pt shell can sustain longer
and also slow down the diffusion of 3d metals [67], it results in the sacrifice of
specific surface area and reduces the gain in mass activity. A more desirable
solution would be modification of the surface structure towards smooth and ordered
Pt surfaces which can better protect the subsurface 3d metals [18]. Studies of
extended surfaces of Pt alloys in ultrahigh vacuum (UHV) show that thermal
annealing (~1000 K) induces the segregation of Pt to the topmost atomic layer,
which is counterbalanced by the depletion of Pt in the next two or three subsurface
layers. The formed pure Pt surface, termed Pt skin (Fig. 6.8) [61], exhibits distinc-
tive electronic and adsorption properties, and much higher catalytic activity than
pure Pt and Pt-skeleton surfaces. Moreover, the formed Pt-skin surface is found to
be able to protect subsurface 3d metals from leaching out and preserve the bene-
ficial surface and catalytic properties under electrochemical conditions [18, 32, 61].
Inspired by the findings on extended surfaces, efforts have been dedicated to
modify the surface structures of Pt-alloy nanocatalysts towards Pt-skin surfaces
[66, 76, 81–85]. For example, in order to induce surface segregation, thermal
annealing was applied to commercial Pt3Co/C nanocatalysts. The observation of
a bright outermost atomic layer followed by a dark second layer was ascribed to the
formation of Pt-skin surface induced by surface segregation (Fig. 6.12) [66]. The
268 D. Raciti et al.

Fig. 6.12 Electron microscopic studies of thermally annealed Pt3Co/C catalysts. (a) Cs-corrected
HAADF-STEM image. (b) Normalized intensity (solid curve) of the particle in (a) along the
dotted line, versus normalized thickness variations for an ideal truncated octahedron particle
(Reprinted with permission from [66]. Copyright 2008 ACS)

a
acid 400°C

PtNi/ C Pt-skeleton PtNi/C Pt-skin PtNi/C


b c d
Pt
Intensity (a. u.)

Intensity (a. u.)


Intensity (a. u.)

Ni

0 1 2 3 4 5 0 1 2 3 4 5 0 1 2 3 4 5
Position (nm) Position (nm) Position (nm)
e f g
60 1.0 0.4
Specific Surface Area (m2/g)

Specific Activity (mA/cm2)

before
Mass Activity (A/mg)

Improvement Factor

before before
Improvement Factor

50 0.8
after after 6 after 6
0.3
40
(vs.Pt/C)

(vs.Pt/C)

0.6
30 4
0.2 4
0.4
20
2
0.1 2
10 0.2

0 0.0 0
0.0 0
Pt/C Pt-skeleton Pt-skin Pt/C Pt-skeleton Pt-skin Pt/C Pt-skeleton Pt-skin
PtNi/C PtNi/C PtNi/C PtNi/C PtNi/C PtNi/C

Fig. 6.13 (a) Schematic and (b–d) composition line profile analysis for the nanostructure
evolution in the synthesis of PtNi/C catalyst with multilayered Pt-skin surface; Summary of
electrochemical durability studies obtained by RDE before and after 4000 potential cycles between
0.6 and 1.1 V for the Pt/C and PtNi/C catalysts in 0.1MHClO4 at 0.95 V and 60  C: (e) specific
surface area, (f) specific activity, and (g) mass activity (Reprinted with permission from
[76]. Copyright 2011 ACS)

annealed catalyst also exhibited ~4 times enhancement in ORR catalytic activity,


versus ~2 times for the acid treated catalyst.
A more promising approach towards the Pt-skin surface other than simple
thermal annealing is a two-step process by successive acid and thermal treatments
of Pt-alloy catalysts (Fig. 6.13a) [76]. Starting with homogeneous Pt0.5Ni0.5 NPs
6 Recent Development of Platinum-Based Nanocatalysts for Oxygen Reduction. . . 269

derived from organic solution synthesis, acid leaching was first applied to form a
Pt-rich skeleton shell. The acid treated catalyst was then annealed at ~400  C. This
mild-temperature treatment may not be sufficient to induce complete surface
segregation to form monolayer Pt-skin, but was found to be able to smooth the
surface, reduce under coordinated defect sites, and form an ordered Pt skin without
inducing particle size change. Composition line profile analysis shows that the
formed Pt-skin structure has a thickness of ~0.5 nm, corresponding to about two
atomic layers (Fig. 6.13b–d) [76]. This multilayered Pt-skin surface was further
found to be capable of preserving the beneficial properties of subsurface Ni, which
exhibited an enhanced ORR catalytic activity ~6 times greater than Pt/C particles
with similar size (~5 nm). Moreover, the multilayered Pt-skin surface also shows
improved durability than the Pt-skeleton surface (Fig. 6.13e–g), preserving more
than 80 % of the specific activity after extensive potential cycling versus less than
50 % for the Pt-skeleton surface. The improved durability was further confirmed by
the elemental analysis of the catalysts after electrochemical studies showing that
more Ni was preserved by the Pt-skin than the Pt-skeleton surface.

6.4 Composite Nanostructures

Besides alloys of Pt and 3d transition metals, composite nanostructures, such as


core/shell [86–92], heterodimer [93, 94], nanotube [95, 96], and nanoporous
[70, 97], have also been explored for electrocatalytic applications. Some of these
nanostructures are still based on bimetallic systems of Pt and 3d transition metals
[77, 87, 88, 90], but with more sophisticated nanostructures. By going beyond
simple particulated catalysts, these composite nanostructures provide additional
dimensions for tuning the surface catalytic properties and have brought about
additional catalytic improvement.

6.4.1 Pt-Based Core/Shell Nanostructures

Core/shell nanostructures are of particular interest for the development of ORR


electrocatalysts considering the potential of reducing the cost of Pt electrocatalysts
by replacing catalytically inaccessible inner Pt atoms by cheaper, less noble metals.
For example, Pd/Pt core/shell nanocatalysts have been synthesized by galvanic
displacement of under potentially deposited (UPD) Cu with Pt on Pd and other
metal or alloy seeds, with the ORR catalytic activity dependent on the shell
thickness (Fig. 6.14) [86, 98–102]. Similar to the case in core/shell nanostructures
formed by dealloying [71], the Pt surface on the Pd/Pt core/shell particles was also
believed to be under compressive strain due to the lattice mismatch between the
core and shell materials, leading to the improved ORR catalytic activity [86].
270 D. Raciti et al.

a b 0.0
ORR in HCIO4: 10 mVs-1

Current density (mA cm-2)


-1.0

-2.0 PdcPt1

-3.0 PdcPt2

-4.0 PdcPt3

-5.0

-6.0

Pt 0.4 0.5 0.6 0.7 0.8 0.9 1.0


Pd
E vs RHE (V)

Fig. 6.14 (a) Projection of the structure model of Pd/Pt core/shell particles formed by galvanic
displacement of UPD Cu; (b) ORR polarization curves measured for Pd/Pt core/shell catalysts
with one- to three-monolayer of Pt shell (Reprinted with permission from [86]. Copyright 2009
ACS)

On the other side, with an electrochemically more stable noble gold metal as the
core, multimetallic Au/FePt3 core/shell NPs were found to exhibit high durability
under potential cycling in the ORR-relevant potential regions (Fig. 6.15)
[91]. While significant changes to the particle size and shape were observed for
Pt/C and Pt-alloy (FePt3/C) catalysts after extensive potential cycling, the changes
to Au/FePt3 were almost negligible. The stabilization of particle morphology and
thus electrocatalytic durability by the Au core was further correlated to a counter
balance between the segregation of Au to the surface and stronger binding of Pt–O
vs. Au–O. It is worth pointing out that stabilization of Pt electrocatalysts by Au was
also observed previously in Au clusters-modified Pt/C catalysts, where the
enhanced durability was ascribed to the raised oxidation potential of Pt [93].

6.4.2 Pt-Based Nanostructured Thin-Films (NSTFs)

Pt-based NSTFs represents another group of electrocatalysts with unique composite


nanostructures. Developed by 3M researchers, NSTF catalysts were synthesized by
sputtering Pt-based metallic films over arrays of polymer nanowires [103, 104]. The
concept of NSTF catalysts is drawn from the fact that the ORR activities on high-
surface-area nanocatalysts are much lower than on their extended-surface counter-
parts, with typical differences of up to one order of magnitude [13]. While extended
surfaces, even planar thin films, cannot be directly adopted as electrocatalysts for
PEMFCs, NSTFs can combine the high surface area of nanocatalysts and high
6 Recent Development of Platinum-Based Nanocatalysts for Oxygen Reduction. . . 271

Fig. 6.15 (a) HAADF-STEM characterization and elemental mapping of Au (green), Pt (red) and
their overlap; (b) TEM characterization of the catalysts before and after the potential cycling
(Reprinted with permission from [91]. Copyright 2011 ACS)

catalytic activity of extended surfaces, achieving advanced catalytic performance


that could be hard to achieve by regular nanoparticulated catalysts.
The as-synthesized NSTFs by sputtering, though already exhibit much high
ORR activities than nanoparticulated Pt/C, still have rough surfaces which are
full of defects (Fig. 6.16a). It was recently found that further annealing of these
NSTFs can smoothen the surface, induce surface segregation and further improve
272 D. Raciti et al.

Fig. 6.16 (a) High-resolution SEM image of Pt-based NSTFs; (b) Specific activities of ORR
measured by RDE in 0.1 M HClO4 with 1600 rpm, 20 mV s1 at 0.95 V with corresponding
improvement factors versus polycrystalline Pt. Reprinted by permission from Macmillan Pub-
lishers Ltd: Nature Materials [96] Copyright 2012

the catalytic activity by another factor of two (Fig. 6.16b) [96]. The ORR specific
activity obtained on the formed mesostructured NSTFs reaches 2.4 mA/cm2 at
0.95 V vs. RHE, representing improvement factors of 8 and 20 versus polycrystal-
line Pt and commercial Pt/C catalysts [18].
Porous materials with nanoscale pore structures have wide applications in
heterogeneous catalysis. Among the various nanoporous catalysts zeolite is prob-
ably the most important system which has been employed in petrochemical industry
for catalytic cracking, isomerization and reforming [105–107]. Nanoporous gold
(NPG) has also been studied as catalysts for selective oxidation [108–110] and
hydrogenation [110–112] reactions. The advantages of nanoporous materials as
catalysts do not only include the distinctive surface structures exposed inside the
pores, but also the unique nanoscale confinement effects which can be tailored to
improve catalytic activity and selectivity.
Erlebacher et al. developed the synthesis of nanoporous Pt3Ni foams (Fig. 6.17a)
[97] as well as NPs (Fig. 6.17b) [70] by controlled electrochemical dealloying. By
impregnation of a hydrophobic and protic ionic liquid with high oxygen solubility
into the nanoporous Pt3Ni foams, they reported a kinetic current density as high as
18.2 mA/cm2 at 0.9 V vs. RHE [97], which approaches the highest activity obtained
on single-crystal extended surfaces [18]. More recently, Stamenkovic et al. applied
this approach to 3D Pt3Ni nanoframes derived from Ni-rich Pt-Ni alloy polyhedral
nanoparticles and reported improvement factors of >20 in specific activity and >30
in mass activity versus commercial Pt/C (Fig. 6.18) [113]. Such improvement
factors are already in line with the previous results obtained on extended surfaces
(Fig. 6.5). Although the catalytic enhancement also arises from the boosted oxygen
mass transport by using the ionic liquid, it is believed that the edges in the
nanoframes possess highly crystalline surfaces that are good mimic of ordered
facets on extended surfaces, likely due to the unique 1D morphology in the
hierarchical nanoscale architecture.
6 Recent Development of Platinum-Based Nanocatalysts for Oxygen Reduction. . . 273

Fig. 6.17 (a) Cross-sectional TEM image of a nanoporous-Pt3Ni foam. The inset shows a lower
magnification view of a sharp interface between the porous dealloyed region (left) and undealloyed
metal (right). Reprinted by permission from Macmillan Publishers Ltd: Nature Materials [97]
copyright 2010; (b) HRTEM image of a 15 nm nanoporous Pt3Ni NP (Reprinted with permission
from [70]. Copyright 2012 ACS)

6.5 Summary and Future Directions

Over the past two decades significant progress has been made in the development of
Pt-based nanomaterials for electrocatalytic applications in fuel cells. We briefly
reviewed the recent development of Pt-based ORR electrocatalysts in terms of
single-component Pt, bimetallic alloys and composite nanostructures. While these
directions may only be a few of the many towards advanced electrocatalysts, we
believe they represent the major efforts in the literature. Further improvements
along these directions could potentially achieve the performance required for the
commercial implementation of PEMFCs in stationary and transportation applica-
tions [1, 3].
Despite the progress that has been made, challenges still exist in Pt-based ORR
electrocatalysts. These include: (i) while the improvement in catalytic activity has
reached or been above the PEMFC’s target [3], the catalyst’s stability has to be
further improved for practical applications, namely much prolonged tests under
PEMFC conditions; (ii) the developed various catalysts need to be incorporated into
the membrane electrode assemblies (MEAs) for fuel cells with structural and
property integrity while maintaining their catalytic performance that are measured
by rotating disk electrodes (RDEs); (iii) scale-up production of the catalysts with
novel nanostructures at gram, or even kilogram scales. To address these challenges
it would require fine balances between catalytic activity and stability, between
sophisticated nanostructures and simple processes easy for scaling up, as well as
between improvement in catalytic performance and cost effectiveness. In this sense,
interdisciplinary efforts combining insights from physics, chemistry, materials
science and chemical engineering will be particularly important.
274 D. Raciti et al.

Fig. 6.18 (a) Synthetic scheme for the Pt3Ni nanoframe/C catalysts; (b, c) TEM analyses of the
seeding PtNi3 polyhedral NPs (b) and Pt3Ni nanoframes produced by aging in air (c); (d) element
mapping for the generated Pt3Ni nanoframes; (e, f) summary of the ORR catalytic performance for
the Pt3Ni nanoframe and nanoframe + ionic liquid composite catalysts (Reprinted with permission
from [113]. Copyright 2014 AAAS)

Acknowledgements The authors at JHU thank the start-up support from the Whiting School of
Engineering, Johns Hopkins University and funding support from NSF/DMR.

References

1. Vielstich W, Lamm A, Gasteiger HA (2003) Handbook of fuel cells: fundamentals, technol-


ogy, and applications. Wiley, Hoboken
2. Richter B, Goldston D, Crabtree G, Glicksman L, Goldstein D, Greene D, Kammen D,
Levine M, Lubell M, Savitz M, Sperling D, Schlachter F, Scofield J, Dawson J (2008) How
6 Recent Development of Platinum-Based Nanocatalysts for Oxygen Reduction. . . 275

America can look within to achieve energy security and reduce global warming. Rev Mod
Phys 80(4):S1–S107. doi:10.1103/RevModPhys.80.S1
3. Gasteiger HA, Kocha SS, Sompalli B, Wagner FT (2005) Activity benchmarks and require-
ments for Pt, Pt-alloy, and non-Pt oxygen reduction catalysts for PEMFCs. Appl Catal B
Environ 56(1–2):9–35
4. Kinoshita K (1990) Particle-size effects for oxygen reduction on highly dispersed platinum in
acid electrolytes. J Electrochem Soc 137(3):845–848
5. Valden M, Lai X, Goodman DW (1998) Onset of catalytic activity of gold clusters on titania
with the appearance of nonmetallic properties. Science 281(5383):1647–1650
6. Arenz M, Mayrhofer KJJ, Stamenkovic V, Blizanac BB, Tomoyuki T, Ross PN, Markovic
NM (2005) The effect of the particle size on the kinetics of CO electrooxidation on high
surface area Pt catalysts. J Am Chem Soc 127(18):6819–6829
7. Bezemer GL, Bitter JH, Kuipers HPCE, Oosterbeek H, Holewijn JE, Xu XD, Kapteijn F, van
Dillen AJ, de Jong KP (2006) Cobalt particle size effects in the Fischer-Tropsch reaction
studied with carbon nanofiber supported catalysts. J Am Chem Soc 128(12):3956–3964.
doi:10.1021/Ja058282w
8. Hvolbaek B, Janssens TVW, Clausen BS, Falsig H, Christensen CH, Norskov JK (2007)
Catalytic activity of Au nanoparticles. Nano Today 2(4):14–18. doi:10.1016/S1748-0132(07)
70113-5
9. Lei Y, Mehmood F, Lee S, Greeley J, Lee B, Seifert S, Winans RE, Elam JW, Meyer RJ,
Redfern PC, Teschner D, Schlogl R, Pellin MJ, Curtiss LA, Vajda S (2010) Increased silver
activity for direct propylene epoxidation via subnanometer size effects. Science 328
(5975):224–228. doi:10.1126/science.1185200
10. Shao MH, Peles A, Shoemaker K (2011) Electrocatalysis on platinum nanoparticles: particle
size effect on oxygen reduction reaction activity. Nano Lett 11(9):3714–3719. doi:10.1021/
Nl2017459
11. Yano H, Inukai J, Uchida H, Watanabe M, Babu PK, Kobayashi T, Chung JH, Oldfield E,
Wieckowski A (2006) Particle-size effect of nanoscale platinum catalysts in oxygen reduc-
tion reaction: an electrochemical and Pt-195 EC-NMR study. PCCP 8(42):4932–4939.
doi:10.1039/B610573d
12. Yamamoto K, Imaoka T, Chun WJ, Enoki O, Katoh H, Takenaga M, Sonoi A (2009) Size-
specific catalytic activity of platinum clusters enhances oxygen reduction reactions. Nat
Chem 1(5):397–402. doi:10.1038/Nchem.288
13. Nesselberger M, Ashton S, Meier JC, Katsounaros I, Mayrhofer KJJ, Arenz M (2011) The
particle size effect on the oxygen reduction reaction activity of pt catalysts: influence of
electrolyte and relation to single crystal models. J Am Chem Soc 133(43):17428–17433.
doi:10.1021/Ja207016u
14. Sheng WC, Chen S, Vescovo E, Shao-Horn Y (2012) Size influence on the oxygen reduction
reaction activity and instability of supported Pt nanoparticles. J Electrochem Soc 159(2):
B96–B103. doi:10.1149/2.009202jes
15. Shao-Horn Y, Sheng WC, Chen S, Ferreira PJ, Holby EF, Morgan D (2007) Instability of
supported platinum nanoparticles in low-temperature fuel cells. Top Catal 46(3–4):285–305.
doi:10.1007/s11244-007-9000-0
16. Tang L, Han B, Persson K, Friesen C, He T, Sieradzki K, Ceder G (2010) Electrochemical
stability of nanometer-scale Pt particles in acidic environments. J Am Chem Soc 132
(2):596–600
17. Markovic NM, Ross PN (2002) Surface science studies of model fuel cell electrocatalysts.
Surf Sci Rep 45(4–6):121–229
18. Stamenkovic VR, Fowler B, Mun BS, Wang GF, Ross PN, Lucas CA, Markovic NM (2007)
Improved oxygen reduction activity on Pt3Ni(111) via increased surface site availability.
Science 315(5811):493–497
276 D. Raciti et al.

19. Wang C, Daimon H, Onodera T, Koda T, Sun SH (2008) A general approach to the size- and
shape-controlled synthesis of platinum nanoparticles and their catalytic reduction of oxygen.
Angew Chem Int Ed 47(19):3588–3591. doi:10.1002/anie.200800073
20. Chen JY, Lim B, Lee EP, Xia YN (2009) Shape-controlled synthesis of platinum nanocrystals
for catalytic and electrocatalytic applications. Nano Today 4(1):81–95. doi:10.1016/j.nantod.
2008.09.002
21. Sanchez-Sanchez CM, Solla-Gullon J, Vidal-Iglesias FJ, Aldaz A, Montiel V, Herrero E
(2010) Imaging structure sensitive catalysis on different shape-controlled platinum
nanoparticles. J Am Chem Soc 132(16):5622–5624. doi:10.1021/Ja100922h
22. Ahmadi TS, Wang ZL, Green TC, Henglein A, ElSayed MA (1996) Shape-controlled
synthesis of colloidal platinum nanoparticles. Science 272(5270):1924–1926
23. Song H, Kim F, Connor S, Somorjai GA, Yang PD (2005) Pt nanocrystals: shape control and
Langmuir-Blodgett monolayer formation. J Phys Chem B 109(1):188–193. doi:10.1021/
Jp0464775
24. Ren JT, Tilley RD (2007) Preparation, self-assembly, and mechanistic study of highly
monodispersed nanocubes. J Am Chem Soc 129(11):3287–3291. doi:10.1021/Ja067636w
25. Wang C, Daimon H, Lee Y, Kim J, Sun S (2007) Synthesis of monodisperse Pt nanocubes and
their enhanced catalysis for oxygen reduction. J Am Chem Soc 129(22):6974–6975
26. Sun SH, Jaouen F, Dodelet JP (2008) Controlled growth of Pt nanowires on carbon
nanospheres and their enhanced performance as electrocatalysts in PEM fuel cells. Adv
Mater 20(20):3900–3904. doi:10.1002/adma.200800491
27. Alia SM, Zhang G, Kisailus D, Li DS, Gu S, Jensen K, Yan YS (2010) Porous platinum
nanotubes for oxygen reduction and methanol oxidation reactions. Adv Funct Mater 20
(21):3742–3746. doi:10.1002/adfm.201001035
28. Wang JX, Ma C, Choi YM, Su D, Zhu YM, Liu P, Si R, Vukmirovic MB, Zhang Y, Adzic RR
(2011) Kirkendall effect and lattice contraction in nanocatalysts: a new strategy to enhance
sustainable activity. J Am Chem Soc 133(34):13551–13557. doi:10.1021/Ja204518x
29. Kibsgaard J, Gorlin Y, Chen ZB, Jaramillo TF (2012) Meso-Structured Platinum Thin Films:
Active and Stable Electrocatalysts for the Oxygen Reduction Reaction. J Am Chem Soc
134 (18):7758-7765. doi:Doi 10.1021/Ja2120162
30. Toda T, Igarashi H, Uchida H, Watanabe M (1999) Enhancement of the electroreduction of
oxygen on Pt alloys with Fe, Ni, and Co. J Electrochem Soc 146(10):3750–3756
31. Stamenkovic V, Mun BS, Mayrhofer KJJ, Ross PN, Markovic NM, Rossmeisl J, Greeley J,
Norskov JK (2006) Changing the activity of electrocatalysts for oxygen reduction by tuning
the surface electronic structure. Angew Chem Int Ed 45(18):2897–2901
32. Stamenkovic VR, Mun BS, Arenz M, Mayrhofer KJJ, Lucas CA, Wang GF, Ross PN,
Markovic NM (2007) Trends in electrocatalysis on extended and nanoscale Pt-bimetallic
alloy surfaces. Nat Mater 6(3):241–247
33. Greeley J, Stephens IEL, Bondarenko AS, Johansson TP, Hansen HA, Jaramillo TF,
Rossmeisl J, Chorkendorff I, Norskov JK (2009) Alloys of platinum and early transition
metals as oxygen reduction electrocatalysts. Nat Chem 1(7):552–556. doi:10.1038/Nchem.
367
34. Peng ZM, Yang H (2009) Designer platinum nanoparticles: control of shape, composition in
alloy, nanostructure and electrocatalytic property. Nano Today 4(2):143–164. doi:10.1016/j.
nantod.2008.10.010
35. Wang C, Markovic NM, Stamenkovic VR (2012) Advanced platinum alloy electrocatalysts
for the oxygen reduction reaction. ACS Catal 2(5):891–898
36. Mayrhofer KJJ, Hartl K, Juhart V, Arenz M (2009) Degradation of carbon-supported Pt
bimetallic nanoparticles by surface segregation. J Am Chem Soc 131(45):16348–16349
37. Ponec V, Bond GC (1995) Catalysis by metals and alloys, vol 95, Studies in surface science
and catalysis. Elsevier, Amsterdam
38. Ertl G, Kn€ozinger H, Weitkamp J (1997) Handbook of heterogeneous catalysis. Weinheim,
VCH
6 Recent Development of Platinum-Based Nanocatalysts for Oxygen Reduction. . . 277

39. Murray CB, Kagan CR, Bawendi MG (2000) Synthesis and characterization of monodisperse
nanocrystals and close-packed nanocrystal assemblies. Annu Rev Mater Sci 30:545–610
40. Cushing BL, Kolesnichenko VL, O’Connor CJ (2004) Recent advances in the liquid-phase
syntheses of inorganic nanoparticles. Chem Rev 104(9):3893–3946
41. Ferrando R, Jellinek J, Johnston RL (2008) Nanoalloys: from theory to applications of alloy
clusters and nanoparticles. Chem Rev 108(3):845–910
42. Sun SH, Murray CB, Weller D, Folks L, Moser A (2000) Monodisperse FePt nanoparticles
and ferromagnetic FePt nanocrystal superlattices. Science 287(5460):1989–1992
43. Shevchenko EV, Talapin DV, Rogach AL, Kornowski A, Haase M, Weller H (2002)
Colloidal synthesis and self-assembly of COPt3 nanocrystals. J Am Chem Soc 124
(38):11480–11485
44. Wang C, Chi MF, Li DG, van der Vliet D, Wang GF, Lin QY, Mitchell JF, More KL,
Markovic NM, Stamenkovic VR (2011) Synthesis of homogeneous Pt-bimetallic
nanoparticles as highly efficient electrocatalysts. ACS Catal 1(10):1355–1359. doi:10.1021/
Cs200328z
45. Wang C, van der Vilet D, Chang KC, You HD, Strmcnik D, Schlueter JA, Markovic NM,
Stamenkovic VR (2009) Monodisperse Pt(3)Co nanoparticles as a catalyst for the oxygen
reduction reaction: size-dependent activity. J Phys Chem C 113(45):19365–19368
46. Ahrenstorf K, Albrecht O, Heller H, Kornowski A, Gorlitz D, Weller H (2007) Colloidal
synthesis of Ni(x)Pt(1-x) nanoparticles with tuneable composition and size. Small 3
(2):271–274
47. Ahrenstorf K, Heller H, Kornowski A, Broekaert JAC, Weller H (2008) Nucleation and
growth mechanism of Ni(x)Pt(1-x) nanoparticles. Adv Funct Mater 18(23):3850–3856
48. Wang C, Chi MF, Wang GF, van der Vliet D, Li DG, More K, Wang HH, Schlueter JA,
Markovic NM, Stamenkovic VR (2011) Correlation between surface chemistry and
electrocatalytic properties of monodisperse Pt(x)Ni(1-x) nanoparticles. Adv Funct Mater 21
(1):147–152
49. Liu ZF, Shamsuzzoha M, Ada ET, Reichert WM, Nikles DE (2007) Synthesis and activation
of Pt nanoparticles with controlled size for fuel cell electrocatalysts. J Power Sources 164
(2):472–480
50. Liu ZF, Ada ET, Shamsuzzoha M, Thompson GB, Nikles DE (2006) Synthesis and activation
of PtRu alloyed nanoparticles with controlled size and composition. Chem Mater 18
(20):4946–4951
51. Lee YH, Lee G, Shim JH, Hwang S, Kwak J, Lee K, Song H, Park JT (2006) Monodisperse
PtRu nanoalloy on carbon as a high-performance DMFC catalyst. Chem Mater 18
(18):4209–4211
52. Chen W, Kim JM, Sun SH, Chen SW (2008) Electrocatalytic reduction of oxygen by FePt
alloy nanoparticles. J Phys Chem C 112(10):3891–3898. doi:10.1021/Jp7110204
53. Mayrhofer KJJ, Blizanac BB, Arenz M, Stamenkovic VR, Ross PN, Markovic NM (2005)
The impact of geometric and surface electronic properties of Pt-catalysts on the particle size
effect in electocatalysis. J Phys Chem B 109(30):14433–14440
54. Zhang J, Yang HZ, Fang JY, Zou SZ (2010) Synthesis and oxygen reduction activity of
shape-controlled Pt(3)Ni nanopolyhedra. Nano Lett 10(2):638–644
55. Wu JB, Gross A, Yang H (2011) Shape and composition-controlled platinum alloy
nanocrystals using carbon monoxide as reducing agent. Nano Lett 11(2):798–802
56. Choi SI, Xie SF, Shao MH, Odell JH, Lu N, Peng HC, Protsailo L, Guerrero S, Park JH, Xia
XH, Wang JG, Kim MJ, Xia YN (2013) Synthesis and characterization of 9 nm Pt-Ni
octahedra with a record high activity of 3.3 A/mg(Pt) for the oxygen reduction reaction.
Nano Lett 13(7):3420–3425
57. Wu JB, Qi L, You HJ, Gross A, Li J, Yang H (2012) Icosahedral platinum alloy nanocrystals
with enhanced electrocatalytic activities. J Am Chem Soc 134(29):11880–11883. doi:10.
1021/Ja303950v
278 D. Raciti et al.

58. Wang C, Wang GF, van der Vliet D, Chang KC, Markovic NM, Stamenkovic VR (2010)
Monodisperse Pt3Co nanoparticles as electrocatalyst: the effects of particle size and
pretreatment on electrocatalytic reduction of oxygen. PCCP 12(26):6933–6939
59. Li DG, Wang C, Tripkovic D, Sun SH, Markovic NM, Stamenkovic VR (2012) Surfactant
removal for colloidal nanoparticles from solution synthesis: the effect on catalytic perfor-
mance. ACS Catal 2(7):1358–1362
60. Zhang J, Yang HZ, Fang JY, Zou SZ (2010) Synthesis and oxygen reduction activity of
shape-controlled Pt3Ni nanopolyhedra. Nano Lett 10(2):638–644. doi:10.1021/Nl903717z
61. Stamenkovic VR, Mun BS, Mayrhofer KJJ, Ross PN, Markovic NM (2006) Effect of surface
composition on electronic structure, stability, and electrocatalytic properties of Pt-transition
metal alloys: Pt-skin versus Pt-skeleton surfaces. J Am Chem Soc 128(27):8813–8819
62. Watanabe M, Tsurumi K, Mizukami T, Nakamura T, Stonehart P (1994) Activity and
stability of ordered and disordered Co-Pt alloys for phosphoric-acid fuel-cells. J Electrochem
Soc 141(10):2659–2668
63. Ball S, Hudson S, Theobald B, Thompsett D (2006) Enhanced stability of PtCo catalysts for
PEMFC. ECS Trans 1(8):141–152
64. Ball S, Hudson S, Theobald B, Thompsett D (2007) PtCo, a durable catalyst for automotive
proton electrolyte membrane fuel cells? ECS Trans 11(1):1267–1278
65. Mani P, Srivastava R, Strasser P (2008) Dealloyed Pt-Cu core-shell nanoparticle
electrocatalysts for use in PEM fuel cell cathodes. J Phys Chem C 112(7):2770–2778
66. Chen S, Ferreira PJ, Sheng WC, Yabuuchi N, Allard LF, Shao-Horn Y (2008) Enhanced
activity for oxygen reduction reaction on “Pt(3)CO” nanoparticles: direct evidence of per-
colated and sandwich-segregation structures. J Am Chem Soc 130(42):13818–13819
67. Chen S, Sheng WC, Yabuuchi N, Ferreira PJ, Allard LF, Shao-Horn Y (2009) Origin of
oxygen reduction reaction activity on “Pt(3)Co” nanoparticles: atomically resolved chemical
compositions and structures. J Phys Chem C 113(3):1109–1125
68. Chen S, Gasteiger HA, Hayakawa K, Tada T, Shao-Horn Y (2010) Platinum-alloy cathode
catalyst degradation in proton exchange membrane fuel cells: nanometer-scale compositional
and morphological changes. J Electrochem Soc 157(1):A82–A97
69. Gan L, Heggen M, Rudi S, Strasser P (2012) Core-shell compositional fine structures of
dealloyed PtxNi1-x nanoparticles and their impact on oxygen reduction catalysis. Nano Lett
12(10):5423–5430. doi:10.1021/Nl302995z
70. Snyder J, McCue I, Livi K, Erlebacher J (2012) Structure/processing/properties relationships
in nanoporous nanoparticles as applied to catalysis of the cathodic oxygen reduction reaction.
J Am Chem Soc 134(20):8633–8645. doi:10.1021/Ja3019498
71. Strasser P, Koh S, Anniyev T, Greeley J, More K, Yu CF, Liu ZC, Kaya S, Nordlund D,
Ogasawara H, Toney MF, Nilsson A (2010) Lattice-strain control of the activity in dealloyed
core-shell fuel cell catalysts. Nat Chem 2(6):454–460. doi:10.1038/Nchem.623
72. Koh S, Strasser P (2007) Electrocatalysis on bimetallic surfaces: modifying catalytic reac-
tivity for oxygen reduction by voltammetric surface dealloying. J Am Chem Soc 129
(42):12624–12625
73. Strasser P, Koha S, Greeley J (2008) Voltammetric surface dealloying of Pt bimetallic
nanoparticles: an experimental and DFT computational analysis. PCCP 10(25):3670–3683.
doi:10.1039/B803717e
74. Cui CH, Gan L, Heggen M, Rudi S, Strasser P (2013) Compositional segregation in shaped Pt
alloy nanoparticles and their structural behaviour during electrocatalysis. Nat Mater 12
(8):765–771
75. Ferreira PJ, La O’ GJ, Shao-Horn Y, Morgan D, Makharia R, Kocha S, Gasteiger HA (2005)
Instability of Pt/C electrocatalysts in proton exchange membrane fuel cells—a mechanistic
investigation. J Electrochem Soc 152(11):A2256–A2271
76. Wang C, Chi MF, Li DG, Strmcnik D, van der Vliett D, Wang GF, Komanicky V, Chang KC,
Paulikas AP, Tripkovic D, Pearson J, More KL, Markovic NM, Stamenkovic VR (2011)
6 Recent Development of Platinum-Based Nanocatalysts for Oxygen Reduction. . . 279

Design and synthesis of bimetallic electrocatalyst with multilayered Pt-skin surfaces. J Am


Chem Soc 133(36):14396–14403
77. Wang C, van der Vliet D, More KL, Zaluzec NJ, Peng S, Sun SH, Daimon H, Wang GF,
Greeley J, Pearson J, Paulikas AP, Karapetrov G, Strmcnik D, Markovic NM, Stamenkovic
VR (2011) Multimetallic Au/FePt(3) nanoparticles as highly durable electrocatalyst. Nano
Lett 11(3):919–926
78. Wagner FT, Gasteiger HA, Makharia R, Neyerlin KC, Thompson EL, Yan SG (2006) ECS
Trans 3(1):19
79. Yu P, Pemberton M, Plasse P (2005) PtCo/C cathode catalyst for improved durability in
PEMFCs. J Power Sources 144(1):11–20
80. Neyerlin KC, Srivastava R, Yu CF, Strasser P (2009) Electrochemical activity and stability of
dealloyed Pt-Cu and Pt-Cu-Co electrocatalysts for the oxygen reduction reaction (ORR). J
Power Sources 186(2):261–267. doi:10.1016/j.jpowsour.2008.10.062
81. Hwang BJ, Kumar SMS, Chen CH, Monalisa CMY, Liu DG, Lee JF (2007) An investigation
of structure-catalytic activity relationship for Pt-Co/C bimetallic nanoparticles toward the
oxygen reduction reaction. J Phys Chem C 111(42):15267–15276
82. Yano H, Kataoka M, Yamashita H, Uchida H, Watanabe M (2007) Oxygen reduction activity
of carbon-supported Pt-M (M ¼ V, Ni, Cr, Co, and Fe) alloys prepared by nanocapsule
method. Langmuir 23(11):6438–6445
83. Malheiro AR, Perez J, Santiago EI, Villullas HM (2010) The extent on the nanoscale of
Pt-skin effects on oxygen reduction and its influence on fuel cell power. J Phys Chem C 114
(47):20267–20271
84. Ingram DB, Christopher P, Bauer JL, Linic S (2011) Predictive model for the design of
plasmonic metal/semiconductor composite photocatalysts. ACS Catal 1(10):1441–1447
85. Mu RT, Fu QA, Xu H, Zhang HI, Huang YY, Jiang Z, Zhang SO, Tan DL, Bao XH (2011)
Synergetic effect of surface and subsurface Ni species at Pt-Ni bimetallic catalysts for CO
oxidation. J Am Chem Soc 133(6):1978–1986
86. Wang JX, Inada H, Wu LJ, Zhu YM, Choi YM, Liu P, Zhou WP, Adzic RR (2009) Oxygen
reduction on well-defined core-shell nanocatalysts: particle size, facet, and Pt shell thickness
effects. J Am Chem Soc 131(47):17298–17302. doi:10.1021/Ja9067645
87. Mazumder V, Chi MF, More KL, Sun SH (2010) Core/shell Pd/FePt nanoparticles as an
active and durable catalyst for the oxygen reduction reaction. J Am Chem Soc 132
(23):7848–7849. doi:10.1021/Ja1024436
88. Chen YM, Liang ZX, Yang F, Liu YW, Chen SL (2011) Ni-Pt core-shell nanoparticles as
oxygen reduction electrocatalysts: effect of Pt shell coverage. J Phys Chem C 115
(49):24073–24079. doi:10.1021/Jp207828n
89. Kuttiyiel KA, Sasaki K, Choi YM, Su D, Liu P, Adzic RR (2012) Nitride stabilized PtNi core-
shell nanocatalyst for high oxygen reduction activity. Nano Lett 12(12):6266–6271. doi:10.
1021/Nl303362s
90. Beard KD, Borrelli D, Cramer AM, Blom D, Van Zee JW, Monnier JR (2009) Preparation
and structural analysis of carbon-supported Co core/Pt shell electrocatalysts using electroless
deposition methods. ACS Nano 3(9):2841–2853. doi:10.1021/Nn900214g
91. Wang C, van der Vliet D, More KL, Zaluzec NJ, Peng S, Sun SH, Daimon H, Wang GF,
Greeley J, Pearson J, Paulikas AP, Karapetrov G, Strmcnik D, Markovic NM, Stamenkovic
VR (2011) Multimetallic Au/FePt3 nanoparticles as highly durable electrocatalyst. Nano Lett
11(3):919–926. doi:10.1021/Nl102369k
92. Yang H (2011) Platinum-based electrocatalysts with core-shell nanostructures. Angew Chem
Int Ed 50(12):2674–2676. doi:10.1002/anie.201005868
93. Zhang J, Sasaki K, Sutter E, Adzic RR (2007) Stabilization of platinum oxygen-reduction
electrocatalysts using gold clusters. Science 315(5809):220–222
94. Wang C, Daimon H, Sun SH (2009) Dumbbell-like Pt-Fe3O4 nanoparticles and their
enhanced catalysis for oxygen reduction reaction. Nano Lett 9(4):1493–1496. doi:10.1021/
Nl8034724
280 D. Raciti et al.

95. Alia SM, Jensen KO, Pivovar BS, Yan YS (2012) platinum-coated palladium nanotubes as
oxygen reduction reaction electrocatalysts. ACS Catal 2(5):858–863. doi:10.1021/
Cs200682c
96. van der Vliet DF, Wang C, Tripkovic D, Strmcnik D, Zhang XF, Debe MK, Atanasoski RT,
Markovic NM, Stamenkovic VR (2012) Mesostructured thin films as electrocatalysts with
tunable composition and surface morphology. Nat Mater 11(12):1051–1058
97. Snyder J, Fujita T, Chen MW, Erlebacher J (2010) Oxygen reduction in nanoporous metal-
ionic liquid composite electrocatalysts. Nat Mater 9(11):904–907. doi:10.1038/Nmat2878
98. Zhang J, Mo Y, Vukmirovic MB, Klie R, Sasaki K, Adzic RR (2004) Platinum monolayer
electrocatalysts for O-2 reduction: Pt monolayer on Pd(111) and on carbon-supported Pd
nanoparticles. J Phys Chem B 108(30):10955–10964
99. Sasaki K, Naohara H, Cai Y, Choi YM, Liu P, Vukmirovic MB, Wang JX, Adzic RR (2010)
Core-protected platinum monolayer shell high-stability electrocatalysts for fuel-cell cath-
odes. Angew Chem Int Ed 49(46):8602–8607. doi:10.1002/anie.201004287
100. Adzic RR, Zhang J, Sasaki K, Vukmirovic MB, Shao M, Wang JX, Nilekar AU,
Mavrikakis M, Valerio JA, Uribe F (2007) Platinum monolayer fuel cell electrocatalysts.
Top Catal 46(3–4):249–262. doi:10.1007/s11244-007-9003-x
101. Sasaki K, Mo Y, Wang JX, Balasubramanian M, Uribe F, McBreen J, Adzic RR (2003) Pt
submonolayers on metal nanoparticles—novel electrocatalysts for H-2 oxidation and O-2
reduction. Electrochim Acta 48(25–26):3841–3849. doi:10.1016/S0013-4686(03)00518-8
102. Shao M, Sasaki K, Marinkovic NS, Zhang L, Adzic RR (2007) Synthesis and characterization
of platinum monolayer oxygen-reduction electrocatalysts with Co-Pd core-shell nanoparticle
supports. Electrochem Commun 9(12):2848–2853
103. Debe MK, Schmoeckel AK, Vernstrorn GD, Atanasoski R (2006) High voltage stability of
nanostructured thin film catalysts for PEM fuel cells. J Power Sources 161(2):1002–1011.
doi:10.1016/j.jpowsour.2006.05.033
104. Debe MK (2012) Electrocatalyst approaches and challenges for automotive fuel cells. Nature
486(7401):43–51. doi:10.1038/Nature11115
105. Marcilly C (2003) Present status and future trends in catalysis for refining and petrochemi-
cals. J Catal 216(1–2):47–62. doi:10.1016/S0021-9517(02)00129-X
106. Vermeiren W, Gilson JP (2009) Impact of zeolites on the petroleum and petrochemical
industry. Top Catal 52(9):1131–1161. doi:10.1007/s11244-009-9271-8
107. Armor JN (2011) A history of industrial catalysis. Catal Today 163(1):3–9. doi:10.1016/j.
cattod.2009.11.019
108. Fajin JLC, Cordeiro MNDS, Gomes JRB (2011) On the theoretical understanding of the
unexpected O-2 activation by nanoporous gold. Chem Commun 47(29):8403–8405. doi:10.
1039/C1cc12166a
109. Wittstock A, Zielasek V, Biener J, Friend CM, Baumer M (2010) Nanoporous gold catalysts
for selective gas-phase oxidative coupling of methanol at low temperature. Science 327
(5963):319–322. doi:10.1126/science.1183591
110. Zielasek V, Jurgens B, Schulz C, Biener J, Biener MM, Hamza AV, Baumer M (2006) Gold
catalysts: nanoporous gold foams. Angew Chem Int Ed 45(48):8241–8244. doi:10.1002/anie.
200602484
111. Zeis R, Lei T, Sieradzki K, Snyder J, Erlebacher J (2008) Catalytic reduction of oxygen and
hydrogen peroxide by nanoporous gold. J Catal 253(1):132–138. doi:10.1016/j.jcat.2007.10.
017
112. Yan M, Jin T, Ishikawa Y, Minato T, Fujita T, Chen LY, Bao M, Asao N, Chen MW,
Yamamoto Y (2012) nanoporous gold catalyst for highly selective semihydrogenation of
alkynes: remarkable effect of amine additives. J Am Chem Soc 134(42):17536–17542.
doi:10.1021/Ja3087592
113. Chen C, Kang YJ, Huo ZY, Zhu ZW, Huang WY, Xin HLL, Snyder JD, Li DG, Herron JA,
Mavrikakis M, Chi MF, More KL, Li YD, Markovic NM, Somorjai GA, Yang PD,
Stamenkovic VR (2014) Highly crystalline multimetallic nanoframes with three-dimensional
electrocatalytic surfaces. Science 343(6177):1339–1343
Chapter 7
Enhanced Electrocatalytic Activity
of Nanoparticle Catalysts in Oxygen
Reduction by Interfacial Engineering

Christopher P. Deming, Peiguang Hu, Ke Liu, and Shaowei Chen

7.1 Introduction

As fuel cells are emerging as an optimal candidate for a renewable, environmentally


friendly energy source, the challenge of large scale commercialization remains at
the forefront of energy research [1–3]. It is well-known that the sluggish kinetics of
electrochemical oxygen reduction is a significant limitation to the efficiency of fuel
cells and therefore a catalyst must be utilized [4–6]. Platinum has been recognized
as the best single metal catalyst for oxygen reduction due to the stability under
the harsh operating conditions of the cell as well as a high intrinsic activity, but
must be used in excess of 50 g for a 100 kW vehicle given the state of the art
PEMFC (polymer electrolyte membrane fuel cell) [3]. This requirement signifi-
cantly limits commercial production given the yearly global output of Pt and the
yearly demand for motor vehicles. Experts predict that the benchmark activity of
160 A/g and 200 μA/cm2 at þ0.90 V must be enhanced by 4–10 folds for this
technology to emerge in the market as a cost efficient alternative to internal
combustion engine vehicles. To improve the intrinsic nature of a metal towards
oxygen reduction, it is beneficial to examine the individual steps that comprise the
complete reduction.
Many ORR pathways have been proposed [7], and in general, each pathway
involves adsorption of oxygen on the catalyst surface followed by electron transfer/
proton addition, ending with the release of water. Depending on the metal surface
the rate limiting step will either be the initial electron transfer or the final release of
product.

C.P. Deming • P. Hu • K. Liu • S. Chen (*)


Department of Chemistry and Biochemistry, University of California, 1156 High Street, Santa
Cruz, CA 95064, USA
e-mail: [email protected]

© Springer International Publishing Switzerland 2016 281


K.I. Ozoemena, S. Chen (eds.), Nanomaterials for Fuel Cell Catalysis,
Nanostructure Science and Technology, DOI 10.1007/978-3-319-29930-3_7
282 C.P. Deming et al.

It has been shown experimentally and computationally that the strength of


interactions between different oxygenated intermediates (OIs) and a catalyst sur-
face cannot be independently optimized. This idea is rationalized by the fact that all
intermediates bind to the catalyst surface through the oxygen atom and that stronger
or weaker surface interactions for one intermediate will yield the same result for all
other surface bound intermediates. Therefore, a surface that binds too strongly to
oxygen will undergo facile initial adsorption (initial step) of oxygen but will not
easily release the final product (final step) resulting in poisoning of active sites and
poor activity. Metals that bind to oxygenated species too weakly will allow for the
rapid release of the final product but will also result in an unfavorable initial step of
oxygen adsorption and consequently will exhibit poor activity. Therefore, the best
catalyst will exhibit moderate adsorptive interactions with OIs to maximize the rate
of the limiting step. The benchmark activity of 160 A/g and 200 μA/cm2 at þ0.90 V
measured for carbon supported platinum nanoparticles is therefore attributed to the
moderate binding to OIs.
Although platinum displays the highest reported activity as a single metal
catalyst, theoretical predictions indicate platinum still binds to OIs too strongly.
This model, in fact, shows a volcano-shaped trend when activity is plotted vs metal/
adsorbate binding energy with the peak of the volcano exhibiting an activity over
20 times that of platinum. This trend was predicted by Sabatier over 100 years ago
[8] and is graphically represented in Fig. 7.1. Here the x axis represents the
difference in binding energy of a sample as compared to that of commercial

Fig. 7.1 Experimental and theoretical data relating electrocatalytic activity and oxygen binding
energy. Reprinted with permission from (Stephens, I. E.; Bondarenko, A. S.; Perez-Alonso, F. J.;
Calle-Vallejo, F.; Bech, L.; Johansson, T. P.; Jepsen, A. K.; Frydendal, R.; Knudsen, B. P.;
Rossmeisl, J.; Chorkendorff, I.: Tuning the activity of Pt(1 1 1) for oxygen electroreduction by
subsurface alloying. J Am Chem Soc 2011, 133, 5485–91). Copyright (2011) American Chemical
Society
7 Enhanced Electrocatalytic Activity of Nanoparticle Catalysts in Oxygen. . . 283

platinum while the y axis represents the enhancement in activity compared to that of
commercial Pt. The blue dashed line represents theoretical calculations while the
black dots represent experimental points that will be discussed in detail later.
Furthermore, to quantitatively describe the adsorptive behaviors of metals, a
model has been developed relating the d band center of a metal to adsorptive
properties. This model states that an upshift of the d band center relative to the
Fermi level will create empty anti-bonding states that result in stronger bonding to
adsorbates. The converse is also true that a downshift of the d band center will result
in weaker interactions with adsorbates. This model is particularly useful in
predicting how a catalyst surface will respond to structural modifications and
when coupled with the dynamics of OIs on the surface, giving a promising outline
for future catalysts to meet and/or exceed the theoretical maximum and afford the
efficiency required for large scale commercialization.
In order to improve the activity of cathode catalysts or reduce the usage amount
of Pt, a number of strategies have been proposed and explored, including manip-
ulation of size [9, 10], morphology [11–13], elemental composition [14, 15], and
surface atomic arrangement [16, 17]. Additionally, deliberate engineering of the
nanoparticle interface has proven an effective method to alter the adsorptive
properties of the particle and thus the electrocatalytic activity [14, 18–22].
Among the pathways of interfacial manipulation, alloying with a second metal
serves to alter the electronic structure of surface atoms by imparting lateral force on
the particle lattice [23] as well as driving electron rearrangements between neigh-
boring atoms with different d band filling and electronegativity [24, 25]. Both
situations will ultimately alter the d band structure of the surface metal and thus
the degree of filling for the metal-adsorbate antibonding state and strength of
adsorptive interaction. The d band model is well established and has allowed for
many predictions and rationalizations of enhanced alloy nanoparticles. Conse-
quently, a great amount of efforts have been devoted to alloy cathode catalysts,
which can be classified into Pt-based alloy catalysts [24–27] and non-Pt alloy
catalysts [15].
Additionally, the functionalization of metal nanoparticles with organic ligands
has afforded activity enhancement through direct manipulation of the electronic
structure of the metal [20, 21, 28–31]. These interactions, along with other opto/
electronic properties, are largely a result of the nature of the interfacial connection
between the metal and capping ligand. For example, metal-sulfur bonds have been
shown to be highly polarized bonds with strongly localized charge while metal-
carbon bonds exhibit more covalent characteristics [21]. Additionally, there are
many examples of π overlap between the ligand p orbitals and the metal d orbitals
which lead to an extensively conjugated particle-ligand system [32–35]. The dif-
ferent interfacial bonding exhibited by each class of ligands will impart a different
effect on the electronic structure of the metal and correspondingly affect the
dynamics of OI adsorption. As models for nanoparticle alloys have linked the d
band center shift to a particle’s characteristics such as composition and type of
alloy, the development of a strong model relating how each type of ligands will
modify a metal d band is essential for understanding the origin of activity
284 C.P. Deming et al.

enhancement from organic ligand functionalization and allow for the proper choice
of capping ligands for a starting material.
Further studies have demonstrated that surface functionalization of metal oxide
nanoparticles will similarly enhance the electrocatalytic activity, rendering them to
be viable catalysts for oxygen reduction [22, 36]. Note that transition metal oxides
are generally cheap and abundant and therefore represent a possible route for
commercially available fuel cell technology, but at present the activity is generally
considered subpar compared to that of platinum group metals [37, 38]. The con-
trolled enhancement of such materials is therefore essential for commercial reali-
zation of this technology, but it must follow a slightly different approach than with
zero-valence metal particles. Metal oxide electrocatalysts are generally tested in
alkaline solution due to high corrosion in acidic media, and therefore a different
mechanism is needed to describe the reduction process. This mechanism involves
the generation of surface OH groups, replacement of these groups with O2, stepwise
reduction through peroxide and oxide intermediates, and the regeneration of surface
OH [38, 39]. For this reduction pathway, the kinetics are limited by the strength of
interaction between the material surface and oxygen. Specifically, an interaction
too weak will limit the initial displacement of OH by O2 and an interaction too
strong would prevent facile regeneration of OH [38, 40]. Again, the optimal surface
would possess properties in between the two extremes. To reach this volcano peak,
many surface manipulation techniques have been performed including direct
manipulation of oxide vacancies [22], surface functionalization with chlorine
[36], and alteration of metal components [40].
In this chapter, we will highlight some recent progress in ORR electrocatalysis,
including Pt-based nanoparticles, non-Pt metal nanoparticles, organically capped
metal nanoparticles, as well as metal oxide-based nanocomposites, with a focus on
the impacts of interfacial engineering on the nanoparticle electrocatalytic activity.

7.2 Interfacial Engineering of Metal Nanoparticle


Catalysts: Metal Compositions

Metal alloying has been proved to be an effective way to realize surface engineering
in the improvement of ORR catalytic activity. It has been widely known that the
catalytic activity of Pt can be effectively improved by alloying with 3d metals
[18, 26, 27, 41, 42]. For instance, Wu et al. [42] prepared truncated-octahedral
Pt3Ni (t,o-Pt3Ni) crystals with dominant exposure of (1 1 1) facets, which showed
remarkably enhanced electrocatalytic activity towards ORR, and the activity
increased with increasing (1 1 1) surface fraction. As shown in the inset to
Fig. 7.2a, the electrochemical surface areas (ECSAs) obtained from the calculated
hydrogen adsorption/desorption charge in the cyclic voltammgrams (CV) were
33.8, 53.7, 62.4, and 65 m2/gPt for the 70, 90, 100 % ((1 1 1) surface fraction) t,o-
Pt3Ni, and commercial Pt (TKK, 50 wt% Pt, 3 nm diameter on Vulcan carbon)
7 Enhanced Electrocatalytic Activity of Nanoparticle Catalysts in Oxygen. . . 285

Fig. 7.2 (a) Polarization curves (at 1600 rpm) and CV curves (inset). (b) area-specific (mA/cm2
Pt), and (c) mass-specific (A/mgPt) ORR activities for the t,o-Pt3Ni and commercial Pt catalysts. (d)
Correlations of the area-specific and mass-specific activities with the fraction of the (1 1 1) surfaces
of these t,o-Pt3Ni catalysts. Reprinted (Adapted) with permission from (Wu, J. B.; Zhang, J. L.;
Peng, Z. M.; Yang, S. C.; Wagner, F. T.; Yang, H., Truncated Octahedral Pt3Ni Oxygen Reduction
Reaction Electrocatalysts. J Am Chem Soc 2010, 132 (14), 4984) Copyright (2010) American
Chemistry Society

catalysts, respectively. The electrocatalytic activities of the t,o-Pt3Ni and commer-


cial Pt catalysts were then evaluated by RDE (rotating disk electrode) voltammetry,
and all three t,o-Pt3Ni catalysts exhibited more positive onset potentials than
commercial Pt/C. Furthermore, as shown in Fig. 7.2b, c, within the potential
range of þ0.84 to þ0.94 V (vs. RHE), the area-specific and mass-specific current
densities were 850 μA/cm2 and 0.53 A/mgPt respectively at þ0.90 V for the 100 % t,
o-Pt3Ni catalysts, both 4-times higher than those (215 μA/cm2 and 0.14 A/mgPt) of
the commercial Pt/C catalysts. In addition, one can see that the higher (1 1 1)
surface fraction of t,o-Pt3Ni, the higher area-specific and mass-specific current
density, as manifested in Fig. 7.2d. In another study, Stamenkovic et al. [27]
demonstrated that the Pt3Ni(1 1 1) surface was 10-fold more active on ORR than
the corresponding Pt(1 1 1) surface and 90-fold more active than commercial Pt/C
catalysts for PEMFC, and further studies were also conducted to explore the
mechanism for the significantly improved activity. It was shown by low energy
ion scattering (LEIS) that the composition of the outermost atomic layer was pure
Pt, a so-called Pt-skin structure. Such first-layer Pt enrichment was suggested to be
286 C.P. Deming et al.

counterbalanced by its depletion in the next two to three layers so that the concen-
tration profile oscillates around the bulk value. They also showed in background-
corrected ultraviolet photoemission spectroscopy (UPS) measurements that such a
near-surface compositional change also resulted in distinctive electronic properties
for PtNi alloys. It is found that the d-band density of states (DOS) is structure-
sensitive, i.e., the d-band center shifts as the surface structure changes (2.70 eV on
Pt3Ni(1 1 0), 3.10 eV on Pt3Ni(1 1 1), and 3.14 eV on Pt3Ni(1 0 0)), which are
also quite different from that of Pt single crystals. Furthermore, the average energy
of the surface atoms influences the chemisorption energies in surface electrochem-
istry when the adsorbate binds onto the surface (the ligand effect). Experimentally,
on the Pt-skin surface of a Pt3Ni single crystal the formation potentials were found
to be negatively shifted by ca. 0.15 V for Hupd (which refers to the underpotentially
deposited hydrogen) and positively shifted by ca. 0.1 V for OHad (which refers to
the adsorbed hydroxyl layer), relative to those on the Pt(1 1 1), and correspondingly,
the fractional coverages of Hupd and OHad on Pt3Ni(1 1 1) were dramatically
reduced by 50 % relative to those on Pt(1 1 1). Both of the observations above
were consistent with the large downshift (0.34 eV) of the d-band center position on
the Pt-skin structure. ORR measurements showed that the ORR kinetics was
accelerated on the Pt3Ni(1 1 1) skin, compared with that on Pt(1 1 1), where one
can find a positive shift of 100 mV in the half-wave potential. The authors ascribed
such improved catalytic activity to the key parameter of reduced coverage of OHad.
Additionally, the lower coverage of Hupd substantially attenuated the production of
peroxide on the Pt-skin surface.
Further studies were also carried out by Stamenkovic et al. on other PtM
(M ¼ Co, Fe, Ti, and V) surfaces in order to investigate the fundamental relation-
ship in catalytic trends between the experimentally determined surface electronic
structure (the d-band center) and the activity for oxygen reduction reaction [26]. As
shown in Fig. 7.3, a volcano-shape relationship was found between the
electrocatalytic activity and the d-band center position on the Pt-skin of PtM
surfaces with the maximum activity from Pt3Co. Such a trend strongly indicates
that two opposite aspects, the strong binding energy of O2 and reaction intermedi-
ates like O 2
2 , O2 , H2O2, and the relatively low coverage by oxygenated species and
specifically adsorbed anions should be taken into consideration and
counterbalanced to obtain better catalysts than Pt. As a result, for metals or alloys
with too high binding energy to oxygen/oxides/anions, like Pt, the d-band center is
too close to the Fermi level and the availability of OHad/anion-free Pt sites is the
key parameter that limits the ORR kinetic rate. In contrast, in the cases like Pt3V
and Pt3Ti of which the d-band center is too far away from the Fermi level, although
the catalyst surface is less occupied by oxygen species and anions, the binding
between oxygen/intermediates and metal/alloy is too weak to enable a turnover rate
of ORR. Consequently, the volcano curve of the electrocatalytic activity can be
interpreted by the Sabatier principle: the reaction is limited by the rate of removing
surface oxides and anions on the surface of catalysts binding oxygen too strongly,
7 Enhanced Electrocatalytic Activity of Nanoparticle Catalysts in Oxygen. . . 287

Specific activity: ik at 0.9 v (mA cm-2real)

Activity enhancement (versus Pt poly)


5
Pt3Co
3
4
Pt3Fe
3 Pt3Ni
2
Pt3V
2
Pt3Ti
1
1 Pt poly

0
–3.4 –3.0 –2.6
d-band centre (ev)

Fig. 7.3 Relationships between the catalytic properties and electronic structure of Pt-skin on Pt3M
alloys. Reprinted by permission from Macmillan Publishers Ltd: Nat Mater (Stamenkovic VR,
Mun BS, Arenz M, Mayrhofer KJJ, Lucas CA, Wang GF, Ross PN, Markovic NM, Trends in
electrocatalysis on extended and nanoscale Pt-bimetallic alloy surfaces. Nat Mater 2007, 6,
241–247), copyright (2007)

but on the surface of catalysts binding oxygen too weakly it is limited by the rate of
electron and proton transfer to the adsorbed oxygen.
Besides ligand effects discussed above, geometrical strain is another important
effect in surface engineering by metal alloying [43]. As a matter of fact, ligands and
geometric effects generally simultaneously occur and impact the catalytic activity
at the same time [44, 45]. In order to understand how geometric strain effect
influence the catalytic activity in an electrocatalytic system, a core-shell catalyst
structure was designed and prepared by Strasser et al. In the structure, the shell
consists of a few atomic monometallic layers, which is supported on the core
substrate with different lattice parameters. Such a structure is supposed to be able
to isolate geometric strain effects since ligand effect on surface catalytic activity is
only limited within two or three atomic layers [23]. To prepare the special catalyst
mentioned above, precursor PtCu alloy nanoparticles were synthesized at first, and
the core-shell structure was obtained via preferential dissolution/removal of the
electrochemically more reactive component (Cu) from the bimetallic alloy, which
is often referred to as “dealloying”. It was then shown that dealloyed PtCu
nanoparticles exhibited uniquely high catalytic activity for ORR.
The ORR catalytic activity of a serial of PtCu dealloyed nanoparticles was both
calculated by density functional theory and measured experimentally to quantify
the relationship between lattice strain, oxygen binding energy, and electrocatalytic
activity, as shown in Fig. 7.4. It is obvious that lattice strain increases as Cu content
increases, and both computational analysis and X-ray spectroscopy data showed a
linear relationship between lattice strain and oxygen binding energy, which is also
288 C.P. Deming et al.

Fig. 7.4 Experimental and


predicted relationships
between electrocatalytic
ORR activity and lattice
strain. Reprinted by
permission from Macmillan
Publishers Ltd: Nat Chem
(Strasser P, Koh S,
Anniyev T, Greeley J,
More K, Yu CF, Liu ZC,
Kaya S, Nordlund D,
Ogasawara H, Toney MF,
Nilsson A, Lattice-strain
control of the activity in
dealloyed core–shell fuel
cell catalysts. Nat Chem
2010, 2, 454–460),
copyright (2010)

consistent with computational calculations in other reports [43]. Based on a


microkinetic model developed for ORR [46], the strain-binding energy relationship
was combined, and a volcano plot was derived for the relationship between the
predicted ORR rate and lattice strain. The volcano plot implies that lattice strain
first accelerates the ORR rate by weakening the bond between surface Pt atoms and
intermediate oxygenated adsorbates, lowering the energy barriers for electron and
proton transfer, and subsequently reducing the coverage of adsorbed oxygen spe-
cies. However, the ORR rate is predicted to decrease after a critical lattice strain,
which can be understood by the fact that the binding energy becomes too weak
between surface Pt atoms and oxygen, leading to an increased energy barrier for
oxygen dissociation and the following formation of oxygenated intermediates.
Experimentally, the ORR catalytic activity of PtCu dealloys, in consistence with
calculation analysis, was also found to be enhanced at first with the increased lattice
strain. However, unlike the prediction in computational studies, the ORR catalytic
activity did not decrease on the left side of the volcano plot, which was rationalized
by the compressive strain relaxation in the Pt shell. That is, Pt atoms adjacent to the
PtCu core adopt the lattice parameters close to that of the core, and in contrast, the
Pt atoms in shell generally relax towards the lattice constants of pure Pt. Therefore,
the real surface strain is less than the calculated value plotted in Fig. 7.4, which is an
average strain in the Pt shell. Thus, it is expected of a data point shift to the right if
the real surface strain is used in Fig. 7.4. Moreover, it is believed that due to the
surface strain relaxation, it is probably hard to obtain sufficient surface strain by
synthesizing dealloyed nanoparticles to achieve the true maximum of the volcano
curve.
Further studies were conducted by Greeley et al. [18] and Xiao et al. [47] to
quantitatively understand the surface structure-activity relationship for fuel cell
cathode catalysts. It turns out that although weakening the surface binding energy to
oxygen by alloying Pt with other metals is responsible for the improvement of the
7 Enhanced Electrocatalytic Activity of Nanoparticle Catalysts in Oxygen. . . 289

Fig. 7.5 (a) The volcano plot between catalytic activity and oxygen binding energy relative to that
of Pt(1 1 1); (b) the linear relationship between the relative lattice size and oxygen binding energy
relative to that of Pt(1 1 1). Reprinted by permission from Macmillan Publishers Ltd: Nat Chem,
(Greeley J, Stephens IEL, Bondarenko AS, Johansson TP, Hansen HA, Jaramillo TF, Rossmeisl J,
Chorkendorff I, Norskov JK (2009) Alloys of platinum and early transition metals as oxygen
reduction electrocatalysts. Nat Chem 1 (7):552–556), Copyright (2009). Reproduced from
(Xiao L, Huang B, Zhuang L, Lu JT (2011) Optimization strategy for fuel-cell catalysts based
on electronic effects. Rsc Adv 1, 1358–1363.) with permission of The Royal Society of Chemistry

electrocatalytic activity of Pt, the surface catalytic activity of PtNi(1 1 1) is not yet
optimal. As shown in Fig. 7.5a, both experimental and theoretical data strongly
suggest that a surface with better ORR catalytic activity than Pt should have a
binding energy to oxygen 0–0.4 eV weaker than that of Pt, with an optimal value at
the binding energy 0.2 eV lower than Pt [18]. According to Fig. 7.5b, it is found by
Xiao et al. [47] that the binding energy to oxygen of Pt decreases linearly as the Pt
lattice compresses. Taken together with the surface electronic structure-surface
reactivity relationship, it is clear of the guideline for catalyst optimization, which
is to control the ligand effect and geometric strain effect to achieve an alloy
structure with the oxygen binding energy around 0.2 eV lower than that of Pt,
and alloy here is not limited to Pt or any other precious metal/alloys, but also
includes non-precious metal/alloys.
In addition to achieving better activity than Pt by surface engineering through
metal alloying, another aspect that should be taken into consideration is the stability
of the surface alloy structures, which means not only the stability in thermodynam-
ics but also the stability under PEMFC working conditions. In order to search for
stable alloy catalysts, DFT calculations have also been conducted by Greeley
et al. with a number of alloys [18]. As shown in Fig. 7.6 (red points), it is clear
that Pt3Y and Pt3Sc turn out to be the most stable among the alloys with ΔEO in the
optimal range (0–0.4 eV). The high stability can be rationalized in terms of
approximately half filled metal-metal d bands, meaning filled bonding states and
empty anti-bonding states. The ORR activity of the catalysts was measured and
shown in Fig. 7.7. It is clear in Fig. 7.7a that both Pt3Y and Pt3Sc exhibit
290 C.P. Deming et al.

0.6 Pt3Nb Pt3V


Pt3Ti

Pd3Nb Pd3V Pt Mo
ΔEO–ΔEOPt (eV)

Pd3Ti
0.4 3
Pt3Fe
Pt3Sc Pd3Zn Pt3Cr
Pd3Sc Pt3Ni Pt3Co Pd3Cr
Pt3Zn Pt3Cu Pt3Mn
0.2
Pd3Cu Pd3Ni
Pd3Y Pd3Co
Pd3Cd
Pt3Y
Pt3Cd
0.0 Pt

-0.2

-1.2 -1.0 -0.8 -0.6 -0.4 -0.2 -0.0 0.2


ΔEalloy (eV)

Fig. 7.6 Output of computational screening procedure, showing the oxygen binding energy,
relative to that of Pt, on a Pt or Pd skin surface, as a function of alloying energy. ΔEalloys is the
calculated energy of formation of the indicated elements from the appropriate bulk elemental
precursors, specified on a per atom basis. For some systems we include results for the cases where
there are 50 % (circles) and 25 % (squares) of the alloying element in the second layer. Reprinted
by permission from Macmillan Publishers Ltd: Nat Chem (Greeley J, Stephens IEL, Bondarenko
AS, Johansson TP, Hansen HA, Jaramillo TF, Rossmeisl J, Chorkendorff I, Norskov JK. Alloys of
platinum and early transition metals as oxygen reduction electrocatalysts. Nat Chem 2009,
1, 552–556), Copyright (2009)

significantly enhanced ORR catalytic activity over Pt, which is consistent with the
prediction in calculations. Moreover, the alloy catalysts are also very stable under
continued cycling conditions, as shown in Fig. 7.7b.
To further reduce the costs of cathode catalysts in fuel cells, a large number of
efforts have also been devoted to non-Pt based alloy catalysts. Greeley et al. [48]
have conducted a density functional theory (DFT)-based combinatorial search for
improved ORR alloy catalysts. In their calculations, oxygen binding energy was
used as an ORR descriptor for the prediction of ORR activity trends, since for
Pt-based alloys the calculation using oxygen binding energy as a descriptor is in
excellent agreement with experimental measurements. According to their calcula-
tions, 85 of 736 transitional metal surface alloys are predicted to be at least as active
as pure Pt for ORR. However, among the 29 most active and stable alloy candidates
for ORR, only 8 (containing primarily Pt, Pd, Ag, or group VA elements in the
surface layer) will survive above þ0.5 V in acidic environment based on the
calculation. What is worse, even this level of stability may not be sufficient to
yield robust catalysts with long-term operation stability in fuel cells.
Among non-Pt based catalysts, Pd is considerably less expensive than Pt and
also less active than Pt for ORR catalysis, but more active and stable than any other
metals [15]. It has been reported that interfacial engineering by alloying can
significantly enhance the catalytic activity of Pd [14, 15, 49]. Therefore, Pd-based
7 Enhanced Electrocatalytic Activity of Nanoparticle Catalysts in Oxygen. . . 291

a 0

j (mA cm–2) –2
Pt3Y
–4 Pt3Sc

–6 Pt

0.0 0.2 0.4 0.6 0.8 1.0


U (V) (RHE)

b 0

–2
j (mA cm–2)

Pt3Y
–4

–6 Pt3Sc

0.0 0.2 0.4 0.6 0.8 1.0


U (V) (RHE)

Fig. 7.7 Anodic sweeps of cyclic voltammograms of Pt, Pt3Sc and Pt3Y in O2-saturated electro-
lyte. (a) The first stable sweep. (b) A comparison of the first steady cycle (solid lines) and after
90 min of cycling (dashed lines), for Pt3Y and Pt3Sc only, under the same conditions as (a).
Measurements were taken at 20 mV/s and 1600 rpm in 0.1 M HClO4 at 23  2  C. Reprinted by
permission from Macmillan Publishers Ltd: Nat Chem (Greeley J, Stephens IEL, Bondarenko AS,
Johansson TP, Hansen HA, Jaramillo TF, Rossmeisl J, Chorkendorff I, Norskov JK, Alloys of
platinum and early transition metals as oxygen reduction electrocatalysts. Nat Chem
2009, 1, 552–556), copyright (2009)

alloy catalysts have been considered as one of the best candidates for non-Pt ORR
catalysts. Suo et al. [50] carried out first-principles studies of the improved
electrocatalytic activity of Pd-based alloys toward ORR. It is found that there is a
volcano relationship (Fig. 7.8) between the degree of alloying (in the case of Pd-Co
alloy) and the catalytic activity, which is believed to arise from the contrary
influences from lattice strain effect and ligand effect by the interfacial engineering.
At a low degree of alloying with Co, the lattice strain effect is predominant because
of the negligible content of Co on the surface, which leads to weakened oxygen
binding energy and increased activity, while at a high degree of alloying, the ligand
effect becomes overwhelming as a result of the high surface content of Co, leading
to increased oxygen binding energy and reduced catalytic activity.
In addition, given that the lattice strain effect on the surface was considered as
the main reason for the enhanced activity, another significant calculation was also
conducted to find the relationship between the oxygen binding energy (denoted as
AE in Fig. 7.9) and the lattice strain on the surface described as the ratio of lattice
292 C.P. Deming et al.

Fig. 7.8 Relationship between surface specific activity (SA) of Pd–Co alloys and degree of
alloying. The kinetic current obtained from RDE experiments (25  C, 5 mV/s, O2-saturated
0.5 M H2SO4, 1600 rpm) was normalized by the electrochemical surface area of Pd–Co alloys
to give SA. Reprinted with permission from (Suo YG, Zhuang L, Lu JT: First-principles consid-
erations in the design of Pd-alloy catalysts for oxygen reduction. Angew Chem Int Edit 2007,
46, 2862–2864. Copyright © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Fig. 7.9 Relationship between the AE of Oads and the strain ratio in the lattice constant of Pd,
which may serve as a guideline for the design of Pd-alloy catalysts for ORR. Reprinted with
permission from (Suo YG, Zhuang L, Lu JT: First-principles considerations in the design of
Pd-alloy catalysts for oxygen reduction. Angew Chem Int Edit 2007, 46, 2862–2864. Copyright
© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

constant, as shown in Fig. 7.9. According to Fig. 7.9, the relationship is linear and
the reduced lattice ratio on the surface is supposed to benefit the catalytic activity. It
is also predicted that Fe would be a better alloy element than Co for interfacial
engineering of Pd for a better catalytic performance, and the activity of Pd11Fe alloy
7 Enhanced Electrocatalytic Activity of Nanoparticle Catalysts in Oxygen. . . 293

may be as comparable to that of Pt since the oxygen binding energy on it is very


close to that on Pt, based on the relationship. Such a prediction, as a matter of fact,
was confirmed by Shao et al. [15]. They prepared a serial of Pd-Fe alloys with
different atomic ratios and found a volcano relationship between the activity and
the atomic ratio. It was reported that different alloying degree of Fe resulted in
different Pd-Pd bond length, generating different lattice strain, further altering the
d-band center position of the alloy, which determined the surface activity of the
alloy. Moreover, the Pd3Fe was found to be very stable, methanol tolerated, and
exhibit even higher activity than Pt.
Ag and Ag-based catalysts have also been gaining increasing interests as a viable
candidate for non-Pt catalysts applied in fuel cells with alkaline electrolytes [51–
53]. Ag is relatively inexpensive as compared with Pd, Au, and more importantly,
although still inferior to Pt, Ag gives relatively high catalytic activity with a
4-electron ORR pathway [52, 53]. Nevertheless, the application of Ag as a
PEMFC catalyst is still limited by both poor activity and stability. Fortunately, in
recent years, remarkable progress has been made in alkaline polymer electrolytes
(APEs), facilitating the development of alkaline polymer electrolyte fuel cell
(APEFC), which is expected to combine the advantages of both PEMFC and
AFC (alkaline fuel cell) [54]. The development of APEFC thus enables the use of
non-Pt metal/alloy catalysts, which are not compatible with the acidic electrolyte in
PEMFC mainly due to the stability problem.
For instance, by simultaneous reduction of Ag and Pd precursors, Slanac
et al. [55] prepared Ag-rich AgPd alloy nanoparticles (~5 nm). The ORR mass
activity per total metal was found to be 60 % higher for the AgPd2 alloy than that of
pure Pd, and the mass activities of Ag9Pd (340 mA/mgmetal) and Ag4Pd (598 mA/
mgmetal) were 2.7 and 3.2 times higher, respectively, than those calculated by a
linear combination of mass activities of Ag (60 mA/mgAg) and Pd (799 mA/mgPd)
nanoparticles. The enhancement of the ORR activity was ascribed to interfacial
engineering, leading to the combined effects of weak binding energy of Ag and
strong binding energy of Pd towards oxygen/oxygen intermediates (ensemble
effect). That is, by creating a AgPd alloy surface, Pd atoms facilitate the adsorption
of oxygen, while Ag atoms desorb the oxygenated products (like OH). Addition-
ally, the electronic structure shifted by the ligand effect, which was confirmed by
XPS measurements, is also believed to contribute to the activity enhancement.

7.3 Metal Nanoparticle Catalysts: Ligand-Mediated


ORR Activity

In addition to metal compositions, another potent method to improve the perfor-


mance of a metal surface for ORR is through deliberate surface functionalization
with organic ligands. Self-assembled monolayers of organic ligands on metal
nanoparticles have been well studied [56, 57] and have been specifically explored
294 C.P. Deming et al.

for improved electroreduction of oxygen on metal surfaces in the form of thiols


[21, 58–60], para-substituted phenyls [20, 21, 28], cyanide ions [61], phenyl
acetylene derivatives [29], n-alkynes [21, 29], phenyl phosphine derivatives [62],
and graphene quantum dots (GQD) [30, 31]. The catalytic performance largely
depends on three factors: the chemical nature of the metal, the structures and
properties of the capping ligands, and the nature of the interfacial bonding interac-
tions between the metals and the ligands. Among these, the structure of the capping
ligands, such as chain length, degree of branching, and functional groups may alter
the nature of the catalytic process through steric hindrance while the nature of
interfacial bonds may affect the catalytic reactions by manipulation of the electron
density in the metal. The primary focus of this section will be to summarize the
effects of interfacial bonding on ORR, specifically, how interactions between the
ligand and metal may be exploited for the enhancement of the kinetics of oxygen
reduction by providing more favorable interactions with OIs.
Thiol ligands have been shown to improve the stability of platinum nanoparticle
catalysts in acidic solution. For instance, for platinum nanoparticles functionalized
by 4-mercapto aniline then further derivatized with 2-thiophencarbonyl chloride, it
was found that no activation process was needed for ligand desorption and the
as-prepared nanoparticles exhibited excellent stability after prolonged electrochem-
ical cycling as indicated by cyclic voltammetry and XPS measurements [58].
For platinum nanoparticles functionalized with para-substituted phenyl frag-
ments, the ORR activity was markedly enhanced, which was correlated with the
electron withdrawing capacity of the phenyl substituents with the most active
sample exhibiting 3 times the specific activity of commercial Pt/C catalysts
[20]. In this study, the para-substituents for the phenyl moiety included –CH3,
–F, –Cl, –OCF3, and –CF3 within a large range of Hammet coefficient (σ). The
ORR activity was evaluated with rotating ring disk electrode (RRDE) voltammetry
in oxygen-saturated 0.1 M HClO4 and it was found that specific activities were very
well correlated with the Hammet coefficient, as shown in Fig. 7.10. The trend in
activity was ascribed to the weakening interaction between platinum and OIs as the
electron density of Pt was diminished by the electron-withdrawing substituting
groups.
Similar effects have also been observed with other metal-ligand interfacial
bonding interactions. In a more recent study [21], a variety of silver nanoparticles
were prepared by surface functionalization with hexanethiol, 1-octyne, and
4-trifluoromethylphenyl fragments (denoted as AgSC6, AgHC8, and AgPhCF3,
respectively), and the impacts of the metal-ligand interfacial bonds were examined,
with silver nanoparticles supported on carbon (Ag/C) as the benchmark material.
The ORR activity was found to increase in the order AgSC6 < Ag/
C < AgHC8 < AgPhCF3. The most active sample was identified as AgPhCF3 that
exhibited a specific current density 13 times that of bare Ag and an onset potential
200 mV more positive. This was accounted for by the manipulation of the electronic
states of silver by the capping ligand, as confirmed by XPS measurements of the
binding energy of the Ag 3d electrons.
7 Enhanced Electrocatalytic Activity of Nanoparticle Catalysts in Oxygen. . . 295

ORR catalytic activity / mA cm-2


R CF3
R R
0.6
R Pt R OCF3
Cl
R R
0.4 R

F
Pt/C
0.2
CH3

0.0
-0.2 -0.1 0.0 0.1 0.2 0.3 0.4 0.5
σ
Fig. 7.10 Activity of Pt nanoparticles functionalized with different para substituted phenyl
ligands as a function of Hammet coefficient (electron withdrawing nature) Reprinted with per-
mission from (Zhou, Z.-Y.; Kang, X.; Song, Y.; Chen, S.: Ligand-Mediated Electrocatalytic
Activity of Pt Nanoparticles for Oxygen Reduction Reactions. J Phys Chem C 2012, 116,
10592–10598). Copyright (2012) American Chemical Society

Triphenylphosphine triphosphonate (TPPTP) capped platinum particles have


also been found to show superior activity as compared to pure Pt particles. The suc-
cessful synthesis of the nanoparticles was confirmed by 1H and 31P NMR spectros-
copy, XPS, TEM, and UV-Vis measurements [63]. Cyclic voltammetry
measurements in 0.1 M HClO4 revealed that the surface oxide reduction peak
shifted to a more positive potential for the TPPTP functionalized particles than
similarly prepared Pt particles without TPPTP functionalization. This suggests
weakened interactions between the metal surface and OIs, which is further con-
firmed by the larger specific activity of the TPPTP-capped particles than that of
uncapped particles. This study highlights the impacts of metal-ligand interfacial
bonds (in this case Pt-P) that will allow for unique electronic interactions and
ultimately result in enhanced electrocatalytic oxygen reduction.
In a recent study, bimetallic AgAu Janus nanoparticles were prepared by gal-
vanic replacement reactions of hexanethiolate-protected silver (AgC6)
nanoparticles with Au(I) mercapto-propanediol complex on one hemisphere of
the particle, leading to phase segregation of both metals and ligands (Fig. 7.11)
[64]. Metallic phase segregation was confirmed by XPS and energy-dispersive
X-ray spectroscopy (EDX, Fig. 7.12) measurements while ligand segregation was
confirmed by contact angle measurements. Electrochemical measurements showed
that the bimetallic AgAu Janus particles exhibited apparent ORR activity that was
seven times that of AgC6 particles and over 4 times that of bulk exchanged AgAu
particles. This enhancement is attributed to the segregation of polar and non-polar
ligands on the particle surface that enhanced the charge transfer from Ag to Au, thus
improving the adsorption of oxygen. Since gold is a metal with a low lying d band
center, it interacts weakly with OIs and is limited by the initial adsorption and
296 C.P. Deming et al.

(i) LB monolayer (ii) Interfacial exchange (iii) Janus nanoparticles

Fig. 7.11 Schematic of bimetallic AgAu Janus nanoparticle synthesis. Reprinted with permission
from (Song, Y.; Liu, K.; Chen, S.: AgAu bimetallic Janus nanoparticles and their electrocatalytic
activity for oxygen reduction in alkaline media. Langmuir 2012, 28, 17143–52). Copyright (2012)
American Chemical Society

Fig. 7.12 EDX measurements of (a) Janus exchanged particles and (b) bulk exchange particles.
Red represents silver domains while green represents gold domains. Reprinted with permission
from (Song, Y.; Liu, K.; Chen, S.: AgAu bimetallic Janus nanoparticles and their electrocatalytic
activity for oxygen reduction in alkaline media. Langmuir 2012, 28, 17143–52). Copyright (2012)
American Chemical Society

electron transfer step. Therefore, the charge transfer from silver to gold will raise
the d band center of gold and result in stronger binding to OIs and is consistent with
the previously stated d band model [25, 65].
Another area gaining much attention is the functionalization of metal
nanoparticles with high surface area supports such as carbon nanotubes [66] and
graphene quantum dots (GQDs) [30, 67]. Research has shown that the function of
such supports can extend beyond merely a means to disperse the catalysts and
connect them to the electrode, but may also serve to enhance the activity
through electronic interactions that result in more favorable binding with OIs
7 Enhanced Electrocatalytic Activity of Nanoparticle Catalysts in Oxygen. . . 297

[67–70]. Furthermore, recent DFT calculations have shown that the defects in the
graphitic structure of GQDs may diminish electron density of the deposited plati-
num nanoparticles leading to enhanced activity through weakened adsorption of
OIs [7]. As carbon black or activated carbon is commonly used for supports, a
Pt/GQD composite is a promising route for improving catalytic activity, for GQDs
will offer the same availability, stability, and high surface to volume ratio as the
common carbon supports in addition to activity enhancement.
In fact, in a recent study focused on the defect-mediated activity of GQD
supported platinum nanoparticles, it was found that the relative amount of GQD
defects can be easily manipulated through both temperature and time of heating.
Platinum salt was dissolved in water with GQDs and hydrothermally reduced in a
Teflon-lined autoclave at controlled temperatures (140, 160, 180, or 200  C) for a
selected period of time (3, 6, or 12 h) to study the relationship between hydrother-
mal treatment and defect concentration.
These defects were largely oxygen containing groups that disrupted the graphitic
nature of GQDs and provided anchoring sites for metal deposition [66] as well as
diminishing electron density. The relative amounts of these groups were examined
by FTIR and XPS measurements, as depicted in Figs. 7.13 and 7.14. It was
found that the C–O–C asymmetric epoxy stretch (1135 cm1), the C¼O stretch
(1685 cm1), and the OH deformation in C–OH (1350 cm1) that were present in
the initial FTIR spectra of the untreated GQDs were markedly diminished and/or

GQD as-prep.
o
140 C, 6 h
o
Transmittance

140 C, 12 h
o
160 C, 3 h
o
160 C, 6 h
o
160 C, 12 h
o
180 C, 3 h
o
180 C, 6 h
o
180 C, 12 h
o
200 C, 3 h
o
200 C, 6 h
o
200 C, 12 h

4000 3500 3000 2000 1500 1000


-1
Wavenumber (cm )

Fig. 7.13 FTIR of platinum functionalized GQDs for different annealing temperatures and times.
Reprinted with permission from (Song, Y.; Chen, S.: Graphene quantum-dot-supported platinum
nanoparticles: defect-mediated electrocatalytic activity in oxygen reduction. ACS Appl Mater
Interfaces 2014, 6, 14050–60). Copyright (2014) American Chemical Society
298 C.P. Deming et al.

140 oC, 6 h

140 oC, 12 h

160 oC, 3 h
160 oC, 6 h
160 oC, 12 h

180 oC, 3 h

180 oC, 6 h

180 oC, 12 h
200 oC, 3 h
200 oC, 6 h
200 oC, 12 h

280 282 284 286 288 290 292


Binding Energy (eV)
Fig. 7.14 XPS of C 1s electrons of platinum functionalized GQDs for different annealing
temperatures and times. Reprinted with permission from (Song, Y.; Chen, S.: Graphene quan-
tum-dot-supported platinum nanoparticles: defect-mediated electrocatalytic activity in oxygen
reduction. ACS Appl Mater Interfaces 2014, 6, 14050–60). Copyright (2014) American Chemical
Society

disappeared with increasing temperature and heating time. In addition,


deconvolution of the C1s signal from the XPS spectra shows a diminishment of
the C¼O and COOH components as temperature and time increase with an almost
complete removal of these groups for samples when treated at 200  C for 12 h.
Furthermore, Raman spectroscopic measurements yielded additional insights
into the nature of the GQD support by examining the intensity ratio of the defect
(D) and graphitic (G) bands. The G band represents the E2g phonon while the D
band represents the breathing mode of κ-point phonons with A1g symmetry
[71, 72]. The intensity ratio of these two bands is well correlated with the XPS
and FTIR results where more defects are removed with a longer heating time and
higher temperature.
The samples were then tested for oxygen reduction and it was found that all
samples facilitated four-electron reduction of oxygen and the onset potentials were
more positive than þ0.80 V vs RHE. The best ORR performance was demonstrated
by the samples treated at 160  C for 12 h and at 180  C for 6 h which had mass
activities of 468.1 A/g and 435.7 A/g, and specific activities of 26.2 A/m2 and
24.6 A/m2, respectively, as depicted in Fig. 7.15.
Furthermore, the activities exhibit a volcano shaped variation with defect con-
centration as a result of direct manipulation of the electronic structure of the Pt
particles by the GQD defects. This trend is in agreement with the dynamics of
7 Enhanced Electrocatalytic Activity of Nanoparticle Catalysts in Oxygen. . . 299

I(D)/I(G)
1.0 0.975 1.050 1.126 1.202 1.277 1.353 1.428 1.504
50
0.9 Pt/G-160-12h
Pt/G-180-6h
40 +0.85 V
+0.90 V
0.8
E (V vs RHE)

140 oC, 6 h
140 oC, 12 h 30

Jk (A/m2)
0.7 160 oC, 3 h
160 oC, 6 h
160 oC, 12 h 20
0.6 180 oC, 3 h

Pt/G-140-12h
Pt/G-200-12h
180 oC, 6 h

Pt/G-160-6h
Pt/G-180-12h

Pt/G-140-6h
Pt/G-200-6h

Pt/G-180-3h
Pt/G-160-3h
Pt/G-200-3h
180 oC, 12 h 10
0.5 200 oC, 3 h
200 oC, 6 h
200 oC, 12 h
0
0.4
0.1 1 10 100 0 10 20 30 40 50 60 70

JK (A/m2) Defect % (XPS)

Fig. 7.15 (Left) Tafel plot of Pt nanoparticle functionalized GQDs prepared at different annealing
temperatures for a different period of times. (Right) Variation of the ORR kinetic current density at
þ0.85 (solid circles) and þ0.90 V (empty circles) with GQD structural defects manifested as the
defect concentrations from XPS measurements and I(D)/I(G) ratio in Raman spectroscopic
measurements. Reprinted with permission from (Song, Y.; Chen, S.: Graphene quantum-dot-
supported platinum nanoparticles: defect-mediated electrocatalytic activity in oxygen reduction.
ACS Appl Mater Interfaces 2014, 6, 14050–60). Copyright (2014) American Chemical Society

oxygen reduction in that a large number of defects will withdraw too much electron
density and limit the first step of reduction while samples with a small amount or no
defects will not diminish significant electron density resulting in the tight binding
between OIs and Pt and an unfavorable release of the final product. Importantly, the
peak performance in this study met the DOE target for 2017.
The results of these studies tie in very well with the d band model proposed
earlier if we allow the downshift of the d band center when electrons are withdrawn
and an upshift when charge is provided. A physical manifestation of this notion is
apparent in Fig. 7.10 where it can be seen that electron withdrawing groups enhance
the activity of Pt when compared to a bare surface while the activity diminishes
when the group is electron-donating. Although much can be concluded from these
results, more research is still needed to bring the level of quantitative understanding
concerning the impact on d band structure from organic capping ligands to that of
the impacts of alloying on d band structure.

7.4 Metal Oxide-Based Catalysts

As the battle for reducing the mass and enhancing the performance of platinum
electrocatalysts continues, metal oxides offer a cheap, abundant alternative to
precious metal catalysis. Although the mechanism of oxygen reduction on a
metal oxide surface is slightly different than that on a traditional metal
electrocatalyst (Pt, Pd), optimization of binding interactions between the surface
and O2 is the key to significant enhancement. Consequently, there has been much
300 C.P. Deming et al.

focus on improving activity and many different models proposed for enhancing
ORR activity, including direct manipulation of oxide vacancies [22], deliberate
surface functionalization with selected chemical ligands [36], and stoichiometry
and properties of the metal(s) [40].
For instance, recently it was shown that the removal of surface oxygen sites on
MnO2 was not only facile and controllable, but led to a vast improvement of the
ORR activity [22]. Experimentally, β-MnO2 nanorods were synthesized via a
hydrothermal route and were further processed at elevated temperatures for differ-
ent periods of times in argon or air atmospheres. The samples were denoted as
Air-250-2 h, Air-400-2 h, Ar-350-2 h, Ar-450-6 h to represent the treatment
conditions. Thermogravimetric analysis (TGA) showed that the loss of lattice
oxygen varied with the hydrothermal conditions, with the most vacancies identified
in the Ar-450-6 h sample and the least oxygen removal for the samples treated in
air. In electrocatalytic testing, the oxides with moderate surface oxygen removal
performed the best as predicted with current models of oxygen reduction in alkaline
media.
Theoretical studies were additionally performed to assess the nature of the
enhancement through oxygen removal as well as to compare samples with different
vacancy concentrations. The reduction pathway for pristine MnO2 and MnO2 with
one and two vacancies was modeled for the strength of adsorption of reaction
intermediates. It was found that O2 adsorbed onto the samples with oxygen
vacancies experienced bond elongation which activated the bond and allowed for
facile breaking of the bond in subsequent steps. Furthermore the barrier for the rate
determining step was 0.47 eV lower for the sample with one vacancy compared that
with two. The stick and ball surface models and the reaction pathway can be seen in
Fig. 7.16. The intermediate level of defects again shows the best adsorptive
properties and best ORR activity, in strong agreement with experimental data.
In another study, functionalization of indium tin oxide (ITO) nanoparticles with
chlorine ions has also been found to lead to marked enhancement of the ORR
activity. ITO nanoparticles were synthesized through a solvothermal method
followed by photo irradiation in o-dichlorobenzene. The resulting particles showed
an average diameter of 7.51 nm from TEM measurements and about 3.9 at.%
chlorine from XPS analysis. The Cl-functionalized particles exhibited enhanced
ORR activity, as compared to the bare particles at all potentials. Although the
mechanism for such enhancement is not fully understood, it is thought that the
improved activity results from the replacement of surface oxygen with chlorine
leading to a lower ratio of O/In and thus quasi-oxygen deficient sites as depicted in
Fig. 7.17 [36]. These sites facilitate the formation of oxide species and provide for
improved reduction kinetics [73].
Additionally, perovskite oxide (ABO3)-based electrocatalysts have displayed
surface property modulation resulting from alteration of the stoichiometry and
properties of the metals. A change in metal ratio or substitution for another metal
has been found to drastically alter the surface properties and thus activity [40]. Sim-
ilar to the d band theory that successfully predicts the activity of metal
nanoparticles, it has been shown than the degree of eg orbital filling of the active
(B) site can accurately describe the catalytic behavior of perovskite oxides. When
7 Enhanced Electrocatalytic Activity of Nanoparticle Catalysts in Oxygen. . . 301

Fig. 7.16 (a) Ball and stick models of perfect MnO2 surfaces and MnO2 with 1 and 2 defects. (b)
Adsorption of O2. (c) Adsorption of OOH. (d) Energy diagram for reduction process. Reprinted
with permission from (Cheng, F.; Su, Y.; Liang, J.; Tao, Z.; Chen, J.: MnO2-Based Nanostructures
as Catalysts for Electrochemical Oxygen Reduction in Alkaline Media. Chem Mater 2010,
22, 898–905) Copyright (2010) American Chemical Society

the eg orbital filling is 1, there is sufficient destabilization of the B-OH bond and an
energetically favorable energy transfer that promotes the facile removal of OH and
addition of O2. Filling greater than one will mean more electrons in the antibonding
orbitals and thus weaker interactions and poorer initial swap between OH and O2.
302 C.P. Deming et al.

Fig. 7.17 Ball and stick model for surface functionalization of indium tin oxide with chlorine.
Reprinted with permission from (Wang, N.; Niu, W.; Li, L.; Liu, J.; Tang, Z.; Zhou, W.; Chen, S.:
Oxygen electroreduction promoted by quasi oxygen vacancies in metal oxide nanoparticles
prepared by photoinduced chlorine doping. Chem Commun 2015, 51, 10620–10623.) Copyright
(2015) Royal Society of Chemistry

Fig. 7.18 Potentials at 25 μA cm2 as a function of eg orbital in perovskite-based oxides. Data


symbols vary with type of B ions (Cr, red; Mn, orange; Fe, grey; Co, green; Ni, blue; mixed
compounds, purple), where x ¼ 0 and 0.5 for Cr, and 0, 0.25 and 0.5 for Fe. Error bars represent
standard deviations. Reproduced with permission from Nat Chem 2011, 3, 546–550. Copyright
2011 Macmillan Publishers Ltd

When eg filling is smaller than 1, there will be little contribution to the formation of
antibonding orbitals and the bonding interaction will be too strong and hence limit
hydroxide regeneration. In fact, the plot of activity vs eg filling shows an apparent
volcano-shaped trend (Fig. 7.18), with the peak activity 4 times that of the poorest
performance in the series.
Interestingly, this model assumes that surface atoms exhibit localized eg orbitals
facing outwards towards the adsorbate rather than a delocalized d band model. The
localization of eg orbitals on B sites is seen as a consequence of the different
bonding symmetry of surface atoms, as compared to the bulk. The site activity is
therefore dependent on the surrounding metals and concentration of each as previ-
ously mentioned. This study is yet another example of surface engineering of an
electrocatalyst for optimal ORR activity.
7 Enhanced Electrocatalytic Activity of Nanoparticle Catalysts in Oxygen. . . 303

7.5 Conclusions and Perspectives

The kinetic limitations of electrocatalytic oxygen reduction has been recognized as


the bottleneck for efficient energy extraction in fuel cell technologies and therefore
many techniques have been developed to optimize catalytic activity. In this chapter
we have reviewed recent examples of interfacial engineering of nanoparticle
catalysts, including metal alloys in the forms of homogenous and segregated
composition, organic functionalization, as well as oxygen vacancy control and
surface orbital manipulation of metal oxides. As the strength of interactions
between the catalyst surface and oxygenated intermediates is largely responsible
for the kinetic limitations, the common theme for all of these techniques is
alteration of the electronic structure of the active metal that will result in more
favorable interactions with oxygenated intermediates. Since the binding interaction
between a metal surface and each individual reaction intermediate can not be
altered independently, an optimal binding interaction exists. This chapter has
outlined the experimental techniques utilized to reach this peak in addition to the
theoretical models to quantify and predict catalytic behaviors. Specifically, the
measure of a metal’s d band center has been directly linked to the metal adsorbate
binding energy. Models relating the effects of alloying on the d band center are well
developed and are in strong agreement with experimental evidence. Additionally
much progress has been made towards understanding the origin of enhancement
from organic ligand functionalization and how that may be exploited for further
catalytic gain. Finally, the rate limiting steps for ORR on oxide surfaces were
recognized as the OH/O2 displacement and the regeneration of surface OH. To
predict and explain the behavior of various perovskite oxides, the filling of the eg
antibonding orbitals serve more than a sufficient indicator. Even as the field of
electrocatalysis is rich in systematic activity enhancements and strong theoretical
models, much work is still needed to bring this technology to commercial settings.

Acknowledgment The authors thank the National Science Foundation for partial support of
the work.

References

1. Wagner FT, Lakshmanan B, Mathias MF (2010) Elecrochemistry and the future of the
automobile. J Phys Chem Lett 1:2204–2219. doi:10.1021/jz100553m|J
2. Cano-Castillo U (2013) Hydrogen and fuel cells: potential elements in the energy transition
scenario. Rev Mex Fis 59(2):85–92
3. Stephens IEL, Bondarenko AS, Grønbjerg U, Rossmeisl J, Chorkendorff I (2012) Understand-
ing the electrocatalysis of oxygen reduction on platinum and its alloys. Energy Environ Sci 5
(5):6744. doi:10.1039/c2ee03590a
4. Norskov JK, Rossmeisl J, Logadottir A, Lindqvist L, Kitchin JR, Bligaard T, Jonsson H (2004)
Origin of the overpotential for oxygen reduction at a fuel-cell cathode. J Phys Chem B
108:17886–17892
304 C.P. Deming et al.

5. Song C, Zhang J (2008) PEM fuel cell electrocatalysis and catalyst layers: fundamentals and
applications. Electrocatalytic oxygen reduction reaction. Springer, New York
6. Rabis A, Rodriguez P, Schmidt TJ (2012) Electrocatalysis for polymer electrolyte fuel cells:
recent achievements and future challenges. ACS Catal 2(5):864–890. doi:10.1021/cs3000864
7. Lim D-H, Wilcox J (2012) Mechanisms of the oxygen reduction reaction on defective
graphene-supported Pt nanoparticles from first-principles. J Phys Chem C 116
(5):3653–3660. doi:10.1021/jp210796e
8. Sabatier P (1911) Announcement. Hydrogenation and dehydrogenation for catalysis. Ber
Dtsch Chem Ges 44:1984–2001. doi:10.1002/cber.19110440303
9. Kinoshita K (1990) Particle-size effects for oxygen reduction on highly dispersed platinum in
acid electrolytes. J Electrochem Soc 137(3):845–848
10. Yano H, Inukai J, Uchida H, Watanabe M, Babu PK, Kobayashi T, Chung JH, Oldfield E,
Wieckowski A (2006) Particle-size effect of nanoscale platinum catalysts in oxygen reduction
reaction: an electrochemical and Pt-195 EC-NMR study. Phys Chem Chem Phys 8
(42):4932–4939. doi:10.1039/B610573d
11. Lim B, Jiang MJ, Camargo PHC, Cho EC, Tao J, Lu XM, Zhu YM, Xia YN (2009) Pd-Pt
bimetallic nanodendrites with high activity for oxygen reduction. Science 324
(5932):1302–1305. doi:10.1126/science.1170377
12. Chen ZW, Waje M, Li WZ, Yan YS (2007) Supportless Pt and PtPd nanotubes as
electrocatalysts for oxygen-reduction reactions. Angew Chem Int Ed 46(22):4060–4063.
doi:10.1002/Anie.200700894
13. Xiao L, Zhuang L, Liu Y, Lu JT, Abruna HD (2009) Activating Pd by morphology tailoring for
oxygen reduction. J Am Chem Soc 131(2):602–608
14. Savadogo O, Lee K, Oishi K, Mitsushima S, Kamiya N, Ota KI (2004) New palladium alloys
catalyst for the oxygen reduction reaction in an acid medium. Electrochem Commun 6
(2):105–109. doi:10.1016/J.Elecom.2003.10.020
15. Shao MH, Sasaki K, Adzic RR (2006) Pd-Fe nanoparticles as electrocatalysts for oxygen
reduction. J Am Chem Soc 128(11):3526–3527. doi:10.1021/Ja060167d
16. Zhang JL, Vukmirovic MB, Sasaki K, Nilekar AU, Mavrikakis M, Adzic RR (2005) Mixed-
metal Pt monolayer electrocatalysts for enhanced oxygen reduction kinetics. J Am Chem Soc
127(36):12480–12481
17. Zhang JL, Vukmirovic MB, Xu Y, Mavrikakis M, Adzic RR (2005) Controlling the catalytic
activity of platinum-monolayer electrocatalysts for oxygen reduction with different substrates.
Angew Chem Int Ed 44(14):2132–2135
18. Greeley J, Stephens IEL, Bondarenko AS, Johansson TP, Hansen HA, Jaramillo TF,
Rossmeisl J, Chorkendorff I, Norskov JK (2009) Alloys of platinum and early transition
metals as oxygen reduction electrocatalysts. Nat Chem 1(7):552–556. doi:10.1038/Nchem.367
19. Bing YH, Liu HS, Zhang L, Ghosh D, Zhang JJ (2010) Nanostructured Pt-alloy
electrocatalysts for PEM fuel cell oxygen reduction reaction. Chem Soc Rev 39
(6):2184–2202. doi:10.1039/B912552c
20. Zhou Z-Y, Kang X, Song Y, Chen S (2012) Ligand-mediated electrocatalytic activity of pt
nanoparticles for oxygen reduction reactions. J Phys Chem C 116(19):10592–10598. doi:10.
1021/jp300199x
21. He G, Song Y, Phebus B, Liu K, Deming CP, Hu P, Chen S (2013) Electrocatalytic activity of
organically functionalized silver nanoparticles in oxygen reduction. Sci Adv Mater 5
(11):1727–1736. doi:10.1166/sam.2013.1624
22. Cheng FY, Zhang TR, Zhang Y, Du J, Han XP, Chen J (2013) Enhancing electrocatalytic
oxygen reduction on MnO2 with vacancies. Angew Chem Int Ed 52(9):2474–2477. doi:10.
1002/anie.201208582
23. Strasser P, Koh S, Anniyev T, Greeley J, More K, Yu CF, Liu ZC, Kaya S, Nordlund D,
Ogasawara H, Toney MF, Nilsson A (2010) Lattice-strain control of the activity in dealloyed
core-shell fuel cell catalysts. Nat Chem 2(6):454–460. doi:10.1038/Nchem.623
7 Enhanced Electrocatalytic Activity of Nanoparticle Catalysts in Oxygen. . . 305

24. Kitchin JR, Norskov JK, Barteau MA, Chen JG (2004) Modification of the surface electronic
and chemical properties of Pt(111) by subsurface 3d transition metals. J Chem Phys 120
(21):10240–10246. doi:10.1063/1.1737365
25. Stephens IE, Bondarenko AS, Perez-Alonso FJ, Calle-Vallejo F, Bech L, Johansson TP, Jepsen
AK, Frydendal R, Knudsen BP, Rossmeisl J, Chorkendorff I (2011) Tuning the activity of Pt
(111) for oxygen electroreduction by subsurface alloying. J Am Chem Soc 133
(14):5485–5491. doi:10.1021/ja111690g
26. Stamenkovic VR, Mun BS, Arenz M, Mayrhofer KJJ, Lucas CA, Wang GF, Ross PN,
Markovic NM (2007) Trends in electrocatalysis on extended and nanoscale Pt-bimetallic
alloy surfaces. Nat Mater 6(3):241–247
27. Stamenkovic VR, Fowler B, Mun BS, Wang GF, Ross PN, Lucas CA, Markovic NM (2007)
Improved oxygen reduction activity on Pt3Ni(111) via increased surface site availability.
Science 315(5811):493–497
28. Zhou ZY, Kang XW, Song Y, Chen SW (2012) Enhancement of the electrocatalytic activity of
Pt nanoparticles in oxygen reduction by chlorophenyl functionalization. Chem Commun 48
(28):3391–3393. doi:10.1039/C2cc17945h
29. Liu K, Kang XW, Zhou ZY, Song Y, Lee LJ, Tian D, Chen SW (2013) Platinum nanoparticles
functionalized with acetylene derivatives: Electronic conductivity and electrocatalytic activity
in oxygen reduction. J Electroanal Chem 688:143–150
30. He GQ, Song Y, Liu K, Walter A, Chen S, Chen SW (2013) Oxygen reduction catalyzed by
platinum nanoparticles supported on graphene quantum dots. ACS Catal 3(5):831–838. doi:10.
1021/Cs400114s
31. Song Y, Chen SW (2014) Graphene quantum-dot-supported platinum nanoparticles: defect-
mediated electrocatalytic activity in oxygen reduction. ACS Appl Mater Interfaces
6(16):14050–14060. doi:10.1021/Am503388z
32. Chen W, Chen SW, Ding FZ, Wang HB, Brown LE, Konopelski JP (2008) Nanoparticle-
mediated intervalence transfer. J Am Chem Soc 130(36):12156–12162. doi:10.1021/
Ja803887b
33. Chen W, Zuckerman NB, Kang XW, Ghosh D, Konopelski JP, Chen SW (2010) Alkyne-
protected ruthenium nanoparticles. J Phys Chem C 114(42):18146–18152
34. Kang XW, Chen W, Zuckerman NB, Konopelski JP, Chen SW (2011) Intraparticle charge
delocalization of carbene-functionalized ruthenium nanoparticles manipulated by selective ion
binding. Langmuir 27(20):12636–12641
35. Zhou ZY, Ren J, Kang X, Song Y, Sun SG, Chen S (2012) Butylphenyl-functionalized Pt
nanoparticles as CO-resistant electrocatalysts for formic acid oxidation. Phys Chem Chem
Phys 14(4):1412–1417. doi:10.1039/c1cp23183a
36. Wang N, Niu W, Li L, Liu J, Tang Z, Zhou W, Chen S (2015) Oxygen electroreduction
promoted by quasi oxygen vacancies in metal oxide nanoparticles prepared by photoinduced
chlorine doping. Chem Commun 51:10620–10623. doi:10.1039/C5CC02808F
37. Cheng F, Su Y, Liang J, Tao Z, Chen J (2010) MnO2-based nanostructures as catalysts for
electrochemical oxygen reduction in alkaline media. Chem Mater 22(3):898–905. doi:10.
1021/cm901698s
38. Cheng FY, Chen J (2012) Metal-air batteries: from oxygen reduction electrochemistry to
cathode catalysts. Chem Soc Rev 41(6):2172–2192. doi:10.1039/c1cs15228a
39. Goodenough JB, Cushing BL (2003) Handbook of fuel cells-fundamentals, technology and
applications, vol 2. Wiley, New York
40. Suntivich J, Gasteiger HA, Yabuuchi N, Nakanishi H, Goodenough JB, Shao-Horn Y (2011)
Design principles for oxygen-reduction activity on perovskite oxide catalysts for fuel cells and
metal-air batteries. Nat Chem 3(7):546–550
41. Xiong L, Kannan AM, Manthiram A (2002) Pt-M (M ¼ Fe, Co, Ni and Cu) electrocatalysts
synthesized by an aqueous route for proton exchange membrane fuel cells. Electrochem
Commun 4(11):898-903. Pii: S1388-2481(02)00485x
306 C.P. Deming et al.

42. Wu JB, Zhang JL, Peng ZM, Yang SC, Wagner FT, Yang H (2010) Truncated octahedral
Pt3Ni oxygen reduction reaction electrocatalysts. J Am Chem Soc 132(14):4984–4985. doi:10.
1021/ja100571h
43. Mavrikakis M, Hammer B, Norskov JK (1998) Effect of strain on the reactivity of metal
surfaces. Phys Rev Lett 81(13):2819–2822
44. Rodriguez JA, Goodman DW (1992) The nature of the metal bond in bimetallic surfaces.
Science 257(5072):897–903. doi:10.1126/Science.257.5072.897
45. Chen MS, Kumar D, Yi CW, Goodman DW (2005) The promotional effect of gold in catalysis
by palladium-gold. Science 310(5746):291–293. doi:10.1126/Science.1115800
46. Stamenkovic V, Mun BS, Mayrhofer KJJ, Ross PN, Markovic NM, Rossmeisl J, Greeley J,
Norskov JK (2006) Changing the activity of electrocatalysts for oxygen reduction by tuning
the surface electronic structure. Angew Chem Int Ed 45(18):2897–2901
47. Xiao L, Huang B, Zhuang L, Lu JT (2011) Optimization strategy for fuel-cell catalysts based
on electronic effects. Rsc Adv 1(7):1358–1363. doi:10.1039/C1ra00378j
48. Greeley J, Norskov JK (2009) Combinatorial density functional theory-based screening of
surface alloys for the oxygen reduction reaction. J Phys Chem C 113(12):4932–4939. doi:10.
1021/Jp808945y
49. Fernandez JL, Raghuveer V, Manthiram A, Bard AJ (2005) Pd-Ti and Pd-Co-Au
electrocatalysts as a replacement for platinum for oxygen reduction in proton exchange
membrane fuel cells. J Am Chem Soc 127(38):13100–13101. doi:10.1021/Ja0534710
50. Suo YG, Zhuang L, Lu JT (2007) First-principles considerations in the design of Pd-alloy
catalysts for oxygen reduction. Angew Chem Int Ed 46(16):2862–2864. doi:10.1002/Anie.
200604332
51. Sleightholme AES, Varcoe JR, Kucernak AR (2008) Oxygen reduction at the silver/hydrox-
ide-exchange membrane interface. Electrochem Commun 10(1):151–155. doi:10.1016/J.
Elecom.2007.11.008
52. Guo JS, Hsu A, Chu D, Chen RR (2010) Improving oxygen reduction reaction activities on
carbon-supported Ag nanoparticles in alkaline solutions. J Phys Chem C 114(10):4324–4330.
doi:10.1021/Jp910790u
53. Chatenet M, Genies-Bultel L, Aurousseau M, Durand R, Andolfatto F (2002) Oxygen reduc-
tion on silver catalysts in solutions containing various concentrations of sodium hydroxide—
comparison with platinum. J Appl Electrochem 32(10):1131–1140. doi:10.1023/
A:1021231503922
54. Varcoe JR, Slade RCT (2005) Prospects for alkaline anion-exchange membranes in low
temperature fuel cells. Fuel Cells 5(2):187–200. doi:10.1002/Fuce.200400045
55. Slanac DA, Hardin WG, Johnston KP, Stevenson KJ (2012) Atomic ensemble and electronic
effects in Ag-rich AgPd nanoalloy catalysts for oxygen reduction in alkaline media. J Am
Chem Soc 134(23):9812–9819. doi:10.1021/Ja303580b
56. Chen S, Templeton AC, Murray RW (2000) Monolayer-protected cluster growth dynamics.
Langmuir 16:3543–3548
57. Templeton AC, Wuelfing MP, Murray RW (2000) Monolayer protected cluster molecules.
Acc Chem Res 33(1):27–36
58. Cavaliere S, Fdr R, Etcheberry A, Herlem M, Perez H (2004) Direct electrocatalytic activity of
capped platinum nanoparticles toward oxygen reduction. Electrochem Solid-State Lett 7(10):
A358. doi:10.1149/1.1792259
59. Baret B, Aubert PH, L’Hermite MM, Pinault M, Reynaud C, Etcheberry A, Perez H (2009)
Nanocomposite electrodes based on pre-synthesized organically capped platinum
nanoparticles and carbon nanotubes. Part I: Tuneable low platinum loadings, specific H upd
feature and evidence for oxygen reduction. Electrochim Acta 54(23):5421–5430. doi:10.1016/
j.electacta.2009.04.033
60. Genorio B, Strmcnik D, Subbaraman R, Tripkovic D, Karapetrov G, Stamenkovic V,
Pejovnik S, Markovic N (2010) Selective catalysts for the hydrogen oxidation and oxygen
7 Enhanced Electrocatalytic Activity of Nanoparticle Catalysts in Oxygen. . . 307

reduction reactions by patterning of platinum with calix[4]arene molecules. Nat Mater


9:998–1003. doi:10.1038/nmat2883
61. Strmcnik D, Escudero-Escribano M, Kodama K, Stamenkovic V, Cuesta A, Markovic N
(2010) Enhanced electrocatalysis of the oxygen reduction reaction based on patterning of
platinum surfaces with cyanide. Nat Chem 2:880–885. doi:10.1038/nchem.771
62. Pietron JJ, Garsany Y, Baturina O, Swider-Lyons KE, Stroud RM, Ramaker DE, Schull TL
(2008) Electrochemical observation of ligand effects on oxygen reduction at ligand-stabilized
Pt nanoparticle electrocatalysts. Electrochem Solid-State Lett 11(8):B161. doi:10.1149/1.
2937448
63. Kostelansky CN, Pietron JJ, Chen MS, Dressick WJ, Swider-Lyons KE, Ramaker DE, Stroud
RM, Klug CA, Zelakiewicz BS, Schull TL (2006) Triarylphosphine-stabilized platinum
nanoparticles in three-dimensional nanostructured films as active electrocatalysts. J Phys
Chem B 110(43):21487–21496. doi:10.1021/Jp062663u
64. Song Y, Liu K, Chen SW (2012) AgAu bimetallic janus nanoparticles and their
electrocatalytic activity for oxygen reduction in alkaline media. Langmuir 28
(49):17143–17152. doi:10.1021/La303513x
65. Lima FHB, Zhang J, Shao MH, Sasaki K, Vukmirovic MB, Ticianelli EA, Adzic RR (2007)
Catalytic activity-d-band center correlation for the O2 reduction reaction on platinum in
alkaline solutions. J Phys Chem C 111:404–410
66. Hull RV, Li L, Xing YC, Chusuei CC (2006) Pt nanoparticle binding on functionalized
multiwalled carbon nanotubes. Chem Mater 18(7):1780–1788. doi:10.1021/Cm0518978
67. Palaniselvam T, Irshad A, Unni B, Kurungot S (2012) Activity modulated low platinum
content oxygen reduction electrocatalysts prepared by inducing nano-order dislocations on
carbon nanofiber through N-2-doping. J Phys Chem C 116(28):14754–14763. doi:10.1021/
Jp300881p
68. Timperman L, Feng YJ, Vogel W, Alonso-Vante N (2010) Substrate effect on oxygen
reduction electrocatalysis. Electrochim Acta 55(26):7558–7563. doi:10.1016/j.electacta.
2009.09.076
69. Vogel W, Timperman L, Alonso-Vante N (2010) Probing metal substrate interaction of Pt
nanoparticles: structural XRD analysis and oxygen reduction reaction. Appl Catal Gen 377
(1–2):167–173. doi:10.1016/j.apcata.2010.01.034
70. Liu X, Yao KX, Meng CG, Han Y (2012) Graphene substrate-mediated catalytic performance
enhancement of Ru nanoparticles: a first-principles study. Dalton Trans 41(4):1289–1296.
doi:10.1039/C1dt11186h
71. Ferrari AC, Basko DM (2013) Raman spectroscopy as a versatile tool for studying the
properties of graphene. Nat Nanotechnol 8(4):235–246
72. Dresselhaus MS, Terrones M (2013) Carbon-based nanomaterials from a historical perspec-
tive. Proc IEEE 101(7):1522–1535
73. Kim J, Yin X, Tsao KC, Fang S, Yang H (2014) Ca(2)Mn(2)O(5) as oxygen-deficient
perovskite electrocatalyst for oxygen evolution reaction. J Am Chem Soc 136
(42):14646–14649. doi:10.1021/ja506254g
Chapter 8
Primary Oxide Latent Storage and Spillover
for Reversible Electrocatalysis in Oxygen
and Hydrogen Electrode Reactions

Milan M. Jaksic, Angeliki Siokou, Georgios D. Papakonstantinou,


and Jelena M. Jaksic

8.1 Introduction

Ever since Sir William Grove invented fuel cells (FCs) [1], electrocatalysts of the
oxygen electrode (ROE) reactions became the main imperative target, challenge
and dream in the whole electrochemical science of aqueous and later PEM media,
while its substantiation is the subject matter of the present paper. The main focus
has always primarily been posed on noble (Pt, Au, Ru) metals, while the reversible
primary (Pt-OH, Au-OH, Ru-OH), and polarizable surface (Pt¼O, Au¼O, Ru¼O)
oxides, along with H-adatoms (Pt-H, Au-H, Ru-H) represent interactive species
defining the overall electrode behavior and properties. In such a respect
potentiodynamic spectra, (Fig. 8.1), usually reveal within a narrow potential
range the highly reversible peaks of primary oxide adsorptive growth and its
backwards desorptive removal [2–4],

Pt þ 2H2 O $ Pt-OH þ H3 Oþ þ e ð8:1Þ

as a typical double layer (DL) charging and discharging pseudo-capacitance [5–11],


while the former soon later irreversibly disproportionates into the strongly polariz-
able, more adsorptive and remarkably more stable surface oxide (Pt¼O) monolayer
[2–4],

M.M. Jaksic
Institute of Chemical Engineering Sciences, CEHT/FORTH, Patras, Greece
Faculty of Agriculture, University of Belgrade, Belgrade, Serbia
e-mail: [email protected]
A. Siokou • G.D. Papakonstantinou • J.M. Jaksic (*)
Institute of Chemical Engineering Sciences, CEHT/FORTH, Patras, Greece
e-mail: [email protected]

© Springer International Publishing Switzerland 2016 309


K.I. Ozoemena, S. Chen (eds.), Nanomaterials for Fuel Cell Catalysis,
Nanostructure Science and Technology, DOI 10.1007/978-3-319-29930-3_8
310 M.M. Jaksic et al.

Fig. 8.1 Cyclic voltammograms of polycrystalline Pt scanned in 0.1 M NaOH at sweep rate
100 mV s1(Reprinted with permission from [7]. Copyright (2014) American Chemical Society)

Pt-OH þ H2 O ! Pt ¼ O þ H3 Oþ þ e ð8:2Þ

With the corresponding pronounced irreversible desorption peak all along the
equivalent reverse potential scan, back and forth. The former, (Eq. 8.1), arises as
the highly reversible reaction, proceeds independently and extremely fast, while the
latter, (Eq. 8.2), as a typical irreversible disproportionation reaction step, is rather
slow and highly polarizable. Such state of the experimental evidence reveals and
definitely defines electrocatalytic fate and properties of all plain Pt(Au), and
non-interactive (Pt/C, Au/C) supported noble and in particular other hypo (n  5)
and hyper (n > 5) d-metals along transition series of elements. In such circum-
stances, there is no plain individual d-metal which might feature electrocatalytic
properties all along the potential axis between hydrogen and oxygen evolving
limits, but partially reversible and partially polarizable.
In fact, in the absence of the (Pt-OH ! 0), therefrom imposes such a critical and
typical, very pronounced the reaction polarization range (Pt¼O ! 1, Eq. 8.2 and
Fig. 8.1), all along, after the primary oxide adsorption/desorption peak, until the
oxygen potential evolving limits, and in both scan directions, back and forth [5–
11]. Such highly reversible (Eq. 8.1), relative to the subsequent strongly irrevers-
ible, and substantially polarizable (Eq. 8.2) transient and coexistence of two
interrelated reaction steps, has been decisive as concerns the overall electrocatalytic
properties and behavior of all d-metals all along transition series, on the first place
for the oxygen electrode reactions (ORR and OER).
The present analysis is best reflected and proved by comparison of Figs. 8.1, 8.2
and 8.3, once when missing Pt-OH spillover (Pt-OH ! 0, Pt¼O ! 1), then when
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 311

Pt/ 0.01M HCHO


0.1M KOH
12
Pt/ 0.1M KOH
i (mA/cm2)

e
c
4
g
f

0
a

b
-4

-8
E/V vs.RHE

0 0.5 1.0 1.5 2.0

-1.0 0.0 1.0


E/Vvs SCE d 0.1M HCHO
Pt
e 0.1M KOH
80
i (mA/cm2)

60

40

20

-20

E/VvsRHE
-0.5 0 0.5 1.0 1.5 2.0

Fig. 8.2 (a, b) Cyclic voltammograms scanned on a polycrystalline Pt wire electrode in alkaline
(0.1 M KOH, dashed lines) solution and in admixture of formaldehyde (0.01 M (a) and 0.1 M
HCHO (b), full lines) at 200 mV s1 sweep rate between hydrogen and oxygen potential evolving
312 M.M. Jaksic et al.

there proceeds simple effusion of self-generated Pt-OH, or finally, the enriched


latent storage spillover enables enormous primary oxide adsorptive deposition and
reverse desorption, (Pt-OH ! 1, Pt¼O ! 0), exactly as big capacitors do, and when
continuously and reversibly present all over the potentiodynamic cycle (see further
downwards detailed stepwise discussion).

8.2 Electrocatalysis of Oxygen Electrode Reactions: Initial


Ideas

Since the equimolar ratio of the primary and surface oxide concentrations defines
the optimal interfering self-catalytic spillover reaction step in cathodic oxygen
reduction reaction mechanism (ORR) [12],

Pt-OH þ Pt ¼ O þ 3H3 Oþ þ 3e ! 5H2 O þ 2Pt ð8:3Þ

and in particular along the reversible (low slope, 30 mV/dec, or even lower) parts of
Tafel line plots, the irreversible disproportionation (Eq. 8.2) imposes an extremely
high reaction polarization barrier (see later downwards) that amounts for even more
than 600 mV s and then, in absence of the external Pt-OH supply, makes plain Pt
(and Au) and non-interactive supported both (Pt/C or Au/C) platinum and gold
irreversible for the oxygen electrode reactions within the broader potential range,
back and forth. The thermodynamic definition of irreversibility then would state
that plain Pt(Au) by no means can feature the reversibility and/or (electro)catalytic
activity all along the potential axis between hydrogen and oxygen evolving limits,
or more specifically, within the whole potentiodynamic circle. The strongly adsorp-
tive and thence highly polarizable Pt¼O (Au¼O), deprived from any local and/or
external Pt-OH (Au-OH) source and supply, then defines one of the most pro-
nounced issues of the reaction polarization in the entire electrochemical science: No
Pt-OH, means that there is no reversible reaction (Eq. 8.3). Quite another story
arises when nanostructured Pt(Au) electrocatalyst is selective interactive grafting
bonded on various, in particular mixed valence hypo-d-(f)-oxide supports.
Whereas hydrogen molecules undergo spontaneous adsorptive dissociation on
plain Pt (Pt/C) yielding H-adatoms (Pt-H) to establish thermodynamic equilibrium
of the RHE (Reversible Hydrogen Electrode, (Pt(H2)/Pt-H/H3O+)), within the




Fig. 8.2 (continued) limits. Labels: (a) reversible hydrogen adsorption peak; (b) irreversible Pt
surface oxide (Pt¼O) desorption peak; (c, d ) successive peaks of anodic aldehyde oxidation; (e)
sudden sharp current jump and reverse peak of repeated HCHO oxidation in the course of
successive cathodic scan; (g, f ) reversible H-adatoms oxidation and desorption peaks, respectively
(Reprinted with permission from [7]. Copyright (2014) American Chemical Society)
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 313

0.20

0.15
2
0.10

0.05 1
I, A

0.00

-0.05

-0.10

-0.15

-0.20
Tcell=80°C, scan rate: 50mV/s

-0.25
0.0 0.5 1.0 1.5
E(RHE), V

Fig. 8.3 Cyclic voltammograms of mixed valence hypo-d-oxides supported nanostructured Pt


electrode (Pt/TiO2,WO3/C), scanned in He stream, ones at negligible moisture content (curve 1)
and at 80  C water vapor saturation (curve 2) (Reprinted with permission from [7]. Copyright
(2014) American Chemical Society)

above identified critical potential range Pt (Au) is deprived (Eq. 8.2), from the
Pt-OH (Au-OH) to provide reversible properties for the ROE ((Pt(O2)/Pt-OH,
Pt¼O)/OH), (Eq. 8.3), and both together assemble the reversible hydrogen fueled
L&MT PEMFCs. In other words, the Pt-OH (or Au-OH) plays the same thermo-
dynamic constitutional role for establishing the equilibrium for the ROE, as Pt-H
does for the RHE, and thence, unavoidably imposes continuous need and requires a
permanent source of the former (Pt-OH) to overcome the reaction polarization by
the latter (Pt¼O), and enables the reversible electrocatalytic properties of the ROE
[7]. In such a respect, the first main step in the present concept towards the ROE
implies that catalytic hyper-d-metals (Pt, Au, Ru) establish with hypo-d-(f)-oxides
(n  5, or their mixed valence compounds, like Nb2O5, TiO2, CeO2; Ta2O5, TiO2,
CeO2; or WO3, Nb2O5, TiO2, GdO2) the interactive Brewer type (d-d or d-f)
bonding effect [13, 14], otherwise well known in heterogeneous catalysis as Tauster
[15–18] SMSI (Strong Metal-Support Interaction), one of the strongest in the entire
chemistry [5]). Consequently, this way interactive inter-bonded composite
electrocatalysts (example, Pt/Nb2O5, TiO2), while strong bonding, impose remark-
ably stretched d-orbitals, and thereby exhibit much weaker adsorptive inter-bond-
ing strengths of intermediates (Pt-H, Pt-OH) in the RDSs (Rate Determining Steps),
thence correspondingly facilitated cleavage of the latter, and thereby, increased the
catalytic activity both for hydrogen and oxygen electrode reactions. Meanwhile,
some other accompanying effects, in particular the ones associated with the Pt-OH
latent storage and spillover, play even more significant and decisive role for the
314 M.M. Jaksic et al.

latter [5–7, 19, 20]. In such a respect, the dramatically pronounced Pt-OH spillover
effect has primarily been already noticed as the remarkable increased CO-tolerance
[5]. Then, hypo-d-oxides and their mixed valence compounds, as based on typical
d-d-(f)-metallic bonds, exhibit extra high stability in both acidic and alkaline media,
and many of them pronounced (above 300 S/cm) electron conductivity. Further-
more, majority of hypo-d-(f)-oxides and in particular of higher altervalent numbers,
feature prevailingly high percentage of dissociative water molecules adsorption
[21, 22], (Eq. (8.1-a)), and thence, the mostly enhanced surface membrane type of
hydroxyl ions migration mass transfer [23–25], (Eqs. (8.1-b) and (8.1-c)),

Nb2 O5 þ 5Hs2 O $ 2NbðOHÞ5 ð8:1-aÞ


NbðOHÞ5 þ Pt ! NbðOHÞ4þ þ Pt-OH þ e ð8:1-bÞ
4þ 5þ þ
NbðOHÞ þ 2H2 O ! NbðOHÞ þ H3 O ð8:1-cÞ
____________________________________
X
Pt þ 2H2 O $ Pt-OH þ H3 Oþ þ e ð8:1Þ

finally, under anodic polarization, ending up with the prevailing electron transfer to
the interactive supported metallic catalyst, so that the Pt-OH behaves as a pro-
nounced dipole species [26], and thus, exhibits the strong spillover surface repul-
sion, transfer and distribution. At the same time, the highly pronounced reversible
potentiodynamic peaks testify for the extremely fast overall spillover (surface)
reaction (Eq. 8.1), in both directions [2–4], primarily used for DL charging and
discharging, and then being ready and available for fast heterogeneous
electrocatalytic reactions. In the same context, it would be significant inferring
that whenever anodic oxidation reactions of the Pt-OH (Au-OH), exceed at least for
an order of magnitude the disproportionation rate (Eq. 8.2), such as the ones with
HCHO [5–11], and other aldehydes, simple alcohols and their acids, then this way
these succeed highly to suppress and remarkably postpone the Pt¼O growth all
along the potential axis till the close proximity of anodic oxygen evolving limits,
and at the same time consequently show as experimentally confirmed enormous
broad extension growth of the Pt-OH (Au-OH) adsorption peak (Fig. 8.2a, b). In
fact, since Pt-OH(Au-OH) (electro)catalytically belong to heterogeneous surface
reactions and are insoluble in aqueous media, to approach the effect similar to the
one just inferred with HCHO, and then partially suppress or completely eliminate
the surface oxide adsorptive growth (Eq. 8.2), along with the corresponding reac-
tion polarization, and create the reversible electrocatalysts for the ROE (ORR,
OER), (Eq. 8.3), there have been employed interactive hypo-d-(f)-oxide supports
yielding the advanced continuous, recoverable and renewable latent storage of
primary oxides [5–11, 19, 20]. These dramatically differ from plain Pt (Au) and
impose quite another quality and behavior in electrocatalysis for the oxygen
electrode reactions (ORR, OER).
In such an overall constellation, individual hypo-d-(f)-oxides, and even more so
their mixed valence composite compounds, when in expanded (sol-gel synthesis
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 315

with thorough liquid CO2 supercritical drying) hydrated surface state (Fig. 8.3),
behave as unlimited latent storage and spillover sources of the primary oxide [5–11,
19, 20], continuously being renewed and recovered simply by water vapor supply
and instantaneously proceeding with spontaneous dissociative adsorption of water
molecules [21, 22].

8.2.1 Adsorptive Dissociation of Water Molecules


and Membrane Type Surface Migration

The first decisive step towards rather fast spillover widespreading phenomena of the
Pt-OH is the consequence of a strong first principle thermodynamic confirmed
evidence (Density Functional Calculations, DFC) by Vittadini et al. [21], that
water molecules undergo prevailing spontaneous dissociative adsorption on anatase
(1 0 1), and even rutile titania, and more so on the higher altervalent oxides [22] of
tungsten, molybdenum, tantalum, niobium and/or cerium, etc. (Fig. 8.4), as the
general oxophilicity properties of hypo-d-metal oxides. In addition, the first-
principles molecular-dynamic simulations showed the existence of a mechanism
for thermodynamically favored spontaneous dissociation of water molecules even

Fig. 8.4 The perspective views of DFT-optimized atomic structures for: (a) the clean anatase
(ADM) ad-molecule model of unreconstructed (0 0 1) surface; (b) the dissociated state of water
(0.5 monolayer) on (0 0 1); (c) the relaxed geometries of molecular state of adsorbed water (1.0
monolayer of hydroxylated anatase) on (0 0 1); and (d) the mixed state of water on (001) with a
half-dissociated coverage of adsorbed monolayer water molecules (courtesy of A. Vittadini,
cf. [21])
316 M.M. Jaksic et al.

at low coverage of oxygen vacancies of the anatase (1 0 1) surface [22], and


consequently at the Magneli phases, as substantially suboxide structure significant
both as highly bulk electronic conductive (up to 1000 S/cm), membrane type
surface transferring hydroxide species and interactive catalyst support. In fact,
this is the status of reversible open circuit dissociative adsorption of water mole-
cules at the equilibrium state, something like capillary phenomena in adsorption
after some critical coverage extents. Meanwhile, in the presence of the nano-sized
metallic part of the catalyst, and continuous enough moisture supply, directional
electric field (or, electrode polarization), further disturbs such an established equi-
librium and dynamically imposes further continuous forced dissociation of water
molecules, and as the consequence, their membrane transport properties (Livage
et al. [23–25], (Eqs. (8.1-a), (8.1-b), (8.1-c) summation (∑) yields Eq. (8.1)),
definitely resulting with the Pt-OH dipole spillover features.
Such an oxide mixed valence network, in particular of polyvalent (high
altervalent numbers) hypo-d-elements, when in hydrous state, distinctly behaves
as an ion exchange membrane even for their own surface hydroxide migration. In
fact, gels (aero and xerogels) are biphasic systems in which solvent molecules are
trapped inside an oxide network, and such a material can be considered as a water-
oxide membrane composite [23–25]. At the same time, the highly pronounced
reversible potentiodynamic peaks testify for the extremely fast and independent
overall spillover reaction (Eq. 8.1), in both directions [2–4], primarily used for DL
charging and discharging pseudo-capacitance [5–8], and being always available for
heterogeneous electrode reactions (see further downwards).

8.2.2 Electrocatalytic Spillover Phenomena

The first spillover phenomenon in heterogeneous catalysis was observed and


defined by Boudart [27–29] for the interactive supported bronze type (Pt/WO3)
catalyst, initially at high temperature (above 400  C) for pure solid oxide system.
Meanwhile, after the dissociative adsorption of water molecules on hypo-d-oxide
supports of Pt [21, 22], the fast interactive effusion of H-adatoms over its hydrated
(W(OH)6) surface becomes dramatically sped up even at ambient conditions in the
ultimate presence of condensed at least monolayer aqueous precipitate, and vice
versa, establishing capillary phenomena of speeding up effusion of the main
catalytic reaction intermediates (Pt-OH). Such a striking sharp wetness impact
upon the overall spillover phenomena associated with hydrated hypo-d-oxides in
aqueous media, implies Ertel [30] auto-catalytic molecular water effect, too, which
states that the catalytic reaction of hydrogen oxidation upon Pt surface, even at deep
low temperatures (140 K), proceeds with remarkable amounts of the Pt-OH, as the
decisive and accumulated intermediate, including the self-catalytic step with
adsorbed water molecules (equivalent to Eq. 8.1),
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 317

2Pt þ Oads þ H2 Oads ! 2Pt-OH ð8:4Þ

In fact, Ertel has pointed out the substantial overall significance of water
molecules in heterogeneous catalysis for oxidation processes that in general pro-
ceed over the Pt-OH generation and spillover, and substantially implies the funda-
mental peak relations (Eq. 8.1) on the same wet way in (electro)catalysis.
This is significant both for the evidence of the extremely fast spillover
widespreading and thereby resulting with imposed the reversible hydrated substrate
reduction. The latter finally leads to the corresponding form of electrocatalytically
active bronze (Pt/H0.35WO3) for cathodic processes, in which non-stoichiometric
incorporated hydrogen obeys the same free reactive properties like adsorptive
(Pt-H), and is the main source for the electrode or heterogeneous catalytic reactions.
In other words, the point is that spontaneous dissociative adsorption of water
molecules [21, 22] imposes much smaller activation energy for transformation of
the resulting hydrated W(OH)6, into corresponding bronze state, and occurs even at
ambient temperature, then behaving remarkably different than the initial solid oxide
WO3, and thereby, dramatically facilitates the overall spillover effect under pro-
nounced wet status (the activation energies thence being in the ratio of 2.2, one with
another). The alterpolar interchanges between the bronze type electrocatalyst and
its hydrated state are correspondingly approved occurring instantaneously and
reversibly fast, exactly because of the substantially facilitated Boudart [27–29]
spillover effect and behave as a thermodynamic equilibrium (Pt/H0.35WO3 $ Pt/
W(OH)6 or, Pt/HxNbO5 $ Pt/Nb(OH)5). In other words, such state of the art
enables to perform (electro)catalytic processes from very high temperature down
at ambient conditions, simply by the wetness effect [5–11].
Hypo-d-electronic transition metal oxides usually feature several altervalent
states giving rise even to interactive mixed valence compounds, such as, for
example, TiO2/WO3, TiO2/Nb2O5, TiO2/Ta2O5, or TiO2/Nb2O5/CeO2, and then
correspondingly increase the overall latent storage and spillover effect of both
H-adatoms and primary oxides (M-OH). The whole spillover and SMSI effect
behave typical synergistic electrocatalytic properties and never any individual
hypo-d-oxide enables that much as mixed altervalent composites. Such coinciding
and interconnected events and phenomena have finally been perfectly and broadly
tuned in mutual interfering phase, almost like a lucky concatenation, and enabled us
to approach and substantiate the reversible electrocatalysts for the oxygen electrode
reactions.
The problem so far was in unattainable nanostructured Pt-bronze, the catalytic
activity of which exponentially increases with decreased Pt nano-size approaching
maximum at monoatomic dispersion [31]. This requirement has now been fulfilled
by the grafting implementation of Pt-acetylacetonate within colloidal particles of
peroxopolytungstic acid, niobia (Nb2O5), tantalia (Ta2O5) and ceria (CeO2),
(Fig. 8.3) [7, 8, 19, 20]. Such homogeneous and even distributions of nanostruc-
tured Pt particles, and such SMSI bonding effectiveness, so far were missing the
experimental evidence in PEMFCs development.
318 M.M. Jaksic et al.

8.2.3 Anodic HCHO Oxidation and Primary Oxide Spillover

Since the heterogeneous reaction of formaldehyde oxidation with Pt-OH, and in


particular Au-OH, proceeds as a fast reversible anodic process mass transfer
limited, and since HCHO is soluble in all ratios in aqueous media, the primary
oxide generation rate and its yielding spillover have primarily been investigated by
potentiodynamic spectra within the broader concentration range and between
hydrogen and oxygen potential evolving limits (Fig. 8.2a, b). Such for an order
faster anodic reaction is able to postpone within unusually broad potential range the
recombination of the primary (Pt-OH), into the more polarizable, more adsorptive
and more stable surface oxide (Pt¼O), (Eq. 8.2). Formaldehyde oxidation starts at
its reversible potential (0.32 V vs. RHE), merges with the second UPD desorption
peak of H-adatoms, and extends as an exaggerated broad twins peak all along the
anodic scan, nearly until the beginning of OER (Fig. 8.2a). In the same sense anodic
Pt-OH striping CO oxidation on composite hypo-d-oxides supported Pt or Pt, Ru
catalysts takes place even within the usual interval of UPD H-adatoms desorption
(see further downwards). In other words, Pt-OH arises available for reaction not
only within its nominal reversible adsorption/desorption peak limits in regular
mineral acid or alkaline aqueous solutions, but depending on the reactant
(HCHO, HCOOH, CO, etc.), its concentration, affinity and the actual reaction
rate, along a broad and extendable potential range. Meanwhile, as the link of DL
charging/discharging, it actually extends and appears available all along the whole
potential axis of cyclic voltammograms. Such an unusually broader charge capacity
area (Fig. 8.2a, b), usually features all the properties of typical under- and over-
potential oxidation (UPO, OPO) peaks, in particular when compared with cathodic
UPD properties of H-adatoms on various metals. Such specific cyclic
voltammograms clearly reveal the interference between Pt-OH and Pt¼O, and
testify for the reaction polarization of the latter along the broader potential range.
Stepwise extension of positive potential limits toward the oxygen evolution
reaction (OER), clearly shows (Fig. 8.2a, b) the absence and/or distinctly reduced
and postponed Pt¼O growth almost until oxygen starts evolving. Then, since
during the reverse potential scan toward the HER, as the result of the former,
there arises and finishes the Pt¼O desorption much earlier and of dramatically
reduced charge capacity, than nominally in the simple acidic or alkaline solutions,
the appearance of a characteristic sharp anodic current jump testifies for the
hysteretic (cathodic sweep, anodic peak) aldehyde oxidation within the former
nominal DL charging/discharging range, and this way reflects and features the
specific and highly reactive properties of the Pt-OH. In other words, as a corollary,
there is no anodic aldehyde, alcohol, their simple acids, and even CO oxidation, nor
the ORR, upon any M¼O (Pt¼O, Au¼O) covered metal surface prior to the
potential of molecular oxygen evolution, when the latter becomes broken, and
thence, occurs only upon prevailing Pt-OH spillover deposits.
Namely, since aldehydes are often soluble in aqueous media almost in all ratios,
their voltammograms at high contents feature imprinted extremely high both charge
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 319

capacities and limiting currents at their peaks (Fig. 8.2b), and thereby testify for
almost unlimited reversible reaction rates (Eq. 8.1), as long as diffusional mass-
transfer supply provides enough reacting species. Even more so, for interactive
supported Pt and Au upon higher altervalent hypo-d-oxides and their mixed valence
compounds, since these behave as highly enriched latent storage capacities of
primary oxide spillover sources [32–36].
This is the cause and reason why within the reversible part of Tafel plots
electrocatalytic metal (Pt,Au) surface is always covered by the interacting Pt-OH/
Pt¼O species, and naturally tends to impose the reversible oxygen electrode
properties, and these are the experimental evidence that the optimal catalytic rate
implies their optimal ratio.
In such a context, meanwhile, one of the most outstanding observations has then
been that cyclic voltammograms of both formaldehyde (Fig. 8.3, [37]) and formic
(muriatic) acid (Fig. 8.3, [38]) anodic oxidation distinctly differ upon plain (Pt/C)
and the same, but more or less enriched supports in specific amounts of interactive
hypo-d-oxide (Pt/Ta2O5/C [37] and/or hypo-f-oxide supports, Pt/CeO2/C [38]). In
other words, the all parameters and conditions being the same, the reaction rate
becomes dramatically different upon various supported electrocatalysts in their
hypo-d-oxide amounts and there resulting latent storage capacity, as the result of
distinctly different the additional Pt-OH spillover feeding and spreading effects
(Fig. 8.5) [32], and as the only distinctly imposed difference. Such a conclusive
observation belongs to the main experimental arguments to prove the theory of the
M-OH interfering self-catalytic spillover contributions in electrocatalysis of aque-
ous media, (Eq. 8.3) [5–11], finally providing the ROE behavior and properties, and
traced the way and entire approach for its substantiation. In such a respect, cyclic
voltammograms deeper enlightened and more clearly revealed the spillover reac-
tion impact properties of the Pt-OH.

8.3 Potentiodynamic Scans of Primary Oxide Latent


Storage and Electrocatalytic Spillover Spreading

Some intermolecular compatible hypo-d-oxide mixed valence architecture


(Pt/WO3,TiO2/C; Pt/Nb2O5,TiO2/C), (Fig. 8.3), as the interactive catalytic
submonolayer (and much more so as the multilayer, Fig. 8.5, [32]) supports of
high altervalent number or capacity, have been investigated by potentiodynamic
scans to reveal both the primary oxide latent storage and resulting Pt-OH spillover
yielding properties, along with same for H-adatoms. In this respect, cyclic
voltammograms scanned at low moisture content of He stream (just enough to
enable basic electrode processes to occur and proceed), insufficient for WO3
(or TiO2) hydration, repeatedly reveal similar potentiodynamic spectra character-
istic for indifferent carbon supported (Pt/C), or plain Pt itself (Fig. 8.3), but with
high double layer charging capacity, because of the accompanying parallel
320 M.M. Jaksic et al.

Fig. 8.5 Cyclic voltammograms scanned at the Pt/GC and Pt/TaOx/GC catalytic electrodes with
2c, 8c, and 20c (cycles) in Ar-saturated 0.5 M H2SO4 solution at scan rate of 50 mV s1, revealing
the effect of proportional increasing of interactive Pt supporting Ta2O5 deposit on the Pt-OH
spillover effect and growth for the ORR (courtesy of T. Ohsaka [32])

charging of Vulcan carbon particles beside the metal (with correspondingly large
charge value, QDL ¼ 1.07 C).
In contrast to such fairly common occurrences, a continuous supply of saturated
water vapor in the He stream at higher temperature (80  C), imposing condensation
(Boudart spillover precondition [27–29]), and leading to the appearance of wet
titania-tungstenia mixed valence oxide composite, along with spontaneous disso-
ciative adsorption of water molecules all over its exposed surface [21, 22], as the
interactive catalytic support, has been accompanied by the unusual phenomenon of
a dramatic blowing expansion of two reversible pairs of peaks of both the primary
oxide (QPt-OH(a) ¼ QPt-OH(c) ¼ 1.453 C), and H-adatoms (247 versus 47 mC cm2, or
in the ratio of about 5.3:1) chemisorptive deposition and desorption (Fig. 8.3), like a
DC capacitance of extremely developed electrode surface charging and
discharging. Since these are highly reversible and evidently behave remarkably
pronounced the latent Pt-OH storage (cf. [5–7, 21]), they keep the same their
extents even after multiple and repeating number of cycles at any other time. The
latter have both been of unusually high spillover charge and discharge capacity
values and for Pt-OH (UPD and OPD) shifted towards both much more negative
and far positive potential limits, in common with Fig. 8.2a, b and discussion
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 321

thereon. In fact, two distinctly different cyclic voltammogram shapes and charge
capacities (Fig. 8.3), appear only as the result of the difference in water vapor
supply, all other parameters being unaltered the same, and as the effect of the
equivalent dipole (Pt-OH) charging and discharging of the double layer, since
nothing else takes place in between. Every cessation in the steam supply instanta-
neously imposes the sudden reversible shrinkage of both such rather exaggerated
pairs of peaks down to the same initial potentiodynamic shape similar to the
nanostructured plain and non-interactive supported Pt/C voltammogram spectra
themselves. Vice versa, the renewed saturate water vapor feeding immediately
leads to their former Pt-OH peaks and the same former charge capacities; namely,
the effect later already noticed and scanned for formaldehyde [37] and formic acid
[38] oxidation. Such an appearance without exception behaves as a typical revers-
ible transient phenomenon by its endless altering repetition [5–7], and never
appears upon the plain Pt/C electrocatalyst, both wet and/or dry, nor with small
and insufficient amounts of catalytic hypo-d-oxide supports (Fig. 8.5, [32]).
The complementary interactive Ta2O5-based electrocatalytic support strongly
reinforces just displayed potentiodynamic features of Pt/WO3,TiO2/C by coincid-
ing and congenial spectral behavior of their cyclic voltammograms: The distinct
growth of Pt-OH and H-adatoms adsorption and desorption peaks, reflecting their
different accumulated latent charge storage capacities, as a function of the amount
(charge density) of interactive composite hypo-d-oxide deposits per unit electrode
surface (Fig. 8.5) [32]. In such a respect, for example, as the effect of much smaller
d-ionic radius, Y2O3 much more does so, than Nb2O5, or WO3. In fact, such Pt-OH
latent storage growth (including the corresponding spillover effect) does not extend
endlessly and usually passes over remarkably pronounced maximum in the ORR
catalytic rate and activity (see Fig. 8.11, [14]).
What is now the substantial difference between voltammograms in Fig. 8.3, wet
state, and Fig. 8.2a, b (or Fig. 8.5)? Even when mostly suppressed in the surface
oxide (Pt¼O) adsorptive growth (Fig. 8.2a, b), the reversal backward cathodic
scans on plain Pt proceed still highly polarized for about 600 mV with negligible
or zero current, exactly corresponding to Fig. 8.1. In other words, there is still no
stored Pt-OH on plain Pt to start either hysteretic HCHO oxidation peak or the self-
catalytic interfering reaction (Eq. 8.3) of the ORR. However, on the hypo-d-oxide
interactive supported Pt catalyst, the a priori latently accumulated initial storage of
the primary oxide from the very beginning is already ready and available, and
thence continuously provides and spillover enhances the latter to proceed as the
uninterrupted and extra fast reversible electrode reaction, as long as there is
continuous water vapor supply. However, for the plain Pt (or Pt/C), the sudden
hysteretic sharp anodic current jump in the course of reversal cathodic sweep
(Fig. 8.2a, b) rearises at and coincides with the classical position of the reversible
peak for the Pt-OH growth, reflects the local interfering self-catalytic effects at such
potential range, provided by the repeated Pt-OH spillover growth (Scheme 8.1).
This is the striking point and the core substance of the present study, while Fig. 8.3,
supported by Fig. 8.5, are the best illustrative issues of the substantiated reversible
electrocatalyst for the oxygen electrode reactions (ORR, OER).
322 M.M. Jaksic et al.

Scheme 8.1 Visual


presentation of novel
spillover latent storage
Pt-OH type interactive
supported electrocatalyst
for oxygen electrode
reactions, exampled for
Pt/Nb2O5,TiO2/Nb(OH)5,Ti
(OH)4 issue (Reprinted with
permission from
[7]. Copyright (2014)
American Chemical
Society)

8.4 Interchangeable Interfering Reversible and Polarizable


Oxygen Electrode Reactions

In general, nanostructured Pt (Au) electrocatalyst selectively interactive grafted


upon hypo-d-(f)-oxide supports, when the latter exist in the hydrated external
surface and/or internal bulk state (Scheme 8.1), thence consequently exhibits
enriched Pt-OH (Au-OH) latent storage, as the feedback oxophilicity effect and
property. The corresponding primary oxide then continuously features renewable
spillover, simply by water vapor supply (surface phenomena, no concentration
polarization), and further continuously yielding spontaneous dissociative adsorp-
tion of aqueous species, so that the reversible anodic oxygen evolution (Eq. 8.5)
now takes place straight from pronouncedly high and also the highly reversible
Pt-OH (Au-OH) adsorptive peak capacities, (Figs. 8.3 and 8.5) [5–7],

4Pt-OH $ O2 þ 2H2 O þ 4Pt þ 4e ð8:5Þ

and, thence, at the thermodynamic equilibrium (ROE) potential value. This is the
substance, this is the point and evidence, this is the achievement of the substantial
reversibility in the oxygen electrode reactions. Quite on the contrary, in classical
issues (Fig. 8.1), the latter occurs from the strong irreversibly deposited and highly
polarizable monolayer of the Pt¼O, (Eq. 8.2), and thereby, while being deprived
from primary oxide presence or external supply, at rather high anodic
overpotentials,
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 323

2Pt ¼ O ! O2 þ 2Pt þ 4e ð8:6Þ

The same, vice versa, consequently alternately occurs both in these reversible
and/or irreversible states, as concerns the ORR, all along the reverse cathodic
scan and polarization (Figs. 8.1, 8.3 and 8.5) [5–7]. The transient between the
reversible and irreversible status features some sharp thresholds both as a function
of hypo-d-(f)-oxide unit content and/or wetness percentage [5–7], and substantially
reflects the Pt-OH (Au-OH) properties (present or absent) for the oxygen electrode
reactions. In such a context, plain and non-interactive supported Pt (Pt/C), in the
absence of primary oxides, and thereby established or a priori existing the broad
reaction polarization range, by no means can feature the overall reversible oxygen
electrode properties along the whole potential axis, in between of the hydrogen and
oxygen evolving limits, or within the entire potentiodynamic circle. Thus, there
exists either the reversible behavior in the presence or continuous external supply of
primary oxides, or highly irreversible properties in the absence of the latter,
(Pt-OH ! 0, Au-OH ! 0), and this is the substance for Pt and Au electrocatalytic
properties!
In such a context, the simple stoichiometric combination of Eqs. 8.1 and 8.5,
reveals that the overall reversible anodic oxygen evolution initiates from water
molecules (Eq. 8.1), and obeys the auto- or self-catalytic Ertel framework [30], with
the general oxidation mechanism based on and catalyzed by M-OH species, or the
primary oxides as the main interfering electrocatalytic species, and substantially
takes place at the reversible oxygen potential value; the same in the reverse
cathodic ORR scans. However, when oxygen evolving initiates from Pt
(Au) monolayer covered by the strongly adsorptive and polarizable surface oxide
(Pt¼O, Au¼O), the anodic, and reverse cathodic reaction have both to overcome
remarkable overpotentials (Eq. 8.6, adjoined with Eq. 8.2). This is the substance
and difference as concerns the irreversible plain (Pt, Au) and non-interactive
supported Pt/C(Au/C), and/or the enriched latent storage and continuous external
spillover feeding of the primary oxide for interactive hypo-d-(f)-oxides supported
Pt electrode. The fascinating and incredible surprising facts, meanwhile, are that
such simple distinct phenomena have been obscure and unknown in electrocatalysis
for many decades now, mainly because of longer missing of cyclic voltammetry.

8.5 Electrocatalytic Consequences

Since Pt(Au) electrocatalyst becomes interactive d-d-(f)-bonded with and grafting


fixed upon hypo-d-(f)-oxide type catalytic supports in new composite nano-
structure, there is no more its metallic nano-particles (cluster) surface diffusion
and agglomeration, nor any Nafion membrane cross-over of hydrogen. Even more
so, from the same reasons, the life-time of such Pt(Au) (electro)catalysts is at least
twice longer and can be guaranteed. Alterpolar spillover interchanges in the
324 M.M. Jaksic et al.

revertible, PEMFC versus WE, proceed reversibly and instantaneous smoothly, and
enable their superior unique operation. There is a great number of parametric and
structural variability for the voltage/current optimization and stability, such as
hypo-d-oxide radius of small, medium and large d-(f)-metals (Y, Ti, Nb, Ta, W,
Ce, Gd, Ho, La, etc.), to keep low Tafel slope within the operating range,
0.0–1.0 A cm2. Since some of hypo-d-(f)-oxide supports of composite
electrocatalysts (Nb2O5, Ta2O5, WO3, TiO2) feature advanced electron conductiv-
ity, nano-particulate carbon carriers and current collectors can be completely
avoided that is substantially significant for WE and mostly for anodic oxygen
evolution. In such a respect, the present paper defines the main substantial and
advantageous frameworks in electrocatalysis of reversible electrocatalysts for oxy-
gen electrode reactions (mostly ORR and AOE), primarily for L&MT PEMFCs,
WE, and their revertible combination, and as the overall main conclusion, hopefully
means one of the most important contribution all along after Sir William Grove. All
stated herein for Pt, means even more so for Au when oxygen electrode reactions
are in consideration (ORR, OER) [6, 31, 39–42].
Mixed altervalent hypo-d-(f)-oxide supports, depending on their preceding ther-
mal treatment, feature advanced electron conductivity from 1000 S cm1, for
Magneli phases, down to about 300 S cm1 for average composites, and this way
enable to replace and even completely remove nanostructured carbon as a
non-interactive catalyst support and current collector. However, except for the
OER, when carbon becomes partially the subject of anodic oxidation, nano-
particulate carbon still quite satisfies requirements for its nominal purposes.
In fact, the main benefit comes from the dissociative adsorption of water
molecules upon mixed valence hypo-d-(f)-oxide supports, enabling us to increase
the latent storage capacity, and therefrom the yielding Pt-OH (Au-OH) spillover
intensity, and/or the higher current density at the more reversible electrode poten-
tials. In the same respect, the supercritical drying within the sol-gel procedure for
hypo-d-(f)-oxide supports development enables to extend remarkably the available
surface for dissociative water molecules adsorption and consequently enlarged the
Pt-OH latent storage and spillover intensity.
The actual attainable target for LT PEMFCs: Tafel slopes of 60 mV/dec, or
η ¼ 180 mV at 1000 mA cm2, or ideally 30 mV/dec, or 90 mV at 1.0 A cm2.

8.6 Hypo-Hyper-d-d-Interelectronic Bonding


and Interfering Spillover Nature of Electrocatalysis
for Hydrogen and Oxygen Electrode Reactions

The majority of physical and chemical properties of elements behave typically


symmetric volcano plots along transition series, observed and well known as
Friedel d-d-electronic correlations [14, 43–48]. These properties substantially
reflect the characteristic periodicity features of d-metals, and when plotted one
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 325

into another, result with simple straight line dependences of deeper fundamental
scientific significance [14, 46–48]. In such a context, one of the most significant
linear interdependences arises as the so called one by sixth law between surface free
and cohesive energies [49], and in a straight forward manner connects them both to
the catalytic activity properties, since the former directly correlates with the
adsorptive bonding strength (or, the free enthalpy of adsorption as the driving
force) of reacting intermediates in the RDS. In the same respect, asymmetric
volcano curves of the electrocatalytic activity (log jo) for the cathodic hydrogen
evolution (HER), (Fig. 8.1, [50]) [43, 51], and/or the free enthalpy of H-adatoms
adsorption, and work function (Fig. 8.2, [14]), [46–48], all being with optimal peak
values of d8-electronic configuration, when plotted with each other, also reveal a
straight line dependence of most significant and substantial meaning in
electrocatalysis. While Trasatti [52] came to such a linear dependence from some
theoretical analysis, our approach follows the overall general fundamental proper-
ties of volcano plots along the transition series also associated with and based on the
d-band state or the d-energy level of individual d-elements [14, 43, 46, 48]. Even
more so, since the fundamental Butler-Volmer equation in electrode kinetics obeys
the typical Maxwell-Boltzmann law of exponential relation between current density
( j), as the reaction rate intensity feature, and polarization or the overvoltage (η), as
the electrode reaction driving force, logarithm of the exchange current ( jo) obeys a
typical linear function dependence of the free activation enthalpy of adsorption of
intermediates at the RDS (log jo  koΔHo) [46–48, 53]. In other words, in more
general form, log j  kΔΗ, and when η ) 0, then j ) jo, and thence, these interre-
late with other volcano plots on quite general basis and simple linearization manner.
Actually, the optimal d-electronic configuration peak for asymmetric volcano plots
introduces a simple kinetic constant correction term, which shifts the initial (or,
real) volcano curves to their nominal symmetric position [14, 43, 46, 48]. In the
same respect, Brewer’s [13, 14] intermetallic bonding theory predicts and con-
firmed that every hypo-hyper-d-d-interelectronic combination (or, the local phase
diagram within the Periodic Table of transition elements) implies even much more
pronounced intermetallic bonding effectiveness, consequently remarkably further
extended their common volcano plots and correspondingly extra high intermetallic
phase stability at peak values of such cohesive energy curves [14, 43, 44, 46, 48,
53]. In other words, every hypo-hyper-d-d-interelectronic phase diagram (Fig. 8.6),
behaves for various physical property plots as a part of the Periodic Table, with
local volcano curves composed from intermetallic phases along the axis in between
that (energetically or by the average d-electronic configuration) correspond to the
set of elements located within the periods of two initial constituents (ingredients) in
the local phase diagram consideration. Thence, as the overall general meaning of
the hypo-hyper-d-d-interelectronic correlations [44], such local intermetallic vol-
cano curves, when plotted one into another, yield again the same type of straight
line interdependence and various characteristic maximum values correspondingly
coincide with each other [43, 46–48]. In such a context of coherent mutual
consequences, the intermetallic Brewer peak bonding effectiveness values corre-
spond to the maximal stability, the highest interatomic and orbital symmetry, and
326 M.M. Jaksic et al.

3 2
a
1850 100 10-1

Ni Zr
Ni Zr 3
1750

Ni5Zr 7 2
1 3
10-1 10-2

NiZr
NiZr2
1500

io/Acm-2
j/Acm-2
T/°C

2 10-2 10-3

1250 4
10-3 10-4

1000
10-4 10-5

750
Zr 10 30 50 70 90 Ni
at, % Ni
b

Fig. 8.6 (a) Electrocatalytic activities of various intermetallic phases (polished below 1.8 in the
roughness factor) along Zr–Ni phase diagram (curve 1) for the HER, taken as the exchanged
current ( jo, close circles, curve 3) and relative current density changes ( j, open circuits, curve 2) at
constant overvoltage (0.2 V), plotted together with the cohesive energy of intermetallic phases
(curve 4). (b) Tafel plots for the HER with indicated extrapolations to assess the relevant exchange
currents, jo; for (a); (c) linear interdependence between cohesive energy (Ec) and electrocatalytic
activity (log jo) for the HER
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 327

c -1.5
ZrNi3
-2

-2.5
Zr2Ni7
ZrNi
log j0

-3
Zr2Ni

-3.5
ZrNi5
- 4

-4.5
370 380 390 400 410 420 430 440
Ec / kJ . mol-1
Fig. 8.6 (continued)

consequently, both the maximal work function values and synergistic


electrocatalytic activities for the hydrogen electrode reactions. Thus, cohesive
and catalytic features are brought in the simple common linear interdependence
with the coinciding maximal values: the stronger the bonding, the higher the
electrocatalytic activity [14, 43, 46–48]. In such a sense, some specific overall
maximum belongs to the HfPd3 symmetric intermetallic phase [13, 14, 43]. Such a
theoretical approach yields the basic guide line in electrocatalysis for hydrogen
electrode reactions amongst individual and hypo-hyper-d-d-interelectronic and
intermetallic composite electrocatalysts, both for macroscopic issues and nano-
structured clusters.
Where then lays the link between the hypo-hyper-d-d-interelectronic bonding
and SMSI synergistic electrocatalytic effects in the broader interfering spillover
sense? The whole electrocatalytic theory [14, 43, 46] relies on the Brewer [13, 14]
intermetallic bonding model and Friedel [44–46] volcanic tape hypo-hyper-d-d-
electronic correlations. They both inferring that the stronger the hypo-hyper-d-d-
intermetallic cohesive bonding, the more stretched, more strengthened and more
exposed arise d-orbitals within the ideal symmetric intermetallic phases, like ZrNi3,
HfPd3, MoPt3, LaNi5, LaxCe(1-x)Ni5, etc., and thereby, the weaker and easier the
cleavage of their adsorptive intermediates (Pt-H, Pt-OH) in the RDS, and conse-
quently, the higher the reaction rate and the overall catalytic activity [5, 9–11, 43,
46, 48]. The same Brewer [13, 14] type d-d-intermetallic bonding model has been
much earlier anticipated and means the preceding basis for Tauster [15–18] pro-
motional SMSI effect, with the far-reaching consequences in both heterogeneous
catalysis and electrocatalysis. The latter systematically predetermined interactive
grafting [9, 10, 19, 20], and the homogeneous, evenly uniform distribution of
individual and prevailing hyper-d-metallic catalysts upon hypo-d-oxide supports
[5–11]. The same Brewer type [13, 14] of hypo-hyper-d-d-interactive bonding
328 M.M. Jaksic et al.

between nanostructured metal particles of such composite catalysts (MoPt3,TiNi3),


and their hypo-d-oxide supports, additionally reinforces the entire electrocatalytic
activity effect based on the overall interactive d-d-interelectronic bonding strength,
similarly as cohesive energy itself defines and advances the electrocatalytic activity
for nanostructured mono-phase systems [5–7, 19, 20, 43]. In fact, the interactive
hypo-hyper-d-d-interelectronic bonding strengths belong to the strongest bonding
effectiveness in the whole chemical science [5–11, 19, 20], sometimes even pro-
ceeding so vigorously with exceptionally intense explosion (HfPd3), and thereby
resulting with extra high intermetallic phase stability [13, 14, 46, 48], and conse-
quently, high electrocatalytic activity for the HER. In other words, the
interelectronic-d-d-(f)-bonding and/or cohesive strength linearly correlates with
the (electro)catalytic activity, [46, 48]. The pronounced cathodic and anodic yields
of interactive spillover contributions within and based on the SMSI, have addition-
ally been significant for the present theory and its embodiment in electrocatalysis of
both hydrogen and oxygen electrode reactions, particularly for Low and Medium
Temperature (L&MT) PEMFCs and WE [5–11, 19, 20]. While in aqueous media
plain nanostructured Pt (Pt/C) features the catalytic surface properties of the active
Pt-H and polarizable Pt¼O, missing any effusion of other interacting species, a new
generation of composite interactive supported (SMSI) electrocatalysts in condensed
wet state primarily characterizes exceptional and extremely fast reversible spillover
interplay of either H-adatoms, or more important the primary oxides (Pt-OH,
Au-OH), as the unique significant and advanced synergistic interactive
electrocatalytic ingredients [5–11, 19, 20].

8.7 Volcano Plots: Their Sense, Properties, Significance


and Uses

Volcano plots of various physical and chemical properties along transition series
reveal the periodicity features of d-metals [14, 44–48], based on the d-d-electronic
correlations [44], with similar symmetric shape, and vice versa. Periodicity features
group transition series of elements into congenial, parallel and symmetric volcano
plots. In the essence, such symmetric volcano plots draw their basic roots from the
hypo-hyper-d-d-electronic variations in the elemental configuration and the
resulting individual atomic energy level states. The ascending part of such volcano
curves comes from hypo-d-elements while filling their semi-d-orbitals, whereas the
descending half reflects hyper-d-electron pairing and yielding the anti-bonding
d-band along with still remaining unpaired semi-d-orbitals, to balance d-states
periodicity on both volcano sides, ascending and descending, along transition
series. Substantially this is the basic reproducible d-electronic symmetric volcano
plot we always imply and start from. All other physical and chemical volcano type
properties simply come as the reflection, reproduction and consequence of such
initial and substantially involved basic hypo-hyper-d-d-electronic distribution of
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 329

individual d-electronic level state of d-metals. Thence, as the consequence of


symmetry and similarity, every common plot of one into other periodicity state,
together yields various linear interdependences of deeper physical meaning and
significance, like cohesive and surface free energy [46–48], or even more so, the
cohesive bonding strength into electrocatalytic activity, and the latter straight into
the surface and/or adsorptive free energy, etc., further unlimited in the general and
universal properties of the matter! As a consequence, the d-band of transition
elements has been confirmed to play a crucial decisive role in the overall behavior
of transition elements, and thereby, is considered for the bonding, adsorptive,
catalytic and electrocatalytic band of transition series [46, 48]. Thus, consequently,
any search for advances and synergism amongst the latter, should be based on the d-
d-interelectronic correlations and modifications [14, 43–48]. Such a state of theo-
retical knowledge and experimental evidence, established on the symmetry basis of
elemental properties, consequently leads to the general statement and universal
conclusion that every hypo-hyper-d-d-interelectronic phase diagram behaves as the
part of the Periodic Table between the two initial periods of the interacting
ingredients, with their intermetallic phases taken for the axis [43, 46–48],
(Fig. 8.6) [14]. The latter means and implies with no exception, so that the partial
volcano plots in the small [14, 15, 44–48], bear a broad and deep catalytic, and in
the substance general physical sense and meaning. Their intermetallic phases in
between two initial ingredients are of the same average d-electronic configuration,
(Fig. 8.6), that way replacing the ‘missing elements’ in between, in their energy
state, level and behavior [46, 48], and consequently, have been used to assess the
synergistically maximally active electrocatalysts for the HER. Meanwhile, since
transition elements themselves establish and reproduce simple individual funda-
mental symmetric ascending/descending volcano plots, depending on basic indi-
vidual d-electronic number of evolved d-electrons along transition series and their
individual configuration of d-band, hypo-hyper-d-d-interelectronic combinations
involve antibonding pairs and consequently their volcano curves are usually asym-
metric (Fig. 8.6) These theoretical and experimental facts can then be employed to
predict the synergistically active electrocatalysts from the peak values of
corresponding local volcano plots along each hypo-hyper-d-d-interelectronic
phase diagram [46–48], and then further catalytically advance by the interactive
hypo-d-oxide interfering spillover contributions (Pt-H for hydrogen, and Pt-OH for
oxygen electrode reactions) [5–11], displayed more systematically and more thor-
oughly in the present and preceding study [7]. Even more so, maximal synergisti-
cally active electrocatalysts can be estimated indirectly with a rater high certainty
and probability, straight from alternative secondary volcano plots for various
correlating physical properties, on the first place now from cohesive energy,
Fig. 8.6 [14, 46, 48]. On the same way Fermi energy plots fundamental elemental
simple symmetric volcano curve dependence along transition series, thence on a
straightforward linear manner correlates with the electrocatalytic activity for the
HER, and since the former is in the linear interdependence with the total cohesive
energy, the overall functional correlation results straight one from another. Thus,
one of the guiding aims should have to be to assess experimentally or theoretically
330 M.M. Jaksic et al.

most promising synergistic electroctalysts, and consequently correlate the actual


hypo-hyper-d-d-interelectronic SMSI effect or hypo-hyper-d-d-interbonding effec-
tiveness for new type interactive supported composite (Pt/Nb2O5,TiO2/C, Pt/Ta2O5,
TiO2/C, or more complex mixed valence compounds for the HER, HfPd3/Nb2O5,
TiO2,CeO2/C) electrocatalysts. The formal difference is that interactive bonding is
interatomic or intermetallic (Brewer-Friedel correlations), while in the more com-
plex and broader Tauster extension SMSI sense [15–18], they are interionic-
intermetallic interactions, or all together being hypo-hyper-d-d-interelectronic
bonding correlations in electrocatalysis. Thus, the interfering Brewer-Tauster
(SMSI) hypo-hyper-d-d-(f)-interbonding effect, along with the resulting primary
oxide (Pt-OH,Au-OH) latent storage and spillover properties, bring electrocatalysis
for hydrogen and oxygen electrode reactions to the common point for mutual
optimization in (electro)catalytic activity.
All attempts to construct volcano plots in heterogeneous catalysis along any
energetic (adsorptive) axis so far failed. Since all (electro)catalytic processes are
governed by, and as the subject matter of the Maxwell-Boltzmann general kinetic
and distribution law, imply and exceptionally impose the exponential dependence
and relations, then consequently, the logarithm of reaction intensity (log
jo ¼ koΔHo) results as a linear, but by no means volcanic function [14, 43, 46–
48]. Volcano plots display without exception the periodicity features along transi-
tion series of elements as a function of d-electron configuration, or, d-electronic
density of states at the Fermi level, which itself reveals the fundamental symmetric
volcanic parabola of hypo-hyper-d-d-electronic structure of individual transition
elements, and reproduces endlessly itself in all physical and chemical property plots
along the Periodic Table [43–48].

8.8 Electrocatalysis and d-Electronic Density of Sates


Versus Fermi Level

The electrocatalytic reaction mechanism by contemporary ab initio DFT calcula-


tions, both for the HER and ORR, implies that a d-band centered near the Fermi
level (EF) can lower the activation energy as the bonding orbital passes EF, the
critical step for reduction processes taking place when the antibonding orbital
passes the Fermi level of the metal from above and picks up electrons to become
filled [54]. In electro-oxidation reactions (HOR, OER), the bonding orbital passes
the Fermi level from below and gets emptied. Consequently, a good catalyst for
these reactions should have a high density of d-states near the Fermi level [40, 55,
56]. Thus, the present concept consists of the proper hypo-hyper-d-d-interelectronic
combinations of transition elements both being of rather high densities of d-states at
the Fermi level (Fig. 8.7a), in which hypo-d-oxide components (W, Nb, Ta, Ti),
besides the substantially high SMSI bonding effect, as typical oxophilic d-metals,
involve their pronounced surface type membrane migration properties for the
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 331

Fig. 8.7 (a) Surface densities of d-band states of some selected hyper-d-electronic transition
metals. The integral over the densities has been normalized to unity; the vertical line indicates the
Fermi level. Indications and labels: dashed-dotted line: Pt(1 1 1); thick line: Au(1 1 1); thin line: Ag
(1 1 1); dotted line: Cu(111); dashed line: Ni(111). (b) Surface densities of d-band states (bulk
values differing from exact surface densities for less than 3 %) of some selected hypo-d-electronic
transition metals (W, Ta, Mo, Nb, color labeled) (Calculated by Professor Wolfgang Schmickler,
University of Ulm, Germany)

primary oxide latent storage, transferring and effusion. In such a state, our aim has
been to keep the composite transition element ingredients with their initial high
densities of d-states, or even to increase the latter by their SMSI interbonding effect,
and at the same time to use the benefits of the primary oxide spillover for the overall
reaction, in particular for the ORR. In this respect, Fig. 8.3b reveals why Nb and Ta,
by the position in the Periodic Table of elements, are electrocatalytically
predestined for even higher activity than W, Mo and Ti.
332 M.M. Jaksic et al.

8.9 Significance and Experimental Evidence for the SMSI


d-d-(f)-Interbonding Effect

The evaluation of Tauster Strong Metal-Support Interaction properties (SMSI) [15–


18], as the d-d-interelectronic bonding effect in thereby advanced catalysis, other-
wise necessary to impose membrane ionic transport, too, and finally yielding with
spillover of dipole type RDS intermediates in (electro)catalysis, implies some
experimental evidence of deeper fundamental theoretical meaning and significance.
In such a context, to study, get some insight in, and assess experimentally the SMSI
bonding effect itself, and further imposed its resulting spillover phenomena, very
fine, nano-sized Au films were deposited by controlled electron beam evaporation
of ultrahigh purity gold metal under high vacuum onto stationary, nano-crystalline
anatase titania (1 0 1) coated microscopic slides [5, 6]. The XP spectra of the Au 4f
electrons reveal the remarkable binding energy shift, (Fig. 8.8), that provides the
insight for the d-d-interelectronic SMSI within the Au/TiO2 interphase and this has
been one of the first reliable precise experimental evidences of that kind in
heterogeneous catalysis. The smaller the nano-particle size or the thickness of the
Au nano-layer, the larger the binding energy shift in the XP spectra with titania, and
the more pronounced the d-d-SMSI effect, with tendency to its maximal d-d-
binding strength at monoatomic dispersion (as extrapolated straight assessed
value), and for 1:1 deposition of Au on available Ti atoms, as theoretically
predicted in the present and preceding papers [8–11, 19, 20, 31]. Due to the nano-
dimensions of the Au layer, the signal originated mainly from Au/TiO2 interface,

Fig. 8.8 XP spectra of Au 4f for vapor deposited nanolayered Au upon a fine thin films of anatase
titania (1 0 1) with deconvolution for lower amounts of deposits to reveal the existence of primary
oxides (Au-OH an AuOOH) [6]
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 333

reflects the bonding status within the latter. In other words, the thinner the nano-
layer, the closer to the interphase itself (Au/TiO2) penetrate spectral beams and
thereby, better and more completely reflect the bonding status within the latter, this
way enabling to get some insight and feeling as concerns the hypo-hyper-d-d-
interelectronic SMSI bonding effect, which is apparently very strong.
The deconvoluted Au 4f peaks with lower Au loadings reveal that Au
nanoparticles in interactive bonding contact with titania appear partially oxidized
[57, 58]. The peak located at 82.15  0.1 eV is attributed to metallic Au, while the
peak at 84.05  0.1 eV corresponds to the gold primary (Au-OH or AuOOH)
oxides. The latter, in accordance with the present theory, appear as the a priori
naturally provided primary oxide latent storage spillover species, associated with
and promoted by the wet anatase titania interaction (Eq. 8.1), and are in advance,
already available and ready for the cathodic ORR and anodic CO, aldehyde group
and other oxidation processes.
Starting with their plain combinations (Au/C versus Pt/C) of nanostructured
electrocatalysts, and in particular with interactive SMSI supported bronze type
(Au/Nb2O5,TiO2/C versus Pt/Nb2O5,TiO2/C) issues, Au is definitely much better
and advanced nano-sized catalyst for the ORR than Pt, primarily because of
remarkably smaller adsorptive strength of corresponding spillover Au-OH versus
Pt-OH reacting species [41].
Haruta et al. [41, 59, 60] have, for example, in such a sense shown that the same
reactants (propylene in admixture with equimolar amounts of hydrogen and oxy-
gen) yield different products upon different nano-sized Au catalysts supported on
anatase titania (Au/TiO2): (i) propane by hydrogenation at nano-particles <2 nm
Au, and (ii) Propylene oxide by epoxidation or oxygen addition for >2 nm
Au. Hydrogenation implies H-adatoms adsorption on Au that should not spontane-
ously occur on any surface of pure massive bulky gold. Haruta [41, 59–64] ascribes
such chemisorptive properties to “forced” or strained Au-d-orbitals within smaller
(below the critical threshold) nanostructured metal particles, in particular when d-d-
SMSI deposited on the interactive anatase titania (cf. [6]). In other words, smaller
nanostructured Au particles (<2 nm), interactively d-d-bonded with anatase titania,
or in particular when interactive supported upon mixed valence hypo-d-d-com-
pounds (Au/Nb2O5,TiO2, Au/Ta2O5,TiO2, or Au/WO3,TiO2), thereby being even
more reinforced in their d-d-interbonding effectiveness, and thence exposed with
the more strained d-orbitals, are qualitatively different than massive Au. Such
highly dispersed nano-particles consequently behave H-adatoms adsorption, and
thus feature provided the reversible behavior of hydrogen electrode in the Nernst
sense [6], and finally, is thereby able to carry out the hydrogenation processes, the
property so far unknown for gold at all. Such defined lower size (<2 nm) Au
nanostructure is something else both in heterogeneous catalysis and particularly
electrocatalysis, than massive Au (“the most noble and most inactive noble metal”
[40–43], primarily because of featuring spontaneous H-adatoms adsorption
[6, 43]. In the same context, it has been further shown that self-reconstructed Au
electrode surface, after multiple potentiodynamic cycles between hydrogen and
oxygen evolving limits, features the pronounced H-adatoms adsorption and even
334 M.M. Jaksic et al.

absorption, and consequently, the reversible electrode properties ([6] and references
therein). The reconstruction effect has been much more pronounced and faster in
heavy water media, because of the stronger interatomic interrelations and deeper
penetrating interphase effect of twice larger deuterium than protium ions and atoms
[6]. Finally, the advanced electrocatalytic properties of sol-gel prepared nanostruc-
tured Au for the HER, and extra superior for the ROE, and even Ag (Ag, AgNi3,
AgNiCo2) for alkaline media, after homogeneous interactive SMSI grafting depo-
sition upon carefully super-critical dried hypo-d- and hypo-d-f-mixed valence
compounds in liquid CO2, (Au/Nb2O5,TiO2/C, or Au/Ta2O5,TiO2,CeO2/C), follow
in the forthcoming studies. The main point, meanwhile, has here substantially been
to show that XPS measurements help approaching and the estimation of the hypo-
hyper-d-d-(f)-metal (Au) versus corresponding oxide SMSI inter-bonding effec-
tiveness, at least in the order and as the limiting values.
In the same context, one of the striking issues and contributions of the present
paper is that every hypo-hyper-d-d-interelectronic bonding effectiveness exceeds
the entering value of individual transition metal cohesive energy strength, and in
particular all hypo-hypo-d-d- and/or hyper-hyper-d-d-interelectronic bonds! This is
the point and substance of the advanced interactive bonding of hypo-d-d-(f)-oxide
supported hyper-d-electronic (Pt, Au, Ru) electrocatalysts.

8.10 Stepwise Survey of the Theory Development


and Experimental Evidence of Its Confirmation

Although Pt best satisfies electrocatalytic requirements of PEMFCs for hydrogen


electrode reactions (HER), but many more common transition elements (Ni,Co),
too, both common energy chart diagrams (Fig. 8.9), and interconnected cyclic
voltammograms (Fig. 8.1) heuristically show that the reversible behavior for
cathodic oxygen reduction (ORR), like a broad plateau in the energy conversion,
starts practically within the potential range of the reversible peak of the Pt-OH
desorption and consequently, overall loses within the preceding critical potential
range of the reaction polarization amount for more than half of thermodynamic
available values; or about and even more than 600 mVs. The main energy barrier
obstacle arises as the imposed reaction polarization between the reversible adsorp-
tion/desorption peaks of the primary oxide growth and removal (Eq. 8.1), and
anodic oxygen evolution, in both potential scan directions, as the result of irrevers-
ible disproportionation (Eq. 8.2), and the overall transition of Pt-OH (Au-OH) into
highly polarizable and strongly adsorptive surface oxide (Pt¼O, Au¼O). Thus,
ever since Sir William Grove [1] invented fuel cells (FCs), the electrocatalytic
search for the reversible oxygen electrode (ROE), and most particularly the ORR,
became the main challenge and imperative target in electrochemistry of aqueous
and PEM media. The substance is to replace the potential range of the strong
reaction polarization of the monolayered surface oxide deposit (Figs. 8.1 and 8.9)
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 335

Fig. 8.9 Classical and 1.4


characteristic energy chart Theoretical Cell Potential (1.23 V at 25°C)
diagram for nanostructured 1.2
Pt/C electrocatalyst in LT
PEMFCs (Reprinted with 1.0

Cell voltage / V
permission from Kinetics losses (mostly at cathode)
[7]. Copyright (2014) 0.8
American Chemical Ohmic losses
Society)
0.6

0.4

0.2 Concentration
losses
0.0
0.0 0.2 0.4 0.6 0.8 1.0
Current density / A cm-2

by the reversible (electro)catalysis of their ideally equimolar concentration ratio


(Pt-OH:Pt¼O  1:1, Eq. 8.3), and make Pt(Au) catalytically active all over the
potential axis between hydrogen and oxygen evolving limits, back and forth,
otherwise Pt(Au) is for a broad potential range substantially and highly polarizable
electrode even when with extremely developed surfaces.

8.10.1 General Theory and Its Substantiation

Under such sharply defined polarization circumstances of a priori existing and


broadly imposed the strong reaction polarization range of the primary oxide
(Pt-OH) disproportionation into the irreversible, rather stable, strongly adsorptive
bonded and highly polarizable (Eq. 8.2) surface oxide (Pt¼O), electrocatalytic
activation certainly implies the external reversible latent storage and continuous
spillover supply of the missing sources of the former.
In such a respect, polarization properties of some typical and characteristic novel
interactive hypo-d-oxides supported nanostructured Pt electrocatalysts for both the
ORR and OER [5–8], enable to select some diagnostic kinetic criterions on the way
towards the ROE, revealing from their Tafel plots: (i) nano-dispersed Pt/C clusters
(10 wt% Pt) non-interactive adhering upon sol-gel developed indifferent nano-
particulate E-tek, Inc., Vulcan-XC-72 carbon (240 sq m/g) carrier and current
collecting species, considered for the classical issue of such an electrocatalytic
activity comparison; (ii) The interactive supported nanostructured Pt particles upon
both the extra stable and electron conductive (300–1000 S/cm), ceramic Magneli
phases (Ebonex®), TinO(2n1), in average Ti4O7, usually defined as a shared rutile
structure, accommodating the oxygen suboxide deficiency in the structure by the
formation of crystal shared planes along the nth layerlike plane of octahedron, so
that Ti4O7, has one TiO for every three TiO2 layers; (iii) Advanced interactive
336 M.M. Jaksic et al.

1.05
0.5 mol dm-3 HClO4
1.00
250C
0.95
1.57 mA cm-2
0.90
E / v vs. RHE

0.85

0.80

0.75 Pt/Nb2O5,TiO2
Pt/C
0.70 Pt/Nb2O5/C 0.08 mA cm-2
Pt/Ebonex
0.65

-3.5 -3.0 -2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5


log j ( j / mA cm
-2
Pt )

Fig. 8.10 Tafel plots for the cathodic ORR scanned on RDE in 0.5 M HClO4 solution at 25  C for
E-tek, Inc., Pt/C (Vulcan XP-72, closed circles); Pt/Ebonex (Magneli phases, closed triangles); Pt
(10 wt%)/Nb2O5 (20 wt%)/C (70 wt%) (open triangles); and Pt/Nb2O5 (5 mol%),TiO2 (95 mol%)
(open circles) (N. V. Krstajic measured and plotted)

selective grafted and homogeneously distributed nanostructured Pt clusters down to


the prevailing (2.2  0.2 nm) nano-size, upon the optimized structure of mixed
valence hypo-d-oxide compounds (Pt/Nb2O5,TiO2/C), and even further extended
(iv) composites with hypo-f-oxides (Pt/Nb2O5,CeO2,TiO2/C, including doped
GdO2, HoO2, and/or LaO2 itself), and their relative combinations of extra high
stability and remarkable electron conductivity, too. Such comparable diagnostic
Tafel polarization interdependences distinctly displays Fig. 8.10, with differences
of more than an order of magnitude in the electrocatalytic activity for the ORR that
arise only as the result of different hypo-d-f-oxide type of interactive catalyst
supports, relative ratios of their amounts versus metallic part of the catalyst (Pt),
and their corresponding SMSI [6].

8.10.2 Leading Idea and Its Confirmations for the ROE


Electrocatalysts

Some subtle potentiodynamic survey of fundamental significance has then been


focused on the actual mixed interfering Pt-OH/Pt¼O coverages at selected points
within the characteristic potentials close to the open circuit value and all along the
reversible potential range of low Tafel line slopes (mostly 30 mV/dec, Fig. 8.11).
The main conclusions then have been that Pt electrodes, both plain and interactive
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 337

60
Ebonex / Pt

50

40
q / mC cm-2 Pt

2
30
4
E/V

- I / mA
1 - 0.86
6 2 - 0.88
20 3 - 0.90
1
8 4 - 0.92
2
5 - 0.94
3
4
6 - 0.98
10 7 - 1.00
10 5 6 7
8 - 1.05
8
12
0,0 0,2 0,4 0,6 0,8 1,0
E / V vs RHE
0
0,85 0,90 0,95 1,00 1,05
E / V vs RHE
b
350
Pt

300

250
q / mC cm-2 Pt

1
1
200 2
- I / mA

E/V
2
3 1 - 0.80
2 - 0.82
3 - 0.86
150 4
4 - 0.88
5
3 5 - 0.90
6
7
6 - 0.94
7 - 0.96

100 4
0,2 0,4 0,6 0,8 1,0
E / V vs. RHE

0,80 0,84 0,88 0,92 0,96


E / V vs. RHE

Fig. 8.11 (a) Charge density that required oxygen species for their reduction, presented as a
function of potential for Pt/(Magneli phases) (5 mg) electrode, same as in Fig. 8.10. The insert
shows potentiodynamic I vs. E relations scanned from different initial potentials (hold) with sweep
rate of 5 V s1.) (N. V. Krstajic conceived and plotted). (b) The same as in (a), but for
polycrystalline Pt metal
338 M.M. Jaksic et al.

supported upon hypo-d-oxides, have been covered by mixed (Pt-OH, Pt¼O) oxides
(besides many other experimental methods, in situ and ex-situ XPS confirmed [5, 6,
8, 43]), while at higher irreversible polarization (120 mV/dec), they were
completely deprived from such adsorptive layers. In other words, while within
the reversible range, the ORR is associated with and proceeds upon mixed oxide
covered Pt surface, as the interfering self-catalytic spillover electrode process,
(Eq. 8.3), otherwise known as the fastest reaction step [12], the oxide-free Pt surface
imposes much higher polarization for the direct electron transfer reaction to occur
[5–11]. Thus, the conclusive observation has now been that Pt oxides play the same
self-catalytic role to establish the ROE properties, as the spontaneous hydrogen
(Pt-H) adsorption does and means for the RHE. In fact, the leading idea consists
now from the extension of the reversible Tafel plot for the ORR all along the
potential axis down from or up to the thermodynamic value (1.29 V vs. RHE) that
implies enriched external latent Pt-OH storage and continuous spillover supply all
along in both scan directions, and particularly within the critical potential range
(Figs. 8.1 and 8.9). Namely, so far the problem and obstacle were the initial highly
reaction polarizable potential range of strongly adsorbed monolayered Pt¼O and
missing the Pt-OH, which cannot be supplied from aqueous solution, but only as the
adsorptive surface species. These rather specific potentiodynamic measurements
even more clearly show that the ORR upon Pt/Ebonex Magnely phase starts and
finishes at remarkably more positive potential values (1.05 down to 0.86 V, versus
0.95 down to 0.8 V, all versus the RHE, Fig. 8.11), relative to polycrystalline Pt
metal and/or nanostructured Pt/C [5, 6]; whereas the other two congenial issues (iii
and iv) initiate with 1.29 V as the completed reversible property of the ROE
(Figs. 8.3 and 8.5), and have been the ones of main substantial conclusive remarks
on the way towards the ROE (electro)catalysts. These afford and enable the reliable
link between the reversible and polarized oxygen electrode properties.
In such a respect, the guiding concept implies homogeneous nanostructured
distribution and selective grafting while interactive hypo-hyper-d-d-interelectronic
bonding of Pt(Au) nano-clusters upon various individual and preferably mixed
valence hypo-d-oxide supports (Fig. 8.8), taken for the reversible external
interconnected latent storage and spillover Pt-OH (Au-OH) sources (Scheme 8.1),
primarily Nb2O5,TiO2 (or Ta2O5,TiO2), because of their much advanced both the
stability and electronic conductivity. In such a constellation, nano-particles of solid
oxides and Pt establish the SMSIs, the ones of strongest in the whole chemistry
(Fig. 8.8), together with the electron conductive transfer, while external surface of
hypo-d-oxide deposit undergoes spontaneous dissociative adsorption of water mol-
ecules and thereby becomes, along with continuous further water vapor supply
(Scheme 8.1), the renewable and dynamically almost unlimited latent storage and
spillover source of the Pt-OH.
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 339

8.11 Nanostructured Characterization of Hypo-d-Oxide


Supported Electrocatalysts

The correlation of XRD spectra and Ultrahigh-Resolution Transmission Electron


Microscopy (UHRTEM) and in particular Scanning Transmission Electron Micros-
copy in High-Angle Annular Dark Field (STEM-HAADF) mode (also known as
Z-contrast) images belongs to the most reliable methods to estimate the nanostruc-
tured particle size distribution of Pt electrocatalysts upon interactive hypo-d-d-(f)-
oxide catalytic supports, and represents substantially the most significant nano-size
analysis manner for the time being now. These along with the SMSI of individual
(Pt, Au), or prevailing hyper-d-metal (MoPt3) catalyst interelectronic grafting
bonded upon hypo-d-d-(f)-oxide (example, Nb2O5, TiO2, CeO2) supports are now
the most reliable complementary sophisticated methods of prediction and estima-
tion of advanced synergistic electrocatalytic activity effects in particular for con-
temporary L&MT PEMFCs.
The XRD spectrum for the Pt/(Nb2O5,TiO2)/C catalyst reveals an overlapping
region between 35 and 50 (Fig. 8.12). In this particular region one expects the
reflection lines for the Pt(1 1 1) and (2 0 0), TiO2 anatase (0 0 4) and (2 0 0), as well
as the graphite (1 0 0) crystallographic planes, indicated with the labeling straight
lines drawn therein. Evidently, all the above mentioned reflections are essentially
overlapping. In such a respect, it is not possible to estimate exactly the Pt and TiO2
particles size, while peaks attributable to Nb specimens are at the noise level and
TiO2 anatase

Pt (111)

TiO2 anatase
(004)

graphite

Pt (200)

(200)
(100)
XRD intensity / a.u.

Pt/Nb2O3/C

Pt/(Nb2O3-TiO2)/C

30 35 40 45 50 55
2 theta / degrees

Fig. 8.12 XRD spectra of the Pt supported on Nb2O5/C and (Nb2O5,TiO2)/C. The vertical lines
drown show the positions of the respective peak reflections (Reprinted with permission from
[7]. Copyright (2014) American Chemical Society)
340 M.M. Jaksic et al.

absent, due to their very small amount. Nb oxide reflection peaks are also absent
even from the XRD spectrum of the enriched Pt/Nb2O5/C catalyst, though the
amount of niobia is now significantly higher as compared with the Pt/(Nb2O5,
TiO2)/C sample. The only visible reflection peak in the Pt/Nb2O5/C sample is
ascribed to the Pt (1 1 1) crystallographic plane. By fitting this particular peak
with a Gaussian function, the average Pt particle size estimated by applying the
Scherrer equation has been found to be approximately 2.2 nm, this way indicating
the highly and uniformly dispersed nature of such a catalyst. From hydrogen gas
B.E.T. chemisorption measurements [65], the average Pt particle size was estimated
to be 3.7 nm (77 m2/g of Pt assuming spherical particles, while the electrochemi-
cally active surface area, UPD H-adatoms estimated by the corresponding
potentiodynamic UPD peaks, is even assessed to be lower, 57.5 m2/g Pt). The
significantly larger Pt particles overestimated by the B.E.T. H2 chemisorption
indicates the interaction between the Pt particles and the oxide support (SMSI)
that hinders the hydrogen gas adsorptive properties of Pt crystallites, leading to
particle size overestimation. On the other hand, the particle size determination by
XRD is usually overestimated (the overlapping effect), and since the size is close to
the instrument limits, such a state of experimental evidence convincingly suggests
that Pt particles are evenly dispersed and strongly bonded on the hypo-d-oxide
catalytic support as shown by many XRD measurements with similar interactive
hypo-d-oxide supported electrocatalysts [5–11]. The absence of niobia reflection
peaks indicates that Nb-oxides are either highly sub-up to mono-layer dispersed on
the carbon particles surface (as essentially the surface species), or less probably in
amorphous state. In fact, the present electrocatalyst has been a priori planned and
calculated by the mass and surface ratio between hypo-d-oxide and Vulcan carbon
particles to be deposited as submonolayer (Nb2O5 having 170 m2/g, while carbon
particles being in higher amount and of much larger surface area, 260 m2/g). Such a
unique XRD experimental insight convincingly testifies to the existence of high Pt
(1 1 1) bronze dispersion upon highly developed hypo-d-oxide support. Such a state
relative to the standard Pt/C electrocatalyst, otherwise suffering from Pt surface
diffusion and agglomeration, characterizes remarkably increased electrocatalytic
activity, a much longer lasting catalyst because of the SMSI bonding effect, while
the interactive structure enables Pt metal recovery, which is another high quality
achievement of the novel nanostructured bronze type electrocatalysts. Meanwhile,
the most significant is the control of homogeneous dispersion of grafted nanostruc-
tured Pt size magnitude: The ratio of available exposed hypo-d-oxide surface
relative to the Pt 10 wt% defines the latter, and in our issues it is usually fixed
between 2.0 and 2.4 nm in average, in particular for NbOx, WOx and TiO2
individual and mixed valence composite hypo-d-oxide supports (SMSI); for TaOx
not yet optimized. However, TEM identifies plentiful of about 1.0 nm Pt/TaOx/C
clusters.
UHRTEM, and in particular STEM imaging in HAADF mode with
aberration correction (FEI Titan 80-300), reveal grafted Pt nano-clusters on hybrid
hypo-d-oxides supports, (Fig. 8.13a–d), with a rather uniform and evenly
homogeneous distribution in average of about 2–2.4 nm in size on the best issues
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 341

Fig. 8.13 Scanning transmission electron microscopy images in high-angle annular dark-field
mode (HAADF-STEM) of TM hypo-d-oxide/carbon support material and nanostructured Pt
electrocatalyst nanoparticles. (a) Composite simple basic Nb-oxide interactive (SMSI) supported
electrocatalyst, {Pt (10 wt%)/20 wt% Nb2O5/C}; (b) Congenial mixed altervalent supported
electrocatalyst, {Pt (10 wt%)/20 wt% Nb2O5,TiO2/C}; (c) Composite interactive (SMSI)
altervalent supported electrocatalyst of the same hypo-d-oxide structure as in (b), but of triple
higher Pt weight percentage, {Pt (30 wt%)/20 wt% Nb2O5,TiO2/C}; and (d) Congenial altervalent
supported electrocatalyst based on mixed W-oxide and anatase titania interactive supported Pt
clusters (Reprinted with permission from [7]. Copyright (2014) American Chemical Society)

{Pt (10 wt%)/20 wt% Nb2O5,TiO2/C}, Fig. 8.13b), {Pt (10 wt%)/20 wt% Nb2O5/
C}, Fig. 8.13a), and {Pt (10 wt%)/20 wt% WO3,TiO2/C}, Fig. 8.13d), obtained so
far and in excellent agreement with size measurement with XRD. Even more so,
with three times larger Pt amount {Pt (30 wt%)/20 wt% Nb2O5,TiO2/C},
Fig. 8.13c), the homogeneous nanostructured Pt distribution keeps the same trend
and in particular the average nano-size. Such a homogeneous, unusually narrow
342 M.M. Jaksic et al.

nano-size level of distribution has never so far been achieved (the interactive SMSI
effect) and confirms the reliable basis for the rather pronounced spillover effect and
the unique electrocatalytic achievements. No single Pt nanostructured cluster has
been noticed on the prevailing carbon nano-particles percentage of their otherwise
highly developed exposed surface area, no surface diffusion and no agglomeration,
either. It would certainly be worthwhile noticing a rather allover homogeneous
widespreading of the interactive hypo-d-oxide support structure and well distinct
inter-d-d-bonded and interactive (SMSI) grafted fine Pt nano-sized clusters upon
them, as a unique nanostructured Pt-bronze substantiation of advanced
electrocatalytic properties, primarily and mostly extended by the pronounced
interfering Pt-OH spillover effect for the cathodic ORR.
In addition, atomic-resolved HAADF-STEM observations strongly support the
interacting (SMSI) and interfering spillover theory. The selective interactive hypo-
hyper-d-d-interelectronic grafting distribution between Pt and Ta2O5 on the carbon
support results in the interdependent Ta oxide network structure, symmetrically
surrounding Pt cluster (Fig. 8.14). High-resolution HAADF images obtained with
aberration-correctors of the electron probe (0.07 nm resolution) even resolve the
individual Ta atoms of the oxide network supported on the carbon (Fig. 8.14b, c).
This structure of Pt surrounded by the Ta oxide network would result, right at the
interface Pt/Ta2O5, in the reinforcement of the overall SMSI effect at such a site,
possibly with the TaPt3 phase, already XPS confirmed for TiPt3 (Fig. 8.15a) [11],
and provide the highest electrocatalytic activity as the result.
The reinforced (SMSI) hypo-hyper-d-d-interelectronic bonded nanostructured
synergistic electrocatalytic composite Pt/TiPt3/TiO2 has already been approved and
tested, while symmetric monoatomic Ta distribution around Pt nano-cluster is the
first step towards the inter-bonding and reinforcing TaPt3 interphase at the interface
Pt/Ta2O5, promising further electrocatalytic SMSI advances. In other words, the
interactive Tauster hypo-hyper-d-d-(f)-supported electrocatalysts synergistically
improve their monophase (Pt) activity, which after longer reduction proceeding
with H-adatoms, or as the overall electrode reduction effect, finally emerge in the
Pt/TaPt3/Ta2O5,TiO2/C state of maximal electrocatalytic activity.
Following the Pt loading on complex support (10 wt% Pt/20 wt% Nb2O5/C),
HAADF, bright-field imaging and elemental mapping (EDXS) demonstrate that
crystalline Pt catalyst nanostructured clusters are homogeneously dispersed on the
hybrid mixed valence hypo-d-d-oxide support (Fig. 8.15b). Combining imaging and
elemental analysis, it was found that the catalyst nanoparticles are embedded into
(and/or properly symmetrically surrounded by) the oxide layer as pointed out using
white arrows in Fig. 8.14. Higher loading on the same support, (30 wt% Pt/20 wt%
Nb2O5/C), provides the evidence of maintained the same preceding uniform dis-
persion (Fig. 8.13). Based on elemental mapping with energy dispersive X-ray
spectrometry, it was enabled to deduce that the Pt nanoparticles distribution is
strongly correlated with the location of the transition metal oxide and rather
avoiding any carbon (no d-d-interactive bonding). When present, the thickness of
the disordered oxide network varies between 1 and 2 nm.
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 343

Fig. 8.14 Atomic-resolved HAADF STEM images analysis taken in different areas of the sample.
Pt catalyst particles and transition metal Ta-oxide network are identified by labels and pointed out
by relevant lines in images. (a) High resolution HAADF STEM general view of the carbon
support, Pt and the Ta-oxide network; (b) detailed view of the Pt and Ta-oxide network showing
the single Ta atoms of the oxide network (the oxygen atoms are not visible in this imaging
condition), (c) overview of intricate Ta-Ox network on carbon support and strong interaction
between the Pt catalysts only located on (or surrounded by) the Ta-oxide (Reprinted with
permission from [7]. Copyright (2014) American Chemical Society)

8.12 Primary Oxide Spillover DL Charging


and Discharging in the Light of the First
Principle Theory

The overall summation (Σ) for the ion exchange properties of hypo-d-oxide
hydrated membrane transferring matrix, under directional electric field effect or
polarization, yields the reversible Pt-OH adsorptive and/or desorptive spillover
344 M.M. Jaksic et al.

a Pt3Ti PtO Pt(CxHyCl)

0.7 eV
75.7 eV
73.6 eV
71.6 eV
XPS Intensity / a.u.

Pt foil

Pt3Ti

78 76 74 72 70 68 66

(1% Pt/TiO2)

84 82 80 78 76 74 72 70 68 66 64
Binding Energy / eV
b

Fig. 8.15 XPS 4f spectra scanned at the interphase between Pt catalyst (1.0 wt%) and interactive
anatase TiO2 support in their d-d-SMSI. Deconvolution indicates the existence of the TiPt3
intermetallic phase at the interphase Pt/TiO2 with all individual spectral properties, as the
prerequisite for the SMSI effect
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 345

upon, or from metal electrocatalyst reactive surface, resulting that way with the
characteristic pronounced reversible effusion peaks of extremely fast electrode
reaction [2–8], (Eq. (8.1-a), (8.1-b) and (8.1-c), yielding by such a summation the
same former Eq. 8.1),

M þ H2 O , M-OH þ Hþ þ e ð8:1Þ

The starting first principle thermodynamic relation for such a SMSI (exampled
as M/TiO2) system and just defined general reversible peak reaction, Eq. 8.1, arises
therefrom in the following form,

μHþ ¼  μMOH  μe þ μH2 O þ μM ð8:7Þ

while from the main general self-explanatory relation for electrochemical potential
definition,

μe ¼ μe  eφ ¼ μe  eψ  eχ ð8:8Þ

and particularly for hydronium ion (H3O+, abbreviated for simplification as H+),
there from Eq. 8.8 further follows an alternative relation

μHþ ¼ μHþ  eφHþ ð8:9Þ

To avoid and eliminate electrochemical potentials (μHþ , μe ), by equalizing the


right-hand sides of Eqs. 8.7 and 8.9, and replacing the corresponding value from
Eq. 8.8, there straightforward yields the following relation,

μHþ ¼ μMOH þ μM þ μH2 O þ eφ  μe þ eφHþ ð8:10Þ

Meanwhile, since by definition, φ ¼ φHþ þ χ ¼ ψ þ χ, and thereby,


φHþ ¼ ψ, there further follows,

μHþ ¼ μMOH  μe þ μM þ μH2 O þ 2eψ þ eχ ð8:11Þ

Finally, by taking the derivative of Eq. 8.10, because Δμe ¼ 0, and since
μM , μH2 O , and ψ ¼ const:, while ΔμHþ  e Δχ, there straightforward results,

ΔμM-OH ¼ eΔχ ð8:12Þ

exactly as cyclic voltammograms (Fig. 8.3) reveal: Every change in spillover of the
M-OH, or of its chemical potential ΔμMOH ), as the driving force of the effusion,
or, simply its concentration variations, each corresponds to the increment of surface
potential (Δχ), or the change in capacity of DL because of M-OH dipole adsorption
package, or desorption (as the spillover capacity), both of them being fast and
substantially reversible, define DL charging or discharging (in its capacity, or
346 M.M. Jaksic et al.

pseudo-capacitance). Namely, since ΔμHþ is logarithmic function of hydronium


ions concentration, and according to the spillover electrode reaction (Eq. 8.1),
primary oxide (M-OH) and H3O+ ions become simultaneously spent or produced
in equivalent (1:1) amounts, while the contribution of surface energy is the linear
function of surface potential (prevailing dipole effect), eΔχ, then the inequality
conclusion, ΔμHþ  eΔχ, emerges quite logically.
Since the fundamental definition of work function (Φ) states [52, 66], that

eΦ ¼ μe þ eχ ð8:13Þ

and taking into account Eq. 8.12, by derivation of the latter (Eq. 8.13), there straight
forward follows the broader fundamental definition for M-OH spillover adsorption
and/or desorption, which is the same as the DL pseudo-capacitance charging and/or
discharging,

eΔΦ ¼ eΔχ ¼ ΔμMOH ð8:14Þ

Simple combination of Eq. 8.13 with Eq. 8.8, then yields the alternative first
principle relation between work function and electrochemical potential, as the well
known equation of general meaning and significance,

μe ¼  eΦ  eψ ð8:15Þ

which, for the considered issue of electrochemical cell (UWR), assembled between
working (W) and stable reference (R) electrode, can be straightforward written to
read [52, 66],

Δμe ¼ ΔU WR ¼ eΔΦ  eΔψ ð8:16Þ

The main contribution both to changes in work function (ΔΦ) and primary oxide
dipole spillover (ΔμMOH ), comes from the surface potential increment (Δχ) itself
(Eq. 8.12), as the prevailing part of inner or Galvani potential (Δφ). Namely,
Vayenas et al. [67, 68] have proved on plentiful systems in solid state electrolyte
[69], aqueous media [70–72] and PEM Nafion 112 [73] the basic NEMCA or EPOC
promotion relation in heterogeneous catalysis,

ΔU WR ¼ eΔΦ ð8:17Þ

that could also be considered as fundamental both in electrode kinetics [74–76] and
spillover phenomena [66, 77]. Meanwhile, the latter (Eq. 8.17), when correlated
with the preceding relation (Eq. 8.16), unambiguously and consequently yields that

Δψ ¼ 0, ψ W ¼ ψ R ¼ const: ð8:18Þ
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 347

which is also implied in above preceding equations (Eqs. 8.11 and 8.12)
[78, 79]. Such conclusive and far-reaching statement is in particular reliably clear
for the electrode steady state and with high concentration supporting electrolyte.
The whole thermodynamic equilibrium for the general issue of interactive
(SMSI) metal/hypo-d-oxide supported catalyst, such as M/TiO2, is now illustra-
tively displayed in all interconnected energy details by Fig. 11 in [9]. As a
consequence, any decrease of the M-OH chemical potential (ΔμMOH ), causes the
work function to decrease correspondingly, and vice versa. In other words, any
consumption of adsorbed M-OH upon metallic catalyst surface causes to decrease
its chemical potential, and this way imposes imbalance within the established
equilibrium. Thus, to keep constant both ΔΦ and cell voltage between working
(W) and reference (R) electrode (ΔUWR), the titania phase, by its enriched latent
storage of the primary oxide, automatically reacts by its membrane transferring
mechanism and supplies hydroxide ions (Scheme 8.1), to emanate further as the
primary oxide spillover adsorptive species on exposed metallic catalyst surface by
corresponding electrocatalytic reaction (Eq. 8.1), both from hydrated anatase struc-
ture and by reacting of continuously supplied water molecules in their spontaneous
adsorptive dissociation [21, 22].
The overall conclusion associated with otherwise broad experience in electro-
chemical promotion (NEMCA, EPOC) [66–68] of various chemical reactions in
heterogeneous catalysis and particularly for Doebereiner (Ertl interpretation [30])
oxidation of hydrogen in water solutions, for catalysts interactive (SMSI) supported
upon hypo-d-oxides, hydroxyl ions play the main membrane transferring and
decisive promotion role, while M-OH, as defined dipole state, undertakes the
spillover and features as the substantially mostly reacting species all over the
metallic catalyst surface, enabling that way to keep established such thermody-
namic equilibrium. In fact, there is no catalytic promotion for hydrogen oxidation in
acidic aqueous media and apparently hydroxide ions impose such a substantial
promoting efficiency, in particular within the interfering (600–900 mV vs. RHE)
potential range, when NEMCA approaches promotion effects of two and even more
orders of magnitude [68, 71–73]. Therefore, any disturbance of such an equilib- 
rium, such as the consumption of spillover adsorbed M-OH species (ΔμMOH , or
simple imposition of an external polarization, as just analytically shown above,
reflects as the driving force (ΔUWR ¼ eΔΦ), and instantaneously tends to
reestablish the former equilibrium, or to keep its steady state.
Since the adsorption of the primary (M-OH) oxide plays a decisive and critical
role in many important (electro)catalytic processes, in particular for CO tolerance
in PEMFCs and the ORR, enthalpies of chemisorption might throw some light for a
priori considerations and estimative assessments. Namely, from the measured
isosteric heat of Pt-OH adsorption (ca 200 kJ/mol), the adsorptive Pt(1 1 1)-OH
bond energy was estimated to be ca 136 kJ/mol, that is much less than the surface
(Pt¼O) oxide chemisorptive bonding energy (ca 350 kJ/mol), and this is taken as
the argument for higher catalytic activity of the former in surface electrochemistry
of CO [80]. In fact, a rough estimation of adsorptive energy for interactive
348 M.M. Jaksic et al.

supported Pt catalysts upon hypo-d-oxides tells that the above ratio might be even
for more than an order higher, as a typical ratio between reversible versus polariz-
able properties, and consequently Pt-OH correspondingly much more active in
heterogeneous catalysis and particularly for the ORR. Otherwise, since two
Pt-oxides, Pt-OH and Pt¼O, distinctly differ in their polarization of desorption
and/or adsorption, and since the latter defines the enthalpy of adsorption, such
pronounced overvoltage differences unavoidably should have to imply their
corresponding adsorptive/desorptive enthalpy values and differences. In the same
context, hydrogen atoms from bronze type interactive supported Pt electrocatalysts
(Pt/H0.3WO3) are distinctly more electrocatalytically and catalytically active than
H-adatoms from plain nanostructured (Pt/C), and feature much weaker adsorptive
bonding strength.

8.13 XPS Evidence for the Primary Oxide Presence


and Spillover

The link between model polycrystalline Pt (85 % in (1 1 1)) and the interactive
supported nanostructured Pt electrocatalysts, on various individual, or mixed
valence hypo-d-oxide compounds (Nb2O5, WO3, MoO3, Ta2O5), composed with
anatase (1 0 1) titania (TiO2) itself, imposed the need for advanced contemporary
surface characterization, in particular as concerns the primary oxide (Pt-OH)
appearance, latent storage and spillover [5, 6, 11, 65]. More specifically, in order
to understand the origin of various oxidation states of Pt deposited on both indi-
vidual (Nb2O5/C, Fig. 8.16) and/or mixed valence (Pt/Nb2O5,TiO2/C, Fig. 8.17)
hypo-d-oxide supports, in situ Pt 4f7/2 and O 1s XP spectra have been scanned and
deconvoluted before and after the exposure in heated hydrogen stream, and by
simple thermal annealing effect under UHV and/or relative to room temperature
(RT) for samples treatment avoiding any exposure to the atmosphere. The shape Pt
4f spectra indicates the existence of more than one oxide type Pt species, so that in
all issues, the photo-peaks can be analysed into three doublets [81]. The first
component at binding energy 71.3 eV is attributed and identified to belong to
metallic platinum, the peak at 72.7 eV is assigned to Pt(OH)x (1  x  2), due to
combination of Pt with OH in the primary oxide (M–OH) type, or some other
congenial O-species, while the third component at 74.9 eV is ascribed to PtOx
[81–83], finally ending-up under longer storage in air, as the stable surface oxide
(PtO2). The partial percentage contribution of each Pt-oxy-component to the total
peak area is equivalently shown for both specimens in Table 8.1. As a strong XP
spectral evidence for the primary oxide (Pt-OH) existence (Eq. 8.1) and interactive
transference (Eq. 2), is quantitatively assessed that both the simple thermal, and in
particular the annealing under hydrogen flow, causes the corresponding distinct
reduction of the Pt oxy-specimens either as the thermal desorptive effect and even
more pronounced by the reactive contribution of Pt-H itself. Similar experimental
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 349

a b
XPS-MgKa
Pt4f XPS-MgKa
Pt/Nb2O3 O1s
metal oxides
Pt/Nb2O3
o Pt(OH)x
Pt PtOx

-OH
XPS Intensity / a.u.

XPS Intensity / .a.u


o o
360 C (H2 flow) 360 C (H2 flow)

o o
360 C (in Vaccum) 360 C (in Vaccum)

RT RT
66 68 70 72 74 76 78 80 82 526 528 530 532 534 536
Binding Energy /eV Binding Energy /eV

Fig. 8.16 (a, b) Pt 4f (a) and O1s (b) XPS core level spectra of the Pt catalyst on individual hypo-
d-oxide, Nb2O5/C support at room temperature (RT), after annealing under UHV conditions at
360  C and after reduction at 360  C under 20 % H2 in N2 flow.

a b
XPS-MgKa o XPS-MgKa
Pt
Pt4f Pt(OH)x
PtO
x O1s metal oxides
Pt/TiO2-Nb2O3 Pt/TiO2-Nb2O5
-OH, C-O

o
360 C (H2 flow)
XPS Intensity / a.u.

XPS Intensity / a.u.

o
360 C (H2 flow)

o
360 C (in Vaccum)

o
360 C (in Vaccum)

RT RT

66 68 70 72 74 76 78 80 82 526 528 530 532 534 536


Binding Energy /eV Binding Energy /eV

Fig. 8.17 (a, b) Pt 4f (a) and O1s (b) XPS core level spectra of the Pt catalyst on the mixed
valence hypo-d-oxides, Nb2O5,TiO2/C support at room temperature, after annealing under UHV
conditions at 360  C and after reduction at 360  C under 20 % H2 in N2 flow
350 M.M. Jaksic et al.

Table 8.1 Percentage contribution of each Pt4f component the total peak area
Pt/Nb2O5-TiO2 Pt/Nb2O5
Pt0 Pt(OH)x PtO2 Pt0 Pt(OH)x PtO2
(71.3 eV) (72.7 eV) (74.9 eV) (71.3 eV) (72.7 eV) (74.9 eV)
RT 63.6 29.5 6.9 58.3 25.6 16.1
360  C in 68.8 18.7 12.5 58.5 30.1 11.4
vacuum
360  C under 75.9 16 8 71 19.1 9.8
H2 flow

evidence has already been observed on the simple basic anatase titania (Pt/TiO2,
Fig. 8.4, [5]) interactive supported Pt electrocatalyst [5, 6, 11, 65]. Furthermore, a
thorough parallel examination of the O 1s along with Pt 4f spectra, shows a
pronounced decrease of the component attributed to the Pt-OH species by the
former (BE ~ 533 eV), which apparently contradicts the enrichment in Pt(OH)x
amounts, observed from the latter, Pt 4f spectrum. In other words, whilst the
annealing causes a straight reduction of Pt oxi-species and desorption of the
primary oxide (Pt-OH), the intensity discrepance in decrease of the O 1s component
attributed to the same Pt-OH species might be compensated by the spillover from
the enriched primary oxide latent storage of mixed valence hypo-d-oxide support.
There is no other source for such remarkable difference and apparent discrepancies.
In fact the same compensation (or, balance feeding spillover) effect enables anatase
titania individually itself, but at correspondingly lower intensity [5, 6, 11, 65].
Meanwhile, all XP spectra distinctly indicate the existence, desorptive and/or Pt-H
reactive removal and compensating latent storage spillover effect of the primary
oxide [7]. In any case, the O 1s spectrum reveals a larger initial partial amounts of
the primary oxide and other similar reactive oxi-species (Pt(OH)x) for the issues of
mixed valence interactive supported electrocatalysts. As a whole, the annealing in
vacuum causes a distinct reduction of Pt oxi-species, in particular the decrease of
the more active primary type Pt(OH)x oxides, and a slight increase of initially
present, (long term of the storage nucleation aging effect) in relatively much lower,
minor extents, or even initially absent, the PtOx (PtO2). At the same time, the O 1s
component attributed to OH-species correspondingly decreases, indicating the
significant desorption of hydroxyl groups in particular pronounced from the
mixed valence hypo-d-oxide supports during annealing, as the strong clear evi-
dence for the initial primary oxide enriched (latent storage) presence and spillover.
In general, the most active primary oxide (Pt-OH) then keeps the low level, or even
the absence of higher oxide in their amount or percentage, (PtOx, PtO2).
XP spectral ex situ measurements, Fig. 8.18a, b, reveal the corresponding Pt 4f
deconvoluted spectra for the originally fresh prepared hypo-d-oxide interactive
(SMSI) Pt electrocatalyst supports by Nb2O5, TiO2-Nb2O5, TiO2-WO3 and TiO2
itself, and for the same specimens after an annual storage. Three distinct
oxo-components, just as above, participate now within the electrocatalytic Pt
surface, appearing at three different binding photo-peaks, but the
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 351

a
XPS-AlKa 72.7 eV
Pt4f Pt(OH)x
74.6 eV
71.1 eV PtOx
Pt0

Pt/Nb2O5/C

b
XPS Intensity /a.u.

XPS-MgKa
Pt4f
Pt/Nb2O5+TiO2/C Pt(OH)x PtOx
Pto
XPS Intensity / a.u.

Pt/ WO3+TiO2/C

Pt/Nb2O5+TiO2/C

Pt(WO3+TiO2)/C

Pt/Nb2O5/C

Pt/TiO2
Pt/TiO2/C

68 70 72 74 76 78 80 68 70 72 74 76 78 80
Binding Energy /eV Binding Energy /eV

Fig. 8.18 (a, b) Pt 4f XPS core level spectra of Nb2O5, TiO2-Nb2O5, TiO2-WO3, and TiO2
supported Pt catalysts at room temperature: ((a) scanned fresh and (b) after aging of one year
exposed in air).

non-stoichiometric PtOx oxides (1  x  2), ending-up by the upper type surface


oxide (PtO2), as the final minor oxidation state [5, 6], (compare Fig. 8.10, [6]). The
latter even does not appear at the Pt 4f spectrum of the fresh prepared Nb2O5-TiO2
supported Pt catalyst, even after many repeating scans within a longer period
(Fig. 8.18a). Such a decisive and conclusive remark is of substantial and funda-
mental significance for the present study: The primary oxide is proved initially
available, in particular in the condensed wet state of the electrocatalysts. In other
352 M.M. Jaksic et al.

Table 8.2 Percentage contribution of each Pt4f component to the total peak area for fresh
prepared specimens and for an annual storage of the same specimens
Percentage contribution of Pt4f components
Catalyst support Pt0 (71.1 eV) Pt(OH) (72.7 eV) PtOx (74.8 eV)
(a)
Nb2O5 67.2 % 23.7 % 9.1 %
Nb2O5-TiO2 65.3 % 34.7 % –
WO3-TiO2 61.9 % 26.7 % 11.4 %
TiO2 56.8 % 31.9 % 11.3 %
(b)
Nb2O5 58.3 % 25.6 % 16.1 %
Nb2O5-TiO2 63.6 % 29.5 % 6.9 %
WO3-TiO2 61.9 % 26.7 % 11.4 %
TiO2 56.8 % 31.9 % 11.3 %

words, for such a state of experimental evidence, to complete the whole story, one
should recall the self-catalytic effect of water molecules on the primary oxide a
priori appearance and existence [30], and/or the Vittadni [21, 22] spontaneous
dissociative water molecular adsorption, as the source of the latent storage for the
Pt-OH spillover. Under such circumstances, highly reversible and electrocata-
lytically extremely active primary oxide, (Pt-OH), by the spontaneous interfering
electrode reaction (Eq. 8.3), spends the exposed surface oxide (Pt¼O), and prevents
its upper level (PtOx ! PtO2) growth, so that the latter is either missing in fresh
prepared specimens, or minor component after longer free air exposure.
The percentage contribution of each component to the total peak area is shown in
Table 8.2a, which has some deeper theoretical significance. The roots and expec-
tancies for the primary oxide spillover are indicated there on an almost quantitative
scale basis. The decisive and pronounced cooperative effect of anatase titania is
clearly indicated by its individual rather high initial primary oxide surface percent-
age, while the Pt/Nb2O5,TiO2 features the most (dynamic latent storage) creative
synergistic properties for the Pt-OH generation that further reflects on and definitely
defines the overall distinctly advanced electrocatalytic activity for the ORR
[7]. Shrinkage of the same XPS Pt-OH peak capacity both by the thermal and
hydrogen reduction effect at higher temperature (Figs. 8.16 and 8.17a, b, see also
Fig. 8.4, [5]), is observed as the additional experimental evidence for the existence
of primary oxide initially available for further instantaneous spillover distribution
under electrode polarization. In the same respect, the complementary DRIFT
imprints (Fig. 8.9, [84]) on a similar manner reveal the same primary oxide thermal
and Pt-H reduction features, but the desorption effect in the present XPS issues
arises proportionally smaller since under UHV the examined specimens are
deprived from water molecules. Meanwhile, the same specimens after unused
open air storage show the nucleation of PtOx, ending with PtO2 (Fig. 8.17b), as
the common transient in particular pronounced with more active the interactive
supported Pt electrocatalysts is [5–11], and/or suppressed in the growth by more
active and when accumulated in larger amounts the primary(Pt-OH) oxide.
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 353

XPS-AlKa 4+ XPS-AlKa
Ti
Ti2p Ti2p
Pt/Nb2O5-TiO2 Pt/TiO2
XPS Intensity /a.u.

XPS Intensity /a.u.


4+
Ti

3+
Ti

452 454 456 458 460 462 464 466 468 456 458 460 462 464 466 468 470
Binding Energy /eV Binding Energy /eV

Fig. 8.19 Deconvoluted Ti 2p XPS core level spectra of TiO2-Nb2O5 (left) and TiO2 (right)
supported Pt catalysts

207.1 eV XPS-AlKa
XPS-AlKa 207.1 eV
5+ Nb3d
Nb3d (Nb ) 5+
Pt/Nb 2O 5 (Nb )
Pt/Nb2O5-TiO2
XPS Intensity /a.u.

XPS Intensity /a.u.

204 206 208 210 212 214 202 204 206 208 210 212 214
Binding Energy /eV Binding Energy /eV

Fig. 8.20 Deconvoluted Nb 3d XPS core level spectra of TiO2-Nb2O5 (left) and TiO2 (right)
supported Pt catalysts

In Fig. 8.19 the deconvoluted Ti 2p spectra of TiO2-Nb2O5 and TiO2 itself


supported Pt catalysts are, in addition, comparatively presented. The main doublet
in both spectra, at binding energy (Ti 2p3/2) 458.8 eV is attributed to Ti4+ species
[81, 85, 86]. In the case of the TiO2-Nb2O5 mixed valence hypo-d-oxide support, a
second doublet at binding energy (Ti 2p3/2) 455 eV is attributed to the contribution
of Ti atoms in the Ti3+ state. Ti atoms in the other mixed hypo-d-oxide support
TiO2-WO3 are detected only in the Ti4+ state [87]. Such an effect might be the cause
for the higher synergistic electrocatalytic activity of the interactive Nb2O5, and
similarly Ta2O5, mixed valence compounds with anatase (1 0 1) titania (TiO2),
relative to other hypo-d-d-oxide composites.
In Fig. 8.20 the deconvoluted Nb 3d XPS core level spectra of TiO2-Nb2O5 and
Nb2O5 supported Pt catalysts are presented. In both cases only one doublet is
apparent at binding energy (Nb 3d5/2) 207.1 eV that is characteristic for the Nb5+
state, which, when inserted into a titanium dioxide network, usually causes the
effect in charge compensation [88]. The addition of such a charge can be compen-
sated either by creating one vacancy of Ti per four introduced Nb ions, or by the
354 M.M. Jaksic et al.

reduction of Ti4+ to Ti3+ per each inserted Nb5+-ion. Both of these effects can occur,
with the latter being much more probable at relatively lower temperatures [88].
Quantitative analysis of the present results, using the Ti 2p and Nb 3d peak
intensities (areas) corrected by the atomic sensitivity factors [89], shows that the
appearance of each Ti3+, corresponds indeed to the introduced one Nb5+-ion,
[90, 91]. In fact, EELS (Electron Energy Loss Spectroscopy) scans, in addition,
clearly point to the valence variations of the oxide support, especially at the Ti
edges (cf. [92, 93]).
Now, there arises a very interesting situation concerning the most promising
electrocatalytic Nb-oxide structure, which at relatively low temperatures of calci-
nation crystallizes as Nb2O5. The latter is well confirmed by the XPS analysis,
while much more stable NbO2 appears above 900  C [94]. Such experimental
evidence reveals the fifth hydroxide (OH) ion for the most easily transferable
within the overall spillover mechanism,

NbðOHÞ5 þ Pt , NbðOHÞ4 þ þ Pt-OH þ e ð8:19Þ

Meanwhile, XPS analysis has also revealed a further interrelating mechanism of


similar exchanges with hydrated anatase titania,

NbðOHÞ4 þ þ TiðOHÞ4 ) NbðOHÞ5 þ TiðOHÞ3 þ ð8:20Þ

or, when summed up,

TiðOHÞ4 þ Pt , TiðOHÞ3 þ þ Pt-OH þ e ð8:21Þ

Then, the entire formalism clears up the equivalence and mutual facilitation
between titania and niobia for the primary oxide spillover, as already concluded
from Table 8.2a.
The conclusive observation from above XPS experimental evidence then has
been that such membrane type ionic migration, both through the bulky network and
over hypo-d-oxide surface exposure, ending up with the Pt-OH(Au-OH) spillover,
is interdependent and can be further remarkably optimized by the common catalyst
interactive support structure, in particular by hypo-f-oxide (CeO2, GdO2, HoO2)
ingredients with much larger, and thence, much more exposed (and thereby easier
interactive and migrating) latent storage hydroxide structure. Let us then add the
straight intimate and much more expanded interactive Pt catalyst hydrated interface
(Pt/Nb(OH)5, Ti(OH)4, Ce(OH)4) contact, and even direct interaction of NaOH
from rich alkaline media [9],

OH þ Pt , Pt-OH þ e ð8:10Þ

which is the same and equivalent to Eq. 8.1 [9], to become aware of the multiple
membrane type migration and even direct spillover Pt-OH generation. In such an
electrocatalytic constellation surface and bulk migration represents now the main
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 355

overall RDS and afforded open optimization area to provide already superior, broad
range reversible electrocatalysts for oxygen electrode reactions (ORR, OER).
The reversible fast adsorptive interacting primary oxides (M-OH) lead up to
adsorbed monolayers (Pt-OH ! 1), and are distinctly marked by a specific
potentiodynamic peaks, as the first step of water oxidation (Eqs. 8.1 and 8.2) [2–
4], immediately after the initial classical double layer charging range of many
metals, first of all on Ru, Au, Pt, Rh, Ag, Ni. Similar behavior is exhibited by
some other transition metal primary type oxides of common formulas and structure,
like MoO(OH), NiOOH, AuOOH, WO2(OH), and can be identified by
corresponding peaks within their potentiodynamic and XPS spectra [53, 65]. All
hypo-d-elements, in particular of high altervalent capacity, afford the reversible
primary oxide type states, usually of pronounced catalytic activity and high elec-
tronic conductive properties (WO(OH), MoO2(OH)), but unfortunately, in their
oxidation sequences end-up with non-conductive and catalytically inactive higher
valence oxide states (MoO3, WO3, Pt¼O, PtO2, Au¼O, NiO2), otherwise the whole
aqueous electrochemistry would feature another and quite different physiognomy!

8.14 Striping Voltammetry Evidence for the Primary


Oxide Spillover Effect

Striping voltammetry represents another indirect complementary experimental


potentiodynamic evidence to scan and prove the primary oxide spillover effects
in electrocatalysis. Interactive hypo-d-oxide supported and non-supported
electrocatalysts (both Pt and RuPt) exhibit dramatically different activity for CO
tolerance in LT PEMFC, and distinctly provide new additional cyclic voltammetry
scanning spectra for the M-OH spillover effect [5, 6]. Ever since Watanabe [95] has
shown that Ru even at submonolayer core-shell deposit, or while alloying with Pt,
shifts the primary oxide growth to a much more negative potential range and
enables CO tolerance, the primary oxide spillover became of substantial signifi-
cance for PEMFCs [96]. Similarly, the hypo-d-oxide supported Pt and Ru (Pt/TiO2/
C, Ru/TiO2/C) in their behavior versus these two plain and non-interactive
supported metals (Pt/C, Ru/C) themselves, or even their interactive unsupported
alloys, RuPt/C and RuPt/TiO2/C, Fig. 8.21, have distinctly different catalytic
properties, too, the interactive alternatives featuring an even more advanced and
much more pronounced primary oxide spillover effect [5, 6]. Such distinct and
clearly pronounced peak capacity difference between the interactive and
non-interactive hypo-d-oxide supported Ru advanced electrocatalysts, belong to
the best experimental evidence for the primary oxide spillover effect extensions by
the latent storage. Since hypo-d-oxides, primarily anatase titania, zirconia and
hafnia, and even more so tungstenia, niobia and tantalia, facilitate the spillover of
M-OH, such facts clearly point to the advanced overall composite effect and
advantages of membrane type OH—transferring within and over TiO2, WO3,
356 M.M. Jaksic et al.

(b)
0,012 (c) (a)

0,010 (d)
0,008

0,006

0,004

0,002
I(A)

0,000

-0,002
o
-0,004
(a) 25 C
o
(b) 60 C
-0,006 o
(c) 80 C
-0,008 o
(d) 60 C
-0,010

-0,012
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,9 1,0 1,1 1,2 1,3 1,4 1,5
V(Volt)

Fig. 8.21 The striping voltammograms for CO desorption from supported 10 wt% (0.4 mg cm2,
2 nm in average size, 1:1 atomic ratio Ru:Pt) RuPt/TiO2/C electrocatalyst CO-saturated at three
different temperatures: 25  C (a); 60  C (b) and 80  C (c), scanned at the scan rate of 2 mV s1; (d)
the same stripping scans for CO desorption at 60  C from unsupported 30 wt% (0.5 mg cm2)
E-tek RuPt/C electrocatalyst of the same RuPt nano-size, atomic ratio and load, and sweep rate
10 mV .s1; its CO saturation at 55  C

Nb2O5, TaO2, and in particular within their substantially wet mixed networks of
mixed altervalent catalyst supports, resulting in the speeded-up primary oxide
effusion, relative to the plain non-interactive carbon (Pt/C).
In other words, while Ru metal itself facilitates both Pt-OH and Ru-OH spillover
transfer even in the intermetallic non-interactive supported RuPt composite
electrocatalyst [95, 96], the supporting effusion effect of titania advances the
same effect for more than 300 mV relative to RuPt/C catalyst (Fig. 8.21). Anodic
CO oxidation upon Ru,Pt/TiO2/C starts even within the potential range of UPD
desorption of H-adatoms and becomes much more pronounced in the charge
capacity relative to Ru, Pt/C. This important result is one of the most significant
confirmations of the present interactive and dynamic spillover catalytic model, as
implemented in electrocatalysis for hydrogen and oxygen electrode reactions. In
fact, cyclic voltammogram spectra of plain Ru feature a unique type of shapes and
in both scan directions stepwise reveal more than any other individual transition
metal various interrelating oxide states and their peaks interdependances [97], and
with such overall potentiodynamic sequences are closer to the reversible properties
for oxygen electrode reactions than any other d-metal.
Mixed anatase (and even rutile) titania, and in particular tungstenia, niobia and
tantalia form intermolecular solid oxide solutions of a high altervalent capacity
(Scheme 8.2), compatible both in amorphous and crystalline forms of the edge
sharing TiO6 and the corner sharing WO6 octahedrons, with pronouncedly
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 357

Scheme 8.2 Model presentation of electrochromic individual WO3 layers and composite combi-
nations of tungstenia and titania with corner and age sharing crystal units within the consistent
mixed valence compounds. SEM and TEM confirmed for amorphous overall structure: (a, b) as
deposited films, (c, d) as colored and (e, f) as bleached hypo d-oxides [92, 93]

increased electrochromic features even at high contents of the former [92, 93].
In fact, highly charged W6+ cations, like Nb5+, additionally favor the reversible
acidic dissociation of water molecules [21, 22], and thereby such electrochromic
layers exhibit well defined ion exchange and electron conductive properties
[23–25]. Thus, one of the fundamental contributions of the present paper is to
show that prevailing anatase titania in the form of a composite mixed valence
catalytic support with tungstenia (and/or even more so niobia, tantalia and
molybdenia), behaves in a compatible way and regarding the (Pt/H0.35WO3 , Pt/
W(OH)6) bronze type equilibrium, so that the consequent reversibility features the
same properties as pure tungsten bronze itself, all of them being further advanced
by hypo-f-oxide ingredients (Ce, Gd, Ho, La, etc.)!
358 M.M. Jaksic et al.

8.15 Superior Bronze Type Electrocatalysts for Reversible


Alterpolar Interchanges in the Revertible (PEMFC vs
WE) Cell Assembly

The striking target issue of the present paper has been to show the development and
substantiation of corresponding spillover electrocatalysts for the superior revertible
cell assembly for spontaneous reversible alterpolar interchanges between PEMFC
and WE. In other words, L&MT PEMFCs, in particular large stationary systems,
are unimaginable without revertible abilities with WE alternation, enabling to
produce hydrogen during (usually overnight) excesses in electrical energy supply,
and vice versa, to use stored amounts of the latter as a fuel in PEMFCs, in the course
of (daily) deficiencies in electricity. In aqueous media classical plain Pt/C
electrocatalysts feature catalytic surface properties of Pt-H and/or Pt¼O, without
any broader effusion of other interacting species, and such situation prevents
reversible alterpolar interchanges, otherwise indispensable for the smooth revert-
ible altering of PEMFC versus WE systems, when both alternatives are mutually
equivalent and operate in their common overall harmony. In such a respect, the
present paper introduces electron conductive and d-d-interactive individual and
composite (mixed valence) hypo-d-(f)-oxide compounds of increased altervalent
capacity, or their suboxides (Magneli phases), as the interactive catalytic supports
and therefrom provides: (i) The SMSI interelectron-d-d-bonding and correspond-
ingly improved catalytic effects, and (ii) Dynamic spillover interactive transfer of
the M-OH, and alternatively the free fast effusion of H-adatoms for further elec-
trode reactions, and thereby advance the overall electrocatalytic and alterpolar
activities. Since hypo-d-oxides feature the exchange membrane properties, the
higher the altervalent capacity, the higher the spillover effect. In fact, altervalent
hypo-d-oxides impose spontaneous dissociative adsorption of water molecules and
then spontaneously pronounced membrane spillover transferring properties instan-
taneously resulting with corresponding bronze type (Pt/HxWO3) under cathodic,
and/or its hydrated state (Pt/W(OH)6), responsible for Pt-OH effusion, under anodic
polarization, this way establishing instantaneous reversibly revertible alterpolar
bronze features (Pt/H0.35WO3 $ Pt/W(OH)6), as the thermodynamic equilibrium,
and thereby substantially advanced electrocatalytic properties of these composite
interactive electrocatalysts, all along the potential axis between hydrogen and
oxygen evolving limits, back and forth. Due to the dual spillover alterpolar inter-
changes, all four electrode reactions in aqueous media are reversibly altered
(HER—Cathodic hydrogen evolution, HOR—Anodic hydrogen oxidation,
ORR—Cathodic oxygen reduction, OER—Anodic oxygen evolution). In other
words, the HER and OER are entirely equivalent in the operational sense to the
couple of ORR and HOR, and in the revertible alterpolar interchanges between
PEMFCs and WE that plain Pt by no means could enable itself, and this is the
substance why we consider such electrocatalytic system for superior, and even ideal
and unique. Such nanostructured type electrocatalysts, even of mixed valence hypo-
d-oxide structures (Pt/H0.35WO3,TiO2/C, or Pt/HxNbO3,TiO2/C), have for the first
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 359

time been synthesized by the sol-gel methods, including the super-critical drying
with liquid CO2, and shown to feature extra high stability, electron conductivity and
non-exchanged initial pure mono-bronze spillover and catalytic properties. In other
words, behaving as the unique, even ideal electrocatalytic system for substantiation
of the revertible cell assembly for spontaneous reversible alterpolar interchanges
between PEMFC and WE, though still being based on unavoidable Pt catalysts.
Meanwhile, since nanostructured Pt clusters are grafted and inter-bonding fixed
upon hypo-d-oxide support, just confirmed by UHRTEM nano-images, their life
time has been dramatically increased, at least doubled, there is no agglomeration,
nor hydrogen crossover through PEM, and there is now provided ability for
prevailing Pt recovery. Namely, the substance consists in the reversible alterpolar
interchangeable electrocatalytic operation with pronounced spillover (both Pt-H
and Pt-OH) abilities; and then the bifunctional (PEMFC versus WE) cell construc-
tion is a simple routine engineering task.

8.16 Striking Electrocatalytic Conclusions

Classical potentiodynamic spectra in aqueous media, (Fig. 8.1), straight after the
reversible adsorptive Pt-OH peak, reveal unusual, anomalous, longer and stronger
anodic polarization even within the thermodynamic equilibrium potential values
(ΔE ¼ ΔEo + ɳr + ɳa), and from the same cause and reasons, along the reverse
cathodic sweep, too. Since the entire amount of reversibly adsorptive Pt-OH sooner,
yet within the same reversible potential range (up to 1.29 V vs. RHE), undergoes
complete disproportionation into highly polarizable and irreversibly deposited
Pt¼O monolayer, (Eq. 8.2), and thereby becomes unavailable for further reaction
within such a longer (about 600 mVs) potential scans, as the result, there imposes the
typical pronounced reaction polarization, (ɳr), which suppresses further proceed-
ing of the interfering Pt-OH/Pt¼O reversible reaction (Eq. 8.3), within both, back
and forth, scan directions. The latter extends further by the activation polarization
(ɳa), and all together they impose very strong and broad polarizable properties,
(Fig. 8.1). On the contrary, interactive (SMSI) supports of mixed valence hypo-d-(f)-
oxides, a priori behave as continuously renewable and recoverable enormous
(Pt-OH) latent storage capacities and spillover sources (Scheme 8.1) all along the
potential range between oxygen and hydrogen evolving limits. Thus, there is now no
reaction polarization, (ɳr ¼ 0), and since the OER occurs straight from the enormous
broad Pt-OH peak, and thereby exactly at the thermodynamic ROE value,
(ΔEo ¼ 1.29 V, ɳr ¼ ɳa ¼ 0), the reversible electrocatalysts of various interactive
hypo-d-(f)-oxide support compositions for oxygen electrode reactions (ORR, OER)
have now, finally been substantiated and tested on Pt and Au. The present conclusive
discussion strongly asserts, implies and reveals why plain, non-modified Pt by no
means can ever become reversible electrocatalyst for the oxygen electrode reactions,
or at least not within the whole potential range between hydrogen and oxygen
evolving limits (Figs. 8.1, 8.3 and 8.4).
360 M.M. Jaksic et al.

The interplay of interconnected potentiodynamic spectra (Figs. 8.1, 8.2, 8.3, and
8.5), as the main electrochemical tools, have been thoroughly employed to inves-
tigate and define spillover phenomena and overall behavior of the primary (Pt-OH),
in particular relative to the surface Pt¼O (Au¼O) oxide, and reveal therefrom the
appearance of the strong reaction polarization range and barrier, as the prevailing
effect of unavoidable broader irreversibility features of plain Pt(Au) along the
potential axis between hydrogen and oxygen evolving limits. Various individual
and/or better mixed valence of altervalent and strong interactive (SMSI) hypo-d-(f)-
oxides, as the complementary catalytic supports have been broadly introduced as
the counterbalancing composite species able to suppress and effectively eliminate
the polarization impacts, by enriched latent storage for the Pt-OH (Au-OH) spill-
over, and achieve provided and long lasting the reversible electrocatalysts for the
oxygen electrode (primarily the ORR) reactions. The entire approach has been
based on intermetallic d-d-, d-f- and f-f-interelectronic bonding features and cor-
relations, and consequently reflects in extra high stability, advanced electron
conductivity and superior (electro)catalytic activity.
The properly conceived potentiodynamic spectral interplay of three complemen-
tary different cyclic voltammograms, as the main electrochemical tools, have been
confirmed as the reliable system to investigate spillover phenomena and define
compatible electrocatalysts for partially, but strongly and broadly polarizable
electrode reaction, and then provide novel advanced electrocatalysts for the entire
scanning and applied cycle. In such a respect the best hyper-d-electrocatalyst
because of a priori advanced individual Ru-OH spillover features would be PtRu
(AuRu not yet tested and optimized), in so far best optimized hypo-d-f-interactive
supports (Nb2O5, Ta2O5, and WO3, in combination with prevailing anatase (1 0 1)
titania, TiO2, and at least doped by CeO2, GdO2, HoO2 mixed and altervalent
composites). The primary oxide spillover interference between Ru, relative to Pt,
the authors infer on, and hypo-d-d-oxides, mostly advanced by Nb2O5, Ta2O5 and
WO3, and promoted by hypo-f-oxide (CeO2, GdO2, HoO2) ingredients, bring the
overall electrocatalytic activity to remarkable higher values at higher current
densities, and mark the general way and manner of further optimization.
Hypo-hyper-d-d-intermetallic bonding (MoPt3, HfPd3), while increasing the
cohesive strength, correspondingly advances electrocatalytic activity for hydrogen
electrode reactions, on the same way being further promoted by interactive (SMSI)
hypo-d-oxide supports. Meanwhile, besides the same mixed valence hypo-hyper-d-
d-(f)-interelectronic bonding effects for the oxygen electrode reactions, the primary
oxide (PtRu-OH, Au-OH) latent storage and spillover imposes the more significant
contribution of continuous renewable and recoverable, the reversible
electrocatalysis for the oxygen electrode reactions (ORR, OER).
All interrelating steps considered, described and carried out in the course of
development of interactive hypo-hyper-d-d-(f)-oxide supported electrocatalysts for
the OER, have substantially been based on the d-d-(f)-ties and bonding strengths,
and d-(f)-electronic density of states, since d-band is stated and proved for cohesive,
adsorptive and catalytic orbital. When stated about their extra high stability in
both acidic and alkaline media5, this implies primarily metallic properties even of
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 361

hypo-d-(f)-oxides, first of all as pronounced metallic type of high electron conduc-


tivity and the overall metallic strengths, features and behavior. This is the substance
in the advanced reversible oxygen electrode electrocatalysis.

Acknowledgements The present chapter has been conceived and carried out at the Institute of
Chemical Engineering Sciences, ICEHT/FORTH, Patras, Greece.

References

1. Grove WR (1842) On a gaseous voltaic battery. Philos Mag 21:417–420


2. Conway BE (1995) Electrochemical oxide film formation at noble metals as a surface-
chemical process. Prog Surf Sci 49:331–452
3. Angerstein-Kozlowska H, Conway BE, Sharp WBA (1973) The real condition of electro-
chemically oxidized platinum surfaces: Part I. Resolution of component processes. J
Electroanal Chem 43:9–36
4. Angerstein-Kozlovska H, Conway BE, Hamelin A, Stoicoviciu L (1986/1987) Elementary
steps of electrochemical oxidation of single-crystal planes of Au. I. Chemical basis of
processes involving geometry of anions and the electrode surfaces, Electrochim Acta
31:1051–1061; II. A chemical and structural basis of oxidation of the (111) plane, J Electroanal
Chem 228:429–453
5. Jaksic JM, Krstajic NV, Vracar LJM, Neophytides SG, Labou D, Falaras P, Jaksic MM (2007)
Spillover of primary oxides as a dynamic catalytic effect of interactive hypo-d-oxide supports.
Electrochim Acta 53:349–361
6. Krstajic NV, Vracar LJM, Radmilovic VR, Neophytides SG, Labou D, Jaksic JM, Tunold R,
Falaras P, Jaksic MM (2007) Advances in interactive supported electrocatalysts for hydrogen
and oxygen electrode reactions. Surf Sci 601:1949–1966
7. Jaksic MM, Botton GA, Papakonstantinou GG, Nan F, Jaksic JM (2014) Primary oxide latent
storage and spillover enabling electrocatalysts with reversible oxygen electrode properties and
the alterpolar revertible (PEMFC versus WE) cell. J Phys Chem C 118:8723–8746
8. Papakonstantinou GD, Jaksic JM, Labou D, Siokou A, Jaksic MM (2011) Spillover phenom-
ena and their striking impacts in electrocatalysis for hydrogen and oxygen electrode reactions.
Adv Phys Chem 2011:1–22, Article ID 412165
9. Neophytides SG, Zafeiratos S, Jaksic MM (2003) Selective interactive grafting of composite
bifunctional electrocatalysts for simultaneous anodic hydrogen and CO oxidation,
I. Theoretical concepts and embodiment of novel type composite catalysts. J Electrochem
Soc 150:E512–E526
10. Neophytides SG, Murase K, Zafeiratos S, Papakonstantinou GD, Paloukis FS, Krstajic NV,
Jaksic MM (2006) Composite hypo-hyper-d-intermetallic phases as supported interactive
electrocatalysts. J Phys Chem B 110:3030–3042
11. Jaksic JM, Labou D, Papakonstantinou GD, Siokou A, Jaksic MM (2010) Novel spillover
interrelating reversible electrocatalysts for oxygen and hydrogen electrode reactions. J Phys
Chem C 114:18298–18312
12. Ma Y, Balabuena PB (2007) Designing oxygen reduction catalysts: insights from
metalloenzymes. Chem Phys Lett 440:130–133
13. Brewer L (1968) Bonding and structures of transition metals. Science 161:115–122
14. Jaksic MM (2000) Hypo-hyper-d-electronic interactive nature of synergism in catalysis and
electrocatalysis for hydrogen reactions. Electrochim Acta 45:4085–4099
15. Tauster SJ, Fung SC (1978) Strong metal-support interactions: occurrence among the binary
oxides of groups IIA–VB. J Catal 55:29–35
362 M.M. Jaksic et al.

16. Tauster SJ, Fung SC, Baker RTK, Horsley JA (1981) Strong-interactions in supported metal
catalysts. Science 211:1121–1125
17. Stevenson SA (1987) Metal-support interaction in catalysis, sintering and redispersion. Van
Nostrand, New York
18. Haller GL, Resasco DE (1989) Metal–support interaction: Group VIII metals and reducible
oxides. In: Eley DD, Pires H, Weisz PB (eds) Advances in catalysis, vol 36. Academic, San
Diego, pp 173–235
19. Neophytides SG, Zafeiratos S, Papkonstantinou GD, Jaksic JM, Paloukis FE, Jaksic MM
(2005) Extended Brewer hypo-hyper-d-interionic bonding theory, I. Theoretical consider-
ations and examples for its experimental confirmation. Int J Hydrogen Energy 30:131–147
20. Neophytides SG, Zafeiratos S, Papakonstantinou GD, Jaksic JM, Paloukis FE, Jaksic MM
(2005) Extended Brewer hypo-hyper-d-interionic bonding theory, II. Strong metal-support
interaction grafting of composite electrocatalysts. Int J Hydrogen Energy 30:393–410
21. Vittadini A, Selloni A, Rotzinger FP, Gratzel M (1998) Structure and energetics of water
adsorbed at TiO2 anatase (101) and (001) surfaces. Phys Rev Lett 81:2954–2957
22. Lazzeri M, Vittadini A, Selloni A (2001) Structure and energetics of stoichiometric TiO2
anatase surfaces. Phys Rev B 63. Article No. 155409
23. Livage J, Henry M, Sanchez C (1988) Sol-gel chemistry of transition metal oxides. Prog Solid
State Chem 18:259–341
24. Judeinstein P, Livage J (1991) Sol-gel synthesis of WO3 thin films. J Mater Chem 1:621–627
25. Livage J, Guzman G (1996) Aqueous precursors for electrochromic tungsten oxide hydrates.
Solid State Ion 84:205–211
26. Koper MTM, Van Santen RA (1999) Interaction of H, O and OH with metal surfaces. J
Electroanal Chem 472:126–136
27. Kohn HW, Boudart M (1964) Reaction of hydrogen with oxygen adsorbed on a platinum
catalyst. Science 145:149–150
28. Benson JE, Kohn HW, Boudart M (1966) On the reduction of tungsten trioxide accelerated by
platinum and water. J Catal 5:307–313
29. Boudart M, Vannice MA, Benson JE (1999) Adlineation, portholes and spillover. Z Phys
Chem NF 64:171–177
30. Volkening S, Bedurftig K, Jacobi K, Wintterlin J, Ertl G (1999) Dual path mechanism for
catalytic oxidation of hydrogen on platinum surface. Phys Rev Lett 83:2672–2675
31. Mavrikakis M, Stoltze P, Norskov JK (2000) Making gold less noble. Catal Lett 64:101–106
32. Awaludin Z, Moo JGS, Okajima T, Ohsaka T (2013) TaOx-capped Pt nanoparticles as active
and durable electrocatalysts for oxygen reduction. J Mater Chem A 1:14754–14765
33. Awaludin Z, Suzuki M, Masud J, Okajima T, Ohsaka T (2011) Enhanced electrocatalysis of
oxygen reduction on Pt/TaOx/GC. J Phys Chem C 115:25557–25567
34. Masuda T, Fukumitsu H, Fugane K, Togasaki H, Matsumura D, Tamura K, Nishihata Y,
Yoshikawa H, Kobayashi K, Mori T, Uosaki K (2012) Role of cerium oxide in the enhance-
ment of activity for the oxygen reduction reaction at Pt-CeOx nanocomposite electrocatalyst—
an in situ electrochemical X-ray absorption fine structure study. J Phys Chem C
116:10098–10102
35. Fugane K, Mori T, Ou DR, Yan P, Ye F, Yoshikawa H, Drennan J (2012) Improvement of
cathode performance on Pt-CeOx by optimization of electrochemical pretreatment conditions
for PEMFC application. Langmuir 8:16692–16700
36. Ou DR, Mori T, Fugane K, Togasaki H, Ye F, Drennan J (2011) Stability of Ceria Supports in
PtCeOx/C catalysts. J Phys Chem C 115:19239–19245
37. Masud J, Alam MT, Okajima T, Ohsaka T (2011) Catalytic electrooxidation of formaldehyde
at Ta2O5-modified Pt electrodes. Chem Lett Jpn 40:252–254
38. Masud J, Alam MT, Miah MR, Okajima T, Ohsaka T (2011) Enhanced electrooxidation of
formic acid at Ta2O5-modified Pt electrode. Electrochem Commun 13:86–89
39. Hammer B, Norskov JK (1995) Why gold is the noblest of all the metals. Nature 376:238–240
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 363

40. Quaino P, Luque NB, Nazmutdinov R, Santos E, Schmickler W (2012) Why is gold such a
good catalyst for oxygen reduction in alkaline media? Angew Chem Int Ed 51:1–5
41. Haruta M (2003) When gold is not noble: catalysis by nanoparticles. Chem Rec 3:75–87
42. Lina C, Song Y, Cao L, Chen S (2013) Oxygen reduction catalyzed by Au-TiO2
nanocomposites in alkaline media. ACS Appl Mater Interfaces 5:13305–13311
43. Jaksic MM (1986) Advances in electrocatalysis for hydrogen evolution in the light of the
Brewer-Engel valence-bond theory. J Mol Catal 38:161–202
44. Friedel J, Sayers CM (1977) On the role of d-d-electron correlations in the cohesion and
ferromagnetism of transition metals, J. Physique 38:697–705
45. Gschneidner KA (1964) Physical properties and interrelations of metallic and semimetallic
elements. In: Seitz F, Turnbull D (eds) Solid state physics, advances in research and applica-
tions, vol 16. Academic, New York, pp 275–427
46. Jaksic MM (2000) Volcano plots along the periodic table, their causes and consequences on
electrocatalysis for hydrogen electrode reactions. J New Mater Electrochem Syst 3:153–168
47. Jaksic MM, Lacnjevac CM, Grgur BN, Krstajic NV (2000) Volcano plots along intermetallic
hypo-hyper-d-electronic phase diagrams and electrocatalysis for hydrogen electrode reactions.
J New Mater Electrochem Syst 3:169–182
48. Jaksic JM, Radmilovic VR, Krstajic NV, Lacnjevac CM, Jaksic MM (2011) Volcanic period-
icity plots along transition series, Hypo-hyper-d-d-interelectronic correlations and
electrocatalysis for hydrogen electrode reactions. Macedonian J Chem Chem Eng 30:3–18
49. Methfessel M, Hennig D, Schefler M (1992) Trends of the surface relaxations, surface
energies, and work functions of the 4d transition metals. Phys Rev B 46:4816–4829
50. Kita H (1966) Periodic variation of exchange current density of hydrogen electrode reaction
with atomic number and reaction mechanism. J Electrochem Soc 113:1095–1111
51. Miles MH (1975) Evaluation of electrocatalysts for water electrolysis in alkaline solutions. J
Electroanal Chem 60:89–96
52. Trasatti S (1977) The work function in electrochemistry. In: Tobias CW, Goerischer H (eds)
Advances in electrochemistry and electrochemical engineering, vol 10. Interscience,
New York, pp 213–321
53. Jaksic JM, Vracar LJM, Neophytides SG, Zafeiratos S, Papakonstantinou GD, Krstajic NV,
Jaksic MM (2005) Structural effects on kinetic properties for hydrogen electrode reactions and
CO tolerance along Mo-Pt phase diagram. Surf Sci 598:156–173
54. Santos E, Schmickler W (2007) Electrocatalysis of hydrogen oxidation—theoretical founda-
tions. Angew Chem Int Ed 46:8262–8265
55. Hammer B, Norskov JK (2000) Theoretical surface science and catalysis—calculations and
concepts. Adv Catal 45:71–129
56. Christoffersen E, Liu P, Ruban A, Skriver HL, Norskov JK (2001) Anode materials for
low-temperature fuel cells: a density functional theory study. J Catal 199:123–131
57. Peuckert M, Coenen FP, Bonzel HP (1984) XPS study of the electrochemical surface oxidation
of platinum in 1N H2SO4 acid electrolyte. Electrochim Acta 29:1305–1314
58. Drawdy JE, Hoflund GB, Gardner SD, Yngvadottir E, Schryer DR (1990) Effect of
pretreatment on a platinized tin oxide catalyst used for low-temperature CO oxidation. Surf
Interface Anal 16:369–374
59. Akita T, Tanaka K, Tsubota S, Haruta M (2000) Analytical high-resolution TEM study of
supported gold catalysts: orientation relationship between Au particles and TiO2 supports. J
Electron Microsc 49:657–662
60. Akita T, Lu P, Ichikawa S, Tanaka K, Haruta M (2001) Analytical TEM study on the
dispersion of Au nanoparticles in Au/TiO2 catalyst prepared under various temperatures.
Surf Interface Anal 31:73–78
61. Haruta M (1997) Size- and support-dependency in the catalysis of gold. Catal Today
36:153–166
62. Date M, Haruta M (2001) Moisture effect on CO oxidation over Au/TiO2 catalyst. J Catal
201:221–224
364 M.M. Jaksic et al.

63. Boccuzzi F, Chiorino A, Tsubota S, Haruta M (1996) FTIR study of carbon monoxide
oxidation and scrambling at room temperature over gold supported on ZnO and TiO2. J Phys
Chem 100:3625–3631
64. Boccuzzi F, Chiorino A, Manzoli M, Lu P, Akita T, Ichikawa S, Haruta M (2001) Au/TiO2
nanosized samples: a catalytic, TEM, and FTIR study of the effect of calcination temperature
on the CO oxidation. J Catal 202:256–267
65. Zafeiratos S, Papakonstantinou G, Jaksic MM, Neophytides SG (2005) The effect of Mo
oxides and TiO2 support on the chemisorption features of linearly adsorbed CO on Pt
crystallites: an infrared and photoelectron spectroscopy study. J Catal 232:127–136
66. Riess I, Vayenas CG (2003) Fermi level and potential distribution in solid electrolyte cells with
and without ion spillover. Solid State Ion 159:313–329
67. Vayenas CG, Bebelis S, Pliangos C, Brosda S, Tsiplakides D (2001) Electrochemical activa-
tion of catalysis: promotion, electrochemical promotion, and metal-support interactions.
Kluwer, New York
68. Vayenas CG, Jaksic MM, Bebelis SI, Neophytides SG (1996) The electrochemical activation
of catalytic reactions. In Bockris JO’M, Conway BE, White RE (eds), Modern aspects of
electrochemistry, vol 29. Plenum Press, New York, pp 57–202
69. Vayenas CG, Bebelis S, Ladas L (1990) The dependence of catalytic activity on catalyst work
function. Nature 343:625–627
70. Tsiplakides D, Nicole J, Vayenas CG, Comninellis C (1998) Work function and catalytic
activity measurements of an IrO2 film deposited on YSZ subjected to in situ electrochemical
promotion. J Electrochem Soc 145:905–908
71. Neophytides S, Tsiplakides D, Stonehart P, Jaksic MM, Vayenas C (1994) Electrochemical
enhancement of a catalytic reaction in aqueous solution. Nature 370:45–47
72. Neophytides SG, Tsiplakides D, Stonehart P, Jaksic MM, Vayenas CG (1996) Non-faradaic
electrochemical modification of the catalytic activity of Pt for H2 oxidation in aqueous alkaline
media. J Phys Chem 100:14803–14814
73. Tsiplakides D, Neophytides SG, Enea O, Jaksic MM, Vayenas CG (1997) Non-faradaic
electrochemical modification of the catalytic activity of Pt-black electrodes deposited on
Nafion 117 solid polymer electrolyte. J Electrochem Soc 144:2072–2078
74. Tsiplakides D, Vayenas CG (2001) Electrode work function and absolute potential scale in
solid-state electrochemistry. J Electrochem Soc 148:E189–E202
75. Tsiplakides D, Archonta D, Vayenas CG (2007) Absolute potential measurements in solid and
aqueous electrochemistry using two Kelvin probes and their implications for the electrochem-
ical promotion of catalysts. Top Catal 44:469–479
76. Trasatti S (1982) The concept of absolute electrode potential. An attempt at a calculation. J
Electroanal Chem 139:1–13
77. Nicole J, Tsiplakides D, Pliangos C, Verykios XE, Comninellis C, Vayenas CG (2001)
Electrochemical promotion and metal-support interactions. J Catal 204:23–34
78. Metcalfe IS (2001) Electrochemical promotion of catalysts, I. Thermodynamic considerations.
J Catal 199:247–258
79. Metcalfe IS (2001) Electrochemical promotion of catalysts, II. The role of a stable spillover
species and prediction of reaction rate modification. J Catal 199:259–272
80. Markovic NM, Ross PN (2002) Surface science studies of model fuel cell electrocatalysts. Surf
Sci Rep 45:117–229
81. Siokou A, Ntais S (2003) Towards the preparation of realistic model Ziegler-Natta catalysts:
XPS study of the MgCl2/TiCl4 interaction with flat SiO2/Si(1 0 0). Surf Sci 540:379–388
82. Suh M, Bagus PS, Pak S, Rosynek MP, Lunsford JH (2000) Reactions of hydroxyl radicals on
titania, silica, alumina, and gold surfaces. J Phys Chem B 104:2736–2742
83. Lang ND (1989) Theory of single-atom imaging in the scanning tunneling microscope.
Comments Condens Matter Phys 14:253–257
84. Jaksic JM, Papakonstantinou GD, Labou D, Siokou A, Jaksic MM (2013) Spillover phenom-
ena in electrocatalysis for oxygen and hydrogen electrode reactions. In: Suib SL (ed) New and
8 Primary Oxide Latent Storage and Spillover for Reversible Electrocatalysis. . . 365

future developments in catalysis: hybrid materials, composites, and organocatalysts. Elsevier,


Amsterdam, pp 175–212
85. Engelhard M, Baer D (2000) Third row transition metals by X-ray photoelectron spectroscopy.
Surf Sci Spectra 7(1):1–68
86. Fuentes RE, Garcia BL, Weidner JW (2008) A Nb-doped TiO2 electrocatalyst for use in direct
methanol fuel cells. ECS Trans 12:239–248
87. Bokhimi MA, Novaro O, Lopez T, Sanchez E, Gomes R (1995) Effect of hydrolysis catalyst on
the Ti deficiency and crystallite size of sol-gel-TiO2 crystalline phases. J Mater Res
10:2788–2796
88. Arbiol J, Cerda J, Dezanneau G, Cirera A, Peiro F, Cornet A, Morante JR (2002) Effects of Nb
doping on the TiO2 anatase-to-rutile phase transition. J Appl Phys 92:853–861
89. Seah MP (1990) Quantification of AES and XPS. In: Briggs D, Seah MP (eds) Practical surface
analysis, vol 1, 2nd edn. Wiley, New York, pp 201–256
90. Simões JAM, Beauchamp JL (1990) Transition metal-hydrogen and metal-carbon bond
strengths: the keys to catalysis. Chem Rev 90:629–688
91. Anderson LC, Mooney CE, Lunsford JH (1992) Hydroxyl radical desorption from polycrys-
talline palladium: evidence for a surface phase transition. Chem Phys Lett 196:445–448
92. Hashimoto S, Matsuoka H (1992) Prolonged lifetime of electrochromism of amorphous WO3–
TiO2 thin films. Surf Interface Anal 19:464–468
93. Hashimoto S, Matsuoka H (1991) Lifetime of electrochromism of amorphous WO3‐TiO2 thin
films. J Electrochem Soc 138:2403–2408
94. Sasaki K, Zhang L, Adzic RR (2008) Niobium oxide-supported platinum ultra-low amount
electrocatalysts for oxygen reduction. Phys Chem Chem Phys 10:159–167
95. Watanabe M, Motoo S (1975) Electrocatalysis by ad-atoms: Part II. Enhancement of the
oxidation of methanol on platinum by ruthenium ad-atoms. J Electroanal Chem 60:267–273
96. Davies JC, Hayden BE, Pegg DJ, Rendall ME (2002) The electro-oxidation of carbon
monoxide on ruthenium modified Pt(1 1 1). Surf Sci 496:110–120
97. Hadzi-Jordanov S, Angerstein-Kozlowska HA, Vukovic M, Conway BE (1978) Reversibility
and growth behavior of surface oxide films at ruthenium electrode. J Electrochem Soc
125:1471–1480
Chapter 9
Metal-Organic Frameworks as Materials
for Fuel Cell Technologies

Henrietta W. Langmi, Jianwei Ren, and Nicholas M. Musyoka

9.1 Introduction

Depleting fossil fuel reserves and increasing environmental awareness have inten-
sified research efforts to develop energy technologies that can meet the escalating
demand for cheap, safe and sustainable clean energy. Among the alternative
technologies, fuel cell technology has attracted considerable attention [1–3]. A
fuel cell is an electrochemical device that directly converts the chemical energy
of a fuel into electrical energy via electrochemical reactions. Fuel cells are attrac-
tive for a number of reasons such as high efficiency and reliability, minimum
impact on environment and distinctive operating characteristics. Among the differ-
ent types of fuel cells being developed, the polymer electrolyte membrane fuel cell
(PEMFC) is considered the most appealing for a range of applications including
portable power, transportation and stationary power systems, due to its unique
features. The components of a PEMFC are an anode where hydrogen oxidation
occurs, a cathode where oxygen reduction occurs, and an electrolyte membrane that
allows protons to move from anode to cathode.
PEMFCs offer many advantages such as a low operating temperature, sustained
operation at a high current density, low weight, compactness, rapid start-ups and
suitability for discontinuous operation [4]. Despite the advantages of PEMFCs over
other types of fuel cells several hurdles including cost, performance and durability,
need to be overcome for PEMFCs to become commercially viable. There is high
catalyst cost associated with the exclusive use of Pt and Pt-based catalysts in the
fuel cell electrodes. In addition, there is inadequate performance durability as a
result of cathode catalyst oxidation, catalyst migration, loss of electrode active

H.W. Langmi (*) • J. Ren • N.M. Musyoka


HySA Infrastructure Centre of Competence, Materials Science and Manufacturing, Council
for Scientific and Industrial Research (CSIR), PO Box 395, Pretoria 0001, South Africa
e-mail: [email protected]

© Springer International Publishing Switzerland 2016 367


K.I. Ozoemena, S. Chen (eds.), Nanomaterials for Fuel Cell Catalysis,
Nanostructure Science and Technology, DOI 10.1007/978-3-319-29930-3_9
368 H.W. Langmi et al.

surface area, and corrosion of the carbon support [2, 5]. Furthermore, PEMFCs are
prone to poisoning by carbon monoxide, which can be found in the fuel in trace
amounts if it is obtained from fossil fuels. The sluggish kinetics of the oxygen
reduction reaction (ORR) necessitates the use of catalysts at the anode, and this
reaction accounts for the main source of voltage loss in low temperature PEMFCs,
even in the presence of Pt-based catalysts [6]. Presently, the state-of-the-art catalyst
for PEMFCs is supported Pt or supported Pt-based catalyst. In order to advance
PEMFC technology research efforts are geared towards developing cheap non-Pt-
based catalysts that are efficient and durable. A substantial amount of research has
also focused on developing a better polymer electrolyte membrane (PEM) for fuel
cells. The requirements for a superior PEM include high proton conductivity to
enable efficient transport of protons, good thermal and chemical stability to retain
electrolyte integrity under fuel cell operating conditions, serve as a gas tight
separator to prevent fuel gases crossover, affordability, scalability to enable high
volume production, and compatibility with other fuel cell components [7].
Fuel cells are a key component of the hydrogen energy value chain and therefore
they are an enabling technology for the Hydrogen Economy. In the same light, other
components of the hydrogen energy value chain such as hydrogen production and
storage are critical enabling technologies for fuel cells. Hydrogen is widely con-
sidered to be an attractive candidate for the replacement of current carbon-based
energy sources. It can be generated from clean and green sources and upon
combustion only produces water as by-product (without any CO2 emitted). In
addition, hydrogen has an energy content that is about three times higher than
that of other chemical fuels such as gasoline or diesel [8]. There are many
technologies for hydrogen production with steam reforming from natural gas
currently the dominant technology. Production of hydrogen from cheap and renew-
able sources is the key factor for the utilisation of hydrogen in fuel cells. However,
technological and economic issues associated with the production and supply of
hydrogen in this way will need to be overcome. Production of hydrogen from water
using renewable energy sources, such as solar and wind, is an ideal method. In
which case, the photocatalytic splitting of water is an important approach and
efforts are being made towards developing low-cost and efficient photocatalysts
for the production of hydrogen from water [9]. Currently, for the storage of
hydrogen compressed and liquefied hydrogen are established technologies which
are being employed in fuel cell vehicles [10]. However, both technologies have
major drawbacks. As very high pressures are required to achieve high storage
capacities this leads to heavy and bulky storage tanks. There are also serious safety
risks associated with both storage technologies. In addition, hydrogen has a low
boiling point (20 K) and therefore requires extensive cooling for liquefaction. This
makes the storage system complex and results in boil off losses and a loss in some
of the energy content of hydrogen [8]. Extensive research has been carried out on
hydrogen storage in solid materials [11–14]. Solid materials-based hydrogen stor-
age has an advantage over compressed and liquid hydrogen storage since hydrogen
is stored in a safer and compact manner. In materials-based storage hydrogen may
be stored through a physical adsorption process where molecular hydrogen binds
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 369

weakly to the material, or through chemisorption where a stronger binding of


atomic hydrogen occurs forming a new compound.
It is clearly evident that the development of effective, safe and affordable
enabling fuel cell technologies as well as low-cost and efficient fuel cell materials
present a serious materials challenge. Metal-organic frameworks (MOFs), also
known as coordination polymers, are an extensive class of crystalline materials
consisting of organic molecules that are bound to metals or metal clusters through
coordination bonds to form one-, two- or three-dimensional structures. They pos-
sess large internal surface areas (experimental value of up to 7140 m2 g1, theo-
retical limit of 14600 m2 g1) [15] and ultra-high porosity (up to 90 % free volume)
with pore sizes in the microporous and mesoporous ranges. The presence of both
inorganic and organic building blocks creates almost a limitless number of possible
structures and allows both the pore size and chemical environment to be tailored to
attain specific properties. These unique characteristics of MOFs make them of
tremendous interest for applications in a number of fields such as gas storage
[16], separation [17], catalysis [18], sensors [19], drug delivery [20], magnetism
[21], electrochemistry [22], clean energy [23] and fuel cells [24]. The huge interest
in MOFs and their significance to the scientific community is reflected in the
marked increasing trend in the number of MOF structures being prepared and
investigated each year [25]. This chapter provides an overview of MOFs as
promising materials for PEM fuel cell components and enabling fuel cell technol-
ogies. It first outlines the key structural characteristics of MOFs including a
statement on their design. Then the use of MOFs as PEM in fuel cells is examined.
This is followed by a review of MOFs as electrocatalysts in fuel cells. After that, the
employment of MOFs in two key enabling technologies for fuel cells i.e. hydrogen
production and storage is discussed. The chapter concludes with a summary and
outlines the prospects for MOFs in fuel cell technologies.

9.2 Structural Features of MOFs

Figure 9.1 shows some examples of MOF structures [26]. Basically, a MOF is
constructed by linking metal-containing units also known as secondary building
units (SBUs) with organic linkers through coordination bonds. The self-assembly of
these components in higher dimensions affords the desired MOF architecture.
Although the complexity of the components can differ significantly from one
assembly to another, in theory, they can usually be simplified. As Cook et al. [27]
pointed out the important fundamental characteristics of a particular MOF include
the number of binding sites available on the metal or metal clusters, the relative
orientation of these binding sites, the number of Lewis-basic sites on a ligand, the
modes of coordination and relative angularity of these sites. An appropriate choice
of inorganic building units can afford diverse SBUs alongside precise control over
predictable network topologies. Although organic building blocks possess
predetermined shape and geometry, their flexibility generally affects the final
370 H.W. Langmi et al.

Fig. 9.1 Crystal structures of a variety of MOFs. The large yellow spheres represent the cavity
space. For each MOF, the framework formula, pore size, and surface area are given. Reprinted
with permission from [26]. Copyright (2005) American Chemical Society
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 371

architecture of the MOF. The organic linkers usually carry coordinating functional
groups, which play a vital role in locking the metal ions tightly in place. Metal ions
provide inadequate directionality which leads to mobility around the metal ion
making it challenging to achieve such MOF structures as more than one structure is
possible [28].
The design of MOFs has been explained through several models including the
node-and-spacer [29], SBU [30, 31] and metal–organic polyhedral [32] models.
However, it is challenging to begin from a particular model and achieve success in
synthesising the corresponding material. In synthesising MOFs, by employing the
principle that symmetrical SBUs attached to simple linkers generates only one high
symmetry structure, enables control and prediction of the MOF structure and
properties [28]. The design of an extended network, such as a MOF, can be attained
by starting with well-defined and rigid molecular building blocks that will maintain
their structural integrity during the construction process [31, 33]. Another promis-
ing approach to the design of these extended structures is the use of well-defined
conditions that result in the formation of such building blocks in situ [31]. Yaghi
et al. [28] highlighted that the practice of logical synthesis must start with knowing
the target network ‘blueprint’ and identifying the building blocks necessary for its
assembly; the implementation of such a process is referred to as reticular synthesis.
Reticular synthesis is defined as the process of carefully assembling designed rigid
molecular building blocks into predetermined ordered structures that are held
together by strong bonding [28].
The SBU approach, which led to the design of rigid frameworks, resulted in the
identification of a small number of preferred topologies that could be targeted in
designed syntheses and was vital to attaining permanent porosity in MOFs [28]. The
geometry of the SBU was shown to be dependent on the structure of the ligand and
type of metal employed, as well as the metal-to-ligand ratio, solvent, and source of
anions to balance the charge of the metal ion [34]. The principles of geometrical
design and of tunability of the pore size by employing a known MOF as prototype
have led to the development of isoreticular MOFs (IRMOFs) [35–37]. A typical
example is the series of 16 highly crystalline IRMOFs based on the same geometry
as that of MOF-5 (also referred to as IRMOF-1) [35]. MOF-5 consists of inorganic
oxide-centred Zn4O tetrahedral clusters coordinated with six carboxylate species of
different terephthalate organic building blocks in an octahedral manner, forming a
three-dimensional cubic porous framework. The IRMOF structures are based on the
skeleton of MOF-5, wherein the pore functionality and size have been altered
without changing the original cubic topology. Therefore, the IRMOFs differ in
the nature of functional groups decorating the pores and in the metrics of their pore
structure determined by the size of the organic building-block. MOFs are also
structurally distinctive in that it is possible to construct frameworks according to
the multivariable or mixed MOFs approach [38–40]. Here the MOF contains two or
more isoreticular organic linkers, each with a different functionality that is dis-
persed within the framework in a random and homogeneous manner. The underly-
ing topology of a MOF crystal structure is described by a net, designated by a three-
letter symbol, and polyatomic groups serve as the vertices and edges of the net [41].
372 H.W. Langmi et al.

Another structural feature of MOFs is the pores, which are the empty spaces
(cavities) generated when guest molecules are removed. It is possible for the
structure to collapse during solvent removal and this is more likely to happen
when the MOF contains large pores. Permanent porosity is attained when the
framework remains intact and is easier to achieve this in microporous MOFs than
in mesoporous MOFs. The removal of weakly coordinating solvent molecules also
generates open metals sites, which can affect the properties of MOFs since open
metal sites act as Lewis acids as well as serve as adsorption sites. The inclusion of
specific functional groups in a MOF can change the chemical properties of the
MOF’s surface. Therefore, different functionalities of organic linkers can lead to
MOFs with properties tailored for specific applications. The functionalisation of a
MOF can be carried out before or after synthesis of the MOF [16]. In the former
approach known as prefunctionalisation, the presence of a specific functional group
on a ligand may sometimes prevent the formation of the targeted MOF.
Postsynthetic modification in which functional groups are incorporated after for-
mation of the desired MOF topology has emerged as a promising approach to alter
the surface of MOFs [42, 43].

9.3 MOFs as Materials for Fuel Cell Components

9.3.1 Polymer Electrolyte Membranes

The development of PEMFC technology and its demonstration in transportation


applications have grown rapidly over the past two decades. For the PEMFC to
function the PEM must (i) conduct protons but not electrons, (ii) prevent fuel
crossover and (iii) be resistant to the reducing environment at the cathode and the
oxidative environment at the anode. Several types of polymers such as Nafion,
polybenzimidazole (PBI) and sulfonated polyether-ether ketones (SPEEK) have
been employed as electrolytes in PEMFCs. Nafion, a sulfonated fluoro-polymer,
is the standard and most widely used with high proton conductivity of up to 101
S m1 at 80  C and 100 % relative humidity (RH) [22]. Proton conduction in these
polymers occurs as a result of the presence of acidic components that create
hydrophilic environments suitable for proton conduction. It is necessary to have
such high conductivities for high power density PEMFCs. A drawback associated
with these polymers arises from the humidification required which introduces
complexity. In addition, the electrolyte may physically deteriorate upon hydration
and dehydration leading to a loss in efficiency [7]. Moreover, the use of Nafion as
membranes is limited to temperatures <100  C. Efforts are being made to produce
alternative lower cost, more robust electrolyte membranes with comparable con-
ductivity to that of polymers such as Nafion. Furthermore, there is a search for novel
proton-conducting membranes with higher conductivity at temperatures above
100  C than currently available membranes. Fuel cells incorporating electrolytes
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 373

with such higher operating temperature will be advantageous in terms of better CO


tolerance, enhanced reaction kinetics and better water management.
Although MOFs are intrinsically porous crystalline materials they show great
promise as materials for membranes in PEMFCs. The reason for this is because the
guest species present in the pores of MOFs are capable of blocking the pores
thereby preventing fuel crossover from anode to cathode and subsequent voltage
loss [44]. Furthermore, the availability of counter ions in the pores of certain MOFs
can concomitantly aid in proton conduction provided they contain water, acidic
groups, or other simple functional groups containing hydrogen atoms. Moreover,
incorporation of anhydrous guest molecules, such as heterocycles, in the pores and
channels of MOFs can provide proton conduction pathways [45]. Therefore proton
conduction in MOFs can occur through water-mediated hydrogen bonding net-
works, which provide pathways for the migration of protons or through non-water
mediated proton-containing guest species that serve as proton conduction path-
ways. These proton-conduction pathways correspond to two temperature regimes;
below 100  C and above 100  C, respectively. Besides exhibiting proton conduc-
tivity as a result of inclusion of protonic charge carriers in the MOF channels the
framework itself can also display proton conductivity [45, 46]. Many research
efforts have been directed towards investigating the proton conductivity of
MOFs, and their potential for use as PEM is intensively being studied in MOFs
research for fuel cells. The reader is referred to the comprehensive reviews by Ren
et al. [24], Yoon et al. [45] and Ramaswamy et al. [7] for further reading.

9.3.1.1 Proton-Conducting MOFs Operating Below 100  C

Although there are limited studies on proton conduction of one-dimensional


(1D) coordination polymers it has been shown that there are some 1D coordination
polymers that display proton conductivity [47–50]. A typical example is ferrous
oxalate dihydrate, Fe(ox)2H2O, commonly referred to as Humboldtine
[49]. Humboldtine consists of 1D chains formed by coordination of oxalate ligands
to the equatorial positions of Fe2+ ions that act as metal nodes. Two water molecules
coordinate axially to the Fe2+ nodes forming a 1D ordered arrangement of water
molecules that interact with the framework via strong hydrogen bonds. It was
suggested that such an arrangement of water molecules could provide a
pathway for proton conduction. The proton conductivity was reported to be
1.3  103 S cm1 at 25  C and 98 % RH, which is high and comparable to Nafion.
The electrical conductivity of Humboldtine was also measured and reported to be
very low (<1010 S cm1), which will favour its application as electrolyte mem-
brane in PEMFCs. In another study, Kanaizuka et al. [47] synthesised a coordina-
tion polymer, [Zn(sbdc)(H2O)3](H2sbdc)H2O (H2sbdc: sulfone
biphenyldicarboxylic acid) consisting of a 1D chain structure of alternating Zn2+
cations and sbdc anions. Non-coordinating uncharged acid (H2sbdc) is present
between the chains with hydrogen bonding networks. The proton conductivity of
the coordination polymer was 2.5  107 S cm1 at room temperature and 95 %
374 H.W. Langmi et al.

RH, and was strongly dependent on the RH. A 1D coordination polymer, M(dhbq)
2H2O (M ¼ Mg, Mn, Co, Ni and Zn; H2(dhbq) ¼ 2,5-dihydroxy-1,4-benzoquinone)
with axially coordinated water molecules was investigated [48, 51]. While the
dihydrate showed a high proton conductivity of 4  105 S cm1, the anhydride
did not display any proton conductivity, indicating that the water molecules were
responsible for the proton conduction.
In 1979, the first investigation of proton conduction in MOFs was reported by
Kanda et al. [52] for the two-dimensional (2D) MOF, [N,N’-bis(2-hydroxyethyl)
dithiooxamido]copper(II), [(HOC2H4)2(dtoa)Cu] (dtoa ¼ dithiooxamide). Thereaf-
ter, very little attention was given to this subject of research until the early 2000s
when Kitagawa and co-workers studied a series of 2D MOFs, which could incor-
porate different concentrations of guest water molecules between the 2D sheets.
The MOFs were variations of the aforementioned 2D Cu-based MOF prepared by
substituting the hydroxyl ethyl group [53–55]. Generally, the concentration of
water molecules increases with increasing RH. Furthermore, proton conductivity
also increases with increasing RH. Therefore, it has been suggested that water
molecules, when present, play a role in proton conduction. At room temperature
and 100 % RH, [(HOC2H4)2(dtoa)Cu] exhibited a proton conductivity of 5  106
S cm1, with the MOF holding approximately 10 water molecules between the 2D
sheets at such RH similarly to the widely used Nafion [56]. The derived series of
MOFs displayed proton conductivity relatively lower than that of Nafion, ranging
from 106 to 105 S cm1 at room temperature and high RH (>75 %) [53–
56]. Higher proton conductivity would have been expected for the MOFs with
longer ligands since more water molecules can occupy the interplanar spacing of
longer ligands [54]. However, it was revealed that [(HO-H4C2)2dtoaCu] and
[(HO-H6C3)2dtoaCu] contain the same amount of water per dimeric unit under
100 % RH. A proton conductivity of 2.0  106 S cm1 was observed for
[(HO-H6C3)2dtoaCu] at 100 % RH implying that the length of the hydrophobic
moiety of the ligand does not significantly enhance proton conductivity.
To further investigate proton conductivity, Kitagawa and co-workers selected
another 2D oxalate-bridged MOF, (NH4)2(adp)[Zn2(ox)3]3H2O (adp ¼ adipic
acid) [57]. Adipic acid, NH4+ ions and water molecules occupy the pores of the
MOF, with adipic acid and NH4+ providing additional protons which favour proton
conductivity. The close proximity of the three species leads to the formation of 2D
hydrogen bond network interactions involving all the guest molecules and the
oxalate ions of the framework. Therefore, a high proton conductivity of 8  103
S cm1 at 298 K and 98 % RH was obtained, which decreased with decreasing RH
indicating that the water molecules present in the pores of the MOFs are critical for
proton conduction. A series of oxalate-based MOFs, NH(prol)3[MCr(ox)3]
(M ¼ Mn(II), Fe(II), Co(II); NH(prol)3+ ¼ tri(3-hydroxylpropyl)ammonium;
ox ¼ oxalate), was also investigated for proton conduction [58]. The MOFs consist
of oxalate-bridged bimetallic layers that are intercalated by NH(prol)3+ ions. The
MOFs showed a dependence of proton conductivity on RH and exhibited proton
conductivity of approximately 1  104 S cm1 at 75 % RH and 298 K, attributed to
the 2D hydrophilic layers formed by the NH(prol)3+ ions. The study demonstrates
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 375

Fig. 9.2 Humidity dependence of the proton conductivity at 298 K of various MOFs. Reprinted
with permission from [59]. Copyright (2012) American Chemical Society

that oxalate-bridged bimetallic complexes that contain hydrophilic ions are poten-
tially good proton conductors. Furthermore, a study was carried out to investigate
the influence of hydrophilicity of cations (containing different substituents) on the
proton conductivity of the 2D oxalate-bridged bimetallic MOF, [NR3(CH2COOH)]
[MCr(ox)3)nH2O+] (R ¼ Me (methyl), Et (ethyl), or Bu (n-butyl); and M ¼ Mn or
Fe) [59]. The RH dependence of the proton conductivity is presented in Fig. 9.2.
The carboxyl groups in the MOFs act as proton carriers with the hydrophilicity of
the cationic species controlled by NR3. The proton conductivity of the MOFs
increased with increasing adsorption of water and increasing hydrophilicity of the
cationic species in the order [NBu3(CH2COOH)]+ < [NEt3(CH2COOH)]+
< [NMe3(CH2COOH)]+. Therefore, the most hydrophilic MOF, [N(Me3)
(CH2COOH)][FeCr(ox)3], adsorbed the largest amount of water and displayed the
highest proton conductivity of 0.8  104 S cm1 even at a low RH value of 65 %
and ambient temperature.
Other 2D MOFs based on phosphonate ligands have been shown to have
significant proton conductivity because phosphonates contain three oxygen
atoms, two of which are available to act as proton acceptors following coordination
of one to the metal node [60]. Taylor et al. [60] investigated the phosphonate-based
proton-conducting MOF, PCMOF-3 (Zn3(L)H2O2H2O; L ¼ [1,3,5-
benzenetriphosphonate]6). The extensive hydrogen bonding network in this
MOF between the water molecules and the phosphonate oxygen atoms gave a
pathway for proton conduction. The proton conductivity was 3.5  105 S cm1
at 25  C and 98 % RH, and decreased to 4  108 S cm1 at 44 % RH. Kundu
et al. [61] attempted to influence proton conductivity by substituting the metal ions
376 H.W. Langmi et al.

in MOFs, and as such constructed three 2D MOFs denoted as Ca-SBBA, Sr-SBBA


and Ba-SBBA (SBBA ¼ 4,4’-sulfobisbenzoic acid). No proton conductivity was
observed in the as-synthesised MOFs. However, the MOFs displayed proton con-
ductivity only after 24 h humidification. While Ca-SBBA and Sr-SBBA exhibited a
proton conductivity of 8.6  106 and 4.4  105 S cm1, respectively at 25  C
under 98 % RH, Ba-SBBA did not display any proton conductivity under similar
conditions. The higher proton conductivity observed for Sr-SBBA was as a result of
a high carrier concentration, arising from the combination of metal clusters and
solvent molecules. It was suggested that upon humidification, water molecules are
absorbed within the crystal through hydrogen bond formation, which occurred with
carboxylate bound metal clusters and also with sulfone groups. The lack of proton
conductivity by Ba-SBBA was ascribed to phase change and loss of crystallinity
upon humidification.
Besides the 1D and 2D MOFs, proton conduction studies have also been carried
out on MOFs with 3D extended structures. An oxalate-bridged bimetallic 3D chiral
quartz-like structured MOF, (NH4)4[MnCr2(ox)6]4H2O, was shown to have a high
proton conductivity of 1.1  103 S cm1 at room temperature and 96 % RH due to
the water molecules in the functionalised channels of the MOFs being in close
proximity with NH4+ ions and water molecules around the framework [62]. The
effect of halogens and helical water chains on proton conductivity was evaluated for
four homochiral MOF isomers; [Zn(l-LCl)(Cl)](H2O)2, [Zn(l-LBr)(Br)](H2O)2, [Zn
(d-LCl)(Cl)](H2O)2, and [Zn(d-LBr)(Br)](H2O)2 (L ¼ 3-methyl-2-(pyridin-4-
ylmethylamino)butanoic acid). Both [Zn(l-LCl)(Cl)](H2O)2 and [Zn(d-LCl)(Cl)]
(H2O)2 consist of a 3D supramolecular network containing water molecules
arranged in a continuous helical chain in 1D channels (Fig. 9.3). The arrangement
of the water molecules in a helical manner is as a result of the repetition of one
water molecule forming a weak hydrogen bond with the Cl atom and another water
molecule forming a hydrogen bond with the first water molecule. Both [Zn(l-LCl)
(Cl)](H2O)2 and [Zn(d-LDCl)(Cl)](H2O)2 showed proton conductivities of
4.45  105 and 4.42  105 S cm1, respectively at 30  C and 98 % RH, which
were higher than the values for the other two MOFs.
The influence of different functional groups on proton conductivity was dem-
onstrated in a series of isostructural 3D MOFs denoted as MIL-53(M)
(MIL ¼ Matériaux de l’Institut Lavoisier; M ¼ Al or Fe) [64]. The MOFs include
Al(OH)(BDC)(H2O), Al(OH)(BDC–NH2)(H2O), Al(OH)(BDC–OH)(H2O)1.5, and
Fe(OH)(BDC–(COOH)2)(H2O) (BDC ¼ 1,4-benzenedicarboxylate). The proton
conductivity was influenced by the acidity of the functional group. In fact, the
order of proton conductivities for the MOFs with different substituents (–NH2 < –
H < –OH < –COOH) correlated well with the order of the pKa values of meta-
substituted benzoic acids (R ¼ –NH2, –H, –OH, and –COOH and pKa ¼ 4.74, 4.19,
4.08, and 3.62, respectively). This is the first example whereby proton conductivity
has been readily controlled by incorporating different functional groups in a family
of isostructural MOFs. Recently, the effect on proton conductivity of node-
coordinated species and pore-filling solvents in MOFs was demonstrated for
HKUST-1 (HKUST ¼ Hong Kong University of Science and Technology), a 3D
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 377

Fig. 9.3 (a) Polyhedral representation of the MOF, [Zn(l-LCl)(Cl)](H2O)2 lattice viewed down the
c-axis. (b) Tiling figure of [Zn(l-LCl)(Cl)](H2O)2, showing zeolitic unh-topology along the c-axis.
The tiling shows one kind of vertices, two kinds of edges, two kinds of faces, and one kind of tiles.
(c) Mirror isomers of helical water chains surrounded by a molecular helix (outer helix). The
molecular helix (outer helix) is shown as pink balls connected via gray bonds, and the helical water
chain (inner helix) is shown as red balls connected via blue rods. Reprinted with permission from
[63] Copyright (2011) American Chemical Society

MOF consisting of Cu (II) paddlewheel-type metal nodes with 1,3,5-benzene


tricarboxylate linkers [65]. Postsynthetic modification of the MOF by coordination
of H2O at open Cu node sites resulted in a large (approximately 75-fold) enhance-
ment in proton conductivity under methanol vapour relative to the MOF whose
nodes were modified by acetonitrile. In fact, the conductivity of H2O-HKUST-1
under methanol vapour was 1.5  105 S cm1. The results could be explained by
the lack of dissociable protons by acetonitrile. As methanol vapour filled the pores,
378 H.W. Langmi et al.

the coordinated‘ H2O molecules served as proton donors, which led to an


increase in the CH3OH2+ carrier concentration and consequently an increase in
the proton conductivity. Replacement of pore-filling solvent, methanol with n-
hexane led to a massive drop in conductivity, which was more than five orders
of magnitude lower than the value for the MOF containing nodes modified by H2O.
Sen et al. [66] studied a MOF, [{(Zn0.25)8(O)}Zn6(L)12(H2O)29(DMF)69(NO3)2]n
{H2L ¼ 1,3-bis(4-carboxyphenyl)imidazolium}, based on a ligand containing both
the carboxylate and imidazolium functionalities. The MOF has a 3D structure and
possesses a unique Zn8O cluster. Within the channels are solvent water and DMF as
well as nitrate anions to balance the charge. The methylene groups of the
imidazolium components are also aligned inside the channels. It was observed that
proton conductivity of the MOF increased with humidity and reached a maximum of
2.3  103 S cm1 at 25  C and 95 % RH. The result was in agreement with water
adsorption data, which revealed plateau regions at two, five and eight water mole-
cules per formula unit. The MOF showed low proton conductivity up to hydration
with four water molecules. Nevertheless, after hydration with the fifth water molecule
the proton conductivity improved considerably. This demonstrates that the extent of
hydration was a key parameter influencing proton conductivity of the MOF. More
recently, Wei et al. [67] investigated composite MOFs containing both the carbox-
ylate functionality and polyoxometalate (POM) anions. Two poly-POM-MOF com-
posites, {H[Cu(Hbpdc)(H2O)2]2[PM12O40]∙nH2O}n, (H2bpdc ¼ 2,20 -bipyridyl-
3,30 -dicarboxylic acid; M ¼ W, Mo) were synthesised and evaluated. The 3D struc-
tures of the composites was built from [Cu(Hbpdc)(H2O)2]2+, [PM12O40]3, protons
and water molecules. The 3D structures contain 1D hydrophilic channels wherein a
continuous water phase is stabilised by hydrogen bonding. Both composites exhibited
a low proton conductivity of approximately 3.0  107 S cm1 at 25  C and 98 %
RH. However, the proton conductivity increased as the temperature increased
reaching 1.25  103 and 1.56  103 S cm1 at 100  C and 98 % RH for M ¼ Mo
and W, respectively.
Costantino et al. [68] investigated two zirconium-MOFs based on phosphonate
ligands designated as Na@Zr and NH4@Zr, and containing two Na+ and two NH4+
cations, respectively in the channels. The channels also contain five water mole-
cules per formula unit, which interact strongly with each other (via hydrogen bonds)
as well as with the P-O groups on the framework surface. The Na+ and NH4+ cations
could be substituted with protons and the MOFs displayed proton conductivity
which increased with RH. Another example of proton conductivity in MOFs based
on phosphonate ligands was demonstrated where tetraphosphonate ligands were
used in the construction of La-based and Mg-based MOFs [68, 69]. Furthermore,
Colodredo et al. [70] employed carboxyphosphonate to produce a series of La-MOFs,
which displayed proton conductivity. Taylor et al. [60] reported proton conduction
in a phosphonate-based 3D MOF, [La(H5L)(H2O)4] (L ¼ 1,2,4,5-tetrakisphosphono-
methylbenzene) designated as PCMOF-5. The flexible, tetrapodal ligand enabled
the formation of a 3D framework with narrow 1D channels. The proton conductivity
was 2.5  103 S cm1 at 60  C and 98 % RH. This was due to the formation of 1D
acidic channels. Within these channels uncoordinated, diprotic phosphonic acid
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 379

groups together with monodentate hydrogen phosphonate and uncoordinated water


were present, which created facile proton conduction pathway. PCMOF-5 was also
found to be very robust demonstrated by its high humidity stability.

9.3.1.2 Proton-Conducting MOFs Operating Above 100  C

The low operating temperature of water-mediated proton conducting MOFs places


a limitation on their use for fuel cells that operate at temperatures higher than
100  C. As such some research efforts have been directed towards anhydrous
proton-conducting MOFs. These MOFs incorporate anhydrous guest species in
the channels that can provide proton conduction pathways above the dehydration
temperature of the MOFs. In this regard, various guest species such as imidazole,
triazole, histamine and ionic liquids have successfully been incorporated in MOFs
and the MOFs investigated for proton conductivity [44, 71–74]. Unlike the case of
water-mediated proton conducting MOFs the conductivity of the anhydrous proton-
conducting MOFs may not be dependent on RH.
For instance, two Al-based MOFs, Al(μ2-OH)(ndc) (ndc ¼ 1,4-naphthalenedi-
carboxylate) and Al(μ2-OH)(bdc) (bdc ¼ 1,4-benzenedicarboxylate) containing
imidazole molecules in their 1D channels were studied [71]. Imidazole is
non-volatile and possesses a high boiling point. It exists as two tautomers with a
proton that migrates between the two nitrogen atoms, giving a proton conduction
pathway. The proton conductivity at room temperature was reported to be in the
range of 108–1010 S cm1 for both imidazole-containing Al-based MOFs.
At 120  C the proton conductivity increased considerably to 2.2  105 and
1.0  107 S cm1 for Al(μ2-OH)(ndc) and Al(μ2-OH)(bdc), respectively. The
difference in the proton conductivity of the two MOFs is substantial and this was
attributed to the dynamic motion of imidazole molecules in the channels. For Al(μ2-
OH)(ndc), the hydrophobic channels interacted weakly with polar imidazole mol-
ecules allowing the imidazole molecules to move freely and conduct protons.
Meanwhile, for Al(μ2-OH)(bdc), since the channel is amphiphilic in nature, the
hydrophilic sites interacted more strongly with some of the imidazole molecules
thereby restricting the mobility of the imidazole molecules in the framework.
The effect of using histamine rather than imidazole as a proton conductor in Al
(μ2-OH)(ndc) was subsequently investigated [73]. Histamine possesses two proton-
hopping sites on the imidazole ring and one on the amine group. The loading
resulted in one histamine molecule per Al3+ ion. The histamine-loaded Al(OH)
(ndc) displayed a high proton conductivity of 1.7  103 S cm1 at 150  C under
anhydrous conditions, which was significantly higher than that of pure histamine. It
was also observed that the proton conductivity dropped to 2.1  104 S cm1 at
150  C when the histamine content was halved. Furthermore, the proton conduc-
tivity of the histamine-loaded MOF was substantially higher than that of the
corresponding aforementioned imidazole-loaded MOF. This massive difference
was ascribed to the much higher concentration of histamine, which was twice as
high as that of imidazole in the Al-based MOFs. Therefore, the concentration of the
380 H.W. Langmi et al.

proton carrier is very important in influencing proton conductivity. The histamine-


loaded MOF is viewed as a superionic conductor and the MOF hosts show promise
in creating hybrid conducting materials. Another study on histamine-loaded
Zn-MOF-74 revealed a low proton conductivity of 4.3  109 S cm1 at 146  C,
which was attributed to the low mobility of histamine in the MOF, which limited
the hopping of protons between imidazole rings [74].
A 3D proton-containing MOF (β-PCMOF2) based on a sulfonate ligand (2,4,6-
trihydroxy-1,3,5-benzenetrisulfonate) with the channels occupied by triazole,
which enhanced the proton conductivity and maximum operating temperature
considerably, was reported [44]. Similarly to imidazole, triazole has a high boiling
point and is non-volatile. The proton conductivity of the MOF could be controlled
by varying the amount of triazole that was loaded in the channels. The conductivity
was in the range 2–5  104 S cm1 at 150  C under anhydrous H2 conditions,
which was higher than the value for either pristine PCMOF2 or triazole. The high
temperature phase MOF with a partial triazole loading (β-PCMOF2(Tz)0.45) was
used as a membrane in a H2/air cell. The membrane was shown to be gas tight and
gave an open circuit voltage of 1.18 V, stable for 72 h at 100  C thus demonstrating
the applicability of MOFs in a membrane electrode assembly of a fuel cell.
A different approach to non-water mediated proton conduction is to synthesise
MOFs which possess intrinsic proton conductivity. This approach was pioneered by
Kitagawa and co-workers, who synthesised a 2D MOF, [Zn(H2PO4)2(TzH)2]n
[46]. In this MOF, extended 2D layer structures are formed by octahedrally-
coordinated Zn2+ ions connected to orthophosphates and bridging triazole mole-
cules. Substantial interlayer hydrogen bonding occurs between triaozles and ortho-
phosphates, yielding a proton conduction pathway. Proton conductivity
measurements conducted on pellets gave a value of 1.2  104 S cm1 at 150  C,
in the absence of guest proton conducting medium. Furthermore, proton conduc-
tivity measurements were also conducted on a single crystal and the results illus-
trated that the MOF conducted protons anisotropically. Kitagawa and co-workers
also investigated another zinc phosphate MOF, [Zn(HPO4)(H2PO4)2](ImH2)2
[75]. The MOF comprises tetrahedrally coordinated Zn2+ ions and two types of
orthophosphates, which form 1D chains, [Zn(HPO4)(H2PO4)2]2. Between the
chains are two crystallographically independent protonated imidazole molecules,
giving an ionic crystal system (Fig. 9.4). Considerable hydrogen bonding occurs
between protonated imidazolium cations and the orthophosphates. Proton conduc-
tivity of the MOF was temperature dependent, increasing with increasing temper-
ature reaching a value of 2.6  104 S cm1 at 130  C. The observed proton
conductivity was ascribed to the local motion of the protonated imidazolium
cations.
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 381

Fig. 9.4 (a) Crystal structure of negatively charged 1D coordination chains of [Zn(HPO4)
(H2PO4)2]2– in [Zn(HPO4)(H2PO4)2](ImH2)2 and (b) packing structure of four 1D chains along
the a axis. (c) Crystal structure of [Zn(HPO4)(H2PO4)2](ImH2)2. The ImH2+ ions are highlighted in
blue and the networks are gray. H atoms in the networks have been omitted. Reprinted with
permission from [75]. Copyright (2012) American Chemical Society

9.3.2 Catalysts or Catalyst Precursors for Oxygen Reduction


Reaction

In PEMFCs oxygen reduction reaction (ORR) is the reaction that occurs at the
cathode. It is a complex 4-electron reaction involving the breaking of a double bond
and the formation of 4 OH-bonds via several stages. The kinetics of the ORR is
sluggish necessitating the use of a catalyst. The sluggish kinetics of the ORR is one
of the major factors limiting the performance and efficiency of fuel cells. Current
fuel cell technologies employ Pt-based materials as cathode catalysts due to the
high activity of Pt for the ORR. However, the high cost, limited reserves and poor
durability of Pt-based catalysts hamper the large-scale commercialization of fuel
cells. Over the past several decades extensive research has been devoted to devel-
oping alternative electrocatalysts, i.e. non-platinum group metal (PGM) catalysts
for the ORR in fuel cells, so as to enhance their viability for practical applications.
Although the non-PGM catalysts are much cheaper and highly abundant they are
typically less efficient. In order to obtain decent efficiency higher catalyst content is
required, which increases the thickness of the cathode layer leading to poor mass
transport properties. Recently, MOFs have attracted attention as promising mate-
rials for ORR catalysts due to their excellent structural properties. The reader is
382 H.W. Langmi et al.

referred to Morozan and Jaouen [22] and Ren et al. [24] for further reading on the
subject.
Mao et al. [76] demonstrated the use of Cu-based MOFs as electrocatalysts for
ORR. Cu-BTC (BTC ¼ 1,3,5-tricarboxylate) was reported to be structurally unsta-
ble in aqueous media though it exhibited electrocatalytic activity. On the other
hand, Cu-bipy-BTC (bipy ¼ 2,20 -bipyridine) was structurally stable in water and
displayed an excellent and stable electrocatalytic activity towards almost 4-electron
reduction of O2. In a phosphate buffer (pH 6.0), Cu-bipy-BTC showed a pair of
well-defined redox peaks at ca. 0.15 V (Fig. 9.5). Upon bubbling of O2 into the

Fig. 9.5 (a) Coordination geometry of Cu atoms in Cu-bipy-BTC. (b) XRD patterns of the
as-prepared Cu-bipy-BTC (red curve) and the simulated one (black curve). (c) Typical SEM
image of fully crystallised Cu-bipy-BTC sample. (d) SEM image of Cu-bipy-BTC after being
immersed in water for 24 h. (e) Typical CVs obtained at the Cu-bipy-BTC-modified GC electrodes
in 0.10 M phosphate buffer (pH 6.0) saturated with N2 (dotted curve) or O2 (solid curve). Scan rate,
20 mV s1. (f) Typical RRDE voltammograms obtained with bare (black curves) and Cu-bipy-
BTC-modified (red curves) GC electrodes as disk electrodes (solid curves) and platinum ring
electrode (dotted curves) in 0.10 M phosphate buffer (pH 6.0) under air-saturated O2. Electrode
rotation rate, 400 rpm. Scan rate, 10 mV s1. Reprinted with permission from [76]. Copyright
(2012) Elsevier B.V.
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 383

buffer the reduction peak current at 0.20 V increased while the reversed oxidation
peak current of the redox wave decreased, indicating that ORR occurred under the
electrocatalysis of Cu-bipy-BTC. A composite graphene–metalloporphyrin MOF
(denoted as (G-dye–FeP)n) prepared from pyridine-functionalised graphene
(denoted as G-dye) and an iron porphyrin (5,10,15,20-tetrakis(4-carboxyl)-iron-
porphyrin; labelled FeP) was investigated for electrocatalytic activity [77]. G-dye
served as a building block in the assembly of the MOF. The reason for using iron
porphyrin was because metallo-porphyrins play a key role in biological reactions
including oxygen transport and reduction reactions. It was reported that the contents
of graphene and Fe porphyrin affected the crystallization process of the MOF. The
composite displayed interesting catalytic activity towards the ORR in alkaline
medium with an onset potential of ca. 0.93 V vs. reversible hydrogen electrode
with a 4-electron ORR pathway. This result was a combined effect of the larger
bond polarity arising from the nitrogen ligands in the G-dye, the catalytically active
iron–porphyrin and the porosity of the MOF.
Several studies have reported the use of MOFs denoted as ZIFs (ZIF ¼ zeolitic
imidazolate framework) as sacrificial templates for the synthesis of ORR catalysts
[78–82]. Ma et al. [78] first demonstrated the use of ZIFs-based ORR
electrocatalyst by using Co-ZIF, [Co(Im)20.5DMA]1 (Im ¼ 3,5-imidazolate;
DMA ¼ N,N-dimethylaniline), as the template. After thermal activation above
600  C the 3D MOF structure was not preserved. However, the structure enabled
the creation of a high density of Co–N4 catalytic sites which could be readily
accessed by oxygen as sufficient porosity and surface area were retained. After
pyrolysis at 750  C in an inert atmosphere, the Co-ZIF exhibited optimal stability
and activity for the ORR at pH 1, with an onset potential of 0.83 V vs. a reversible
hydrogen electrode and ORR mechanism characterised by 3.2–3.5 electrons per O2
molecule. A direct relationship existed between the surface area and the catalytic
activity. The mesopores accounted for most of the surface area formed during heat
treatment.
Proietti et al. [80] derived an electrocatalyst from a combination of iron acetate,
phenanthroline and a Zn-based MOF, ZIF-8. The catalyst displayed increased
volumetric activity and enhanced mass transport properties. ZIF-8 acted as a host
for iron acetate and phenanthroline to produce a catalyst precursor that later
underwent pyrolysis. ZIF-8 was selected as host due to its nitrogen-rich composi-
tion and large microporous surface area (1700 m2 g1), which are both crucial for
ORR activity of the Fe/N/C catalysts. After the heat treatment processes ZIF-8 was
transformed into a nitrogenated microporous carbon structure hosting Fe–Nx active
sites. A cathode made with the best Fe/N/C electrocatalyst in this study and tested in
a H2-O2 fuel cell yielded a power density of 0.75 W cm2 at 0.6 V. This value was
comparable with that of a commercial Pt-based cathode that was tested under same
conditions. The unprecedented catalytic activity was ascribed to enhanced mass
transport properties due to the porous carbon nanostructure obtained following heat
treatment. The microporous surface area of the optimised catalyst was approxi-
mately 1000 m2 g1. However, the results for durability and stability for the
384 H.W. Langmi et al.

MOF-derived Fe/C/N catalyst were relatively poor hindering its applicability on a


commercial scale.
A polymetallic Fe-based ZIF was used as a precursor to prepare Fe-based
electrocatalysts, which demonstrated excellent activity towards ORR in acidic
medium [81]. Further enhancement of the catalytic activity was achieved by
combining the electrocatalyst with ZIF-8 followed by pyrolysis. The newly derived
catalyst as cathode showed an onset potential of 0.977 V and measured volumetric
current density of 12 A cm3 at 0.8 V in a single cell test. Very recently, Zhao
et al. [82] demonstrated a facile synthesis approach for electrocatalysts involving
pyrolysis of one-pot synthesised ZIF decorated with a uniformly dispersed iron
complex. The electrocatalysts displayed excellent activity, which was influenced by
the imidazole ligands, with the highest volumetric current density of 88.1 A cm3
measured at 0.8 V in polymer electrolyte fuel cell tests. The cell performance at
higher cell currents was affected by the surface area and porosity of the catalysts.
This approach enables the feasibility of scale-up of ZIFs-based electrocatalysts and
also offers the possibility of tailored synthesis of electrocatalysts in the burgeoning
search for non-PGM catalysts.
A variation of the aforementioned study by Proietti et al. [80] was later carried
out by the same research group where either the MOF was changed to Fe-BTC
(specific surface area ¼ 1500 m2 g1; BTC ¼ 1,3,5-tricarboxylate) or the Fe pre-
cursor changed to Fe–phthalocyanine or the organic precursor substituted with
polyaniline [83]. It was observed that the changes in the electrochemical perfor-
mance were small when ZIF-8 was present in the catalyst precursor mixture.
However when ZIF-8 was substituted with Fe-BTC substantial adverse effects
were reported. The reason for this is twofold. Firstly, ZIF-8 is a Zn-based MOF
and Zn could easily evaporate at the high temperatures employed during pyrolysis.
On the other hand, Fe present in Fe-BTC accumulated during thermal treatment and
promoted graphitisation, which led to a decreased microporous surface area. In
addition, the absence of nitrogen in Fe-BTC implies that the sole source of nitrogen
in the catalyst precursor mixture for the active sites was the organic precursor. This
study demonstrates the significance of MOF composition when used to derive
non-PGM electrocatalysts by pyrolysis.

9.3.3 Anode Catalysts for Alcohol Electrooxidation

The use of alcohols is an alternative to the direct use of hydrogen in fuel cells.
Alcohols, such as methanol and ethanol, possess high energy densities, and are
stable liquids that can easily be transported and stored. Direct alcohol fuel cells
(DAFCs) are being targeted for portable electronic devices and electric vehicles due
to their high power density output and low pollutant emissions. PGM catalysts are
highly active and stable particularly in acidic medium, which makes them attractive
for use as anode catalysts for electrooxidation in fuel cells. However, PGMs are
easily poisoned by intermediate products, such as CO, from alcohol oxidation.
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 385

Moreover, PGMs are expensive and limited in supply, which hampers the
commercialisation of DAFCs. Consequently, there is need to find alternative
low-cost, abundant catalysts with high efficiency and durability. Although very
limited, mixed conduction of electrons and protons can be obtained in MOFs, which
is desirable for their use as anode catalyst for alcohol oxidation.
Yang et al. [84] demonstrated the first example of the utilization of a noble-
metal-free MOF material, [(HOC2H4)2dtoaCu] (H2dtoa ¼ dithiooxamide), as an
electrocatalyst for ethanol electrooxidation reaction. The MOF was coated on a
glassy-carbon electrode, and evaluated in 0.5 M H2SO4 solution. Cyclic
voltammetry revealed one redox peak centred at 0.35 V vs. Ag/AgCl attributed to
CuI/CuII. The ethanol electrooxidation on this MOF was dependent on the ethanol
concentration. Upon increasing the ethanol concentration there was a continuous
increase in current density. It was also noted that even when the ethanol concen-
tration was as high as 2 M, no plateau was observed in the cyclic voltammograms,
illustrating that the MOF catalyst was tolerant to the oxidation products of ethanol.
Acetaldehyde was detected as the oxidation product of ethanol. The MOF catalyst
performance in terms of oxidation potential and current density was comparable
with Pt-based catalysts, even though ethanol was only partially oxidised to acetal-
dehyde under the conditions of investigation. Recently, the electrooxidation reac-
tion mechanism of ethanol on the MOF electrocatalyst, [(HOC2H4)2dtoaCu], was
elucidated theoretically for the first time by using the density functional theory
method [85]. The indirect proton-transfer reaction from ethanol to the N atom via
the hydroxyl group in HOC2H4 was suggested to be the key mechanism controlling
the reactivity of the ethanol oxidation, since the direct proton-transfer reaction
displayed approximately 44 kcal mol1 higher activation energy. This novel etha-
nol electrooxidation reaction mechanism in the MOF is of fundamental importance
to the application of noble-metal free MOF as electrocatalysts.

9.4 MOFs as Materials for Enabling Fuel Cell


Technologies

9.4.1 Hydrogen Production

9.4.1.1 Photocatalysts for Hydrogen Production from Water

Much of the hydrogen produced in the world today is obtained from steam
reforming of natural gas, which is composed predominantly of methane. In the
future, hydrogen production from water is desirable to meet environmental and
sustainability needs. Such a supply of hydrogen will be beneficial to the fuel cells
industry. In particular, the reduction of water into hydrogen molecules in the
presence of a photocatalyst is not dependent on fossil fuels, and is therefore ideally
suited for clean energy generation. Some MOFs have been shown to display
semiconducting behaviour whereby photon absorption generates a state of charge
386 H.W. Langmi et al.

separation with a positive hole in the valence band and an electron in the conduction
band [86, 87]. Semiconductors are generally applied in photocatalysis whereby
light energy is converted into chemical energy [88]. Recently, MOFs have been
explored as photocatalysts for generating hydrogen from water.
For instance, Kataoka et al. [89] reported the first example of MOFs (Ru-based
MOFs) that function as an activity site for the reduction of water into hydrogen
molecules in the presence of Ru(bpy)32+ (bpy ¼ 2,20 -bipyridine), MV2+ (methyl
viologen; N,N0 -dimethyl-4,40 -bipyridinium), and EDTA–2Na
(EDTA ¼ ethylenediaminetetraacetic acid) under visible light irradiation. The
results showed that the highest turnover number was 8.16 based on Ru–MOFs
and 81.6 based on Ru(bpy)32+ after 4 h of light irradiation. The apparent quantum
yield was 4.82 % at 450 nm. It was also shown that a porous MOF structure was
valuable for improving the activity of the photochemical water reduction. A further
study was conducted on the Ru-MOFs modified by changing the counter-ions
within the framework to yield [Ru2( p-bdc)2X]n (bdc ¼ 1,4-benzenedicarboxylate;
X ¼ Cl,Br or BF4) [90]. Hydrogen was generated when the Ru-MOFs were
combined with Ru(bpy)32+, MV2+ and EDTA. The observed trend in catalytic
activity was as follows: [Ru2( p-bdc)2Br]n > [Ru2(p-bdc)2BF4]n > [Ru2( p-
bdc)2Cl]n. The difference in the catalytic activity was ascribed to the surface
modification, which was linked to the choice of counter ions. The MV2+ species
were physically adsorbed on the Ru-MOFs surface and not within the cavities. The
most active catalyst, [Ru2( p-bdc)2Br]n, displayed the highest catalytic activity
generating 46.7 mmol H2 with a turnover number of 18.7 after 4 h of irradiation.
The study also revealed that the catalytic activity of MOFs with counter-ions
changed significantly when the counter-ions were manipulated by molecular
catalysts.
Two porphyrin-based MOFs, PCL-1 (Ru2(H2tccp)BF4) and PCL-2 (Ru2(Zntccp)
BF4) (PCL ¼ porphyrin coordination lattice, tccp ¼ tetrakis(4-carboxyphenyl-
porphyrin)), were studied by the same group [91]. It was also shown here that
physical adsorption of MV2+ on the surface of the PCLs was necessary for
photocatalysis with the adsorbed MV2+ acting as intermolecular electron relays
from the porphyrin sites to the Ru metal centers. Irradiation of MV2+, PCL and
EDTA at 320 nm afforded turnover numbers of 20.8 and 29.9 for PCL-1 and PCL-2,
respectively. The high turnover numbers is an indication of the positive effect
porphyrin might have on MOF photocatalysts as porphyrin in this case
functioned as a photosensitiser eliminating the need for Ru(bpy)32+. In another
report, water-stable porphyrin-based MOFs, Al-PMOF (H2TCPP
[AlOH]2(DMF3(H2O)2) (H2TCPP ¼ meso-tetra(4-carboxyl-phenyl) porphyrin))
and Al/Zn-PMOF (Zn0.986(12)TCPP[AlOH]2) were investigated as photocatalysts
for hydrogen generation from water [92]. Two different hydrogen generation
systems, MOF/MV2+/EDTA/Pt and MOF/EDTA/Pt were evaluated. For
Al/Zn-PMOF and Al-PMOF hydrogen was generated at a rate of 100 and
200 mmol g1 h1, respectively after an induction period of about 3 h.
Silva et al. [93] described the activity for the photocatalytic hydrogen generation
of two Zr-based MOFs: UiO-66 ([Zr6O4(OH)4(bdc)12], where bdc ¼ 1,4-benzene
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 387

dicarboxylate; UiO ¼ University of Oslo) and UiO-66(NH2) ([Zr6O4(OH)4(ata)12],


where ata ¼ 2-aminoterephthalate). The water-stable MOFs displayed
photocatalytic activity for hydrogen generation in water/methanol upon irradiation
at wavelength longer than 300 nm. In fact, the apparent quantum yield for H2
generation using monochromatic light at 370 nm in a 3:1 water/methanol mixture
was 3.5 % for UiO-66(NH2). It was shown that the photocatalytic activity can be
increased by rational design of the MOF and by addition of a hydrogen evolution
co-catalyst. Laser flash photolysis enabled the photocatalytic activity for hydrogen
generation to be correlated with the creation of charge separated states. The
presence of the -NH2 group in the bdc ligand did not affect the photochemistry
but produced a bathochromic shift in the optical spectrum, showing the prospects of
developing more efficient MOFs for generating hydrogen from water.
In a recent report, Wang et al. [94] loaded Pt nanoparticles in phosphorescent
MOFs (Fig. 9.6) and demonstrated that the resulting Pt@MOF materials
photocatalysed the generation of hydrogen by synergistic photoexcitation of the
MOFs and electron injection into the Pt nanoparticles. In another study, MOF-253-
Pt was synthesised by immobilising a platinum complex in 2,20 -bipyridine-based

CO2H CO2H +

ZrCl4
N
HOAc K2PtCl4
+
N
DMF TEA
N
100°C hv
CO2H CO2H

1 Pt@1
H2bpdc H2L1 Zr6(O)4(OH)4(bpdc)5.94(L1)0.06

CO2H

+
ZrCl4
N TEA K2PtCl4

N DMF TEA
N hv
100°C

CO2H
2
H2L2 Zr6(O)4(OH)4(L2)6·64DMF Pt@2

Zr6(O)4(OH)4(carboxylate)12 bpdc L1 L2

Fig. 9.6 Synthesis of phosphorescent Zr-carboxylate MOFs (1 and 2) of the fcu topology and
subsequent loading of Pt nanoparticles inside MOF cavities via MOF-mediated photoreduction of
K2PtCl4 to form the Pt@1 and Pt@2 assemblies. Reprinted with permission from [94]. Copyright
(2012) American Chemical Society
388 H.W. Langmi et al.

microporous MOF (MOF-253) using a post-synthesis modification approach


[95]. The functionalised MOF-253-Pt had a dual function; as a photosensitiser
and a photocatalyst for hydrogen evolution under visible-light irradiation. The
photocatalytic activity of MOF-253-Pt was about five times higher than that of
the corresponding complex. The presence of the short Pt  Pt interactions played a
key role in enhancing the photocatalytic activity of the resulting MOF. Even more
recently, UiO-66 sensitised by adsorbed or directly added rhodamine B dye
displayed photocatalytic activity for hydrogen evolution under visible-light illumi-
nation. In the presence of Pt as a co-catalyst, the adsorbed and directly added dye
increased the photocatalytic activity remarkably to 30 and 26 times the value of
neat Pt@UiO-66, respectively [96].
Besides serving as photocatalysts or hosts for photocatalysts MOFs have been
employed as sacrificial templates or precursors of photocatalysts. For example,
deKrafft et al. [97] developed a novel MOF-templated approach to prepare a mixed
metal oxide nanocomposite material. Core-shell particles were first created and
then calcined to decompose the core, producing crystalline octahedral nanoshells
consisting of hematite Fe2O3 nanoparticles embedded in anatase TiO2 with some Fe
doping. The Fe2O3@TiO2 material possessed interesting photophysical properties
since it allowed photocatalytic hydrogen production from water using visible light,
which could not be achieved by either Fe2O3 or TiO2 alone, or a mixture of the two.
More recently, He et al [98] demonstrated that embedding CdS in MOFs could
considerably increase the photocatalytic efficiency of CdS for visible light-driven
hydrogen production.

9.4.1.2 Catalysts or Catalyst Supports for Hydrogen Production from


Hydrolysis of Chemical Hydrides

Hydrolysis of chemical hydrides such as NH3BH3 and NaBH4 for the production of
hydrogen for fuel cells is of great interest to many researchers because hydrolysis of
these hydrides combines the best properties for storage and generation of hydrogen,
such as solubility in water, fast controllable hydrolysis, overall stability and mod-
erate exothermicity [99]. In the presence of an appropriate catalyst hydrogen can be
generated from the catalytic hydrolysis of the chemical hydrides. Catalytic hydro-
lysis of chemical hydrides does not only generate hydrogen from the chemical
hydride but also the hydrogen in water can be released, providing a promising
method for on-site hydrogen supply [100]. Generally, the reactions for generation
of hydrogen from hydrolysis of NH3BH3 and NaBH4 are presented below,
respectively:

NH3 BH3 þ 2H2 O ! NH4 BO2 þ 4H2


NaBH4 þ 2H2 O ! NaBO2 þ 4H2

In one report, a Co-based MOF, ZIF-9 (ZIF ¼ zeolitic imidazolate framework), was
synthesised via solvothermal method and the feasibility of the ZIF-9 as catalyst for
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 389

30°C
600
40°C
50°C
500
H2 generation (ml)

400

300

200

100

0
0 20 40 60 80 100 120 140 160
Time (min)

Fig. 9.7 The effect of reaction temperature on hydrogen production. NaBH4: 0.5 wt%; NaOH:
5 wt.%; H2O: 50 mL, catalyst: 0.025 g. Reprinted with permission from [99] Copyright (2012)
Elsevier B.V.

hydrogen production from NaBH4 hydrolysis was reported [99]. It was revealed
that initial hydrogen generation rate of the ZIF-9 catalyst was relatively slow
because of the gradual formation of CoB active centers. However, after the forma-
tion of CoB the hydrogen generation rate increased rapidly. Indeed, the hydrogen
generation rate at 40  C reached 3641.69 mL min 1 g 1(Co) (Fig. 9.7). It was
further demonstrated that the addition of NaOH increased the hydrolysis rate of
NaBH4 considerably. Upon cycling no obvious decrease of hydrogen production
rate was observed implying ZIF-9 catalyst possessed relatively high stability.
Furthermore, ZIF-9 maintained its basic crystal structure and crystallinity, but its
long range order was altered to a certain degree during the operation.
Li et al. [101] successfully immobilised Ni nanoparticles in the MOF, ZIF-8, and
demonstrated that the immobilised Ni nanoparticles acted as the catalytic sites. High
catalytic activity and long durability for hydrogen generation from hydrolysis of
aqueous NH3BH3 at room temperature was reported. It was suggested that the
catalytic process occurred through the formation of an activated complex species
originating from interaction between NH3BH3 molecules and the surface of Ni
nanoparticles. For Ni@ZIF-8 prepared by chemical liquid deposition NH3BH3
hydrolysis could be completed with the release of hydrogen of H2/NH3BH3 ¼ 3.0
in 19 min (Ni/NH3BH3 ¼ 0.019), corresponding to a turnover frequency of 8.4 min1.
Meanwhile, Ni@ZIF-8 prepared using chemical vapour deposition displayed a higher
activity for which the reaction could be completed in 13 min (Ni/NH3BH3 ¼ 0.016),
corresponding to a turnover frequency of 14.2 min1. In another study by the same
390 H.W. Langmi et al.

group, ultrafine Pt nanoparticles were successfully immobilised inside the pores of


the MOF, MIL-101, without aggregation of Pt nanoparticles on the external surfaces
of the MOF [102]. The obtained Pt@MIL-101 was tested for the catalytic hydrolysis
of NH3BH3 and the results showed that hydrogen was generated from aqueous
NH3BH3 at room temperature in the presence of Pt@MIL-101 with different Pt
loadings. All the Pt@MIL-101 catalysts were highly active for the hydrolysis of
NH3BH3 with hydrogen release of H2/NH3BH3 ¼ 3. When 2 wt% Pt@MIL-101
(Pt/AB ¼ 0.0029 in molar ratio) was employed, the hydrolysis of NH3BH3 was
completed within 2.5 min, corresponding to 1.0  104 LH2 molPt1 min1, twice
as high as the value of 2 wt% Pt/γ-Al2O3, which is the most active Pt catalyst for this
reaction that has been reported.
MOFs have also been employed as precursors for hydrogen generation catalysts
[103, 104]. For instance, an amorphous Co(0) catalyst was synthesised by the
reduction of a MOF precursor, Co2(bdc)2(dabco) (bdc ¼ 1,4-benzenedicarboxylate;
dabco ¼ 1,4-diazabicyclo[2.2.2]octane) [104]. The active Co(0) sites were gener-
ated in situ with the reducing reaction through the micropores and channels of the
MOF. The catalytic effective Co(0) sites were surrounded by the organic linkers
and stabilised in the residue of the framework. The amorphous Co(0) catalyst
displayed highly efficient activity in the hydrolysis of NH3BH3. Hydrogen gener-
ation of a 0.32 M aqueous NH3BH3 solution completed in 1.4 min at room
temperature. The porous structure in the MOF was suggested to play a major
role. This report demonstrated that MOFs exhibiting high surface area and suffi-
cient pore structures, may ideally be suited as precursors for highly efficient
heterogeneous catalysts.

9.4.2 Hydrogen Storage

9.4.2.1 Materials for Adsorptive Hydrogen Storage

Numerous studies have revealed that MOFs can store considerable amounts of
hydrogen at liquid nitrogen temperature. At this temperature hydrogen adsorption
occurs by a physical process involving weak van der Waals forces. Adsorptive
hydrogen storage offers the benefits of fast kinetics of hydrogen uptake and release
and complete reversibility of the adsorption process. There are many structural
factors that determine hydrogen storage in MOFs such as surface area, pore size,
pore volume‘, open metal sites and ligand functionalisation. Many research efforts
are being directed towards enhancing hydrogen storage properties of MOFs so that
they can become more viable for practical hydrogen storage. For a recent review of
hydrogen storage in MOFs the reader is referred to Langmi et al. [105].
It has been reported that H2 uptake in MOFs is influenced by pore size at low
pressures, surface area at moderate pressures and free volume at high pressures
[106]. It has also been established that at 77 K and a low pressure of 1 bar, hydrogen
uptake can directly correlate with BET surface area in the range 100–2000 m2 g1
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 391

Fig. 9.8 A plot showing the relationship between H2 uptake capacities at 77 K and BET surface
areas of various MOFs. Low pressure is approximately 1 bar and high pressures are in the range of
10–90 bar. Reprinted with permission from [16]. Copyright (2012) American Chemical Society

(Fig. 9.8). Nevertheless, above 2000 m2 g1 the correlation breaks down since the
MOF surface cannot be fully covered with hydrogen molecules [16]. A series of
NbO-type MOFs (NOTT-101, PCN-46 and NOTT-102, NOTT ¼ Nottingham;
PCN ¼ porous coordination network) was studied and a correlation between hydro-
gen uptake and length of ligand was reported [107, 108]. The longer the length of
the ligand the higher the specific surface area, pore volume and high pressure
hydrogen uptake. In fact, NOTT-101 possessed the shortest ligand (5.77 Å), lowest
surface area (2316 m2 g1), pore volume (0.886 cm3 g1) and total H2 uptake at
77 K and 60 bar (6.2 wt%). Meanwhile, NOTT-102 exhibited the highest values of
10.098 Å, 2942 m2 g1, 1.138 cm3 g1, 6.7 wt% respectively.
Generally, the interaction energy of hydrogen with adsorbent surfaces is low
[109]. However, if the pores of the adsorbent are sufficiently small the potential
fields from opposite walls can overlap, enhancing the interaction potential. Wang
and Johnson [110] calculated the optimal pore size to be approximately 6 Å, which
is roughly double the kinetic diameter of the hydrogen molecule (2.89 Å). This
ideal pore size results in optimal interactions between hydrogen and the framework
as it may allow the formation of a monolayer of hydrogen molecules on opposite
walls thereby maximising the total van der Waals forces acting on the hydrogen
molecules [111]. One strategy that has been employed to tailor pore size and
enhance hydrogen interaction with the framework is catenation. Catenation is
392 H.W. Langmi et al.

defined as the intergrowth of two or more identical frameworks. In the case where
there is maximal displacement of frameworks from each other the catenation is
known as interpenetration. Meanwhile, where there is minimal displacement of the
frameworks it is regarded as interweaving [112, 113]. A typical example of the
effect of catenation was illustrated for the isomer pair PCN-60 (noncatenated) and
PCN-6 (catenated) both with formula Cu3(TATB)2 (TATB ¼ 4,40 ,400 -s-triazine-
2,4,6-triyltribenzoate) [114, 115]. At 1 bar and 77 K, hydrogen uptake for
PCN-6’ activated at 50  C was 1.35 wt% while PCN-6 activated at the same
temperature adsorbed 1.74 wt%, indicating that catenation can lead to enhancement
in H2 uptake [114]. Ma et al. [115] further illustrated from inelastic neutron
scattering studies that for both isomers hydrogen initially bound to open Cu centers
of the paddlewheel moieties with similar interaction energies. With increasing
hydrogen pressure the H2 molecules bound to or around the organic linkers. The
latter interaction was substantially stronger in catenated PCN-6 than in
non-catenated PCN-60 . Therefore, H2 adsorption at 50 bar of the catenated frame-
work PCN-6 was higher than that of non-catenated PCN-6’ (i.e. 6.7 and 4.0 wt%,
respectively). In non-catenated PCN-60 , the distance between opposite pore walls
was 21.4 Å, which was long and did not favour overlap of potential fields, leaving
an open structure with space in the middle of the pores that decreased the number of
effective binding sites.
It has been demonstrated that open metal sites in MOFs have high affinities for
hydrogen and as such play a key role in hydrogen storage [106, 116, 117]. The
isosteric heat of hydrogen adsorption gives an indication of the strength of interac-
tion between H2 molecules and a MOF. Many of the reported MOFs generally
display isosteric heat of adsorption in the range 4–12 kJ mol1 [118, 119]. Although
the inclusion of unsaturated metal sites in MOFs can enhance the binding of
hydrogen to the MOFs [117, 120, 121], it has also been shown that the alignment
of the unsaturated metal sites in relation to the H2 molecules plays a part
[122]. There have been numerous studies conducted on the NOTT-nnn series of
MOFs [107, 118, 122, 123] (Fig. 9.9). For instance, Lin et al. [107] investigated
isostructural MOFs constructed from binuclear Cu(II) paddlewheel nodes each
bridged by four carboxylate. Total hydrogen adsorption of NOTT-103 reached
77.8 mg g1 (7.22 wt%) at 77 K and 60 bar. Generally, the high hydrogen storage
capacities of this series of MOFs were as a result of a vacant coordination site at
each Cu(II) center and the large pore volumes of the MOFs. Although it was
observed that the open Cu(II) sites were the strongest adsorption sites the isosteric
heat of hydrogen adsorption for Cu(II) sites marginally exceeded those for other
sites in the MOF [107]. Similarly, the high H2 adsorption capacities for NOTT-112
and NOTT-116 were attributed to the presence of open metal sites in the
cuboctahedral cages [123]. Another polyhedral MOF, NOTT-140, was reported
[118]. The desolvated MOF with open Cu(II) sites had a total H2 adsorption of
6.0 wt% at 20 bar and 77 K and an isosteric heat of 4.15 kJ mol1 at zero coverage.
This value is low in comparison with other Cu(II)-based polyhedral MOFs due to
the different alignment of open Cu(II) sites [118]. A strategy was devised by Wang
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 393

Fig. 9.9 Synthesis of ligands and their corresponding Cu complexes [Cu(L)2].. (i) Mg, Et2O, B
(OMe)3; (ii) tBuOH/H2O, KMnO4; (iii) EtOH, H2SO4; (iv) BrRBr, Pd(PPh3)4, 80  C, 3 days; (v)
2 M NaOH, HCl. Reprinted with permission from [107]. Copyright (2009) American Chemical
Society

et al. [122] to increase the number of closest neighbouring open metal sites of each
void that hosts hydrogen, and to align these sites toward the H2 molecules, so as to
make the interaction between hydrogen and the framework stronger.
Doping of MOFs with metal ions has attracted attention as several calculations
have predicted enhancement in hydrogen uptake for MOFs doped with metal ions
such as Li+ ions. Such enhancement is associated with the stronger binding of
hydrogen with the dopant cations [124–128]. Although these predictions have
not been confirmed experimentally, a few experimental studies have been carried
out to investigate the influence of chemical doping of MOFs on hydrogen storage
[129–132]. For instance, the effect of doping MOFs with alkali metals (i.e. Li,
Na, and K) was evaluated [130, 131]. In one study, a 2-fold interwoven
MOF, Zn2(NDC)2(diPyNI) (NDC ¼ 2,6-naphthalenedicarboxylate; diPyNI ¼ N,
N0 -di-(4-pyridyl)-1,4,5,8-naphthalenetetracarboxydiimide) was first chemically
reduced and then doped with Li+, Na+ and K+ [131]. In another study, MOF-5 was
doped with the same alkali metal ions [130]. There was a considerable increase in
hydrogen uptake for the doped MOFs compared to the undoped MOFs [130, 131]. At
77 K and 1 bar, for the same amount of dopant, H2 uptake followed the trend: Li+
< Na+ < K+, in agreement with the trend of increasing size of dopant cation. How-
ever, the H2 binding strength followed the reverse trend Li+ > Na+ > K+. These
observations were ascribed to the structural changes accompanying framework
reduction, such as ligand polarizability and framework displacement, rather than
the creation of particular metal-based adsorption sites [131, 133].
Apart from the metal sites, which are preferential sites for hydrogen adsorption,
organic linkers have also been demonstrated to be favourable sites for hydrogen
adsorption. An inelastic neutron scattering spectroscopy study of the rotational
transitions of adsorbed H2 molecules on MOF-5, Zn4O(BDC)3 (BDC ¼ 1,4-
benzenedicarboxylate) showed that there are two distinct binding sites attributed
394 H.W. Langmi et al.

to hydrogen binding to Zn (II) metal centers and the BDC linkers [134]. It was
reported that polarisable linkers promote H2 uptake through improved interactions
with H2 molecules at higher surface coverage [135]. It was shown that the heat of
H2 adsorption decreased gradually from 7.1 kJ mol1 at zero surface coverage to
4.5 kJ mol1 at 7 mg g1 (7 wt%) surface coverage for MOF-326, which has
positively charged B and negatively charged N in the linkers. The polarised B-N
bonds seemed to enhance the binding of H2 to the MOF in comparison with simple
aromatic units of other MOFs [135]. Nevertheless, aromatic rings in organic ligands
can enhance the interactions between H2 molecules and the framework. In partic-
ular, the presence of tetrazine rings can result in an even stronger interaction
between H2 molecules and the framework as tetrazine rings contain more electrons
to generate an electron-rich conjugated π system for improved interaction with
hydrogen [136]. Wang et al. [137] demonstrated that incorporation of aromatic
moieties in IRMOF-3 (IR ¼ isoreticular) positively influenced the H2 binding
capacity as a result of specific interactions between H2 molecules and the added
phenyl groups. Three isostructural MOFs denoted as NOTT-113, NOTT-114 and
NOTT-115 (NOTT ¼ Nottingham) with the same cuboctahedral cage structure
constructed from 24 isophthalates from the ligands and 12 {Cu2(RCOO)4}
paddlewheel moieties were investigated [138]. The MOFs differed only in the
functionality of the central core of the hexacarboxylate ligands with
trimethylphenyl, phenylamine and triphenylamine moieties in NOTT-113,
NOTT-114 and NOTT-115, respectively. Amongst the three MOFs, NOTT-115
displayed the highest heat of adsorption, indicating that functionalisation of the
cage walls with more aromatic rings (i.e. triphenylamine) led to favourable inter-
actions between H2 and the framework. Conversely, amine functionalisation of the
ligand core in NOTT-114 weakened the H2–framework interaction.
One of the major challenges associated with MOFs is their low hydrogen storage
capacity at ambient temperature. Although MOFs display high hydrogen storage
capacities at cryogenic temperatures (for example, up to 9 wt% was reported
experimentally at 56 bar and 77 K for NU-100; NU ¼ Northwestern University
[139]), at ambient temperature hydrogen storage is typically less than 1 wt%
because of the weak binding (4–12 kJ mol1 [118, 119] between hydrogen and
the MOFs. This weak binding involves van der Waals forces that are responsible for
physisorption leading to low hydrogen storage capacities at ambient temperature. In
one theoretical study, it was predicted that the material should have an isosteric heat
of adsorption of 15.1 kJ mol1 in order to store hydrogen at ambient temperature
and approximately 30 bar, and release at about 1.5 bar [140]. In another theoretical
study on the effect of heat of adsorption on hydrogen storage and delivery between
1.5–120 bar, it was shown that the optimal isosteric heat of hydrogen adsorption is
about 20 kJ mol1 [141]. Therefore, it is crucial to increase the interactions between
hydrogen and MOFs to foster their applicability in practical hydrogen storage
systems. Even though theoretical calculations have predicted enhanced interactions
of hydrogen with MOFs and high hydrogen storage capacity at room temperature
when MOFs are doped with Li+ ion, these predictions have not been confirmed by
experiments [124–126].
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 395

An approach that has been debated in the literature for enhancement of hydrogen
uptake at room temperature is ‘spillover’. Hydrogen spillover is a well-documented
phenomenon in catalysis. It involves the dissociation of hydrogen molecules into
atoms on a supported metal surface followed by migration of the hydrogen atoms to
the support also known as primary receptor. The hydrogen atoms may further
migrate to a secondary receptor and this movement may be facilitated by
constructing a bridge between the two receptors [142]. Hydrogen spillover has
been investigated for improvement of hydrogen storage by incorporation of noble
metals (such as Pt and Pd) in MOFs [143–147]. Despite earlier reports of encour-
aging hydrogen storage capacities for MOFs by spillover at room temperature
ambiguity still exists about the mechanism involved in hydrogen storage enhance-
ment from/by spillover. While a massive improvement in hydrogen uptake was
reported in some studies no enhancement as a result of spillover was reported in
other studies. As a result spillover for hydrogen storage in MOFs has been a subject
of debate [148–151]. It is therefore important to exercise caution before attributing
hydrogen storage in MOFs to spillover from a noble metal.

9.4.2.2 Hosts for Chemical Hydrogen Storage Materials

An alternative materials-based storage method to adsorptive hydrogen storage is


chemical hydrogen storage. In this approach, hydrogen is stored via formation of
chemical bonds such as in complex hydrides. The high hydrogen storage capacities
of these materials make them particularly attractive for hydrogen storage
(e.g. NH3BH3 and LiBH4 have theoretical hydrogen storage capacities of 19.6
and 18.3 wt%, respectively). However, the poor kinetics of hydrogen uptake and
release and/or unfavourable thermodynamics for these materials presents severe
drawbacks [152, 153]. For instance, materials such as LiBH4 and MgH2 are too
stable while others like AlH3 are known to be too unstable for practical applica-
tions. In addition, materials such as MgH2 are also known to possess sluggish
kinetics. Nanoconfinement, a technique for compartmentalising nanoparticles
within porous scaffolds, is a promising approach to enhance the kinetics and
thermodynamic properties of chemical hydrogen storage materials
[154]. Nanoconfinement is attractive due to the large atomic population at the
particle surface that leads to an increase in surface energy. Such increase in the
ratio between the surface and bulk atoms has been reported to afford higher surface
area and also lead to enhanced surface reaction kinetics, particularly those involv-
ing diffusion within the particles [155]. MOFs have been employed as scaffolds for
nanoconfining a range of chemical hydrogen storage materials. Their nano-sized
pores, tunability and crystalline nature make them highly attractive for this purpose.
The imposed nanostructural geometry of the hydrides in the scaffold matrix appar-
ently generates many defects and specific interactions between the metal-hydrogen
bond and the nanoporous walls may favour hydrogen release [156]. The inclusion
of chemical hydrogen storage materials into the pores of MOFs may not only
improve the kinetics, thermodynamics and reversibility of hydrogen desorption,
396 H.W. Langmi et al.

Fig. 9.10 Schematic


representation of the
(NaAlH4)8 clusters inside
the MOF-74(Mg) pores.
The space inside the sphere
corresponds to the available
space outside of magnesium
van der Waals radius
(1.3 Å). Reprinted with
permission from
[161]. Copyright (2012)
American Chemical Society

but may also prevent the release of unwanted gases, making the materials more
practical for hydrogen storage [157–159].
Recently, Sun et al. [160] reported that when a Cu-based MOF, HKUST-1, was
used as the host for loading LiBH4, redox reactions occurred between LiBH4 and
Cu–O units causing dehydrogenation to occur at a much lower temperature. In fact,
dehydrogenation of LiBH4@Cu-MOF began at about 60  C, which is substantially
lower than the temperature for pristine LiBH4 (380  C). Remarkably, during the
loading of LiBH4 into the hydrated HKUST-1, the coordinated water molecules
present in HKUST-1 could react with LiBH4 to release H2 at room temperature. In
another report, Bhakta et al. [158] incorporated nanoclusters of NaAlH4 (4 wt%) in
HKUST-1. Nanoconfinement in this manner led to faster H2 desorption kinetics, as
such, decomposition occurred at about 100  C lower for NaAlH4@HKUST-1 than
bulk NaAlH4. Bulk NaAlH4 released 70 % of its hydrogen at about 250  C while the
nanoconfined NaAlH4 began to release hydrogen at 70  C, with 80 % of its
hydrogen released at 155  C. Therefore, NaAlH4 was destabilised in terms of H2
desorption by nanoscaling through confinement in the MOF material. Reversible
hydrogen storage by NaAlH4 confined within the pores of Ti-functionalised
MOF-74(Mg) was later reported [161]. Nanoconfinement of NaAlH4 within the
MOF pores (Fig. 9.10) led to a modification of the decomposition pathway of
NaAlH4. Both Ti-doped and undoped nano-NaAlH4@MOF-74(Mg) started to
release hydrogen at a much lower temperature of about 50  C, which was roughly
100  C lower than the onset desorption temperature for bulk NaAlH4. The presence
of Ti catalyst enabled an almost total reversibility of rehydrogenation.
Li et al. [159] used an yttrium-based MOF to nanoconfine NH3BH3. Enhanced
H2 release kinetics and lower H release temperature were reported. Nanoconfined
NH3BH3 started to desorb hydrogen at 50  C reaching a peak at 85  C, which was
30  C lower than that of neat NH3BH3. At 85  C nanoconfined NH3BH3 desorbed
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 397

8.0 wt% hydrogen within 10 min and 11 wt% within 3 h. Meanwhile, no hydrogen
desorption occurred for neat NH3BH3 at this temperature. At 95  C, H2 desorption
from nanoconfined NH3BH3 even reached 10.2 wt% in only 10 min and 13 wt% in
3 h. Moreover, the release of NH3 alongside H2 was eliminated due to the interac-
tion between open Y3+ sites and NH3BH3. In another report, Wahab et al. [162] also
pointed out that the formation of ammonia in nanoconfined NH3BH3 could be
eliminated if there are metallic sites in the host material such as MOF. Clean H2
generation and enhanced kinetics has further been demonstrated for NH3BH3
nanoconfined in Fe-MIL-53 [163]. Gadipelli et al. [164] examined
nanoconfinement of NH3BH3 within the pores of MOF-74(Mg). It was also
revealed that nanoconfinement of NH3BH3 within the pores of MOF-74(Mg) led
to significant enhancement of the dehydrogenation kinetics at temperatures below
100  C. The release of unwanted by-products (i.e. ammonia, borazine and diborane)
was also prevented. In a recent investigation, Lim et al. [157] reported the incor-
poration of magnesium nanocrystals (Mg NCs) in the pores of SNU-90’, [Zn4O
(atb)2] (SNU ¼ Seoul National University; atb ¼ aniline-2,4,6-tribenzoate). The
study showed that the hybrid hydrogen storage material possessed both
physisorption and chemisorption properties, displaying synergistic behaviour by
increasing the isosteric heat of H2 physisorption for the MOF, and decreasing the
temperatures for H2 chemisorption and desorption for magnesium. The thermal
stability of MOFs is important for their application as hosts for chemical hydrogen
storage materials. In this regard, the thermal stabilities of HKUST-1(Cu), MIL-53
(Al) and ZIF-8(Zn) in a hydrogen environment have been evaluated [165, 166].

9.5 Conclusions and Future Prospects

Fuel cells are an integral component in the hydrogen energy value chain serving as
conversion devices to convert the chemical energy of hydrogen into electrical
energy via electrochemical reactions. MOFs, a class of inorganic-organic hybrid
porous crystalline materials, are attractive for a variety of applications due to their
many unique features such as tunable porosity and functionality, extraordinarily
high surface area and huge structural diversity as a result of an almost limitless
possibility of combining metal centres and organic linkers. This chapter has
reviewed MOFs as materials for fuel cell technologies including their application
in fuel cell components (i.e. polymer electrolyte membrane and electrocatalysts)
and enabling technologies for fuel cells (i.e. hydrogen production and storage).
Several properties of polymeric membranes are required to meet certain stan-
dards for application in fuel cells, such as high proton conduction, good strength (i.
e. mechanical, chemical and thermal strength) and low gas permeability. Develop-
ment of a robust inexpensive substitute to state-of-the-art membrane like Nafion is
desirable. MOFs exhibiting significant proton conduction below 100  C mediated
by water where the hydrogen-bonding networks provide a proton-conduction path-
way, have been explored in many studies. Proton conductivity in these MOFs
398 H.W. Langmi et al.

relates directly to relative humidity. In order for MOFs to act as proton conductors
for practical applications, it is desirable that the MOFs function at relatively high
operating temperatures (>100  C) and under anhydrous conditions. In this light,
guest molecules such as imidazole, triazole, histamine and ionic liquids have been
incorporated in MOFs and these molecules play a key role in attaining high proton
conductivity in the MOFs. Significant progress has been made in proton-conducting
MOFs but there are still challenges that need to be overcome and these provide
opportunities for future research. Proton-conducting MOFs with proton conductiv-
ities matching that of Nafion have been reported. Further research on the long-term
performance of proton-conducting MOFs is warranted for their practical applica-
tion in fuel cells. The flexibility in design, and tunability of pore size, functionality,
chemical and thermal stability, will all play a role in the development of MOFs for
use as proton-conducting polymer electrolyte membranes.
One of the barriers to overcome before PEMFC can be properly commercialised
is associated with the electrocatalyst for the ORR. For operation with pure hydro-
gen and air, Pt is the most active material that is currently employed. The use of
non-PGMs as catalysts for the ORR will reduce the cost but the cell performance
will need to be improved. Although there have been limited reports on the appli-
cability of MOFs for the ORR a few studies have examined the direct use of MOFs
or MOF composites as electrocatalysts for the ORR. These studies reported inter-
esting catalytic activity of these materials for the ORR. However, there is need to
investigate a wider range of MOFs for this role before any concrete conclusions can
be made about their direct applicability as electrocatalysts for the ORR. It seems
that MOFs will have more potential as precursors or sacrificial templates for
electrocatalysts for the ORR. The studies that have been carried out using this
approach show promise. In this regard, the surface area and porosity of the MOF
will play a role in determining the properties of the derived catalysts. A wide variety
of MOFs with pores in the microporous to mesoporous range should be exploited
for this purpose, and the activity, stability and durability of the derived catalyst as
well as potential for scalability are all issues that will need to be addressed.
Hydrogen production and storage are known to be key enabling technologies for
fuel cells. Without these technologies the transition to the Hydrogen Economy
cannot be fully realised. For hydrogen generation the modifiable pore size, surface
area, metal centres and organic linkers of MOFs are major contributing factors. The
role of MOFs in hydrogen production has been demonstrated through
photocatalytic splitting of water and hydrolysis of chemical hydrides such as
NaBH4 and NH3BH3. For photocatalytic hydrogen generation, when catalysis
occurs at the inorganic cluster the reported catalytic activities are low, but when
the MOF is used as a scaffold or for light harvesting higher catalytic performances
are realised. Hydrolysis of chemical hydrides is advantageous over pyrolysis as
production of unwanted gases alongside hydrogen is eliminated. While significant
hydrogen generation has been achieved by both photocatalytic water splitting and
hydrolysis of chemical hydrides, it still remains a challenge to improve the catalytic
activity of MOFs for hydrogen production. By appropriate design of active sites and
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 399

corresponding functionalities MOFs are sure to make their mark in hydrogen


generation.
In the context of hydrogen storage, MOFs have been employed for adsorptive
hydrogen storage and as hosts for chemical hydrides. For adsorptive hydrogen
storage, the storage capacities of MOFs at cryogenic temperatures are quite prom-
ising especially at high pressures. However, as the temperature increases the
capacities decrease. Various structural and compositional factors influence hydro-
gen storage in MOFs such as surface area, pore volume, pore size, catenation, open
metal sites, dopant cations and ligand functionalisation. Despite the encouraging
storage capacities of MOFs at cryogenic temperatures the main challenge facing
MOFs for practical applications is how to enhance hydrogen adsorption at ambient
temperature. Further research should focus on optimising existing MOFs through
various modifications and functionalisation or even formation of MOF composites
to enhance storage capacities at ambient temperatures. The potential of MOFs to act
as host for chemical hydrides offers the benefits of reduced dehydrogenation
temperature and enhanced hydrogen release kinetics for the hydride, as well as
suppression of unwanted gaseous products. The design of suitable MOFs with high
thermal stability for this purpose is one hurdle that will need to be overcome.
In summary, MOFs are receiving increasing interest in fuel cells and their
enabling technologies. In this regard, significant progress has been made but
many challenges are yet to be overcome. The key to using MOFs in real fuel cell
applications may lie in modification of existing MOFs to obtain desirable properties
or synthesis of novel MOFs with tailored properties. With the many attributes of
MOFs such as the vast structural diversity, tunability of pore size and surface
functionality, extraordinary porosity and surface area, and their rich chemistry, it
can be said that the future of MOFs in fuel cell technologies is certainly promising.

References

1. Steele BCH, Heinzel A (2001) Materials for fuel-cell technologies. Nature 414:345–352
2. Bashyam R, Zelenay P (2006) A class of non-precious metal composite catalysts for fuel
cells. Nature 443:63–66
3. Kamarudin MZF, Kamarudin SK, Masdar MS, Daud WRW (2013) Review: direct ethanol
fuel cells. Int J Hydrogen Energy 38:9438–9453
4. Wee J (2007) Applications of proton exchange membrane fuel cell systems. Renew Sustain
Energy Rev 11:1720–1738
5. Xu H, Kunz R, Fenton JM (2007) Investigation of platinum oxidation in PEM fuel cells at
various relative humidities. Electrochem Solid State Lett 10:B1–B5
6. Gasteiger HA, Kocha SS, Sompalli B, Wagner FT (2005) Activity benchmarks and require-
ments for Pt, Pt-alloy, and non-Pt oxygen reduction catalysts for PEMFCs. Appl Catal B
Environ 56:9–35
7. Ramaswamy P, Wong NE, Shimizu GKH (2014) MOFs as proton conductors-challenges and
opportunities. Chem Soc Rev 43:5913–5932
8. Schlapbach L, Z€ uttel A (2001) Hydrogen-storage materials for mobile applications. Nature
414:353–358
400 H.W. Langmi et al.

9. Ni M, Leung MKH, Leung DYC, Sumathy K (2007) A review and recent developments in
photocatalytic water-splitting using TiO2 for hydrogen production. Renew Sustain Energy
Rev 11:401–425
10. Eberle U, M€uller B, Von Helmolt R (2012) Fuel cell electric vehicles and hydrogen
infrastructure: status 2012. Energy Environ Sci 5:8780–8798
11. Mandal TK, Gregory DH (2009) Hydrogen storage materials: present scenarios and future
directions. Annu Rep Prog Chem Sect A 105:21–54
12. McWhorter S, Read C, Ordaz G, Stetson N (2011) Materials-based hydrogen storage:
attributes for near-term, early market PEM fuel cells. Curr Opin Solid State Mater Sci
15:29–38
13. Langmi HW, McGrady GS (2008) Ternary nitrides for hydrogen storage: Li-B-N, Li-Al-N
and Li-Ga-N systems. J Alloys Compd 466:287–292
14. Ramirez-Cuesta AJ, Mitchell PCH, Ross DK, Georgiev PA, Anderson PA, Langmi HW,
Book D (2007) Dihydrogen in zeolite CaX-An inelastic neutron scattering study. J Alloys
Compd 446–447:393–396
15. Farha OK, Eryazici I, Jeong NC, Hauser BG, Wilmer CE, Sarjeant AA, Snurr RQ, Nguyen
ST, Yazaydin AO, Hupp JT (2012) Metal-organic framework materials with ultrahigh surface
areas: is the sky the limit? J Am Chem Soc 134:15016–15021
16. Suh MP, Park HJ, Prasad TK, Lim D (2012) Hydrogen storage in metal-organic frameworks.
Chem Rev 112:782–835
17. Li J, Sculley J, Zhou H (2012) Metal-organic frameworks for separations. Chem Rev
112:869–932
18. Yoon M, Srirambalaji R, Kim K (2012) Homochiral metal-organic frameworks for asym-
metric heterogeneous catalysis. Chem Rev 112:1196–1231
19. Kreno LE, Leong K, Farha OK, Allendorf M, Van Duyne RP, Hupp JT (2012) Metal-organic
framework materials as chemical sensors. Chem Rev 112:1105–1125
20. Horcajada P, Chalati T, Serre C, Gillet B, Sebrie C, Baati T, Eubank JF, Heurtaux D,
Clayette P, Kreuz C, Chang J, Hwang YK, Marsaud V, Bories P, Cynober L, Gil S,
Férey G, Couvreur P, Gref R (2010) Porous metal-organic-framework nanoscale carriers as
a potential platform for drug delivery and imaging. Nat Mater 9:172–178
21. Kurmoo M (2009) Magnetic metal-organic frameworks. Chem Soc Rev 38:1353–1379
22. Morozan A, Jaouen F (2012) Metal organic frameworks for electrochemical applications.
Energy Environ Sci 5:9269–9290
23. Li S, Xu Q (2013) Metal-organic frameworks as platforms for clean energy. Energy Environ
Sci 6:1656–1683
24. Ren Y, Chia GH, Gao Z (2013) Metal-organic frameworks in fuel cell technologies. Nano
Today 8:577–597
25. Furukawa H, Cordova KE, O’Keeffe M, Yaghi OM (2013) The chemistry and applications of
metal-organic frameworks. Science 341:1230444-1–1230444-1
26. Millward AR, Yaghi OM (2005) Metal-organic frameworks with exceptionally high capacity
for storage of carbon dioxide at room temperature. J Am Chem Soc 127:17998–17999
27. Cook TR, Zheng Y, Stang PJ (2013) Metal-organic frameworks and self-assembled supra-
molecular coordination complexes: comparing and contrasting the design, synthesis, and
functionality of metal-organic materials. Chem Rev 113:734–777
28. Yaghi OM, O’Keeffe M, Ockwig NW, Chae HK, Eddaoudi M, Kim J (2003) Reticular
synthesis and the design of new materials. Nature 423:705–714
29. Moulton B, Zaworotko MJ (2001) From molecules to crystal engineering: supramolecular
isomerism and polymorphism in network solids. Chem Rev 101:1629–1658
30. O’Keeffe M, Eddaoudi M, Li H, Reineke T, Yaghi OM (2000) Frameworks for extended
solids: geometrical design principles. J Solid State Chem 152:3–20
31. Eddaoudi M, Moler DB, Li H, Chen B, Reineke TM, O’Keeffe M, Yaghi OM (2001) Modular
chemistry: secondary building units as a basis for the design of highly porous and robust
metal-organic carboxylate frameworks. Acc Chem Res 34:319–330
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 401

32. Perry VIJJ, Perman JA, Zaworotko MJ (2009) Design and synthesis of metal-organic
frameworks using metal-organic polyhedra as supermolecular building blocks. Chem Soc
Rev 38:1400–1417
33. Yaghi OM, Li H, Davis C, Richardson D, Groy TL (1998) Synthetic strategies, structure
patterns, and emerging properties in the chemistry of modular porous solids. Acc Chem Res
31:474–484
34. Collins CS, Sun D, Liu W, Zuo J, Zhou H (2008) Reaction-condition-controlled formation of
secondary-building-units in three cadmium metal-organic frameworks with an orthogonal
tetrakis(tetrazolate) ligand. J Mol Struct 890:163–169
35. Eddaoudi M, Kim J, Rosi N, Vodak D, Wachter J, O’Keeffe M, Yaghi OM (2002) Systematic
design of pore size and functionality in isoreticular MOFs and their application in methane
storage. Science 295:469–472
36. Cavka JH, Jakobsen S, Olsbye U, Guillou N, Lamberti C, Bordiga S, Lillerud KP (2008) A
new zirconium inorganic building brick forming metal organic frameworks with exceptional
stability. J Am Chem Soc 130:13850–13851
37. Surblé S, Serre C, Mellot-Draznieks C, Millange F, Férey G (2006) A new isoreticular class
of metal-organic-frameworks with the MIL-88 topology. Chem Commun 3:284–286
38. Burrows AD, Frost CG, Mahon MF, Richardson C (2008) Post-synthetic modification of
tagged metal-organic frameworks. Angew Chem Int Ed 47:8482–8486
39. Kleist W, Jutz F, Maciejewski M, Baiker A (2009) Mixed-linker metal-organic frameworks
as catalysts for the synthesis of propylene carbonate from propylene oxide and CO2. Eur J
Inorg Chem 24:3552–3561
40. Deng H, Doonan CJ, Furukawa H, Ferreira RB, Towne J, Knobler CB, Wang B, Yaghi OM
(2010) Multiple functional groups of varying ratios in metal-organic frameworks. Science
327:846–850
41. O’Keeffe M, Peskov MA, Ramsden SJ, Yaghi OM (2008) The Reticular Chemistry Structure
Resource (RCSR) database of, and symbols for, crystal nets. Acc Chem Res 41:1782–1789
42. Tanabe KK, Cohen SM (2011) Postsynthetic modification of metal-organic frameworks—a
progress report. Chem Soc Rev 40:498–519
43. Wang Z, Cohen SM (2007) Postsynthetic covalent modification of a neutral metal-organic
framework. J Am Chem Soc 129:12368–12369
44. Hurd JA, Vaidhyanathan R, Thangadurai V, Ratcliffe CI, Moudrakovski IL, Shimizu GKH
(2009) Anhydrous proton conduction at 150 C in a crystalline metal-organic framework. Nat
Chem 1:705–710
45. Yoon M, Suh K, Natarajan S, Kim K (2013) Proton conduction in metal-organic frameworks
and related modularly built porous solids. Angew Chem Int Ed 52:2688–2700
46. Umeyama D, Horike S, Inukai M, Itakura T, Kitagawa S (2012) Inherent proton conduction
in a 2D coordination framework. J Am Chem Soc 134:12780–12785
47. Kanaizuka K, Iwakiri S, Yamada T, Kitagawa H (2010) Design and characterization of a
polarized coordination polymer of a zinc(II) biphenyldicarboxylate bearing a sulfone group.
Chem Lett 39:28–29
48. Morikawa S, Yamada T, Kitagawa H (2009) Crystal structure and proton conductivity of a
one-dimensional coordination polymer, {Mn(DHBQ)(H2O)2}. Chem Lett 38:654–655
49. Yamada T, Sadakiyo M, Kitagawa H (2009) High proton conductivity of one-dimensional
ferrous oxalate dihydrate. J Am Chem Soc 131:3144–3145
50. Dey C, Kundu T, Banerjee R (2012) Reversible phase transformation in proton conducting
Strandberg-type POM based metal organic material. Chem Commun 48:266–268
51. Yamada T, Morikawa S, Kitagawa H (2010) Structures and proton conductivity of
one-dimensional M(dhbq)nH2O (M ¼ Mg, Mn, Co, Ni, and Zn, H2(dhbq) ¼ 2,5-
dihydroxy-1,4-benzoquinone) promoted by connected hydrogen-bond networks with
absorbed water. Bull Chem Soc Jpn 83:42–48
52. Kanda S, Yamashita K, Ohkawa K (1979) A proton conductive coordination polymer. I. [N,
N0 -Bis(2 hydroxyethyl)dithiooxamido]copper(II). Bull Chem Soc Jpn 52:3296–3301
402 H.W. Langmi et al.

53. Nagao Y, Ikeda R, Kanda S, Kubozono Y, Kitagawa H (2002) Complex-plane impedance


study on a hydrogen-doped copper coordination polymer: N,N0 -bis-(2-hydroxyethyl)
dithiooxamidato-copper(II). Mol Cryst Liq Cryst Sci Technol Sect A Mol Cryst Liq Cryst
379:89–94
54. Nagao Y, Ikeda R, Iijima K, Kubo T, Nakasuji K, Kitagawa H (2003) A new proton-
conductive copper coordination polymer, (HOC3H6)2dtoaCu (dtoa ¼ dithiooxamide).
Synth Met 135–136:283–284
55. Nagao Y, Fujishima M, Ikeda R, Kanda S, Kitagawa H (2003) Highly proton-conductive
copper coordination polymers. Synth Met 133–134:431–432
56. Kitagawa H, Nagao Y, Fujishima M, Ikeda R, Kanda S (2003) Highly proton-conductive
copper coordination polymer, H2dtoaCu (H2dtoa ¼ dithiooxamide anion). Inorg Chem
Commun 6:346–348
57. Sadakiyo M, Yamada T, Kitagawa H (2009) Rational designs for highly proton-conductive
metal-organic frameworks. J Am Chem Soc 131:9906–9907
58. Okawa H, Shigematsu A, Sadakiyo M, Miyagawa T, Yoneda K, Ohba M, Kitagawa H (2009)
Oxalate-bridged bimetallic complexes {NH(prol)3}[MCr(ox)3] (M ¼ MnII, FeII, CoII; NH
(prol)3+ ¼ tri(3-hydroxypropyl)ammonium) exhibiting coexistent ferromagnetism and proton
conduction. J Am Chem Soc 131:13516–13522
59. Sadakiyo M, Okawa H, Shigematsu A, Ohba M, Yamada T, Kitagawa H (2012) Promotion of
low-humidity proton conduction by controlling hydrophilicity in layered metal-organic
frameworks. J Am Chem Soc 134:5472–5475
60. Taylor JM, Mah RK, Moudrakovski IL, Ratcliffe CI, Vaidhyanathan R, Shimizu GKH (2010)
Facile proton conduction via ordered water molecules in a phosphonate metal-organic
framework. J Am Chem Soc 132:14055–14057
61. Kundu T, Sahoo SC, Banerjee R (2012) Alkali earth metal (Ca, Sr, Ba) based thermostable
metal-organic frameworks (MOFs) for proton conduction. Chem Commun 48:4998–5000
62. Pardo E, Train C, Gontard G, Boubekeur K, Fabelo O, Liu H, Dkhil B, Lloret F, Nakagawa K,
Tokoro H, Ohkoshi S, Verdaguer M (2011) High proton conduction in a chiral ferromagnetic
metal-organic quartz-like framework. J Am Chem Soc 133:15328–15331
63. Sahoo SC, Kundu T, Banerjee R (2011) Helical water chain mediated proton conductivity in
homochiral metal-organic frameworks with unprecedented zeolitic unh -topology. J Am
Chem Soc 133:17950–17958
64. Shigematsu A, Yamada T, Kitagawa H (2011) Wide control of proton conductivity in porous
coordination polymers. J Am Chem Soc 133:2034–2036
65. Jeong NC, Samanta B, Lee CY, Farha OK, Hupp JT (2012) Coordination-chemistry control
of proton conductivity in the iconic metal-organic framework material HKUST-1. J Am
Chem Soc 134:51–54
66. Sen S, Nair NN, Yamada T, Kitagawa H, Bharadwaj PK (2012) High proton conductivity by
a metal-organic framework incorporating Zn 8O clusters with aligned imidazolium groups
decorating the channels. J Am Chem Soc 134:19432–19437
67. Wei M, Wang X, Duan X (2013) Crystal structures and proton conductivities of a MOF and
Two POM-MOF composites based on CuII ions and 2,20 -bipyridyl-3,30 -dicarboxylic acid.
Chem Eur J 19:1607–1616
68. Costantino F, Donnadio A, Casciola M (2012) Survey on the phase transitions and their effect
on the ion-exchange and on the proton-conduction properties of a flexible and robust Zr
phosphonate coordination polymer. Inorg Chem 51:6992–7000
69. Colodrero RMP, Olivera-Pastor P, Losilla ER, Hernández-Alonso D, Aranda MAG, Leon-
Reina L, Rius J, Demadis KD, Moreau B, Villemin D, Palomino M, Rey F, Cabeza A (2012)
High proton conductivity in a flexible, cross-linked, ultramicroporous magnesium tetrapho-
sphonate hybrid framework. Inorg Chem 51:7689–7698
70. Colodrero RMP, Papathanasiou KE, Stavgianoudaki N, Olivera-Pastor P, Losilla ER, Aranda
MAG, León-Reina L, Sanz J, Sobrados I, Choquesillo-Lazarte D, Garcı́a-Ruiz JM,
Atienzar P, Rey F, Demadis KD, Cabeza A (2012) Multifunctional luminescent and
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 403

proton-conducting lanthanide carboxyphosphonate open-framework hybrids exhibiting crys-


talline-to-amorphous- to-crystalline transformations. Chem Mater 24:3780–3792
71. Bureekaew S, Horike S, Higuchi M, Mizuno M, Kawamura T, Tanaka D, Yanai N, Kitagawa
S (2009) One-dimensional imidazole aggregate in aluminium porous coordination polymers
with high proton conductivity. Nat Mater 8:831–836
72. Chen W, Xu H, Zhuang G, Long L, Huang R, Zheng L (2011) Temperature-dependent
conductivity of Emim+ (Emim+ ¼ 1-ethyl-3-methyl imidazolium) confined in channels of a
metal-organic framework. Chem Commun 47:11933–11935
73. Umeyama D, Horike S, Inukai M, Hijikata Y, Kitagawa S (2011) Confinement of mobile
histamine in coordination nanochannels for fast proton transfer. Angew Chem Int Ed
50:11706–11709
74. Inukai M, Horike S, Umeyama D, Hijikata Y, Kitagawa S (2012) Investigation of post-
grafted groups of a porous coordination polymer and its proton conduction behavior. Dalton
Trans 41:13261–13263
75. Horike S, Umeyama D, Inukai M, Itakura T, Kitagawa S (2012) Coordination-network-based
ionic plastic crystal for anhydrous proton conductivity. J Am Chem Soc 134:7612–7615
76. Mao J, Yang L, Yu P, Wei X, Mao L (2012) Electrocatalytic four-electron reduction of
oxygen with copper(II)-based metal-organic frameworks. Electrochem Commun 19:29–31
77. Jahan M, Bao Q, Loh KP (2012) Electrocatalytically active graphene-porphyrin MOF
composite for oxygen reduction reaction. J Am Chem Soc 134:6707–6713
78. Ma S, Goenaga GA, Call AV, Liu D (2011) Cobalt imidazolate framework as precursor for
oxygen reduction reaction electrocatalysts. Chem Eur J 17:2063–2067
79. Palaniselvam T, Biswal BP, Banerjee R, Kurungot S (2013) Zeolitic imidazolate framework
(ZIF)-derived, hollow-core, nitrogen-doped carbon nanostructures for oxygen-reduction
reactions in PEFCs. Chem Eur J 19:9335–9342
80. Proietti E, Jaouen F, Lefèvre M, Larouche N, Tian J, Herranz J, Dodelet J (2011) Iron-based
cathode catalyst with enhanced power density in polymer electrolyte membrane fuel cells.
Nat Commun 2:416
81. Zhao D, Shui J, Chen C, Chen X, Reprogle BM, Wang D, Liu D (2012) Iron imidazolate
framework as precursor for electrocatalysts in polymer electrolyte membrane fuel cells.
Chem Sci 3:3200–3205
82. Zhao D, Shui J, Grabstanowicz LR, Chen C, Commet SM, Xu T, Lu J, Liu D (2014) Highly
efficient non-precious metal electrocatalysts prepared from one-pot synthesized zeolitic
imidazolate frameworks. Adv Mater 26:1093–1097
83. Lefèvre M, Dodelet JP (2012) Recent advances in non-precious metal electrocatalysts for
oxygen reduction in PEM fuel cells. ECS Trans 45:35–44
84. Yang L, Kinoshita S, Yamada T, Kanda S, Kitagawa H, Tokunaga M, Ishimoto T, Ogura T,
Nagumo R, Miyamoto A, Koyama M (2010) A metal-organic framework as an
electrocatalyst for ethanol oxidation. Angew Chem Int Ed 49:5348–5351
85. Ishimoto T, Ogura T, Koyama M, Yang L, Kinoshita S, Yamada T, Tokunaga M, Kitagawa H
(2013) A key mechanism of ethanol electrooxidation reaction in a noble-metal-free metal-
organic framework. J Phys Chem C 117:10607–10614
86. Llabrés i Xamena FX, Corma A, Garcia H (2007) Applications for Metal-Organic Frame-
works (MOFs) as quantum dot semiconductors. J Phys Chem C 111:80–85
87. Alvaro M, Carbonell E, Ferrer B, Llabrés I, Xamena FX, Garcia H (2007) Semiconductor
behavior of a metal-organic framework (MOF). Chem Eur J 13:5106–5112
88. Fox MA, Dulay MT (1993) Heterogeneous photocatalysis. Chem Rev 93:341–357
89. Kataoka Y, Sato K, Miyazaki Y, Masuda K, Tanaka H, Naito S, Mori W (2009)
Photocatalytic hydrogen production from water using porous material [Ru2(p-BDC)2]n.
Energy Environ Sci 2:397–400
90. Kataoka Y, Miyazaki Y, Sato K, Saito T, Nakanishi Y, Kiatagwa Y, Kawakami T,
Okumura M, Yamaguchi K, Mori W (2011) Modification of MOF catalysts by manipulation
of counter-ions: experimental and theoretical studies of photochemical hydrogen production
404 H.W. Langmi et al.

from water over microporous diruthenium (II, III) coordination polymers. Supramol Chem
23:287–296
91. Sato K, Kataoka Y, Mori W (2012) Photochemical production of hydrogen from water using
microporous porphyrin coordination lattices. J Nanosci Nanotechnol 12:585–590
92. Fateeva A, Chater PA, Ireland CP, Tahir AA, Khimyak YZ, Wiper PV, Darwent JR,
Rosseinsky MJ (2012) A water-stable porphyrin-based metal-organic framework active for
visible-light photocatalysis. Angew Chem Int Ed 51:7440–7444
93. Silva CG, Luz I, Llabrés I, Xamena FX, Corma A, Garcı́a H (2010) Water stable
Zr-benzenedicarboxylate metal-organic frameworks as photocatalysts for hydrogen genera-
tion. Chem Eur J 16:11133–11138
94. Wang C, Dekrafft KE, Lin W (2012) Pt nanoparticles@photoactive metal-organic frame-
works: efficient hydrogen evolution via synergistic photoexcitation and electron injection. J
Am Chem Soc 134:7211–7214
95. Zhou T, Du Y, Borgna A, Hong J, Wang Y, Han J, Zhang W, Xu R (2013) Post-synthesis
modification of a metal-organic framework to construct a bifunctional photocatalyst for
hydrogen production. Energy Environ Sci 6:3229–3234
96. He J, Wang J, Chen Y, Zhang J, Duan D, Wang Y, Yan Z (2014) A dye-sensitized Pt@UiO-
66(Zr) metal-organic framework for visible-light photocatalytic hydrogen production. Chem
Commun 50:7063–7066
97. Dekrafft KE, Wang C, Lin W (2012) Metal-organic framework templated synthesis of Fe2O3/
TiO2 nanocomposite for hydrogen production. Adv Mater 24:2014–2018
98. He J, Yan Z, Wang J, Xie J, Jiang L, Shi Y, Yuan F, Yu F, Sun Y (2013) Significantly
enhanced photocatalytic hydrogen evolution under visible light over CdS embedded on
metal-organic frameworks. Chem Commun 49:6761–6763
99. Li Q, Kim H (2012) Hydrogen production from NaBH4 hydrolysis via Co-ZIF-9 catalyst.
Fuel Process Technol 100:43–48
100. Bai Y, Wu C, Wu F, Yi B (2006) Carbon-supported platinum catalysts for on-site hydrogen
generation from NaBH4 solution. Mater Lett 60:2236–2239
101. Li P, Aranishi K, Xu Q (2012) ZIF-8 immobilized nickel nanoparticles: highly effective
catalysts for hydrogen generation from hydrolysis of ammonia borane. Chem Commun
48:3173–3175
102. Aijaz A, Karkamkar A, Choi YJ, Tsumori N, R€ onnebro E, Autrey T, Shioyama H, Xu Q
(2012) Immobilizing highly catalytically active Pt nanoparticles inside the pores of metal-
organic framework: a double solvents approach. J Am Chem Soc 134:13926–13929
103. Li Y, Song P, Zheng J, Li X (2010) Promoted H2 generation from NH3BH3 thermal
dehydrogenation catalyzed by metal-organic framework based catalysts. Chem Eur J
16:10887–10892
104. Song P, Li Y, Li W, He B, Yang J, Li X (2011) A highly efficient Co(0) catalyst derived from
metal-organic framework for the hydrolysis of ammonia borane. Int J Hydrogen Energy
36:10468–10473
105. Langmi HW, Ren J, North B, Mathe M, Bessarabov D (2014) Hydrogen storage in metal-
organic frameworks: a review. Electrochim Acta 128:368–392
106. Lin X, Jia J, Zhao X, Thomas KM, Blake AJ, Walker GS, Champness NR, Hubberstey P,
Schr€oder M (2006) High H2 adsorption by coordination-framework materials. Angew Chem
Int Ed 45:7358–7364
107. Lin X, Telepeni I, Blake AJ, Dailly A, Brown CM, Simmons JM, Zoppi M, Walker GS,
Thomas KM, Mays TJ, Hubberstey P, Champness NR, Schr€ oder M (2009) High capacity
hydrogen adsorption in Cu(II) tetracarboxylate framework materials: the role of pore size,
ligand functionalization, and exposed metal sites. J Am Chem Soc 131:2159–2171
108. Zhao D, Yuan D, Yakovenko A, Zhou H (2010) A NbO-type metal-organic framework
derived from a polyyne-coupled di-isophthalate linker formed in situ. Chem Commun
46:4196–4198
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 405

109. Zhao X, Xiao B, Fletcher AJ, Thomas KM, Bradshaw D, Rosseinsky MJ (2004) Hysteretic
adsorption and desorption of hydrogen by nanoporous metal-organic frameworks. Science
306:1012–1015
110. Wang Q, Johnson JK (1999) Molecular simulation of hydrogen adsorption in single-walled
carbon nanotubes and idealized carbon slit pores. J Chem Phys 110:577–586
111. De La Casa-Lillo MA, Lamari-Darkrim F, Cazorla-Amorós D, Linares-Solano A (2002)
Hydrogen storage in activated carbons and activated carbon fibers. J Phys Chem B
106:10930–10934
112. Batten SR, Robson R (1998) Interpenetrating nets: ordered, periodic entanglement. Angew
Chem Int Ed 37:1460–1494
113. Chen B, Eddaoudi M, Hyde ST, O’Keeffe M, Yaghi OM (2001) Interwoven metal-organic
framework on a periodic minimal surface with extra-large pores. Science 291:1021–1023
114. Ma S, Sun D, Ambrogio M, Fillinger JA, Parkin S, Zhou H (2007) Framework-catenation
isomerism in metal-organic frameworks and its impact on hydrogen uptake. J Am Chem Soc
129:1858–1859
115. Ma S, Eckert J, Forster PM, Ji WY, Young KH, Chang J, Collier CD, Parise JB, Zhou H
(2008) Further investigation of the effect of framework catenation on hydrogen uptake in
metal-organic frameworks. J Am Chem Soc 130:15896–15902
116. Dinca M, Long JR (2007) High-enthalpy hydrogen adsorption in cation-exchanged variants
of the microporous metal-organic framework Mn3[(Mn4CI) 3(BTT)8(CH3OH)10]2. J Am
Chem Soc 129:11172–11176
117. Liu Y, Eubank JF, Cairns AJ, Eckert J, Kravtsov VC, Luebke R, Eddaoudi M (2007)
Assembly of Metal-Organic Frameworks (MOFs) based on indium-trimer building blocks:
a porous MOF with soc topology and high hydrogen storage. Angew Chem Int Ed
46:3278–3283
118. Tan C, Yang S, Champness NR, Lin X, Blake AJ, Lewis W, Schr€ oder M (2011) High capacity
gas storage by a 4,8-connected metal-organic polyhedral framework. Chem Commun
47:4487–4489
119. Chen B, Zhao X, Putkham A, Hong K, Lobkovsky EB, Hurtado EJ, Fletcher AJ, Thomas KM
(2008) Surface interactions and quantum kinetic molecular sieving for H2 and D2 adsorption
on a mixed metal-organic framework material. J Am Chem Soc 130:6411–6423
120. Dinca M, Dailly A, Liu Y, Brown CM, Neumann DA, Long JR (2006) Hydrogen storage in a
microporous metal-organic framework with exposed Mn2+ coordination sites. J Am Chem
Soc 128:16876–16883
121. Latroche M, Surblé S, Serre C, Mellot-Draznieks C, Llewellyn PL, Lee J, Chang J, Sung HJ,
Férey G (2006) Hydrogen storage in the giant-pore metal-organic frameworks MIL-100 and
MIL-101. Angew Chem Int Ed 45:8227–8231
122. Wang X, Shengqian M, Rauch K, Simmons JM, Yuan D, Wang X, Yildirim T, Cole WC,
López JJ, De Meijere A, Zhou H (2008) Metal-organic frameworks based on double-bond-
coupled Di-isophthalate linkers with high hydrogen and methane uptakes. Chem Mater
20:3145–3152
123. Yan Y, Telepeni I, Yang S, Lin X, Kockelmann W, Dailly A, Blake AJ, Lewis W, Walker GS,
Allan DR, Barnett SA, Champness NR, Schr€ oder M (2010) Metal-organic polyhedral frame-
works: high H2 adsorption capacities and neutron powder diffraction studies. J Am Chem Soc
132:4092–4094
124. Blomqvist A, Araújo CM, Srepusharawoot P, Ahuja R (2007) Li-decorated metal-organic
framework 5: a route to achieving a suitable hydrogen storage medium. Proc Natl Acad Sci U
S A 104:20173–20176
125. Sang SH, Goddard WA III (2007) Lithium-doped metal-organic frameworks for reversible
H2 storage at ambient temperature. J Am Chem Soc 129:8422–8423
126. Mavrandonakis A, Tylianakis E, Stubos AK, Froudakis GE (2008) Why Li doping in MOFs
enhances H2 storage capacity? A multi-scale theoretical study. J Phys Chem C
112:7290–7294
406 H.W. Langmi et al.

127. Mavrandonakis A, Klontzas E, Tylianakis E, Froudakis GE (2009) Enhancement of hydrogen


adsorption in metal-organic frameworks by the incorporation of the sulfonate group and Li
cations. A multiscale computational study. J Am Chem Soc 131:13410–13414
128. Klontzas E, Mavrandonakis A, Tylianakis E, Froudakis GE (2008) Improving hydrogen
storage capacity of MOF by functionalization of the organic linker with lithium atoms.
Nano Lett 8:1572–1576
129. Mulfort KL, Farha OK, Stern CL, Sarjeant AA, Hupp JT (2009) Post-synthesis alkoxide
formation within metal-organic framework materials: a strategy for incorporating highly
coordinatively unsaturated metal ions. J Am Chem Soc 131:3866–3868
130. Chu C, Chen J, Lee T (2012) Enhancement of hydrogen adsorption by alkali-metal cation
doping of metal-organic framework-5. Int J Hydrogen Energy 37:6721–6726
131. Mulfort KL, Hupp JT (2008) Alkali metal cation effects on hydrogen uptake and binding in
metal-organic frameworks. Inorg Chem 47:7936–7938
132. Himsl D, Wallacher D, Hartmann M (2009) Improving the hydrogen-adsorption properties of
a hydroxy-modified MIL-53(A1) structural analogue by lithium doping. Angew Chem Int Ed
48:4639–4642
133. Mulfort KL, Hupp JT (2007) Chemical reduction of metal-organic framework materials as a
method to enhance gas uptake and binding. J Am Chem Soc 129:9604–9605
134. Rosi NL, Eckert J, Eddaoudi M, Vodak DT, Kim J, O’Keeffe M, Yaghi OM (2003) Hydrogen
storage in microporous metal-organic frameworks. Science 300:1127–1129
135. Tranchemontagne DJ, Park KS, Furukawa H, Eckert J, Knobler CB, Yaghi OM (2012)
Hydrogen storage in new metal-organic frameworks. J Phys Chem C 116:13143–13151
136. Chang Z, Zhang D, Chen Q, Li R, Hu T, Bu X (2011) Rational construction of 3D pillared
metal-organic frameworks: synthesis, structures, and hydrogen adsorption properties. Inorg
Chem 50:7555–7562
137. Wang Z, Tanabe KK, Cohen SM (2010) Tuning hydrogen sorption properties of metal-
organic frameworks by postsynthetic covalent modification. Chem Eur J 16:212–217
138. Yan Y, Blake AJ, Lewis W, Barnett SA, Dailly A, Champness NR, Schr€ oder M (2011)
Modifying cage structures in metal-organic polyhedral frameworks for H2 storage. Chem Eur
J 17:11162–11170
139. Farha OK, Yazaydin AÖ, Eryazici I, Malliakas CD, Hauser BG, Kanatzidis MG, Nguyen ST,
Snurr RQ, Hupp JT (2010) De novo synthesis of a metal-organic framework material
featuring ultrahigh surface area and gas storage capacities. Nat Chem 2:944–948
140. Bhatia SK, Myers AL (2006) Optimum conditions for adsorptive storage. Langmuir
22:1688–1700
141. Bae Y, Snurr RQ (2010) Optimal isosteric heat of adsorption for hydrogen storage and
delivery using metal-organic frameworks. Micropor Mesopor Mater 132:300–303
142. Lachawiec AJ Jr, Qi G, Yang RT (2005) Hydrogen storage in nanostructured carbons by
spillover: bridge-building enhancement. Langmuir 21:11418–11424
143. Cheon YE, Suh MP (2009) Enhanced hydrogen storage by palladium nanoparticles fabricated
in a redox-active metal-organic framework. Angew Chem Int Ed 48:2899–2903
144. Li Y, Yang RT (2006) Hydrogen storage in metal-organic frameworks by bridged hydrogen
spillover. J Am Chem Soc 128:8136–8137
145. Li Y, Yang RT (2006) Significantly enhanced hydrogen storage in metal-organic frameworks
via spillover. J Am Chem Soc 128:726–727
146. Li Y, Yang RT (2008) Hydrogen storage in metal-organic and covalent-organic frameworks
by spillover. AIChE J 54:269–279
147. Liu Y, Zeng J, Zhang J, Xu F, Sun LX (2007) Improved hydrogen storage in the modified
metal-organic frameworks by hydrogen spillover effect. Int J Hydrogen Energy
32:4005–4010
148. Hirscher M (2010) Remarks about spillover and hydrogen adsorption—comments on the
contributions of A.V. Talyzin and R.T. Yang. Micropor Mesopor Mater 135:209–210
9 Metal-Organic Frameworks as Materials for Fuel Cell Technologies 407

149. Li YW, Wang L, Yang RT (2010) Response to “hydrogen adsorption in Pt catalyst/MOF-5


materials” by Luzan and Talyzin. Micropor Mesopor Mater 135:206–208
150. Luzan SM, Talyzin AV (2011) Comment to the “response to ‘hydrogen adsorption in Pt
catalyst/MOF-5 materials ” by Li et al. Micropor Mesopor Mater 139:216–218
151. Szilágyi PA, Callini E, Anastasopol A, Kwakernaak C, Sachdeva S, Van De Krol R,
Geerlings H, Borgschulte A, Z€ uttel A, Dam B (2014) Probing hydrogen spillover in
Pd@MIL-101(Cr) with a focus on hydrogen chemisorption. Phys Chem Chem Phys
16:5803–5809
152. Bérubé V, Radtke G, Dresselhaus M, Chen G (2007) Size effects on the hydrogen storage
properties of nanostructured metal hydrides: a review. Int J Energy Res 31:637–663
153. Fichtner M (2011) Nanoconfinement effects in energy storage materials. Phys Chem Chem
Phys 13:21186–21195
154. Rude LH, Nielsen TK, Ravnsbæk DB, B€ osenberg U, Ley MB, Richter B, Arnbjerg LM,
Dornheim M, Filinchuk Y, Besenbacher F, Jensen TR (2011) Tailoring properties of boro-
hydrides for hydrogen storage: A review. Phys Status Solidi A Appl Mater Sci
208:1754–1773
155. Zlotea C, Latroche M (2013) Role of nanoconfinement on hydrogen sorption properties of
metal nanoparticles hybrids. Colloids Surf A Physicochem Eng Asp 439:117–130
156. Wu H (2008) Strategies for the improvement of the hydrogen storage properties of metal
hydride materials. ChemPhysChem 9:2157–2162
157. Lim D, Yoon JW, Ryu KY, Suh MP (2012) Magnesium nanocrystals embedded in a metal-
organic framework: hybrid hydrogen storage with synergistic effect on physi- and chemi-
sorption. Angew Chem Int Ed 51:9814–9817
158. Bhakta RK, Herberg JL, Jacobs B, Highley A, Behrens R Jr, Ockwig NW, Greathouse JA,
Allendorf MD (2009) Metal-organic frameworks as templates for nanoscale NaAIH4. J Am
Chem Soc 131:13198–13199
159. Li Z, Zhu G, Lu G, Qiu S, Yao X (2010) Ammonia borane confined by a metal-organic
framework for chemical hydrogen storage: Enhancing kinetics and eliminating ammonia. J
Am Chem Soc 132:1490–1491
160. Sun W, Li S, Mao J, Guo Z, Liu H, Dou S, Yu X (2011) Nanoconfinement of lithium
borohydride in Cu-MOFs towards low temperature dehydrogenation. Dalton Trans
40:5673–5676
161. Stavila V, Bhakta RK, Alam TM, Majzoub EH, Allendorf MD (2012) Reversible hydrogen
storage by NaAlH4 confined within a titanium-functionalized MOF-74(Mg) nanoreactor.
ACS Nano 6:9807–9817
162. Wahab MA, Zhao H, Yao XD (2012) Nano-confined ammonia borane for chemical hydrogen
storage. Front Chem Sci Eng 6:27–33
163. Srinivas G, Ford J, Zhou W, Yildirim T (2012) Zn-MOF assisted dehydrogenation of
ammonia borane: enhanced kinetics and clean hydrogen generation. Int J Hydrogen Energy
37:3633–3638
164. Gadipelli S, Ford J, Zhou W, Wu H, Udovic TJ, Yildirim T (2011) Nanoconfinement and
catalytic dehydrogenation of ammonia borane by magnesium-metal-organic-framework-74.
Chem Eur J 17:6043–6047
165. Grzech A, Yang J, Dingemans TJ, Srinivasan S, Magusin PCMM, Mulder FM (2011)
Irreversible high-temperature hydrogen interaction with the metal organic framework Cu3
(BTC)2. J Phys Chem C 115:21521–21525
166. Chen H, Wang L, Yang J, Yang RT (2013) Investigation on hydrogenation of metal-organic
frameworks HKUST-1, MIL-53, and ZIF-8 by hydrogen spillover. J Phys Chem C
117:7565–7576
Chapter 10
Sonoelectrochemical Production of Fuel Cell
Nanomaterials

Bruno G. Pollet and Petros M. Sakkas

10.1 Introduction

When oil, one of the most important energy sources in the history of mankind, was
first discovered in Pennsylvania (USA) almost 150 years ago, the fuel cell had
already been known for 20 years in 1839 by Sir William Grove who developed the
first fuel cell based on reversing the electrolysis of water by accident [1]. However,
the first operational commercial fuel cell was developed in 1950 by Francis Bacon
at Cambridge University who demonstrated a 5 kW alkaline fuel cell (AFC) and in
the 1970s International Fuel Cells developed a 12 kW AFC for NASA’s space
shuttle orbiter which provided continuous and reliable power.
Since the 1960’s until today, USA, Canada and Japan have received large
governmental R&D funding programmes (unlike the UK and Europe who have
so far received little funding in comparison) to develop further fuel cell systems for
stationary and transport applications [2]. Today, fuel cells are widely considered to
be an efficient (offering a much higher energy density and energy efficiency
compared to any other current energy storage devices), a non-polluting power
source and therefore a promising energy device for the transport, mobile and
stationary sectors [3]. However, the deployment of fuel cells has, so far, not been
very successful due to its high cost and limited durability.
A fuel cell is an ‘electrochemical’ device operating at various temperatures
(up to 1000  C) that transforms the chemical energy of a fuel (hydrogen, methanol
or natural gas etc) and an oxidant (air or pure oxygen) in the presence of a catalyst

B.G. Pollet (*)


Eau2Energy, Nottingham NG14 6DX, England, UK
e-mail: [email protected]
P.M. Sakkas
School of Chemical Engineering, National Technical University of Athens,
Zografou Campus 15780, Greece

© Springer International Publishing Switzerland 2016 409


K.I. Ozoemena, S. Chen (eds.), Nanomaterials for Fuel Cell Catalysis,
Nanostructure Science and Technology, DOI 10.1007/978-3-319-29930-3_10
410 B.G. Pollet and P.M. Sakkas

Fig. 10.1 Schematic of various fuel cells

into water, heat and electricity. Furthermore, the power generated by a fuel cell
depends largely upon catalytic electrodes and materials used [1–4].
There are currently six main groupings of fuel cell available namely [5]
(Fig. 10.1): (i) Proton Exchange Membrane Fuel Cell (PEMFC) including Direct
Methanol Fuel Cell (DMFC), (ii) Alkaline Fuel Cell (AFC), (iii) Phosphoric Acid
Fuel Cell (PAFC), (iv) Molten Carbonate Fuel Cell (MCFC), (v) Solid Oxide Fuel
Cell (SOFC) and (vi) Microbial Fuel Cell (MFC). PEMFC, AFC, PAFC, and MFC
operate at low temperatures in the range of [50–200  C] while SOFC and MCFC at
high operating temperatures in the range of [650–1000  C].
The heart of fuel cells consists of a non-conductive electrolyte material
sandwiched between two electrodes; the anode and cathode (Fig. 10.1). The fuel
and the oxidant are fed continuously to the anode and the cathode sides respec-
tively. At the anode side the fuel is decomposed into ions and electrons. The
insulator electrolyte material allows only ions to flow from both the anode and
cathode sides. Free electrons generated at the anode flow to the cathode side
through an external electrical circuit. The recombination of the ions with oxidant
occurs at the cathode to form pure water. Contrarily to water electrolysis, the
polarity of a fuel cell on the anode is negative and on the cathode is positive [5].
Currently, the focus in Fuel Cell technologies is to offer significant quantitative
improvements in: (a) lowering costs (by a factor of 10–100) and significantly
improving reliability, durability and performance for the stationary, transport and
portable markets; (b) addressing the challenges related to hydrogen generation,
storage and utilisation; and (c) the acceleration of fuel cell technologies deployment
to various markets such as cars and buildings [1].
10 Sonoelectrochemical Production of Fuel Cell Nanomaterials 411

As all fuel cells are normally distinguished by the materials used, e.g., the
electrolyte and catalyst materials, although the manufacturing of the fuel cell
electrodes is also different in each case, the main objective in fuel cell technologies
is to develop low-cost, high-performance and durable materials. Currently, fuel cell
systems are too expensive and not durable. For example, taking the case of
PEMFCs, the platinum (Pt) catalyst accounts for more than half [the average
price of platinum has risen more than four times in 10 years between 2000 and
2010] of the total stack cost with Pt tending to aggregate or/and dissolve when used
in long operating conditions [4].
There are several ways to reduce the cost and increasing the performance of a
fuel cell by [1–5]:
(i)decreasing the catalyst loading in fuel cell electrodes,
(ii)decreasing the catalyst nanoparticle size,
(iii)developing Pt-free electrocatalysts,
(iv) using novel fabrication methods to synthesize catalysts and producing better
catalyst dispersion on fuel cell electrodes yielding better utilisation,
(v) using new techniques to increase mass-transport at the fuel cell electrode
surface.
One of the most promising method and technique for producing performing and
efficient fuel cell catalysts, electrodes and electrolyte materials is to employ
ultrasound.

10.2 Ultrasound

Many of us are familiar with the use of ultrasound [6, 7], for example in biochem-
istry ultrasound is commonly used for the disruption of cells and living tissues in
order to extract effectively important constituents. Ultrasound is also used as a
medical imaging tool (e.g. pre-natal image scanning) and as a diagnostic tool
(e.g. non-destructive testing of materials) in the frequency range of 2–10 MHz,
but recently there has been an upsurge of interest in the application of
low-frequency high-energy power ultrasound (20 kHz–2 MHz). First observed in
the nineteenth century with the discovery of the piezoelectric effect by Curie and
the ultrasonic whistle by Galton, ultrasound is usually defined as a sound wave with
a frequency above 16 kHz (16,000 Hz or 16,000 cycles per second) with the upper
limit usually taken to be 5 MHz for gases and 500 MHz for liquids and solids
(Fig. 10.2).
The application of ultrasound in chemical, physical and biological sciences can
be divided into two main groups: (i) low frequency or power ultrasound
(20–100 kHz), also known in the 1950s as “Macrosonics” and (ii) high frequency
or diagnostic ultrasound (2–10 MHz).
412 B.G. Pollet and P.M. Sakkas

Fig. 10.2 The frequency ranges of (ultra)sound

Diagnostic ultrasound is often used in chemical analysis, medical scanning and


in the study of relaxation phenomena [6, 7]. Low amplitude waves are used to
determine the velocity and absorption coefficient of the sound wave by the medium,
i.e. the effect of the medium on the ultrasonic wave. However, power ultrasound
can be regarded as the effect of the sound wave on the medium. Power ultrasound
used in liquid systems causes (i) a zone of extreme ‘mixing’ close to the ultrasonic
source (i.e. the ultrasonic transducer), (ii) degassing, (iii) surface cleaning and
pitting (erosion) and, (iv) an increase in bulk temperature (Fig. 10.3). It is for
these reasons that low frequency and high energy waves are used in ultrasonic
cleaning, drilling, soldering, chemical processes and emulsification [1].
Ultrasonic cleaning is probably the most common and known application of
ultrasound. However, there are several areas where power ultrasound has been
successfully employed (Fig. 10.4) such as: water and soil remediation
(e.g. destruction of bacteria and organics, heavy metals removal, etc.), manufactur-
ing of food ingredients and products (e.g. emulsification of oil/water based fluids,
flavourings and vitamins extraction, etc.), impregnation of various materials,
crystallisation and precipitation of organic and inorganic compounds, nano-sized
materials production, polymerisation, drilling, soldering, cutting, plastic welding,
surface treatment and preparation (for activation and modification) prior to plating
and electroplating, metal finishing and precision engineering (e.g. the aerospace
industry) [6, 7].
More recently, power ultrasound has been successfully demonstrated as an
effective industrial process intensification technique to treat water effluents and
produce pharmaceutical materials. The systems are used as non-chemical processes
to either (i) control and eradicate microbes and bacteria (microbial activity) in
10 Sonoelectrochemical Production of Fuel Cell Nanomaterials 413

Fig. 10.3 The effect of ultrasound in a liquid and near a surface

Fig. 10.4 The use of ultrasound in industrial applications

contaminated soils and waters (in other words, power ultrasound acting as a
powerful bactericide) mainly in the water treatment industry, or (ii) produce
micro- and nano-sized pharmaceutical ingredients (process known as Sonocrystal-
lisation) in the fine chemicals and pharmaceutical industries. The systems are safe,
robust and proven technologies.
414 B.G. Pollet and P.M. Sakkas

10.3 What Is Sonochemistry?

Over the past few years the use of power ultrasound has found wide applications in
the chemical and processing industries where it is used to enhance both synthetic
and catalytic processes and to generate new products. This area of research has been
termed sonochemistry, which mainly concerns reactions involving a liquid leading
to an increase in reaction rates, product yields and erosion of surfaces [6, 7]. How-
ever, the main reason for most of the observed effects of ultrasound on surfaces and
chemical reactions is recognised as being due to ‘cavitation’ effect which occurs as
a secondary effect when an ultrasonic wave passes through a liquid medium.
Cavitation was first reported in 1895 by Thornycroft and Barnaby [8] when they
observed that the propeller of a submarine eroded over short operating times,
caused by collapsing bubbles induced by hydrodynamic cavitation in turns gener-
ating intense pressure and temperature gradients locally. In the late 1920s, Lord
Rayleigh [9] published the first mathematical model describing ‘cavitation’ in
incompressible fluids. It is not until 1927 that the use of ultrasound on chemical
and biological systems was first observed and recognised as a useful tool by Richard
and Loomis [10].
As an ultrasonic wave passes through a liquid; fluctuating pressures are rapidly
set up as a result of the alternate periods of compression and rarefaction associated
with the wave [6–10]. During the compression cycle, the liquid is subjected to a
sufficiently positive pressure which pushes the molecules of the liquid together,
whilst during the following rarefaction cycle, the liquid is subjected to an equal but
negative pressure which pulls the molecules of the liquid away from each other.
Increasing the amplitude of the wave leads to an increase in the magnitude of the
positive and negative pressures. If the liquid is subjected to a sufficiently large
negative pressure during the rarefaction cycle, the molecules are torn away from
each other producing very small cavities, called ‘microbubbles’. In other words
these negatives pressures are strong enough to overcome the intermolecular forces
binding the liquid. The process of tearing the liquid apart is known as ‘cavitation’
and the microbubbles are called ‘cavitation bubbles’ (Fig. 10.5) [6–10]. The cav-
itation threshold is the limit of sound intensity below which cavitation does not
occur in a liquid.
Cavitation phenomenon is well known to cause erosion, emulsification, molec-
ular degradation, sonoluminescence and sonochemical enhancements of reactivity
solely attributed to the collapse of cavitation bubbles [6–10]. It is now well
accepted in the field that the cavitation bubble collapse leads to near adiabatic
heating of the vapour that is inside the bubble, creating the so-called “hot-spot” in
the fluid, where:
(1) High temperatures (ca. 5000 K) and high pressures (ca. 2000 atm/200 MPa) are
generated with a collision density of 1.5 kg cm2 and pressure gradients of
2 TPa cm1, with lifetimes shorter than 0.1 μs and cooling rates above 109–10
K s1 during the collapsing of cavitation bubbles are observed. Here, water
vapour is ‘pyrolyzed’ into hydrogen radicals (H•) and hydroxyl radicals (OH•),
10 Sonoelectrochemical Production of Fuel Cell Nanomaterials 415

Fig. 10.5 Representation of a cavitation bubble imploding

Fig. 10.6 Schematic of water sonolysis

known as water sonolysis (Fig. 10.6). Note that the exact temperatures and
pressures generated during cavity implosion are difficult both to calculate
theoretically and to determine experimentally.
(2) The interfacial region between the cavitation bubbles and the bulk solution is
paramount. The temperature is lower in the interior of the bubbles than the
exterior but high enough for thermal decomposition of the solutes to take place
with greater local hydroxyl radical concentrations in this region.
(3) The reactions of solute molecules with hydrogen atoms and hydroxyl radicals
occur in the bulk solution at ambient temperature.
416 B.G. Pollet and P.M. Sakkas

Because of ultrasound’s ‘extraordinary’ effects, extensive work has been carried


out in which high power ultrasound (20 kHz–2 MHz) was applied to various
chemical processes leading to several industrial applications and many publications
over a wide range of subject areas [6, 7]. It has been shown that the effects of high
intensity ultrasonic irradiation on chemical processes lead to both chemical
and physical effects, for example, mass-transport enhancement, surface cleaning
and radical formation (HO•, HO2• and H•) due to homolytic cleavage via sonolysis
due to cavitation phenomena (Fig. 10.6) [6, 7].
Although, the use of ultrasound in chemistry went through a period of neglect
until the 1980s when laboratory equipment in the form of cleaning baths and
biological cell disrupters became more available. Consequently, interest has been
revived and nowadays ultrasound is applied to a wide range of subject areas within
chemistry such as nanochemistry, analytical chemistry, organometallic chemistry,
organic chemistry, inorganic chemistry, electrochemistry, material science, envi-
ronmental chemistry and polymer science [6, 7].
Cavitation Bubble Formation. There are two kinds of cavitation—transient
and stable cavitation [6, 7, 11]. Transient cavitation bubbles are those which exist
for one, or at most a few, acoustic cycles expanding to a radius of at least double
their initial size before collapsing violently into several smaller bubbles. These tiny
bubbles may act as nuclei for other bubbles. Transient bubbles occur mainly in
liquids subjected to ultrasound intensities which are greater than 10 W cm2. It was
thought initially that the effects such as erosion, emulsification, molecular degra-
dation, sonoluminescence and sonochemical enhancement of reactivity were solely
attributable to the collapse of transient cavities.
Stable cavities are those which contain gas and vapour and are known to be
generated at low ultrasonic intensities (1–2 W cm2). These cavities oscillate
non-linearly about an equilibrium size over a number of acoustic cycles. Stable
cavities are capable of being transformed into transient cavities and are known now
to be responsible for numerous chemical effects.

10.3.1 Transient Cavitation Bubble

As the name implies, transient cavitation bubbles have a short life-time (ca. 105 s
in a 20 kHz ultrasonic field) before they collapse violently on compression and
disintegrate into smaller bubbles. These smaller bubbles may act as nuclei for
further bubbles, or if they are of sufficiently small radius, they can dissolve into
the bulk of the solution. During the short lifetime of the transient bubble, it is
assumed that little or no diffusion of dissolved gas can take place from the solution
bulk into the cavity, whereas condensation and evaporation of liquid is assumed to
take place freely. Since there is no gas present in the bubble to act as a cushion, the
implosion leads to a very violent collapse.
It was assumed by many authors, such as Noltingk and Neppiras [12], that
adiabatic collapse of the bubbles would allow for a calculation of the temperature
10 Sonoelectrochemical Production of Fuel Cell Nanomaterials 417

or pressure within the bubble (Eq. 10.1 and 10.2). For example, if it is assumed that
the vibrations of the bubbles occur so very fast that little heat exchange can occur
with the surrounding liquid environment, then the vapour inside the bubble is
heated during the compression cycle and one may deduce, under these adiabatic
conditions, that the maximum pressure, Pmax and temperature, Tmax, can be given as
follows:
 γ
Pm ðγ  1Þ γ1
Pmax ¼ P ð10:1Þ
P
 
ðγ  1 Þ
T max ¼ T o Pm ð10:2Þ
P

where To is the ambient temperature,


γ is the ratio of the specific heat capacities of the gas (or gas vapour) mixture,
P is the pressure in the bubble at its maximum size (usually assumed to be equal to
the vapour pressure, Pv, of the liquid), and
Pm is the pressure in the liquid at the moment of transient collapse.
At ambient temperature and pressure, Noltingk and Neppiras [12] were able to
deduce that extremely high temperatures (Tmax) and pressures (Pmax) were pro-
duced in the final phase of implosion equivalent to 5000 K and 2000 atm. The
release of the pressure as a shock wave is a factor which has been used to explain
increased chemical reactivity, due to increased molecular collision, and polymer
degradation; the high temperatures within the bubble have formed the basis for the
explanation of radical production and sonoluminescence [6, 7, 11].

10.3.2 Stable Cavitation Bubble

A stable bubble is one of which is thought to exist for many cycles. Stable bubbles
contain mainly gas and vapour. They are produced at low intensities (1–3 W cm2)
and oscillate about an equilibrium size for several acoustic cycles [6, 7, 11]. The
time-scale over which they exist is long enough for both mass transfer and thermal
diffusion to occur freely. This process occurs as follows:
In the rarefaction phase of the sound wave gas diffuses from the liquid into the
bubble causing the bubble to expand, whilst in the compression phase gas diffuses
out of the bubble into the liquid. The increased surface area increases the rate of gas
and vapour diffusion into the bubble. As a succeeding compression wave passes
through the liquid, the bubble is compressed. Therefore, gas and vapour diffuse out.
As a result, the rate of inward diffusion will become greater than the rate of
diffusion back to the liquid and this will lead to an overall growth of the bubble.
As the bubble grows, the bubble will become more compressible due to changes in
the acoustical and environmental conditions of the medium.
418 B.G. Pollet and P.M. Sakkas

The stable bubble may be transformed to a transient bubble and undergo


collapse, the violence of the collapse being less than that of a vapour-filled transient
bubble since the gas present cushions the implosion. There are two possible fates
for stable cavitation bubbles. They may either grow sufficiently large to be capable
of rising to the surface of the liquid—this is the process of ultrasonic degassing—
they can become unstable due to differences between the resonant frequency and
the driving frequency (i.e. the frequency from the transducer) and be transformed
into transient bubbles.
It was also found that the maximum temperatures and pressures of these cavi-
tation bubbles generated upon collapse are less than for transient cavitation bubbles
due to the ‘cushioning’ effect [6, 7, 11]. However, it was shown that contracted
bubbles cause temperature rise within the bubble and changes in local hydrostatic
pressure [6, 7, 11]. Calculations of the local pressures due to these resonance
vibrations have resulted in values which exceed the hydrostatic pressure by a factor
of 150,000. There is no doubt that the intense local strains in the vicinity of the
resonating bubble are the cause of the many disruptive mechanical effects of
ultrasound.
In summary, cavitation bubble formation is a three step process consisting of
(1) nucleation, (2) bubble growth and (3) collapse of gas vapour filled bubbles in a
liquid phase. These bubbles transform the low energy density of a sound field into a
high energy density by absorbing energy from the sound waves over one or several
cycles and releasing it during very short intervals. During rarefaction cycles,
negative pressures developed by the high power ultrasound are strong enough to
overcome the intermolecular forces binding the fluid. The succeeding compression
cycles can cause the microbubles to collapse almost instantaneously with the
release of a large amount of energy. When the cavitation bubble collapses close
to or on a solid surface it can only do so asymmetrically due to the surface which
leads to a microjet of liquid being directed towards the surface of the material at
speeds of up to 200 m s1 [6, 11, 13].

10.4 Sonoelectrochemistry: The Use of Ultrasound


in Electrochemistry

The use of ultrasound on electrochemical systems or Sonoelectrochemistry was first


observed by Moriguchi as early as 1934 [14] and since then, continues to be an
active and exciting research area. In the 1930s, Schmid and Ehert [15] also studied
the ultrasonic effects on the passivity of metals and the generation of gases by
electrolysis [16]. Nearly 30 years later, in 1956, Nyborg et al. [17] demonstrated
that the movement of liquid induced by ultrasound, known as acoustic streaming
and in the same year Penn et al. [18] studied the effect of ultrasound on concentra-
tion gradient in the electrolyte and at the electrode surface. In 1965, for the first
time, Bard [19] showed that ultrasound caused an increase in mass transport of
10 Sonoelectrochemical Production of Fuel Cell Nanomaterials 419

electroactive species from the bulk solution to the electrode surface in controlled
potential coulometry experiments. Since then, extensive work has been carried out
in which high power ultrasound (20–100 kHz) was applied to various electrochem-
ical processes leading to several industrial applications and many publications over
a wide range of subject areas such as electrodeposition, electroplating, electro-
chemical dissolution, corrosion testing and nanotechnology.
For over 80 years, nearly a thousand papers have been written on the subject with
many original work, general reviews [6] in sonoelectrochemistry (with the first one
from Mason et al. [20] in 1990, including the effects of power ultrasound on special
media), organic sonoelectrosynthesis, sonoelectroanalysis, sonoelectrochemical
production of nanomaterials, and recently the sonoelectrochemical production of
fuel cell materials [6].
In all these papers and reviews, it was clearly shown that the effects of high
intensity ultrasonic irradiation on electrochemical processes lead to both chemical
and physical effects, for example, mass-transport enhancement, surface cleaning
and radical formation. Many workers have also investigated the distribution of
ultrasonic waves or energy in various electrochemical reactors operating in the
lower ultrasonic frequency range (20–100 kHz) and at high ultrasonic powers.
Several methods for such determination have been proposed e.g. aluminium foil
erosion, sonoluminescence, calorimetric methods, chemical dosimetry [7] and
laser-sheet visualization [6].
In sonoelectrochemistry, ultrasound is known for its capacity to promote espe-
cially heterogeneous reactions mainly through intense mass-transport, interfacial
cleaning and thermal effects. In addition, homogeneous chemical reactions have
been reported to be affected and the generation of highly radical species [6, 7] [e.g.
the production of H• and OH• radicals by sonolysis in intense sound fields is an
important aspect for the use of ultrasound. For example, in the detoxification of
environmentally harmful wastes [containing heavy metals, hydrocarbons (PAH’s)
and chlorinated compounds (PCB’s)], ultrasound has been found to be extremely
beneficial.
The vast variety of ultrasonically induced effects observed in sonoelectro-
chemistry may be ascribed to the generation, pulsation and collapse of cavitation
bubbles in the electrolyte medium near the electrode surface. This ultrasonic
cavitation occurs at low to high ultrasonic intensities (ultrasonic power per tip or
transducer or cell base area). A pulsating cavitation bubble close to the electrode
surface generates microstreaming. When the cavitation bubble reaches a resonant
size, it collapses asymmetrically leading to the formation of high velocity jet of
liquid toward the surface. This phenomenon leads to a thinning of the diffusion
layer and can improve the overall mass-transfer and hence reaction rates [6, 7].
Since most of the observed effects of ultrasound in electrochemical processes are
thought to be due to the cavitation effect together with micro-streaming, the
application of ultrasound is known to be very beneficial in the electrochemical
industry. This has led to investigations into mass-transport, electron-transfer pro-
cesses and electrode surface adsorption [6, 7]. The reader is invited to consult the
following book ‘Power Ultrasound in Electrochemistry: From Versatile Laboratory
420 B.G. Pollet and P.M. Sakkas

Tool to Engineering Solution’ by BG Pollet (ed) [6]. This book deals with the aspect
of electrochemistry combined with ultrasound and explains the various electro-
chemical phenomena occurring at the electrode surface when a potential is applied
across it. For this purpose, electrode kinetic and mass-transport parameters are
defined. The book also outlines the theory, principles and applications of sonoelec-
trochemistry in various branches of chemistry. Finally it is shown how coupling
ultrasound with electrochemistry could be used to improve electrochemical pro-
cesses, enhance detection limits in the electroanalysis of toxic samples and produce
nanomaterials.
In 1998, Pollet et al. [6, 13] showed, with the aid of mathematical models based
on mass-balance equations and using the quasi-reversible redox couple Fe(CN)63/
Fe(CN)64 as an electrochemical model, that a Levich-like equation relating the
limiting current, the inverse square root of the electrode radius and the inverse
square root of the electrode-horn distance, the transmitted ultrasonic intensity (and
thus the transmitted ultrasonic power) may be generated for ultrasonic frequencies
of 20 and 40 kHz (probe systems only and at 298 K) using Eq. (10.3) also known as
the Pollet equation [6, 13]:

I lim ¼ 0:84nFADo 2=3 n1=6 r e 1=2 Auht 1=2 d 1=2 C*PT 1=2 ð10:3Þ

where Ilim is the limiting current (A), n is the number of electrons transferred during
the electrochemical process, F is the Faraday constant (C mol1), A is the electrode
area (cm2), Do is the diffusion coefficient (cm2 s1) of the electroactive species, d is
the ultrasonic horn-electrode distance (cm), ν is the kinematic viscosity (cm2 s1),
re is the working electrode radius (cm), Auht is the ultrasonic horn tip area (cm2), C*
is the bulk concentration of the electroactive species (mol cm3) and PT is the
ultrasonic power transmitted (W).
From Eq. 10.3, experimental limiting current values and assuming that all
the above parameters are known, the transmitted ultrasonic power (PT) can be
calculated.

10.5 Sonoelectrochemical Production of Noble Metals


and Fuel Cell Electrocatalysts

The sonoelectrochemical method is the use of power ultrasound applied to various


electrochemical processes leading to both chemical and physical effects, such as
increase in mass-transport, improved surface cleaning, increase in deposit hardness,
deposit thickness, deposition rates and efficiencies as well as improved deposit
adhesion and radical formation [6].
There are a few studies reporting the use of the sonoelectrochemical method to
produce noble metals and fuel cell electrocatalysts. Here, the sonoelectrochemical
technique involves either depositing a metal under continuous electrical current and
ultrasound or producing nanosize metals at various currents and ultrasonic pulses
(a few ms) at a vibrating electrode (Fig. 10.7).
10 Sonoelectrochemical Production of Fuel Cell Nanomaterials 421

Fig. 10.7 Sonoelectrochemical experimental set up

10.5.1 For PEM Fuel Cells

Shen et al. [21] produced uniform spherical three-dimensional dendritic Pt


nanostructures (DPNs) with an average dimension of 2.5  0.5 nm, at room tem-
perature sonoelectrochemically (20 kHz; 20 W; pulse width of the current ¼1.0 s,
resting time of the current pulse ¼ 0.5 s; duration of the ultrasonic pulse ¼ 0.3 s) by
using mixtures of hexachloroplatinum(IV) acid hydrate in various surfactants [PVP,
poly(ethylene glycol)20-poly(propylene glycol)20-poly(ethyleneglycol)20 (P123),
SDS and poly-diallyl-dimethyl ammonium chloride (PDDA)] at current densities
up to 40 mA cm2. They showed that at low current densities the reduction rate of
Pt is slower, the number of nuclei is small and the rate of growth is faster than that
of nucleation, which lead to large nanoparticles (5–7 nm). As the current density
increases the reduction rate of Pt4þ increases and the nucleation rate is faster than
that of growth leading to the generation of more nuclei and the formation of smaller
primary Pt nanoparticles. They also showed that stabilized Pt nanoparticles do not
aggregate in PVP due to stronger bonds between the Pt precursors with the C¼O
groups of the PVP in other words PVP adheres to the nanoparticles through a
charge-transfer interaction between the pyrrolidone rings and Pt atoms. The DPNs
showed improved electrocatalytic activity towards MOR due to monodisperse Pt
nanoparticles, and improved porosity structure leading to large effective surface
422 B.G. Pollet and P.M. Sakkas

area. They proposed a mechanism whereby Pt ions are reduced by the electrical
current and formed Pt primary nanoparticles which are subsequently dislodged by
the vibrating electrode. The primary nanoparticles then in solution spontaneously
assemble together and formed small spherical DPNs. They showed that ultrasound
leads to small primary nanoparticles in favour of the crystallite reorganization and
growth of a stable near-single crystal.
Zin et al. [22] produced platinum nanoparticles from aqueous chloroplatinic
solutions by the pulse sonoelectrochemical (20 kHz, up to 118 W cm2) method on
titanium alloy electrodes in the absence of any surfactants, alcohols and polymers
by producing short applied current pulses triggered and followed immediately by
ultrasonic pulses at the working electrode (in this case the cathode). The time
management sequence employed was as follows [Fig. 10.7]:
1. A short current pulse of |i| ¼ 50 mA cm2 was applied to the sonoelectrode,
whereby the titanium horn acted as an electrode only (tON); the time of this phase
typically varied between 0.3 and 0.5 s.
2. Immediately after the electrochemical pulse was turned off, an ultrasonic pulse
was sent to the sonoelectrode and here it acted only as a vibrating ultrasonic horn
(tUS); this second phase lasted no more than 0.5 s.
3. A rest time, tp, followed the two previous phases (this was useful to restore the
initial electrolyte conditions close to the sonoelectrode).
They showed that Pt mean grain size ranging from 11 to 15 nm was produced
and globular clusters had a mean size ranging between 100 and 200 nm which in
turn aggregated and built complex structures.
Shen et al. [23] showed that by using the sonoelectrochemical (20 kHz; 20 W;
pulse on time of the current ¼ 0.5 s; pulse off time of the current ¼ 0.5 s, duration of
the ultrasonic pulse ¼ 0.3 s) method, it is possible to realize the morphology-
controlled synthesis of palladium nanostructures [spherical (SNP), multitwinned
(MTP) and spongelike (SSP)] at room temperature in the presence of various
surfactants and polymers [cethyltrimethylammoniumbromide (CTAB), PVP and
PDDA]. They showed that the size and shape of the Pd nanostructures may be
controlled by varying the current density and the precursor solution pH value.
Furthermore, the electrocatalytic activities of the produced spongelike Pd
nanostructures for direct alcohol oxidation in alkaline media showed higher elec-
trochemical active surface than other Pd nanostructures. Qui et al. [24] also
synthesized highly dispersed spherical Pd nanoparticles of a dendritic superstruc-
ture in the presence of CTAB by the pulse sonoelectrochemical (20 kHz) method at
room temperature and a reaction time above 2.5 h. They explained that the
dendritic-structured Pd nanoparticles had a treelike structure and agreed with the
diffusion-limited aggregation (DLA) model, involving cluster formation by the
adhesion of particles to a selected seed on contact and allowing the particle to
diffuse and stick to the growing structure. They stipulated that it is possible that Pd
particles reach the anode and grow into a dendritic structure. They concluded that
the shape and size of spherical nanocrystalline Pd may be controlled by varying the
current density (8–13 mA cm2), the interval between continuous ultrasonic pulses,
10 Sonoelectrochemical Production of Fuel Cell Nanomaterials 423

Table 10.1 A summary of the sonoelectrochemical production of noble mono-metallics at


various ultrasonic frequencies and powers in several surfactants
Ultrasonic
Noble frequency Ultrasonic Particle size
metals (kHz) power Surfactant Solvent (nm) Authors
Pt 20 20 W PVP, – 2.5  0.5 Shen
P123, et al. [21]
SDS,
PDDA
Pt 20 Up to – – <200 Zin
118 W cm2 et al. [22]
Pd 20 – PVP, – <200 Shen
PDDA, et al. [23]
CTAB
Pd 20 20–120 W cm2 CTAB – 5–10 (depending Qui
on current densi- et al. [24]
ties and CTAB
concentrations)

ultrasonic intensity (20–120 W cm2) and the CTAB concentration. For example,
they observed that the shape of the nanoparticles appeared irregular and agglom-
erated below 20 and 120 W cm2.
For all the sonoelectrochemical production of nano-size metals, a mechanism
has been proposed where metallic ions are reduced by a short current pulse to
produce metallic nanoparticles on the sonoelectrode surface, which are then
dislodged by the ultrasonic pulse. The metallic nanoparticles in solution tend to
spontaneously assemble together whereby under insonation, the Ostwald ripening
process is accelerated, leading to smaller primary nanoparticles [6].
Table 10.1 shows a summary of the sonoelectrochemical production of noble
mono- metallics at various ultrasonic frequencies and powers in several surfactants.

10.5.2 For Solid Oxide Fuel Cells (SOFC)

10.5.2.1 Basic Components

SOFC technology has the most remarkable property of running using a wide variety
of gas fuels, from pure H2, CO and H2/CO rich reformate gases, to gaseous or
gasified liquid hydrocarbons such as natural gas [25]. The electrode exposed to fuel
is anode and plays a multi-functional role. On the side where anode is attached to
solid electrolyte (typically yttria-stabilized zirconia, YSZ) the former receives
oxygen ions delivered by the later at temperatures beyond 600  C. On the side
where anode is exposed to fuel gaseous phase it receives reactants (typically H2)
and yields products (typically H2O). Finally the solid electrode phase takes up
electrons, establishing one of the electric current flow poles. This is the concept of
424 B.G. Pollet and P.M. Sakkas

three-phase boundary (TPB), where transport pathways of ions, gaseous chemical


species and electrons intercept. Electrochemistry can only take place in locations
where all three components are spatially close to each other and thus be enhanced
when a high volumetric TPB length is achieved [26].
Such a composite material, consisting of ceramic and metal species, is referred
to as cermet electrode. The latter has to retain a porous structure with the metal
particles well dispersed in its bulk, so as to deliver high electrochemical perfor-
mance, i.e. let O2 oxidize the fuel that resides at the porous anode side. So far,
Ni/YSZ fulfils most of the requirements for an SOFC anode. This composite acts as
an electronic conductor, due to Ni, as well as an ionic conductor, due to YSZ
[27]. An alternative to YSZ ionic conductor is gadolinium-doped ceria, GDC, as it
exhibits high activity for hydrocarbon oxidation and high ionic conductivity at the
same time [28]. Since the first attempts to use ceria for the SOFC anodes in the ‘60s,
introduction of CeO2-δ based additives, e.g. Ni/GDC, is widely considered among
the most promising directions in anodes development [29].
As far as the electrons conductive part of an anode is concerned, Pt electrodes
were originally used due to their stability, high conductivity, and a substantial
catalytic activity, while later it was substituted with less expensive components,
primarily Ni-based cermets for the anodes [30]. Although Ni has been the standard
primary metallic component in anode cermets for decades, Ni/YSZ is thermally
unstable and results easily to Ni agglomeration and island formation. Nevertheless,
Ni/YSZ systems demonstrate a tendency towards accumulation of impurities and
formation of secondary phases at two- and three-phase boundaries [26]. As a result,
continuous efforts are being made since the 70’s in order to develop optimized
anode formulations, particularly via the use of transition metal components [29].

10.5.2.2 Importance of Nanomaterials Addition in SOFC

One of the most complex parts of an SOFC is the anode, as it has to (a) let as many
O2 as possible migrate from the electrolyte towards its porous surface, (b) fulfill
fuel oxidation at intermediate (typically 700–850  C) or high temperatures (beyond
900  C) and (c) let current be conducted through its bulk and be collected at its
surface. At the same time, it should ideally (d) remain porous for as long as possible
and (e) not experience any considerable sintering of its catalysis aiding metallic
components. To date, the superior catalytic activity of Ni for electrochemical
oxidation of hydrocarbon fuels renders cermets containing Ni/YSZ or Ni/GDC as
the preferred anode materials. As in all catalytic applications, the smaller and the
more dispersed these components are, the better the fuel utilization is expected
to be.
Moreover, several combinations of metal components are utilized on the anode
side particularly for the case of direct non-pure hydrogen oxidation [31, 32]. A good
selection of materials combinations may be the key for achieving the best perfor-
mance characteristics for a SOFC. For instance, it has been shown that part of the
anode overpotential is enhancing the Ni particles sintering and that this process can
10 Sonoelectrochemical Production of Fuel Cell Nanomaterials 425

be significantly delayed by increasing the work of adhesion between Ni and YSZ


particles. The latter can be achieved by trace amounts of additives, such as CeO2 or
TiO2 at the Ni/YSZ interface [33]. Other studies have also shown that composites of
Ni/CeO2 have additional significance because of the possible application in high
temperature fuel cells [34]. Moreover, the important problem of SOFC cermet
anodes degradation has been studied via a theoretical model from a fundamental
viewpoint. The model describes the gradual degradation of the anode due to nickel
particle sintering and the concomitant loss of TPB length [35].
Among recent developments many studies concern the direct feeding of SOFCs
with hydrocarbons from fossil or renewable sources. Even though the above
strategy appears more attractive compared to the state-of-the-art technology of
using hydrogen as fuel, there still remain some major challenges to be addressed
for extended fuel application in SOFCs. One of them is anode poisoning caused by
unfavorable reactions of catalytic anode materials with C and/or S species, which
are present in readily available hydrocarbon fuels such as natural gas, biogas or coal
syngas [36]. In recent years there have been several reviews that tackle the
development challenges of SOFC anode materials [37–40].
Carbon deposition in SOFCs may originate from the internal steam or dry
reforming of hydrocarbons or biofuels. Many factors that influence the likelihood
of such formation have been identified. The catalyst nature is one of the most
important [41]. For example, carbon formation is more likely over Ni than, for
instance, over Au, Ag, Cu, Ru or LSCM [27, 41–43]. In a more recent study Sn/Ni
surface alloy has been identified as much more carbon tolerant than monometallic
Ni under conditions of steam reforming of methane, propane and isooctane at
moderate steam to carbon ratios [44]. As a result, a state of the art Ni-based cermet
enhanced with a second or more metal compounds could operate under steam or dry
reforming conditions with high carbon tolerance in the intermediate temperature
range.
Sulfur poisoning of SOFC systems operating at intermediate temperatures can be
irreversible while at high temperatures in tends to be reversible [45, 46]. Inhibition
of sulfur adsorption leading to sulfides formation has been reported in reviews that
investigate different anodes and conditions [36, 47]. Several studies prove that by
decorating Ni surface with elements such as Mo, Re, W, Pt, Ag or their oxycarbides,
can modify the surface electronic properties [48–53].

10.5.2.3 Preparation of SOFC Materials

Sonoelectrochemical synthetic route has an extended nanoparticles preparation


showcase of many metallic, metal oxides or alloy species. All these compounds
can be used for cermets preparation either by (a) a step-by-step approach where the
sole nanoparticles synthesis is followed by impregnation or decoration on the fuel
cells ceramic materials or (b) a synchronous approach where the nanoparticles are
co-deposited on the working electrode along with the ceramic substrates.
426 B.G. Pollet and P.M. Sakkas

Sole Synthesis of Metal Species

Numerous research groups worldwide have extensively investigated metal, metal


oxides or metallic alloys sonoelectrochemical synthesis. As a result, there have
been published some very comprehensive reviews that compile and compare
several methods [54–57]. From all species reviewed Ag, Ni, Cu, Pt and Au have
been multiple times studied.
A typical method to categorize the various synthesis routes is according to the
position of the ultrasound emitter according to the working electrode. Three types
are reported: (a) the conventional face-on where the sonohorn is opposite to the
electrode, (b) the side-on where the electrode is placed perpendicular to the
ultrasonic waves, and finally (c) the sonotrode where the actual ultrasonic horn is
at the same time the electrode [58]. Use of the sonotrode is suited for nanoparticles
synthesis that will be dispersed in the electrochemical bath usually with the help of
surfactandts also dissolved in the bath. Furthermore, the sonoelectrochemical
technique may follow either (a) the typical sequence of electrochemical
(potentio-/galvano-static) and ultrasound pulses or (b) electrochemical pulses
under continuous sonication.
Nickel is the typical element used in SOFC anodes because it enhances catalytic
activity and offers good electrical conductivity being a much more economic
alternative to Pt. Originally, Delplancke et al. prepared crystalline Ni powders
with a sharp diameter distribution around 100 nm combining pulsed current and
pulsed ultrasound on the same electrode surface [59]. Later on, Davis et al. studied
the advantageous effects that the acoustic streaming and cavitation brings in Ni
sonoelectroanalysis [60]. Jia et al. synthesized for the first time first time
nanoporous Ni particles in dimethyl sulphoxide (DMSO) [61]. It was found that
these nanopores were produced by the aggregation and/or melting of primary
nanoparticles of about 4–5 nm.
Gold nanoparticles are of the largest interest for SOFC anode modification after
Ni. Liu et al. synthesized Au nanoparticles from 2 to 15 nm under cathodic
overpotentials and ultrasonication [62]. These experiments revealed that adjusting
the sonoelectrochemical reduction time could control the ratio of Au nanoparticles
to Au-containing nanocomplexes in solutions. On the other hand, Aqil et al. made a
study of Au nanoparticles synthesized under potentiostatic conditions and by
experimenting with several stabilizing agents [63]. They claimed that the mean
size could be easily tuned between 5 and 35 nm by simple control of the electro-
chemical parameters. Shen et al. managed to prepare Au nanoparticles with no use
of stabilizers. Size and shape was controlled by adjusting current density, reaction
time and pH of bath. Particles sizes of 10–50 nm were synthesized with an obvious
tendency to form agglomerates [64]. Sakkas et al. conducted a series of experiments
for Au nanoparticles synthesis at various potentiostatic pulses and continuous
sonication by using PVP as surfactant [65]. The average size distributions ranged
from 10 to 35 nm according to the electrochemical pulse. No crystallinity was
evident under HR-TEM study. This confirms previous studies, which proved that
10 Sonoelectrochemical Production of Fuel Cell Nanomaterials 427

sonoelectrodeposition under continuous ultrasonication leads to the formation of


amorphous materials [66].
Copper or its oxides have been investigated as an alternative to the more
expensive Au or Ag. Haas et al. used galvanostatic conditions and PVP as stabilizer.
They experimented with current duration, experiment duration, PVP concentration,
temperature, and ultrasonic power to obtain particles ranging from 10 to 55 nm. A
suggested possible bonding framework for the interaction between metallic Cu and
PVP is also presented [67]. On the other hand, Sáez et al. present a study of Cu
electrocrystallization for obtaining a narrow size distribution. They used both
galvanostatic and potentiostatic conditions that yielded mean particles sizes of
200 nm and 81 nm respectively. In some of the experiments PVP was also used
[68]. Furthermore, Sakkas et al. prepared Cu nanoparticles using galvanostatic
conditions in presence of PVP. Their study yielded that for brief electrodeposition
pulses the mean particle size distribution is 8 nm, smaller than for longer pulses
where they obtained 50 nm Cu nanoparticles [65]. Finally, Schneider et al. conducted,
among others, a series of synthesis experiments at room temperature using a copper
sulfate based electrolyte and following the face-on route [69]. They grew a fine Cu
film on a typical electronic quartz crystal microbalance (EQCM) using continuous
ultrasonic irradiation at different intensities. Their attempts revealed that the surface
of Cu deposition gets significantly coarser compared to silent conditions, which in
turn leads to higher area of the film grown [69].
Silver has also been investigated by many research schemes. Zhu et al. worked
with a typical Ag electrochemical bath adding nitrilotriacetate (NTA). They
reported that varying the electrosonic time, the Ag precursor and NTA concentra-
tion affects the nanoparticles shape [70]. Liu et al. from the same group attempted
Ag nanoparticles synthesis from a different electrochemical solution and this time
in the presence of gelatin. They prepared the electrochemical bath following a quite
intricate route and keeping as low light exposure as possible. The reduction was
done with galvanostatic conditions. The powder that yielded was amorphous and
the size distribution from 12.2 nm to more than 60 nm [71]. Liu et al. followed a
potentiostatic approach of Ag nanoparticles synthesis, and by increasing the
cathodic overpotential they managed to control the size distribution from as low
as 2–20 nm [72]. Jiang et al. followed galvanostatic conditions with the presence of
PVP as surfactant. They were able to achieve Ag nanoparticles of 5–35 nm as well
as dendritic formations by varying the current density or the PVP presence
[73]. Finally, Vu et al. managed to synthesize not only spherical but also rod-like
Ag nanoparticles in presence of SDS as surfactant. In their studies they used
galvanostatic conditions and a stainless steel foil as anode. Ultrasonic horn was
immersed in the bath in the space between the opposite facing electrodes. By
varying the precursor as well as surfactant concentration they ended up with fine
particles of 5–14 nm or rods of 100–500 nm [74].
Platinum has been so far synthesized in two different nanoformations,
i.e. dendritic Pt nanostructures (DPNs) and monodisperse. Shen et al. managed to
synthesize DNPs that exhibited remarkable porosity structure and enhanced effec-
tive surface area, thus adding to the electrocatalytic activity of this compound
428 B.G. Pollet and P.M. Sakkas

[21]. On the other hand, Zin et al. prepared Pt nanoagregates with the Pt
nanoparticles ranging from 11 to 15 nm. Their approach involved galvanostatic
synthesis with use of no surfactant [22].
Tungsten nanoparticles has been rarely synthesized via sonoelectrochemistry. In
their work, Lei et al. managed to prepare body center cubic W nanoparticles
galvanostatically by varying the current density as well as the ultrasound pulse
period and intensity. Spherical nanoparticles of about 30 nm yielded [75].

Synchronous Deposition of Ceramic and Metal Species

A combined action of electrochemical metal deposition and ultrasonication for


SOFC cermet composites preparation was utilized for the first time by Argirusis
et al. [76]. In their experiments they used state of the art doped oxide nanoparticles
(GDC) suspended in a typical metal precursor-bearing electrolyte for Ni/GDC
synthesis. They worked with a Ni sulfamate electrolyte at 50  C and 5 g/L of
GDC in suspension. GDC crystallite size was 5–10 nm while the pre ultrasonication
agglomerates reached up to 149 μm. Ultrasound was used prior and during electro-
deposition, continuously, for achieving deaglomeration and full dispersion respec-
tively. An ultrasonic horn tip of 13 mm diameter was placed at a distance of 22 mm
face on [58] to the Au working electrode and operated at 20 kHz and 25 W cm2.
With the application of ultrasound, currents increased and became noisier com-
pared to sole Ni depositions conducted as reference. Co-deposited Ce amount was
reported to have reached 2.5 wt% according to SEM/EDX data.
On the other hand, Ni/CeO2 cermets were synthesized by Xue et al. [77]. They
conducted, among others, a study on Ni electrochemical pulsed current deposition
and synchronous CeO2 nanoparticles deposition under continuous ultrasonication.
In their experiments they used a typical sulfamate electrolyte at 45  C and 40 g/L of
suspended CeO2 with particles average size estimated at 40 nm. The two electrodes
were placed vertically and the bath was immersed in an ultrasonic bath with the aid
of magnetic stirring at 1000 rpm. Current density was kept at 4 A dm2 in all
samples. SEM analysis revealed that grain size of cermets prepare via this method is
reduced significantly, with the CeO2 nanoparticles uniformly distributed.
An attempt of Cu/CeO2 nanocomposite cermet was reported by Lee
et al. [78]. They used a mixture of copper(II) sulfate pentahydrate with sulfuric
acid as electrolyte at 50  C and 20 g/L ceria with approximate mean particles
diameter of 20 nm. Anode and cathode were dipped in bath in parallel with a
distance of 20 mm. A 3 mm diameter ultrasonic tip was also immersed in the
suspension at a position where it was out of the inter-electrode area and used
synchronously to the electrodeposition. At an ultrasonic intensity of 120 W cm2
and diluted electrolyte concentration, a higher volume fraction of CeO2
snanoparticles in Cu film of about 20.5 vol.% was observed compared to the silent
deposition.
10 Sonoelectrochemical Production of Fuel Cell Nanomaterials 429

Decoration of Ceramic Species with Nanoparticles

Most of the times, nanoparticles prepared via the sonoelectrochemical route remain
in a colloidal solution state that is prone to slow or rapid agglomeration depending
on whether a suitable surfactant was used. In order to take advantage of the
synthesized species minuscule size and a good dispersion of them on the ceramic
powders one has to proceed to their decoration on ceramic substrate material. This
has been achieved in the past via impregnation. According to it, substrate powders
are dispersed in the colloidal solution and accordingly agitated until successful
residing of nanoparticles in nanopores or cavities of the substrate powders.
However the use of ultrasonication has demonstrated remarkable results in deco-
ration. Gedanken has reported a study of decoration on several substrates via
ultrasonication [79]. Especially in case of Au nanoparticles decoration on state-
of-the-art GDC powders, Sakkas et al. have used direct high power ultrasonication
with successful results [65, 80].

10.5.3 For Other Fuel Cells

Molten carbonate fuel cells experience cathode material dissolution under extreme
conditions. Nickel is used widely as cathode material. Addition of LiCoO2 on the
surface of the cathode is known to reduce the precipitation of Ni in the bath [81]. It
is also reported that CeO2 layers and CoO/CeO2 layers reduce NiO dissolution, and
catalyze the Li incorporation into the NiO cathode [82]. As a result, there is strong
interest in Co incorporation in CeO2 composites for MCFCs cathodes protection.
Sonoelectrochemical synthetic route has been used many times for preparation of
Co, Ni or their alloys. Argirusis et al. achieved a co-deposition of Co and GDC with
CeO2 content of up to 8 wt% [76]. Nevertheless, Reisse et al. reported the formation
of a new alloy that consists of 75 %Co-25 %Ni [54]. Delplancke et al. have also
reported Co nanoparticles synthesis on average size distribution of 100 nm
[59]. Finally, Dabala et al. reported the synthesis of Co-Fe alloys [83].

10.6 Conclusions

The sonoelectrochemical methods for the preparation of efficient mono- and bi-
metallic nanoparticle electrocatalysts have an advantage over many other methods
due to the unusual experimental conditions caused by cavitation, water sonolysis
and enhanced mass-transport phenomenon. These methods are expected to be
promising in fuel cell technologies as they are cost-effective, easy to use and less
time-consuming than any other conventional methods.
430 B.G. Pollet and P.M. Sakkas

References

1. Sammes N (ed) (2006) Fuel cell technology: reaching towards commercialization. Springer,
London
2. Blomen LJMJ, Mugerwa MN (1993) Fuel cell systems. Plenum Press, New York
3. Mench MM (2008) Fuel cell engines. Wiley, London
4. O’Hayre R, Colella W, Cha S-W, Prinz FB (2009) Fuel cell fundamentals. Wiley, London
5. Srinivasan S (2006) Fuel cells: from fundamentals to applications. Springer, London
6. Pollet BG (ed) (2012) Power ultrasound in electrochemistry: from versatile laboratory tool to
engineering solution. Wiley, Chichester
7. Mason TJ, Lorimer JP (1998) Sonochemistry, theory, applications and uses of ultrasound in
chemistry. Ellis Horwood, Chichester
8. Thorneycroft J, Barnaby SW (1895) Torpedo-boat destroyers. Inst Civ Eng 122
9. Rayleigh L (1917) On the pressure developed in a liquid during the collapse of a spherical
cavity. Philos Mag 34(199-04):94–98
10. Richards WT, Loomis AL (1927) Chemical effects of high frequency sound waves I. A
preliminary survey. J Am Chem Soc 49:3086–3100
11. Pollet BG (1998) The effect of ultrasound upon electrochemical processes. Dissertation,
Coventry University, England, UK
12. Noltingk BE, Neppiras EA (1950) Cavitation produced by ultrasonics. Proc Phys Soc B 63:674
13. Pollet BG, Hihn J-Y, Doche M-L, Lorimer JP, Mandroyan A, Mason TJ (2007) Transport
limited currents close to an ultrasonic horn equivalent flow velocity determination. J
Electrochem Soc 154:E131–E138
14. Moriguchi N (1934) The influence of supersonic waves on chemical phenomena. III. The
influence on the concentration polarisation. J Chem Soc Jpn 55:749–750
15. Schmid G, Ehret L (1937) Beeinflussung der Metallpassivität durch Ultraschall. Z
Elektrochem 43:408–415
16. Schmid G, Ehret L (1937) Beeinflussung der Elektrolytischen Abscheidungspotentiale von
Gasen durch Ultraschall. Z Elektrochem 43:597–608
17. Kolb J, Nyborg W (1956) Small‐scale acoustic streaming in liquids. J Acoust Soc Am
28:1237–1242
18. Penn R, Yager E, Hovorka F (1959) Effect of ultrasonic waves on concentration gradients. J
Acoust Soc Am 31:1372
19. Bard A (1965) High speed controlled potential coulometry. Anal Chem 35:1125–1128
20. Mason TJ, Lorimer JP, Walton DJ (1990) Sonoelectrochemistry. Ultrasonics 28:333–337
21. Shen Q, Jiang L, Zhang H, Min Q, Hou W, Zhu J-J (2008) Three-dimensional dendritic Pt
nanostructures: sonoelectrochemical synthesis and electrochemical applications. J Phys Chem
C 112:16385–16392
22. Zin V, Pollet BG, Dabalá M (2009) Sonoelectrochemical (20 kHz) production of platinum
nanoparticles from aqueous solutions. Electrochim Acta 54:7201–7206
23. Shen Q, Min Q, Shi J, Jiang L, Zhang J-R, Hou W, Zhu J-J (2009) Morphology-controlled
synthesis of palladium nanostructures by sonoelectrochemical method and their application in
direct alcohol oxidation. J Phys Chem C 113:1267–1273
24. Qiu X-F, Xu J-Z, Zhu J-M, Zhu J-J, Xu S, Chen HY (2003) Controllable synthesis of palladium
nanoparticles via a simple sonoelectrochemical method. J Mater Res 18:1399–1404
25. Steele BCH (1999) Fuel cell technology: running on natural gas. Nature 400:619–620
26. Bessler WG, Vogler M, St€ ormer H, Gerthsen D, Utz A, Weber A, Ivers-Tiffée E (2010) Model
anodes and anode models for understanding the mechanism of hydrogen oxidation in solid
oxide fuel cells. Phys Chem Chem Phys 12:13888–13903
27. Park S, Vohs JM, Gorte RJ (2000) Direct oxidation of hydrocarbons in a solid-oxide fuel cell.
Nature 404:265–267
28. Trovarelli A (1996) Catalytic properties of ceria and ceria-containing materials. Catal Rev Sci
Eng 38:439–520
10 Sonoelectrochemical Production of Fuel Cell Nanomaterials 431

29. Tsipis EV, Kharton VV (2008) Electrode materials and reaction mechanisms in solid oxide
fuel cells: a brief review—I.Performance-determining factors. J Solid State Electrochem
12:1039–1060
30. Minh NQ, Takahashi T (1995) Science and technology of ceramic fuel cells. Elsevier,
Amsterdam
31. Ford DC, Nilekar AU, Xu Y, Mavrikakis M (2010) Partial and complete reduction of O2 by
hydrogen on transition metal surfaces. Surf Sci 604:1565
32. Peng G, Mavrikakis M (2015) Adsorbate diffusion on transition metal nanoparticles. Nano
Lett 15:629
33. Presvytes D, Vayenas CG (2007) Mathematical modeling of the operation of SOFC Nickel-
cermet anodes. Ionics 13:9–18
34. Qu NS, Zhu D, Chan KC (2006) Fabrication of Ni–CeO2 nanocomposite by electrodeposition.
Scr Mater 54:1421–1425
35. Faes A, Hessler-Wyser A, Presvytes D, Vayenas CG, Van herle J (2009) Nickel-zirconia anode
degradation and triple phase boundary quantification from microstructural analysis. Fuel Cells
9:841–851
36. Gong M, Liu X, Trembly J, Johnson C (2007) Sulfur-tolerant anode materials for solid oxide
fuel cell application. J Power Sources 168:289–298
37. Mark Ormerod R (2003) Solid oxide fuel cells. Chem Soc Rev 32:17–28
38. Brandon NP, Skinner S, Steele BCH (2003) Recent advances in materials for fuel cells. Annu
Rev Mater Res 33:183–213
39. Tao S, John Irvine JTS (2004) Catalytic properties of the perovskite oxide
La0.75Sr0.25Cr0.5Fe0.5O3-δ in relation to its potential as a solid oxide fuel cell anode material.
Chem Mater 16:4116–4121
40. Myung J-H, Ko H-J, Lee J-J, Lee J-H, Hyun S-H (2012) Synthesis and characterization of NiO/
GDC-GDC dual nano-composite powders for high performance methane fueled solid oxide
fuel cells. Int J Hydrogen Energy 37:11351–11359
41. Gavrielatos I, Drakopoulos V, Neophytides SG (2008) Carbon tolerant Ni-Au SOFC elec-
trodes operating under internal steam reforming conditions. J Catal 259:75–84
42. Bebelis S, Neophytides SG, Kotsionopoulos N, Triantafyllopoulos N, Colomer MT, Jurado J
(2006) Methane oxidation on composite ruthenium electrodes in YSZ cells. Solid State Ion
177:2087–2091
43. Tao S, Irvine JTS (2004) Synthesis and Characterization of (La0.75Sr0.25)Cr0.5Mn0.5O3-δ, a
Redox-Stable, Efficient Perovskite Anode for SOFCs. J Electrochem Soc 151:A252
44. Nikolla E, Schwank J, Linic S (2009) Comparative study of the kinetics of methane steam
reforming on supported Ni and Sn/Ni alloy catalysts: the impact of the formation of Ni alloy on
chemistry. J Catal 263:220–227
45. Sasaki K, Susuki K, Iyoshi A, Uchimura M, Imamura N, Kusaba H, Teraoka Y, Fuchino H,
Tsujimoto K, Uchida Y, Jingo N (2006) H2S poisoning of solid oxide fuel cells. J Electrochem
Soc 153:A2023–A2029
46. Matsuzaki Y, Yasuda I (2000) The poisoning effect of sulfur-containing impurity gas on a
SOFC anode: Part I. Dependence on temperature, time, and impurity concentration. Solid State
Ion 132:261–269
47. Trembly JP, Marquez AI, Ohrn TR, Bayless DJ (2006) Effects of coal syngas and H2S on the
performance of solid oxide fuel cells: single-cell tests. J Power Sources 158:263–273
48. Liu M, Wei G, Luo J, Sanger AR, Chuang KT (2003) Use of metal sulfides as anode catalysts
in H2S-Air SOFCs. J Electrochem Soc 150:A1025–A1029
49. Hahn K, Mavrikakis M (2014) Atomic and molecular adsorption on Re(0001). Top Catal 57:54
50. McEvoy AJ, Smith MJ (2007) Regeneration of anodes exposed to sulfur. ECS Trans
7:373–380
51. Yentekakis IV, Vayenas CG (1989) Chemical cogeneration in solid electrolyte cells. The
oxidation of formula to formula. J Electrochem Soc 136:996–1002
432 B.G. Pollet and P.M. Sakkas

52. Nilekar AU, Sasaki K, Farberow CA, Adzic RR, Mavrikakis M (2011) Mixed-metal Pt
monolayer electrocatalysts with improved CO tolerance. J Am Chem Soc 133:18574
53. Wei GL, Liu M, Luo JL, Sanger AR, Chuang KT (2003) Influence of gas flow rate on
performance of H2S/air solid oxide fuel cells with MoS2-NiS-Ag anode. J Electrochem Soc
150:A463–A469
54. Reisse J, Caulier T, Deckerkheer C, Fabre O, Vandercamrnen J, Delplancke JL, Winand R
(1996) Quantitative sonochemistry. Ultrason Sonochem 3:147–151
55. Sáez V, Mason TJ (2009) Review—sonoelectrochemical synthesis of nanoparticles. Mole-
cules 14:4284–4299
56. González-Garcı́a J, Esclapez MD, Bonete P, Hernández YV, Garretón LG, Sáez V (2010)
Current topics on sonoelectrochemistry. Ultrasonics 50:318–322
57. Pollet BG (2010) The use of ultrasound for the fabrication of fuel cell materials. Int J Hydrogen
Energy 35:11986–12004
58. Compton RG, Eklund JC, Marken F, Rebbitt TO, Akkermans RP, Waller DN (1997) Dual
activation: coupling ultrasound to electrochemistry—an overview. Electrochim Acta
42:2919–2927
59. Delplancke J-L, Di Bella V, Reisse J, Winand R (1994) Production of metal nanopowders by
sonoelectrochemistry. MRS Proc 372:75
60. Davis J, Vaughan DH, Stirling D, Nei L, Compton RG (2002) Cathodic stripping voltammetry
of nickel: sonoelectrochemical exploitation of the Ni(III)/Ni(II) couple. Talanta 57:1045–1051
61. Jia F, Hu Y, Tang Y, Zhang L (2007) A general nonaqueous sonoelectrochemical approach to
nanoporous Zn and Ni particles. Powder Technol 176:130–136
62. Liu Y-C, Lin L-H, Chiu W-H (2004) Size-controlled synthesis of gold nanoparticles from bulk
gold substrates by sonoelectrochemical methods. J Phys Chem B 108:19237–19240
63. Aqil A, Serwas H, Delplancke JL, Jérôme R, Jérôme C, Canet L (2008) Preparation of stable
suspensions of gold nanoparticles in water by sonoelectrochemistry. Ultrason Chem
15:1055–1061
64. Shen Q, Min Q, Shi J, Jiang L, Hou W, Zhu J-J (2011) Synthesis of stabilizer-free gold
nanoparticles by pulse sonoelectrochemical method. Ultrason Sonochem 18:231–237
65. Sakkas P, Schneider O, Martens S, Thanou P, Sourkouni G, Argirusis C (2012) Fundamental
studies of sonoelectrochemical nanomaterials preparation. J Appl Electrochem 42:763–777
66. Rao CNR, Muller A, Cheetan AK (2008) The chemistry of nanomaterials synthesis, properties
and applications, vol 1. Wiley-VCH Verlag GmbH & Co., Weinheim, p 151
67. Haas I, Shanmugam S, Gedanken A (2006) Pulsed sonoelectrochemical synthesis of size-
controlled copper nanoparticles stabilized by poly(N-vinylpyrrolidone). J Phys Chem B
110:16947–16952
68. Sáez V, Graves J, Paniwnyk L, Mason TJ (2010) Copper electrocrystallization on titanium
electrodes: controlled growth of copper nuclei using a potential step technique. Phys Procedia
3:111–115
69. Schneider O, Matić S, Argirusis C (2008) Application of the electrochemical quartz crystal
microbalance technique to copper sonoelectrochemistry Part 1. Sulfate-based electrolytes.
Electrochim Acta 53:5485–5495
70. Zhu J, Liu S, Palchik O, Koltypin Y, Gedanken A (2000) Shape‐controlled synthesis of silver
nanoparticles by pulse sonoelectrochemical methods tools. Langmuir 16:6396–6399
71. Liu S, Huang W, Chen S, Avivi S, Gedanken A (2001) Synthesis of X-ray amorphous silver
nanoparticles by the pulse sonoelectrochemical method. J Non Cryst Solids 283:231–236
72. Liu YC, Lin L-H (2004) New pathway for the synthesis of ultrafine silver nanoparticles form
bulk silver substrates in aqueous solutions by sonoelectrochemical methods. Electrochem
Commun 6:1163–1168
73. Jiang L-P, Wang A-N, Zhao Y, Zhang J-R, Zhu J-J (2004) Novel route for the preparation of
monodisperse silver nanoparticles via a pulsed sonoelectrochemical technique. Inorg Chem
Commun 7:506–509
10 Sonoelectrochemical Production of Fuel Cell Nanomaterials 433

74. Vu LV, Long NN, Doanh SC, Trung BQ (2009) Preparation of silver nanoparticles by pulsed
sonoelectrochemical method and studying their characteristics. J Phys Conf Ser 187:012077
75. Lei H, Tang Y-J, Wei J-J, Li J, Li X-B, Shi H-L (2007) Synthesis of tungsten nanoparticles by
sonoelectrochemistry. Ultrason Sonochem 14:81–83
76. Argirusis C, Matić S, Schneider O (2008) An EQCM study of ultrasonically assisted electro-
deposition of Co/CeO2 and Ni/CeO2 composites for fuel cell applications. Phys Status Solidi A
205:2400–2404
77. Xue Y-J, Liu H-B, Lan M-M, Li J-S, Li H (2010) Effect of different electrodeposition methods
on oxidation resistance of Ni-CeO2 nanocomposite coating. Surf Coat Technol 204:3539–3545
78. Lee D, Gan YX, Chen X, Kysar JW (2007) Influence of ultrasonic irradiation on the micro-
structure of Cu/Al2O3, CeO2 nanocomposite thin films during electrocodeposition. Mater Sci
Eng A 447:209–216
79. Gedanken A (2007) Doping nanoparticles into polymers and ceramics using ultrasound
radiation. Ultrason Sonochem 14:418–430
80. Sakkas PM, Schneider O, Sourkouni G, Argirusis C (2014) Sonochemistry in the service of
SOFC research. Ultrason Sonochem 21:1939–1947
81. Brenscheidt T, Nitschke F, S€ ollner O, Wendt H (2001) Molten carbonate fuel cell research
II. Comparing the solubility and the in-cell mobility of the nickel oxide cathode material in
lithium: potassium and lithium: sodium carbonate melts. Electrochim Acta 46:783
82. Kim MH, Hong MZ, Kim Y-S, Park E, Lee H, Ha H-W, Kim K (2006) Cobalt and cerium
coated Ni powder as a new candidate cathode material for MCFC. Electrochim Acta 51:6145
83. Dabala M, Pollet BG, Zin V, Campadello E, Mason TJ (2008) Sonoelectrochemical (20 kHz)
production of Co65Fe35 alloy nanoparticles from Aotani solutions. J Appl Electrochem
38:395–402
Chapter 11
Direct Ethanol Fuel Cell on Carbon
Supported Pt Based Nanocatalysts

T.S. Almeida, N.E. Sahin, P. Olivi, T.W. Napporn, G. Tremiliosi-Filho,


A.R. de Andrade, and K.B. Kokoh

11.1 Introduction

One of the biggest problems nowadays comes from the progressive increase in the
concentration of toxic gases due to combustion of fossil fuels, which represents a
large part of the energy consumed by the society. The greenhouse gas emissions,
especially in large urban centers, reach every year increasingly levels. Besides the
large industries, urban transport is another main factor responsible for pollution.
The need for cleaner energy sources and converter systems that combine high
efficiency and reduction of environmental impacts has attracted increasing interest
from the scientific community. In this context, fuel cells, system capable of
converting chemical energy into electrical energy, are receiving special attention
mainly due to its ability to generate electricity efficiently and not aggressive to the
environment. The versatility of such systems is the feasible use of various fuels,
efficient energy conversion, silent operation, and high applicability. These make
this system an attractive technology for a “clean” future [1].

T.S. Almeida
Universidade Federal do Tri^angulo Mineiro—Campus Iturama,
Iturama 382800-000, MG, Brazil
N.E. Sahin • T.W. Napporn • K.B. Kokoh (*)
Université de Poitiers, IC2MP UMR CNRS 7285, “Equipe SAMCat”,
4 rue Michel Brunet B27, TSA 51106, Poitiers Cedex 09 86073, France
e-mail: [email protected]
P. Olivi • A.R. de Andrade
Departamento de Quı́mica, Faculdade de Filosofia Ciências e Letras de Ribeir~ao Preto,
Universidade de S~ao Paulo, Ribeir~ao Preto 14040-901, SP, Brazil
G. Tremiliosi-Filho
Instituto de Quimica de S~ao Carlos, Universidade de S~ao Paulo,
S~ao Carlos 13560-970, SP, Brazil

© Springer International Publishing Switzerland 2016 435


K.I. Ozoemena, S. Chen (eds.), Nanomaterials for Fuel Cell Catalysis,
Nanostructure Science and Technology, DOI 10.1007/978-3-319-29930-3_11
436 T.S. Almeida et al.

Table 11.1 Types of fuel cells classified according to the electrolyte used
Working Efficiency
Type Electrolyte temperature ( C) (%)
Alkaline (AFC) KOH 60–90 50–60
Polymeric electrolyte (PEMFC) Polymer: Nafion® (H+) 50–90 50–60
Phosphoric acid (PAFC) H3PO4 (H+) 160–200 55
Molten Carbonate (MCFC) Molten carbonate (CO3) 650–700 60–65
Solid oxide (SOFC) ZrO2 (O2) 800–900 55–65

Fuel cell systems have a higher thermodynamic efficiency than internal com-
bustion engines or turbines, which convert chemical energy into heat to perform
useful work. Electrical devices produce energy without combustion, and thus there
is no need of moving parts. Additionally, fuel cells also reduce the emission of
pollutants. There are several types of fuel cells which are classified according to the
electrolytes employed [2]. Table 11.1 shows the main types of fuel cells.
Among the fuel cells cited above the proton exchange membrane fuel cell
(PEMFC) is one of the most promising converter systems for portable devices
[3, 4]. This system operates at low temperatures (50–90  C) using as electrolyte a
proton exchange membrane, with operational simplicity, high efficiency and low
gas pollutant emissions which makes its use even more attractive [2]. Nowadays the
best performance of PEMFCs is observed for those that use hydrogen as the fuel.
Hydrogen can be supplied to the cell either directly or indirectly by the reform of
liquid substances [5, 6]. However, the problems in the storage, handling, distribu-
tion, and high cost of the equipment used for the hydrogen generation represent a
major drawback to its implementation [7, 8]. For this reason, liquid fuels such as
methanol and ethanol are being investigated as a convenient alternative to the
replacement of hydrogen.
Ethanol becomes particularly attractive as an alternative fuel for use in direct
ethanol fuel cell (DEFC) because it can be easily produced from the biomass such
as sugar cane, corn and others, it is not harmful and presents high theoretical energy
density (8.6 kWh kg1) compared to methanol (6.1 kWh kg1) [9]. However, this
fuel presents challenging oxidation kinetics, requiring the development of catalysts
capable of breaking down the C-C bond [10, 11].
The principle of PEM-DEFC operation is illustrated in Fig. 11.1. The anode is
fed by an ethanol solution, while the cathode is composed by humidified oxygen, so
that good conductivity is maintained in the proton exchange membrane (PEM).
Ethanol is oxidized to carbon dioxide, and protons and electrons are produced on
the anode. Protons are then transferred through the polymeric electrolyte to the
cathode side, where they react with oxygen and electrons to produce water. The
equations are shown below:

Anode: CH3 CH2 OH þ 3H2 O ! 2CO2 þ 12Hþ þ 12e E ¼ 0:084 V vs: SHE
ð11:1Þ
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 437

Fig. 11.1 Schematic


diagram of a DEFC

Cathode: 3O2 þ 12Hþ þ 12e ! 6H2 O E ¼ 1:229 V vs: SHE


ð11:2Þ

Global Equation: CH3 CH2 OH þ 3O2 ! 2CO2 þ 3H2 O E ¼ 1:313 V vs: SHE
ð11:3Þ

There are two challenge that must be overcome to implement this technology:
(i) on the anode side, one must find a way to achieve the complete oxidation of
ethanol to CO2, to release the theoretical 12 electrons and also beat the strong
adsorption of the intermediates of ethanol oxidation onto the active sites of the
catalyst, which leads to a loss of efficient energy [10, 12]; (ii) on the cathode side,
enhance the slow kinetics of oxygen reduction at the temperature in which the
membrane is thermically stable (below 120  C).

11.2 Bi and Tri-Metallic Materials: Effect


of the Co-catalyst on the Pt-Based Anode

Pt-based catalysts are considered as the main material for ethanol electro-oxidation
in acidic medium. However, pure Pt does not display good catalytic activity due to
poisoning of its active sites with intermediates generated during ethanol
electrooxidation [13–15]. Bimetallic and trimetallic Pt-based catalysts can enhance
the oxidation of CO and small organic molecules due to a bifunctional mechanism
[16, 17] or to electronic effects [18]. For instance, metals such as Sn, Ni, Rh, Ru,
438 T.S. Almeida et al.

and W are largely investigated as co-catalysts in the composition of the electrode


materials in order to minimize the poisoning of Pt sites [13, 14, 19–29].
Of these, Pt-Sn composites display excellent activity for the oxidation of etha-
nol, as reported by Purgato et al. [30, 31], Neto et al. [13], and Almeida
et al. [22, 32]. This improvement is attributed to the bifunctional mechanism
exerted by the Sn and its oxides form which are responsible for the H2O activation
at their surfaces at low overpotentials as compared to pure Pt, which leads to a
reduced onset potential and enhanced catalytic activity of ethanol oxidation
[20, 33–35]. Nevertheless, even if PtSn has shown better results than pure Pt
catalyst, most of the products formed at this bimetallic composite still contain
C-C bonds, which suggests that the activity of PtSn/C, Pt and PtSn must be further
enhanced by the addition of a third element to improve the dehydrogenation
reaction and the cleavage of the C-C bond.
Ru is another metal largely investigated as a co-catalyst and plays a similar role
as Sn in ethanol electrooxidation. The adsorption and decomposition of ethanol and
its intermediate reaction products occur on Pt active sites, while the dissociative
adsorption of water occurs over Sn or Ru sites to form oxygen-containing surface
species. As reported by Emerson et al. [36] the addition of Ru to PtSn-based
catalysts showed good performance in the electrochemical oxidation of ethanol.
Its presence lowered the onset potential of ethanol oxidation (near 0.2 V), but
acetaldehyde was the major byproduct of the reaction. Antolini et al. [14] also
found that Ru addiction to PtSn catalysts enhances the catalytic activity and this
effect is related to the Ru/Sn ratio in the alloy and the synergistic effect of Ru and
Sn oxides.
More recently Ni has been studied as the second and/or third element added to
Pt, PtSn or PtRu catalysts. The main advantage of the introduction of this metal is
the shift of the oxidation potential of ethanol to lower values, coupled with an
increase in current density. The literature reveals that Sn and Ni decrease the energy
of the chemisorption of ethanol and its oxidation intermediates, such as CO on
Pt-sites [37, 38]. The addition of a third metal improves the oxophilic character of
the surface, thus raising the strength of the Sn-O bond and contributing to increased
acidity of the surface of the Sn-OH sites. This accentuates the bifunctional mech-
anism facilitating the ethanol electrooxidation [39].
The beneficial effect of Ni addition has been reported by Ribadeneira and Hoyos
[40] who evaluated the catalytic effect of Ni on binary PtSn/C and PtRu/C and
concluded that the addition of Ni on PtRu/C and PtSn/C catalysts increases signif-
icantly their catalytic activities to ethanol electro-oxidation. Bonesi et al. [41] also
studied the Ni effect on PtSn/C and PtRu/C and the chronoamperometry character-
ization showed a higher current density for the ternary composition PtSnNi/C, again
demonstrating the beneficial effect of Ni addition.
Rh has received great attention as a co-catalyst in ethanol electrooxidation due to
its ability to break the C-C bond more efficiently than other metals studied so far.
Souza et al. [42] carried out DEMS and in-situ FTIR studies and monitored the
catalytic activity of Pt, Rh and PtRh. They showed that Pt-Rh combination
decreased significantly the concentration of acetaldehyde and increased CO2
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 439

concentration during the ethanol electrooxidation as compared pure platinum. They


attributed this behavior to the formation of Rh-O species that act as oxygen supply
promoting the formation of CO-O bonds and finally CO2, therefore, releasing
12 electrons from the ethanol molecule. Sagar et al. [43] investigated Pt:Rh
(90:10, 70:30 and 35:65) catalysts and have shown an activity strongly dependent
on the Rh amount. Its presence above 30 % weight has a positive contribution to the
ethanol oxidation because it assists in breaking the C-C bond. Lima and Gonzalez
[19] investigated Pt-Ru, Pt-Rh, Pt-Ru-Rh activities and observed by XANES
characterization a pronounced electronic modification of the Pt 5d-band in the
presence of Rh and Ru. This consequently reduces the strong adsorption of ethanol
intermediates on Pt, increasing their rates of oxidation and decreases its oxidation
onset potential. The presence of Rh forms a larger quantity of acetic acid and CO2.
Rh has presented promising results for the C-C bond cleavage [44, 45]; its high cost
represents a barrier since it means more expensive catalysts.
Less investigated than other metals described so far, W has been reported in
some studies towards ethanol electrooxidation. Maillard [46] investigated the
tolerance to CO on Pt-WOx composites supported on carbon, where the character-
ization by X-ray absorption showed that WOx promotes slight modification of the
electronic structures of the platinum surface. The presence of WOx species assist
the electrooxidation of ethanol acting as a Bronstëd acid in the process of dehy-
drogenation and water molecules dissociation. Ribeiro et al. [25] and dos Anjos
et al. [26] observed a higher catalytic activity for the oxidation of ethanol at PtSnW
than at PtSn/C. They attributed the improvement of the catalytic activity to the
presence of Sn and W that decreased the overpotential for the intermediate species
removal.
A considerable number of different metals were evaluated as possible Pt struc-
ture modifiers. However, the metals described above have shown more significant
effects. With the large amount of materials already studied for ethanol
electrooxidation since the beginning of the development of fuel cells to date, it
can be found in the literature a large library of materials with different compositions
and metal ratios.

11.3 Catalyst Synthesis Methods

The synthesis methods are an essential part to obtain efficient catalysts for applying
in DEFC. It is known that the catalyst activity is dependent on their composition,
morphology, and particle size [47, 48] and the methodology applied to their
synthesis can lead to better control of these parameters [41, 49, 50]. In this context,
many research groups have dedicated efforts to develop efficient methods for the
synthesis of nanostructured catalysts, aiming at obtain materials with homogeneous
metallic distribution, small particle size, and high catalytic activity [49, 51–53].
Below we describe some of the methods employed in the catalysts synthesis.
440 T.S. Almeida et al.

Microemulsion method: This method is convenient to synthesize nanoparticles by


using a reducing agent, normally NaBH4, in combination with different strategies
for better control of nanoparticle size and composition. More precisely, the
microemulsion method is based on the use of two immiscible solvents in the
presence of a surfactant forming stable microscopic structures called micelles or
reverse micelles. On a macroscopic scale, the resulting solution seems to be
homogeneous but in fact, on a nanometer scale, the solution is heterogeneous and
it is composed of nanodroplets dispersed in a continuous organic phase. The
micellar medium is responsible to control the chemical reduction rate of metallic
cations and particles size. The solvent and the surfactant used must be chosen
carefully to allow a good control of the micelle volume which is directly related to
the particle size, and the reducing agent should be stable in the medium (solvent)
avoiding parallel reaction with the micelle components [54].
Pt-based catalysts can be easily synthesized by preparing a microemulsion by
dissolving the metal precursors in ultrapure water then adding an organic solvent
such as n-heptane following by the addition of polyethylene glycol dodecyl ether
(Brij® 30, surfactant). After mixing the solutions to form reverse micelles the
carbon support is added following by the reducing agent sodium borohydride
(NaBH4). Finally after nanoparticle formation, the material is filtered, washed
and ground [54, 55]. For example, Godoy et al. [21] have prepared PtSn/C catalysts
with sizes around 2.5 nm that were well dispersed on the carbon support. The
authors claimed that the reduction of metals trapped in the micelles enabled the
achievement of well-distributed particles.
Alcohol-reduction method: Developed by Toshima and Yonezawa [56, 57] this
method provides colloidal dispersions of nanoparticles with uniform particle size
and homogeneous distribution. The method consists in refluxing an alcohol solution
that acts as a solvent and reducing agent [52], usually ethylene glycol, containing
the metallic ions and a stabilizing agent (polymer or surfactant), providing homo-
geneous colloidal dispersions of the corresponding metal nanoparticles. This
method offers some advantages such as good reproducibility, satisfactory distribu-
tion, and small particle size.
Neto et al. [13] have investigated PtRu/C, PtSn/C, and PtSnRu/C catalysts
prepared by the alcohol-reduction method using an ethylene glycol/water solution.
Particle of the order of 2.7 nm were achieved for the catalysts. Spinacé et al. [23]
have used this method for the production of PtSn/C, PtRh and PtSnRh/C catalysts.
They also obtained catalysts with small particle size (2.0 nm), which culminated in
significant catalytic activity for ethanol electrooxidation.
Formic acid method: This method was developed with the purpose of obtaining
dispersed platinum catalysts supported on carbon for PEMFC fuel cells [58]. Basi-
cally, the catalyst is obtained by the addition of a high surface area carbon support
(Vulcan XC-72, carbon nanotube, etc.) in a formic acid solution, which is used as a
reducing agent. The resulting mixture is heated to 80  C followed by addition of
aliquots of a solution containing platinum and other metal salts to the reaction
vessel. After complete metal reduction, the nanocatalyst is filtered, dried, and
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 441

grounded. Antolini et al. [14] have looked into the effects of introducing ruthenium
into PtSn/C catalysts prepared by the formic acid method. They obtained the
required catalytic composition with particles of small size (3.5 nm) homogeneously
distributed onto the carbon support.
B€onnemann method [55, 59]: This method consists in the production of a stable
colloid in an inert and dry atmosphere using non-hydrated metallic chlorides and
solvents. This approach can be successfully applied to obtain Pt-based catalysts
containing different metals and metal oxides, in particular Sn, V, W and Mo, and
other transition-metal elements such as Cu, Fe, Co and Ni [60].
The anhydrous salts are dissolved in anhydrous tetrahydrofuran (THF) with an
appropriate amount of tetraoctylammonium bromide [N(oct)4Br]. The reducing
agent is produced by dissolving [N(oct)4Br] in THF and adding [KHBr(Et)3],
resulting in the formation of a stark reducing agent, triethylhydroborate of
tetraoctylammonium [N(oct)4HBr(Et)3] (Eq. 11.4). The reducing agent will also
act as a surfactant after metal reduction, preventing any agglomeration of the
metallic particles.

N ðoctÞ4 Br þ KHBðetÞ3 ! N ðoctÞ4 HBðetÞ3 þ KBr ð11:4Þ

For the reduction of metallic ions, a 50 % excess over the stoichiometric amount
of [N(oct)4HBr(Et)3] is added to the solution of metal salts and heated under
stirring. The reduction occurs with hydrogen evolution, as shown in Eq. 11.5, and
the solution darkening.
 þ
MeXn þ N ðoctÞ4 HBr ðetÞ3 ! M* N ðoctÞ4 þ nBðetÞ3 þ n=2 H2 þ nX ð11:5Þ

After the synthesis of the stable colloid, a suspension of carbon black in THF is
prepared and the colloid solution is added in a dropwise fashion to the carbon black
dispersion, finishing the preparation of the powder catalyst with filtration and
several washes using THF and ethanol. The catalyst is calcined at 300  C for 1 h
under air atmosphere to remove the organic surfactant.
Thermal decomposition of polymeric precursors (DPP) method: The method-
ology based on the Pechini method [61] was initially developed for the preparation
of ceramic materials and later adapted for the preparation of electrode materials.
Firstly, citric acid (CA) is dissolved in a polyhydroxylic alcohol (ethylene glycol,
EG) at 60  C. Metallic salt (Pt or another metal) is added and the temperature is
raised to 90  C, forming a polymeric resin containing the metallic cation [22, 25,
30, 31]. To obtain the catalyst, the resins are stoichiometrically mixed with carbon
Vulcan XC-72 or any other support, and fired at high temperatures; e.g.,
300–400  C. The major advantage of this route is the attainment of robust catalysts
with experimental composition close to the nominal one. This method has been
employed in the synthesis of a series of Pt-based materials [25, 26, 30, 62].
It is important to know that the degree of esterification, hardness and porosity of
the resin can be controlled by taking in account the molar ratio between the
complexing agent [CA], the metal cation [METAL], and the ethylene glycol
442 T.S. Almeida et al.

[EG]. Thus, two molar ratios: [CA]/[METAL] ¼ CM and [CA]/[EG] ¼ CE can be


defined. The microstructure of the resulting powder is related with these two ratios.
CM describes the degree of the chelation process of the metal in the organic
product. For solutions with low CM ratios (CM < 1), there may be precipitation
of the metals salts in the solution, and therefore, no uniformity in the chelation of
the metals [63], which is due to an insufficient amount of organic molecules to
chelate metallic cations. On the other hand, for relatively high CM ratios, it is easy
to dissolve the metal ions but the amount of organics to be removed is excessive. CE
describes the degree of esterification between the chelating agent and the ethylene
glycol and affects powder morphologies. Equimolar ratio leads to porous resin [64]
generating a rigid polyester net that reduces any segregation of metals during the
polymer decomposition process at high temperatures [65].
Microwave-assisted heating method: This method has gained importance due to
its operational simplicity, efficiency, and reduced preparation time [23, 53, 66, 67].
It promotes rapid reduction of the metallic precursors and is responsible for the
formation of nanometric particles size [68]. However, its application must meet
some criteria; the solvent acts as a reducing agent and must show high reduction
conversion of metallic ions, high capacity for conversion of electromagnetic energy
into thermal energy, and suitable temperature profile for reduction for all metallic
ions. Figure 11.2 depicts a voltammogram of a PtSn/C electrode material obtained
from this synthetic method.
The efficiency of reducing agents in stabilizing the nanoparticles and preventing
formation of metal clusters is related to the oxidizing species that is generated
during the heating process [53]. Oxidation of small-chain alcohols occurs through
interaction between the hydroxyl group(s) present in these molecules and the
metallic ions, which are subsequently reduced to metallic particles forming
carbonyl and/or carboxyl species which adsorbs onto the particle surface promoting
their stabilization and preventing their agglomeration, which culminates in
formation of small particles [69].

Fig. 11.2 Cyclic


voltammogram of the
Pt70Sn30/C (20 wt.% metal
loading) electrode material
recorded in supporting
electrolyte 0.05 mol L1
H2SO4 and in the presence
of 1.0 mol L1 ethanol at
10 mV s1. Reprinted with
permission from
[82]. Copyright (2012),
Elsevier
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 443

Ethanol may be used as a reducing agent and its oxidation under microwave-
assisted heating leads to formation of acetaldehyde and acetic acid, which are the
stabilizing species. When ethylene glycol is applied as the reducing agent, there is
also the formation of acetaldehyde, which acts also as a reducing agent, and acetic
acid as stabilizing agent [70, 71]. In another proposal, Bock et al. [69] claimed that
the stabilizing agent from microwave-assisted heating of ethylene glycol should be
oxalic and glycolic acid. The use of propylene glycol differs from ethylene glycol
only in the final obtained stabilizer that are lactic and pyruvic acids [72, 73].
Besides the formation of stabilizing species during the heating processes, some
stabilizing compounds can be used in the nanoparticle synthesis that improve the
size control and the particle distribution [66, 74, 75]. These stabilizing agents can
act in two different ways depending on their chemical structure. When the stabi-
lizing species is a surfactant it can hinder the nanoparticle growth due to steric
hindrance, where the presence of a large carbon chain in the structure of these
molecules prevents the formation of metal clusters during the reduction process
[76]. In another approach, the use of anions such as citrate and acetate assists in
electrostatic stabilization. These molecules have the ability to form chelates that
electrostatically repels species with similar charge preventing the formation of
metallic clusters [70, 77]. Electrostatic stabilizing agents act in two ways: (i) in
the formation of stabilizing agent-metal ion pairs; and (ii) adsorbing on fresh-
formed nanoparticles surface to prevent further reduction and cluster formation.
Thus, the more efficient the stabilizing agents the smaller the particle size and more
homogeneous will be its distribution onto de carbon support [78].
Table 11.2 summarizes different catalyst compositions synthesized by various
methodologies, highlighting the particle and crystallite sizes. As can be seen in

Table 11.2 Average particle size and crystallite size for different catalysts prepared by different
literature methods
Particle size Crystallite size
Methodology Composition (TEM) (nm) (DRX) (nm) Ref.
Microemulsion Pt75Sn25 3.0 – [21]
Alcohol reduction Pt52Sn48 3.0 2.5 [23]
Pt52Sn42Ni06 2.4 – [79]
Pt67Sn11Rh22 2.0 3.0 [23]
Pt49Sn29Ru22 3.3 – [13]
Formic acid Pt77Sn33 4.5 3.9 [52]
Pt40Sn45Pd15 2.5 3.1 [58]
Pt30Sn37Rh32 1.8 3.0 [80]
DPP Pt78Sn22 5.0 7.4 [25]
Pt72Sn28 8.2 6.5 [22]
Pt3Sn1Ni1 5.1 3.7 [81]
Pt83Sn09Ni06 4.3 3.3 [22]
Pt78Sn13Ru09 5.7 4.8 [36]
MW Pt35Sn65 4.2 3.9 [62]
Pt73Sn27 2.8 2.0 [82]
444 T.S. Almeida et al.

Table 11.2, the particle and crystallite size is virtually the same for all the methods
listed. However it is important to notice that each method has specific advantages
that may be useful depending on the desired catalyst structure and morphology.
Here we listed only usual methods of synthesis of nanoparticles dispersed on
carbon. However, for specific structures, such as core-shell, single crystals and
others more specific methods must be applied.

11.4 Oxygen Reduction Reaction on Carbon Supported


Pt Catalyst

The electrocatalytic reduction of molecular oxygen has been intensively researched


for many practical systems because this reaction plays a key role for achieving the
overall reaction in fuel cells. A multi-electron transfer electrode for molecular
oxygen reduction in acid solution is of primary concern because of its possible
utilization as cathode in fuel cells. The mechanism of the electrochemical oxygen
reduction reaction is quite complicated and involves many intermediates, primarily
depending on the nature of the electrode material, catalyst and electrolyte.
Stonehart and Ross [83] proposed the use of rotating thin layer electrode (RTLE)
for the direct measurement of electrochemical reaction rate constant for the first
time. To elucidate the oxygen reduction mechanism, the rotating disk electrode
(RDE) and rotating ring disk electrode (RRDE) have been widely used since RDE
and RRDE enable the determination of kinetics and the quantitative detection of the
intermediate reaction species [84–90]. In these methods the oxygen reduction
reaction (ORR) occurs on the disk electrode and the produced hydrogen peroxide
is detected at the ring.

11.4.1 Oxygen Reduction Reaction Pathways

As the ORR involves four electrons, four protons and a O-O bond cleavage, it has a
complex mechanism in acidic [91–93], alkaline [88, 94, 95] and neutral [90]
electrolytes at low and moderate temperatures. Depending on the electrode material
where the electrochemical oxygen reduction reaction takes place, H2O2 may be
generated during the ORR as an intermediate or a final product [96–99].
Since the first reaction scheme proposed for the oxygen reduction reaction by
Damjanovic et al. [100], several reaction schemes have been proposed in order to
identify the mechanistic reaction pathways. Besides the oxygen reduction reaction
schemes proposed by Damjanovic et al. [100], Wroblowa et al. [101], Appleby
et al. [102] Anastasijevic et al. [88], and Zurilla et al. [103], also suggested
modification on the ORR mechanism. Figure 11.3 shows a simplified scheme that
gives a diagnostic criterion to discriminate between the different pathways.
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 445

O2 (bulk) k1 (+ 4e )
-

diff.
- -
k5 k2 (+2e ) k3 (+2e )
O2 (sur) O2 (ads) (HOOH)(ads) 2H2O
-
k-5 k-2 (-2e )
k-6 k
6
k4

diff.
(HOOH)(sur) H2O2 (bulk)

Fig. 11.3 Scheme showing the oxygen reduction reaction pathways proposed by Wroblowa
et al. [101] in acidic media. Reprinted with permission from [101]. Copyright (1976), Elsevier

According to the scheme shown in Fig. 11.3, one O2 molecule is adsorbed on a


catalytic site at the electrode surface, and it may either follow the direct four-
electron reduction pathway to form water by means of Eq. 11.6 and k1, or
two-electron pathway to generate hydrogen peroxide by means of Eq. 11.7 and
k2. Hydrogen peroxide either desorbs from the surface (k6), is further reduced to
water by means of Eq. 11.8 and k3, moreover it can desorb from the surface or may
chemically decompose to form water and oxygen by means of Eq. 11.9 and k4. For
the hydrogen peroxide decomposition, two pathways (electrochemical and radical)
were envisaged. In the case of platinum, hydrogen peroxide decomposition is a
combination of electrochemical reduction and oxidation. Hydrogen peroxide
decomposition can also be defined as a radical reaction [97, 104, 105]. Molecular
oxygen may be converted into oxygenated (O*), hydroxyl (OH*) and
superhydroxyl (OOH*) intermediate species, which are not detectable experimen-
tally. Adsorption and coverage of O2 on the platinum active surface may affect the
reaction pathway depending on the dissociative and associative mechanism
[106, 107]. According to the density functional theory (DFT) calculations [106],
oxygen reduction tends to follow an associative mechanism where superhydroxyl
(OOH*) is first formed and then O-O bond is cleaved in the case of two-electron
pathway. Additionally, in the direct four-electron reduction pathway, dissociative
mechanism is dominant at low oxygen coverage where O-O bond is cleaved before
superhydroxyl (OOH*) formation.

O2 þ 4H þ þ 4e ! 2H 2 O E ¼ 1:229 V vs: SHE ð11:6Þ

O2 þ 2Hþ þ 2e ⇆ H 2 O2 E ¼ 0:682 V vs: SHE ð11:7Þ

H 2 O2 þ 2H þ þ 2e ! 2H 2 O E ¼ 1:776 V vs: SHE ð11:8Þ

2H 2 O2 ! 2H2 O þ O2 ð11:9Þ
446 T.S. Almeida et al.

From the thermodynamic data based on Nernst equation for the electrochemical
reactions (in Eqs. 11.6–11.8), the corresponding equilibrium potentials are 1.229,
0.695, 1.763 V versus the standard hydrogen electrode (SHE), respectively
(Eqs. 11.10–11.12).

RT a2  
EO2 =H2 O ¼ EOo 2 =H2 O  ln H2 O4 EOo 2 =H2 O ¼ þ1:229 V ð11:10Þ
4F aO2 aHþ

RT aH2 O2  
EO2 =H2 O2 ¼ EOo 2 =H2 O2  ln EOo 2 =H2 O2 ¼ þ0:682 V ð11:11Þ
2F aO2 a2Hþ

RT a2H2 O  
EH2 O2 =H2 O ¼ EHo 2 O2 =H2 O  ln EHo 2 O2 =H2 O ¼ þ1:776 V ð11:12Þ
2F aH2 O2 a2Hþ

where the Eon is defined as standard potential of electrochemical reaction n (V), R is


the ideal gas constant (8.314 J mol1 K1), T is the absolute temperature (K),
F is the Faraday constant (96,485 C mol1) and ai is the activity of the compound
i [108, 109].
In the case of alkaline media, while direct four-electron pathway takes place
according to Eq. 11.13, peroxide pathway is carried out via Eq. 11.14, and followed
by either the reduction of peroxide ion through Eq. 11.15 or decomposition of
peroxide via Eq. 11.16.
The peroxide pathways related to acidic medium (Eqs. 11.7 and 11.8) and
alkaline medium (Eqs. 11.14 and 11.15) may contain adsorbed-states of superoxide
  
2 , superhydroxide (HO2 ) and peroxide ðO2 H Þ, (H2O2). As the adsorbed
O*
peroxide species tend to desorb from the electrode surface, the mechanism can be
distinguished [97, 110, 111].

O2 þ 2H2 O þ 4e ! 4OH  E ¼ 0:401 V ð11:13Þ

O2 þ H 2 O þ 2e ! HO
2 þ OH

E ¼ 0:076 V ð11:14Þ

HO 
2 þ H 2 O þ 2e ! 3OH

E ¼ 0:867 V ð11:15Þ

2HO 
2 ! 2OH þ O2 ð11:16Þ

Thanks to its high energy conversion, avoiding of corrosion of carbon supports, and
the elimination of possible degradation of electrode material because of the perox-
ide formation and free radicals, the most efficient pathway is the direct four-
electron reduction of adsorbed oxygen to water without peroxide formation. More-
over, peroxide formation tends to attack the membrane and accelerate the dissolu-
tion of platinum, which causes a deterioration of the durability. Therefore, oxygen
reduction with high selectivity for the direct four-electron reduction pathway is
desired. Great effort has been put into the development of selectively catalyzing the
four-electron reduction of oxygen to produce water with low overpotential and low
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 447

H2O2-production. That’s why, it is important to understand the intrinsic character-


istics of the surface of the platinum catalyst by quantifying the yield of H2O2, that is
the molar ratio of produced H2O2 to the total of all reduced oxygen [99, 106, 112].

11.4.2 RDE Study on Oxygen Reduction Kinetic

Metal nanoparticles have attracted a great deal of interest in scientific research and
industrial applications due to their unique properties based on large surface to
volume ratio and quantum size effects [113–115]. It is clear that geometric (Pt-Pt
bond distance) [116] and electronic (Pt d-band vacancy) [117] effects are of great
importance in the electrocatalytic activity of the Pt/C electrocatalyst. The most
efficient catalyst for oxygen reduction reaction is platinum, as its surface reactivity
supply catalytic process mainly depending on O and OH bonding energy on the
platinum surface [118]. With sluggish reaction kinetics the oxygen reduction
reaction includes a multi-electron transfer process involving different reaction
intermediates, thus the oxygen reduction requires high overpotentials. To overcome
this electrochemical voltage loss, catalysis of oxygen reduction requires a high
platinum loading (typically 0.1–0.5 mg cm2) in fuel cell cathodes [119].
The oxygen reduction reaction has been analyzed on carbon supported platinum
(Pt/C) electrocatalysts in 0.1 mol L1 HClO4 through cyclic voltammetry
(CV) [120, 121] and rotating disk electrode (RDE) [60, 84, 122–124] techniques.
Electrochemical rotating disk electrode (RDE) experiments were conducted in a
conventional three-electrode electrochemical cell in 0.1 mol L1 HClO4. The
working electrode was composed of 25 μg of catalyt powders by dropping a
catalytic ink onto the freshly polished glassy carbon disk. A saturated calomel
electrode (0.244 V vs. RHE) and the glassy carbon plate (5 cm2 geometric surface
area) were the reference electrode and the counter electrode, respectively. To avoid
the electrolyte contamination by chlorides the reference electrode was separated
from the working electrode compartment by a closed Luggin-Haber ionic bridge.
The working electrode potential is controlled by an applied driving potential
difference between the reference and working electrodes. Owing to the very short
distance between the Luggin-Haber capillary tip and the working electrode, in most
cases ohmic (iR) corrections are not required. All potentials used were converted to
reversible hydrogen electrode (RHE) scale.

11.4.3 Levich and Koutecky´-Levich Plots

Rotating disk electrode (RDE) measurements allow analysis of the electrocatalytic


activity of the oxygen reduction reaction. Under the steady-state conditions and
considering the electrode process is described in Eq. 11.17, which gives a sigmoidal
polarization curves with a diffusion-limited current density ( jd).
448 T.S. Almeida et al.

O þ n e ! R ð11:17Þ

Figure 11.4a displays the polarization curves plotted by current density versus disk
potential for the Pt/C in 0.1 mol L1 HClO4 recorded at 5 mV s1 and 20  C as a
function of the rotation rates. It is clearly seen that the oxygen reduction reaction
takes place under a mixed kinetic and diffusion control region in the potential range
0.94 and 0.70 V vs. RHE followed by a purely diffusion-limited region. By using
Levich plots ( j versus ω1/2), the diffusion-limited current density ( jd) without any
film formation can be calculated using Eq. 11.18 from the Levich law [125].

jd ¼ 0:20 nFDO2 CO2 v1=6 ω1=2 ¼ Bω1=2


2=3
ð11:18Þ

where B is the Levich constant, n is the number of transferred electrons per oxygen
molecule, F is the Faraday constant (96,485 C mol1); DO2 , is the coefficient
diffusion of oxygen in 0.1 mol L1 HClO4 (1.7  105 cm2 s1), CO2 is the oxygen

Fig. 11.4 RDE measurement results performed with O2 saturated 0.1 mol L1 HClO4 recorded at
5 mV s1 and 20  C. (a) polarization curves on Pt/C electrocatalyst at various rotation rates
(0, 400, 900, 1600, 2500 rpm), (b) Koutecký-Levich plots at different potentials, in which the
reaction is controlled by both diffusion and kinetic effects (0.81, 0.79, 0.77, 0.75, 0.73 V vs. RHE),
(c) exponential plot of limiting current density, and (d) Tafel plots for low current density (l.c.d)
and high current density (h.c.d) regions
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 449

concentration in the considered electrolyte (1.3  103 mol dm3) in the electro-
lyte, v is the kinetic viscosity of the electrolyte (1.0  102 cm2 s1), ω is the
rotation rate of the disk electrode, and 0.20 is the coefficient used when the ω is
expressed in revolution per minute (rpm) [122]. The calculated jd value from the
Levich law according to the Eq. (11.18) is 7.14 mA cm2 for four transferred
electrons. As can be seen in Fig. 11.4a, the current density plateau increases
proportionally by increasing the rotation rate in diffusion controlled area and the
current density at 2500 rpm is nearly 7.2 mA cm2, corresponding to an oxygen
reduction pathway involving mainly direct four-electrons pathway. Additionally,
from the slopes of the Levich plots, the number of transferred electrons involved in
oxygen reduction can be calculated by using Eq. 11.18 [126, 127].
For the reactions controlled by both diffusion and kinetics through the hydrody-
namic voltammetry, the measured current density ( j) can be written as dependent
on the diffusion-limited current density ( jd) and the kinetic current density ( jk) in
Eq. 11.19 as well-known Koutecký-Levich (K-L) approach [124, 128–130], which
provides the kinetic parameters as regards the catalyst surface. Jahn and Vielstich
illustrated the Koutecký-Levich approach on the rates of the Fe2+/Fe3+ redox
system [131]. A typical Koutecký-Levich plot is a linear function between inverse
of the current densities versus inverse of the square root of the rotation rates (1/j
versus 1/ω1/2) for the Pt/C catalyst as shown in Fig. 11.4b. The plots of 1/j versus
1/ω1/2 are anticipated to yield straight lines with the intercept corresponding to jk in
Eq. 11.20 and the slopes reflecting the so-called B constant value.

1 1 1
¼ þ ð11:19Þ
jjj jjd j jjk j

In the case of a catalytic film, the kinetic current density ( jk) can be represented by
Eq. 11.20.

1 1 1 1 1
¼ þ þ þ ð11:20Þ
jjj jjd j jf jjads j jjo ðθ=θe Þ  eðη=bÞ j

where jf and jads correspond to diffusion-limited current density in the catalytic film
and to the adsorption-limited current density of molecular oxygen, respectively.
Additionally, η¼E-Eeq is the overpotential, b is the Tafel slope, jo is the exchange
current density, θ and θe are the coverage of platinum surface by species are caused
oxygen adsorption at potential E and at the equilibrium potential Eeq [124, 132]. If
assuming that the electron transfer is the rate determining step, then the adsorption
step is faster than the electron transfer reaction, so θ  θe for all electrode potentials
[111, 128]. As the film diffusion-limited current density ( jf) and the adsorption-
limited current density ( jads) do not depend on the disk electrode rotation rate and
applied potential (E), the kinetic current density can be expressed with the limiting
current density ( jlim) in Eq. 11.21.
450 T.S. Almeida et al.

Table 11.3 RDE and RRDE results: number of electrons (ne), limiting current density (jlim),
exchange current density (jo), Tafel slopes (b), kinetic current density from intercept value at
0.90 V (jk), and real kinetic current density normalized with the active surface area ( jk) for the Pt/C
catalyst
jlim b jo b jo jk @0.90 V
Pt/C ne (mA cm2) (mV dec1) (mA cm2) (mV dec1) (mA cm2) (mA cm2)
RDE 3.91 33.25 68.92 1.58  105 114.50 2.59  103 0.012
RRDE 3.93 38.50 66.74 1.65  105 116.44 3.84  103 0.011

1 1 1
¼ þ ð11:21Þ
jjk j jjlim j jjo ðθ=θe Þ  eðη=bÞ j

By extrapolating the intercepts of the K-L plots as a function of the electrode


potential, the limiting current density ( jlim) can be estimated as shown in Fig. 11.4c.
Then, in order to access Tafel slope, Eq. 11.21 is written in Eq. 11.22. Tafel plots of
jk
the kinetic current densities giving a straight line are drawn between log jl jk versus
electrode potential in Fig. 11.4d. From the Tafel slope and the intercept at the
equilibrium electrode potential, one could estimate ½logðjlim =jo Þ value. Then the
exchange current density is evaluated with the calculated (jlim) value by using
Eq. 11.22. The kinetic parameters (ne, jlim, b, jo) obtained from K-L analysis were
listed in Table 11.3. The RDE measurements indicated that the Pt/C catalyst is
active for molecular oxygen reduction in acid media with multi-electron charge
transfer (n ¼ 4), to form water.
    
j jk
η ¼ b log lim þ log ð11:22Þ
jo jlim  jk

Tafel slope is expressed in unit of mV/dec. The unit “dec” refers to decades of
current density. Two distinct Tafel slopes are obtained with the same values,
indicating that mechanism of oxygen adsorption and the first electron transfer are
similar for RDE and RRDE experiments. Typically, the Tafel slopes are 60 mV/dec
and 120 mV/dec for the oxygen reduction reaction at low current density (l.c.d.) and
high current density (h.c.d.) regions, respectively [133–135]. Tafel slopes were
close to 67.0  1.0 mV/dec at low current density regions and 115.0  1.0 mV/dec
at high current density regions for the Pt/C electrocatalyst on RDE and RRDE
experiments, indicating that Tafel slope regions change from Temkin to Langmuir
conditions for changing coverage of surface by adsorbed oxygen species.
The intrinsic catalytic activity of an electrochemical reaction is defined by the
exchange current density, which is an important kinetic parameter representing the
electrochemical reaction rate at equilibrium. For an electrochemical reaction, both
forward and backward reactions can occur. At equilibrium, the net current density
of the reaction is zero. The current density of the forward reaction and equals that of
the backward reaction. This current density is called exchange current density.
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 451

The magnitude of the exchange current density determines how rapidly the
electrochemical reaction can occur. The catalytic activity of the Pt/C catalyst
can be usually evaluated in terms of the current density per geometric area j, and
the kinetic current density ( jk), which is free from the mass transport, can be
evaluated by active surface area under identical operating conditions
[136, 137]. The calculated values of the kinetic current density on RDE and
RRDE listed in Table 11.3.

11.4.4 Oxygen Reduction Reaction Order

Different potentials in the region of mass transport control can be selected to


determine the reaction order with respect to O2, and the limiting current at each
rotation rate. The plots of log(i) versus log[(id  i)/id] at measured potentials where
the ORR is expected to be under the mixed kinetic/diffusion control should yield
straight and parallel lines, confirming that the oxygen reduction on Pt/C catalyst
surface is the first order reaction [85, 93, 103].
 
i
logðiÞ ¼ m log 1  þ logik ð11:23Þ
id

where i is the measured current, id is the diffusion-limiting current, ik is the kinetic


current and m is the reaction order. The slope of the log(i) versus log[(id  i)/id] plot
in Fig. 11.5 found about 0.99 for Pt/C electrocatalyst, confirming that kinetics of
oxygen reduction in HClO4 electrolyte is the first order reaction.

Fig. 11.5 Reaction order


(m) plots of log (i) versus
log [(id  i)/id] at various
rotation rates at 0.81 V
vs. RHE for oxygen
reduction reaction on the
Pt/C electrocatalyst in
0.1 mol L1 HClO4
452 T.S. Almeida et al.

11.4.5 RRDE Study on Oxygen Reduction Pathways

Rotating ring disk electrode (RRDE) consisted of a glassy carbon disk (0.196 cm2
geometric area) and a Pt ring (0.11 cm2 geometric area) sealed in a polytetrafluor-
oethylene holder. The electrochemical measurements were performed with a
bipotentiostat and rotation control. The RRDE technique provides further informa-
tion than RDE during the oxygen reduction reaction. The Pt ring electrode potential
was kept at 1.2 V vs. RHE, at which all H2O2 molecules that reach at the ring were
oxidized to H2O as a function of the applied potential. In addition to the oxidation of
H2O2 at the ring, which produces an anodic current, decomposition of the H2O2
without current flow may occur according to reaction in Eq. 11.9.

11.4.5.1 Collection Efficiency (N )

A typical ferrocyanide/ferricyanide redox reversible couple was used for determi-


nation of the collection efficiency in 0.1 mol L1 NaOH supporting electrolyte,
containing 0.01 mol L1 K3[Fe(CN)6]. The constant ring potential of 1.20 V
vs. RHE enables the oxidation of the [Fe(CN)6]4 at the disk electrode, to
[Fe(CN)6]3 in 0.1 mol L1 NaOH electrolyte. That is, while the reduction of ferric
ions to ferrous ions on the disk electrode according to Eq. 11.24, the oxidation of
ferrous ions occurs via mass transport limited rate at the ring electrode according to
Eq. 11.25.

Fe3þ þ e ! Fe2þ ð11:24Þ

Fe2þ ! Fe3þ þ e ð11:25Þ

The collection efficiency was determined by averaging the ratio of the disk current
between 0.4 and 1.0 V vs. RHE and the corresponding ring current according to
Eq. 11.26 [112, 138].

N ¼ ir =id ð11:26Þ

Figure 11.6 shows (a) ring and (b) disk currents regarding the collection efficiency
related to RRDE at various rotation rates. The experimental collection efficiency
value decreases with increasing the rotation rates because of the combined effects
of the disk current that increases more rapidly than ring electrode response,
depending on the laminar flow. A plot of the ring to disk currents (ir/id) versus
disk potential E is shown in Fig. 11.6c. From the six independent measurements, the
collection efficiency was evaluated as N ¼ 0.20.
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 453

Fig. 11.6 Linear scan


voltammograms obtained
on RRDE in 0.1 mol L1
NaOH containing
0.01 mol L1 K3[Fe(CN)6]
recorded at (a) ring, (b) disk
electrode with 100, 400,
900, 1600 and 2500 rpm
at a constant ring potential
of 1.20 V vs. RHE and
20 mV s1, and (c)
the collection efficiency
obtained by the ring to
disk current ratio

11.4.5.2 H2O2 Detection and Kinetic Parameters by RRDE

From rotating ring disk electrode measurements as shown in Fig. 11.7, it is possible
to assume that oxygen reduction reaction on Pt/C catalyst involves the four-electron
transfer per oxygen molecule by producing mainly H2O. The calculations based on
Eq. 11.27 [87] demonstrate that H2O proportion efficiency p(H2O) is 99.71 % listed
in Table 11.4. N represents the collection efficiency, jd represents the disk current
density and jr,l represent the limiting current density for the disk potentials Ed lower
than 0.9 V vs. RHE and jor,l represents the limiting current density for disk
potentials Ed higher than 0.9 V vs. RHE.
 
N  jjd  jro, l  1
r, l
pð H 2 O Þ ¼   ð11:27Þ
jd
N j  jro, l þ 1
r, l

From the calculated data using Eq. 11.27 in Fig. 11.7, a very small amount of H2O2
was detected at high overpotentials, i.e. below 0.8 V vs. RHE for the cathodic scan
direction, corresponding to n ¼ 3.93 as shown in Fig. 11.8. During the oxygen
reduction reaction on the disk, two-electron and four-electron reduction take
place forming peroxide as intermediate, which is detected on the ring electrode.
The current corresponds to the two-electron reduction on the disk is equal to |ir/N|,
454 T.S. Almeida et al.

Fig. 11.7 Polarization


curves (a) ring, (b) disk
electrode oxygen reduction
reaction on Pt/C
electrocatalyst recorded at
1600 rpm in O2 saturated
0.1 mol L1 HClO4 and
5 mV s1

Table 11.4 For RRDE—collection efficiency (N ), the number of exchanged electrons (ne), H2O
proportion efficiency p(H2O), the percentage of peroxide formation (% H2O2) at 1600 rpm and the
total flux of consumed oxygen (Fd) at disk electrode for Pt/C catalyst
N% ne % pH2O % H2O2 Fd  105
RRDE 20.55 3.93 99.71 0.29 2.84

Fig. 11.8 Number of


transferred electrons
determined from ring and
disk currents by using
Eq. 11.29 for the oxygen
reduction reaction on Pt/C
electrocatalyst recorded at
1600 rpm in O2 saturated
0.1 mol L1 HClO4 and
5 mV s1
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 455

while the current corresponds to the four-electron reduction is the difference


between total measured disk current |id| and |ir/N|. Therefore, the total flux of
oxygen consumed at the disk electrode (Fd) is determined using Eq. 11.28
[93, 112, 139, 140].

jir =N j id  jir =N j id
Fd ¼ þ ¼ ð11:28Þ
2FA 4FA nFA

Figure 11.8 represents the average number of electron transfer (ne), which is
evaluated by using Eq. 11.29 as a function of the disk potential as regards the
Pt/C electrode catalyst. There is a decrease around 0.75 V vs. RHE and a further
decrease towards lower potentials, where the peroxide formation occurs. Equa-
tion 11.29 is used to calculate the average number of transferred electrons involved
in oxygen reduction reaction. The results show that the value of ne is very close to
4, indicating that the oxygen reduction reaction on Pt/C surface takes place through
direct four-electron reaction to form water.

4
ne ¼ ð11:29Þ
1 þ jir =N  id j

According to the polarization curves of the Pt/C in Fig. 11.9, the region of the mixed
kinetic and diffusion control (0.95 V < E < 0.70 V) is followed by diffusion limited
current region (0.70 V < E < 0.30 V). The kinetic data regarding the rotating ring
disk electrode experiments are summarized in Table 11.4. Moreover, the behavior
of the ring electrode clearly denotes the amount of H2O2 formed, which increases
when the disk potential decreases [98]. When the potential range decrease from
0.55 to 0.30 V, a decrease of ring current and an increase of disk current are
observed with increasing the rotation rate, indicating that the intermediate H2O2
continues to reduce to water. The H2O2 produced at the disk electrode is oxidized at
the ring by the reverse reaction in Eq. 11.7. The percentage of the current to
generate water is calculated by using the collection efficiency, ring and disk
currents in Eq. 11.30. The amount of peroxides was 0.29 %, which is in good
agreement with the data obtained from Eq. 11.27 as summarized in Table 11.4.

N jI d j  I r
pðH 2 OÞ ¼ ð11:30Þ
N jI d j þ I r

By using the polarization curves of the Pt/C obtained from RRDE results in
Fig. 11.10, Koutecký-Levich plots (Fig. 11.10a), limiting current density
(Fig. 11.10b), Tafel plots (Fig. 11.10c) and peroxide amount (Fig. 11.10d) formed
during the oxygen reduction reaction obtained from Eq. 11.30 were confirmed by
rotating disk electrode results in Table 11.3.
We can conclude from the data related to Tables 11.3 and 11.4 that O2 reduction
proceeds principally through the four-electron reduction pathway with 99.71 %
percentage of H2O and 0.29 % H2O2 formation, as shown in Fig. 11.10d.
456 T.S. Almeida et al.

Fig. 11.9 Polarization


curves (a) ring, (b) disk
electrode response for the
oxygen reduction reaction
on Pt/C electrocatalyst
recorded at different
rotation rates in O2
saturated 0.1 mol L1
HClO4 and 5 mV s1

11.5 Analytical Determination of Intermediates


and Reaction Products Resulted from Ethanol
Oxidation

In order to achieve the development of suitable electrocatalysts for the


electrooxidation of organic molecules it is important to know what are the products
and intermediates formed in that kind of reaction. These parameters are necessary
to determine the reaction mechanism and the role of the distinct catalytic materials.
Different techniques may be employed in order to identify intermediates and
reactions products. In situ techniques must be used in order to determine interme-
diates that can be formed and consumed during the reaction and that exist only into
the double layer region. This kind of determination is difficult to be done in the fuel
cell and they are applied normally using electrochemical cells with special config-
uration. The formed products can be determined ex situ using both fuel cells and
electrochemical cells by the application of different techniques.
Several studies on the electrooxidation of ethanol have been devoted to the
identification and quantification of the adsorbed intermediates and by-products onto
the electrode by means of various techniques, such as Differential Electrochemical
Mass Spectrometry (DEMS) [141–146], in situ Fourier-Transform Infrared Spec-
troscopy (FTIR) [24, 147, 148], and High Performance Liquid Chromatography
(HPLC) [22, 62, 149, 150].
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 457

Fig. 11.10 Polarization curves (a) ring, (b) disk electrode responses for the oxygen reduction
reaction on Pt/C electrocatalyst recorded at different rotation rates in O2 saturated 0.1 mol L1
HClO4 and 5 mV s1

11.5.1 In Situ Infrared Spectroscopy

In situ FTIR technique allows tracking of the formation and evolution of the
products during ethanol oxidation. The technique also provides information about
the reaction intermediates adsorbed on the surface of the catalyst or the reaction
products present in the immediate vicinity of the electrode.
The ethanol oxidation involves a large number of intermediates and reaction
products such as acetaldehyde, acetic acid, carbon monoxide and carbon dioxide.
The use of FTIR techniques is also useful to compare the selectivity of the various
catalysts. FTIR studies can be accomplished on the adsorption and oxidation of key
reaction intermediates such as CO, acetaldehyde or acetic acid in relation to the
applied potential to refine the understanding of the reaction mechanism.
In these techniques, the cell geometry is adapted to the spectroscopic apparatus
in order to allow the external reflection and the recording of the reflected beam
[151]. The working electrode is typically a glassy carbon or gold disk in which
surface the catalytic material is dispersed. The electrode is inserted against a
window that is transparent to incident radiation in the spectral range concerned.
In the investigation of ethanol electrooxidation, a calcium fluoride (CaF2) window
458 T.S. Almeida et al.

which is transparent in the 1000–3000 cm1 wavenumber range [152], is usually


employed.
There are two main in situ FTIR reflection techniques generally used in order to
investigate ethanol oxidation or any organic molecule oxidation. The SPAIRS
(Single Potential Alteration Infrared Reflectance Spectroscopy) technique is used
to monitor the presence of reaction intermediates in situ, and the formation of
reaction products, such as CO2, on and in the vicinity of the electrode. On other
hand, the SNIFTIRS (Subtractively Normalized Interfacial Fourier Transform
Infrared Reflectance Spectroscopy) technique is more suitable to the in situ detec-
tion of adsorbed species from dissociative chemisorption of ethanol. These two
techniques can give us complementary information.
The SPAIRS method is based on the acquisition of reflectivity spectra at
different potentials recorded during a slow variation of the potential (1 mV s1)
in a linear voltammetry or first voltammetric cycle. The spectra are recorded every
50 mV and the potential range depends on the catalyst composition considering that
some metals such as Ru, Mo, Rh and others are soluble at high potential values.
Normally the studies are done in between 0.05 and 1.4 V vs. RHE. The results
recorded consist in a spectrum obtained from the accumulation of n interferograms
processed by Fourier transform subtracting recorded reference potential spectrum
as showed by the expression below (Eq. 11.31):

ΔR REi  REref
Δabsorbance ¼ ¼ ð11:31Þ
REref REref

where REi represents the reflectivity on the potential where the spectra accumula-
tion were recorded and REref is the spectrum of reference taken on the beginning or
on the end of the positive potential variation.
When the calculation of ΔR/R is done with Ei > Eref, each negative band
observed on spectrum corresponds to the formation of a species while a positive
band corresponds to the consumption of a reagent.
Figure 11.11 shows representative SPAIRS spectra during ethanol oxidation.
The interpretation of SPAIRS can be easily made knowing the respective
wavenumber of the all intermediates and byproducts from ethanol oxidation.
Table 11.5 summarizes the species formed during the oxidation of ethanol and
their respective bands observed in the infrared spectrum.
The intense band at 1640 cm1 is attributed to interfacial water [154] and the
strong band at 1200 cm1 is characteristic of HSO4 ions in the thin layer of
solution between the working electrode and the CaF2 window. The anion is formed
due to the reaction of SO42 with H+ released during the oxidation of ethanol
[30, 155].
The band at 1100–1120 cm1 is assigned to the vibration of acetaldehyde C-H
bonds, δ-(C-H) [24]. The formation of acetic acid can be seen by the appearance of
two band at 1280 cm1 and 1380 cm1 [156–158] arising from the C-O stretch and
–OH deformation, respectively. The presence of this molecule is also evidenced by
the broad band at 2615-2650 cm1 that is assigned to the asymmetric stretching
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 459

Fig. 11.11 Representative SPAIRS spectra of the species coming from ethanol oxidation on PtSn/
C. 0.1 mol L1 Ethanol in 0.5 mol L1 H2SO4, T ¼ 25  C

Table 11.5 Different species and their respective bands in the infrared spectrum [24, 153]
Wavenumber (cm1) Functional group Assignment
2983, 2906 CH3, CH2 C-H stretch
2632 –CH3 (acetic acid) C-H stretch
2341 CO2 C-O asymmetric stretch
2055–2060 –CO Linearly bonded
1725 COOH or CHO C¼O stretch
1402 –CH3COO C-O asymmetric stretch
1370/1280 COOH C-O stretch + OH deformation
1600 Interfacial water H-O-H stretch

ν(C-H) of the –CH3 group [24]. The band at 1725 cm1 corresponds to the
stretching of the carbonyl group ν(C¼O) in both acetaldehyde and acetic acid,
indicating their formation [141, 156, 159]. The band between 2030 and 2050 cm1
corresponds to CO adsorbed linearly on the Pt sites [160–162] and can be observed
appearing weakly in 2035 cm1 at 300 mV vs. RHE. Above 400 V, COL begins to
be oxidized to CO2, as evidenced by the appearance of the band at 2343 cm1,
attributed to the asymmetrical stretching (O¼C¼O).
The use of SPAIRS techniques provides access to the potential in which different
species appear and disappear and allow comparison of different catalyst materials.
The example below (Fig. 11.12) compares different PtSn-based catalyst towards
ethanol oxidation by monitoring the evolution of the COL and CO2 bands.
460 T.S. Almeida et al.

Fig. 11.12 Evolution of


CO (a) and CO2 (b) band
intensity against the
potential for different
catalyst composition. Inset
corresponds to the band
intensities of the catalysts
PtSn, PtSnRh, and PtSnRu

It can be seen in Fig. 11.12 that the catalyst containing Ni exhibits the most
intense COL band, justifying their liability to break the C-C bond. COL appears
around 300 mV and has a maximum at 500 mV vs. RHE, and then there is a drop in
intensity due to its oxidation to CO2. The formation of CO2 as evidenced by the
band at 2340 cm1 was observed for all compositions and this molecule begins to
be formed at 500–600 mV vs. RHE. As a conclusion, one can say that the catalysts
containing Ni, Rh and Ru enhanced the CO2 formation in comparison with binary
catalysts which showed a low amount of CO2 indicating that these catalysts are less
active for breaking the C-C bond.
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 461

11.5.2 High Performance Liquid Chromatography

The quantification of ethanol oxidation products can be obtained using HPLC


technique with two different approaches. First, a conventional electrochemical
cell is used in a semi-cell condition in which the anodic reaction is the ethanol
electrooxidation. This approach allows investigation of different conditions such as
temperature, ethanol concentration, catalyst composition, and so on, even in the
potentiostatic or galvanostatic modes. The products are analyzed after the electrol-
ysis or small aliquots can be withdrawn in order to determine the product concen-
tration and evolution. Volatile compounds can be determined by trapping them in
appropriate solutions. In this experiment, a proton exchange membrane separates
the electrolytic cell containing two compartments. The anode side must remain
sealed to avoid losses of volatile compounds like CO2, acetaldehyde and even
ethanol (Fig. 11.13).
The second method is to analyze the solution from the anode side of a direct
ethanol fuel cell according to the scheme shown in Fig. 11.14a. In both systems, the
products are separated with an HPX-87H column under isocratic conditions using
3.33 mmol L1 H2SO4 and a flow rate of 0.6 mL min1 (Fig. 11.14b).
At the end of the long-term electrolysis, N2 is bubbled through the solution, to
quantify the volatile compounds produced during the process in separate trap
compartments. Acetaldehyde is trapped in a 0.2 wt% of 2,4-dinitrophenyl-hydra-
zine solution in 2.0 mol L1 HCl. Subsequently, the concentration of the solid

Fig. 11.13 Electrolytic cell scheme used to study ethanol oxidation and determine its products by
chromatography analysis. Reprinted with Permission from [149]
462 T.S. Almeida et al.

Nitrogen Acetaldehyde CO2


Acetic acid
Acetaldehyde
DEFC
NaOH
solution

Anode side 2,4 DNPH solution

b Eluent: H2SO4 3.33 mM


Column: HPX-87H Ethanol
Flow rate: 0.6 mL min-1

Acetaldehyde

Acetic acid

10 15 20 25
time / min

Fig. 11.14 (a) Set-up for recovering reaction products of DEFC for chromatography analysis;
Reprinted with permission from [149]. Copyright (2006), with permission from Elsevier. (b)
separation of the reaction products in the bulk solution. Reprinted from [30]. Copyright (2009),
with permission from Elsevier

hydrazone formed therein is quantified after dissolution in ethyl acetate, by means


of an NH2P-50 (Asanhipak NH2P series) column or any column used to determine
amine compounds, under isocratic elution using a mobile phase consisting of
acetonitrile: H2O (40:60 v/v), at a 0.6 mL min1 flow rate.
The derivatization of acetaldehyde to allow its quantification is shown in the
equation below (Eq. 11.32):
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 463

The released CO2 was trapped in 0.1 mol L1 NaOH. The possibly formed
carbonate is quantitatively analyzed by comparison with a Na2CO3 reference
prepared under the same conditions.
Based on the results of the techniques described above the global mechanism of
ethanol oxidation may be summarized according to the following scheme [10]:

CH3 CH2 OH ! ½CH3 CH2 OHads ! CO2 ðtotal oxidationÞ ð11:33Þ


CH3 CH2 OH ! ½CH3 CH2 OHads ! CH3 CHO
! CH3 COOH ðpartial oxidationÞ ð11:34Þ

From this general equation, highlights of the detailed mechanisms have been
proposed by different groups [34, 141, 159, 161] and can be summarized as
described below [25, 30]:

Pt þ CH3 CH2 OH ! Pt  ðOCH2 CH3 Þads þ e þ Hþ ð11:35Þ


 þ
Pt  ðOCH2 CH3 Þads ! Pt þ CH3 CHO þ e þ H ð11:36Þ

or,

Pt þ CH3 CH2 OH ! Pt  ðCHOH  CH3 Þads þ e þ Hþ ð11:37Þ


 þ
Pt  ðCHOH  CH3 Þads ! Pt þ CH3 CHO þ e þ H ð11:38Þ
 þ
Pt þ CH3 CHO ! Pt  ðCO  CH3 Þads þ e þ H ð11:39Þ
Pt þ H2 O ! Pt  OHads þ e þ Hþ ð11:40Þ
Pt  ðCO  CH3 Þads þ Pt  OHads ! 2Pt þ CH3 COOH ð11:41Þ
Pt þ CH3 CH2 OH ! Pt  COads þ Pt  CHx ð11:42Þ
 þ
Pt  COads þ Pt  OHads ! CO2 þ e þ H ð11:43Þ
464 T.S. Almeida et al.

11.6 Electrical Performances of a Single DEFC

Finding optimum operation condition of DEFC must be analyzed in order to have


realistic information about the practical use of this device. Heysiattalab et al. [163]
have reported a systematic study of key parameters that affect the DEFC perfor-
mance. Among them, they investigated temperature, ethanol concentration and
oxygen pressure effects. These authors have shown that the increase of temperature
not only contributes to the kinetic process but also influences the ethanol oxidation
and oxygen reduction rates, because the increase of the diffusion will increase
concentration of ethanol and oxygen in the catalyst layer, leading to a rise of ethanol
oxidation and oxygen reduction. Also, temperature increase enhances the mem-
brane conductivity, which will increase the current density and fuel cell perfor-
mance. The increase of ethanol concentration leads to an enhanced fuel cell
performance, because it improves diffusion and consequently the ethanol concen-
tration in the catalyst layer. However, they noticed that high ethanol concentration
leads to a decrease in the cell performance, mainly because of ethanol crossover
through the membrane that causes a decrease in the oxygen reduction rate in the
cathode side. In their study, they have found that 0.5 mol L1 of ethanol was the
best concentration to reach the maximum current density and power. The pressure
in the cathode side has the same effect of ethanol concentration; by increasing O2
pressure the cell performance improves because more O2 would be introduced in
the catalyst layer that facilitates its reduction.

Fig. 11.15 Electrical performances of a 5 cm2 single DEFC recorded at 90  C using tin modified
platinum catalyst. (2 mg cm2 catalyst loading, 40 % catalyst on Vulcan XC-72 carbon); cathode
material: 2 mg cm2 (Pt/C); membrane: Nafion® 117; [Ethanol] ¼ 2 mol L1; Pethanol ¼ 1 bar;
PO2 ¼ 3 bar. Reproduced and adapted with permission from [31]. Copyright (2012), Elsevier
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 465

It is important to notice that the catalyst composition and the deposited amount
on the diffuse layer also significantly affect the performance of the fuel cell
(Fig. 11.15). There is still no consensus between the different groups that investi-
gate DEFC performances related to the control of the parameters that influence the
performance of the cell. Table 11.6 summarizes some catalyst compositions,
operating conditions of the cell and the observed power to give us an overview of
results obtained so far.

Table 11.6 Summary of PEM-DEFC performances


Catalyst Operating conditions Power (mW cm2) Ref.
Pt3Sn1 PO2 ¼ 2 bar; [EtOH] ¼ 1.5 mol L1; 82 [164]
T ¼ 90  C; anode (2.0 mgPt cm2)
Pt1Sn1 PO2 ¼ 2 bar; [EtOH] ¼ 2.0 mol L1; 60 [165]
T ¼ 100  C; anode (1.0 mgPt cm2)
Pt3Sn1 PO2 ¼ 2 bar; [EtOH] ¼ 1.5 mol L1; 55 [166]
T ¼ 90  C; anode (1.5 mgPt cm2)
Pt3Sn1 PO2 ¼ 2 bar; [EtOH] ¼ 1.0 mol L1; 80 [167]
T ¼ 90  C; anode (1.5 mgPt cm2)
Pt3Sn2 PO2 ¼ 2 bar; [EtOH] ¼ 1.0 mol L1; 40 [168]
T ¼ 75  C; anode (1.3 mgPt cm2)
Pt1Sn1 PO2 ¼ 2 bar; [EtOH] ¼ 1.0 mol L1; 55 [20]
T ¼ 90  C; anode (1.0 mgPt cm2)
Pt89Sn11 PO2 ¼ 3 bar; [EtOH] ¼ 2.0 mol L1; 35 [169]
T ¼ 90  C; anode (2.0 mgPt cm2)
Pt90Sn10 PO2 ¼ 3 bar; [EtOH] ¼ 2.0 mol L1; 72 [62]
T ¼ 110  C; anode (2.0 mgPt cm2)
Pt80Sn20 PO2 ¼ 3 bar; [EtOH] ¼ 2.0 mol L1; 37 [31]
T ¼ 90  C; anode (2.0 mgPt cm2)
Pt75Sn25 PO2 ¼ 3 bar; [EtOH] ¼ 1.0 mol L1; 20 [52]
T ¼ 90  C; anode (1.0 mgPt cm2)
Pt50Ru50 PO2 ¼ 2 bar; [EtOH] ¼ 1.0 mol L1; 45 [20]
T ¼ 90  C; anode (2.0 mgPt cm2)
Pt52Ru48 PO2 ¼ 1 bar; [EtOH] ¼ 2.0 mol L1; 60 [39]
T ¼ 80  C; anode (3.0 mgPt cm2)
Pt50Ru50 PO2 ¼ 2 bar; [EtOH] ¼ 2.0 mol L1; 20 [75]
T ¼ 90  C; anode (1.0 mgPt cm2)
Pt86Sn10Ru04 PO2 ¼ 3 bar; [EtOH] ¼ 2.0 mol L1; 50 [149]
T ¼ 80  C; anode (3.0 mgPt cm2)
Pt65Sn19Ru15 PO2 ¼ 3 bar; [EtOH] ¼ 2.0 mol L1; 10 [170]
T ¼ 80  C; anode (2.0 mgPt cm2)
Pt51Sn38Rh11 PO2 ¼ 2 bar; [EtOH] ¼ 2.0 mol L1; 42 [23]
T ¼ 100  C; anode (1.0 mgPt cm2)
(continued)
466 T.S. Almeida et al.

Table 11.6 (continued)


Catalyst Operating conditions Power (mW cm2) Ref.
Pt86Sn07Rh07 PO2 ¼ 3 bar; [EtOH] ¼ 1.0 mol L1; 10 [171]
T ¼ 80  C; anode (2.0 mgPt cm2)
Pt67Sn20Rh13 PO2 ¼ 3 bar; [EtOH] ¼ 2.0 mol L1; 7 [170]
T ¼ 80  C; anode (2.0 mgPt cm2)
Pt75Sn14Pd11 PO2 ¼ 3 bar; [EtOH] ¼ 1.0 mol L1; 8 [171]
T ¼ 90  C; anode (2.0 mgPt cm2)
Pt40Sn45Pd15 PO2 ¼ 3 bar; [EtOH] ¼ 2.0 mol L1; 14 [170]
T ¼ 90  C; anode (2.0 mgPt cm2)
Pt81Sn11Ni08 PO2 ¼ 3 bar; [EtOH] ¼ 1,0 mol L1; 34 [171]
T ¼ 80  C; anode (2.0 mgPt cm2)
Pt81Sn11Ni08 PO2 ¼ 3 bar; [EtOH] ¼ 2.0 mol L1; 45 [172]
T ¼ 80  C; anode (2.0 mgPt cm2)
Pt82Sn09Ni09 PO2 ¼ 3 bar; [EtOH] ¼ 2.0 mol L1; 45 [170]
T ¼ 80  C; anode (2.0 mgPt cm2)

11.7 Conclusion

Based on fundamental aspects, investigations were carried out on both the ethanol
oxidation reaction at various Pt-based catalysts and the oxygen reduction reaction
on Pt/C. The helpful results obtained gave useful information on the electrochem-
ical behaviors of the fuel (ethanol) and the oxidant (O2) in acid medium for the
construction of a DEFC device. Indeed, ethanol is a sustainable molecule; its
oxidation at the surface of mono, bi and trimetallic Pt-based anodes showed that
a number of transition metals must be used as co-catalysts to decrease the content of
the precious noble metal, to enhance the current densities and to shift the onset
potential toward lower values. Otherwise, the kinetics of the ORR have been
introduced with details on the main pathway of this reduction process at the Pt/C
surface. Note that it proceeds with a nearly 4-electron process, indicating the
formation of hydrogen peroxide in low amount. For the moment and in the better
case, the delivered power density in the single DEFC performed was inferior to
100 mW cm2 whatever the catalyst composition, which is mainly due to a weak
C-C bond cleavage. It is hoped that the oxophilics transition metals such as Sn, Ru,
Rh were active to remove CO to CO2, leading to notice that the main reaction
products of ethanol oxidation were acetaldehyde and acetic acid. Anyway the
findings collected in this chapter open an excellent window of many studies on
various renewable and friendly environmentally fuels that will pave the way for
further developments in direct alcohol fuel cell technology.
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 467

References

1. Wu J, Yuan XZ, Martin JJ, Wang H, Zhang J, Shen J, Wu S, Merida W (2008) A review of
PEM fuel cell durability: degradation mechanisms and mitigation strategies. J Power Sources
184:104–119
2. Wendt H, G€otz M, Linardi M (2000) Tecnologia de células a combustı́vel. Quı́m Nova
23:538–546
3. Wang CY (2004) Fundamental models for fuel cell engineering. Chem Rev 104:4727–4765
4. Song SQ, Tsiakaras P (2006) Recent progress in direct ethanol proton exchange membrane
fuel cells (DE-PEMFCs). Appl Catal B Environ 63:187–193
5. Perna A (2007) Hydrogen from ethanol: theoretical optimization of a PEMFC system
integrated with a steam reforming processor. Int J Hydrogen Energy 32:1811–1819
6. Chutichai B, Authayanun S, Assabumrungrat S, Arpornwichanop A (2013) Performance
analysis of an integrated biomass gasification and PEMFC (proton exchange membrane
fuel cell) system: hydrogen and power generation. Energy 55:98–106
7. Str€obel R, Oszcipok M, Fasil M, Rohland B, J€ orissen L, Garche J (2002) The compression of
hydrogen in an electrochemical cell based on a PE fuel cell design. J Power Sources
105:208–215
8. Takeichi N, Senoh H, Yokota T, Tsuruta H, Hamada K, Takeshita HT, Tanaka H,
Kiyobayashi T, Takano T, Kuriyama N (2003) “Hybrid hydrogen storage vessel”, a novel
high-pressure hydrogen storage vessel combined with hydrogen storage material. Int J
Hydrogen Energy 28:1121–1129
9. Lamy C, Lima A, LeRhun V, Delime F, Coutanceau C, Léger J-M (2002) Recent advances in
the development of direct alcohol fuel cells (DAFC). J Power Sources 105:283–296
10. Antolini E (2007) Catalysts for direct ethanol fuel cells. J Power Sources 170:1–12
11. Calegaro ML, Suffredini HB, Machado SAS, Avaca LA (2006) Preparation, characterization
and utilization of a new electrocatalyst for ethanol oxidation obtained by the sol-gel method. J
Power Sources 156:300–305
12. Lamy C, Rousseau S, Belgsir EM, Coutanceau C, Léger J-M (2004) Recent progress in the
direct ethanol fuel cell: development of new platinum–tin electrocatalysts. Electrochim Acta
49:3901–3908
13. Neto AO, Dias RR, Tusi MM, Linardi M, Spinacé EV (2007) Electro-oxidation of methanol
and ethanol using PtRu/C, PtSn/C and PtSnRu/C electrocatalysts prepared by an alcohol-
reduction process. J Power Sources 166:87–91
14. Antolini E, Colmati F, Gonzalez ER (2007) Effect of Ru addition on the structural charac-
teristics and the electrochemical activity for ethanol oxidation of carbon supported Pt-Sn
alloy catalysts. Electrochem Commun 9:398–404
15. Tripkovic AV, Lovic JD, Popovic KD (2010) Comparative study of ethanol oxidation at
Pt-based nanoalloys and UPD-modified Pt nanoparticles. J Serb Chem Soc 75:1559–1574
16. Watanabe M, Motoo S (1975) Electrocatalysis by ad-atoms. 2. Enhancement of oxidation of
methanol on platinum by ruthenium ad-atoms. J Electroanal Chem 60:267–273
17. Watanabe M, Motoo S (1975) Electrocatalysis by ad-atoms. 3. Enhancement of oxidation of
carbon-monoxide on platinum by ruthenium ad-atoms. J Electroanal Chem 60:275–283
18. Grant JL, Fryberger TB, Stair PC (1985) Charge-transfer electronic effects on chemically
modified Mo(100) surfaces. Surf Sci 159:333–352
19. Lima FHB, Gonzalez ER (2008) Ethanol electro-oxidation on carbon-supported Pt-Ru, Pt-Rh
and Pt-Ru-Rh nanoparticles. Electrochim Acta 53:2963–2971
20. Li H, Sun G, Cao L, Jiang L, Xin Q (2007) Comparison of different promotion effect of PtRu/
C and PtSn/C electrocatalysts for ethanol electro-oxidation. Electrochim Acta 52:6622–6629
21. Godoi DRM, Perez J, Villullas HM (2010) Alloys and oxides on carbon-supported Pt-Sn
electrocatalysts for ethanol oxidation. J Power Sources 195:3394–3401
22. Almeida TS, Kokoh KB, De Andrade AR (2011) Effect of Ni on Pt/C and PtSn/C prepared by
the Pechini method. Int J Hydrogen Energy 36:3803–3810
468 T.S. Almeida et al.

23. Spinacé EV, Dias RR, Brandalise M, Linardi M, Neto AO (2010) Electro-oxidation of
ethanol using PtSnRh/C electrocatalysts prepared by an alcohol-reduction process. Ionics
16:91–95
24. Garcia-Rodriguez S, Rojas S, Pena MA, Fierro JL, Baranton S, Léger J-M (2011) An FTIR
study of Rh-PtSn/C catalysts for ethanol electrooxidation: effect of surface composition.
Appl Catal B Environ 106:520–528
25. Ribeiro J, dos Anjos DM, Léger J-M, Hahn F, Olivi P, de Andrade AR, Tremiliosi-Filho G,
Kokoh KB (2008) Effect of W on PtSn/C catalysts for ethanol electrooxidation. J Appl
Electrochem 38:653–662
26. Dos Anjos DM, Hahn F, Léger J-M, Kokoh KB, Tremiliosi-Filho G (2008) Ethanol
electrooxidation on Pt-Sn and Pt-Sn-W bulk alloys. J Braz Chem Soc 19:795–802
27. Shen PK, Xu C (2006) Alcohol oxidation on nanocrystalline oxide Pd/C promoted
electrocatalysts. Electrochem Commun 8:184–188
28. Xu C, Shen PK, Liu Y (2007) Ethanol electrooxidation on Pt/C and Pd/C catalysts promoted
with oxide. J Power Sources 164:527–531
29. Xu C, Tian Z, Shen P, Jiang SP (2008) Oxide (CeO2, NiO, Co(3)O(4) and Mn3O4)-promoted
Pd/C electrocatalysts for alcohol electrooxidation in alkaline media. Electrochim Acta
53:2610–2618
30. Purgato FLS, Olivi P, Léger J-M, de Andrade AR, Tremiliosi-Filho G, Gonzalez ER,
Lamy C, Kokoh KB (2009) Activity of platinum-tin catalysts prepared by the Pechini-
Adams method for the electrooxidation of ethanol. J Electroanal Chem 628:81–89
31. Purgato FLS, Pronier S, Olivi P, de Andrade AR, Léger J-M, Tremiliosi-Filho G, Kokoh KB
(2012) Direct ethanol fuel cell: electrochemical performance at 90  C on Pt and PtSn/C
electrocatalysts. J Power Sources 198:95–99
32. Palma LM, Almeida TS, De Andrade AR (2013) high catalytic activity for glycerol
electrooxidation by binary Pd-based nanoparticles in alkaline media. ECS Trans 58:651–661
33. Su BJ, Wang KWE, Tseng CJE, Wang CHA, Hsueh YU (2012) Synthesis and catalytic
property of PtSn/C toward the ethanol oxidation reaction. Int J Electrochem Sci 7:5246–5255
34. Arenz M, Stamenkovic V, Blizanac BB, Mayrhofer KJ, Markovic NM, Ross PN (2005)
Carbon-supported Pt–Sn electrocatalysts for the anodic oxidation of H2, CO, and H2/CO
mixtures: Part II: The structure–activity relationship. J Catal 232:402–410
35. Koper MTM (2004) Electrocatalysis on bimetallic and alloy surfaces. Surf Sci 548:1–3
36. Cunha EM, Ribeiro J, Kokoh KB, de Andrade AR (2011) Preparation, characterization and
application of Pt–Ru–Sn/C trimetallic electrocatalysts for ethanol oxidation in direct fuel cell.
Int J Hydrogen Energy 36:11034–11042
37. Moghaddam RB, Pickup PG (2012) Support effects on the oxidation of ethanol at Pt
nanoparticles. Electrochim Acta 65:210–215
38. Balakrishnan K, Schwank J (1992) FTIR study of bimetallic Pt-Sn/Al2O3 catalysts. J Catal
138:491–499
39. Liu ZL, Ling XY, Su XD, Lee JY, Gan LM (2005) Preparation and characterization of Pt/C
and Pt-Ru/C electrocatalysts for direct ethanol fuel cells. J Power Sources 149:1–7
40. Ribadeneira E, Hoyos BA (2008) Evaluation of Pt–Ru–Ni and Pt–Sn–Ni catalysts as anodes
in direct ethanol fuel cells. J Power Sources 180:238–242
41. Bonesi A, Garaventa G, Triaca WE, Castro Luna AM (2008) Synthesis and characterization
of new electrocatalysts for ethanol oxidation. Int J Hydrogen Energy 33:3499–3501
42. De Souza JPI, Queiroz SL, Bergamaski K, Gonzalez ER, Nart FC (2002) Electro-oxidation of
ethanol on Pt, Rh, and PtRh electrodes. A study using DEMS and in-situ FTIR techniques. J
Phys Chem B 106:9825–9830
43. Sen Gupta S, Datta J (2006) A comparative study on ethanol oxidation behavior at Pt and
PtRh electrodeposits. J Electroanal Chem 594:65–72
44. Kowal A, Li M, Shao M, Sasaki K, Vukmirovic MB, Zhang J, Marinkovic NS, Liu P, Frenkel
AI, Adzic RR (2009) Ternary Pt/Rh/SnO2 electrocatalysts for oxidizing ethanol to CO2. Nat
Mater 8:325–330
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 469

45. Kowal A, Gojkovic SL, Lee KS, Olszewski P, Sung YE (2009) Synthesis, characterization
and electrocatalytic activity for ethanol oxidation of carbon supported Pt, Pt-Rh, Pt-SnO2 and
Pt-Rh-SnO2 nanoclusters. Electrochem Commun 11:724–727
46. Maillard F, Peyrelade E, Soldo-Olivier Y, Chatenet M, Chainet E, Faure R (2007) Is carbon-
supported Pt-WOx composite a CO-tolerant material? Electrochim Acta 52:1958–1967
47. Nores-Pondal FJ, Vilella IMJ, Troiani H, Granada M, De Miguel SR, Scelza OA, Corti HR
(2009) Catalytic activity vs. size correlation in platinum catalysts of PEM fuel cells prepared
on carbon black by different methods. Int J Hydrogen Energy 34:8193–8203
48. Liu ZL, Ling XY, Su XD, Lee JY (2004) Carbon-supported Pt and PtRu nanoparticles as
catalysts for a direct methanol fuel cell. J Phys Chem B 108:8234–8240
49. Lim D-H, Choi D-H, Lee W-D, Lee H-I (2009) A new synthesis of a highly dispersed and CO
tolerant PtSn/C electrocatalyst for low-temperature fuel cell; its electrocatalytic activity and
long-term durability. Appl Catal B Environ 89:484–493
50. Spinacé EV, Farias LA, Linardi M, Neto AO (2008) Preparation of PtSn/C and PtSnNi/C
electrocatalysts using the alcohol-reduction process. Mater Lett 62:2099–2102
51. Vidal-Iglesias FJ, Al-Akl A, Watson DJ, Attard GA (2006) A new method for the preparation
of PtPd alloy single crystal surfaces. Electrochem Commun 8:1147–1150
52. Colmati F, Antolini E, Gonzalez ER (2007) Ethanol oxidation on a carbon-supported
Pt75Sn25 electrocatalyst prepared by reduction with formic acid: effect of thermal treatment.
Appl Catal B Environ 73:106–115
53. Zhao J, Chen W, Zheng Y, Li X, Xu Z (2006) Microwave polyol synthesis of Pt/C catalysts
with size-controlled Pt particles for methanol electrocatalytic oxidation. J Mater Sci
41:5514–5518
54. Cushing BL, Kolesnichenko VL, O’Connor CJ (2004) Recent advances in the liquid-phase
syntheses of inorganic nanoparticles. Chem Rev 104:3893–3946
55. Coutanceau C, Brimaud S, Lamy C, Léger J-M, Dubau L, Rousseau S, Vigier F (2008)
Review of different methods for developing nanoelectrocatalysts for the oxidation of organic
compounds. Electrochim Acta 53:6865–6880
56. Wang Z-B, Yin G-P, Zhang J, Sun Y-C, Shi P-F (2006) Investigation of ethanol
electrooxidation on a Pt-Ru-Ni/C catalyst for a direct ethanol fuel cell. J Power Sources
160:37–43
57. Toshima N, Yonezawa T (1998) Bimetallic nanoparticles-novel materials for chemical and
physical applications. New J Chem 22:1179–1201
58. Antolini E, Colmati F, Gonzalez ER (2009) Ethanol oxidation on carbon supported
(PtSn)alloy/SnO2 and (PtSnPd)alloy/SnO2 catalysts with a fixed Pt/SnO2 atomic ratio: effect
of the alloy phase characteristics. J Power Sources 193:555–561
59. Franco EG, Neto AO, Linardi M, Aricó E (2002) Synthesis of electrocatalysts by the
B€onnemann method for the oxidation of methanol and the mixture H2/CO in a proton
exchange membrane fuel cell. J Braz Chem Soc 13:516–521
60. Jeon MK, Zhang Y, McGinn PJ (2010) A comparative study of PtCo, PtCr, and PtCoCr
catalysts for oxygen electro-reduction reaction. Electrochim Acta 55:5318–5325
61. Pechini MP (1967) USA Patent 3,330,697
62. Simoes FC, dos Anjos DM, Vigier F, Léger J-M, Hahn F, Coutanceau C, Gonzalez ER,
Tremiliosi-Filho G, de Andrade AR, Olivi P, Kokoh KB (2007) Electroactivity of tin
modified platinum electrodes for ethanol electrooxidation. J Power Sources 167:1–10
63. Galceran M, Pujol MC, Aguiló M, Dı́az F (2007) Sol-gel modified Pechini method for
obtaining nanocrystalline KRE(WO4)2 (RE ¼ Gd and Yb). J Sol-Gel Sci Technol 42:79–88
64. Laberty-Robert C, Ansart F, Deloget C, Gaudon M, Rousset A (2001) Powder synthesis of
nanocrystalline ZrO2–8%Y2O3 via a polymerization route. Mater Res Bull 36:2083–2101
65. Kwon SW, Park SB, Seo G, Hwang ST (1998) Preparation of lithium aluminate via polymeric
precursor routes. J Nucl Mater 257:172–179
470 T.S. Almeida et al.

66. Li X, Chen WX, Zhao J, Xing W, Xu ZD (2005) Microwave polyol synthesis of Pt/CNTs
catalysts: effects of pH on particle size and electrocatalytic activity for methanol electroox-
idization. Carbon 43:2168–2174
67. Kadirgan F, Beyhan S, Atilan T (2009) Preparation and characterization of nano-sized Pt-Pd/
C catalysts and comparison of their electro-activity toward methanol and ethanol oxidation.
Int J Hydrogen Energy 34:4312–4320
68. Tsuji M, Hashimoto M, Nishizawa Y, Kubokawa M, Tsuji T (2005) Microwave-assisted
synthesis of metallic nanostructures in solution. Chem Eur J 11:440–452
69. Bock C, Paquet C, Couillard M, Botton GA, MacDougall BR (2004) Size-selected synthesis
of PtRu nano-catalysts: reaction and size control mechanism. J Am Chem Soc
126:8028–8037
70. Yang J, Deivaraj TC, Too HP, Lee JY (2004) Acetate stabilization of metal nanoparticles and
its role in the preparation of metal nanoparticles in ethylene glycol. Langmuir 20:4241–4245
71. Yu WY, Tu WX, Liu HF (1999) Synthesis of nanoscale platinum colloids by microwave
dielectric heating. Langmuir 15:6–9
72. Tsujino T, Ohigashi S, Sugiyama S, Kawashiro K, Hayashi H (1992) Oxidation of propylene
glycol and lactic acid to pyruvic acid in aqueous phase catalyzed by lead-modified palladium-
on-carbon and related systems. J Mol Catal 71:25–35
73. Pinxt H, Kuster BFM, Marin GB (2000) Promoter effects in the Pt-catalysed oxidation of
propylene glycol. Appl Catal A Gen 191:45–54
74. Zhao J, Wang P, Chen W, Liu R, Li X, Nie Q (2006) Microwave synthesis and characteri-
zation of acetate-stabilized Pt nanoparticles supported on carbon for methanol electro-
oxidation. J Power Sources 160:563–569
75. Neto AO, Verjulio-Silva RWR, Linardi M, Spinacé EV (2009) Preparation of PtRu/C
electrocatalysts using citric acid as reducing agent and OH ions as stabilizing agent for
direct alcohol fuel cell (DAFC). Int J Electrochem Sci 4:954–961
76. Tu WX, Liu HF (2000) Rapid synthesis of nanoscale colloidal metal clusters by microwave
irradiation. J Mater Chem 10:2207–2211
77. Viau G, Brayner R, Poul L, Chakroune N, Lacaze E, Fievet-Vincent F, Fievet F (2003)
Ruthenium nanoparticles: size, shape, and self-assemblies. Chem Mater 15:486–494
78. Ozkar S, Finke RG (2002) Nanocluster formation and stabilization fundamental studies:
ranking commonly employed anionic stabilizers via the development, then application, of
five comparative criteria. J Am Chem Soc 124:5796–5810
79. Spinacé EV, Linardi M, Neto AO (2005) Co-catalytic effect of nickel in the electro-oxidation
of ethanol on binary Pt–Sn electrocatalysts. Electrochem Commun 7:365–369
80. Colmati F, Antolini E, Gonzalez ER (2008) Preparation, structural characterization and
activity for ethanol oxidation of carbon supported ternary Pt–Sn–Rh catalysts. J Alloys
Compd 456:264–270
81. Parreira LS, da Silva JCM, D’Villa -Silva M, Simões FC, Garcia S, Gaubeur I, Cordeiro
MAL, Leite ER, dosSantos MC (2013) PtSnNi/C nanoparticle electrocatalysts for the ethanol
oxidation reaction: Ni stability study. Electrochim Acta 96:243–252
82. Almeida TS, Palma LM, Leonello PH, Morais C, Kokoh KB, De Andrade AR (2012) An
optimization study of PtSn/C catalysts applied to direct ethanol fuel cell: effect of the
preparation method on the electrocatalytic activity of the catalysts. J Power Sources
215:53–62
83. Stonehart P, Ross PN Jr (1976) The use of porous electrodes to obtain kinetic rate constants
for rapid reactions and adsorption isotherms of poisons. Electrochim Acta 21:441–445
84. Schmidt TJ, Gasteiger HA, Stäb GD, Urban PM, Kolb DM, Behm RJ (1998) Characterization
of high surface area electrocatalysts using a rotating disk electrode configuration. J
Electrochem Soc 145:2354–2358
85. Vesovic V, Anastasijevic N, Adzic RR (1987) Rotating disk electrode: a re-examination of
some kinetic criteria with a special reference to oxygen reduction. J Electroanal Chem
Interfacial Electrochem 218:53–63
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 471

86. Hsueh KL, Chin DT, Srinivasan S (1983) Electrode kinetics of oxygen reduction: a theoret-
ical and experimental analysis of the rotating ring-disc electrode method. J Electroanal Chem
Interfacial Electrochem 153:79–95
87. Jakobs RCM, Janssen LJJ, Barendrecht E (1985) Oxygen reduction at polypyrrole elec-
trodes—I. Theory and evaluation of the rrde experiments. Electrochim Acta 30:1085–1091
88. Anastasijević NA, Vesović V, Adžić RR (1987) Determination of the kinetic parameters of
the oxygen reduction reaction using the rotating ring-disk electrode: Part I. Theory. J
Electroanal Chem Interfacial Electrochem 229:305–316
89. Anastasijević NA, Vesović V, Adžić RR (1987) Determination of the kinetic parameters of
the oxygen reduction reaction using the rotating ring-disk electrode: Part II. Applications. J
Electroanal Chem Interfacial Electrochem 229:317–325
90. Zečević S, Dražić DM, Gojković S (1989) Oxygen reduction on iron: Part III. An analysis of
the rotating disk-ring electrode measurements in near neutral solutions. J Electroanal Chem
Interfacial Electrochem 265:179–193
91. Maruyama J, Inaba M, Ogumi Z (1998) Rotating ring-disk electrode study on the cathodic
oxygen reduction at Nafion®-coated gold electrodes. J Electroanal Chem 458:175–182
92. Anastasijević NA, Dimitrijević ZM, Adžić RR (1986) Oxygen reduction on a ruthenium
electrode in acid electrolytes. Electrochim Acta 31:1125–1130
93. Jiang T, Brisard GM (2007) Determination of the kinetic parameters of oxygen reduction on
copper using a rotating ring single crystal disk assembly (RRDCu(hkl)E). Electrochim Acta
52:4487–4496
94. Damjanovic A, Genshaw MA, Bockris JOM (1967) The mechanism of oxygen reduction at
platinum in alkaline solutions with special reference to H2O2. J Electrochem Soc
114:1107–1112
95. Appel M, Appleby AJ (1978) A ring-disk electrode study of the reduction of oxygen on active
carbon in alkaline solution. Electrochim Acta 23:1243–1246
96. Markovic N (2002) Surface science studies of model fuel cell electrocatalysts. Surf Sci Rep
45:117–229
97. Yeager E (1984) Electrocatalysts for O2 reduction. Electrochim Acta 29:1527–1537
98. Antoine O, Durand R (2000) RRDE study of oxygen reduction on Pt nanoparticles inside
Nafion®: H2O2 production in PEMFC cathode conditions. J Appl Electrochem 30:839–844
99. Ke K, Hatanaka T, Morimoto Y (2011) Reconsideration of the quantitative characterization
of the reaction intermediate on electrocatalysts by a rotating ring-disk electrode: The intrinsic
yield of H2O2 on Pt/C. Electrochim Acta 56:2098–2104
100. Damjanovic A, Genshaw MA, Bockris JOM (1966) Distinction between intermediates
produced in main and side electrodic reactions. J Chem Phys 45:4057–4059
101. Wroblowa HS, Yen Chi P, Razumney G (1976) Electroreduction of oxygen: a new mecha-
nistic criterion. J Electroanal Chem Interfacial Electrochem 69:195–201
102. Appleby AJ, Savy M (1978) Kinetics of oxygen reduction reactions involving catalytic
decomposition of hydrogen peroxide: application to porous and rotating ring-disk electrodes.
J Electroanal Chem Interfacial Electrochem 92:15–30
103. Zurilla RW, Sen RK, Yeager E (1978) The kinetics of the oxygen reduction reaction on gold
in alkaline solution. J Electrochem Soc 125:1103–1109
104. Sánchez-Sánchez CM, Bard AJ (2009) Hydrogen peroxide production in the oxygen reduc-
tion reaction at different electrocatalysts as quantified by scanning electrochemical micros-
copy. Anal Chem 81:8094–8100
105. Olson TS, Pylypenko S, Fulghum JE, Atanassov P (2010) Bifunctional oxygen reduction
reaction mechanism on non-platinum catalysts derived from pyrolyzed porphyrins. J
Electrochem Soc 157:B54–B63
106. Nørskov JK, Rossmeisl J, Logadottir A, Lindqvist L, Kitchin JR, Bligaard T, Jónsson H
(2004) Origin of the overpotential for oxygen reduction at a fuel-cell cathode. J Phys Chem B
108:17886–17892
107. Guo S, Zhang S, Sun S (2013) Tuning nanoparticle catalysis for the oxygen reduction
reaction. Angew Chem Int Ed Engl 52:8526–8544
472 T.S. Almeida et al.

108. Katsounaros I, Schneider WB, Meier JC, Benedikt U, Biedermann PU, Auer AA, Mayrhofer
KJ (2012) Hydrogen peroxide electrochemistry on platinum: towards understanding the
oxygen reduction reaction mechanism. Phys Chem Chem Phys 14:7384–7391
109. Katsounaros I, Cherevko S, Zeradjanin AR, Mayrhofer KJ (2014) Oxygen electrochemistry
as a cornerstone for sustainable energy conversion. Angew Chem Int Ed Engl 53:102–121
110. Jiang R, Dong S (1990) Rotating ring disk electrode (RRDE) theory dealing with non
stationary electrocatalysis: study of the electrocatalytic reduction of dioxygen at cobalt
protoporphrin modified electrode. J Phys Chem 94:7471–7476
111. Demarconnay L, Coutanceau C, Léger J-M (2004) Electroreduction of dioxygen (ORR) in
alkaline medium on Ag/C and Pt/C nanostructured catalysts—effect of the presence of
methanol. Electrochim Acta 49:4513–4521
112. Paulus UA, Schmidt TJ, Gasteiger HA, Behm RJ (2001) Oxygen reduction on a high-surface
area Pt/Vulcan carbon catalyst: a thin-film rotating ring-disk electrode study. J Electroanal
Chem 495:134–145
113. Kawabata A (1996) Electronic properties of metallic fine particles. Surf Rev Lett 03:9–12
114. Volokitin Y, Sinzig J, de Jongh LJ, Schmid G, Vargaftik MN, Moiseevi II (1996) Quantum-
size effects in the thermodynamic properties of metallic nanoparticles. Nature 384:621–623
115. Halperin WP (1986) Quantum size effects in metal particles. Rev Mod Phys 58:533–606
116. Jalan V, Taylor EJ (1983) Importance of interatomic spacing in catalytic reduction of oxygen
in phosphoric acid. J Electrochem Soc 130:2299–2302
117. Toda T, Igarashi H, Uchida H, Watanabe M (1999) Enhancement of the electroreduction of
oxygen on Pt alloys with Fe, Ni, and Co. J Electrochem Soc 146:3750–3756
118. Lim D-H, Wilcox J (2012) Mechanisms of the oxygen reduction reaction on defective
graphene-supported Pt nanoparticles from first-principles. J Phys Chem C 116:3653–3660
119. Rabis A, Rodriguez P, Schmidt TJ (2012) Electrocatalysis for polymer electrolyte fuel cells:
recent achievements and future challenges. ACS Catal 2:864–890
120. Angerstein-Kozlowska H, Conway BE, Sharp WBA (1973) The real condition of electro-
chemically oxidized platinum surfaces: Part I. Resolution of component processes. J
Electroanal Chem Interfacial Electrochem 43:9–36
121. Pozio A, De Francesco M, Cemmi A, Cardellini F, Giorgi L (2002) Comparison of high
surface Pt/C catalysts by cyclic voltammetry. J Power Sources 105:13–19
122. Grolleau C, Coutanceau C, Pierre F, Léger J-M (2010) Optimization of a surfactant free
polyol method for the synthesis of platinum–cobalt electrocatalysts using Taguchi design of
experiments. J Power Sources 195:1569–1576
123. Chen W, Kim J, Sun S, Chen S (2008) Electrocatalytic reduction of oxygen by FePt alloy
nanoparticles. J Phys Chem C 112:3891–3898
124. Grolleau C, Coutanceau C, Pierre F, Léger J-M (2008) Effect of potential cycling on structure
and activity of Pt nanoparticles dispersed on different carbon supports. Electrochim Acta
53:7157–7165
125. Bard AJ, Faulkner LR (2001) Electrochemical methods: fundamentals and applications.
Wiley, New York
126. Choi SI, Lee SU, Kim WY, Choi R, Hong K, Nam KM, Han SW, Park JT (2012)
Composition-controlled PtCo alloy nanocubes with tuned electrocatalytic activity for oxygen
reduction. ACS Appl Mater Interfaces 4:6228–6234
127. Lima FHB, Ticianelli EA (2004) Oxygen electrocatalysis on ultra-thin porous coating
rotating ring/disk platinum and platinum–cobalt electrodes in alkaline media. Electrochim
Acta 49:4091–4099
128. Coutanceau C, Croissant MJ, Napporn T, Lamy C (2000) Electrocatalytic reduction of
dioxygen at platinum particles dispersed in a polyaniline film. Electrochim Acta 46:579–588
129. Wang JX, Markovic NM, Adzic RR (2004) Kinetic analysis of oxygen reduction on Pt(111)
in acid solutions: intrinsic kinetic parameters and anion adsorption effects. J Phys Chem B
108:4127–4133
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 473

130. Wang JX, Brankovic SR, Zhu Y, Hanson JC, Adz̆ić RR (2003) Kinetic characterization of
PtRu fuel cell anode catalysts made by spontaneous Pt deposition on Ru nanoparticles. J
Electrochem Soc 150:A1108–A1117
131. Jahn D, Vielstich W (1962) Rates of electrode processes by the rotating disk method. J
Electrochem Soc 109:849–852
132. Lebègue E, Baranton S, Coutanceau C (2011) Polyol synthesis of nanosized Pt/C
electrocatalysts assisted by pulse microwave activation. J Power Sources 196:920–927
133. Prakash J, Tryk DA, Yeager EB (1999) Kinetic investigations of oxygen reduction and
evolution reactions on lead ruthenate catalysts. J Electrochem Soc 146:4145–4151
134. Murthi VS, Urian RC, Mukerjee S (2004) Oxygen reduction kinetics in low and medium
temperature acid environment: correlation of water activation and surface properties in
supported Pt and Pt alloy electrocatalysts. J Phys Chem B 108:11011–11023
135. Bakir ÇC, Şahin N, Polat R, Dursun Z (2011) Electrocatalytic reduction of oxygen on
bimetallic copper–gold nanoparticles–multiwalled carbon nanotube modified glassy carbon
electrode in alkaline solution. J Electroanal Chem 662:275–280
136. Diabaté D, Napporn TW, Servat K, Habrioux A, Arrii-Clacens S, Trokourey A, Kokoh KB
(2013) Kinetic study of oxygen reduction reaction on carbon supported Pd-based
nanomaterials in alkaline medium. J Electrochem Soc 160:H302–H308
137. Mayrhofer KJJ, Strmcnik D, Blizanac BB, Stamenkovic V, Arenz M, Markovic NM (2008)
Measurement of oxygen reduction activities via the rotating disc electrode method: from Pt
model surfaces to carbon-supported high surface area catalysts. Electrochim Acta
53:3181–3188
138. Gojković SL, Gupta S, Savinell RF (1999) Heat-treated iron(III) tetramethoxyphenyl por-
phyrin chloride supported on high-area carbon as an electrocatalyst for oxygen reduction:
Part III. Detection of hydrogen-peroxide during oxygen reduction. Electrochim Acta
45:889–897
139. Maruyama J, Inaba M, Morita T, Ogumi Z (2001) Effects of the molecular structure of
fluorinated additives on the kinetics of cathodic oxygen reduction. J Electroanal Chem
504:208–216
140. Vogel W, Lundquist L, Ross P, Stonehart P (1975) Reaction pathways and poisons—II: The
rate controlling step for electrochemical oxidation of hydrogen on Pt in acid and poisoning of
the reaction by CO. Electrochim Acta 20:79–93
141. Iwasita T, Pastor E (1994) A dems and FTir spectroscopic investigation of adsorbed ethanol
on polycrystalline platinum. Electrochim Acta 39:531–537
142. Garcı́a G, Tsiouvaras N, Pastor E, Pe~ na MA, Fierro JLG, Martı́nez-Huerta MV (2012)
Ethanol oxidation on PtRuMo/C catalysts: in situ FTIR spectroscopy and DEMS studies.
Int J Hydrogen Energy 37:7131–7140
143. Schmiemann U, M€ uller U, Baltruschat H (1995) The influence of the surface structure on the
adsorption of ethene, ethanol and cyclohexene as studied by DEMS. Electrochim Acta
40:99–107
144. Ianniello R, Schmidt VM, Rodrı́guez JL, Pastor E (1999) Electrochemical surface reactions
of intermediates formed in the oxidative ethanol adsorption on porous Pt and PtRu. J
Electroanal Chem 471:167–179
145. Wang H, Jusys Z, Behm RJ (2006) Ethanol electro-oxidation on carbon-supported Pt, PtRu
and Pt3Sn catalysts: a quantitative DEMS study. J Power Sources 154:351–359
146. Sun S, Halseid MC, Heinen M, Jusys Z, Behm RJ (2009) Ethanol electrooxidation on a
carbon-supported Pt catalyst at elevated temperature and pressure: a high-temperature/high-
pressure DEMS study. J Power Sources 190:2–13
147. Pastor E, Iwasita T (1994) D/H exchange of ethanol at platinum electrodes. Electrochim Acta
39:547–551
148. Camara GA, de Lima RB, Iwasita T (2005) The influence of PtRu atomic composition on the
yields of ethanol oxidation: a study by in situ FTIR spectroscopy. J Electroanal Chem
585:128–131
474 T.S. Almeida et al.

149. Rousseau S, Coutanceau C, Lamy C, Léger J-M (2006) Direct ethanol fuel cell (DEFC):
electrical performances and reaction products distribution under operating conditions with
different platinum-based anodes. J Power Sources 158:18–24
150. Palma LM, Almeida TS, de Andrade AR (2012) Development of plurimetallic
electrocatalysts prepared by decomposition of polymeric precursors for EtOH/O2 fuel cell.
J Braz Chem Soc 23:555–564
151. Eneau-Innocent B, Pasquier D, Ropital F, Léger J-M, Kokoh KB (2010) Electroreduction of
carbon dioxide at a lead electrode in propylene carbonate: a spectroscopic study. Appl Catal
B Environ 98:65–71
152. Bewick A, Kunimatsu K, Pons BS, Russell JW (1984) Electrochemically modulated infrared
spectroscopy (EMIRS): experimental details. J Electroanal Chem Interfacial Electrochem
160:47–61
153. Xia XH, Liess HD, Iwasita T (1997) Early stages in the oxidation of ethanol at low index
single crystal platinum electrodes. J Electroanal Chem 437:233–240
154. Rodes A, Pastor E, Iwasita T (1994) An FTIR study on the adsorption of acetate at the basal
planes of platinum single-crystal electrodes. J Electroanal Chem 376:109–118
155. Batista EA, Malpass GRP, Motheo AJ, Iwasita T (2004) New mechanistic aspects of
methanol oxidation. J Electroanal Chem 571:273–282
156. Vigier F, Coutanceau C, Hahn F, Belgsir EM, Lamy C (2004) On the mechanism of ethanol
electro-oxidation on Pt and PtSn catalysts: electrochemical and in situ IR reflectance spec-
troscopy studies. J Electroanal Chem 563:81–89
157. Corrigan DS, Krauskopf EK, Rice LM, Wieckowski A, Weaver MJ (1988) Adsorption of
acetic acid at platinum and gold electrodes: a combined infrared spectroscopic and radio-
tracer study. J Phys Chem 92:1596–1601
158. Leung LWH, Weaver MJ (1988) Real-time FTIR spectroscopy as a quantitative kinetic probe
of competing electrooxidation pathways for small organic molecules. J Phys Chem
92:4019–4022
159. Iwasita T, Rasch B, Cattaneo E, Vielstich W (1989) A sniftirs study of ethanol oxidation on
platinum. Electrochim Acta 34:1073–1079
160. Beden B, Lamy C, Bewick A, Kunimatsu K (1981) Electrosorption of methanol on a
platinum electrode. IR spectroscopic evidence for adsorbed CO species. J Electroanal
Chem 121:343–347
161. Perez JM, Beden B, Hahn F, Aldaz A, Lamy C (1989) “In situ” infrared reflectance
spectroscopic study of the early stages of ethanol adsorption at a platinum electrode in acid
medium. J Electroanal Chem Interfacial Electrochem 262:251–261
162. Léger J-M, Beden B, Lamy C, Bilmes S (1984) Carbon monoxide electrosorption on low
index platinum single crystal electrodes. J Electroanal Chem Interfacial Electrochem
170:305–317
163. Heysiattalab S, Shakeri M, Safari M, Keikha MM (2011) Investigation of key parameters
influence on performance of direct ethanol fuel cell (DEFC). J Ind Eng Chem 17:727–729
164. Zhu M, Sun G, Xin Q (2009) Effect of alloying degree in PtSn catalyst on the catalytic
behavior for ethanol electro-oxidation. Electrochim Acta 54:1511–1518
165. Crisafulli R, Antoniassi RM, Neto AO, Spinacé EV (2014) Acid-treated PtSn/C and PtSnCu/
C electrocatalysts for ethanol electro-oxidation. Int J Hydrogen Energy 39:5671–5677
166. Zhu M, Sun G, Li H, Cao L, Xin Q (2008) Effect of the Sn(II)/Sn(IV) redox couple on the
activity of PtSn/C for ethanol electro-oxidation. Chin J Catal 29:765–770
167. Jiang L, Sun G, Sun S, Liu J, Tang S, Li H, Zhou B, Xin Q (2005) Structure and chemical
composition of supported Pt–Sn electrocatalysts for ethanol oxidation. Electrochim Acta
50:5384–5389
168. Tsiakaras PE (2007) PtM/C (M ¼ Sn, Ru, Pd, W) based anode direct ethanol–PEMFCs:
structural characteristics and cell performance. J Power Sources 171:107–112
11 Direct Ethanol Fuel Cell on Carbon Supported Pt Based Nanocatalysts 475

169. Ribeiro J, dos Anjos DM, Kokoh KB, Coutanceau C, Léger J-M, Olivi P, de Andrade AR,
Tremiliosi-Filho G (2007) Carbon-supported ternary PtSnIr catalysts for direct ethanol fuel
cell. Electrochim Acta 52:6997–7006
170. Almeida TS, Palma LM, Morais C, Kokoh KB, De Andrade AR (2013) Effect of adding a
third metal to carbon-supported PtSn-based nanocatalysts for direct ethanol fuel cell in acidic
medium. J Electrochem Soc 160:F965–F971
171. Beyhan S, Coutanceau C, Léger J-M, Napporn TW, Kadırgan F (2013) Promising anode
candidates for direct ethanol fuel cell: carbon supported PtSn-based trimetallic catalysts
prepared by B€onnemann method. Int J Hydrogen Energy 38:6830–6841
172. Beyhan S, Léger J-M, Kadırgan F (2013) Pronounced synergetic effect of the nano-sized
PtSnNi/C catalyst for ethanol oxidation in direct ethanol fuel cell. Appl Catal B Environ
130–131:305–313
Chapter 12
Direct Alcohol Fuel Cells: Nanostructured
Materials for the Electrooxidation
of Alcohols in Alkaline Media

Hamish Andrew Miller, Francesco Vizza, and Alessandro Lavacchi

12.1 Introduction: The Potential of Biomass Derived


Alcohols in Direct Fuel Cells

Renewable biomass derived alcohols are attractive fuels for application in direct
fuel cells due to their low toxicity and their potential as part of carbon neutral
energy transformation processes [1, 2]. Among these ethanol has been the focus of
most interest, due to the possibility of production from the fermentation of biomass
or the steam explosion of lignocelluloses with reasonably low energy cost
[3, 4]. Alcohols with a higher molecular weight are also attracting attention as
they offer high energy densities, low vapor pressure and some of them can be
obtained from renewable resources (Table 12.1). Included in this group are ethylene
glycol [5, 6] (EG) and glycerol [7, 8] (G). Ethylene glycol has a volumetric energy
density of 5.9 kWh L1 and can be produced by the heterogeneous hydrogenation
of cellulose derivatives [9, 10]. Glycerol has a volumetric energy density of
6.3 kWh L1, and is a by-product of biodiesel production and, as such, is inexpensive
(0.3 US$ kg1) and readily available (2.4 million tonnes produced per year) [11].
Biomass production is a virtuous cycle in terms of greenhouse emissions. Indeed
CO2 emitted during the production process of ethanol and the CO2 eventually
released by ethanol oxidation in direct ethanol fuel cells (DEFCs) can be reverted
into biomasses by the plants from which the fuel is produced. Hence, ethanol is
potentially a sustainable energy source. On the other hand we should consider that a
life cycle analysis of bioethanol renders an energy payback ratio of the fuel in the
range of 1 or lower (i.e. a negative return on energy invested) if the production is
from corn for example [12]. Hence, it is not at all convenient to use corn for
producing ethanol. Alternatively, sugar-rich plants such as rapeseeds and sugarcane

H.A. Miller (*) • F. Vizza • A. Lavacchi


ICCOM-CNR, Via Madonna del Piano 10, Sesto Fiorentino (Firenze) 50019, Italy
e-mail: [email protected]

© Springer International Publishing Switzerland 2016 477


K.I. Ozoemena, S. Chen (eds.), Nanomaterials for Fuel Cell Catalysis,
Nanostructure Science and Technology, DOI 10.1007/978-3-319-29930-3_12
478 H.A. Miller et al.

Table 12.1 Physical properties of various alcohol fuels


Specific
energy density Density at Volumetric energy
Fuel (kWh kg1) 20  C g cm3 density (kWh dm3) Ecell (V) n
Methanol 6.1 0.79 4.8 1.21 6e
Ethanol 8.0 0.79 6.3 1.15 12e
Ethylene glycol 5.2 1.11 5.8 1.22 10e
Glycerol 5.0 1.26 6.4 1.09 14e
Propan-1-ol 9.1 0.81 7.4 1.13 18e-
Propan-2-ol 9.0 0.79 7.1 1.12 18e

can produce an energy return ratio for derived bioethanol that ranges from 4 to
8. More importantly the production of bioethanol from unused nonagricultural
lands from spontaneously grown plants such as arundo donax also known as
giant cane, has also been reported [13]. In such cases it should be possible to get
an energy payback ratio even larger than ten. Other alcohols such as glycerol and
ethylene glycol can also be derived by a variety of processes from biomass. These
two alcohols have received much attention recently as they can potentially be
oxidized in direct alcohol fuel cells to render not only electrical energy but also
valuable chemicals from the formation of partially oxidized intermediates. This will
form an important part of the discussion in this chapter.
Anion exchange membrane (AEM) direct alcohol fuel cells (DAFCs) represent
devices in which the exploitation of these bio-alcohols can be achieved [2, 14–17].
In alkaline media the kinetics of the oxygen reduction reaction (ORR) are greatly
enhanced due to improved charge transfer [1]. The high pH also offers a much less
corrosive environment and hence non-precious metal catalysts can be employed
especially at the cathode [18–20]. Polarization losses due to alcohol crossover from
the anode to the cathode are also alleviated by using these catalysts, which are highly
selective for the ORR.
To date for alcohols like EtOH, G and EG, complete oxidation to CO2 is yet to be
achieved and remains a very challenging task [21, 22]. To the best of our knowl-
edge, no known transition metal based electrocatalyst promotes complete oxidation
in either acidic or alkaline media. The major reaction pathways involve the forma-
tion of various carboxylic acids or carboxylates while C-C scission and formation
of CO2 or carbonate is only a minor route. Indeed, the C-C bond cleavage of EtOH
has only been observed on Pd at pHs less than 13 [23]. For EG complete oxidation is
also difficult with the major pathways involving the formation of oxalate and
glycolate [8, 24]. In the case of G the electrooxidation mechanism is much more
complex and significant amounts of carbonate can be formed. On nanostructured Pd
electrocatalysts this occurs through the oxidation of the glycolate intermediate
[8]. Partial and selective oxidation of poly-alcohols to important fine chemicals is
potentially a route to the realization of the so-called bio-refinery. One of the main
goals in such processes is increasing the selectivity towards a given partial oxida-
tion product. Tartronate, dihydroxyacetone, hydroxyperuvate, mexoxylate and
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 479

lactate are all value added products of industrial interest. In order to realize such
processes the anode catalyst employed in the fuel cell must guarantee two aspects:
(i) selectivity in the partial oxidation and (ii) high activity. The recent development
of molecular anode electrocatalysts employed in Organo Metallic Fuel Cells
(OMFCs) with tunable single catalytic sites has enabled this to be achieved for a
number of alcohols of industrial interest. In fact for the first time the oxidation of
substrates such as 1,3-propandiol and 1,4-butandiol have been exploited in fuel
cells.
The challenge facing researchers in this field is to use nanotechnology to tune
anode materials to reach such performance targets. In this chapter, we discuss the
most recent strategies in nanostructuring of anode materials for DAFCs. After
looking at the mechanisms of alcohol electrooxidation in alkaline media we exam-
ine the most recently developed materials on the basis of the active metal present.
This is followed by examples of nanostructured metal catalysts with shape and
structure control where the formation of high index faceted nanoparticles has led to
activity enhancements. Next we take a look at a molecular approach to alcohol
electrooxidation in particular the development of organometallic complexes as
anode catalysts which show remarkable selectivity in the formation of partially
oxidized valuable chemicals (e.g. lactate from 1,2-propandiol). The emphasis in the
discussion here is to highlight materials that have been employed in DAFCs with
anion exchange membranes, thus showing the state of the art in alkaline DAFC
performance.

12.2 Alcohol Electrooxidation in Alkaline Media

In this overview of alcohol electrooxidation we have excluded methanol which is


more than sufficiently reviewed elsewhere. Under acidic conditions platinum-
ruthenium alloys exhibit the lowest overpotentials and resistance to poisoning due
to the synergistic effect of the two metals. Platinum is extremely active for the
dissociative chemisorption of methanol, while the oxidation of carbonaceous
adsorbates (e.g. CO) is favored by the presence of ruthenium oxides [25]. Indeed,
carbon supported PtRu alloys are presently the state of the art for methanol
electrooxidation in acidic conditions like those found in DMFCs. We will discuss
in detail higher alcohols of which selected physical properties are detailed in
Table 12.1.

12.2.1 Ethanol Electrooxidation

Ethanol is the most obvious candidate to replace methanol in DAFCs. Scheme 12.1
shows the variety of oxidation products attainable in principle through
electrooxidation. Complete oxidation to CO2 renders 12e but as mentioned
480 H.A. Miller et al.

Scheme 12.1 Ethanol


electrooxidation reaction
products

previously is difficult to obtain as this requires a breakage of the C-C bond. Usually,
partial oxidation is obtained with the major products being acetaldehyde and acetic
acid which deliver 2e and 4e respectively [26].
Detailed studies of the reaction products of ethanol electrooxidation in acidic
media on Pt electrodes by chromatographic techniques and Differential Electro-
chemical Mass Spectroscopy (DEMS) has allowed a detailed description of the
complex mechanisms involved. Reaction (Eq. 12.1) occurs at higher electrode
potentials (E < 0.6 V) whereas reaction (Eq. 12.2) is favored at lower potentials
(E < 0.6 V).

CH 3 CH 2 OH þ H2 O ! CH 3 COOH þ 4Hþ þ 4e ð12:1Þ

CH 3 CH 2 OH ! CH 3 CHO þ 2H þ þ 2e ð12:2Þ

At intermediate potentials (0.6 V  E  0.8 V), the dissociative adsorption of water


occurs (Eq. 12.3):

Pt þ H 2 O ! Pt  OH ads þ Hþ þ e ð12:3Þ

The oxidation of adsorbed CH3CHO may then produce acetic acid as follows
(Eq. 12.4):

CH 3 CHO þ Pt ! Pt  ðCH 3 CHOÞads and


ð12:4Þ
Pt  ðCH 3 CHOÞads þ Pt  ðOH Þads ! CH 3 COOH þ H þ þ e

It is difficult to obtain further oxidation to CO2 at low temperatures but both CO


(as poisoning species), CO2 and methane have been observed as reaction products.
The mechanisms are as follows:

Pt þ PtðCO  CH 3 Þ ! Pt  ðCOÞads þ Pt  ðCH 3 Þads ð12:5Þ

Equation 12.5 illustrates the dissociation of the adsorbed aldehyde into adsorbed
CO and CH3. Hence, the adsorbed methyl may recombine with adsorbed hydrogen
to produce methane and refresh catalytic metal sites (Eqs. 12.6 and 12.7).
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 481

2Pt þ H2 O ! Pt  Hads þ Pt  ðOH Þads ð12:6Þ

Pt  ðCH 3 Þads þ Pt  H ads ! CH 4 þ 2Pt ð12:7Þ

In addition, CO may react with the hydroxyl adsorbed at the platinum surface to
produce CO2.

Pt  ðCOÞads þ Pt  ðOH Þads ! CO2 þ H þ þ e þ 2Pt ð12:8Þ

The formation of strongly adsorbing intermediates like CO that block the Pt


catalyst surface sites together with the difficulty in C-C bond breaking are the two
key obstacles to overcome for exploiting as much as possible of the energy density
of ethanol as fuel. For this the role that the adsorbed hydroxyl species play in the
oxidation of ethanol is fundamental. The presence of adsorbed CO species at the
platinum surface hampers catalytic activity. Nevertheless, its occurrence is essen-
tial to produce a full oxidation to CO2. Indeed it is the presence of adsorbed
hydroxyl which allows CO to be oxidized to CO2 (Eq. 12.8). Coupling materials
capable of increasing the rate of formation of the adsorbed hydroxyl at the platinum
surface is indeed a key for increasing the effectiveness of ethanol electrooxidation.
Ethanol electrooxidation in alkaline media has become increasingly investigated
as Pt can be effectively substituted by other transition metals in particular Pd and
Au. Nevertheless, C-C bond cleavage in alkaline environments has been shown to
not occur for ethanol at pHs larger than 13 [23], acetate being the only oxidation
product according to Eq. 12.9.

C2 H 5 OH þ 5OH  ! CH 3 COO þ 4H 2 O þ 4e ð12:9Þ

which proceeds according to the mechanism illustrated in Scheme 12.2.

Scheme 12.2 Ethanol electrooxidation in alkaline media


482 H.A. Miller et al.

In order to produce full oxidation to acetate the adsorption of the hydroxyl anion
is essential. Indeed, it has been demonstrated that at low over potentials on Pd
catalysts, hydroxyl adsorption is the rate-determining step [27] (Eq. 12.10).

Pd þ OH  ! Pd  OH ads þ e ð12:10Þ

The nanostructuring of electrocatalytic materials which increases the hydroxyl


adsorption rate of palladium materials has been shown to be effective in enhancing
ethanol electrooxidation. In one prominent example, the addition of CeO2 to a
Vulcan-XC72 carbon supported Pd catalyst improved significantly the kinetics of
ethanol electrooxidation. The enhanced activity was attributed to the phenomenon
known as primary oxide spillover [28].
The importance of OHads on the ethanol electrooxidation mechanism on Pd in
alkaline media has also been supported by ab initio molecular dynamics simulations
[29]. These studies have shown that within the inner Helmholtz layer (IHL), OHads
is more active for α-CH bond breakage than Pd surface sites indicating that OHads is
the active centre for dehydrogenation. After the formation of OHads, the predomi-
nant channel for the generation of electric current occurs through water dissociation,
with new OHads formed for further dehydrogenation. Ethanol decomposition and
water dissociation hence form a complete electrocatalytic cycle, transferring H to
the aqueous solvent on a continuous basis. Consequently, the concentration of
OHads (OH– anion) on or near the surface should have a great impact on the
electroactivity of ethanol oxidation. Compared to the typical multistep mechanisms,
the mechanism proposed by these researchers (shown in Fig. 12.1) involves limited
intermediates that concur with experimental observations.

Fig. 12.1 General scheme describing ethanol electrooxidation in a DEFC on a Pd electrode in the
presence of the electrical double layer. The red arrows indicate ethanol dehydrogenation while the
purple arrows indicate water dissociation. Reprinted with permission from [29] Copyright (2014)
American Chemical Society
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 483

Scheme 12.3 Reaction pathways for interfacial ethanol oxidation on a Pd electrode in alkaline
media (The solid-line arrows denote the findings described in [30]). Reprinted with permission
from [30] Copyright (2014) American Chemical Society

Our understanding of the ethanol electrooxidation mechanism on Pd in alkaline


media is still much less complete as compared to Pt in acidic media despite the
number of recent in-situ investigations. In an important example, Yang and
co-workers used in-situ attenuated total reflection surface enhanced infrared
adsorption spectroscopy (ATR-SEIRAS) to study ethanol electrooxidation on Pd
in alkaline media and proposed a duel pathway (Scheme 12.3) [30]. This study
showed that at OCV adsorbed acyl species (CH3COads) form via a non electro-
chemical pathway. The adsorbed acyl species thus formed can be further
electrooxidized either through a C2 pathway to form acetate (at high anode poten-
tials) or a C1 pathway to COad and CHx and eventually CO2 (at low potentials). The
real-time spectral measurements with and without isotope labeling indicated that
ethanol may undergo dehydrogenation at α-C to form adsorbed acyl rather than
acetaldehyde, followed by a successive decomposition to C1 species, including
COad and CHx at open circuit potential or lower potentials. As the potential
becomes more positive, the adsorbed acetyl may be converted to either acetate in
the C2 pathway or CO2 via the C1 species described above; adsorbed CHx may also
be converted to COad species at high potentials. Figure 12.2 clearly shows that the
intensity of the vs(OCO) acetate band observed in the ATR-SEIRA spectra
(D) largely traces the CV current density curve (C), suggesting a predominant
share of the C2 pathway in the overall EOR current density.
It is worthwhile to note that both the C1 and C2 pathways are often used to
account for EOR on both Pt and Pd electrodes, and in the majority of cases the C2
pathway remains the predominant one. One of the interesting conclusions of this
study was the suggestion that inhibiting direct oxidation of the adsorbed acyl should
be an effective way of enhancing the C1 pathway of EOR.
484 H.A. Miller et al.

Fig. 12.2 In situ ATR-SEIRA spectra obtained with a time resolution of 5 s on a Pd electrode in
0.1 M NaOH/0.5 M EtOH solution, taking the single-beam spectrum at 0.5 V (a) and 0.3 V (b) as
the reference spectrum. (c) Corresponding CV of the Pd electrode in a 0.1 M NaOH/0.5 M EtOH
solution at 5 mV/s, together with (d) potential-dependent band intensities for ν(COad) (red), νs
(OCO) of adsorbed acetate (pink), and ν(C═O)acetyl þ δ(HOH)free (blue). Reprinted with
permission from [30] Copyright (2014) American Chemical Society

12.2.2 Ethylene Glycol and Glycerol

The number of polyalcohols which can be potentially employed in alkaline DAFCs


is quite extensive with recent extension to 1,2-propandiol, 1,3-propandiol and
1,4-butandiol good examples [31]. Here we limit the discussion of the mechanisms
to the electrooxidation of EG and G as there are many more reports of alkaline fuel
cells employing these fuels in the recent literature [32]. As with ethanol complete
oxidation of these species to CO2 has not yet been achieved although for EG or G
some C-C bond scission may occur in alkaline media with eventual formation of
carbonate although this remains a minor reaction pathway as compared to the
formation of various carboxylates. Such reactions are, in turn, not completely
selective with usually a mixture of products formed. As shown in Scheme 12.4,
in the case of EG on Pd-based anodes in alkaline media, carbonate can be produced
by the oxidation of glycolate (route c), whereas the major product, oxalate, is
produced by the direct oxidation of EG (path b).
The oxidation of G is not surprisingly more complex than that of
EG. Scheme 12.5 illustrates the possible reaction mechanisms proposed for G
oxidation on Pd, Pt and Au based anode electrocatalysts. On Pt the main product
is glycerate produced via glyceraldehyde. Glycerate is the primary oxidation
product on Au electrodes. At higher potentials (>0.8 V), glyceric acid can be
further oxidized to glycolate, formate and CO2 [33].
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 485

Scheme 12.4 Direct and sequential routes for the electrooxidation of EG on Pd in alkaline media.
Reproduced from [32] with permission of John Wiley and Sons

Scheme 12.5 Main oxidation pathways of glycerol oxidation in alkaline media

Glycerol on Pd-based anodes is first oxidized to the aldehyde (path a), which in
turn is quickly oxidized to glycerate in a subsequent two-electron transfer step.
Glycerate is further oxidized to tartronate (path a2) and, by cleavage of C-C bond,
into glycolate and formate (path a2 and b1). The latter species leads to carbonate
and glycolate is oxidized to oxalate (path a4) [33–36]. G is also oxidized directly to
tartronate (path b) by chelating adsorption on catalytic surfaces. The oxidation of G
yields significant amounts of carbonate. Since oxalate is very slowly oxidized on
486 H.A. Miller et al.

Pd-based electrodes [34, 35, 37, 38] CO2 is prevalently a by-product of the
oxidation of either glycerate or glycolate (paths a1 and a3).
Longer chain alcohols have also been proposed for use as fuels in DAFCs.
Isopropanol oxidation on Pt and Pd in alkaline media has been shown to produce
acetone as the only soluble product [39]. Other more complex poly-alcohols that
have up till now not been reported in DAFCs, such as 1,2-propandiol,
1,3-propandiol and 1,4-butandiol have only been reported using organometallic
anode electrocatalysts in OMFCs [31]. This will be discussed in the section devoted
to molecular electrocatalysts.

12.3 Direct Alkaline Alcohol Fuel Cells

Alcohols can be oxidized both in acidic and alkaline environments. However, the
employment of higher alcohols (e.g. EtOH, G and EG) in proton exchange mem-
brane fuels cells (despite intensive investigation) is limited by a number of factors
including; (i) extremely sluggish anode electrode kinetics requiring high noble
metal loadings, (ii) incomplete oxidation [24] and (iii) severe poisoning of Pt
based catalysts by carbonaceous intermediates. Consequently, acidic DAFC
devices have been realized that produce power densities in the range of only a
few tens of mW cm2 and with very low fuel efficiencies.
At present the best performing DAFCs [40] are those operating in alkaline
environments [14] employing anion exchange polymeric membranes (AEMs) com-
bined with a liquid electrolyte in the fuel solution (e.g. KOH) (Fig. 12.3). Both the
kinetics of alcohol oxidation reactions and the ORR are faster at high pH. It has also
been shown that when an acid electrolyte is changed to alkaline media, the fuel cell
efficiency increases [41–43].

Fig. 12.3 Schematic


Load
representation of a DAFC
e- e-
fueled with ethanol
operating with an anion Cathode Anode
exchange membrane OH- CH3COOK
O2
(AEM) and KOH
electrolyte
H2O
CH3CH2OH
+ H2O
+ KOH
AEM
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 487

12.3.1 Cathode Catalysts and Anion Exchange Membranes

Regarding the cathode electrocatalyst in alkaline environments, platinum can be


readily replaced by non noble metal materials [20, 44–47]. This is an important
advantage as these catalysts are often not active toward alcohol oxidation hence
limiting the consequences of fuel cross-over to the cathode. The most commonly
utilized non PGM cathode catalysts employed in alkaline DAFCs have been
transition metal (Fe, Co, Ni, Mn, Cr) phthalocyanine based macrocycles [48, 49]
as well as Fe/C [50], Cu/C [51], FeCu/C [52–54] FeAg/C [19] and FeCo/C [18]
based catalysts produced from carbon supported heat treated (600–1000  C) metal
phthalocyanine precursors. A number of other non PGM cathodes have now been
reported derived from separate metal and nitrogen precursors brought together with
a carbonaceous support in a heat treatment step under inert atmosphere
(600–1000  C) [55, 56]. Nitrogen doped carbon materials that may or may not
contain transition metals have also attracted much attention as alternatives to Pt
[57, 58]. For example nitrogen containing carbon nanotube arrays (NCNTs) have
been shown by Gong et al. to be highly stable and have an onset potential for the
ORR comparable to Pt/C in 0.1 M KOH [59]. Unlike Pt/C the NCNTs were shown
to be immune to the presence of methanol, H2 and CO.
The range of anion exchange membranes that have been investigated for elec-
trochemical systems including DAFCs has been recently nicely reviewed by
Varcoe et al. [17]. However, for the most part commercially available AEMs
have been employed to obtain the best performance including those produced by
Tokuyama, Solvay and Fumatech.

12.3.2 Fuel Cell Performance

In Tables 12.2, 12.3, and 12.4 a selection of the best recent DAFC performance data
reported in the literature is shown for the three fuels EtOH, EG and G. Apart from
isolated examples the best performance has been obtained using FeCo/C or FeCu/C
cathode catalysts combined with Tokuyama AEMs (A201 or A901). Fuel concentra-
tions ranging from 1 to 3 M alcohol and up to 6 M KOH liquid electrolyte have been
used. The importance of the KOH electrolyte to performance is vital not only to
maintain ionic conductivity in the electrode layer and membrane but also provides
sufficient OH for the alcohol oxidation reaction as the AOR is highly pH dependent.
Keeping the pH above 14 favors the direct 4e oxidation to carboxylates avoiding
poisoning intermediates that can be formed at lower pHs (e.g. CO).
Ethanol has been by far the most investigated fuel. As can be seen from
Table 12.2 the most successful anode catalysts are based primarily on
Pd. Enhancements in performance with respect to simple Pd/C catalysts have
been achieved through both alloy formation (e.g. PdIrNi/C and Pd3Ru/C) and by
exploiting strong metal-support interactions using metal oxide supports (e.g. CeO2,
488 H.A. Miller et al.

Table 12.2 Selected DAFC performance with ethanol


Pmax Tcell OCV
Anode Cathode Membrane (mW cm2) ( C) (mV) Fuel Ref.
Pd/C-CeO2 FeCo/C A201 140 80 870 10 % EtOH, [60]
1 mg cm2 2 mg cm2 2 M KOH
Pd/TNTA-web FeCo/C A201 210 80 600 10 % EtOH, [61]
1.5 mg cm2 2 mg cm2 2 M KOH
Pd/TNTA-web FeCo/C A201 335 80 900 10 % EtOH, [61]
2 2 2 M KOH
6 mg cm 2 mg cm
Pd/Ni-foam FeCu/C A201 164 60 870 EtOH 3 M, [62]
3 mg cm2 2 mg cm2 3 M KOH
PdNi/C FeCo/C A201 130 80 900 EtOH 3 M, [63]
2 mg cm2 2 mg cm2 5 M KOH
PdIrNi/C FeCo/C A201 92 60 900 EtOH 3 M, [64]
1 mg cm2 2 mg cm2 5 M KOH
Pd3Ru/C MnO2 A201 176 80 800 EtOH 3 M, [65]
nanotube 3 M KOH
1 mg cm2 2 mg cm2

Table 12.3 Selected DAFC performance with ethylene glycol


Pmax Tcell OCV
Anode Cathode Membrane (mW cm2) ( C) (mV) Fuel Ref.
Pt/C FeCu/C A201 71 50 868 1 M EG, [66]
1 mg cm2 1 mg cm2 2 M KOH
Au/C FeCu/C A201 7.3 50 478 1 M EG, [66]
1 mg cm2 1 mg cm2 2 M KOH
Pd/TNTA-web FeCo/C A201 170 80 920 10 % EG, [61]
1.5 mg cm2 2 mg cm2 2 M KOH
PdNi FeCo/C PBI 112 90 1 M EG, [67]
1 mg cm2 1 mg cm2 7 M KOH

Table 12.4 Selected DAFC performance with glycerol fuel


Pmax (mW Tcell OCV
Anode Cathode Membrane cm2) ( C) (mV) Fuel Ref.
SD-PtCo/CNT FeCu/C A901 268.5 80 860 3 M G, [68]
0.5 mg 3 mg cm2 6 M KOH
Pt/C FeCu/C A901 169 80 790 3 M G, [68]
0.5 mg 3 mg cm2 6 M KOH
Pt/CNT FeCu/C A901 221 80 840 3 M G, [68]
3 mg cm2 6 M KOH
Pd/TNTA-web FeCo/C A201 160 80 940 10 % G, [61]
1.5 mg cm2 2 mg cm2 2 M KOH
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 489

TiO2 and Ni/NiO). Ma and coworkers using a carbon supported Pd3Ru anode and
MnO2 nanotube cathode achieved a peak power with ethanol at 80  C of
176 mW cm2 while Bambagioni et al. obtained 140 mW cm2 peak power also
at 80  C using a Pd nanoparticle catalyst supported on a mixed carbon-CeO2
support. Even higher power densities have been obtained using Pd nanoparticles
supported on titania nanotube arrays (TNTAs). Chen et al. was able to reach
335 mW cm2 peak power by using a 6 mgPd cm2 loaded TNTA-web material
at 80  C.
Reports of complete cells using EG and G are much less prolific. In most cases
simple mono-metal nanoparticle anodes have been used. Unlike with EtOH the
oxidation kinetics on Pd are more sluggish and as a consequence good performance
has been obtained also with Pt/C catalysts. Xin et al. compared Pt/C and Au/C
anodes with EG and was able to obtain 7.3 mW cm2 with Au while using Pt/C
produced ten times better power output. With glycerol as fuel higher power outputs
have been reported. Qi and coworkers using CNT supported Pt catalysts obtained
>200 mW cm2 at 80  C with a very impressive 268.5 mW cm2 obtained using
surface dealloyed PtCo nanoparticles supported on CNTs (SD-PtCo/CNT).
Nickel has been cited often as a possible low cost alternative to precious metals
for alcohol electrooxidation [69–74]. Despite many research efforts the
overpotential for alcohol oxidation on nickel catalysts is too high for practical
application in fuel cells. In the next section we will discuss how nanostructuring
metal catalysts have been used in improving the kinetics and overpotential of
alcohol electrooxidation in alkaline media. Emphasis will be given to materials
actually employed in DAFCs and the selectivity for the preparation of partially
oxidized intermediates will also be discussed.

12.4 Nanostructured Materials for Alcohol


Electrooxidation in DAFCs

12.4.1 Palladium Based Nanocatalysts

A recent direct comparison of the ethanol electrooxidation behaviors of Pt/C and


Pd/C in alkaline media demonstrated that Pd is superior for two main reasons;
(i) the higher oxophilic nature of Pd that promotes the adsorption of OH groups and
(ii) the relatively inert nature of Pd to C-C bond cleavage [75]. Pd is consequently
more stable and less susceptible to poisoning in the direct oxidation of ethanol to
acetate. Indeed, the number and diversity of Pd based compounds that have been
developed for alcohol electrooxiation under alkaline conditions is large. As we have
seen in the previous section DAFCs equipped with Pd anodes can produce peak
power densities of over 100 mW cm2 [1, 60, 76].
The catalytic activity and selectivity of palladium electrocatalysts can be
improved by the addition of a second metal or metal oxide with a promotional
490 H.A. Miller et al.

effect. The activity of Pd/C catalysts for ethanol electrooxidation (EOR) has been
enhanced by a number of researchers by alloy formation with non noble metals.
Shen and co-workers obtained a peak power density of 90 mW cm2 using a
Pd2Ni3/C alloy as anode catalyst in a DAFC operating at 60  C. X-ray photoelec-
tron spectroscopy (XPS) analysis of the anode catalyst revealed the chemical states
of Ni, including metallic Ni, NiO, Ni(OH)2 and NiOOH [77]. In another recent
report Pt-Pd alloy nanostructures exhibited enhanced activity and stability for
ethanol electrooxidation (EOR) through an electronic, synergistic alloy effect
between Pt and Pd atoms, and the presence of unique interconnected nanostructures
[78]. The direct oxidation pathway from ethanol to acetate was found to be favored
by this catalyst. The addition of Ni based compounds to Pd catalysts generally leads
to an enhancement in the EOR. Jiang and coworkers prepared a Pd–Ni–P compound
for ethanol oxidation. It was found that the presence of Ni and P atoms modify the
crystal structure, and charge transfer to the nearby Pd atoms thus enhancing the
EOR. Compared to a Pd black catalyst a 110 mV decrease in overpotential was
observed for the EOR on the Pd-Ni-P catalyst [79]. The four electron direct
oxidation to acetate was also confirmed for this material.
The combination of Pd and Au in alloyed and core-shell materials either carbon
supported or without support has been the subject of a number of recent reports
regarding enhancements in EOR activity [80–84]. The addition of gold to palladium
nanostructures has an important ameliorative effect on the EOR. Firstly, because
the co-metal Au favors the adsorption of OHads onto the Pd surface of the catalyst.
Secondly, as the lattice constant of Pd is smaller than that of Au, the addition of Au
to Pd causes a tensile strain in the structure of surface Pd and an up shift of the
d-band center of Pd [85]. According to d band theory of Nørskov and coworkers the
trend of reactivity follows the trend in d-band center values of overlayer and
impurity atoms [86–90]. When metals with small lattice constants are overlayed
or alloyed on metals with larger lattice constants, the d-band center shifts up and
vice versa, which subsequently affects the reaction rate. Since the lattice parameter
for Au (4.08 Å) is larger than that for Pd (3.89 Å), the lattice mismatch between Au
and Pd leads to an expansive strain of Pd and an up-shift of d-band center, which
weakens the interaction between the adsorbed intermediate species CH3COads and
the surface of Pd and thus results in an enhanced activity toward EOR [82, 91]. This
effect has been supported by XPS results that show the binding energies of Pd shift
to higher values as the Au/Pd ratios increased [80]. Two recent examples of
bimetallic PdAu non supported catalysts in the form of nanoporous solids [83]
and nanowire networks [80] (NNW) have been shown to have in addition to
enhanced activity also greater stability. Such nano-networks can possess the advan-
tages of the one-dimensional structure such as improved electron transport charac-
teristics during catalysis as a result of the path-directing effects of structural
anisotropy [92, 93] combined with the fact that self-supported materials avoid the
severe corrosion and oxidation of the carbon supports [94, 95].
Researchers incorporating Sn or SnO2 into Pd based catalysts have obtained
remarkable improvements in ethanol oxidation activity in alkaline media
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 491

[96–99]. In Pd-Sn binary alloys prepared by Du et al. using a polyol method, the Sn
was found to be partially oxidized in the otherwise homogeneous alloy [96]. Using
DFT calculations the optimum Sn content of 14 wt% was determined. DFT
calculations also confirmed that Pd–Sn alloy structures would result in lower
reaction energies for the dehydrogenation of ethanol, compared to the pure Pd
crystal. All of the PdSn/C catalysts produced at least twice the peak current density
for ethanol oxidation in CVs with respect to Pd/C. The results also confirmed the
formation of acetate as major product of ethanol oxidation. Atanassov and
co-workers recently studied the mechanism of ethanol electrooxidation on SnO2
supported Pd nanoparticles. In situ IR reflection absorbance spectroscopy (IRRAS)
studies showed that in 1 M KOH the SnO2 support acted as a co-catalyst providing
hydroxide ions to the interface layer, increase the turnover rate, and limiting the
final product to acetate whereas in 0.1 M KOH some CO2 was formed at high
potentials [100].
The nature of the catalyst support material can have a dramatic effect on
performance. Increasingly, researchers are investigating novel materials both car-
bon and non-carbon based. An important recent example is reduced graphene oxide
(RGO). Catalysts prepared with this material have ultra-high dispersion of
nanoparticles combined with strong SMSIs (Strong Metal Support Interactions).
Zhang et al. used a one-pot solvothermal synthesis to prepare Pd-on-Cu
nanoparticles evenly distributed on reduced graphene oxide (Pd-on-Cu/RGO)
[101]. The as-prepared nanocomposites exhibited a large EASA of 64.98 m2 gPd1
and as a consequence improved electrocatalytic activity, and better stability for
ethanol oxidation in alkaline media. RGO modified with various organic materials
such as Dimethyldiallylammonium chloride (DMDAAC) [102] has been used as a
support for a number of highly active Pd based EOR catalysts including Pd [103],
PdCo [104], PdCu [105], NixPd100-x [106] and Ni@Pd [107]. All showed enhanced
activity and stability for the EOR. A number of other exotic supports have been used
to support Pd nanoparticles for EOR. An anode consisting of Pd supported on
hydroxyapatite [108] combined with a Pt/C cathode and anion exchange membrane
produced a peak power density of 50 mW cm2 at room temperature with 1 M
ethanol and 1 M KOH (Pd loading of anode was 0.5 mg cm2). MWCNT supported
Pd nanoparticles have been investigated in DAFCs. Together with a FeCo/C cathode
and anion exchange membrane (Tokuyama A201) peak power densities of 95, 75
and 80 were obtained at 80  C with methanol, ethanol and glycerol respectively
[109]. The same authors also supported Pd on a composite FLG-CNT material where
the CNTs were shown to improve Pd dispersion by inhibiting aggregation of the
graphene sheets [110].
Another strategy to prepare Pd catalysts without carbon supports has been to
prepare self supported materials which combine active phase and high surface area
support. A number of self supported bimetallic Pd catalysts have been developed
for the electrooxidation of glycerol and ethylene glycol such as Pd-In [111] and
PdxBi [112]. These catalysts were prepared using the Sacrificial Support Method
(SSM) developed at the University of New Mexico. This synthetic method pro-
duces self-supported, porous nanostructured materials. The reaction products of
492 H.A. Miller et al.

glycerol oxidation were studied using the PdxBi material. This important study
showed how the catalyst composition and morphology such as the formation of
pores that act as nanoreactors influence the product selectivity between the forma-
tion of aldehydes and ketones at low potentials as opposed to hydroxypyruate and
CO2 at high potentials. In another recent example, the synergistic effect of Au-Pd
bimetallic surfaces in Au-covered Pd and Pt nanowires for ethanol electrooxidation
in alkaline media has been reported [113, 114]. Cherevko and co-workers prepared
highly ordered Pt, Pd and Au nanowire arrays using a home-made AAO (Anodic
Aluminum Oxide) template by electrodeposition [114]. After which the AAO
template was removed. Decoration of Pt and Pd on the as prepared Au nanowire
array was achieved using chemical reduction with ascorbic acid. The ethanol
electrooxidation activities of the decorated materials were several times greater
than the Pd or Pt only arrays although the synergistic effect of the two metals was
not determined but could be due to an up-shift in the energy of the d-states.

12.4.1.1 Product Selectivity on Palladium

As described previously ethanol electrooxidation at high pHs on Pd proceeds


through the direct formation of acetate. The product selectivity of electrooxidation
of G and EG in alkaline fuel cells is not the same and can be largely influenced by
the nanostructure of Pd electrodes. This was shown by Vizza and co-workers
comparing a simple Pd/C catalyst with a nanostructured Pd/(Ni-Zn)-C anode in
DAFCs operating at room temperature [32]. Both the conversion and the product
selectivity varied between the two catalysts. The improvement in conversion with
the Pd/(Ni-Zn)-C anode was assigned to the ability of the (Ni-Zn)-C support to
increase the amount of OHads groups on the Pd catalyst surface. On the other hand
the Pd/C catalyst was much more selective in EG oxidation, yielding 89.5 % of
glycolate whereas Pd/(Ni-Zn)-C yielded in addition to glycolate both oxalate and
carbonate. With G there was little difference in the product selectivities between the
two catalysts with a mixture of glycolate, glycerate, tartronate, oxalate, formate,
and carbonate formed. No secondary alcohol oxidation products were detected with
either catalyst (dihydroxyacetone, hydroxypiruvate, or mesoxalate).

12.4.2 Nanostructured Gold Electrocatalysts

Gold electrodes in acidic media typically exhibit very low activity for alcohol
oxidation. In alkaline media however Au is substantially more active for alcohol
oxidation and the electrooxidation of a wide range of substrates have been studied
on Au electrodes (e.g. D-glucose, xylose, etc.) [115]. The improved oxidation of
alcohols on gold electrodes in alkaline media has been attributed to the
deprotonation of the R–OH group that occurs at sufficiently high pHs, and leads
to the formation of the alkoxide species (RO), that are much more active in
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 493

oxidation reactions [115]. Varella and de Lima studied the catalytic oxidation of
ethanol on a polycrystalline gold electrode in alkaline media [116]. Cyclic
voltammetry combined with in-situ FTIR, showed that adsorbed CO formed at
very low overpotentials (circa 600 mV vs. RHE), indicating that a small amount of
C-C bond breakage occurs. Acetate ions were however identified as the major
product of electrooxidation.
The electrooxidation of ethanol on Au(1 1 1), Au(1 0 0), Au(1 1 0) and Au(2 1 0)
single crystal electrodes was studied by Beyhan and co-workers as well as on a Au
polycrystalline electrode [117]. The activity for the electrooxidation of ethanol in
the low potential region was ranked as follows: Au(1 1 0) > Au(2 1 0) > Au
(1 0 0) > Au(poly) > Au(1 1 1), suggesting that defect sites on the Au surface dom-
inate. Adsorbed OH was shown to play an important role in the ethanol oxidation
reaction and FTIR spectroscopy measurements showed that the main reaction
product is acetate. Ethanol electrooxidation has been widely investigated on
carbon supported gold nanoparticle catalysts; including gold-conducting
polymer nanocomposites (Polyaniline, polypyrrole, polythiophene and poly
(3,4-ethylenedioxythiophene) [118]. In another recent example Kumar
et al. prepared a nanostructured Au decorated graphene catalyst. The onset of
ethanol electrooxidation was shown to shift favorably [119].
Despite being extensively studied, the overpotential for ethanol electrooxidation
on nanostructured Au electrodes is too high to have practical use in DAFCs. By
contrast gold based catalysts have been more readily applied as anodes in fuel cells
and electrolysis cells fed with glycerol [120] as the onset potential for glycerol
oxidation has been shown to be similar to palladium. On gold the glycerol
electrooxidation reaction has been shown to be structure sensitive (Au(1 1 0) >
Au(1 0 0) and (Au(1 1 1)) and the rate determining step in alkaline media is the
adsorption of hydroxyl ions with partial charge transfer (Eq. 12.11).

Au þ OH  ! AuOH λ þ ð1  λÞe ð12:11Þ

The research group of Li at Michigan Technological University has extensively


applied gold and gold alloy catalysts in DAFCs using both glycerol [7, 121–126]
and ethylene glycol [66] as fuels. A nanoparticle Au/C (3.5 nm) catalyst was
prepared and studied for EG and G electrooxidation in alkaline media. The Elec-
trochemical Active Surface Area (EASA) of the Au/C catalyst was calculated as
26.8 m2 g1. An anode prepared with this catalyst (1 mgAu cm2) and combined
with a Tokuyama A201 membrane and FeCu/C cathode was studied in a DAFC
fueled with 1 M alcohol and 2 M KOH. The behavior of the cell was investigated
with G, EG and methanol under the same test conditions at 80  C. When fed with
2.0 M KOH þ 1.0 M methanol, the cell yielded an OCV of 0.29 V and an peak
power density of 0.8 mW cm2 (at 8 mA cm2). Much better performance was
achieved with EG; OCV 0.58 V PP 20.3 mW cm2 and G OCV 0.67 V and PP
57.9 mW cm2.
494 H.A. Miller et al.

12.4.2.1 Product Selectivity on Gold

Li’s group also studied the product distribution of working DAFCs using the Au/C
anode. Product analysis showed that on Au/C high selectivity of EG
electrooxidation to glycolic acid (>98 %) was obtained while the electrooxidation
of glycerol favored the production of deeper-oxidized chemicals: i.e. tartronate,
mesoxalate and oxalate. The same authors showed that the product distribution
could be changed by tuning the anode potential; by moving from 0.35 to 0.65 V, the
selectivity to tartronate dropped from 79 to 26 %, while that to mesoxalate
increased from 0 to 57 % [123]. Further tuning of the electrode structure and
reaction conditions helped to produce >60 % yield of tartronate from glycerol. A
gold nanoparticle catalyst supported on CNTs was also used in an electrolyzer set at
1.6 V (SHE). Under these conditions 85 % selectivity for the formation of glycolate
was achieved avoiding the tartronate route confirming that the anode potential has a
large effect on selectivity for electrooxidation on Au [121].

12.4.3 Fine Tuning the Surface Structure of Nanocatalysts at


the Atomic Level

The electrooxidation of alcohols like ethanol, EG and G in DAFCs is only partial


and this is the case in both acid and alkaline media. Further oxidation requires the
cleavage of the CC bond and as we have seen for ethanol the main products are
acetaldehyde and acetic acid, while CO2 production contributes to less than 2 % of
the total current even on the most effective electrocatalysts (e.g. Pt  Sn alloys)
[127]. Consequently, the fuel utilization efficiency of an ethanol fed DAFC is
actually very low. Higher conversion to CO2 has been demonstrated only working
with very dilute ethanol solutions in acidic DEFCs (Faradaic efficiency 64 % for
0.1 M ethanol at 80  C) [128].
A number of studies on the reactivity of Pt single-crystal planes with hydrocar-
bons (e.g. hydrogenolysis and isomerization) have revealed that the product selec-
tivity depends highly on surface atomic arrangements and the presence of low
coordinated step atoms, especially kink atoms, which can promote the breaking of
CC bonds. Surfaces with a high number of low coordinated surface atoms have
high-index facets. High index planes have at least one Miller index larger than
1. The (1 1 1) and (1 0 0) planes are atomically flat with closely packed highly
coordinated surface atoms. Pt high-index planes exhibit higher reactivity than these
low index planes. This knowledge has been very helpful when designing catalysts
used for the electrooxidation of small organic molecules that contain CC bonds.
For example Tarnowski et al. [129] studied the effects of step atoms on the
selectivity of ethanol electrooxidation by using Pt(5 3 3), Pt(7 5 5), and Pt(1 1 1)
single crystal electrodes. The yield of CO2 was increased by introducing the (1 0 0)
step on a (1 1 1) terrace. At the same time the yield of acetic acid on the Pt(5 3 3)
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 495

electrode was only approx. 25–30 % of that on the Pt(1 1 1). Sun and co-workers
[130] in a similar way studied the electrooxidation of isopropanol on platinum and
found that Pt(6 1 0) was the most active surface for yielding CO2. The order of
activity for producing CO2 was Pt(6 1 0) > Pt(1 1 1) > Pt(1 0 0) > Pt(2 1 1) > Pt
(1 1 0). These results clearly demonstrate that high-index surfaces with a high
density of step and kink atoms do significantly promote the complete
electrooxidation of fuels containing CC bonds.
In some cases this strategy has been applied to the EOR in alkaline media. For
example Zhang and coworkers successfully prepared concave palladium polyhedra
(size >50 nm) by an in situ facet-selective etching growth route. Representative
electron microscopic images of the as-obtained Pd nanoparticles are shown in
Fig. 12.4 [131]. Figure 12.4c depicts model images of ideal concave polyhedra in
different orientations, and the corresponding nanocrystals (NCs) are indicated in
Fig. 12.4b. Each particle is comprised of 32 facets in total, including 24 identical
five-edge facets and 8 identical triangular facets. Aberration-corrected high-reso-
lution (HR)TEM was used to characterize the surface structure (Fig. 12.4d). It can
be seen that the concave polyhedral Pd NCs possess a high density of atomic steps
on their surface (see the border atoms highlighted in 12.4e, f). Due to the presence
of a high density of atomic steps, these nanoparticles exhibited an enhanced specific
electrocatalytic activity towards ethanol oxidation in alkaline media.
Sang Woo Han and co-workers have recently reported the synthesis of Au-Pd
alloyed nanoparticles enclosed exclusively by high-index facets {5 4 1}
[132]. These nanocrystals were realized using a simple one-pot aqueous synthesis
that did not require seed or additional metal ions as structure regulating agent. The
nanocrystals with an average size of 114 nm displayed an hexaoctahedral structure
(i.e. a polyhedron bound by 48 triangular high index facets) and exhibited higher
catalytic performance toward the electro-oxidation of ethanol in alkaline media
than equivalent Au-Pd nanocrystals bound by low-index facets.
The majority of bottom-up methods for shape and structure control of bimetallic
core-shell or faceted metal nanoparticles such as “seeded growth” or “kinetic
controlled growth” tend to lead to particles sizes larger than 50 nm. At the same
time we know that electrocatalytic activity is optimized on nanoparticles in the sub
10 nm size range [133, 134]. Such size control has been achieved by some
researchers using extremely slow growth. For example Tilley and coworkers
prepared Pd shell on Au core-shell nanocrystals with precise layer by layer control
of the shell thickness [135]. More facile methods to obtain particle sizes in the sub
10 nm range involve top-down milling. Top-down nanostructuring with precious
metal decoration of core-shell particle formation has been achieved by Ozoemena
et al. A precursor material FeCo@Fe/C (Fe shell on FeCo core) with large particle
sizes (>210 nm), subjected to a rapid solvothermal microwave reaction in the
presence of a Pd salt, lead to the downsizing of the core particles and decoration
of Pd (particle sizes 3–7 nm) [136]. The resulting FeCo@Fe@Pd/C core-shell
particles exhibited enhanced activity for EtOH, EG and G electrooxidation in
alkaline media when compared to a Pd/C catalyst with the same metal
loading [137].
496 H.A. Miller et al.

Fig. 12.4 (a) TEM and (b) SEM images of concave polyhedral Pd NCs. (c) Model images in
different orientations corresponding to those of the nanoparticles marked with the same number in
(b). (d–f) Aberration‐corrected HRTEM images of concave polyhedral Pd NCs oriented along the
‹110› zone axis. Panel (e) and (f) is the magnified HRTEM image taken from the selected area
marked by the red and blue line in panel (d), respectively. The inset in (e) and (f) is the
corresponding FFT pattern. Reproduced from [131] with permission of John Wiley and Sons

Sun and co-workers in 2007 announced the development of a novel electro-


chemical method for the synthesis of small (2–10 nm) Pt nanocrystals enclosed with
high-index facets [138, 139]. The process involves an electrochemical square-wave
potential deposition process of Pt onto a glassy carbon (GC) electrode in a solution
containing 0.1 M H2SO4 and 30 mM ascorbic acid. This electrochemical shape
control method was also used to obtain high index faceted Pt nano-crystals
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 497

supported on carbon black with a size (2–10 nm) comparable to that of standard
commercial Pt/C catalysts [140]. A much higher density of low coordinate atomic
steps was obtained and this led to a doubling of the cleavage of the C-C bond in the
electrooxidation of ethanol as evidenced by in situ FTIR spectroscopy. Other
researchers have followed suit using square wave potential treatments to obtain
shape transformations. For example Zhou et al. transformed Pt nanocubes to
tetrahexahedra with a size of around 10 nm [141]. During the process the surface
structure was changed from {1 0 0} low-index facets to {3 1 0} high-index facets.
As a consequence the electrocatalytic activity for ethanol oxidation was also greatly
enhanced.
The electrochemical square-wave potential method has also been used, with
some modifications, for the preparation of Palladium nanocrystals with high-index
facets [142]. In this work Pd nanocrystals were electrodeposited (from solution)
onto a GC electrode from a 0.2 mM PdCl2 þ 0.1 M HClO4 solution by programmed
electrodeposition. A SEM image of the Pd nanoparticles is shown in Fig. 12.5
(mean particle diameter 61 nm).

Fig. 12.5 (a) SEM image of Pd nano-crystals. The inset is a high magnification SEM image.
(b) TEM image of a Pd NC recorded along the [0 0 1] direction. (c) HRTEM image recorded from
the boxed area in (b), showing some {2 1 0} and {3 1 0} steps that have been marked by red dots.
(d) Cyclic voltammograms of Pd NCs (solid line) and Pd black catalyst (dashed line) at 10 mV s1
in 0.1 M ethanol þ 0.1 M NaOH. Reprinted with permission from [142]. Copyright (2011)
American Chemical Society
498 H.A. Miller et al.

Fig. 12.6 (A) TNTAs with as-deposited Pd and (a) the corresponding SEM image. (B) TNTAs
with Pd after heavy ECMF and (b) the corresponding SEM image. (C) TNTAs with Pd after heavy
and mild ECFM and (c) the corresponding SEM image. False coloring of the SEM images shows
Pd NPs (light blue) and TNTA support (violet). The white scale bars in (a–c) are 200 nm.
Reprinted from [143] with permission from John Wiley and Sons

The exposed facets (mainly {7 3 0}) were determined by HRTEM and SAED.
The {2 1 0} and {3 1 0} steps are visible on the border atoms in the HRTEM image
Fig. 12.5b, c. The presence of such a high density of surface active sites, yielded
approx 4–6 times higher catalytic activity per unit surface area compared to a
commercial Pd black catalyst (Johnson Matthey, Inc.) for ethanol electrooxidation
in alkaline media (Fig. 12.5d). This Pd catalyst also showed enhanced stability.
After 1000 potential cycles, 75.0–95.5 % of the initial catalytic activity was
maintained.
An adaption of the square wave method has been recently reported with the
name Electrochemical Milling and Faceting (ECMF) [143]. Large Pd nanoparticles
(35 nm) with low-index facets were first supported on TiO2 nanotube arrays
(TNTAs) (Fig. 12.6). A two-step square wave electrochemical treatment was then
applied. In the first “heavy” step, a palladium oxidation was applied at 4.55 V
(vs. RHE) for 180 s, followed by the reduction of the Pd oxides at 1.95 V
(vs. RHE) for 180 s (Fig. 12.6b). This was followed by a milder treatment with a
frequency of 0.025 Hz for 3 h between þ3.35 and 0.75 V (vs. RHE) (Fig. 12.6).
The overall treatment resulted in not only a net reduction in mean particle size to
7 nm particles (milling) but also the formation of high index facets, multiple twins
and a high density of step atoms (Faceting) (Fig. 12.7). Cyclic voltammetry with
ethanol under alkaline conditions showed a peak current density of 201 mA cm2
(Fig. 12.8, curve 3), corresponding to a normalized mass-specific activity for Pd of
8965 Ag1. The onset potential for the oxidation of ethanol shifted 170 mV more
negative than the potential obtained for the as-deposited sample.
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 499

Fig. 12.7 (a) TEM image of the Pd-loaded TNTA electrode after heavy and mild ECMF (scale
bar ¼ 50 nm). (b) Pd nanoparticles found in the electrolyte after heavy and mild ECMF (scale
bar ¼ 35 nm). (c) HRTEM image (scale bar ¼ 2 nm) and (d) atomic models with face assignment
of the TNTA-supported Pd nanoparticle along the ‹100› direction. (e) HRTEM image (scale
bar ¼ 2 nm) and (f) face assignment of the TNTA-supported Pd nanoparticles along the ‹110›
direction. Reprinted from [143] with permission from John Wiley and Sons

Fig. 12.8 Cyclic voltammograms of TNTAs with deposited Pd recorded in (a) 0.1 M HClO4 and
(b) 2 M KOH with 10 wt% EtOH. Scan rate: 50 mV s1. Curve 1: TNTA-Pd as deposited. Curve 2:
TNTA-Pd after heavy ECMF. Curve 3: TNTA-Pd after heavy and mild ECMF. Reprinted from
[143] with permission from John Wiley and Sons

The same authors have applied analogous square wave methods to treat poly-
crystalline Pd and Pt electrodes [144, 145]. Constant potential oxidation reduction
cycles applied to both electrodes resulted in an increase in activity for alcohol
electrooxidation, which was attributed to a combination of increased electro active
surface area (EASA) and an increased concentration of low coordination surface
atoms. Electrochemical in-situ FTIR spectroscopy showed that on the treated Pt
500 H.A. Miller et al.

electrode C-C bond breaking was actually limited when compared to the pristine
electrode thus leading to more stable performance as poisoning intermediates such
as CO are avoided [145].

12.4.4 Nanostructured Metal Oxide Supports

Up till now we have described a wide range of Pd and Pd alloyed catalysts that show
excellent initial activity for alcohol electrooxidation under alkaline conditions both
in half cell investigations and in DAFCs. Despite the large quantity of materials
reported to date very few investigations have dealt with the actual fuel utilization
efficiency of these catalysts employed in alkaline DAFCs. In one rare example
Bambagioni and coworkers using a Pd/C catalyst in a room temperature DEFC
found a fuel energy efficiency of circa 2.6 % [60]. A number of factors have an
effect on the fuel efficiency including incomplete conversion to CO2, catalyst
poisoning, fuel consumption as well as consumption of supporting electrolyte. In
addition palladium based electrocatalysts also suffer from the fact that at anode
potentials higher than 0.6 V (RHE) in half cells and 0.15 in monoplanar DAFCs, the
surface adsorbed Pd-OHads species (Eq. 12.12), active for the oxidation of ethanol,
start converting into inactive Pd-O according to Eqs. 12.13 and 12.14. As a result,
the number of active sites on the electrode surface decreases, and fuel conversion is
slowed down and ultimately inhibited.

Pd þ OH  ! Pd  OH ads þ e ð12:12Þ

Pd  OH ads þ OH  ! Pd  O þ H2 O þ e ð12:13Þ

Pd  OH ads þ Pd  OH ads ! Pd  O þ H2 O ð12:14Þ

In one strategy developed to diminish the extent of Pd oxidation to Pd-O on


DAFC anodes, a small amount of NaBH4 was introduced to the ethanol fuel
solution. This reagent was able to reduce any Pd-O formed back to Pd metal
[146]. An alternative approach involves anticipating the oxidation of Pd0 to PdI-
OHads through the use of co-supports that promote the transfer of adsorbed OH
groups to the Pd catalyst surface. One of the most effective materials used so far is
ceria. CeO2 is a mixed conductor, showing both electronic and ionic conduction,
with many applications in catalysis in conjunction with transition metals. The first
investigation of DAFCs containing anode electrocatalysts made of Pd nanoparticles
supported on a mixture of carbon and ceria showed that, at the same metal loading
and experimental conditions, the energy efficiency of a DEFC assembled with the
Pd–CeO2/C anode electrocatalyst was twice as much as that supplied by the cells
with an analogous Pd/C electrocatalyst [60]. It was proposed that ceria promotes the
formation at low potentials of PdI-OHads, the species responsible for ethanol
oxidation. In DEFCs using the Pd-CeO2/C anode power densities as high as
66 mW cm2 at 25  C and 140 mW cm2 at 80  C were obtained.
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 501

Carbon supported nanoparticle electrocatalysts suffer from limited long term


durability [147, 148]. Dissolution, and metal particle growth (aggregation and
Ostwald ripening), and carbon corrosion account for most of the loss of the
EASA during fuel cell operation [94, 149, 150]. Carbon materials are susceptible
to corrosion/oxidation at high potentials (>0.7 V vs. NHE) under fuel cell operating
conditions, processes which accelerate further the above degradation mechanisms
[151]. The hydrophobicity of carbon also limits the mass transport of electro-active
species inside the catalyst layer which impacts negatively on the kinetics of
electrochemical reactions in the liquid phase.
Alternative catalyst supports have indeed shown improved electrode stability
compared to carbon supported nanoparticles during ethanol electrooxidation in
alkaline media. For example a PdAg alloy supported on a Nb-doped-TiO2 support
exhibited a remarkable improvement in stability compared to Pd/C under fixed
potential and cycling tests [152]. In another example, Pd-PANI-Pd sandwich
nanotubes used for ethanol electrooxidation showed the advantage of an open
electrode structure that favors fast mass transport in the liquid phase [153]. Chen
and coworkers have recently reported a novel electrode structure comprised of a
2 μm thick layer of TiO2 nanotube arrays (TNTA) that cover the surface of a titanium
fiber non-woven web electrode (Fig. 12.9) [154]. Onto this material Pd nanoparticles
are supported by deposition and chemical reduction with NaBH4. This material was
applied as an anode electrode in two electrochemical devices (i) in electrochemical
reforming of alcohols for hydrogen production [154] and (ii) as anode electrode in
DAFCs [61]. The fibrous open structure of the web electrode, provides in the liquid
phase, rapid mass transport and short diffusion paths for electroactive species
(Fig. 12.9). This material, employed as anode electrode in alkaline DAFCs, com-
bined with an anion exchange membrane (A201) and a FeCo/C cathode produced
peak power densities at 80  C of 210, 170 and 160 mW cm2 with EtOH, EG and G
respectively (Fig. 12.10). The flexibility of this material allowed an increase in Pd
loading (from 1.5 to 6 mg cm2) by combining four single electrodes on the anode
side of the DAFC. With ethanol at 80  C the cell produced a peak power of
335 mW cm2 and a maximum current density of 2.25 A cm2 at 0.1 V cell potential
(Fig. 12.10).

12.5 Molecular Anode Catalysts for DAFCs

Increasingly, researchers in this field are striving to use DAFCs to combine the
combination of the production of renewable energy (with no CO2 emissions)
together with the production of industrially relevant feedstocks (such as aldehydes,
ketones and carboxylic acids) [7, 31, 34, 66, 76, 121, 123, 126]. In this way the free
energy of an alcohol is converted into electrical energy and the alcohol itself is
transformed into an oxidation product or products which are different from CO2.
Two types of DAFCs have been developed for this purpose: (1) traditional devices,
where the anode and cathode are separated by an anion-exchange membrane and
502 H.A. Miller et al.

Fig. 12.9 Representative FESEM images of Pd/TNTA-web (a, b). TEM (d) and HRTEM (e)
analysis of Pd/TNTA-web. Particle size distribution (c) from FESEM analysis of Pd particles
around the opening of the nanotubes and (f) particle size distribution from TEM analysis of
particles situated along the tube walls. Reprinted from [61] with permission from John Wiley and
Sons

Fig. 12.10 (Left) Potentiodynamic and power density curves for DAFCs fuelled with 2.0 M KOH
and alcohol (10 wt%) solutions Pd/TNTA-web (1.5 mgPd cm2) and (right) potentiodynamic and
power density curves for combined Pd/TNTA-web anodes (6 mgPd cm2) ethanol 10 wt% in 2 M
KOH. Reprinted from [61] with permission from John Wiley and Sons

are coated with an electrocatalyst, generally a nanostructured noble metal,


supported on a conductive material [36, 76, 109, 155–157] and (2) enzymatic
biofuel cells (EBFC) utilizing oxidation enzymes such as dehydrogenases in con-
junction with an electron transfer mediator [158, 159]. Efficient devices of this type
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 503

Scheme 12.6 Proposed mechanism for the reactions occurring on the anode of the OMFC.
Reprinted from [31] with permission from John Wiley and Sons

have been recently developed for a variety of renewable alcohols and polyalcohols
such as ethylene glycol, glycerol and 1,2-propandiol [36, 156].
Recently a third type of fuel cell has been developed, known as an OrganoMe-
tallic Fuel Cell (OMFC), operating in alkaline media where the anode catalyst is a
molecular metal complex. Several research groups have demonstrated this concept
showing that electrocatalysts based on specific metal complexes can oxidize alco-
hols [160–162] and glucose [163, 164] at low overpotentials. In the most successful
example, an organometallic complex [Rh(OTf)(trop2NH)(PPh3)] ((12.1) in
Scheme 12.6) has been shown to selectively catalyze the oxidation of alcohols to
carboxylic compounds [28, 31, 165].
The metal complex evolves through fast chemical equilibria to form specific
catalysts for ethanol dehydrogenation, aldehyde dehydrogenation and Hþ/electron
transfer. The [Rh(OTf)(trop2NH)(PPh3)] complex (12.1) can be deposited intact
onto a conductive carbon support such as Vulcan XC-72 or Ketjenblack EC 600 JD
to form the anode electrode. The mechanism for alcohol oxidation is shown in
Scheme 12.6.
504 H.A. Miller et al.

Fig. 12.11 Polarization ( filled circle) and power density (circle) curves of OMFCs fueled with:
( filled circle) 1,4-B; ( filled diamond) 1,3-P; ( filled triangle) 1,2-P; ( filled square) EG aqueous
solutions 5 wt% in 2 M KOH. (a) Passive air-breathing OMFC at 22  C and (b) active OMFC at
80  C. Reprinted from [31] with permission from John Wiley and Sons

On the electrode surface, the precursor 1@C is rapidly converted into the
hydroxo complex, [Rh(OH)(trop2NH)(PPh3)]@C (2@C), which is in a rapid
equilibrium with the amide [Rh(trop2N)(PPh3)]@C (3@C) and water. The amide
3@C dehydrogenates diols to aldehydes. Under the basic reaction conditions, the
aldehydes are rapidly further converted promoted by catalyst 2@C to form carbox-
ylate ions and 5@C. The latter complex is oxidized at the electrode, releasing two
Hþ (neutralized to give water under basic conditions) and two electrons with
regeneration of the amide 3@C. There is some resemblance with the enzymatic
biofuel cell, but the main characteristic of this system is that one molecular rhodium
complex is capable of evolving through fast chemical equilibria in the course of the
catalytic cycle to form a specific catalyst for alcohol dehydrogenation (the amide
3@C), a specific catalyst for aldehyde dehydrogenation (the hydroxo complex
2@C), and a specific catalyst for the Hþ/ electron transfer (the hydride 5@C).
MEAs were fabricated comprising of a nickel foam anode coated with 1@C
(ca. 1 mg cm2 Rh), a carbon paper cathode coated with a Fe-Co/C electrocatalyst
and a Tokuyama A-201 anion-exchange membrane. Figure 12.11a shows typical
polarization and power density curves for passive cells fueled with aqueous solu-
tions of ethylene glycol (EG), 1,2-propanediol (1,2-P), 1,3-propanediol (1,3-P), and
1,4-butanediol (1,4-B) in 2 M KOH at room temperature. The cell containing 1,4-B
exhibits the highest peak power density (8.5 mW cm2 at 0.14 V) compared to the
other diols. The power density supplied by these OMFCs increases with
cell temperature (Fig. 12.11b). Indeed, 12.4 mW cm2, 26.9 mW cm2,
24.3 mW cm2, and 42.2 mWcm 2, were obtained with EG, 1,2- P, 1,3-P, and
1,4-B, respectively at 80  C. To date this is the first report of fuel cells that are fed
with 1,2-P, 1,3-P, and 1,4-B.
Room temperature OMFCs were subjected to galvanostatic experiments
(20 mA cm2) and the fuel exhausts were quantitatively and qualitatively analyzed
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 505

Table 12.5 Electrooxidation selectivity data for OMFCs at 22  C and 80  C with various alcohol
fuels
Max power Galvanostatic
Temperature density Conversion time
Fuel ( C) (mW cm2) TON % (mmol) (20 mA cm2) Products
EG 22 2.7 86 38 % (3.2) 17 h 14 m 100 %
glycolate
1,2-P 22 3.1 92 50 % (3.4) 18 h 17 m 100 % lactate
1,3-P 22 5.9 173 83 % (6.1) 34 h 16 m 92 %
3-hydroxy-
propanate
8 % malonate
1,4-B 22 8.5 114 78 % (114) 28 h 38 m 79 %
4-hydroxy-
butanoate
21 % succinate
EG 80 12.4 329 28 % (12.1) 12 h 45 m 100 %
glycolate
1,2-P 80 26.9 318 34 % (12.0) 12 h 39 m 100 % lactate
1,3-P 80 24.3 445 45 % (16.3) 17 h 29 m 96 %
3-hydroxy-
propanate
4 % malonate
1,4-B 80 42.2 486 79 % (18.4) 22 h 11 m 74 %
4-hydroxy-
butanoate
26 % succinate
Reprinted from [31] with permission from John Wiley and Sons

by 13C{1H} NMR spectroscopy and HPLC (Table 12.5 summarizes the selectivity
data). The OMFCs fed with EG and 1,2-P, yielded selectively glycolate and lactate.
The galvanostatic experiment with 1,3-P gave 3-hydroxy-propanoate (92 %) and
malonate (8 %) with a total conversion of 83 %. Using 1,4-B, 4-hydroxy-butanoate
(79 %) and succinate (21 %) were obtained. In general this molecular anode
electrocatalyst was shown to convert short chain diols (e.g. EG and 1,2-P) into
the mono-carboxylate whereas diols with longer spacer chains produced also a
small amount of the di-carboxylate (e.g. 1,3-P and 1,4-B).

12.6 Conclusions and Future Perspectives

The recent flurry of activity in this field attests to this fact that the application of bio-
mass derived alcohols in DAFCs is an attractive target for researchers working
towards a system of sustainable energy production and distribution. The move to
study devices that work in alkaline media rather than under corrosive acidic
506 H.A. Miller et al.

conditions is advantageous as many more materials can be investigated and the


prohibitive costs associated with the use of large amounts of platinum based
catalysts can be avoided. The availability of commercial anion exchange mem-
branes together with highly active non precious metal cathode electrocatalysts
means that work can be focused on improving the anode electrocatalyst. This
chapter has shown that the most active metals for the alcohol oxidation reaction
are Pt, Pd and Au. Palladium stands out as the most active material. The addition of
a second transition metal (e.g. Ni, Pt, Cu, Co, Ru, In, Bi) or oxide (e.g. CeO2, TiO2,
SnO2) to Pd catalysts leads to dramatic improvements in activity. Researchers have
also manipulated the surface structure of Pd nanoparticles on an atomic level. A
high concentration of low coordinated surface atoms on high index faceted
nanoparticles have been prepared by both chemical and electrochemical means.
These faceted particles exhibit enhanced activity for alcohol oxidation. Important
advances in stability have been achieved by replacing carbon supports with inor-
ganic conductive oxides. The most remarkable results have been obtained by
employing Pd doped titania nanotube arrays (TNTA) where peak power densities
in DAFCs of up to 335 mW cm2 have been obtained with ethanol.
As the alcohol used as fuel becomes more complex (e.g. 1,2-propandiol and
1,4-butandiol) the electrooxidation pathways also become complicated with a wide
range of possible pathways and intermediates. Indeed, the most challenging aspect
of developing DAFCs as power sources is to obtain complete oxidation of the fuel
molecule to CO2, a feat yet to be achieved even for the simplest, ethanol. Taking
advantage of this aspect some investigators have worked at tuning selectivity of the
oxidation process toward the formation of a valuable partially oxidized intermedi-
ate (e.g. lactate from 1,2-propandiol). For this the most successful anode catalysts
have been organometallic complexes which are single-site molecular catalysts that
have been shown to selectively oxidize a range of polyalcohols to their mono-
carboxylates.
In summary, as a power source both catalyst stability and complete oxidation of
alcohol fuels to CO2 are essential in order for DAFCs to be exploited commercially.
As part of the bio-refinery concept i.e. cogeneration of electrical power and
industrially relevant intermediates from renewable bio-alcohols, DAFCs will
have an important role to play. However, in such devices better selectivity and
higher conversion is required from the anode electrocatalyst.

References

1. Brouzgou A, Podias A, Tsiakaras P (2013) PEMFCs and AEMFCs directly fed with ethanol:
a current status comparative review. J Appl Electrochem 43(2):119–136. doi:10.1007/
s10800-012-0513-2
2. Yu EH, Krewer U, Scott K (2010) Principles and materials aspects of direct alkaline alcohol
fuel cells. Energies 3(8):1499–1528. doi:10.3390/En3081499
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 507

3. Scordia D, Cosentino SL, Jeffries TW (2013) Effectiveness of dilute oxalic acid pretreatment
of Miscanthus  giganteus biomass for ethanol production. Biomass Bioenerg 59:540–548.
doi:10.1016/j.biombioe.2013.09.011
4. Shi AM, Du ZY, Ma XC, Cheng YL, Min M, Deng SB, Chen P, Li D, Ruan R (2013)
Production and evaluation of biodiesel and bioethanol from high oil corn using three
processing routes. Bioresour Technol 128:100–106. doi:10.1016/j.biortech.2012.10.007
5. Livshits V, Philosoph A, Peled E (2008) Direct ethylene glycol fuel-cell stack—study of
oxidation intermediate products. J Power Sources 178(2):687–691. doi:10.1016/j.jpowsour.
2007.07.054
6. Kaplan D, Burstein L, Rosenberg Y, Peled E (2011) Comparison of methanol and ethylene
glycol oxidation by alloy and core-shell platinum based catalysts. J Power Sources 196
(20):8286–8292. doi:10.1016/j.jpowsour.2011.06.023
7. Xin L, Zhang ZY, Wang ZC, Li WZ (2012) Simultaneous generation of mesoxalic acid and
electricity from glycerol on a gold anode catalyst in anion-exchange membrane fuel cells.
Chemcatchem 4(8):1105–1114. doi:10.1002/cctc.201200017
8. Marchionni A, Bevilacqua M, Bianchini C, Chen YX, Filippi J, Fornasiero P, Lavacchi A,
Miller H, Wang LQ, Vizza F (2013) Electrooxidation in alkaline media of ethylene glycol
and glycerol on Pd-(Ni-Zn)/C anodes in direct alcohol fuel cells. Chemsuschem 6
(3):390–390. doi:10.1002/cssc.201300154
9. Ji N, Zhang T, Zheng MY, Wang AQ, Wang H, Wang XD, Chen JGG (2008) Direct catalytic
conversion of cellulose into ethylene glycol using nickel-promoted tungsten carbide cata-
lysts. Angew Chem Int Ed 47(44):8510–8513. doi:10.1002/anie.200803233
10. Zheng MY, Wang AQ, Ji N, Pang JF, Wang XD, Zhang T (2010) Transition metal-tungsten
bimetallic catalysts for the conversion of cellulose into ethylene glycol. Chemsuschem 3
(1):63–66. doi:10.1002/cssc.200900197
11. Van Gerpen J (2005) Biodiesel processing and production. Fuel Process Technol 86
(10):1097–1107. doi:10.1016/j.fuproc.2004.11.005
12. Pimentel D, Patzek T (2005) Ethanol production using corn, switchgrass, and wood; biodiesel
production using soybean and sunflower. Nat Resour Res 14(1):65–76. doi:10.1007/s11053-
005-4679-8
13. Scordia D, Cosentino SL, Lee J-W, Jeffries TW (2011) Dilute oxalic acid pretreatment for
biorefining giant reed (Arundo donax L.). Biomass Bioenergy 35(7):3018–3024. http://dx.
doi.org/10.1016/j.biombioe.2011.03.046
14. Antolini E, Gonzalez ER (2010) Alkaline direct alcohol fuel cells. J Power Sources 195
(11):3431–3450. doi:10.1016/j.jpowsour.2009.11.145
15. Wang YJ, Qiao JL, Baker R, Zhang JJ (2013) Alkaline polymer electrolyte membranes for
fuel cell applications. Chem Soc Rev 42(13):5768–5787. doi:10.1039/C3cs60053j
16. Tarasevich MR, Korchagin OV, Kuzov AV (2013) Electrocatalysis of anodic oxidation of
ethanol. Russ Chem Rev 82(11):1047–1065. doi:10.1070/Rc2013v082n11abeh004276
17. Varcoe JR, Atanassov P, Dekel DR, Herring AM, Hickner MA, Kohl PA, Kucernak AR,
Mustain WE, Nijmeijer K, Scott K, Xu TW, Zhuang L (2014) Anion-exchange membranes in
electrochemical energy systems. Energy Environ Sci 7(10):3135–3191. doi:10.1039/
C4ee01303d
18. Bambagioni V, Bianchini C, Filippi J, Lavacchi A, Oberhauser W, Marchionni A, Moneti S,
Vizza F, Psaro R, Dal Santo V, Gallo A, Recchia S, Sordelli L (2011) Single-site and
nanosized Fe-Co electrocatalysts for oxygen reduction: synthesis, characterization and cata-
lytic performance. J Power Sources 196(5):2519–2529. doi:10.1016/j.jpowsour.2010.11.030
19. Miller HA, Bevilacqua M, Filippi J, Lavacchi A, Marchionni A, Marelli M, Moneti S,
Oberhauser W, Vesselli E, Innocenti M, Vizza F (2013) Nanostructured Fe-Ag
electrocatalysts for the oxygen reduction reaction in alkaline media. J Mater Chem A 1
(42):13337–13347. doi:10.1039/C3ta12757e
20. Kruusenberg I, Matisen L, Shah Q, Kannan AM, Tammeveski K (2012) Non-platinum
cathode catalysts for alkaline membrane fuel cells. Int J Hydrogen Energy 37
(5):4406–4412. doi:10.1016/j.ijhydene.2011.11.143
508 H.A. Miller et al.

21. Shen SY, Zhao TS, Wu QX (2012) Product analysis of the ethanol oxidation reaction on
palladium-based catalysts in an anion-exchange membrane fuel cell environment. Int J
Hydrogen Energy 37(1):575–582. doi:10.1016/j.ijhydene.2011.09.077
22. He QG, Shyam B, Macounova K, Krtil P, Ramaker D, Mukerjee S (2012) Dramatically
enhanced cleavage of the C-C bond using an electrocatalytically coupled reaction. J Am
Chem Soc 134(20):8655–8661. doi:10.1021/Ja301992h
23. Fang X, Wang LQ, Shen PK, Cui GF, Bianchini C (2010) An in situ Fourier transform
infrared spectroelectrochemical study on ethanol electrooxidation on Pd in alkaline solution.
J Power Sources 195(5):1375–1378. doi:10.1016/j.jpowsour.2009.09.025
24. Lamy C, Lima A, LeRhun V, Delime F, Coutanceau C, Leger JM (2002) Recent advances in
the development of direct alcohol fuel cells (DAFC). J Power Sources 105(2):283–296.
doi:10.1016/S0378-7753(01)00954-5, Pii: S0378-7753(01)00954-5
25. Gasteiger HA, Markovic N, Ross PN, Cairns EJ (1993) Methanol electrooxidation on well-
characterized Pt-Rn alloys. J Phys Chem 97(46):12020–12029. doi:10.1021/J100148a030
26. Lamy C, Lima A, LeRhun V, Delime F, Coutanceau C, Léger J-M (2002) Recent advances in
the development of direct alcohol fuel cells (DAFC). J Power Sources 105(2):283–296.
doi:10.1016/S0378-7753(01)00954-5
27. Liang ZX, Zhao TS, Xu JB, Zhu LD (2009) Mechanism study of the ethanol oxidation
reaction on palladium in alkaline media. Electrochim Acta 54(8):2203–2208. doi:10.1016/j.
electacta.2008.10.034
28. Bevilacqua M, Bianchini C, Marchionni A, Filippi J, Lavacchi A, Miller H, Oberhauser W,
Vizza F, Granozzi G, Artiglia L, Annen SP, Krumeich F, Grutzmacher H (2012) Improve-
ment in the efficiency of an organometallic fuel cell by tuning the molecular architecture of
the anode electrocatalyst and the nature of the carbon support. Energy Environ Sci 5
(9):8608–8620. doi:10.1039/C2ee22055e
29. Sheng T, Lin WF, Hardacre C, Hu P (2014) Role of water and adsorbed hydroxyls on ethanol
electrochemistry on Pd: new mechanism, active centers, and energetics for direct ethanol fuel
cell running in alkaline medium. J Phys Chem C 118(11):5762–5772. doi:10.1021/
Jp407978h
30. Yang YY, Ren J, Li QX, Zhou ZY, Sun SG, Cai WB (2014) Electrocatalysis of ethanol on a
Pd electrode in alkaline media: an in situ attenuated total reflection surface-enhanced infrared
absorption spectroscopy study. ACS Catal 4(3):798–803. doi:10.1021/Cs401198t
31. Bellini M, Bevilacqua M, Filippi J, Lavacchi A, Marchionni A, Miller HA, Oberhauser W,
Vizza F, Annen SP, Grutzmacher H (2014) Energy and chemicals from the selective
electrooxidation of renewable diols by organometallic fuel cells. Chemsuschem 7
(9):2432–2435. doi:10.1002/cssc.201402316
32. Marchionni A, Bevilacqua M, Bianchini C, Chen Y-X, Filippi J, Fornasiero P, Lavacchi A,
Miller H, Wang L, Vizza F (2013) Electrooxidation of ethylene glycol and glycerol on
Pd-(Ni-Zn)/C anodes in direct alcohol fuel cells. Chemsuschem 6(3):518–528. doi:10.1002/
cssc.201200866
33. Kwon Y, Schouten KJP, Koper MTM (2011) Mechanism of the catalytic oxidation of
glycerol on polycrystalline gold and platinum electrodes. Chemcatchem 3(7):1176–1185.
doi:10.1002/cctc.201100023
34. Bambagioni V, Bevilacqua M, Bianchini C, Filippi J, Marchionni A, Vizza F, Wang LQ,
Shen PK (2010) Ethylene glycol electrooxidation on smooth and nanostructured Pd elec-
trodes in alkaline media. Fuel Cells 10(4):582–590. doi:10.1002/fuce.200900120
35. Bambagioni V, Bianchini C, Filippi J, Oberhauser W, Marchionni A, Vizza F, Psaro R,
Sordelli L, Foresti ML, Innocenti M (2009) Ethanol oxidation on electrocatalysts obtained by
spontaneous deposition of palladium onto nickel-zinc materials. Chemsuschem 2(1):99–112.
doi:10.1002/cssc.200800188
36. Bianchini C, Shen PK (2009) Palladium-based electrocatalysts for alcohol oxidation in half
cells and in direct alcohol fuel cells. Chem Rev 109(9):4183–4206. doi:10.1021/cr9000995
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 509

37. Christensen PA, Hamnett A (1989) The oxidation of ethylene glycol at a platinum electrode
in acid and base: an in situ FTIR study. J Electroanal Chem Interfacial Electrochem 260
(2):347–359. http://dx.doi.org/10.1016/0022-0728(89)87149-9
38. Hahn F, Beden B, Kadirgan F, Lamy C (1987) Electrocatalytic oxidation of ethylene glycol:
Part III. In-situ infrared reflectance spectroscopic study of the strongly bound species
resulting from its chemisorption at a platinum electrode in aqueous medium. J Electroanal
Chem Interfacial Electrochem 216(1–2):169–180. http://dx.doi.org/10.1016/0022-0728(87)
80205-X
39. Santasalo-Aarnio A, Kwon Y, Ahlberg E, Kontturi K, Kallio T, Koper MTM (2011) Com-
parison of methanol, ethanol and iso-propanol oxidation on Pt and Pd electrodes in alkaline
media studied by HPLC. Electrochem Commun 13(5):466–469. doi:10.1016/j.elecom.2011.
02.022
40. An L, Zhao TS, Wu QX, Zeng L (2012) Comparison of different types of membrane in
alkaline direct ethanol fuel cells. Int J Hydrogen Energy 37(19):14536–14542. http://dx.doi.
org/10.1016/j.ijhydene.2012.06.105
41. An L, Zhao TS, Shen SY, Wu QX, Chen R (2010) Performance of a direct ethylene glycol
fuel cell with an anion-exchange membrane. Int J Hydrogen Energy 35(9):4329–4335. http://
dx.doi.org/10.1016/j.ijhydene.2010.02.009
42. An L, Zhao TS, Chen R, Wu QX (2011) A novel direct ethanol fuel cell with high power density.
J Power Sources 196(15):6219–6222. http://dx.doi.org/10.1016/j.jpowsour.2011.03.040
43. Li YS, Zhao TS, Liang ZX (2009) Performance of alkaline electrolyte-membrane-based
direct ethanol fuel cells. J Power Sources 187(2):387–392. http://dx.doi.org/10.1016/j.
jpowsour.2008.10.132
44. Lee J, Jeong B, Ocon JD (2013) Oxygen electrocatalysis in chemical energy conversion and
storage technologies. Curr Appl Phys 13(2):309–321. doi:10.1016/j.cap.2012.08.008
45. Othman R, Dicks AL, Zhu ZH (2012) Non precious metal catalysts for the PEM fuel cell
cathode. Int J Hydrogen Energ 37(1):357–372. doi:10.1016/j.ijhydene.2011.08.095
46. Brouzgou A, Song SQ, Tsiakaras P (2012) Low and non-platinum electrocatalysts for
PEMFCs: current status, challenges and prospects. Appl Catal B Environ 127:371–388.
doi:10.1016/j.apcatb.2012.08.031
47. Jaouen F, Proietti E, Lefevre M, Chenitz R, Dodelet JP, Wu G, Chung HT, Johnston CM,
Zelenay P (2011) Recent advances in non-precious metal catalysis for oxygen-reduction
reaction in polymer electrolyte fuel cells. Energy Environ Sci 4(1):114–130. doi:10.1039/
C0ee00011f
48. Guo JS, Zhou J, Chu D, Chen RR (2013) Tuning the electrochemical interface of Ag/C
electrodes in alkaline media with metallophthalocyanine molecules. J Phys Chem C 117
(8):4006–4017. doi:10.1021/Jp310655y
49. Ding L, Xin Q, Zhou XJ, Qiao JL, Li H, Wang HJ (2013) Electrochemical behavior of
nanostructured nickel phthalocyanine (NiPc/C) for oxygen reduction reaction in alkaline
media. J Appl Electrochem 43(1):43–51. doi:10.1007/s10800-012-0503-4
50. Lalande G, Faubert G, Cote R, Guay D, Dodelet JP, Weng LT, Bertrand P (1996) Catalytic
activity and stability of heat-treated iron phthalocyanines for the electroreduction of oxygen
in polymer electrolyte fuel cells. J Power Sources 61(1–2):227–237. doi:10.1016/S0378-7753
(96)02356-7
51. Ding L, Qiao JL, Dai XF, Zhang J, Zhang JJ, Tian BL (2012) Highly active electrocatalysts
for oxygen reduction from carbon-supported copper-phthalocyanine synthesized by high
temperature treatment. Int J Hydrogen Energy 37(19):14103–14113. doi:10.1016/j.
ijhydene.2012.07.046
52. He QG, Yang XF, Ren XM, Koel BE, Ramaswamy N, Mukerjee S, Kostecki R (2011) A
novel CuFe-based catalyst for the oxygen reduction reaction in alkaline media. J Power
Sources 196(18):7404–7410. doi:10.1016/j.jpowsour.2011.04.016
53. He QG, Yang XF, He RH, Bueno-Lopez A, Miller H, Ren XM, Yang WL, Koel BE (2012)
Electrochemical and spectroscopic study of novel Cu and Fe-based catalysts for oxygen
510 H.A. Miller et al.

reduction in alkaline media. J Power Sources 213:169–179. doi:10.1016/j.jpowsour.2012.04.


029
54. Piana M, Boccia M, Filpi A, Flammia E, Miller HA, Orsini M, Salusti F, Santiccioli S,
Ciardelli F, Pucci A (2010) H-2/air alkaline membrane fuel cell performance and durability,
using novel ionomer and non-platinum group metal cathode catalyst. J Power Sources 195
(18):5875–5881. doi:10.1016/j.jpowsour.2009.12.085
55. Lefevre M, Proietti E, Jaouen F, Dodelet JP (2009) Iron-based catalysts with improved
oxygen reduction activity in polymer electrolyte fuel cells. Science 324(5923):71–74.
doi:10.1126/science.1170051
56. Proietti E, Jaouen F, Lefevre M, Larouche N, Tian J, Herranz J, Dodelet JP (2011) Iron-based
cathode catalyst with enhanced power density in polymer electrolyte membrane fuel cells.
Nat Commun 2:416. doi:10.1038/Ncomms1427
57. Trogadas P, Fuller TF, Strasser P (2014) Carbon as catalyst and support for electrochemical
energy conversion. Carbon 75:5–42. doi:10.1016/j.carbon.2014.04.005
58. Yang Z, Nie HG, Chen X, Chen XH, Huang SM (2013) Recent progress in doped carbon
nanomaterials as effective cathode catalysts for fuel cell oxygen reduction reaction. J Power
Sources 236:238–249. doi:10.1016/j.jpowsour.2013.02.057
59. Gong KP, Du F, Xia ZH, Durstock M, Dai LM (2009) Nitrogen-doped carbon nanotube
arrays with high electrocatalytic activity for oxygen reduction. Science 323(5915):760–764.
doi:10.1126/science.1168049
60. Bambagioni V, Bianchini C, Chen YX, Filippi J, Fornasiero P, Innocenti M, Lavacchi A,
Marchionni A, Oberhauser W, Vizza F (2012) Energy efficiency enhancement of ethanol
electrooxidation on Pd-CeO2/C in passive and active polymer electrolyte-membrane fuel
cells. Chemsuschem 5(7):1266–1273. doi:10.1002/cssc.201100738
61. Chen Y, Bellini M, Bevilacqua M, Fornasiero P, Lavacchi A, Miller HA, Wang L, Vizza F
(2014) Direct alcohol fuel cells: toward the power densities of hydrogen-fed proton exchange
membrane fuel cells. Chemsuschem 8(3):524–533. doi:10.1002/cssc.201402999
62. Li YS, He YL (2014) Layer reduction method for fabricating Pd-coated Ni foams as high-
performance ethanol electrode for anion-exchange membrane fuel cells. RSC Adv 4
(32):16879–16884. doi:10.1039/C4ra01399a
63. Li YS, Zhao TS (2011) A high-performance integrated electrode for anion-exchange mem-
brane direct ethanol fuel cells. Int J Hydrogen Energy 36(13):7707–7713. doi:10.1016/j.
ijhydene.2011.03.090
64. Shen SY, Zhao TS, Xu JB, Li YS (2011) High performance of a carbon supported ternary
PdIrNi catalyst for ethanol electro-oxidation in anion-exchange membrane direct ethanol fuel
cells. Energy Environ Sci 4(4):1428–1433. doi:10.1039/C0ee00579g
65. Ma L, He H, Hsu A, Chen RR (2013) PdRu/C catalysts for ethanol oxidation in anion-
exchange membrane direct ethanol fuel cells. J Power Sources 241:696–702. doi:10.1016/j.
jpowsour.2013.04.051
66. Xin L, Zhang ZY, Qi J, Chadderdon D, Li WZ (2012) Electrocatalytic oxidation of ethylene
glycol (EG) on supported Pt and Au catalysts in alkaline media: reaction pathway investiga-
tion in three-electrode cell and fuel cell reactors. Appl Catal B Environ 125:85–94. doi:10.
1016/j.apcatb.2012.05.024
67. An L, Zeng L, Zhao TS (2013) An alkaline direct ethylene glycol fuel cell with an alkali-
doped polybenzimidazole membrane. Int J Hydrogen Energy 38(25):10602–10606. doi:10.
1016/j.ijhydene.2013.06.042
68. Qi J, Xin L, Zhang ZY, Sun K, He HY, Wang F, Chadderdon D, Qiu Y, Liang CH, Li WZ
(2013) Surface dealloyed PtCo nanoparticles supported on carbon nanotube: facile synthesis
and promising application for anion exchange membrane direct crude glycerol fuel cell.
Green Chem 15(5):1133–1137. doi:10.1039/C3gc36955b
69. Suleimanov NM, Khantimerov SM, Kukovitsky EF, Matukhin VL (2008) Electrooxidation
of ethanol on carbon nanotubes-nickel nanoparticles composites in alkaline media. J Solid
State Electrochem 12(7–8):1021–1023. doi:10.1007/s10008-008-0519-1
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 511

70. Xu CW, Hu YH, Rong JH, Jiang SP, Liu YL (2007) Ni hollow spheres as catalysts for
methanol and ethanol electrooxidation. Electrochem Commun 9(8):2009–2012. doi:10.1016/
j.elecom.2007.05.028
71. Jin GP, Baron R, Xiao L, Compton RG (2009) Ultrasonic synthesis of nickel nanostructures
on glassy carbon microspheres and their application for ethanol electrooxidation. J Nanosci
Nanotechnol 9(4):2719–2725. doi:10.1166/Jnn.2009.462
72. Yi Y, Uhm S, Lee J (2010) Electrocatalytic oxidation of ethanol on nanoporous Ni electrode
in alkaline media. Electrocatalysis 1(2–3):104–107. doi:10.1007/s12678-010-0015-0
73. Wang ZH, Du YL, Zhang FY, Zheng ZX, Zhang YZ, Wang CM (2013) High electrocatalytic
activity of non-noble Ni-Co/graphene catalyst for direct ethanol fuel cells. J Solid State
Electrochem 17(1):99–107. doi:10.1007/s10008-012-1855-8
74. Muench F, Oezaslan M, Rauber M, Kaserer S, Fuchs A, Mankel E, Brotz J, Strasser P,
Roth C, Ensinger W (2013) Electroless synthesis of nanostructured nickel and nickel-boron
tubes and their performance as unsupported ethanol electrooxidation catalysts. J Power
Sources 222:243–252. doi:10.1016/j.jpowsour.2012.08.067
75. Ma L, Chu D, Chen RR (2012) Comparison of ethanol electro-oxidation on Pt/C and Pd/C
catalysts in alkaline media. Int J Hydrogen Energy 37(15):11185–11194. doi:10.1016/j.
ijhydene.2012.04.132
76. Bianchini C, Bambagioni V, Filippi J, Marchionni A, Vizza F, Bert P, Tampucci A (2009)
Selective oxidation of ethanol to acetic acid in highly efficient polymer electrolyte
membrane-direct ethanol fuel cells. Electrochem Commun 11(5):1077–1080. doi:10.1016/j.
elecom.2009.03.022
77. Shen SY, Zhao TS, Xu JB, Li YS (2010) Synthesis of PdNi catalysts for the oxidation of
ethanol in alkaline direct ethanol fuel cells. J Power Sources 195(4):1001–1006. doi:10.1016/
j.jpowsour.2009.08.079
78. Liu XY, Zhang Y, Gong MX, Tang YW, Lu TH, Chen Y, Lee JM (2014) Facile synthesis of
corallite-like Pt-Pd alloy nanostructures and their enhanced catalytic activity and stability for
ethanol oxidation. J Mater Chem A 2(34):13840–13844. doi:10.1039/C4ta02522a
79. Jiang RZ, Tran DT, McClure JP, Chu D (2014) A class of (Pd-Ni-P) electrocatalysts for the
ethanol oxidation reaction in alkaline media. ACS Catal 4(8):2577–2586. doi:10.1021/
Cs500462z
80. Hong W, Wang J, Wang EK (2014) Facile synthesis of highly active PdAu nanowire
networks as self-supported electrocatalyst for ethanol electrooxidation. ACS Appl Mater
Interfaces 6(12):9481–9487. doi:10.1021/Am501859k
81. Geraldes AN, da Silva DF, Pino ES, da Silva JCM, de Souza RFB, Hammer P, Spinace EV,
Neto AO, Linardi M, dos Santos MC (2013) Ethanol electro-oxidation in an alkaline medium
using Pd/C, Au/C and PdAu/C electrocatalysts prepared by electron beam irradiation.
Electrochim Acta 111:455–465. doi:10.1016/j.electacta.2013.08.021
82. Feng YY, Liu ZH, Xu Y, Wang P, Wang WH, Kong DS (2013) Highly active PdAu alloy
catalysts for ethanol electro-oxidation. J Power Sources 232:99–105. doi:10.1016/j.jpowsour.
2013.01.013
83. Chen LY, Chen N, Hou Y, Wang ZC, Lv SH, Fujita T, Jiang JH, Hirata A, Chen MW (2013)
Geometrically controlled nanoporous PdAu bimetallic catalysts with tunable Pd/Au ratio for
direct ethanol fuel cells. ACS Catal 3(6):1220–1230. doi:10.1021/Cs400135k
84. Huang ZY, Zhou HH, Li CH, Zeng FY, Fu CP, Kuang YF (2012) Preparation of well-
dispersed PdAu bimetallic nanoparticles on reduced graphene oxide sheets with excellent
electrochemical activity for ethanol oxidation in alkaline media. J Mater Chem 22
(5):1781–1785. doi:10.1039/C1jm13024b
85. Nguyen ST, Law HM, Nguyen HT, Kristian N, Wang SY, Chan SH, Wang X (2009)
Enhancement effect of Ag for Pd/C towards the ethanol electro-oxidation in alkaline
media. Appl Catal B Environ 91(1-2):507–515. doi:10.1016/j.apcatb.2009.06.021
86. Hammer B, Norskov JK (2000) Theoretical surface science and catalysis—calculations and
concepts. Adv Catal 45:71–129
512 H.A. Miller et al.

87. Greeley J, Norskov JK (2005) A general scheme for the estimation of oxygen binding
energies on binary transition metal surface alloys. Surf Sci 592(1-3):104–111. doi:10.1016/
j.susc.2005.07.018
88. Greeley J, Norskov JK, Mavrikakis M (2002) Electronic structure and catalysis on metal
surfaces. Annu Rev Phys Chem 53:319–348. doi:10.1146/annurev.physchem.53.100301.
131630
89. Stamenkovic V, Mun BS, Mayrhofer KJJ, Ross PN, Markovic NM, Rossmeisl J, Greeley J,
Norskov JK (2006) Changing the activity of electrocatalysts for oxygen reduction by tuning
the surface electronic structure. Angew Chem Int Ed 45(18):2897–2901. doi:10.1002/anie.
200504386
90. Ruban A, Hammer B, Stoltze P, Skriver HL, Norskov JK (1997) Surface electronic structure
and reactivity of transition and noble metals. J Mol Catal A Chem 115(3):421–429. doi:10.
1016/S1381-1169(96)00348-2
91. He QG, Chen W, Mukerjee S, Chen SW, Laufek F (2009) Carbon-supported PdM (M ¼ Au
and Sn) nanocatalysts for the electrooxidation of ethanol in high pH media. J Power Sources
187(2):298–304. doi:10.1016/j.jpowsour.2008.11.065
92. Xia YN, Yang PD, Sun YG, Wu YY, Mayers B, Gates B, Yin YD, Kim F, Yan YQ (2003)
One-dimensional nanostructures: synthesis, characterization, and applications. Adv Mater 15
(5):353–389. doi:10.1002/adma.200390087
93. Zhang LG, Li N, Gao FM, Hou L, Xu ZM (2012) Insulin amyloid fibrils: an excellent
platform for controlled synthesis of ultrathin superlong platinum nanowires with high
electrocatalytic activity. J Am Chem Soc 134(28):11326–11329. doi:10.1021/Ja302959e
94. Wang YJ, Wilkinson DP, Zhang JJ (2011) Noncarbon support materials for polymer elec-
trolyte membrane fuel cell electrocatalysts. Chem Rev 111(12):7625–7651. doi:10.1021/
Cr100060r
95. Lavacchi A, Miller H, Vizza F (2013) Other support nanomaterials. Nanotechnol Electrocatal
Energy 170:145–187. doi:10.1007/978-1-4899-8059-5_6
96. Du WX, Mackenzie KE, Milano DF, Deskins NA, Su D, Teng XW (2012) Palladium-tin
alloyed catalysts for the ethanol oxidation reaction in an alkaline medium. ACS Catal 2
(2):287–297. doi:10.1021/Cs2005955
97. Wang H, Liu ZY, Ji S, Wang KL, Zhou TB, Wang RF (2013) Ethanol oxidation activity and
structure of carbon-supported Pt-modified PdSn-SnO2 influenced by different stabilizers.
Electrochim Acta 108:833–840. doi:10.1016/j.electacta.2013.07.061
98. Mao HM, Wang LL, Zhu PP, Xu QJ, Li QX (2014) Carbon-supported PdSn SnO2 catalyst for
ethanol electro-oxidation in alkaline media. Int J Hydrogen Energy 39(31):17583–17588.
doi:10.1016/j.ijhydene.2014.08.079
99. Wang RF, Liu ZY, Ma YJ, Wang H, Linkov V, Ji S (2013) Heterostructure core PdSn-SnO2
decorated by Pt as efficient electrocatalysts for ethanol oxidation. Int J Hydrogen Energy 38
(31):13604–13610. doi:10.1016/j.lihydene.2013.08.044
100. Martinez U, Serov A, Padilla M, Atanassov P (2014) Mechanistic insight into oxide-
promoted palladium catalysts for the electro-oxidation of ethanol. Chemsuschem 7
(8):2351–2357. doi:10.1002/cssc.201402062
101. Zhang QL, Zheng JN, Xu TQ, Wang AJ, Wei J, Chen JR, Feng JJ (2014) Simple one-pot
preparation of Pd-on-Cu nanocrystals supported on reduced graphene oxide for
enhanced ethanol electrooxidation. Electrochim Acta 132:551–560. doi:10.1016/j.electacta.
2014.03.159
102. Zhang MM, Xie JM, Sun Q, Yan ZX, Chen M, Jing JJ, Hossain AMS (2013) In situ synthesis
of palladium nanoparticle on functionalized graphene sheets at improved performance for
ethanol oxidation in alkaline media. Electrochim Acta 111:855–861. doi:10.1016/j.electacta.
2013.08.135
103. Gao LN, Yue WB, Tao SS, Fan LZ (2013) Novel strategy for preparation of graphene-Pd, Pt
composite, and its enhanced electrocatalytic activity for alcohol oxidation. Langmuir 29
(3):957–964. doi:10.1021/La303663x
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 513

104. Wang Y, Zhao Y, Yin J, Liu MC, Dong Q, Su YQ (2014) Synthesis and electrocatalytic
alcohol oxidation performance of Pd-Co bimetallic nanoparticles supported on graphene. Int
J Hydrogen Energy 39(3):1325–1335. doi:10.1016/j.ijhydene.2013.11.002
105. Dong Q, Zhao Y, Han X, Wang Y, Liu MC, Li Y (2014) Pd/Cu bimetallic nanoparticles
supported on graphene nanosheets: facile synthesis and application as novel electrocatalyst
for ethanol oxidation in alkaline media. Int J Hydrogen Energy 39(27):14669–14679. doi:10.
1016/j.ijhydene.2014.06.139
106. Ahmed MS, Jeon S (2014) Highly active graphene-supported NixPd100-x binary alloyed
catalysts for electro-oxidation of ethanol in an alkaline media. ACS Catal 4(6):1830–1837.
doi:10.1021/Cs500103a
107. Zhang MM, Yan ZX, Sun Q, Xie JM, Jing JJ (2012) Synthetic core-shell Ni@Pd
nanoparticles supported on graphene and used as an advanced nanoelectrocatalyst for meth-
anol oxidation. New J Chem 36(12):2533–2540. doi:10.1039/C2nj40651a
108. Cui Q, Chao SJ, Bai ZY, Yan HY, Wang K, Yang L (2014) Based on a new support for
synthesis of highly efficient palladium/hydroxyapatite catalyst for ethanol electrooxidation.
Electrochim Acta 132:31–36. doi:10.1016/j.electacta.2014.03.129
109. Bambagioni V, Bianchini C, Marchionni A, Filippi J, Vizza F, Teddy J, Serp P, Zhiani M
(2009) Pd and Pt-Ru anode electrocatalysts supported on multi-walled carbon nanotubes and
their use in passive and active direct alcohol fuel cells with an anion-exchange membrane
(alcohol ¼ methanol, ethanol, glycerol). J Power Sources 190(2):241–251. doi:10.1016/j.
jpowsour.2009.01.044
110. Machado BF, Marchionni A, Bacsa RR, Bellini M, Beausoleil J, Oberhauser W, Vizza F,
Serp P (2013) Synergistic effect between few layer graphene and carbon nanotube supports
for palladium catalyzing electrochemical oxidation of alcohols. J Energy Chem 22
(2):296–304
111. Serov A, Martinez U, Atanassov P (2013) Novel Pd-In catalysts for alcohols electrooxidation
in alkaline media. Electrochem Commun 34:185–188. doi:10.1016/j.elecom.2013.06.003
112. Zalineeva A, Serov A, Padilla M, Martinez U, Artyushkova K, Baranton S, Coutanceau C,
Atanassov PB (2014) Self-supported PdxBi Catalysts for the electrooxidation of glycerol in
alkaline media. J Am Chem Soc 136(10):3937–3945. doi:10.1021/Ja412429f
113. Cheng FL, Dai XC, Wang H, Jiang SP, Zhang M, Xu CW (2010) Synergistic effect of Pd-Au
bimetallic surfaces in Au-covered Pd nanowires studied for ethanol oxidation. Electrochim
Acta 55(7):2295–2298. doi:10.1016/j.electacta.2009.11.076
114. Cherevko S, Xing XL, Chung CH (2011) Pt and Pd decorated Au nanowires: extremely high
activity of ethanol oxidation in alkaline media. Electrochim Acta 56(16):5771–5775. doi:10.
1016/j.electacta.2011.04.052
115. Rodriguez P, Koper MTM (2014) Electrocatalysis on gold. Phys Chem Chem Phys 16
(27):13583–13594. doi:10.1039/C4cp00394b
116. de Lima RB, Varela H (2008) Catalytic oxidation of ethanol on gold electrode in alkaline
media. Gold Bull 41(1):15–22
117. Beyhan S, Uosaki K, Feliu JM, Herrero E (2013) Electrochemical and in situ FTIR studies of
ethanol adsorption and oxidation on gold single crystal electrodes in alkaline. J Electroanal
Chem 707:89–94. doi:10.1016/j.jelechem.2013.08.034
118. Pandey RK, Lakshminarayanan V (2012) Ethanol electrocatalysis on gold and conducting
polymer nanocomposites: a study of the kinetic parameters. Appl Catal B Environ
125:271–281. doi:10.1016/j.apcatb.2012.06.002
119. Kumar VL, Siddhardha RSS, Kaniyoor A, Podila R, Molli M, Kumar VSM,
Venkataramaniah K, Ramaprabhu S, Rao AM, Ramamurthy SS (2014) Gold decorated
graphene by laser ablation for efficient electrocatalytic oxidation of methanol and ethanol.
Electroanalysis 26(8):1850–1857. doi:10.1002/elan.201400244
120. Simoes M, Baranton S, Coutanceau C (2012) Electrochemical valorisation of glycerol.
Chemsuschem 5(11):2106–2124. doi:10.1002/cssc.201200335
121. Zhang ZY, Xin L, Qi J, Wang ZC, Li WZ (2012) Selective electro-conversion of glycerol to
glycolate on carbon nanotube supported gold catalyst. Green Chem 14(8):2150–2152. doi:10.
1039/C2gc35505a
514 H.A. Miller et al.

122. Zhang ZY, Xin L, Qi J, Wang ZC, Chadderdon D, Li WZ (2012) Simultaneous generation of
valuable chemicals and electricity from selective electrocatalytic oxidation of glycerol in
anion exchange membrane fuel cell reactors. Abstr Pap Am Chem S 244
123. Zhang ZY, Xin L, Qi J, Chadderdon DJ, Sun K, Warsko KM, Li WZ (2014) Selective electro-
oxidation of glycerol to tartronate or mesoxalate on Au nanoparticle catalyst via electrode
potential tuning in anion-exchange membrane electro-catalytic flow reactor. Appl Catal B
Environ 147:871–878. doi:10.1016/j.apcatb.2013.10.018
124. Zhang ZY, Xin L, Qi J, Chadderdon DJ, Li WZ (2013) Supported Pt, Pd and Au nanoparticle
anode catalysts for anion-exchange membrane fuel cells with glycerol and crude glycerol
fuels. Appl Catal B Environ 136:29–39. doi:10.1016/j.apcatb.2013.01.045
125. Zhang ZY, Xin L, Li WZ (2012) Supported gold nanoparticles as anode catalyst for anion-
exchange membrane-direct glycerol fuel cell (AEM-DGFC). Int J Hydrogen Energy 37
(11):9393–9401. doi:10.1016/j.ijhydene.2012.03.019
126. Qi J, Xin L, Chadderdon DJ, Qiu Y, Jiang YB, Benipal N, Liang CH, Li WZ (2014)
Electrocatalytic selective oxidation of glycerol to tartronate on Au/C anode catalysts in
anion exchange membrane fuel cells with electricity cogeneration. Appl Catal B Environ
154:360–368. doi:10.1016/j.apcatb.2014.02.040
127. Wang Q, Sun GQ, Jiang LH, Xin Q, Sun SG, Jiang YX, Chen SP, Jusys Z, Behm RJ (2007)
Adsorption and oxidation of ethanol on colloid-based Pt/C, PtRu/C and Pt3Sn/C catalysts: In
situ FTIR spectroscopy and on-line DEMS studies. Phys Chem Chem Phys 9(21):2686–2696.
doi:10.1039/B700676b
128. James DD, Pickup PG (2012) Measurement of carbon dioxide yields for ethanol oxidation by
operation of a direct ethanol fuel cell in crossover mode. Electrochim Acta 78:274–278.
doi:10.1016/j.electacta.2012.05.120
129. Tarnowski DJ, Korzeniewski C (1997) Effects of surface step density on the electrochemical
oxidation of ethanol to acetic acid. J Phys Chem B 101(2):253–258. doi:10.1021/Jp962450c
130. Sun SG, Lin Y (1998) Kinetics of isopropanol oxidation on Pt(111), Pt(110), Pt(100), Pt(610)
and Pt(211) single crystal electrodes—studies of in situ time-resolved FTIR spectroscopy.
Electrochim Acta 44(6–7):1153–1162. doi:10.1016/S0013-4686(98)00218-7
131. Zhang ZC, Nosheen F, Zhang JC, Yang Y, Wang PP, Zhuang J, Wang X (2013) Growth of
concave polyhedral Pd nanocrystals with 32 facets through in situ facet-selective etching.
Chemsuschem 6(10):1893–1897. doi:10.1002/cssc.201300346
132. Lee YW, Kim D, Hong JW, Kang SW, Lee SB, Han SW (2013) Kinetically controlled growth
of polyhedral bimetallic alloy nanocrystals exclusively bound by high-index facets: AuPd
hexoctahedra. Small 9(5):660–665. doi:10.1002/smll.201201813
133. Bell AT (2003) The impact of nanoscience on heterogeneous catalysis. Science 299
(5613):1688–1691. doi:10.1126/science.1083671
134. Shao MH, Peles A, Shoemaker K (2011) Electrocatalysis on platinum nanoparticles: particle
size effect on oxygen reduction reaction activity. Nano Lett 11(9):3714–3719. doi:10.1021/
Nl2017459
135. Henning AM, Watt J, Miedziak PJ, Cheong S, Santonastaso M, Song MH, Takeda Y,
Kirkland AI, Taylor SH, Tilley RD (2013) Gold-palladium core-shell nanocrystals with
size and shape control optimized for catalytic performance. Angew Chem Int Ed 52
(5):1477–1480. doi:10.1002/anie.201207824
136. Fashedemi OO, Julies B, Ozoemena KI (2013) Synthesis of Pd-coated FeCo@Fe/C core–
shell nanoparticles: microwave-induced ‘top-down’ nanostructuring and decoration. Chem
Commun 49:2034–2036
137. Fashedemi OO, Ozoemena KI (2014) Comparative electrocatalytic oxidation of ethanol,
ethylene glycol and glycerol in alkaline medium at Pd-decorated FeCo@Fe/C core-shell
nanocatalysts. Electrochim Acta 128:279–286. doi:10.1016/j.electacta.2013.10.194
138. Tian N, Zhou ZY, Sun SG, Ding Y, Wang ZL (2007) Synthesis of tetrahexahedral platinum
nanocrystals with high-index facets and high electro-oxidation activity. Science 316
(5825):732–735. doi:10.1126/science.1140484
12 Direct Alcohol Fuel Cells: Nanostructured Materials for the Electrooxidation. . . 515

139. Ding Y, Gao Y, Wanga ZL, Tian N, Zhou ZY, Sun SG (2007) Facets and surface relaxation of
tetrahexahedral platinum nanocrystals. Appl Phys Lett 91(12):121901. doi:10.1063/1.
2785953
140. Zhou ZY, Huang ZZ, Chen DJ, Wang Q, Tian N, Sun SG (2010) High-index faceted platinum
nanocrystals supported on carbon black as highly efficient catalysts for ethanol
electrooxidation. Angew Chem Int Ed 49(2):411–414. doi:10.1002/anie.200905413
141. Zhou ZY, Shang SJ, Tian N, Wu BH, Zheng NF, Xu BB, Chen C, Wang HH, Xiang DM, Sun
SG (2012) Shape transformation from Pt nanocubes to tetrahexahedra with size near 10 nm.
Electrochem Commun 22:61–64. doi:10.1016/j.elecom.2012.05.023
142. Tian N, Zhou ZY, Yu NF, Wang LY, Sun SG (2010) Direct electrodeposition of
tetrahexahedral Pd nanocrystals with high-index facets and high catalytic activity for ethanol
electrooxidation. J Am Chem Soc 132(22):7580–7581. doi:10.1021/Ja102177r
143. Chen YX, Lavacchi A, Chen SP, di Benedetto F, Bevilacqua M, Bianchini C, Fornasiero P,
Innocenti M, Marelli M, Oberhauser W, Sun SG, Vizza F (2012) Electrochemical milling and
faceting: size reduction and catalytic activation of palladium nanoparticles. Angew Chem Int
Ed 51(34):8500–8504. doi:10.1002/anie.201203589
144. Wang L, Bevilacqua M, Chen Y-X, Filippi J, Innocenti M, Lavacchi A, Marchionni A,
Miller H, Vizza F (2013) Enhanced electro-oxidation of alcohols at electrochemically treated
polycrystalline palladium surface. J Power Sources 242:872–876
145. Wang LQ, Bevilacqua M, Filippi J, Fornasiero P, Innocenti M, Lavacchi A, Marchionni A,
Miller HA, Vizza F (2015) Electrochemical growth of platinum nanostructures for enhanced
ethanol oxidation. Appl Catal Environ 165:185–191. doi:10.1016/j.apcatb.2014.10.009
146. Wang L, Bambagioni V, Bevilacqua M, Bianchini C, Filippi J, Lavacchi A, Marchionni A,
Vizza F, Fang X, Shen PK (2010) Sodium borohydride as an additive to enhance the
performance of direct ethanol fuel cells. J Power Sources 195(24):8036–8043. doi:10.1016/
j.jpowsour.2010.06.101
147. Kocha SS (2012) Electrochemical degradation: electrocatalyst and support durability. In:
Polymer electrolyte fuel cell degradation. Academic, Boston, pp 89–214
148. Chen ZW, Waje M, Li WZ, Yan YS (2007) Supportless Pt and PtPd nanotubes as
electrocatalysts for oxygen-reduction reactions. Angew Chem Int Ed 46(22):4060–4063.
doi:10.1002/anie.200700894
149. Borup R, Meyers J, Pivovar B, Kim YS, Mukundan R, Garland N, Myers D, Wilson M,
Garzon F, Wood D, Zelenay P, More K, Stroh K, Zawodzinski T, Boncella J, McGrath JE,
Inaba M, Miyatake K, Hori M, Ota K, Ogumi Z, Miyata S, Nishikata A, Siroma Z,
Uchimoto Y, Yasuda K, Kimijima KI, Iwashita N (2007) Scientific aspects of polymer
electrolyte fuel cell durability and degradation. Chem Rev 107(10):3904–3951. doi:10.
1021/Cr050182l
150. Shao-Horn Y, Sheng WC, Chen S, Ferreira PJ, Holby EF, Morgan D (2007) Instability of
supported platinum nanoparticles in low-temperature fuel cells. Top Catal 46(3-4):285–305.
doi:10.1007/s11244-007-9000-0
151. Cao MN, Wu DS, Cao R (2014) Recent advances in the stabilization of platinum electrocatalysts
for fuel-cell reactions. Chemcatchem 6(1):26–45. doi:10.1002/cctc.201300647
152. Nguyen ST, Yang YH, Wang X (2012) Ethanol electro-oxidation activity of Nb-doped-TiO2
supported PdAg catalysts in alkaline media. Appl Catal B Environ 113:261–270. doi:10.
1016/j.apcatb.2011.11.046
153. Wang AL, Xu H, Feng JX, Ding LX, Tong YX, Li GR (2013) Design of Pd/PANI/Pd
sandwich-structured nanotube array catalysts with special shape effects and synergistic
effects for ethanol electrooxidation. J Am Chem Soc 135(29):10703–10709. doi:10.1021/
Ja403101r
154. Chen YX, Lavacchi A, Miller HA, Bevilacqua M, Filippi J, Innocenti M, Marchionni A,
Oberhauser W, Wang L, Vizza F (2014) Nanotechnology makes biomass electrolysis more
energy efficient than water electrolysis. Nat Commun 5:4036. doi:10.1038/ncomms5036
516 H.A. Miller et al.

155. Antolini E (2007) Catalysts for direct ethanol fuel cells. J Power Sources 170(1):1–12. doi:10.
1016/j.jpowsour.2007.04.009
156. Matsuoka K, Iriyama Y, Abe T, Matsuoka M, Ogumi Z (2005) Alkaline direct alcohol fuel
cells using an anion exchange membrane. J Power Sources 150:27–31. doi:10.1016/j.
jpowsour.2005.02.020
157. Vigier F, Rousseau S, Coutanceau C, Leger JM, Lamy C (2006) Electrocatalysis for the direct
alcohol fuel cell. Top Catal 40(1–4):111–121. doi:10.1007/s11244-006-0113-7
158. Cracknell JA, Vincent KA, Armstrong FA (2008) Enzymes as working or inspirational
electrocatalysts for fuel cells and electrolysis. Chem Rev 108(7):2439–2461. doi:10.1021/
cr0680639
159. Vincent KA, Barton SC, Canters GW, Heering HA (2009) Electrocatalysis for fuel cell at
enzyme-modified electrodes. In: Koper MTM (ed) Fuel cell catalysis. Wiley, New York.
doi:10.1002/9780470463772.ch17
160. Yamazaki S, Yao M, Fujiwara N, Siroma Z, Yasuda K, Ioroi T (2012) Electrocatalytic
oxidation of alcohols by a carbon-supported Rh porphyrin. Chem Commun 48
(36):4353–4355. doi:10.1039/C2cc30888f
161. Brownell KR, McCrory CCL, Chidsey CED, Perry RH, Zare RN, Waymouth RM (2013)
Electrooxidation of alcohols catalyzed by amino alcohol ligated ruthenium complexes. J Am
Chem Soc 135(38):14299–14305. doi:10.1021/Ja4055564
162. Weiss CJ, Das P, Miller DL, Helm ML, Appel AM (2014) Catalytic oxidation of alcohol via
nickel phosphine complexes with pendant amines. ACS Catal 4(9):2951–2958. doi:10.1021/
Cs500853f
163. Elouarzaki K, Haddad R, Holzinger M, Le Goff A, Thery J, Cosnier S (2014) MWCNT-
supported phthalocyanine cobalt as air-breathing cathodic catalyst in glucose/O-2 fuel cells. J
Power Sources 255:24–28. doi:10.1016/j.jpowsour.2013.12.109
164. Elouarzaki K, Le Goff A, Holzinger M, Thery J, Cosnier S (2012) Electrocatalytic oxidation
of glucose by rhodium porphyrin-functionalized MWCNT electrodes: application to a fully
molecular catalyst-based glucose/O-2 fuel cell. J Am Chem Soc 134(34):14078–14085.
doi:10.1021/Ja304589m
165. Annen SP, Bambagioni V, Bevilacqua M, Filippi J, Marchionni A, Oberhauser W,
Schonberg H, Vizza F, Bianchini C, Grutzmacher H (2010) A biologically inspired organo-
metallic fuel cell (OMFC) that converts renewable alcohols into energy and chemicals.
Angew Chem Int Ed 49(40):7229–7233. doi:10.1002/anie.201002234
Chapter 13
Effects of Catalyst-Support Materials
on the Performance of Fuel Cells

Paul M. Ejikeme, Katlego Makgopa, and Kenneth I. Ozoemena

13.1 Introduction

In the twenty-first century, the energy crisis and environmental pollution are both
stark challenges faced by human society because of speedily rising energy demand
and enormous combustion of fossil derived fuels. It is generally believed that the
conventional notions of energy generation and conversion are not suitable for the
world’s sustainable development [1, 2]. Renewable energy technologies therefore
hold the promise to meet the increasing energy demands of over a seven billion
people planet without depleting our natural resources and compromising the quality
of our environment [3]. Recently, the emerging need for clean energy and high
speed electronic has motivated researchers across the globe to discover, develop,
and assemble new classes of nanomaterials in unconventional device architectures.
A great deal of research has thus been carried out to explore clean and renewable
energy sources as well as their devices [4, 5]. Among the different energy systems,
hydrogen and DAFCs involving mainly methanol and ethanol have been

P.M. Ejikeme (*)


Department of Chemistry, University of Pretoria, Pretoria 0002, South Africa
Department of Pure and Industrial Chemistry, University of Nigeria, Nsukka 410001, Nigeria
e-mail: [email protected]
K. Makgopa
Department of Chemistry, University of Pretoria, Pretoria 0002, South Africa
K.I. Ozoemena (*)
Council for Scientific and Industrial Research, Pretoria, South Africa
Energy Materials Unit, Materials Science and Manufacturing, Council for Scientific
and Industrial Research (CSIR), Pretoria 0001, South Africa
Molecular Sciences Institute, School of Chemistry, University of the Witwatersrand,
Johannesburg 2050, South Africa
e-mail: [email protected]

© Springer International Publishing Switzerland 2016 517


K.I. Ozoemena, S. Chen (eds.), Nanomaterials for Fuel Cell Catalysis,
Nanostructure Science and Technology, DOI 10.1007/978-3-319-29930-3_13
518 P.M. Ejikeme et al.

extensively studied as ideal energy converters that convert chemical energy of


alcohols and an oxidant directly to electrochemical energy [6, 7].
Fuel cells (FCs) are ‘electrochemical’ devices that convert directly the chemical
energy / heat of combustion of a fuel (e.g., hydrogen, methanol, ethanol, natural
gas, hydrocarbons, etc.) and an oxidant (i.e., air or pure oxygen) in the presence of a
catalyst into electricity, heat and water [8–10], at an efficiency that is typically 2–3
times that of an internal combustion engine [9]. A German scientist, Christian
Friedrich Schoenbein (1799–1868), first discovered the fuel cell effect in January
1839, and reported observing electric current induced by the combination of
hydrogen and oxygen, (i.e., the inverse electrolysis process). The best known
early fuel cell experiments were however performed in 1842 by the British physicist
and lawyer, Sir William R. Grove (1811–1896), who produced an electric current
by connecting a hydrogen anode and an oxygen cathode [11, 12]. Fuel cell
technologies are only classed as renewable energy systems if the fuel used is
renewable (i.e., biofuels, hydrogen from wind or solar conversion of water)
[3]. They show promise as high-efficient energy conversion technology for both
mobile and stationary power applications, and have shown even greater promise as
a substitute to the conventional combustion engine for transportation applications
[13]. The FC process, just as any other redox process, involves electron transfer
during the oxidation and reduction reactions with an essentially invariant electrode-
electrolyte system.

13.1.1 Classification of Fuel Cells

Fuel cells are generally classified into six (6) broad groups based on the choice of
operating fuel and electrolyte [14] as shown in Fig. 13.1 and described elsewhere in
this book. Alkaline fuel cells (AFC) utilise an aqueous potassium hydroxide (KOH)
solution as the alkaline electrolyte while proton exchange membrane fuel cells
(PEMFC) and direct methanol fuel cells (DMFC) use polymer membranes as the
electrolyte for conducting protons. Furthermore, phosphoric acid fuel cells (PAFC)
use pure phosphoric acid as the electrolyte, molten carbonate fuel cells (MCFC) use
a molten mixture of lithium, sodium, and potassium carbonates for conducting
carbonate ions while solid oxide fuel cells (SOFC) make use of ceramic (solid
oxide) material as electrolytes, to conduct negative oxygen ions from the cathode to
the anode. It is also common to classify fuel cells according to their operating
temperature regimes (Fig. 13.1). Thus, low temperature (i.e., AFC, PEMFC,
DMFC); medium temperature (PAFC); and high-temperature (i.e., MCFC,
DCFC, and SOFC) fuel cells are also known.
Today, the low-temperature fuel cells (PEMFC and AFC) shown in Fig. 13.2 are
quite popular. While the PEMFC operates in acidic conditions and uses proton-
exchange membranes (PEM) and platinum catalysts, the AFC operates in alkaline
media with anion-exchange membrane (AEM) and uses low-Pt or Pt-free catalysts.
DIRECT METHANOL O2
H+
CH3OH 50 −
FUEL CELL DMFC
CO2 H2O 120 °C

POLYMER ELECTROLYTE O2
MEMBRANE FC PEM H2 H+ 80 °C
H2O
ALKALINE
OH-
H2 90 −
FUEL CELL AFC O2
H2O 100 °C

PHOSPHORICACID O2
H+
100 −
FUEL CELL PAFC H2 250 °C
H2O
MOLTEN CARBONATE H2 O2 600 −
FUEL CELL MCFC
H2O
CO3-2 700 °C
CO2
SOLID OXIDE H2
O2
700 −
FUEL CELL SOFC O2
H2 O 1000 °C

FUEL OXYGEN

ELECTROLYTE
ANODE CATHODE

Fig. 13.1 Schematic representation of different types of fuel cells, electrode reactions and
products, and operating temperatures (Source: http://www.fuelcells.org/uploads/FuelCell Types.
jpg, Accessed on 13th June, 2015)

PEM FUEL CELL ALKALINE FUEL CELL


Electrical Current Electrical Current

Excess e- e- Water and e- e-


Fuel Heat Out Hydrogen In Oxygen In

e- H
2 e- O
2
e-
H+
H O
e-
2
H+ e- e-
H e-
2 H+ OH-
O e-
2 H O
H+ 2
e-
Water and e-
Heat Out e-
e-
Fuel In Air In
Anode Cathode Anode Cathode
Electrolyte Electrolyte

Fig. 13.2 Schematic representation of PEMFC and AFC (Source: Green Car Congress, Accessed
on 13 June 2015, http://www.greencarcongress.com/2005/05/zap_and_apollo_.html)
520 P.M. Ejikeme et al.

Fig. 13.3 Schematic representation of a direct methanol fuel cell (DMFC) operating in (a) acid
and (b) alkaline conditions (Source: http://clipart-finder.com/oxygen-clipart.html, Accessed 13th
June 2015)

A growing area of fuel cell research and development is the direct alcohol fuel
cell wherein hydrogen in PEMFC and AFC is replaced with alcohol (such as
methanol, ethanol, ethylene glycol and glycerol). Alcohols are considered alterna-
tives to hydrogen since the power density of the alcohols (liquids with no pressure
requirement during storage) in terms of energy by volume of fuel is much higher
than that of hydrogen at standard conditions [15]. Figure 13.3 shows a generalised
DAFC scheme using methanol as fuel. DAFCs (especially alkaline-based systems
involving higher alcohols) yield products that are valuable intermediates for the fine
chemical, pharmaceutical, and agrochemical sectors.

13.2 Catalyst-Supports for Fuel Cells

Electrocatalysts are the main drivers of the reactions that occur at the anode and
cathode sides of a fuel cell. The performance of each of the electrode reaction is
predominantly dependent on the specific surface area of the catalyst (SSA, i.e., real
catalyst surface area per gram). To increase the value of the SSA, the catalysts must
be dispersed as nanoparticles (2–5 nm in diameter) on a high-surface-area support
(10–50 nm in diameter). An ideal support for the electrocatalyst should, amongst
other considerations, possess the following properties [6, 14]:
(i)a good electrical conductivity,
(ii)good catalyst–support interaction,
(iii)a large surface area,
(iv) a mesoporous structure enabling the ionomer and polymer electrolyte to bring
the catalyst nanoparticles close to the reactants (suitable porosity to allow
good reactant and product flux), i.e. to maximise the triple-phase boundary
(TPB),
(v) a good water handling capability to avoid flooding,
13 Effects of Catalyst-Support Materials on the Performance of Fuel Cells 521

(vi) good corrosion resistance/high stability in fuel cell environments, and


(vii) ease of catalyst recovery.
It has been recognized [14] that a good interaction between the catalyst and the
support not only improves catalysts’ effectiveness while decreasing catalyst loss,
but also governs charge transfer, assist in sufficiently enhancing the catalyst
performance and robustness by reducing catalyst deactivation, and in some cases
affects the catalyst particle size. From the foregoing, the choice of support material
is vital in shaping the behaviour, performance, durability and cost effectiveness of
the catalyst in particular, and the overall fuel cell performance in general [15].

13.2.1 Carbon-Based Supports

The reputation of active carbon as a precious metal support in industrial chemistry


is well documented. The existence of carbon in various allotropes [16] (Fig. 13.4),
coupled with the established advantages of graphitized carbon as support material,

Fig. 13.4 Naturally existing and man-made carbon allotropes (Source: Li et al., Chem. Soc. Rev.,
2014, 43, 2572–2586, Reproduced with permission from [16], copyright 2014 Royal Society of
Chemistry)
522 P.M. Ejikeme et al.

Fig. 13.5 SEM images of USP porous carbons. Reaction conditions: 1.5 M solutions, 700  C, Ar
at 1.0 standard litre per minute (slpm). Product from (a) lithium chloroacetate (b) sodium
chloroacetate (c) potassium chloroacetate (d) lithium dichloroacetate (e) sodium dichloroacetate
and (f) potassium dichloroacetate (Source: Skrabalak and Suslick, J. Am. Chem. Soc., 2006,
128, 12642–12643. Reproduced with permission from [17], copyright 2006 American Chemical
Society)

with different modifications available, has however continued to make it a choice


material for the support of FC catalysts.
Depending on the synthesis protocol adopted, carbon supports come in different
morphologies (Fig. 13.5a–f) [17] ranging from core-shells, spheres, hollow spheres,
nanotubes, onion-like, to mention but a few. Generally, they can either be
mesoporous (pore diameters between 2 and 50 nm) or microporous (pore diameters
of less than 2 nm) or macroporous (pore diameters of greater than 50 nm) or a
mixture of the three types of pore sizes. Figure 13.5 exemplifies the presence of
hollow core-shell spheres (Fig. 13.5a), mesoporous (Fig. 13.5d) and macroporous
carbon networks (Fig. 13.5b, c, e, f). Carbons are stable in both acidic and basic
media and can easily be burnt off; a simple technique for recovering the metal
catalysts from the carbon support. Supported catalysts are of special interest as they
allow for the fine dispersion and stabilisation of small metallic particles. They
provide access to a much larger number of catalytically active atoms than in the
corresponding bulk metal, even when the latter is ground to a fine powder [18, 19].
Carbons remain the most important support material for electrocatalysts used in
fuel cells. The importance of carbon as support material stems from its high
availability and low-cost. However, carbon suffers shortcomings that limit its
rate performance as catalyst-support, which include:
13 Effects of Catalyst-Support Materials on the Performance of Fuel Cells 523

(i) Severe corrosion/oxidation with use under normal operating conditions (dur-
ing fuel cell start-up and shut-down) at high potentials (>0.7 V vs. NHE),
resulting in poor durability of the electrocatalyst because of dissolution,
Ostwald ripening, and aggregation [14, 20, 21];
(ii) Presence of high amount of micropores (<1 nm), which can impede fuel
supply to the surface, and present low accessible surface area for the deposi-
tion of metal particles;
(iii) Low polarity and high hydrophobicity, which impede permeability of gases
and liquids;
(iv) Poor stability at temperatures higher than 373 K and lack of proton conduc-
tivity [22, 23].
The degradation mechanism of carbon-supported platinum catalyst in ORR was
investigated in a very recent study by Zhang et al. [24]. From a combination of
methods based on accelerated stress test (AST) protocols, the authors observed
activity loss in electrochemical active surface area (ECSA) and ORR which they
attributed to the Pt dissolution/redeposition, agglomeration, detachment and carbon
corrosion.
The corrosion reaction of carbon material in aqueous acid electrolytes including
proton exchange membranes is generalized as follows [25, 26]:

C þ 2H2 O ! CO2 þ 4Hþ þ 4e E ¼ 0:207 V ðvs SHEÞ ð13:1Þ

Equation 13.1 indicates that carbon can be thermodynamically oxidized at poten-


tials above 0.207 V. Although this reaction is thermodynamically feasible at the
potentials which the fuel cell cathode operates, it is believed to be almost
non-existent or negligible in low-temperature fuel cells (60–90  C). However,
despite the sluggishness of the reaction, long-term operations can cause a decrease
in carbon content in the catalyst layers [27]. Furthermore, under certain circum-
stances when electrode potentials are raised extremely high, there is a rapid
degradation of carbon supports as well as the catalyst. Electrochemical corrosion
of carbon materials as catalyst-supports of PEMFCs leads to two major effects:
(i) electrical isolation of the catalyst particles as they are separated from the support
and (ii) aggregation of catalyst small particles (Ostwald ripening) and consequently
to the evolution of an inhomogenous structure over time (Fig. 13.6) [28].
This stems from the fact that molecules on the surface of a particle are energet-
ically less stable than the ones already well ordered and packed in the interior.
These dual effects result in a decrease in the electrochemical active surface area
(EASA) of the catalyst and an increase in the hydrophilicity of the surface. This can,
in turn, result in a decrease in gas permeability as the pores become more likely to
be filled with liquid water films that can hinder gas transport [25].
This problem associated with catalyst-supports has been extensively studied by
several workers [29–31]. For example, Mathias et al. [29] subjected MEA cathodes
to accelerated carbon support corrosion at 1.2 V and 80  C at various times and
demonstrated that carbon weight losses between 5 and 10 % occurred (Fig. 13.7),
524 P.M. Ejikeme et al.

Fig. 13.6 Schematic representation of the effect of corrosion of carbon on (1) agglomeration, (2)
coalescence, and (3) loss of catalyst particles in the membrane electrode assembly (MEA) during
operation of PEMFCs: (a) normal (corrosion-resistant) electrode and (b) corroded electrode
(Source: Huang et al., J. Am. Chem. Soc., 2009, 131, 13898–13899. Reproduced with permission
from [28], copyright 2009 American Chemical Society)

Fig. 13.7 Impact of accelerated carbon support corrosion on the E/I performance curves of 50 cm2
MEAs with both a ~50 % Pt/C standard and a  30 wt% Pt alloy/Ccorr.-resist. cathode catalyst
(Source: Matthias et al., The Electrochemical Society Interface, 2005, 14(3), 24–35. Reproduced
with permission from [29], copyright 2005 The Electrochemical Society)
13 Effects of Catalyst-Support Materials on the Performance of Fuel Cells 525

Fig. 13.8 Characterization of MEAs operated under “real-life” PEMFC conditions: (a) Raman
spectra, (b) scanning electron microscopy (SEM) and transmission electron microscopy (TEM)
images, and (c) XPS-derived data. PEMFC operating conditions: j ¼ 0.25 A cm2,
0.65 < Ecell < 0.77 V, 333 < T < 338 K, t ¼ 12,860 h, 250 start/stop events (Source: Castanheira
et al., ACS Catal., 2014, 4, 2258–2267. Reproduced with permission from [31], copyright 2014
American Chemical Society)

leading to unacceptably large performance deterioration under studied operating


conditions; implying that standard carbon supports (Vulcan-XC72 and Ketjen
black) do not satisfy automotive requirements with respect to either start-up /
shut-down cycles or prolonged idle conditions. Similar findings were reported by
Wang et al. [30] in their study with graphene. In a related study [31], it was
demonstrated that the disordered domains of the carbon supports were preferen-
tially oxidized at voltages related to the PEMFC cathode (0.40 < E < 1.0 V).
Experiments at electrode potentials greater than 1.0 V (obtained during start-up,
fuel starvation events and shut-down of PEMFC systems) were reported [31] to
have accelerated this degradation process (Fig. 13.8). The decrease of the D1-band
intensity and the thinning of the G- and D2-bands at 1610 cm1 (Fig. 13.8a)
confirmed the preferential degradation of the disordered graphite domains of the
supports while the emergence of the carbonyl band at 1760 cm1 showed that the
ordered graphite crystallites were also corroded. The SEM images of the MEA
(Fig. 13.8b) show that the cathode thickness estimated from SEM images decreased
(Fig. 13.8b, yellow arrows) from 13.1 μm at the fresh state to 3.8 μm after 12,860 h
of operation while the TEM images revealed an increased fraction of agglomerated
Pt nanoparticles, as well as changes of the structure of the carbon particles. The
larger value of the O1s/C1s ratio indicates that, on the average, the surface was
more oxidized after 12,860 h of operation than at the fresh state.
526 P.M. Ejikeme et al.

13.2.1.1 Physico-Chemical Properties of Carbon Supports

The performance of carbon supports is dependent on their unique physico-chemical


properties such as size, shape, pore structure/distribution, surface defects and
chemical properties. Several types of carbons have been studied as catalyst-
supports for fuel cell and they include carbon blacks, carbon nanotubes (CNTs),
carbon microspheres, carbon nanofibers and graphenes. The shape and structure of
carbons are some of the factors that can impact on their ability to strongly adsorb
catalysts on their surfaces for enhanced fuel cell performance. For example, Cuong
et al. [32] used density functional theory to unravel the interplay between the
carbon supports (i.e., graphene sheet, a metallic single wall carbon nanotube
(SWCNT) and a series of semiconducting SWCNTs) and Pt nano-clusters in
determining the stability and electronic properties of Pt nanoclusters. They found
that the Pt cluster could be best stabilized by adsorption on carbon nanotube
supports than on graphene support, and the optimal curvature of SWCNTs for
adsorption can be found. Considering that the dissolution and/or loss of
electrocatalyst’s surface area over a period of time due to sintering remains one
of the most fundamental problems in fuel cells, this finding may explain the
superiority in catalytic activity of carbon nanotube supported-Pt catalyst. The
increased adsorption energy of Pt13 cluster on the carbon nanotubes, 4.0 eV
SWCNTs vs 2.21 eV for graphene, which is an indication of an enhanced stability
was interpreted in term of geometric factors. According to the authors, “due to the
curvature-induced pyramidalization and misalignment of p-orbitals of carbon
atoms in SWCNT, carbon atoms on the outside wall of the SWCNT have more sp3
nature than those on a flat graphene sheet”. Also, considering that the Pt-C bonding
is due to hybridization between d-states of Pt atoms and p-states of adjacent C
atoms, it means that the Pt atoms are able to strongly interact with carbon atoms in
the case of carbon nanotubes, leading to an increase in the adsorption energy. The
high curvature of the carbon nanotube can promote the overlap between p and d
wave functions, thus increasing the adsorption energy of Pt on SWCNTs.
Importantly, the physico-chemical properties of carbon materials are intricately
linked with the methods used in their production or surface modification. Generally,
surface functionalization of carbons is aimed at improving the hydrophilicity (e.g.,
treatment with acids and bases) or hydrophobicity (e.g., treatment with organic
compounds such as benzene) of the carbon supports. The functional groups, such as
the oxygen-containing groups, provide the binding sites for the growth of the metal
catalyst ions. The most popular surface modification of carbon support is the
functionalization with acids or bases, a simple treatment that allows for the intro-
duction of oxygen-containing groups such as carboxyl, hydroxyl or phenolic groups
on the carbon surfaces. In their study, Kim and Park [33] interrogated the effects of
chemical treatment of carbon supports on electrochemical behaviours for platinum
catalysts of fuel cells. Three different carbon supports were studied with Pt cata-
lysts; virgin carbon blacks (CBs), neutral-treated carbon blacks (NCBs obtained by
treating CBs with 0.2 M benzene), base-treated carbon blacks (BCBs, obtained by
13 Effects of Catalyst-Support Materials on the Performance of Fuel Cells 527

treating CBs with 0.2 M KOH), and acid-treated carbon blacks (ACBs, obtained by
treating CBs with 0.2 M H3PO4). The authors found that the size and the Pt-loading
were strongly dependent on surface characteristics of the carbons blacks: BCBs-
supported Pt gave the smallest particle size (2.65 nm) and the highest loading
(97 %) compared to other carbons investigated. The electroactivity of the Pt cata-
lysts was enhanced with BCBs and NCBs, but decayed with ACBs.
CNTs are increasingly becoming attractive supports for electrocatalysts for fuel
cells due to their excellent mechanical strength, high conductivity, and high surface
areas. However, for the CNTs to be effectively used as support materials, they need
to be functionalized to enhance their dispersibility in solvents, and to permit
anchorage of the metal. Some notable methods for surface-functionalization of
CNTs include acid-treatments [34, 35] and sulfonation processes [36, 37]. The
sulfonation process was introduced by Peng et al. [38]. To improve the sulfonation
process, Ozoemena and co-workers [39] first functionalised the pristine CNTs to
increase the concentration of the oxo-groups (mainly the –COOH) by the three-step
acid-treatment process: (i) refluxing in 2.6 M HNO3 for 24 h; (ii) sonicating in
H2SO4/HNO3 mixture (3:1 ratio) for 24 h; and (iii) stirring in H2SO4/H2O2 mixture
(4:1 ratio) at 70  C. The acid-functionalised CNTs (MWCNT-COOH) were then
sulfonated using a mixture of H2SO4 and acetic anhydride at 70  C for 2 h using the
procedure reported by Sun et al. [36]. The catalysts were supported onto the
sulfonated-CNTs (MWCNT-SO3H) using the fast microwave-assisted polyol syn-
thesis, as shown in Fig. 13.9.
Recently, Ozoemena and co-workers introduced a new technique called “micro-
wave-induced top-down nanostructuring and decoration (MITNAD)” to synthesize
Pd-based ternary core–shell-shell (FeCo@Fe@Pd) nanocatalyst [40–43],
Figs. 13.10, 13.11, and 13.12. To understand the effect of chemical treatments on
the CNTs, they studied the catalytic properties of the FeCo@Fe@Pd nanocatalyst
supported on MWCNT-COOH and MWCNT-SO3H.
It was clearly proved that the surface functional groups (mainly –COOH and –
SO3H) on the MWCNT support played a critical role on the physico-chemical
properties of the FeCo@Fe@Pd nanocatalyst towards the electrocatalytic oxidation
of ethylene glycol and glycerol. The FeCo@Fe@Pd/MWCNT-COOH gave smaller
particle size (ca. 7.4 nm), more uniform dispersion or loading of the catalyst on the
support, higher electrochemically-active surface area (ECSA, ca. 75 m2 g1) and
more enhanced electrocatalytic activity than the FeCo@Fe@Pd/MWCNT-SO3H
(~11 nm particle size, with ECSA of ca. 42 m2 g1). The result is in agreement with
the findings of Kim and Park [33] as already discussed above, where carbon black
treated with acid showed poor catalyst loading compared to the base-treated or
neutral carbon blacks.
The MWCNT-COOH support (a weak acid) might have resulted in complete
reduction of the Pd, thus allowing a more uniform dispersion and higher loading of
catalysts than the MWCNT-SO3H support (strong acid). In a nutshell, the high
catalytic performance of the FeCo@Fe@Pd/MWCNT-COOH may be related to the
improved electronic properties of the catalysts coupled with the high affinity of its –
COOH surface with the catalyst species.
528 P.M. Ejikeme et al.

Fig. 13.9 Schematic of the preparation SF-MWCNT-M (M ¼ Pd, Ni or Sn) nanocomposites using
microwave-assisted strategy (Source: Ramulipho et al., Electrochimica Acta 2012, 59, 310–320.
Reproduced with permission from [39], copyright 2012 Elsevier LTD)

Chemical oxidation of carbon supports provides oxygen-containing functional


groups as the binding sites for the growth of metal ions. However, the oxidised sites
of the carbon supports may accelerate the degradation process of the support
material, thereby inducing significant aggregation of the catalyst nanoparticles.
13 Effects of Catalyst-Support Materials on the Performance of Fuel Cells 529

Fig. 13.10 (a) High-angle annular dark-field (HAADF)- and (b) bright-field (BF)-STEM images
simultaneously acquired for the FeCo@Fe@Pd/MWCNT-COOH (Source: Fashedemi et al.,
J. Mater. Chem. A, 2015, 3, 7145–7156. Reproduced from [43], copyright 2015 Royal Society
of Chemistry)

Fig. 13.11 FESEM images of (a) FeCo@Fe/MWCNT-COOH, (b) FeCo@Fe@Pd/ MWCNT-


COOH, (c) FeCo@Fe/MWCNT-SO3H and (d) FeCo@Fe@Pd/MWCNT-SO3H (Source:
Fashedemi et al., J. Mater. Chem. A, 2015, 3, 7145–7156. Reproduced from [43], copyright
2015 Royal Society of Chemistry)
530 P.M. Ejikeme et al.

Fig. 13.12 Chronoamperometric curves of the various catalysts in (a) 0.5MEG/1.0 M KOH and
(b) 0.5MGLY/1MKOH at a fixed potential of 0.16 V vs. Ag|AgCl (3 M KCl) (Source: Fashedemi
et al., J. Mater. Chem. A, 2015, 3, 7145–7156. Reproduced from [43], copyright 2015 Royal
Society of Chemistry)

To curb these potential oxidation problems, it is possible to wrap the carbon support
with polymeric materials that contain appropriate functional groups for the binding
of the catalyst metal ions. For example, Fujigaya et al. [44–50] have elegantly
wrapped CNTs with polymeric materials that contain nitrogen atoms (i.e., pyridine-
doped polybenzimidazole [50]), sulfonic acid (i.e., sulfonated polysulfone and
sulfonated polyimide [46]) and phosphonic acid groups (i.e., poly(vinylpho-
sphonicacid)-doped polybenzimidazole [44]). In a recent study [48], the group
wrapped CNTs with poly[2,29-(2,6-pyridine)-5,59-bibenzimidazole] (PyPBI;
Fig. 13.13), a polymer with N-atoms as the binding sites for the growth of Pt ions.
The wrapping was made possible through the π-π interactions between the CNTs
and PyPBI. A series of CNT/PyPBI/Pt composite were prepared by changing the
initial amount of Pt salt used in each reaction with the view of controlling the
13 Effects of Catalyst-Support Materials on the Performance of Fuel Cells 531

Fig. 13.13 Schematic drawing describing the preparation of series of CNT/PyPBI/Pt composites
(Source: Hafez et al., Scientific Reports 2014, 4, 6295. doi:10.1038/srep06295. Reproduced with
permission from [48], copyright 2014 Nature Publishing Group)

diameter of the Pt catalysts. The composite with the smallest Pt nanoparticle size of
2.3 nm diameter gave about 8 times higher mass activity than the fuel cell
containing Pt with a 3.7 nm diameter. This represents the first example of the
ability to control the diameter of Pt on polymer-wrapped carbon supporting mate-
rials. The catalysts displayed excellent stability with durability of 10 times higher
than that of the commercial carbon black/Pt.

13.2.2 Non-carbon Supports

As already discussed, the major drawback associated with the use of carbon
electrodes is corrosion. The corrosion problem of carbon-based support can be
ameliorated by the use of non-carbon support materials that are corrosion-resistant
and still maintain the desirable properties of carbon such as high-surface-area and
high electrical conductivity. To this end, porous metals and metal alloys, inorganic
metal oxides and mixed oxides, metal carbides and nitrides amongst others are
being studied and used as briefly described hereunder.

13.2.2.1 Conducting Polymers

The use of heterocyclic conducting polymers (Fig. 13.14) as support materials has
gained wide acceptance in fuel cells. Conducting polymers (CPs) such as
532 P.M. Ejikeme et al.

Fig. 13.14 Chemical structures of some common heterocyclic conducting polymers ((a)
polypyrrole, (b) polythiophene, (c) polyaniline, (d) poly-(3-methylthiophene))

polypyrrole (PPY), polythiophene, polyaniline (PANI), and poly-


(3-methylthiophene) possess conjugated backbones and are, therefore, convenient
for proton transport [51], since the repeating cyclic groups are weak Lewis base
sites favourable for proton hopping. In addition, the extensive conjugation-induced
structural co-planarity reduces the steric hindrance in proton transport.
These common conducting polymers and their derivatives used as supports for
catalysts are not only conjugated but most of them also possess heteroatoms which
could easily serve as the site for catalyst attachments, thus avoiding the necessity of
additional functionalization. In an a recent review by Dutta et al. [52], the authors
elegantly discussed the use of CPs as catalyst-supporting matrices, especially their
properties that have permitted them to provide better dispersion, distribution and
anchoring of catalysts for enhanced electrochemical performance. A number of
reports exist in the literature on various conducting polymers that have been used as
support materials [51, 53–58]. Antolini and Gonzalez [59] established that porous
polymeric matrixes not only increase the specific area of catalysts but, most
importantly, present higher tolerance to poisoning due to the adsorption of CO
species compared to carbon supports. CPs are highly electron-conducting (in the
range 106 to 103 S cm1, which is about two orders of magnitude higher than that
of conventionally used carbon supports [60, 61]) and also proton-conducting
materials. The inherent proton- and electron-conducting properties of CPs make
them good candidates for replacing Nafion ionomers in the catalyst layer of fuel cell
electrodes providing enhanced performance in the process. The morphology of the
CPs also play a role in their performance as support materials. For example, the
electrocatalytic oxidation of methanol on nanofibular polyaniline (nPANI) and
granular polyaniline (gPANI) electrodes modified with Pt microparticles showed
that nPANI had better conductivity and higher specific surface area, exhibiting a
considerably higher electrocatalytic activity on the methanol oxidation than the
gPANI electrode [53].
Huang et al. [62] investigated polypyrrole (PPy) samples, obtained by in-situ
chemical oxidative polymerization, as an alternate cathode catalyst support mate-
rial (i.e., for oxygen reduction reaction, ORR) in PEMFC. The electrochemical
stabilities of the PPy samples and commercial carbon supports (Vulcan XC-72)
were examined by cyclic voltammetry (CV). In contrast to the carbon supports
(Vulcan XC-72), the PPy only showed a small anodic current up to 1.8 V, indicating
resistance towards oxidation under high positive potentials. Pt/PPy was electro-
chemically more stable than the Pt/C electrocatalyst. Tafel plots for the ORR
activity showed two-fold higher activity for Pt/PPy catalyst than that of Pt black
catalyst at 0.9 V. The Pt/PPy catalyst demonstrated good ORR kinetics and
comparable fuel cell performance with that of the commercial E-TEK Pt/C catalyst.
13 Effects of Catalyst-Support Materials on the Performance of Fuel Cells 533

Some of the advantages of the use of the CPs as support materials in fuel cells
(e.g., DMFC) include: (i) ability to significantly reduce CH3OH crossover, with a
slight compromise on proton conductivity; (ii) ability to allow the diffusion of gases
out of the MEA due to the porous nature of CPs; (iii) replacement of the costly
Nafion, thus reducing the overall cost of the fuel cells, and (iv) CPs are thermally
stable in the temperature range within which DMFC operates, thus rendering
stability to the overall system [51]. Other major advantages of CPs (especially the
aromatic CPs) as support constituent include the prospect of (i) adequate catalyst
binding on the inherently present heteroatoms in common ACPs, and its deriva-
tives, thus avoiding the necessity for additional functionalization steps [54, 58, 61],
and (ii) fine dispersion and uniform distribution of expensive catalysts, resulting in
their exhibiting the maximum active surface area for catalytic reactions leading to
their reduced consumption and maximum utilization [54, 63].
Another important CP is the poly(3,4-ethylenedioxythiophene)-polystyrene
sulphonic acid (PEDOT-PSSA) which has been shown to perform better than the
traditional Vulcan XC-72R carbon, as shown in a solid-polymer-electrolyte direct
methanol fuel cell (SPE-DMFC using Pt-Ru catalyst) [61]. PEDOT-PSSA
performed better than Vulcan XC-72 carbon in terms of electrical conductivity
and current response. The high performance of the PEDOT-PSSA was attributed to
its mixed-conducting nature that enhanced proton and electron transport within the
anode catalyst, improving the activity of the Pt-Ru catalyst.
Recently, Wang et al. [30] investigated the gold nanoparticles supported on
various supports for glycerol electro-oxidation. The authors found that Au
nanocatalysts supported on extended poly(4-vinylpyridine) functionalized
graphene (Au-P4P/G) exhibited much higher activity and selectivity for three
carbon products than those on carbon black, P4P functionalized reduced graphene
oxide (Au-P4P/rGO) and poly(m-aminophenol) (PmAP) wrapped graphene
(Au-PmAP/G).

13.2.2.2 Metal Oxides and Hydroxides

Metal oxides and hydroxides are among the non-carbon materials used as support
for direct alcohol fuel cell electrocatalysis in a bid to circumvent the perennial
problem of carbon corrosion. The following metal oxides have been incorporated as
co-catalysts for fuel cells (Table 13.1): NbO, NbO2 and Nb2O5 [64–75], niobium-
doped titanium dioxide [76–79], ZrO2 [80–85], CeO2 used alone [86–90] or in
conjunction with other oxides or carbon [91–96]. Lei et al. recently showed that the
excellent durability of Pt-CeO2 nanocubes-graphene oxide catalyst was attributable
to the free radical scavenging activity of CeO2 that significantly slowed down the
chemical degradation of Nafion binder in catalytic layers, consequently alleviating
the decay of Pt catalysts, resulting in catalyst’s excellent cycle life [97]. Other
metallic oxides which have been found useful include MoOx [98–103], RuO2 and
Ru-mixed oxides [104–110], titanium oxides (TiO2) and their composites [28, 106,
111–118], MnO2, Mn2O3, Mn3O4 and MnO [89, 90, 119–122], cobalt oxides and
534 P.M. Ejikeme et al.

Table 13.1 Other oxide support materials employed in DAFCs


Active catalyst Oxide type Process involved Reference
Pt ZnO MeOH oxidation [144]
PdCo tGO# ORR [145]
Pd-Zn Al2O3 MeOH reforming [146]
Pd Al2O3 MeOH reforming [146]
Au γFe2O3 CO oxidation [147]
Pt γFe2O3 CO oxidation [148]
Pt RuO2-TiO2 ROH oxidation [149]
Pt RuO2/xH2O-TiO2 ROH oxidation [149]
Pt SnO2 ORR [150]
Pt SnO2/Ketjen black ORR [151]
Pt SnO2-Nb I-V performance [152]
Pt MgO/C EtOH oxidation [153]
Pt MgO/C EtOH oxidation [154]
Pd NiO/MgO@C EtOH oxidation [155]
Pd In2O3/CNTs EtOH oxidation [156]
Pd TiO2-C MeOH oxidation [157]
Pt MgO/CNT EtOH oxidation [158]
Mn CaMn3O6 ORR [159]
Mn CaMnO3 ORR [159]
Pt Mn3(PO4)2@BSA MeOH oxidation [119]
Key: tGO# ¼ thiolated grapheme oxide

hydroxides [89, 90, 123–132] as well as silica, a unique oxide that forms hydroxyl
species even in acidic pH and their hybrids [133–143]. The vast number of
transition metal oxides arising from the variability in the oxidation states of the
transition metals has made possible the design of numerous transition metal-based
oxides for FC catalyst support. Other promising alternative metal oxide support
materials employed in the search for ways of alleviating the problem of corrosion of
the carbon support of electrocatalysts under cathode conditions in FCs are summa-
rized in Table 13.1. Bimetallic catalysts combined with some oxides, like TiO2 and
exhibiting definite morphologies (Fig. 13.15) have shown promising results in
methanol oxidation.
Metal oxides are considered to be more appealing as they spontaneously form
surface hydroxyl groups that helps to circumvent adsorption of catalyst poisons,
CO, albeit most of them in an alkaline medium.

13.2.2.3 Transition Metal Carbides

Transition metal carbides (TMCs) are produced by incorporating C atoms into the
interstitial sites of transition metal atoms, including all 3d elements, as well as 4d
and 5d elements of group’s 4–6 early transition metals. It was reported [160–162]
13 Effects of Catalyst-Support Materials on the Performance of Fuel Cells 535

Fig. 13.15 Scanning electron micrograph showing the carbon fiber paper and its modification
with TiO2 and Pt catalyst: (a) CFE, (b) CFE/TiO2, (c) Pt-Ru/CFE, and (d) TiO2/CFE/Pt-Ru.
(Magnification: Bar in left column represents 50 μm, bar in right column represents 6 μm).
(Source: Drew et al., J. Phys. Chem. B 2005, 109, 11851–11857. Reproduced with permission
from [111], copyright 2005 American Chemical Society)

that the groups 4–6 TMCs typically display properties characteristic of three
different classes of materials that include the extreme hardness and brittleness of
covalent solids, the ordered bulk structure and high melting temperatures of ionic
solids, as well as electric conductivity and heat capacity of metals. Since the
electrochemical properties of metal carbides can vary greatly depending on oper-
ating conditions (i.e., electrochemical potential, current density) as well as pH
environment (i.e., acidic, neutral, basic) [163] it is therefore of particular concern
to recognize the electrochemical stability of transition metal carbide catalysts over
536 P.M. Ejikeme et al.

a wide range of chemical environments to determine their usefulness for different


electrochemical applications. The use of these metal carbides either as catalysts or
catalyst supports in a variety of electrochemical devices including fuel cells, photo-
electrochemical cells, etc. is expected to lead to reduction in the amount of noble
metals used in electrochemical applications. The literatures are replete with con-
clusions that the interaction between carbides and noble metals increases with the
decrease of carbide particle size, [164, 165] (i.e., the larger carbide particles having
lower specific surface area, resulting in poor dispersion of loaded noble metal
nanoparticles) [166]. The ability of the transition metals of roughly groups 4–6 to
form stable carbides, including multiple stoichiometries and polymorphs, that are
resistant to electrochemical corrosion in addition to being more abundant and
orders of magnitude less expensive than the Pt-group metal catalysts [167, 168]
have ultimately lead to increased interest in these materials for electrochemical
applications. The establishment of syntheses methods such as salt flux synthesis and
temperature programmed reduction developed to increase the range of carbide
materials available and to lower the cost of synthesizing these carbides [167]
have been added incentive in the electrochemical application of these materials.
A comparison of the valence band X-ray photoelectron spectrum of tungsten
carbide with the spectra of tungsten and platinum shows that, near the Fermi level,
the electronic density of states of tungsten carbide more nearly resembles that of
platinum than that of tungsten [169, 170]. This overall resemblance in the electronic
as well as in catalytic properties of Pt and tungsten monocarbide (WC) [171], has
made tungsten carbide one of the most studied [172] amongst these transition
metals, either as electrocatalyst or as support material for low loadings of Pt. It
should be noted though that contrary to the hypothesis that the platinum-like
catalytic activity of tungsten carbide results from the contribution of carbon valence
electrons to the 5d band of tungsten, the width of the unfilled portion of the d band
increases continuously from tungsten to tungsten carbide, according to Houston
et al. [173]. From the foregoing, a number of reports on the use of transition metal
carbides as electrocatalysts exists [163, 167, 171, 172, 174, 175] while even more
could be found in the literature on their use as catalyst support materials [164, 166,
171, 176–179]. For further reading on this subject matter, reference should be made
to some extensive and excellent reviews in the literature [13, 160, 174, 175].

13.2.2.4 Transition Metal Nitrides

The use of transition metal nitrides as catalyst support in direct alcohol fuel cells
electrocatalysis has become a subject of great interest due to their inherent good
catalyst support interaction, thermal stability, high corrosion resistance and high
conductivity compared to carbon supports. Some of the nitrides investigated
include the mesoporous binary nitrides (e.g., CrN, TiN, Fig. 13.16 [3, 180])
and ternary nitride (e.g., Ti0.5Nb0.5N [181]) complexes. In all cases, Yang
et al. [180–183] reported that these nitride supports for Pt showed excellent
electronic conductivity, high surface area, and better performance towards
13 Effects of Catalyst-Support Materials on the Performance of Fuel Cells 537

Fig. 13.16 SEM images of TiN samples from ammonolysis of Zn2TiO4 at (a) 500 1C, (b) 600 1C,
(c) 700 1C and (d) 800 1C for 8 h (Source: Yang et al., Phys. Chem. Chem. Phys., 2013,
15, 1088–1092. Reproduced with permission from [183], copyright 2013 Royal Society of
Chemistry)

methanol electro-oxidation in both acid and alkaline media than commercial Pt/C.
In addition, they showed better electrochemical stability and higher tolerance to
corrosion than the commercial Pt/C. Unlike other nitrides investigated, Mo2N was
found to be electrochemically unstable even at a relative low potential (0.5 V,
vs. SCE) [184]. This finding suggests that Mo2N may not be able to promote ORR,
especially if one considers that ORR occurs at reasonable rates between ~0.45 and
0.65 V (RHE) and its standard equilibrium potentials is ~0.80 V (RHE).
Thotiyl and co-workers [185] indicated that titanium nitride nanoparticles (TiN
NPs) are nitrides of choice because it does not only exhibit excellent corrosion
resistance and electrochemical stability, but also acts as a co-catalyst adsorbing OH
groups and facilitating the oxidation of the poisonous intermediates (CO) adhered
on adjacent Pt surface, enhancing the methanol oxidation reaction (MOR) activity
in the process (especially for high temperature PEM fuel cells, to replace carbon
materials [165]). The authors observed that the onset of ethanol oxidation was
negatively shifted by at least 100 mV on Pt-TiN as compared to Pt-C and that the
peak currents were more than doubled when Pt was loaded onto TiN showing the
positive effect of TiN support for Pt towards ethanol oxidation reaction (EOR). A
positive chemical shift of Pd (3d) binding energy was used by the authors to confirm
the existence of metal-support interaction between Pd and TiN, which in turn
538 P.M. Ejikeme et al.

helped to weaken the Pd-CO synergetic bonding interaction during the electro-
oxidation process [186].
A number of other researchers have also probed into the possibilities of using
metallic nitrides as electro-oxidation catalysts, with varying degrees of success
[165, 187–192] and recommendations [193–198]. The possibility of whether the
introduction of molybdenum element into TiN might improve the electrocatalytic
activity of supported Pt catalyst on one hand, and preserve the intrinsic electro-
chemical stability of the TiN nanostructures on the other towards methanol oxida-
tion was the subject of a recent investigation [199]. The titanium molybdenum
nitride (Ti0.8Mo0.2N) supported Pt catalyst did not only exhibit a much higher mass
activity and durability than that of the conventional Pt/C (E-TEK) electrocatalysts
for MOR but also led to a significant decrease of Pt density on carbon support after
the accelerated durability test (ADT) test, demonstrating that the major cause for
ECSA loss of Pt/C was Pt detaching, ripening and the corrosion of the carbon
support. DAFC support for Pt nanoparticles with high surface area, hollow and
interior porous structure, titanium nitride nanotubes (TiN NTs), was obtained by
Xiao et al. [200] via a combination of solvothermal alcoholysis and post-nitriding
method. The support reportedly displayed enhanced activity and durability during
MOR compared with the commercial Pt/C (E-TEK) catalyst. The performance of
the TiN NTs was attributed to the smaller size and better dispersion of the Pt NPs
(providing high EASA), the unique hollow and porous structure (Fig. 13.17) of the
TiN NTs (speeding up the diffusion rate and oxidation of the carbonaceous species),
as well as co-catalyst effect, and strong interaction with Pt nanoparticles.

13.2.2.5 Dual-Purpose Supports

Some materials play dual role as support as well as current collector. Some of the
materials that have served this dual purpose include glassy carbon rods [201],
nickel foam and carbon cloths [202]. For example, Modibedi et al. [202] employed

Fig. 13.17 (a) SEM image of TiN NTs and (insert) magnified SEM image of the vertical section
of some nanotubes. (b) A typical enlarged SEM image of a TiN NT. [Xiao et al., Electrochim.
Acta, 2014, 141, 279–285. Reproduced with permission from [200], copyright 2014 Elsevier Ltd]
13 Effects of Catalyst-Support Materials on the Performance of Fuel Cells 539

Ni foam and carbon cloth as both supports and current collectors for Pt and Pd
catalysts obtained by electrochemical atomic layer deposition (EC-ALD). They
observed that methanol oxidation proceeded better on carbon paper than on Ni
foam; while the reverse was the case with ethanol oxidation and that the two
supports produced good quality deposits.

13.3 Summary and Future Prospects

In this book chapter, we have reviewed the concept of fuel cell generally and
particularly various types of catalyst supports that have been developed over time
for use in fuel cell eletrocatalysis. The direct conversion of the chemical energy/
heat of combustion of fuels and oxidants in the presence of a catalyst into electric-
ity, heat and water at an efficiency higher than that of internal combustion engine
have revolutionised the search for environmentally benign fuel. The challenges of
the use of carbon supports in fuel cells are gradually being addressed by the
modification of conventional carbon supports as well as the use of non-carbon
supports. Significant progress has been made in the commercialization of fuel cells
for various applications to major corporations while fuel cells for aircraft auxiliary
power units (APUs) that provide energy for functions other than propulsion as well
as fuel cells for fire prevention are very high potential outlets.
Due to the high price premiums associated with fuel cells, added value features
need to be exploited in order to make them more appealing and increase unit sales
and market penetration. With the limited availability and high cost of platinum and
platinum group metals, research should be sustained in the search for low cost
materials that may uniquely exhibit catalytic and stability properties equivalent to,
or surpassing that of platinum. Irrespective of the effort already made in developing
modern fuel cell technologies, there is need for newer technologies that shall
alleviate the prevailing economic challenges and exceed the advantages of the
existing technologies to be fully acceptable for mass production. It is also
recommended that research effort should be channelled towards the development
of fuel cell materials that will address most of the challenges highlighted in this
chapter and elsewhere in this book as well as full optimization of fuel cell operating
conditions. It is equally important that the new catalysts being developed keep their
electrochemical performance for ORR and their alcohol tolerance ability over a
long usage time.

References

1. Huang H, Wang X (2014) Recent progress on carbon-based support materials for


electrocatalysts of direct methanol fuel cells. J Mater Chem A 2:6249–6670. doi:10.1039/
c3ta14754a
540 P.M. Ejikeme et al.

2. Wang H, Dai H (2013) Strongly coupled inorganic-nano-carbon hybrid materials for energy
storage. Chem Soc Rev 42:3088–3113. doi:10.1039/c2cs35307e
3. Linares N, Silvestre-Albero AM, Serrano E, Silvestre-Albero J, Garcia-Martinez J (2014)
Mesoporous materials for clean energy technologies. Chem Soc Rev 43:7681–7717. doi:10.
1039/c3cs60435g
4. Guo YG, Hu JS, Wan LJ (2008) Nanostructured materials for electrochemical energy
conversion and storage devices. Adv Mater 20:2878–2887. doi:10.1002/adma.200800627
5. Choi NS, Chen Z, Freunberger SA, Ji X, Sun YK, Amine K et al (2012) Challenges facing
lithium batteries and electrical double-layer capacitors. Angew Chem Int Ed 51:9994–10024.
doi:10.1002/anie.201201429
6. Zhang Z, Shimizu T, Senz S, G€ osele U (2009) Ordered high-density Si [100] nanowire arrays
epitaxially grown by bottom imprint method. Adv Mater 21:2824–2828. doi:10.1002/adma.
200802156
7. Jariwala D, Sangwan VK, Lauhon LJ, Marks TJ, Hersam MC (2013) Carbon nanomaterials
for electronics, optoelectronics, photovoltaics, and sensing. Chem Soc Rev 42:2824–2860.
doi:10.1039/c2cs35335k
8. Aiyejina A, Sastry MKS (2012) PEMFC flow channel geometry optimization: a review. J
Fuel Cell Sci Technol 9:011011. doi:10.1115/1.4005393
9. Grujicic M, Chittajallu KM (2004) Optimization of the cathode geometry in polymer
electrolyte membrane (PEM) fuel cells. Chem Eng Sci 59:5883–5895. doi:10.1016/j.ces.
2004.07.045
10. Carcadea E, Stefanescu I, Ionete RE, Ene H, Ingham DB, Ma L (2008) PEM fuel cell
geometry optimisation using mathematical modelling. Int J Multiphys 2:313–326. doi:10.
1260/175095408786927462
11. Grove WR (1839) On a small voltaic battery of great energy; some observations on voltaic
combinations and forms of arrangement; and on the inactivity of a copper positive electrode
in nitro-sulphuric acid. Philos Mag 15:287–293. doi:10.1080/14786443908649881
12. Grove WR (1839) On voltaic series and the combination of gases by platinum. Philos Mag
14:127–130. doi:10.1080/14786443908649684
13. Li L, Hu L, Li J, Wei Z (2015) Enhanced stability of Pt nanoparticle electrocatalysts for fuel
cells. Nano Res 8:418–440. doi:10.1007/s12274-014-0695-5
14. Sharma S, Pollet BG (2012) Support materials for PEMFC and DMFC electrocatalysts—a
review. J Power Sources 208:96–119. doi:10.1016/j.jpowsour.2012.02.011
15. Martı́nez-Huerta M, Tsiouvaras N, Garcı́a G, Pe~ na M, Pastor E, Rodriguez J et al (2013)
Carbon-supported ptrumo electrocatalysts for direct alcohol fuel cells. Catalysts 3:811–838.
doi:10.3390/catal3040811
16. Li Y, Xu L, Liu H, Li Y (2014) Graphdiyne and graphyne: from theoretical predictions to
practical construction. Chem Soc Rev 43:2572–2586. doi:10.1039/c3cs60388a
17. Skrabalak SE, Suslick KS (2006) Porous carbon powders prepared by ultrasonic spray
pyrolysis. J Am Chem Soc 128:12642–12643. doi:10.1021/ja064899h
18. Bianchi CL, Biella S, Gervasini A, Prati L, Rossi M (2003) Gold on carbon: influence of
support properties on catalyst activity in liquid-phase oxidation. Catal Lett 85:91–96. doi:10.
1023/A:1022176909660
19. Auer E, Freund A, Pietsch J, Tacke T (1998) Carbons as supports for industrial precious metal
catalysts. Appl Catal A Gen 173:259–271. doi:10.1016/S0926-860X(98)00184-7
20. Qiu Z, Huang H, Du J, Tao X, Xia Y, Feng T et al (2014) Biotemplated synthesis of bark-
structured TiC nanowires as Pt catalyst supports with enhanced electrocatalytic activity and
durability for methanol oxidation. J Mater Chem A 2:8003. doi:10.1039/c4ta00277f
21. Wilson MS (1993) Surface area loss of supported platinum in polymer electrolyte fuel cells. J
Electrochem Soc 140:2872. doi:10.1149/1.2220925
22. Philippot K, Serp P (2012) Concepts in nanocatalysis. In: Nanomaterials in catalysis, 1st edn.
Wiley-VCH, Weinheim, pp 1–54
13 Effects of Catalyst-Support Materials on the Performance of Fuel Cells 541

23. Yuan X, Ding X-L, Wang C-Y, Ma Z-F (2013) Use of polypyrrole in catalysts for low
temperature fuel cells. Energy Environ Sci 6:1105–1124. doi:10.1039/C3EE23520C
24. Zhang Y, Chen S, Wang Y, Ding W, Wu R, Li L et al (2015) Study of the degradation
mechanisms of carbon-supported platinum fuel cells catalyst via different accelerated stress
test. J Power Sources 273:62–69. doi:10.1016/j.jpowsour.2014.09.012
25. Borup R, Meyers J, Pivovar B, Kim YS, Mukundan R, Garland N et al (2007) Scientific
aspects of polymer electrolyte fuel cell durability and degradation. Chem Rev
107:3904–3951
26. Kinoshita K (1988) Carbon: electrochemical and physicochemical properties. Wiley,
New York, ISBN: 978-0471848028
27. Sasaki K, Shao M, Adzic R (2009) Dissolution and stabilization of platinum in oxygen
cathodes. In: Proton exchange membrane fuel cell durability. Springer, New York, pp 7–28
28. Huang S, Ganesan P, Park S, Popov BN (2009) Development of a titanium dioxide supported
platinum catalyst with ultrahigh stability for polymer electrolyte membrane fuel cell appli-
cations. J Am Chem Soc 131:13898–13899. doi:10.1021/ja904810h
29. Mathias MF, Makharia R, Gasteiger HA, Conley JJ, Fuller TJ, Gittleman CJ et al (2005) Two
fuel cell cars in every garage? Electrochem Soc Interface 14:24–35
30. Wang H, Thia L, Li N, Ge X, Liu Z, Wang X (2015) Selective electro-oxidation of glycerol
over Au supported on extended poly(4-vinylpyridine) functionalized graphene. Appl Catal B
Environ 166–167:25–31. doi:10.1016/j.apcatb.2014.11.009
31. Castanheira L, Dubau L, Mermoux M, Berthomé G, Caqué N, Rossinot E et al (2014) Carbon
corrosion in proton-exchange membrane fuel cells: from model experiments to real-life
operation in membrane electrode assemblies. ACS Catal 4:2258–2267. doi:10.1021/
cs500449q
32. Cuong NT, Fujiwara A, Mitani T, Chi DH (2008) Effects of carbon supports on Pt nano-
cluster catalyst. Comput Mater Sci 44:163–166. doi:10.1016/j.commatsci.2008.01.061
33. Kim S, Park SJ (2006) Effects of chemical treatment of carbon supports on electrochemical
behaviors for platinum catalysts of fuel cells. J Power Sources 159:42–45. doi:10.1016/j.
jpowsour.2006.04.041
34. Siswana MP, Ozoemena KI, Nyokong T (2006) Electrocatalysis of asulam on cobalt phtha-
locyanine modified multi-walled carbon nanotubes immobilized on a basal plane pyrolytic
graphite electrode. Electrochim Acta 52:114–122. doi:10.1016/j.electacta.2006.03.090
35. Liu J, Rinzler AG, Dai H, Hafner JH, Bradley RK, Boul PJ et al (1998) Fullerene pipes.
Science 280:1253–1256. doi:10.1126/science.280.5367.1253
36. Sun ZP, Zhang XG, Liang YY, Li HL (2009) A facile approach towards sulfonate functiona-
lization of multi-walled carbon nanotubes as Pd catalyst support for ethylene glycol electro-
oxidation. J Power Sources 191:366–370. doi:10.1016/j.jpowsour.2009.01.093
37. Du CY, Zhao TS, Liang ZX (2008) Sulfonation of carbon-nanotube supported platinum
catalysts for polymer electrolyte fuel cells. J Power Sources 176:9–15. doi:10.1016/j.
jpowsour.2007.10.016
38. Peng F, Zhang L, Wang H, Lv P, Yu H (2005) Sulfonated carbon nanotubes as a strong
protonic acid catalyst. Carbon N Y 43:2405–2408. doi:10.1016/j.carbon.2005.04.004
39. Ramulifho T, Ozoemena KI, Modibedi RM, Jafta CJ, Mathe MK (2012) Fast microwave-
assisted solvothermal synthesis of metal nanoparticles (Pd, Ni, Sn) supported on sulfonated
MWCNTs: Pd-based bimetallic catalysts for ethanol oxidation in alkaline medium.
Electrochim Acta 59:310–320. doi:10.1016/j.electacta.2011.10.071
40. Fashedemi OO, Julies B, Ozoemena KI (2013) Synthesis of Pd-coated FeCo@Fe/C core-shell
nanoparticles: microwave-induced “top-down” nanostructuring and decoration. Chem
Commun (Camb) 49:2034–2036. doi:10.1039/c3cc38672d
41. Fashedemi OO, Ozoemena KI (2013) Enhanced methanol oxidation and oxygen reduction
reactions on palladium-decorated FeCo@Fe/C core-shell nanocatalysts in alkaline medium.
Phys Chem Chem Phys 15:20982–20991. doi:10.1039/c3cp52601a
542 P.M. Ejikeme et al.

42. Fashedemi OO, Ozoemena KI (2014) Comparative electrocatalytic oxidation of ethanol,


ethylene glycol and glycerol in alkaline medium at Pd-decorated FeCo@Fe/C core-shell
nanocatalysts. Electrochim Acta 128:279–286. doi:10.1016/j.electacta.2013.10.194
43. Fashedemi OO, Miller HA, Marchionni A, Vizza F, Ozoemena KI (2015) Electro-oxidation
of ethylene glycol and glycerol at palladium-decorated FeCo@Fe core–shell nanocatalysts
for alkaline direct alcohol fuel cells: functionalized MWCNT supports and impact on product
selectivity. J Mater Chem A 3(2015):7145–7156. doi:10.1039/C5TA00076A
44. Berber MR, Fujigaya T, Sasaki K, Nakashima N (2013) Remarkably durable high tempera-
ture polymer electrolyte fuel cell based on poly(vinylphosphonic acid)-doped polybenzi-
midazole. Sci Rep 3:1764. doi:10.1038/srep01764
45. ChemCatChem (2013) Interfacial engineering of Pt catalysts for fuel cells—MeOH oxidation
is dramatically improved by polymer coating on a Pt catalyst. 7:1701–1704. doi:10.1002/
cctc.201300157
46. Fujigaya T, Kim CR, Matsumoto K, Nakashima N (2013) Effective anchoring of
Pt-nanoparticles onto sulfonated polyelectrolyte-wrapped carbon nanotubes for use as a
fuel cell electrocatalyst. Polym J 45:326–330. doi:10.1038/pj.2012.145
47. Fujigaya T, Nakashima N (2013) Fuel cell electrocatalyst using polybenzimidazole-modified
carbon nanotubes as support materials. Adv Mater 25:1666–1681. doi:10.1002/adma.
201204461
48. Hafez IH, Berber MR, Fujigaya T, Nakashima N (2014) Enhancement of platinum mass
activity on the surface of polymer-wrapped carbon nanotube-based fuel cell electrocatalysts.
Sci Rep 4:6295. doi:10.1038/srep06295
49. Matsumoto K, Fujigaya T, Sasaki K, Nakashima N (2011) Bottom-up design of carbon
nanotube-based electrocatalysts and their application in high temperature operating polymer
electrolyte fuel cells. J Mater Chem 21:1187. doi:10.1039/c0jm02744h
50. Fujigaya T, Okamoto M, Nakashima N (2009) Design of an assembly of pyridine-containing
polybenzimidazole, carbon nanotubes and Pt nanoparticles for a fuel cell electrocatalyst with
a high electrochemically active surface area. Carbon N Y 47:3227–3232. doi:10.1016/j.
carbon.2009.07.038
51. Dutta K, Kumar P, Das S, Kundu PP (2014) Utilization of conducting polymers in fabricating
polymer electrolyte membranes for application in direct methanol fuel cells. Polym Rev
54:1–32. doi:10.1080/15583724.2013.839566
52. Dutta K, Das S, Rana D, Kundu PP (2015) Enhancements of catalyst distribution and
functioning upon utilization of conducting polymers as supporting matrices in DMFCs: a
review. Polym Rev 55:1–56. doi:10.1080/15583724.2014.958771
53. Zhou HH, Jiao SQ, Chen JH, Wei WZ, Kuang YF (2004) Effects of conductive polyaniline
(PANI) preparation and platinum electrodeposition on electroactivity of methanol oxidation.
J Appl Electrochem 34:455–459. doi:10.1023/B:JACH.0000016635.35555.04
54. Dutta K, Kundu PP (2014) A review on aromatic conducting polymers-based catalyst
supporting matrices for application in microbial fuel cells. Polym Rev 54:401–435. doi:10.
1080/15583724.2014.881372
55. Rajesh B, Thampi KR, Bonard J-M, Mathieu HJ, Xanthopoulos N, Viswanathan B (2003)
Conducting polymeric nanotubules as high performance methanol oxidation catalyst support.
Chem Commun (Camb) 21:2022–2023. doi:10.1039/b305591d
56. Maiyalagan T (2008) Electrochemical synthesis, characterization and electro-oxidation of
methanol on platinum nanoparticles supported poly(o-phenylenediamine) nanotubes. J Power
Sources 179:443–450. doi:10.1016/j.jpowsour.2008.01.048
57. Chen Z, Xu L, Li W, Waje M, Yan Y (2006) Polyaniline nanofibre supported platinum
nanoelectrocatalysts for direct methanol fuel cells. Nanotechnology 17:5254–5259. doi:10.
1088/0957-4484/17/20/035
58. Ma Y, Jiang S, Jian G, Tao H, Yu L, Wang X et al (2009) CNx nanofibers converted from
polypyrrole nanowires as platinum support for methanol oxidation. Energy Environ Sci
2:224. doi:10.1039/b807213m
13 Effects of Catalyst-Support Materials on the Performance of Fuel Cells 543

59. Antolini E, Gonzalez ER (2009) Polymer supports for low-temperature fuel cell catalysts.
Appl Catal A Gen 365:1–19. doi:10.1016/j.apcata.2009.05.045
60. Qi Z, Pickup PG (1998) High performance conducting polymer supported oxygen reduction
catalysts. Chem Commun 1998:2299–2300. doi:10.1039/a805322g
61. Tintula KK, Pitchumani S, Sridhar P, Shukla AK (2010) A solid-polymer-electrolyte direct
methanol fuel cell (DMFC) with Pt-Ru nanoparticles supported onto poly
(3,4-ethylenedioxythiophene) and polystyrene sulphonic acid polymer composite as anode.
J Chem Sci 122:381–389. doi:10.1007/s12039-010-0043-6
62. Huang SY, Ganesan P, Popov BN (2009) Development of conducting polypyrrole as
corrosion-resistant catalyst support for polymer electrolyte membrane fuel cell (PEMFC)
application. Appl Catal B Environ 93:75–81. doi:10.1016/j.apcatb.2009.09.014
63. Liu FJ, Huang LM, Wen TC, Li CF, Huang SL, Gopalan A (2008) Effect of deposition
sequence of platinum and ruthenium particles into nanofibrous network of polyaniline-poly
(styrene sulfonic acid) on electrocatalytic oxidation of methanol. Synth Met 158:603–609.
doi:10.1016/j.synthmet.2008.04.002
64. Kremliakova N (2013) Stable, durable carbon supported catalyst composition for fuel cell.
US patent 2013/0164655 A1
65. Campbell SA, Kremliakova N (2014) Synthesis of stable and durable catalyst composition for
fuel cell. US patent 8722284 B2
66. Hardman S, Chandan A, Steinberger-Wilckens R (2014) Fuel cell added value for early
market applications. J Power Sources 287:297–306. http://dx.doi.org/10.1016/j.jpowsour.
2015.04.056
67. Brosha EL, Blackmore KJ, Burrell AK, Henson NJ, Phillips J (2010) Engineered nano-scale
ceramic supports for PEM fuel cells. In: 2010 fuel cell seminar and exposition, San
Antonio, TX
68. Merzougui B, Shao M, Protsailo LV (2011) Fuel cell catalyst support with boron
carbide-coated metal oxides/phosphates and method of manufacturing same. US patent
20110136047 A1
69. Zhang L, Wang L, Holt CMB, Zahiri B, Li Z, Malek K et al (2012) Highly corrosion resistant
platinum–niobium oxide–carbon nanotube electrodes for the oxygen reduction in PEM fuel
cells. Energy Environ Sci 5:6156. doi:10.1039/c2ee02689a
70. Morozan A, Jousselme B, Palacin S (2011) Low-platinum and platinum-free catalysts for the
oxygen reduction reaction at fuel cell cathodes. Energy Environ Sci 4:1238. doi:10.1039/
c0ee00601g
71. Tripković V, Abild-Pedersen F, Studt F, Cerri I, Nagami T, Bligaard T et al (2012) Metal
oxide-supported platinum overlayers as proton-exchange membrane fuel cell cathodes.
ChemCatChem 4:228–235. doi:10.1002/cctc.201100308
72. Bonakdarpour A, Tucker RT, Fleischauer MD, Beckers NA, Brett MJ, Wilkinson DP (2012)
Nanopillar niobium oxides as support structures for oxygen reduction electrocatalysts.
Electrochim Acta 85:492–500. doi:10.1016/j.electacta.2012.08.005
73. Huang K, Li Y, Yan L, Xing Y (2014) Nanoscale conductive niobium oxides made through
low temperature phase transformation for electrocatalyst support. RSC Adv 4:9701. doi:10.
1039/c3ra47091a
74. Chris MB, Michael H, Zhang L, Wang L, Holt CMB, Navessin T et al (2010) NRC
Publications Archive (NPArC) Archives des publications du CNRC (NPArC) oxygen reduc-
tion reaction activity and electrochemical stability of thin-film bilayer systems of platinum on
niobium oxide. J Phys Chem C 114:16463–16474
75. Xu C, Pietrasz P, Yang J, Soltis R, Sun K, Sulek M, et al (2013) Pt-based ORR catalyst on
carbon-supported amorphous niobium oxide support. In: 224th ECS meeting. The Electro-
chemical Society
76. Huang S-Y, Ganesan P, Popov BN (2010) Electrocatalytic activity and stability of niobium-
doped titanium oxide supported platinum catalyst for polymer electrolyte membrane fuel
cells. Appl Catal B Environ 96:224–231. doi:10.1016/j.apcatb.2010.02.025
544 P.M. Ejikeme et al.

77. Do TB, Cai M, Ruthkosky MS, Moylan TE (2010) Niobium-doped titanium oxide for fuel
cell application. Electrochim Acta 55:8013–8017. doi:10.1016/j.electacta.2010.03.027
78. Rosenfeld D, Schmid P, Széles S, Lévy F, Demarne V, Grisel A (1996) Electrical transport
properties of thin-film metal-oxide-metal Nb2O5 oxygen sensors. Sensors Actuators B Chem
37:83–89. doi:10.1016/S0925-4005(96)01991-0
79. Morris D, Dou Y, Rebane J, Mitchell C, Egdell R, Law D et al (2000) Photoemission and
STM study of the electronic structure of Nb-doped TiO2. Phys Rev B 61:13445–13457.
doi:10.1103/PhysRevB.61.13445
80. Iordache C, Blair S, Lycke D, Huff S (2008) Catalysts including metal oxide for organic fuel
cells. US 2008/0014494 A1
81. Suzuki Y, Ishihara A, Mitsushima S, Kamiya N, Ota K-I (2007) Sulfated-zirconia as a support
of Pt catalyst for polymer electrolyte fuel cells. Solid-State Lett 10:B105–B107. doi:10.1149/
1.2730625
82. Subban C, Zhou Q, Leonard B, Ranjan C, Edvenson HM, Disalvo FJ et al (2010) Catalyst
supports for polymer electrolyte fuel cells. Philos Trans A Math Phys Eng Sci
368:3243–3253. doi:10.1098/rsta.2010.0116
83. Lv H, Mu S (2014) Nano-ceramic support materials for low temperature fuel cell catalysts.
Nanoscale 6:5063–5074. doi:10.1039/c4nr00402g
84. Ji S, Cho GY, Yu W, Su P-C, Lee MH, Cha SW (2015) Plasma-enhanced atomic layer
deposition of nanoscale yttria-stabilized zirconia electrolyte for solid oxide fuel cells with
porous substrate. ACS Appl Mater Interfaces 7:2998–3002. doi:10.1021/am508710s
85. Kim J-W (1999) Polarization effects in intermediate temperature, anode-supported solid
oxide fuel cells. J Electrochem Soc 146:69. doi:10.1149/1.1391566
86. Sun C, Xie Z, Xia C, Li H, Chen L (2006) Investigations of mesoporous CeO2-Ru as a
reforming catalyst layer for solid oxide fuel cells. Electrochem Commun 8:833–838. doi:10.
1016/j.elecom.2006.03.018
87. Tada M, Bal R, Mu X, Coquet R, Namba S, Iwasawa Y (2007) Low-temperature PROX
(preferential oxidation) on novel CeO(2)-supported Cu-cluster catalysts under fuel-cell
operating conditions. Chem Commun (Camb) 2:4689–4691. doi:10.1039/b709176a
88. Sun C, Stimming U (2007) Recent anode advances in solid oxide fuel cells. J Power Sources
171:247–260. doi:10.1016/j.jpowsour.2007.06.086
89. Xu C, Tian Z, Shen P, Jiang SP (2008) Oxide (CeO2, NiO, Co3O4 and Mn3O4)-promoted
Pd/C electrocatalysts for alcohol electrooxidation in alkaline media. Electrochim Acta
53:2610–2618. doi:10.1016/j.electacta.2007.10.036
90. Ye KH, Zhou SA, Zhu XC, Xu CW, Shen PK (2013) Stability analysis of oxide (CeO2, NiO,
Co3O4 and Mn3O4) effect on Pd/C for methanol oxidation in alkaline medium. Electrochim
Acta 90:108–111. doi:10.1016/j.electacta.2012.12.012
91. Scibioh MA, Kim S-K, Cho EA, Lim T-H, Hong S-A, Ha HY (2008) Pt-CeO2/C anode
catalyst for direct methanol fuel cells. Appl Catal B Environ 84:773–782. doi:10.1016/j.
apcatb.2008.06.017
92. Avgouropoulos G, Ioannides T, Papadopoulou C (2002) A comparative study of Pt/γ-Al2O3,
Au/α-Fe2O3 and CuO-CeO2 catalysts for the selective oxidation of carbon monoxide in
excess hydrogen. Catal Today 75:157–167
93. Yu HB, Kim J-H, Lee H-I, Scibioh MA, Lee J, Han J et al (2005) Development of nanophase
CeO2-Pt/C cathode catalyst for direct methanol fuel cell. J Power Sources 140:59–65. doi:10.
1016/j.jpowsour.2004.08.015
94. Guo JW, Zhao TS, Prabhuram J, Chen R, Wong CW (2006) Development of PtRu-CeO2/C
anode electrocatalyst for direct methanol fuel cells. J Power Sources 156:345–354. doi:10.
1016/j.jpowsour.2005.05.093
95. Ou DR, Mori T, Togasaki H, Takahashi M, Ye F, Drennan J (2011) Microstructural and
metal-support interactions of the Pt-CeO2/C catalysts for direct methanol fuel cell applica-
tion. Langmuir 27:3859–3866. doi:10.1021/la1032898
13 Effects of Catalyst-Support Materials on the Performance of Fuel Cells 545

96. Liu Y, Fu Q, Stephanopoulos MF (2004) Preferential oxidation of CO in H2 over CuO-CeO2


catalysts. Catal Today 93–95:241–246. doi:10.1016/j.cattod.2004.06.049
97. Lei M, Wang ZB, Li JS, Tang HL, Liu WJ, Wang YG (2014) CeO2 nanocubes-graphene
oxide as durable and highly active catalyst support for proton exchange membrane fuel cell.
Sci Rep 4:7415. doi:10.1038/srep07415
98. Ioroi T, Akita T, Yamazaki SI, Siroma Z, Fujiwara N, Yasuda K (2006) Comparative study of
carbon-supported Pt/Mo-oxide and PtRu for use as CO-tolerant anode catalysts. Electrochim
Acta 52:491–498. doi:10.1016/j.electacta.2006.05.030
99. Vellacheri R, Unni SM, Nahire S, Kharul UK, Kurungot S (2010) Pt-MoOx-carbon nanotube
redox couple based electrocatalyst as a potential partner with polybenzimidazole membrane
for high temperature polymer electrolyte membrane fuel cell applications. Electrochim Acta
55:2878–2887. doi:10.1016/j.electacta.2010.01.012
100. Ugalde-Reyes O, Hernandez-Maya R, Ocampo-Flores AL, Ramirez FA, Sosa-Hernandez E,
Angeles-Chavez C et al (2014) Study of the electrochemical activities of Mo-modified Pt
catalysts, for application as anodes in direct methanol fuel cells: effect of the aggregation
route. J Electrochem Soc 162:H132–H141. doi:10.1149/2.0521503jes
101. Yan Z, Xie J, Jing J, Zhang M, Wei W, Yin S (2012) MoO2 nanocrystals down to 5 nm as Pt
electrocatalyst promoter for stable oxygen reduction reaction. Int J Hydrogen Energy
37:15948–15955. doi:10.1016/j.ijhydene.2012.08.033
102. Zhang Z, Liu J, Gu J, Cheng L (2014) An overview of metal oxide materials as
electrocatalysts and supports for polymer electrolyte fuel cells. Energy Environ Sci
7:2535–2558. doi:10.1039/c3ee43886d
103. Ma L, Zhao X, Si F, Liu C, Liao J, Liang L et al (2010) A comparative study of Pt/C and
Pt-MoOx/C catalysts with various compositions for methanol electro-oxidation. Electrochim
Acta 55:9105–9112. doi:10.1016/j.electacta.2010.08.034
104. Antolini E, Gonzalez ER (2009) Ceramic materials as supports for low-temperature fuel cell
catalysts. Solid State Ion 180:746–763. doi:10.1016/j.ssi.2009.03.007
105. Parrondo J, Han T, Niangar E, Wang C, Dale N, Adjemian K et al (2014) Platinum supported
on titanium-ruthenium oxide is a remarkably stable electrocatayst for hydrogen fuel cell
vehicles. Proc Natl Acad Sci U S A 111:45–50. doi:10.1073/pnas.1319663111
106. Lo CP, Wang G, Kumar A, Ramani V (2013) TiO2-RuO2 electrocatalyst supports exhibit
exceptional electrochemical stability. Appl Catal B Environ 140–141:133–140. doi:10.1016/
j.apcatb.2013.03.039
107. Shinde VM, Madras G (2013) Synthesis of nanosized Ce0.85 M0.1Ru0.05O2-delta (M ¼ Si,
Fe) solid solution exhibiting high CO oxidation and water gas shift activity. Appl Catal B
Environ 138–139:51–61. doi:10.1016/j.apcatb.2013.02.021
108. Kumar A, Ramani VK (2013) RuO2-SiO2 mixed oxides as corrosion-resistant catalyst
supports for polymer electrolyte fuel cells. Appl Catal B Environ 138–139:43–50. doi:10.
1016/j.apcatb.2013.02.015
109. Thanh Ho VT, Pillai KC, Chou H-L, Pan C-J, Rick J, Su W-N, et al (2011) Robust non-carbon
Ti0.7Ru0.3O2 support with co-catalytic functionality for Pt: enhances catalytic activity and
durability for fuel cells. Energy Environ Sci 4:4194. doi:10.1039/c1ee01522b
110. Ho VTT, Nguyen NG, Pan CJ, Cheng JH, Rick J, Su WN et al (2012) Advanced nanoelec-
trocatalyst for methanol oxidation and oxygen reduction reaction, fabricated as
one-dimensional pt nanowires on nanostructured robust Ti0.7Ru0.3O2 support. Nano Energy
1:687–695. doi:10.1016/j.nanoen.2012.07.007
111. Drew K, Girishkumar G, Vinodgopal K, Kamat P (2005) Boosting fuel cell performance wit a
semiconductor photocatalyst: TiO2/Pt-Ru hybrid catalyst for methanol oxidation. J Phys
Chem B 109:11851–11857
112. Gojković SL, Babić BM, Radmilović VR, Krstajić NV (2010) Nb-doped TiO2 as a support of
Pt and Pt-Ru anode catalyst for PEMFCs. J Electroanal Chem 639:161–166. doi:10.1016/j.
jelechem.2009.12.004
546 P.M. Ejikeme et al.

113. Naeem R, Ahmed R, Ansari MS (2014) TiO2 and Al2O3 promoted Pt/C nanocomposites as
low temperature fuel cell catalysts for electro oxidation of methanol in acidic media. IOP
Conf Ser Mater Sci Eng 60:012031. doi:10.1088/1757-899X/60/1/012031
114. Park KW, Han SB, Lee JM (2007) Photo(UV)-enhanced performance of Pt-TiO2 nanostruc-
ture electrode for methanol oxidation. Electrochem Commun 9:1578–1581. doi:10.1016/j.
elecom.2007.02.020
115. Park K-W, Seol K-S (2007) Nb-TiO2 supported Pt cathode catalyst for polymer electrolyte
membrane fuel cells. Electrochem Commun 9:2256–2260. doi:10.1016/j.elecom.2007.06.
027
116. Sun S, Zhang G, Sun X, Cai M, Ruthkosky M (2012) Highly stable and active Pt/Nb-TiO2
carbon-free electrocatalyst for proton exchange membrane fuel cells. J Nanotechnol
2012:13–15. doi:10.1155/2012/389505
117. Zhao G, Zhao TS, Yan XH, Zeng L (2015) A high catalyst-utilization electrode for direct
methanol fuel cells. Electrochim Acta 164:337–343. doi:10.1016/j.electacta.2015.02.181
118. Zhao L, Wang Z-B, Liu J, Zhang J-J, Sui X-L, Zhang L-M et al (2015) Facile one-pot
synthesis of Pt/graphene-TiO2 hybrid catalyst with enhanced methanol electrooxidation
performance. J Power Sources 279:210–217. doi:10.1016/j.jpowsour.2015.01.023
119. Zhang Z, Zhang Y, He L, Yang Y, Liu S, Wang M et al (2015) A feasible synthesis of
Mn3(PO4)2@BSA nanoflowers and its application as the support nanomaterial for Pt catalyst.
J Power Sources 284:170–177. doi:10.1016/j.jpowsour.2015.03.011
120. Deleebeeck L, Ippolito D, Hansen KK (2015) Catalytic enhancement of carbon black and
coal-fueled hybrid direct carbon fuel cells. J Electrochem Soc 162:F327–F339. doi:10.1149/
2.0761503jes
121. Dong X, Takahashi M, Nagao M, Hibino T (2011) Compact bipolar plate-free direct
methanol fuel cell stacks. Chem Commun (Camb) 47:5292–5294. doi:10.1039/c1cc10493d
122. Dong H-Q, Chen Y-Y, Han M, Li S-L, Zhang J, Li J-S et al (2014) Synergistic effect of
mesoporous Mn2O3-supported Pd nanoparticle catalysts for electrocatalytic oxygen reduction
reaction with enhanced performance in alkaline medium. J Mater Chem A 2:1272. doi:10.
1039/c3ta13585c
123. Menezes PW, Indra A, Sahraie NR, Bergmann A, Strasser P, Driess M (2015) Cobalt-
manganese-based spinels as multifunctional materials that unify catalytic water oxidation
and oxygen reduction reactions. ChemSusChem 8:164–171. doi:10.1002/cssc.201402699
124. Selcuk S, Selloni A (2015) DFT þ U study of the surface structure and stability of Co3O4
(110): dependence on U. J Phys Chem C 4:150427132340009. doi:10.1021/acs.jpcc.5b02298
125. Zou X, Goswami A, Asefa T (2013) Efficient noble metal-free (electro)catalysis of water and
alcohol oxidations by zinc-cobalt layered double hydroxide. J Am Chem Soc
135:17242–17245. doi:10.1021/ja407174u
126. Yeo BS, Bell AT (2011) Enhanced activity of gold-supported cobalt oxide for the electro-
chemical evolution of oxygen. J Am Chem Soc 133:5587–5593. doi:10.1021/ja200559j
127. Frydendal R, Busch M, Halck NB, Paoli EA, Krtil P, Chorkendorff I et al (2015) Enhancing
activity for the oxygen evolution reaction: the beneficial interaction of gold with manganese
and cobalt oxides. ChemCatChem 7:149–154. doi:10.1002/cctc.201402756
128. Walton AS, Fester J, Bajdich M, Arman MA, Osiecki J, Knudsen J et al (2015) Interface
controlled oxidation states in layered cobalt oxide nanoislands on gold. ACS Nano
9:2445–2453
129. Jiao F, Frei H (2010) Nanostructured cobalt and manganese oxide clusters as efficient water
oxidation catalysts. Energy Environ Sci 3:1018. doi:10.1039/c002074e
130. Wang HF, Kavanagh R, Guo YL, Guo Y, Lu G, Hu P (2012) Origin of extraordinarily high
catalytic activity of Co3O4 and its morphological chemistry for CO oxidation at low temper-
ature. J Catal 296:110–119. doi:10.1016/j.jcat.2012.09.005
131. Larmier K, Chizallet C, Raybaud P (2015) Tuning the metal-support interaction by structural
recognition of cobalt-based catalyst precursors. Angew Chem Int Ed 54(23):6824–6827.
doi:10.1002/anie.201502069
13 Effects of Catalyst-Support Materials on the Performance of Fuel Cells 547

132. Jiao F, Frei H (2009) Nanostructured cobalt oxide clusters in mesoporous silica as efficient
oxygen-evolving catalysts. Angew Chem Int Ed 48:1841–1844. doi:10.1002/anie.200805534
133. Melvin AA, Joshi VS, Poudyal DC, Khushalani D, Haram SK (2015) Electrocatalyst on
insulating support?: hollow silica spheres loaded with Pt nanoparticles for methanol oxida-
tion. ACS Appl Mater Interfaces 7:6590–6595. doi:10.1021/am508830h
134. Duval Y, Mielczarski JA, Pokrovsky OS, Mielczarski E, Ehrhardt JJ (2002) Evidence of the
existence of three types of species at the quartz-aqueous solution interface at pH0-10: XPS
surface group quantification and surface complexation modeling. J Phys Chem B
106:2937–2945. doi:10.1021/jp012818s
135. Seger B, Kongkanand A, Vinodgopal K, Kamat PV (2008) Platinum dispersed on silica
nanoparticle as electrocatalyst for PEM fuel cell. J Electroanal Chem 621:198–204. doi:10.
1016/j.jelechem.2007.09.037
136. Liu B, Chen JH, Zhong XX, Cui KZ, Zhou HH, Kuang YF (2007) Preparation and
electrocatalytic properties of Pt-SiO2 nanocatalysts for ethanol electrooxidation. J Colloid
Interface Sci 307:139–144. doi:10.1016/j.jcis.2006.11.027
137. Lin C, Hsu S, Ho W (2015) Using SiO2 nanopowders in anode catalyst layer to improve the
performance of a proton exchange membrane fuel cell at low humidity. J Mater Sci Chem
Eng 3:72–79
138. Vu THT, Tran TTT, Le HNT, Tran LT, Nguyen PHT, Nguyen HT et al (2015) Solvothermal
synthesis of Pt -SiO2/graphene nanocomposites as efficient electrocatalyst for methanol
oxidation. Electrochim Acta 161:335–342. doi:10.1016/j.electacta.2015.02.100
139. Uzunoglu A, Dundar F, Ata A (2015) Modification of vulcan XC-72 for enhanced durability
of PEMFC catalyst layer. Int JRenew Energy Res 5:111–120
140. Pinchuk OA, Dundar F, Ata A, Wynne KJ (2012) Improved thermal stability, properties, and
electrocatalytic activity of sol-gel silica modified carbon supported Pt catalysts. Int J Hydro-
gen Energy 37:2111–2120. doi:10.1016/j.ijhydene.2011.10.093
141. Pinchuk OA, Dundar F, Ata A, Wynne KJ (2012) Improved thermal stability, properties, and
electrocatalytic activity of sol-gel silica modified carbon supported Pt catalysts. Int J Hydro-
gen Energy 37:2111–2120. doi:10.1016/j.ijhydene.2011.10.093
142. Inaba M, Suzuki T, Hatanaka T, Morimoto Y (2015) Fabrication and cell analysis of a
Pt/SiO2 platinum thin film electrode. J Electrochem Soc 162:F634–F638. doi:10.1149/2.
0201507jes
143. Dundar F, Uzunoglu A, Ata A, Wynne KJ (2012) Durability of carbon-silica supported
catalysts for proton exchange membrane fuel cells. J Power Sources 202:184–189. doi:10.
1016/j.jpowsour.2011.12.010
144. Su C, Hsueh Y, Kei C, Lin C, Perng T (2013) Fabrication of high-activity hybrid Pt@ZnO
catalyst on carbon cloth by atomic layer deposition for photoassisted electro-oxidation of
methanol. J Phys Chem C
145. Yun M, Ahmed MS, Jeon S (2015) Thiolated graphene oxide-supported palladium cobalt
alloyed nanoparticles as high performance electrocatalyst for oxygen reduction reaction. J
Power Sources 293:380–387. doi:10.1016/j.jpowsour.2015.05.094
146. Ilinich O, Liu Y, Castellano C, Koermer G, Moini A, Farrauto R (2008) A new palladium-
based catalyst for methanol steam reforming in a miniature fuel cell power source. Platin Met
Rev 52:134–143. doi:10.1595/147106708X324403
147. Chen LY, Chang BK, Lu Y, Yang WG, Tatarchuk BJ (2002) Selective catalytic oxidation of
CO in H2: fuel cell applications. Abstr Pap Am Chem Soc 224:249–254. doi:10.1016/S0920-
5861(00)00426-0
148. Korotkikh O, Farrauto R (2000) Selective catalytic oxidation of CO in H-2: fuel cell
applications. Catal Today 62:249–254
149. Lo C-P, Wang G, Kumar A, Ramani V (2011) RuO2/xH2O-TiO2 as catalyst support for PEM
fuel cells. ECS Trans 41:1249–1255. doi:10.1149/1.3635656
150. Horiguchi D, Tsukatsune T, Noda Z, Hayashi A, Sasaki K (2014) Pt/SnO2 electrocatalysts on
conductive fillers. ECS Trans 64:215–220. doi:10.1149/06403.0215ecst
548 P.M. Ejikeme et al.

151. Kinumoto T, Kitayama S, Matsuoka M, Tsumura T, Toyoda M (2014) Correlation between


preparation condition and performance of Pt/SnO2/KB for cathode catalyst of PEMFC. ECS
Trans 64:199–205
152. Nagamatsu Y, Kanda K, Noda Z, Hayashi A, Sasaki K (2014) Electrochemical performance
of MEAs with Pt/SnO2 mixed with conductive fillers. ECS Trans 64:207–213
153. Xu C, Shen PK, Ji X, Zeng R, Liu Y (2005) Enhanced activity for ethanol electrooxidation on
Pt-MgO/C catalysts. Electrochem Commun 7:1305–1308. doi:10.1016/j.elecom.2005.09.015
154. Li N, Zeng Y, Chen S, Xu C, Shen P (2014) Ethanol oxidation on Pd/C enhanced by MgO in
alkaline medium. Int J Hydrogen Energy 39:1–5. doi:10.1016/j.ijhydene.2013.12.122
155. Mahendiran C, Maiyalagan T, Scott K, Gedanken A (2011) Synthesis of a carbon-coated
NiO/MgO core/shell nanocomposite as a Pd electro-catalyst support for ethanol oxidation.
Mater Chem Phys 128:341–347. doi:10.1016/j.matchemphys.2011.02.067
156. Chu D, Wang J, Wang S, Zha L, He J, Hou Y et al (2009) High activity of Pd-In2O3/CNTs
electrocatalyst for electro-oxidation of ethanol. Catal Commun 10:955–958. doi:10.1016/j.
catcom.2008.12.041
157. Liang R, Hu A, Persic J, Zhou YN (2013) Palladium nanoparticles loaded on carbon modified
TiO2 nanobelts for enhanced methanol electrooxidation. Nano Micro Lett 5:202–212. doi:10.
5101/nml.v5i3.p202-212
158. Liu B, Chen JH, Xiao CH, Cui KZ, Yang L, Pang HL et al (2007) Preparation of Pt/MgO/
CNT hybrid catalysts and their electrocatalytic properties for ethanol electrooxidation.
Energy Fuels 21:1365–1369. doi:10.1021/ef060452i
159. Han X, Zhang T, Du J, Cheng F, Chen J (2013) Porous calcium–manganese oxide micro-
spheres for electrocatalytic oxygen reduction with high activity. Chem Sci 4:368–376. doi:10.
1039/c2sc21475j
160. Stottlemyer AL, Kelly TG, Meng Q, Chen JG (2012) Reactions of oxygen-containing
molecules on transition metal carbides: surface science insight into potential applications in
catalysis and electrocatalysis. Surf Sci Rep 67:201–232. doi:10.1016/j.surfrep.2012.07.001
161. Chen JG (1996) Carbide and nitride overlayers on early transition metal surfaces: prepara-
tion, characterization, and reactivities. Chem Rev 96:1477–1498. doi:10.1021/cr950232u
162. Oyama ST (1992) Preparation and catalytic properties of transition metal carbides and
nitrides. Catal Today 15:179–200. doi:10.1016/0920-5861(92)80175-M
163. Weidman MC, Esposito DV, Hsu Y-C, Chen JG (2012) Comparison of electrochemical
stability of transition metal carbides (WC, W2C, Mo2C) over a wide pH range. J Power
Sources 202:11–17. doi:10.1016/j.jpowsour.2011.10.093
164. Yan Z, He G, Cai M, Meng H, Shen PK (2013) Formation of tungsten carbide nanoparticles
on graphitized carbon to facilitate the oxygen reduction reaction. J Power Sources
242:817–823. doi:10.1016/j.jpowsour.2013.05.161
165. Wang YJ, Wilkinson DP, Zhang JJ (2011) Noncarbon support materials for polymer elec-
trolyte membrane fuel cell electrocatalysts. Chem Rev 111:7625–7651
166. Yan Z, Gao L, Zhang M, Xie J, Chen M (2015) Angstrom-scale vanadium carbide rods as Pt
electrocatalyst support for efficient methanol oxidation reaction. RSC Adv 5:9561–9564.
doi:10.1039/C4RA11798K
167. Regmi YN, Waetzig GR, Duffee KD, Schmuecker SM, Thode JM, Leonard BM (2015)
Carbides of group IVA, VA and VIA transition metals as alternative HER and ORR catalysts
and support materials. J Mater Chem A 3:10085–10091. doi:10.1039/C5TA01296A
168. Liu X, Fechler N, Antonietti M (2013) Salt melt synthesis of ceramics, semiconductors and
carbon nanostructures. Chem Soc Rev 42:8237–8265. doi:10.1039/c3cs60159e
169. Bennett LH, Cuthill JR, McAlister AJ, Erickson NE, Watson RE (1974) Electronic structure
and catalytic behavior of tungsten carbide. Science 181:563–565. doi:10.1126/science.184.
4136.563
170. Levy RB, Boudart M (1973) Platinum-like behavior of tungsten carbide in surface catalysis.
Science 181(1973):547–549. doi:10.1126/science.181.4099.547
13 Effects of Catalyst-Support Materials on the Performance of Fuel Cells 549

171. Kimmel YC, Esposito DV, Birkmire RW, Chen JG (2012) Effect of surface carbon on the
hydrogen evolution reactivity of tungsten carbide (WC) and Pt-modified WC electrocatalysts.
Int J Hydrogen Energy 37:3019–3024. doi:10.1016/j.ijhydene.2011.11.079
172. Gong XB, You SJ, Wang XH, Gan Y, Zhang RN, Ren NQ (2013) Silver-tungsten carbide
nanohybrid for efficient electrocatalysis of oxygen reduction reaction in microbial fuel cell. J
Power Sources 225:330–337. doi:10.1016/j.jpowsour.2012.10.047
173. Houston JE, Laramore GE, Park RL (1974) Surface electronic properties of tungsten,
tungsten carbide, and platinum. Science 185:258–260
174. Chen W-F, Muckerman JT, Fujita E (2013) Recent developments in transition metal carbides
and nitrides as hydrogen evolution electrocatalysts. Chem Commun (Camb) 49:8896–8909.
doi:10.1039/c3cc44076a
175. Ham DJ, Lee JS (2009) Transition metal carbides and nitrides as electrode materials for low
temperature fuel cells. Energies 2:873–899. doi:10.3390/en20400873
176. Esposito DV, Chen JG (2011) Monolayer platinum supported on tungsten carbides as
low-cost electrocatalysts: opportunities and limitations. Energy Environ Sci 4:3900. doi:10.
1039/c1ee01851e
177. Hsu IJ, Kimmel YC, Dai Y, Chen S, Chen JG (2012) Rotating disk electrode measurements of
activity and stability of monolayer Pt on tungsten carbide disks for oxygen reduction reaction.
J Power Sources 199:46–52. doi:10.1016/j.jpowsour.2011.10.024
178. Lu JL, Li ZH, Jiang SP, Shen PK, Li L (2012) Nanostructured tungsten carbide/carbon
composites synthesized by a microwave heating method as supports of platinum catalysts for
methanol oxidation. J Power Sources 202:56–62. doi:10.1016/j.jpowsour.2011.11.018
179. Yan Z, Cai M, Shen PK (2013) Nanosized tungsten carbide synthesized by a novel route at
low temperature for high performance electrocatalysis. Sci Rep 3:1646. doi:10.1038/
srep01646
180. Yang M, Cui Z, DiSalvo FJ (2013) Mesoporous chromium nitride as a high performance
non-carbon support for the oxygen reduction reaction. Phys Chem Chem Phys 15:7041–7044.
doi:10.1039/c3cp51109j
181. Cui Z, Burns RG, DiSalvo FJ (2013) Mesoporous Ti0.5Nb0.5N ternary nitride as a novel non
carbon support for oxygen reduction reaction in acid and alkaline electrolytes. Chem Mater
25:3782–3784. doi:10.1021/cm4027545
182. Yang M, Guarecuco R, Disalvo FJ (2013) Mesoporous chromium nitride as high performance
catalyst support for methanol electrooxidation. Chem Mater 25:1783–1787. doi:10.1021/
cm400304q
183. Yang M, Cui Z, DiSalvo FJ (2013) Mesoporous titanium nitride supported Pt nanoparticles as
high performance catalysts for methanol electrooxidation. Phys Chem Chem Phys
15:1088–1092. doi:10.1039/C2CP44215A
184. Cui Z, Yang M, Di Salvo FJ (2013) Mo2N/C hybrid material as a promising support for the
electro-oxidation of methanol and formic acid. Electrochem Commun 33:63–67. doi:10.
1016/j.elecom.2013.04.017
185. Thotiyl MMO, Sampath S (2011) Electrochemical oxidation of ethanol in acid media on
titanium nitride supported fuel cell catalysts. Electrochim Acta 56:3549–3554. doi:10.1016/j.
electacta.2010.12.091
186. Thotiyl MMO, Kumar TR, Sampath S (2010) Pd supported on titanium nitride for efficient
ethanol oxidation. J Phys Chem C 114:17934–17941. doi:10.1021/jp1038514
187. Cesiulis H, Ziomek-Moroz M (2000) Electrocrystallization and electrodeposition of silver on
titanium nitride. J Appl Electrochem 30:1261–1268. doi:10.1023/A:1026553712521
188. Evans SAG, Terry JG, Plank NOV, Walton AJ, Keane LM, Campbell CJ et al (2005)
Electrodeposition of platinum metal on TiN thin films. Electrochem Commun 7:125–129.
doi:10.1016/j.elecom.2004.11.014
189. Musthafa OTM, Sampath S (2008) High performance platinized titanium nitride catalyst for
methanol oxidation. Chem Commun (Camb) 7:67–69. doi:10.1039/b715859a
550 P.M. Ejikeme et al.

190. Kakinuma K, Wakasugi Y, Uchida M, Kamino T, Uchida H, Deki S et al (2012) Preparation


of titanium nitride-supported platinum catalysts with well controlled morphology and their
properties relevant to polymer electrolyte fuel cells. Electrochim Acta 77:279–284. doi:10.
1016/j.electacta.2012.06.001
191. Avasarala B, Murray T, Li W, Haldar P (2009) Titanium nitride nanoparticles based
electrocatalysts for proton exchange membrane fuel cells. J Mater Chem 19:1803. doi:10.
1039/b819006b
192. Higgins DC, Choi J-Y, Wu J, Lopez A, Chen Z (2012) Titanium nitride–carbon nanotube
core–shell composites as effective electrocatalyst supports for low temperature fuel cells. J
Mater Chem 22:3727. doi:10.1039/c2jm15014j
193. Jia Y, Wang Y, Dong L, Huang J, Zhang Y, Su J et al (2015) A hybrid of titanium nitride and
nitrogen-doped amorphous carbon supported on SiC as a noble metal-free electrocatalyst for
oxygen reduction reaction. Chem Commun 51:2625–2628. doi:10.1039/C4CC08007F
194. Pan Z, Xiao Y, Fu Z, Zhan G, Wu S, Xiao C et al (2014) Hollow and porous titanium nitride
nanotubes as high performance catalyst support for oxygen reduction reaction. J Mater Chem
A 2:13966–13975. doi:10.1039/c4ta02402h
195. Chen J, Takanabe K, Ohnishi R, Lu D, Okada S, Hatasawa H et al (2010) Nano-sized TiN on
carbon black as an efficient electrocatalyst for the oxygen reduction reaction prepared using
an mpg-C3N4 template. Chem Commun (Camb) 46:7492–7494. doi:10.1039/c0cc02048f
196. Alhajri NS, Yoshida H, Anjum DH, Garcia-Esparza AT, Kubota J, Domen K et al (2013)
Synthesis of tantalum carbide and nitride nanoparticles using a reactive mesoporous template
for electrochemical hydrogen evolution. J Mater Chem A 1:12606. doi:10.1039/c3ta12984e
197. Dong Y, Wu Y, Liu M, Li J (2013) Electrocatalysis on shape-controlled titanium nitride
nanocrystals for the oxygen reduction reaction. ChemSusChem 6:2016–2021
198. Xiao Y, Zhan G, Fu Z, Pan Z, Xiao C, Wu S et al (2015) Titanium cobalt nitride supported
platinum catalyst with high activity and stability for oxygen reduction reaction. J Power
Sources 284:296–304. doi:10.1016/j.jpowsour.2015.03.001
199. Xiao Y, Fu Z, Zhan G, Pan Z, Xiao C, Wu S et al (2015) Increasing Pt methanol oxidation
reaction activity and durability with a titanium molybdenum nitride catalyst support. J Power
Sources 273:33–40. doi:10.1016/j.jpowsour.2014.09.057
200. Xiao Y, Zhan G, Fu Z, Pan Z, Xiao C, Wu S et al (2014) Robust non-carbon titanium nitride
nanotubes supported Pt catalyst with enhanced catalytic activity and durability for methanol
oxidation reaction. Electrochim Acta 141:279–285. doi:10.1016/j.electacta.2014.07.070
201. Mkwizu TS, Mathe MK, Cukrowski I (2010) Electrodeposition of multilayered bimetallic
nanoclusters of ruthenium and platinum via surface-limited redox-replacement reactions for
electrocatalytic applications. Langmuir 26:570–580. doi:10.1021/la902219t
202. Modibedi RM, Louw EK, Mathe MK, Ozoemena KI (2013) The electrochemical atomic layer
deposition of Pt and Pd nanoparticles on Ni foam for the electro oxidation of alcohols. ECS
Trans 50:9–18
Chapter 14
Applications of Nanomaterials in Microbial
Fuel Cells

R. Fogel and J. L. Limson

14.1 Principles of Microbial Fuel Cells

Microbial fuel cells are biotechnological devices that transform the chemical
energy present in biodegradable organic compounds into electrical energy, primar-
ily through the use of microorganisms as catalysts for oxidation/reduction reactions
[1, 2]. As is the convention in electrochemical cells, a fuel cell can be broadly
categorised into possessing an anodic compartment i.e. the site of the electron
source reactions and a cathodic compartment, where reducing reactions accept
the electron. The most commonly-reported configurations of MFCs make use of
biologically modified anodes (bioanodes), but research into the use of biocathodes
are increasingly common [1, 2]. Anodes and cathodes can be either physical
separation of the electrode types, or conceptual divisions of the reactions that
occur in MFCs. A physical separation is considered to be useful in that it provides
separate environmental conditions required for efficient anodic and cathodic reac-
tions to occur. Typically, the anode and the cathode are separated from one another
by a cation-exchange membrane, or CEM, in order to facilitate the transport of
biogenerated protons from the anode to the cathode (Fig. 14.1).
The primary purpose of microbial fuel cells has traditionally been perceived to
be the recovery of energy, either through direct electrical energy generation, or to
energistically support the production of energy-rich compounds e.g. H2 from waste
[2, 3]. Due to the inclusion of metabolising bacteria within this technology, they
bear the additional promise of combining bioremediation processes with power
generation [2, 3]. Increasing emphasis has been placed on the utilisation of different
wastewaters and the potential that MFCs hold for bioremediation of specific wastes,
either as a standalone technology or coupled to other wastewater treatment

R. Fogel • J.L. Limson (*)


Biotechnology Innovation Centre, Rhodes University, Grahamstown 6140, South Africa
e-mail: [email protected]; [email protected]

© Springer International Publishing Switzerland 2016 551


K.I. Ozoemena, S. Chen (eds.), Nanomaterials for Fuel Cell Catalysis,
Nanostructure Science and Technology, DOI 10.1007/978-3-319-29930-3_14
552 R. Fogel and J.L. Limson

Fig. 14.1 Schematic diagram of general components of a conventional bioanodic microbial fuel
cells fuelled with oxygen, one of the most common examples of microbial fuel cell configurations
available in current literature. Electrons generated through catabolic microbial processes enter the
anode and travel via the external circuit to the cathode. Protons generated through microbial
processes diffuse through the Cation Exchange Membrane (CEM) to the cathode to react in
reduction reactions

processes such as anaerobic digesters. As such, a very large list of dissolved organic
compounds, ranging from complex lignocellulose fractions [4], abundant and
nutrient-rich domestic [2] and industrial [4] wastewaters, toxic phenols [5], as
well as inorganic nitrates, sulphates, and heavy metals [3] have been studied as to
their ability to support waste bioremediation technologies. Given this broad scope,
research into the application of nanomaterials to address enduring concerns in MFC
design and architecture takes this duality of purpose into consideration, with several
such studies conducted using specific wastewaters in research aimed at real
applications.
Considering the core purpose of microbial fuel cells, research is primarily
addressed at improving both power output and to a lesser extent, enhancing
bioremediation. Here issues of the nature of the bacteria, viability and lifespan,
nature of the electrode materials with respect to cost, stability and biocompatibility
are frequently considered along with foci aimed at MFC architecture around issues
of scalability in particular.
Currently, as a direct means of generating electrical power, microbial fuel cells
are too uneconomic to manufacture or operate at the scale required to process the
volumes of wastewater to make this technology viable. Several key aspects limiting
this technology have been identified across several different publications and can be
summarised as: the reliance on expensive catalysts/prohibitive amounts of
14 Applications of Nanomaterials in Microbial Fuel Cells 553

catalysts, especially platinum [6, 7]; ineffective use of catalysts e.g. poor dispersal
[7]; the reliance on expensive reagents/materials for optimum function, especially
the addition of electron transfer mediators and electrode materials [6]; toxicity of
electron transfer mediators [8]; low power generation [8]; poor durability of
electrode materials [6, 7]; poor scaleability; poor mass transport across the system
and within compartments [7] and the requirement for external power input for
optimum function e.g. agitation/heat.
Electrode materials for any low-cost industrial process, such as MFCs, should be
mainly comprised of low-cost materials that remain chemically stable during
operating lifetimes, have a high surface area and be easily scaled-up [2]. Due to
the biological nature of microbial fuel cells, electrodes interacting with the
biocatalysts should, ideally, promote cellular adhesion (or, at least, not be inimical
to bacteria in the anode) and possess a limited tendency towards chemo- and
bio-fouling, to improve their long-term operational times. For these outlined rea-
sons, carbonaceous electrodes are often included as electrode materials.
Given the cost-effectiveness and ease-of-production, combined with their chem-
ical stability, good biocompatibility and good conductivity of many of the carbon-
based materials, there is an obvious research interest in incorporating them as
macroscale electrode materials for fuel cells. However, compared to their metal
counterparts, the carbonaceous electrodes suffer from relatively poor conductivity
and high electrochemical overpotentials.
Considering the above, the application of nanomaterials and research within this
area thus encompasses several aspects, most notably, electrode modification and the
potential thereof for impacting (1) anodic electron transferpower processes,
(2) mass transport within the system and (3) cathodic electron transfer processes.
As such, in this chapter, we overview how nanomaterials may have a positive
impact on key limitations in terms of electrode materials and surface area, as
catalysts in the oxygen reduction reaction, and as mediators for enhancing bacterial
electron transfer efficiency.

14.2 Nanostructured Electrode Materials in Microbial


Fuel Cells

Table 14.1 provides a brief list of samples of the available microbial fuel cells
reported in recent literature, emphasising those where nanomaterial modifications
were investigated. Where available, power density values prior to and after
nanomaterial inclusion at the anode or cathode (as indicated) of the MFC config-
urations are listed, with associated percentage increases noted.
Given the diversity of electrode materials used for MFC fabrication, microbial
catalysts selected and organic substrates investigated as well as nanomaterials
utilised in modifying one or both of the electrodes, it is difficult to create any but
the most general comparisons from data presented in Table 14.1. Indeed, this lack
of standardisation was identified previously [23], but is still limiting the ability of
Table 14.1 outputs of various reported MFCs present in the current literature (2014 and previous)
554

Anolyte (bacteria)
Catholyte Anode modifiers Cathode modifiers P.D. (mW m2)a % increaseb Ref.
Anaerobic sludge consortium Carbon cloth Carbon cloth 468 189 [9]
50 mM Fe(CN)63/4 in PBS GO/PANI NFs Carbon cloth 1390
Palm oil mill effluent(consort.) Carbon paper Carbon paper 38 [10]
– Copper Pc 118 210
– Pt 121 218
PB with NH4Cl – Ni NPs 94 147
Unreported Carbon Carbon 32 272 [11]
Air cathode – MnO2 119
Urban wastewater (consort.) PB Graphite Pd/TiO2/Ti 200 [12]
Domestic wastewater (consort.) Carbon Carbon 306 18 [13]
PB  Fe(CN)63/4 Au NPs – 363
(Escherichia coli K12) Polymer/Pt FePc NPs þ carbon 2957 [14]
PB, pH 3.5
Synthetic nutrient medium (consort.) Carbon paper Pt carbon paper Carageenan 80 15 [15]
PBS with NH4Cl – NPs þ polypyrrole 72
(Clostridium butyricum) PTFE þ Platinised Pt. Graphitised carbon 260,000 [16]
Fe(CN)63/4 in PB, pH 7.0
1 g/L glucose PBS (E. coli K12) Toray carbon paper Pt. carbon 468 32 [17]
– Co/Fe/N/graphite 161
Air cathode – Co/Fe/N/CNTs 751
2 g/L glucose medium (E. coli) Carbon felt Pt/carbon paper ~80 [18]
MoO2C – ~170 113
CNTs – ~180 125
Air cathode MoO2C/CNTs – ~1050 1215
R. Fogel and J.L. Limson
14

Synthetic medium (Enterobacter cloacae) Carbon paper Carbon paper 4 [19]


– Carbon Black/FePc 17 323
PB, pH 7 Carbon Black Carbon Black/FePc 43 975
Synthetic medium (E. cloacae) PB Carbon paper Carbon paper 2.3 [20]
CNTs/FePc CNTs/FePc 44 1913
Synthetic medium (sludge consortium) Carbon cloth Carbon cloth 430 330 [21]
Air cathode – MnO2 NT/GO 1850
Glucose (E. coli) Glassy carbon Pt 457 [22]
MWCNT – 699 53
PB  Fe(CN6)3/4 MWCNT/SnO2 – 1421 211
Underlined text in column 1 indicates the catholyte
Abbreviations: AuNPs gold nanoparticles, CNTs carbon nanotubes, consort consortium, FePc Iron(II) Phthalocyanine, GO graphene oxide, NFs nanofibers,
NPs nanoparticles, NT nanotubes, PB/S phosphate-buffered/saline, PANI polyaniline
a
Maximum power density (P.D.) reported during the study
b
Percentage increase of maximum power density, as compared to electrode configuration lacking nanomaterial modification reported in that study
Applications of Nanomaterials in Microbial Fuel Cells
555
556 R. Fogel and J.L. Limson

the MFC research community to unify findings and provide an underlying sense of
progress. As demonstrated in Table 14.1, microbial fuel cells are generally
characterised by having quite low reported power densities than other contempo-
rary electrical energy generating devices. Reported peak power densities obtained
at microbial fuel cells using conventional electrodes typically range from the
mW m2, and rarely, to the low W m2 of reported electrode surface area
(Table 14.1) and is an often-cited limitation to the broader application of MFC
[24]. This range of power generation is cited variously as 2–3 orders of magnitude
lower [25] than other energy-generating technologies, including their next coun-
terparts—chemical fuel cells. Compared to reported power density of catalysed
direct methanol fuel cells operating under STP (~14,000 mW m2 in [26]), the
power generating ability of microbial fuel cells are currently very low. The pro-
posed target of recoverable energy of ~1 kWhr kg1 of organic pollutant removed
and generation capacity of 1 kW m3 [3] can be achieved with very few of these
systems.
This is as a result of multiple intersecting factors that lower the overall obtain-
able power densities (Fig. 14.1). Major discrepancies are notable in the literature,
mainly attributed to the selection of the bacteria and overall differences to the
configuration and design of the fuel cells. This uniformly low power response
however remains the single greatest challenge to the field. Arguably, this is the
area of MFC research to which nanotechnology can yield the greatest results and
has already demonstrated great promise. Overall, in the literature and as indicated
with specific examples in Table 14.1, the rational inclusion of nanomaterials as
electrode materials/modifiers appears to have significantly improved the power
generation of the MFC systems. From the data indicated in Table 14.1, for example,
the utilisation of carbon nanotubes and FePc at the cathode serves to synergistically
enhance power density by 1913 % relative to the use of unmodified carbon paper
electrodes (e.g. [20] in Table 14.1). Similarly, increases in power density were
observed for modifications of the anode, with values of 125 % increase at metal
oxide modified electrodes with the inclusion of carbon nanotubes by Wang et al. in
2014 ([18] in Table 14.1).
Considering this, lowering the cost of catalysts is an area where nanomaterials
also play a particular role, platinum-based catalysts being a case in point. Indeed, the
cost-benefit of including nanomaterials based catalysts are frequently touted as a
reason for their study. Given the role of platinum in catalysing the oxygen reduction
reaction, several papers compare the application of metallophthalocyanine-based
catalysts (in the presence and absence of nanomaterials) as alternatives to platinum.
Reported power densities obtained at MFCs operated with cathodes modified with
copper phthalocyanines [10] and iron phthalocyanines [20] were comparable to
those operating using platinum-modified cathodes. The cost of nanomaterials is
also an important consideration in any scale up operation. For this purpose,
Mshoperi et al., 2014 [19] examined untreated carbon black as an alternative support
for iron phthalocyanine to carbon nanotubes, obtaining near identical power density
values to studies using acid functionalised carbon nanotubes. Such lowering in cost
and complexity of pre-treatment holds obvious utility in the scale up and real
application of MFC technology.
14 Applications of Nanomaterials in Microbial Fuel Cells 557

Microbial Fuel cells


Mass transfer

Direct energy Assisted generaon


generaon of fuels e.g. H2
Catalyst dispersal
and support

Bioc reacons Abioc reacons

Direct microbe-
electrode electron Redox catalysts
transfer

Fig. 14.2 Overview of the current limitations of the major divisions of fuel cell functions and
processes (black text, white background) and the limitations in fuel cell technology addressable by
nanotechnology (white text, grey background)

In order to further contextualise the role that nanomaterials may play, it is


necessary to consider the inherent limitations of MFC technology, summarised
briefly in Fig. 14.2.

14.3 Inherent Limitations of Microbial Fuel Cells

14.3.1 Thermodynamic Limitations of Using


Biological Systems

Part of the power limitations reported at bioanodic microbial fuel cells and noted in
Table 14.1 arises due to inherent constraints of the technology, but may be
addressable by futuristic technologies. This limitation arises because living organ-
isms are responsible for the initial oxidation reactions, liberating electrons from
carbonaceous molecules. The final processes of energy generation in biological
systems is mainly electrochemical—the chemical energy present in nutrient com-
pounds is converted to reducing equivalents (e and H+) and stored in the reduced
forms of the high-energy redox intermediates, NAD(H) and FAD(H2). A series of
biochemical electron-transfer reactions serves to convert this electrical energy into
a proton imbalance across membranes, the relaxation of which drives ATP synthe-
sis [23, 27].
558 R. Fogel and J.L. Limson

A portion, typically the majority, of the harvestable energy arising from catab-
olism is retained and used by the biocatalyst to maintain cellular processes, rather
than entering the external circuit via the anode. Despite billions of years of
evolution, the energy yield from the metabolism of as simple a compound as
glucose is only between 55 and 65 % energy efficient [27]. When one considers
that the majority of electron-transport reactions occur on the intracellular face of
cell membranes and are (currently) unavailable for transfer to the external circuit,
the very nature of cellular metabolism imposes a severe restriction on the amount of
energy that is harvestable from the system. Conversely, chemical and enzymatic
fuel cells do not suffer from similar restrictions, with the additional benefit in
chemical fuel cells that temperature and pressure of the reactants can be varied
for optimal performances without degrading the catalysts.
The potential energy of electrons exiting the bacteria is low. Comparison of the
formal reduction potentials between the penultimate cellular electron acceptors
(members of the cytochrome protein family, cytochromes a and cyotochrome c,
have formal reduction potentials of þ0.633 and þ0.703 V vs. SHE, respectively)
and the most common terminal electron acceptor molecule, O2 (formal reduction
potential of þ1.229 V [27]), sets the limit of maximum potential difference reported
to be between 0.5 and 0.8 V at OCV, under physiological conditions [6]. While
there might indeed be technological interventions in the future that might address
this limitation, these are currently relegated to conjectural stages of research.
Efforts at improving the economic sustainability and the cost-effectiveness of
MFCs is rather a more achievable and immediate manner of improving the imple-
mentation and deployment of MFC technology [8]. Despite this, engineering and
design interventions have demonstrated power densities of >2.4 W/m2 electrode
surface area [8, 23], effectively demonstrating the amount of harvestable energy
that is available from these systems without complex electrode engineering and
materials sciences.
Alterations to the rate and efficiency of electron transfer have been investigated
and optimised under several conditions: electrode selection and configuration [2],
microbial fuel cell design and configuration, composition and concentration of
electrolytes [23], presence, composition and type of permiselective membranes
(if any).

14.3.2 Mass Transfer Limitations

As with any heterogeneous chemical process, microbial fuel cells suffer from mass
transfer limitations. Two separate classes of reactions of typical MFCs have been
determined to limit power generation: mass transfer supplying substrates to cata-
lysts and removing products within an electrode’s compartment [28] and the
transfer of charge carriers e.g. H+ between compartments, usually separated
through a permiselective membrane, e.g., Nafion [29, 30]. Mass transfer limitations
14 Applications of Nanomaterials in Microbial Fuel Cells 559

affect both cell voltage (as described by the classical Nernst equation) and the
overall power obtainable, by decreasing the obtainable current [28].
An important criterion in microbial fuel cell technology is the surface- and
system-engineering of energy-efficient ways of improving desirable mass-transfer
reactions, such as those mentioned above, while minimising undesirable mass
transfer reactions, e.g., the diffusion of oxygen from aerobic compartments to
anaerobic compartments [29, 30], the diffusion of microbial cells to abiotic com-
partments, etc. Solution agitation, while noted to produce significant improvements
on the power generation of fuel cells e.g. [19], requires too much energy input to
produce a sustainable solution.
One of the most-often cited methods of improving the effectiveness of MFC
power generation overall is the improvement of effective electrode surface areas.
This not only increases the catalyst loading of electrodes, but also improves the
mass transfer rates between the electrode and the compartment space, through
enhanced surface area to volume ratios. It is in this area that the use of conductive
nanoparticles and nanomaterials play significant roles as non-catalytic electrode
modifiers. The translation of materials from the macro- and microscale to the
nanoscale results in improvements of surface areas of several orders of magnitude.
Table 14.2 shows that carbon materials can range from a macropscopic size of
0.005 m2 g1 (carbon cloth, Table 14.2, [32]) to 1662 m2 g1 (reported for activated
carbon powder [33]), an impressive span of five orders of magnitude. Indeed it is
this increase in effective surface area that is most commonly attributed with the
increases in power density with the nanomodification of electrodes recorded in
Table 14.1, particularly when studying carbonaceous nanomaterials.

Table 14.2 Reported and calculated surface areas of common electrode materials and nanoscaled
electrode modifiers
Material Surface area m2 g1(method) Reference
Micro- and macroscopic materials
Activated carbon granules 0.125 (BET, N2) [31]
Carbon cloth 0.005 (cited) [32]
Carbon fibers 7.11 (cited) [32]
Graphite rods 0.007 (cited) [32]
Graphite felt 0.020 (cited) [32]
Nanoscopic materials
Graphene 264 [32]
Activated carbon powder, SPC-01 1662 (BET) [33]
Graphitic carbon nanostructures 112 (BET) [34]
Graphene oxide—AuNP (φ 20–30 nm) 984 (BET) [33]
Pristine carbon nanotubes 34 (BET) [35]
Acid-functionalised CNTs 85 (BET) [35]
Hydrothermally-synthesised MnO2 nanoparticles 112 (BET) [11]
20 % Pt dispersed on Vulcan XC-72 65 (RRDE Voltammetry) [36]
560 R. Fogel and J.L. Limson

A combination of improved mass transport conditions, and the rational design of


electrodes supporting more facile electron-transfer reactions and nanotechnological
enhancements have all yielded significant overall improvements on the efficiency
and output capacity of microbial fuel cells as evidenced in Table 14.1.
In the rest of this chapter we detail further how these limitations impact on the
MFC system for both the biotic and abiotic processes commonly reported in the
literature, highlighting in each section how nanomaterials can be harnessed to
address these.

14.4 Nanomaterials Supporting Biotic Processes


in Microbial Fuel Cells

Microbial fuel cells can make use of microbial catalysts in both the anodic and
cathodic sections. The most commonly-encountered microbial fuel cells make use
of microbes as anodic catalysts, primarily used in order to biodegrade organic
matter and act as the proton and electron sources which participate in the rest of
the MFC process [2, 3]. Considered as cathodic catalysts, microbes are capable of
acting as intermediate electron acceptors, transferring the incoming electrons to a
variety of researched pollutants and acceptors, such as nitrates, sulphates, perchlo-
rates, metal ions and dioxygen [2, 3, 28, 37]. For either purpose, the method by
which microbial-electrode electron transfer occurs are much the same. As such,
they are discussed simultaneously.
Due to their integral function within the MFC system, the selection of the
microbial catalyst has a profound influence on the nature and the extent of the
performance of the microbial fuel cell. This is due to, firstly, the ability of a
biocatalyst to thrive in the conditions present in the anodic/cathodic chambers
e.g. nutrient availability and the absence/presence of oxygen, pH, pollutant pres-
ence, etc. Secondly, the selection of microbial catalyst influences the manner by
which electrons transfer to the anode and enter the external circuit [3]. Cells owe
much of their metabolic function to electrochemical reactions, through the use of
oxidorectase enzymes. This class of enzymes catalyses the transfer of electron
between reactants, via redox centres embedded in their protein structures, and
participate greatly in the electrochemical reactions that generate cellular energy
[38]. As mentioned previously, the means by which energy is gathered during
catabolism reactions is by a series of electrochemical reactions.
In eukaryote cells, the majority of the catabolic electron-transport metabolic
processes are conducted within the specialised organelles of the mitochondria. This
limits the accessibility of the electrodes to interact with electrons arising from these
processes. Hence, the majority of the microbes studied for microbial fuel cell
technologies are primarily bacterial or archaeal organisms. Nonetheless, research
on some facultative anaerobic eukaryotes, primarily yeasts, are proceeding. The
majority of researched microbes are either obligate anaerobes and facultative
14 Applications of Nanomaterials in Microbial Fuel Cells 561

Table 14.3 Common microbial catalysts investigated in current research and pertinent charac-
teristics—2014
Organism name Cell wall Cited #a Respiration Electron transfer [Ref]
Shewanella spp. Gram-negative 277 Facultative Mixed [23, 40, 41]
Pseudomonas spp. Gram-negative 159 Facultative Self-mediated [42,43]
Geobacter sulfurreducens Gram-negative 131 Anaerobic Direct [23]
Escherichia coli Gram-negative 127 Facultative Mixed [18, 44]
Clostridium spp. Gram-positive >100 Anaerobic Mixed [23, 45]
Saccharomyces cereviseae Fungal (yeast) 87 Facultative Not reported
Archae family Archaea 39 Varied Not reported
Enterobacter cloacae Gram-negative 17 Facultative Direct [46]
a
Cited # established via keyword search in Scopus.com in December, 2014

anaerobes i.e. use terminal electron acceptors other than O2—the rationale behind
this is that the reliance on other, less abundant and cell-permeable, electron
acceptors acts as a selective pressure for extracellular electron transfer
reactions [23].
Gram-negative bacteria, compared to the Gram-positive bacteria, maintain a
thinner outer cell membrane, a point which has been implicated in providing greater
extracellular electron transfer efficiencies, both when direct and indirect bacteria-
anode electron transfer is attempted, e.g. [39]. They therefore are more commonly
researched for MFC purposes (e.g. Table 14.3). However, the Clostridium genus is
often included in microbial fuel studies, due to its historical role in biofuel fermen-
tation and have been shown to successfully produce power e.g. [23]. Other studies
have also found no influence on the outer cell wall structure with power generation
e.g. [47], but no concerted attempt to distinguish between direct electron transfer
and mediated electron transfer has yet been made between the two cell wall classes.
When cultured in a consortium, Gram-positive bacteria have been exhibited to
allow electron-transfer reactions when co-cultured with Gram-negative bacteria
that secrete electron transport mediator compounds [42].
Exoelectrogenesis, the processes by which electrons are transported from out-
side their liberating cells, can occur via two pathways: direct and indirect. A
schematic of this appears in Fig. 14.3. Direct cell-electrode electron transfer is
possible and is discussed in greater detail below.
The electrical connection between the cells and electrode by indirect transport
occurs with the deliberate addition, or microbial secretion, of electroactive medi-
ator compounds that diffusively shuttle electrons between the microbial catalyst
and the anode. Some examples of mediators produced by cultures operating under
mediated electron transport include flavins, quinones and phenazines [39, 42,
48]. Depending on the system under examination, the dissolved electron transport
mediators interface with the outer portion of the cell wall, i.e., cytochromes, or can
penetrate through the cell wall to react with the inner metabolic machinery of the
microbe, in order to abstract electrons from the system. A much wider range of
tested microbes can make use of mediator compounds as electron acceptors than
have been observed to produce them [3].
562 R. Fogel and J.L. Limson

Fig. 14.3 Overview of


general processes occurring
at most bioanodes reported
in literature. In the case of
biocathodes, electron
transport originates at the
electrode and the cell, or
produced metabolites, act as
electron acceptor

Overall, the means by which exoelectrogenesis occurs appears to be more


dependant on the system present in the anode, rather than be a key feature of the
microbe selected e.g. [18]. It is quite likely that microbes use a variety of means in
order to transfer electrons to their metabolic electron acceptor, depending on local
environmental conditions, in order to maintain metabolic flexibility.

14.4.1 Nanomaterials Supporting Direct Electron Transport

The transfer of electrons from the bacteria to electrode via direct, or “mediatorless”,
mechanisms is highly desirable for electrochemical fuel cells: the absence of added
mediator improves scaleability of the technology, decreases mass transfer losses,
lowers the operational costs and circumvents the toxicity considerations of includ-
ing certain electron transport mediators e.g. ferricyanide [49].
Since both Gram-positive and Gram-negative bacteria cell walls possess a net
negative charge at physiological pHs [50], both direct and indirect electron transfer
14 Applications of Nanomaterials in Microbial Fuel Cells 563

has been reported to be passively improved by enhancing cell adhesion to the


electrode material through alteration of the electrostatic charge of the electrode
e.g. ammonia treatment [49]. Two separate catalyst classes are, therefore,
distinguishable when considering improving anodic processes—those that improve
direct electron transfer between the immobilised microbe and the electrode and
those that catalyse oxidation reactions of reduced microbial mediators or
metabolites.
Direct electron transport between microbe and solid electrode appears to rely
primarily on the interaction of the electrode materials with cellular c-type cyto-
chrome proteins bound to the outer cell membrane. These heme-containing proteins
are widely distributed among bacteria and archaea [48]. c-Type cytochromes have
been noted to be expressed at high levels under certain growth conditions with
certain microbial species e.g. Shewanella putrefasciens and Geobacter
sulferreducens [23], which make use of these proteins to transfer electrons between
cytochromes and extracellular insoluble metals, e.g., Fe3+ and Mn4+ oxides, as
terminal electron transfers during metabolism. This phenomenon can be further
aided by the microbial extrusion of pili containing a high surface density of these
proteins, resulting in a conductive bacterial “nanowire”, which maintains electro-
chemical activity via the cytochromes (Fig. 14.4 serves as an example) Naturally,
this phenomenon is of great interest to researchers seeking to achieve better direct
electrical connectivity between microbial catalysts and proximal solid electrodes to
lower mass transfer losses [2]; more so when pili have been noted to connect
microbial cells to one-another, both intra- and inter-species [51], allowing for the

0.4
Current density (mA cm-2)

0.3
c
0.2
0.1 b

0.0
a
-0.1
-0.2
-0.3
-0.4
-0.6 -0.4 -0.2 0.0 0.2 0.4 0.6
Potential (V) v.s. Ag/AgCl
Fig. 14.4 Cyclic voltammograms of carbon felt bianodes modified with carbon nanotubes/
MoO2C nanoparticles/ PTFE nanocomposite. Cyclic voltammograms display electrodes following
immersion in growth medium containing E. coli for: (a) 0 h, (b) 48 h and (c) 120 h. Peaks observed
are attributed to the reduction/oxidation of c-type cytochromes, demonstrating successful
colonising of the electrode and direct bacteria-anode electron transfer. Reproduced from [18]
with permission from Elsevier
564 R. Fogel and J.L. Limson

possibility of electrically connecting an entire biofilm to solid electrodes via this


mechanism, to improve power generation [6].
Electrode engineering research aimed at improving electrical connectivity
between the biocatalyst and the anode, as well as improving the surface area of
the electrode system has resulted in several physical and chemical modifications
proposed to enhance microbe-electrode attachment. Polypyrrole- and polyaniline-
based nanomaterials also demonstrate significant improvement of cellular adhe-
sion, by reasons of increased surface roughness, coupled with the presence of
cationic nitrogenous groups to improve cellular adhesion electrostatically [24]. A
more targeted approach, through the covalent attachment of electron transfer
mediators to the surfaces of electrodes, in order to stimulate direct electron transfer
between membrane-bound electroactive cellular proteins and the electrode have
also met with some success [49].
Nanoparticles, through a combination of enhanced electrode surface area, alter-
ation of surface chemistry and the presentation of electroactive moieties to the
microbial cells, have been reported to play a role in the improvement of power
generation at microbial fuel cells [3, 25, 49]. To this effect, Ni, Pd, Au and Fe2O3
nanoparticles have been employed to enhance direct electron transfer [49]. In the
case of carbon-based nanoparticles, the additional improvements in conductivity
between nanoscopic and macroscopic materials is also cited as a contributing factor
[25]. The inherent biocompatibility of carbonaceous nanomaterials has been cited
as a major influence on cell adherence to electrode materials [49]: for this reason,
numerous research articles study the inclusion of carbon nanotubes [49], carbon
blacks [19] and graphite/graphene nanomaterials [49] to improve microbe-
electrode interactions. Metal nanoparticles have been researched for their ability
to improve direct electron transfer, chiefly by mimicking the natural electron
acceptors used by pilli-forming microbes.
The improvements in current response in electrodes modified with nanoparticles
are often attributed to an increase in electrode surface area. Related benefits of
improving surface area using nanomaterials includes increasing the mass-transport
of solutes around the electrode and improving the surface area for cell adhesion at
anodes [52]. Certainly, anodic modifications with nanoparticles (Table 14.1)
(Fig. 14.5) are responsible for significant improvements in power generation at
MFCs, when compared to control MFC configurations with unmodified anode
e.g. [9, 18–20]. In most cases, improved biocompatibility, resulting in higher levels
of microbial colonisation is cited as a major contributing factor behind power
generation increases, as well as improved cell-electron electron transfer efficiency.
Figure 14.6 provides a typical example of improved direct electron transfer effi-
ciencies, and subsequent improvements in power generation from nanomodified
bioanodes in the manner commonly reported in the literature.
In addition to improving electron transfer efficiency, recent research has uncov-
ered the direct influence of nanoparticles to the microbial catalysts’ behaviour. Iron
nanoparticle-decorated graphite anodes were noted to produce a significant increase
in bioanodic compartment power generation [54] and improve direct electron
transfer (Fig. 14.6) [53]. Examination of the gene expression differences between
14 Applications of Nanomaterials in Microbial Fuel Cells 565

Fig. 14.5 Scanning


electron micrograph of
carbon black: FePc
nanocomposite coated
carbon microfiber paper
electrode

a Escherichia coli b 0.4


Escherichia coli
160
Only carbon Only carbon
Fe(10wt%)/CNT
120 Fe(10wt%)/CNT
Fe(30wt%)/CNT
Fe(50wt%)/CNT
Fe(30wt%)/CNT
80 0.3
Fe(80wt%)/CNT Fe(50wt%)/CNT
Fe(80wt%)/CNT
Current(mA)

40
Current(mA)

0 0.2

-40

-80 0.1

-120

-160 0.0
-0.6 -0.4 -0.2 0.0 0.2 0.4 0.6 0 20 40 60 80 100 120
Potential vs. Ag/AgCl (V) Time(hr)

Fig. 14.6 Electrochemical characteristics of carbon paper anodes modified with


bionanocomposite mixture of carbon nanotubes, Fe3O4 nanoparticles and Escherichia coli cells:
(a) cyclic voltammetric analysis of anode of bionanocomposite with varying Fe3O4 content, with
peak formation attributed to microbial-electrode electron transfer. (b) MFC performance using
nanomodified bioanodes, with varying Fe3O4 content. Reproduced from [53], copyright 2014 with
permission from Elsevier

populations of Shewanella oneidasis, grown in the presence or absence of the


nanoparticles, indicated increased expression of biofilm-generating genes, notably
those implicated in organism-electrode transfer, such as those regulating flavin
production and cytochrome c proteins [54] for the reasons discussed above. In a
similar study, a mixed consortium of microbial species grown either on unmodified
carbon paper bioanodes or on ones modified with gold nanoparticles showed
566 R. Fogel and J.L. Limson

significant difference between colonising species present in either biofilms, con-


comitant with an increase in power generation [13]. Both phenomenons were
demonstrated to alter with alteration in surface density of the nanoparticles.

14.5 Nanomaterials Supporting Abiotic Processes

As means of improving power generation at bioanodes, the use of catalysts to


improve mediated electron transfer between cells and solid electrodes has previ-
ously been investigated, but given the uncertainty regarding the mechanism by
which electrons are transferred to the external circuit, power improvements and
electrochemical behaviour cannot be conclusively attributed to an improvement of
direct, or mediated electron transfer reactions. Certainly, the ability of nanoparticles
to act as catalysts and mediators for the oxidation of NADH [55], flavins [56] and
other molecules implicated in microbe-electrode mediated transport e.g. phenolic
compounds [57] are fairly well documented in the literature and have been explored
for these roles in enzymatic fuel cells (e.g. [56] and biosensors [55].
As part of the ongoing integration of this technology towards bioremediation,
several water pollutants have been studied as potential electron sinks. Nitrates,
sulphates and perchlorates are among the pollutants studied, but diooxygen reduc-
tion is still the most commonly-researched electron acceptor in cathodes for MFC
purposes. The addition of microbial and nanoscale catalysts and modifiers have
been consistently studied for their application in improving cathodic reactions and
to increase the overall economic viability of microbial fuel cells. The remainder of
this section is devoted to an examination of the role of nanomaterials in improving
the oxygen reduction reaction, for microbial fuel cell purposes

14.5.1 Overview of the Oxygen Reduction Reaction

Diooxygen (O2) is already well established as an effective electron acceptor. The


overall reaction rate for oxygen reduction to water is well known to proceed via
several different mechanisms—either through 1-electron, 2-electron, or 4-electron
processes, as outlined in Table 14.4.

Table 14.4 Standard Half-reaction Εo0 (V vs. SHE)


reduction potentials of
½O2 þ 2H+ þ 2e ! H2O 1.229
oxygen reduction reactions
O2 þ 2H+ þ 2e ! H2O2 0.703
O2 þ 2H2O þ 4 e ! 4OH 0.633
H2O2þ 2H+ þ 2e ! 2H2O 0.473
Reduction potentials drawn from [58]
14 Applications of Nanomaterials in Microbial Fuel Cells 567

Given the strong oxidising potentials of diooxygen and its associated reducible
hydrides, from þ1.3 to þ0.6 V vs. SHE (Table 14.4), as well as its relative
abundance in the environment, it is unsurprising that a large number of research
publications have been generated with a focus on its successful integration into
conventional MFC technologies. Indeed, the majority of microbial fuel cell
research publications make use of oxygen as the terminal electron acceptor in
their reported systems.
It is tacitly accepted that there are three major obstacles that must be overcome
when considering oxygen as an electron acceptor—the relatively poor solubility of
diooxygen in water (~8.2 mg/L maximum solubility in pure water, at STP using
atmospheric air) imposes a large mass transfer limitation on the system overall; the
large overpotentials of oxygen reduction reactions observed at unmodified carbo-
naceous electrodes [59] and the tendency of carbon electrodes to produce peroxide
intermediates under overpotential conditions. These are undesirable for a further
number of reasons:
The two-electron reduction of oxygen to form hydrogen peroxide has a lower
formal reduction potential than the four-electron process due to the lowered number
of electrons transferred. This decreases the overall emf of the MFC.
The four-electron and two-electron reduction pathways are distinguished by the
O-O bond in diooxygen. In the case of the four-electron reduction pathway, this
bond is broken as the initial reaction of the pathway, forming two adsorping oxygen
atoms on the electrode surface which then undergo two separate 2H+/2e reduc-
tions [59] to form H2O as in Table 14.4. The O-O bond in the diooxygen molecule
remains whole while undergoing a 2H+/2e reduction to produce H2O2 [59]; this
occurs more readily at carbonaceous electrodes than metal electrodes. Due to the
more negative formal potential of the subsequent reduction reaction to generate
water from H2O2 (Table 14.4), this allows hydrogen peroxide to accumulate at
appreciable levels in these systems [59]. Peroxides, and their reaction progenitors,
are corrosive and may damage the electrode and vessel materials.
Hydrogen peroxides undergo spontaneous oxidation to form water and O2
i.e. this side-reaction abstracting protons without electrogenerated electron transfer
reactions and lowers the overall efficiency of MFCs.
Significant effort has been expended in order to prevent the diffusion of oxygen
into the anodic compartment, primarily through engineering research into the MFC
cation exchange membrane [60]. Even if the selection of microbial catalysts is not
limited to the strict anaerobes (who would die if exposed to aerobic conditions),
oxygen, with its high oxidation potential, would compete with the anode as the
electron acceptor.
Regardless of the exact reduction mechanism employed by the system, oxygen
reduction is generally agreed to initiate by a rate-determining adsorption step at
both metallic and non-metallic surfaces [61]. Enhancing the rate at which this
occurs, largely through increasing the surface area of electrodes and the dispersion
of catalyst, is as important as thermodynamic improvements to the system through
further catalytic steps.
568 R. Fogel and J.L. Limson

14.5.2 Nanomaterials Supporting Oxygen Reduction


in Microbial Fuel Cells

The platinum group metals, iridium, osmium, palladium, platinum, rhodium and
ruthenium, are well known as hydrogenation reaction catalysts [62] and it is
unsurprising to note their common inclusion as catalysts for lowering the oxygen
reduction reaction overpotential in the literature. Indeed, these form amongst the
best electrocatalysts for both hydrogen formation from H+ and the reduction of
water [63]. There is good evidence in the literature that oxygen reduction via
platinides proceeds along a four-electron transfer route described in Table 14.4,
validated both for the bulk metal and for Pt nanoparticles dispersed along carbon
electrodes, measured via RDE studies [64]. Due to the typically large separation of
overpotential between platinum-catalysed and carbon electro-reduction of oxygen,
almost complete separation of reactions has been noted to be able to occur,
substantially decreasing the amount of hydroxyl anion radicals and hydrogen
peroxide generated near the cathode surface.
The porphyrin-like prosthetic groups in proteins, e.g., heme, have long been
known to possess both oxygen-binding capabilities and strong redox catalyst
behaviour. It is well known that the synthetic porphyrin macro-class not only
exhibit strong chelating properties towards oxygen, but also strong catalytic activity
towards oxygen reduction processes, as part of a wider diversity of reduction/
oxidation catalysis reactions [65]. Porphyrins and phthalocyanines e.g. Fig. 14.7,
both metal-free [66] and metal-containing [20] are known to have strong catalytic
activity towards oxygen reduction reactions—their investigation as cost-effective

Fig. 14.7 Iron(II)


phthalocyanine, a nitrogen-
carbon macrocyclic
organometallic compound
widely researched as a
low-cost catalyst for the
oxygen reduction reaction
14 Applications of Nanomaterials in Microbial Fuel Cells 569

alternatives to the platinum-group metal catalysts is still a major focus of ongoing


investigation [14, 19].
Depending on the metal centres chelated within these rings, four-electron, or
two-electron oxygen reduction reactions have been noted, with the nitrogen periph-
ery at the metal substitution moiety appearing to be responsible for ORR catalysis
[61, 67]. As nanoscale functional analogues to porphyrins, nitrogen- and nitrogen-
metal-doped carbon nanomaterials demonstrate application as metal-free ORR
catalysts [61, 68]. Several examples of metal-based and carbon-based
nanocomposite cathodes for ORR are presented in Fig. 14.8.
Apart from acting as catalysts themselves, carbon nanoparticles serve a separate,
but equally important, role as low-cost methods of catalyst dispersal at a nanoscale.

Fig. 14.8 Nanocomposite carbon-based oxygen reduction reaction catalysts found in literature. (i)
Fe-SEM micrograph of MnO2/Graphene oxide nanocomposite. Reproduced from [21], copyright
2014, with permission from Elsevier; (ii) cyclic voltammetry demonstrating ORR obtained at (a)
unmodified carbon cloth working electrode Oxygen reduction and (e) the MnO2/ Graphene oxide
nanocomposite. Reproduced from [21], copyright 2014, with permission from Elsevier; (iii)
rotating ring-disk electrode voltammograms for the ORR reaction obtained at glassy carbon
working electrodes modified with either graphene (Graphene), platinised Vulcan XC-72R carbon
(Pt/C) and nitrogen-containing graphene (N-graphene). Reprinted with permission from [68],
copyright 2010, American Chemical Society; (iv) Carbon black/Iron(II)phthalocyanine
nanocomposite, demonstrating the ORR profile obtained at unmodified GCE working electrodes
(asterisk, right) and at the nanocomposite-modified electrode (asterisk, left). Reproduced from
[19], copyright 2014, with permission from Elsevier. For all presented voltammograms, note the
movement of the ORR to more anodic potentials
570 R. Fogel and J.L. Limson

As mentioned earlier, adequate support and dispersion of catalysts is essential to


achieve uniform deposition of catalysts along electrode surfaces in order to max-
imise mass transfer to the electrode [17]. In this area, carbon nanomaterials in
particular, have been employed to great effect. Metal nanoparticles have been
dispersed using carbon nanotubes to produce catalyst-modified conductive
nanomaterial cathodes for improved ORR [10]. Given the high degree of aroma-
ticity along pristine stretches of their walls and potential for π-π electron stacking,
multi-walled carbon nanotubes have been demonstrated to evenly disperse
binuclear phthalocyanines [17, 69], resulting in superior electrocatalytic perfor-
mance of the nanocomposite towards ORR, in turn improving the obtainable power
from the MFC [20].

14.5.3 Other Electron Acceptors

To mimic the heterogeneity of the overall denitrification process making use of


nitrate anions as electron acceptors, a multi-metal heterogenous electrocatalyst,
consisting of gold nanoparticles, heme proteins and cadmium ions deposited onto
glassy carbon electrodes has shown some weak denitrification [70] catalytic activ-
ity, albeit to a lesser extent than previously reported.

14.6 Conclusions

The ultimate goal of research in MFC technology is the realisation of large scale
[8], cost-effective systems that generate appreciable levels of power. Given the
nature of the system towards remediation or the utilisation of waste materials as an
energy source, such large-scale efforts aim for power generated to be sufficient to
power the operations of the treatment facility, at least. That MFC technology will
not match the power density outputs of chemical fuel cell technology without
futuristic approaches identified, owing to the thermodynamic limitations of the
biological systems engaged, is clear. Notwithstanding other significant areas of
progress in recent years, nanomaterials offer substantial opportunities to enhance
baseline power density output through improved electron transfer and mass transfer
processes that limit the power generation of MFC, both considering the biotic and
abiotic processes present in this technology. Their inclusion resulting in increased
surface area and as a dispersal agent or immobilisation matrix for catalysts, as has
been shown. Nanomaterial catalysts used show substantial benefits for lowering the
operating cost of MFC technology, an important curtailment in scale up and
widespread adoption of such technology. Fifty percent of the capital costs of
MFC systems are estimated to arise from catalyst inclusion [24] and the ongoing
research into effective, low-cost catalysts and catalyst dispersants to replace
14 Applications of Nanomaterials in Microbial Fuel Cells 571

expensive platinum-group metal catalysts can only improve the cost-effectiveness


of this technology.
Several research groups have expended efforts on the study of different waste-
waters, and it is here where the biggest leaps in terms of the future practical
adoption and large scale utilisation of MFC technology may yet come from. This
is stated in light of increasing legislation globally on industrial, municipal and
domestic wastewater as well as tenuous access to potable water, and increasingly so
in so-called first world nations. Here beneficiation of wastewater to generate a
usable product, e.g. potable or agricultural grade water is an attractive reality, in
addition to the system generating both direct electricity and achieving remediation
of a wastewater. Such opportunities are easily represented in areas with food or
domestic wastewater rather than industrial wastewater, given the complexity of the
latter. However, industrial wastewater, given legislative requirements may repre-
sent financially viable opportunities given the premium on remediation. The wide
variety of manufacturing industry however means that the realisation of such a
usable system would need to be developed with consideration of chemical and
biological complexity of the waste, and on a waste-by-waste basis. Curtailments in
efficiency of MFC systems are obvious, given differences in environmental param-
eters and chemical composition. In particular the nature of the biomaterials used in
the anode and cathode are particularly vulnerable to fluctuating conditions. In order
to limit such vulnerability, for example of bacteria to toxins in industrial waste
streams, efforts at bacterial encapsulation to isolate and protect these from the
immediate environment are needed, for example, sol-gel encapsulation [71].
Nanomaterials and nanotechnology approaches may yet have further roles to
play in enhancing aspects of remediation and is an area currently open for further
exploration, particularly as this applies to the realisation of future large-scale MFC
systems.

References

1. Logan BE, Rabaey K (2012) Review: conversion of wastes into bioelectricity and chemicals
by using microbial electrochemical technologies. Science 337:686–690. doi:10.1126/science.
1217412
2. Mohan SV, Velvizhi G, Modestra JA, Srikanth S (2014) Microbial fuel cell: critical factors
regulating bio-catalyzed electrochemical process and recent advancements. Renew Sustain
Energy Rev 40:779–797. doi:10.1016/j.rser.2014.07.109
3. Rabaey K, Rodrı́guez J, Blackall LL, Keller J, Gross P, Batstone D, Verstraete W, Nealson KH
(2007) Microbial ecology meets electrochemistry: electricity-driven and driving communities.
ISME J 1:9–18. doi:10.1038/ismej.2007.4
4. Pant D, Van Bogaert G, Diels L, Vanbroekhoven K (2010) A review of the substrates used in
microbial fuel cells (MFCs) for sustainable energy production. Bioresour Technol
101:1533–1541. doi:10.1016/j.biortech.2009.10.017
5. Friman H, Schechter A, Ioffe Y, Nitzan Y, Cahan R (2013) Current production in a microbial
fuel cell using a pure culture of Cupriavidus basilensis growing in acetate or phenol as a carbon
source. J Microbial Biotechnol 6:425–434. doi:10.1111/1751-7915.12026
572 R. Fogel and J.L. Limson

6. Logan BE, Regan JM (2006) Microbial fuel cells—challenges and applications. Environ Sci
Technol 40:5172–5180. doi:10.1021/es0627592
7. Pollet BG (2010) The use of ultrasound for the fabrication of fuel cell materials. Int J Hydrogen
Energy 35:11986–12004. doi:10.1016/j.ijhydene.2010.08.021
8. Logan BE (2010) Scaling up microbial fuel cells and other bioelectrochemical systems. Appl
Microbiol Biotechnol 85:1665–1671. doi:10.1007/s00253-009-2378-9
9. Hou J, Liu Z, Zhang P (2013) A new method for fabrication of graphene/polyaniline
nanocomplex modified microbial fuel cell anodes. J Power Sources 224:139–144. doi:10.
1016/j.jpowsour.2012.09.091
10. Ghasemi M, Daud WRW, Rahimnejad M, Rezayid M, Fatemi A, Jafari Y, Somalu MR, Manzour
A (2013) Copper-phthalocyanine and nickel nanoparticles as novel cathode catalysts in micro-
bial fuel cells. Int J Hydrogen Energy 38:9533–9540. doi:10.1016/j.ijhydene.2013.01.177
11. Haoran Y, Lifang D, Tao L, Yong C (2014) Hydrothermal synthesis of nanostructured
manganese oxide as cathodic catalyst in a microbial fuel cell fed with leachate. Sci World
J. ID: 791672. doi:10.1155/2014/791672
12. Hosseini MG, Ahadzadeh I (2012) A dual-chambered microbial fuel cell with Ti/nano-TiO2/Pd
nano-structure cathode. J Power Sources 220:292–297. doi:10.1016/j.jpowsour.2012.07.096
13. Alatraktchi FA, Zhang Y, Angelidaki I (2014) Nanomodification of the electrodes in microbial
fuel cell: impact of nanoparticle density on electricity production and microbial community.
Appl Energy 116:216–222. doi:10.1016/j.apenergy.2013.11.058
14. Zhao F, Harnischa F, Schr€ oder U, Scholz F, Bogdanoff P, Hermann I (2005) Application of
pyrolysed iron(II) phthalocyanine and CoTMPP based oxygen reduction catalysts as cathode
materials in microbial fuel cells. Electrochem Commun 7:1405–1410. doi:10.1016/j.elecom.
2005.09.032
15. Esmaeili C, Ghasemi M, Heng LY, Hassan SHA, Abda MM, Daud WRW, Ilbeygi H, Ismail
AF (2014) Synthesis and application of polypyrrole/carrageenan nano-biocomposite as a
cathode catalyst in microbial fuel cells. Carbohydr Polym 114:253–259. doi:10.1016/j.
carbpol.2014.07.072
16. Niessen J, Schr€oder U, Scholz F (2004) Exploiting complex carbohydrates for microbial
electricity generation—a bacterial fuel cell operating on starch. Electrochem Commun
6:955–958. doi:10.1016/j.elecom.2004.07.010
17. Deng L, Zhou M, Liu C, Liu L, Liu C, Dong S (2010) Development of high performance of
Co/Fe/N/CNT nanocatalyst for oxygen reduction in microbial fuel cells. Talanta 81:444–448.
doi:10.1016/j.talanta.2009.12.022
18. Wang Y, Li B, Cui D, Xiang X, Li W (2014) Nano-molybdenum carbide/carbon nanotubes
composite as bifunctional anode catalyst for high-performance Escherichia coli-based micro-
bial fuel cell. Biosensors Bioelectron 51:349–355. doi:10.1016/j.bios.2013.07.069
19. Mshoperi E, Fogel R, Limson J (2014) Application of carbon black and iron phthalocyanine
composites in bioelectricity production at a brewery wastewater fed microbial fuel cell.
Electrochim Acta 128:311–317. doi:10.1016/j.electacta.2013.11.016
20. Edwards SL, Fogel R, Mtambanengwe K, Togo C, Laubscher R, Limson JL (2012)
Metallophthalocyanine/carbon nanotube hybrids: extending applications to microbial fuel
cells. J Porphyrins Phthalocyanines 16:917–926. doi:10.1142/S1088424612501027
21. Kumar GG, Awan Z, Nahm KS, Xavier JS (2014) Nanotubular MnO2/graphene oxide com-
posites for the application of open air-breathing cathode microbial fuel cells. Biosensors
Bioelectron 53:528–534. doi:10.1016/j.bios.2013.10.012
22. Mehdinia A, Ziaei E, Jabbari A (2014) Multi-walled carbon nanotube/SnO2 nanocomposite: a
novel anode material for microbial fuel cells. Electrochim Acta 130:512–518. doi:10.1016/j.
electacta.2014.03.011
23. Rinaldi A, Mecheri B, Garavaglia V, Licoccia S, Di Nardo P, Traversa E (2008) Review—
engineering materials and biology to boost performance of microbial fuel cells: a critical
review. Energy Environ Sci 1:417–429. doi:10.1039/b806498a
24. Ghasemi M, Daud WRW, Hassan SHA, Oh S-E, Ismail M, Rahimnejad M, Jahim JM (2013)
Review—nano-structured carbon as electrode material in microbial fuel cells: a comprehen-
sive review. J Alloys Compd 580:245–255. doi:10.1016/j.jallcom.2013.05.094
14 Applications of Nanomaterials in Microbial Fuel Cells 573

25. Gadhamshetty V, Koratkar N (2012) Nano-engineered biocatalyst-electrode structures for next


generation microbial fuel cells. Nano Energy 1:3–5. doi:10.1016/j.nanoen.2011.11.003
26. Baranton S, Coutanceau C, Léger J-M, Roux C, Capron P (2005) Alternative cathodes based
on iron phthalocyanine catalysts for mini- or micro-DMFC working at room temperature.
Electrochim Acta 51:517–525. doi:10.1016/j.electacta.2005.05.010
27. Garret RH, Grisham CM (1999) Electron transport and oxidative phosphorylation. In: Garret
RH, Grisham CM (eds) Biochemistry, 2nd edn. Saunders College Publishing, Philadelphia, pp
673–689
28. Rismani-Yazdi H, Carver SM, Christy AD, Tuovinen OH (2008) Cathodic limitations in micro-
bial fuel cells: an overview. J Power Sources 180:683–694. doi:10.1016/j.jpowsour.2008.02.074
29. Chae KJ, Choi M, Ajayi FF, Park W, Chang IS, Kim IS (2008) Mass transport through a proton
exchange membrane (nafion) in microbial fuel cells. Energy Fuel 22:169–176. doi:10.1021/
ef700308u
30. Zhang X, Cheng S, Wang X, Huang X, Logan BE (2009) Separator characteristics for
increasing performance of microbial fuel cells. Environ Sci Technol 43:8456–8461. doi:10.
1021/es901631p
31. Huggins T, Wang H, Kearns J, Jenkins P, Ren ZJ (2014) Biochar as a sustainable electrode
material for electricity production in microbial fuel cells. Bioresour Technol 157:114–119.
doi:10.1016/j.biortech.2014.01.058
32. Kumar GG, Sarathi VGS, Nahm KS (2013) Recent advances and challenges in the anode
architecture and their modifications for the applications of microbial fuel cells. Biosensors
Bioelectron 43:461–475. doi:10.1016/j.bios.2012.12.048
33. Peng X, Yu H, Lina A, Li N, Wang X (2013) Time behavior and capacitance analysis of nano-
Fe3O4 added microbial fuel cells. Bioresour Technol 114:689–692. doi:10.1016/j.biortech.
2013.07.037
34. Sevilla M, Sanchı́s C, Valdés-Solı́s T, Morallón E, Fuertes AB (2007) Synthesis of graphitic
carbon nanostructures from sawdust and their application as electrocatalyst supports. J Phys
Chem C 111:9749–9756. doi:10.1021/jp072246x
35. Vix-Guterl C, Couzi M, Dentzer J, Trinquecoste M, Delhaes P (2004) Surface characteriza-
tions of carbon multiwall nanotubes: comparison between surface active sites and raman
spectroscopy. J Phys Chem B 108:19361–19367. doi:10.1021/jp047237s
36. Paulus UA, Schmidt TJ, Gasteiger HA, Behm RJ (2001) Oxygen reduction on a high-surface
area Pt/Vulcan carbon catalyst: a thin-film rotating ring-disk electrode study. J Electroanal
Chem 495:134–145. doi:10.1016/S0022-0728(00)00407-1
37. Kelly PT, He Z (2014) Nutrients removal and recovery in bioelectrochemical systems: a
review. Bioresour Technol 153:351–360. doi:10.1016/j.biortech.2013.12.046
38. Xu F (2000) Applications of oxidoreductases: recent progress. Ind Biotechnol 1:38–50. doi:10.
1089/ind.2005.1.38
39. Read ST, Dutta P, Bond PL, Keller J, Rabaey K (2010) Initial development and structure of
biofilms on microbial fuel cell anodes. BMC Microbiol 10:98. doi:10.1186/1471-2180-10-98
40. Wu Y, Liu T, Li X, Li F (2014) Exogenous electron shuttle-mediated extracellular electron
transfer of Shewanella putrefaciens 200: electrochemical parameters and thermodynamics.
Environ Sci Technol 48:9306–9314. doi:10.1021/es5017312
41. Biffinger JC, Pietron J, Bretschger O, Nadeau LJ, Johnson GR, Williams CC, Nealson KH,
Ringeisen BR (2008) The influence of acidity on microbial fuel cells containing Shewanella
oneidensis. Biosensors Bioelectron 24:900–905. doi:10.1016/j.bios.2008.07.034
42. Pham TH, Boon N, De Maeyer K, H€ ofte M, Rabaey K, Verstraete W (2008) Use of Pseudo-
monas species producing phenazine-based metabolites in the anodes of microbial fuel cells to
improve electricity generation. Appl Microbiol Biotechnol 80:985–993. doi:10.1007/s00253-
008-1619-7
43. Jayapriya J, Ramamurthy V (2014) The role of electrode material in capturing power generated
in Pseudomonas catalysed fuel cells. Can J Chem Eng 92:610–614. doi:10.1002/cjce.21895
574 R. Fogel and J.L. Limson

44. Chalenko Y, Shumyantseva V, Ermolaeva S, Archakov A (2012) Electrochemistry of


Escherichia coli JM109: direct electron transfer and antibiotic resistance. Biosensors
Bioelectron 31:219–223. doi:10.1016/j.bios.2011.12.015
45. Park HS, Kim BH, Kim HS, Kim HJ, Kim GT, Kim M, Cahn IS, Park YK, Chang HI (2001) A
novel electrochemically active and Fe(III)-reducing bacterium phylogenetically related to
Clostridium butyricum isolated from a microbial fuel cell. Anaerobe 7:297–306. doi:10.
1006/anae.2001.0399
46. Feng C, Li J, Qin D, Chen L, Zhao F, Chen S, Hongbo H, Yu C-P (2014) Characterization of
exoelectrogenic bacteria enterobacter strains isolated from a microbial fuel cell exposed to
copper shock load. PLoS One 9:e11379. doi:10.1371/journal.pone.0113379
47. Juang DF, Yang PC, Lee CH, Hsueh SC, Kuo TH (2011) Electrogenic capabilities of Gram
negative and Gram positive bacteria in microbial fuel cell combined with biological waste-
water treatment. Int J Environ Sci Technol 8:781–792. doi:10.1007/BF03326261
48. Yang Y, Xu M, Guo J, Sun G (2012) Bacterial extracellular electron transfer in bioelectro-
chemical systems. Process Biochem 47:1707–1714. doi:10.1016/j.procbio.2012.07.032
49. Du J, Catania C, Bazan GC (2014) Modification of abioticbiotic interfaces with small
molecules and nanomaterials for improved bioelectronics. Chem Mater 26:686–697. doi:10.
1021/cm401912j
50. Devi UV, Puri P, Sharma NN, Ananthasubramanian M (2014) Electrokinetics of cells in
dielectrophoretic separation: a biological perspective. BioNanoScience 4:276–287. doi:10.
1007/s12668-014-0140-y
51. Kato S, Watanebe K (2010) Ecological and evolutionary interactions in syntrophic
methanogenic consortia. Microbes Environ 25:145–151. doi:10.1264/jsme2.ME10122
52. Inoue S, Parra EA, Higa A, Jiang Y, Wang P, Buie CR, Coates JD, Lin L (2012) Structural
optimization of contact electrodes in microbial fuel cells for current density enhancements.
Sensor Actuat A Phys 177:30–36. doi:10.1016/j.sna.2011.09.023
53. Park IH, Christy M, Kim P, Nahm KS (2014) Enhanced electrical contact of microbes using
Fe3O4/CNT nanocomposite anode in mediator-less microbial fuel cell. Biosensors Bioelectron
58:75–80. doi:10.1016/j.bios.2014.02.044
54. Xu S, Liu H, Fan Y, Schaller R, Jia J, Chaplen F (2012) Enhanced performance and mechanism
study of microbial electrolysis cells using Fe nanoparticle-decorated anodes. Appl Microbiol
Biotechnol 93:871–880. doi:10.1007/s00253-011-3643-2
55. Gai P-P, Zhao C-E, Wang Y, Abdel-Halim ES, Zhang J-R, Zhu J-J (2014) NADH
dehydrogenase-like behaviour of nitrogen-doped graphene and its application in NAD+-
dependent dehydrogenase biosensing. Biosensors Bioelectron 62:170–176. doi:10.1016/j.
bios.2014.06.043
56. Yehezkeli O, Tel-Vered R, Raichlin S, Willner I (2011) Nano-engineered flavin-dependent
glucose dehydrogenase/gold nanoparticle-modified electrodes for glucose sensing and biofuel
cell applications. ACS Nano 5:2385–2391. doi:10.1021/nn200313t
57. Wang G, He X, Zhou F, Li Z, Fang B, Zhang X, Wang L (2012) Application of gold
nanoparticles/TiO2 modified electrode for the electrooxidative determination of catechol in
tea samples. Food Chem 135:446–451. doi:10.1016/j.foodchem.2012.04.139
58. Harvey D (2000) Appendix 3D: standard reduction potentials. In: Harvey D (ed) Modern
analytical chemistry, 1st edn. McGraw-Hill Higher Education, New York, pp 743–747
59. Bagotsky VS (2006) Reactions at nonconsumable electrodes. In: Bagotsky VS (ed), Funda-
mentals of electrochemistry, 2 edn. Wiley, New York, pp 261–296
60. Leong JX, Daud WRW, Ghasemi M, Liew KB, Ismail M (2013) Ion exchange membranes as
separators in microbial fuel cells for bioenergy conversion: a comprehensive review. Renew
Sustain Energy Rev 28:575–587. doi:10.1016/j.rser.2013.08.052
61. Liew KB, Daud WRW, Ghasemi M, Leong JX, Lim SS, Ismail M (2014) Non-Pt catalyst as
oxygen reduction reaction in microbial fuel cells: a review. Int J Hydrogen Energy
39:4870–4883. doi:10.1016/j.ijhydene.2014.01.062
14 Applications of Nanomaterials in Microbial Fuel Cells 575

62. Patnaik P (2003) Platinum. In: Patnaik P (ed) Handbook of inorganic chemicals. McGraw-Hill
Higher Education, New York, pp 719–722
63. Tarasevich MR, Korchagin OV (2013) Electrocatalysis and pH (a review). Russ J Electrochem
49:600–618. doi:10.1134/S102319351307015X
64. Geniès L, Faure R, Durand R (1998) Electrochemical reduction of oxygen on platinum
nanoparticles in alkaline media. Electrochim Acta 44:1317–1327. doi:10.1016/S0013-4686
(98)00254-0
65. Zagal JH, Griveau S, Silva JF, Nyokong T, Bedioui F (2010) Metallophthalocyanine-based
molecular materials as catalysts for electrochemical reactions. Coord Chem Rev
251:2755–2791. doi:10.1016/j.ccr.2010.05.001
66. Masa J, Zhao A, Xia W, Muhler M, Schuhmann W (2013) Metal-free catalysts for oxygen
reduction in alkaline electrolytes: influence of the presence of Co, Fe, Mn and Ni inclusions.
Electrochim Acta 128:271–278. doi:10.1016/j.electacta.2013.11.026
67. Kobayashi M, Niwa H, Saito M, Harada Y, Masaharu O, Ofuchi H, Terakura K, Ikeda T,
Koshigoe Y, Ozaki J, Miyata S (2012) Indirect contribution of transition metal towards oxygen
reduction reaction activity in iron phthalocyanine-based carbon catalysts for polymer electro-
lyte fuel cells. Electrochim Acta 74:254–259. doi:10.1016/j.electacta.2012.04.075
68. Qu L, Liu Y, Baek J-B, Dai L (2010) Nitrogen-doped graphene as efficient metal-free
electrocatalyst for oxygen reduction in fuel cells. ACS Nano 4:1321–1326. doi:10.1021/
nn901850u
69. Hu X, Xia D, Zhang L, Zhang J (2013) High crystallinity binuclear iron phthalocyanine
catalyst with enhanced performance for oxygen reduction reaction. J Power Sources
231:91–96. doi:10.1016/j.jpowsour.2012.12.018
70. Chen Y, Zhu H, Rasmussen M, Scherson D (2010) Rational design of electrocatalytic
interfaces: the multielectron reduction of nitrate in aqueous electrolytes. J Phys Chem Lett
1:1907–1911. doi:10.1021/jz1005253
71. Ghach W, Etienne M, Billard P, Jorand FPA, Walcarius A (2013) Electrochemically assisted
bacteria encapsulation in thin hybrid sol-gel films. J Mater Chem B 1:1052–1059. doi:10.1039/
C2TB00421F
Index

A 373, 384–385, 410, 423–428, 436–439,


Acetaldehyde, 61, 385, 438–439, 443, 461, 465, 466, 479, 484, 486, 488–492,
457–459, 461, 462, 466, 480, 483, 494 494, 500, 502–506, 518, 520, 533,
Acetylene, 294 551–557, 560–562, 564, 565, 567, 570
Activation polarization, 359 Antibonding, 283, 302, 303, 328–330
Adatoms, 34–36, 50, 55, 71, 192–194, 204, Autoclave, 180, 297
210, 262
Adsorbate, 103, 145, 170, 282, 283, 286, 288,
302, 303, 479 B
Adsorption, 2, 17, 21, 33, 36, 39, 40, 49, 66, 73, Bacteria, 412, 551–553, 555, 557, 562, 570
74, 169, 170, 178, 189, 195, 203, 207, Bending mode, 238
310, 312, 314, 318, 324, 325, 333, 334, Bifunctional effect, 49–51, 101
338, 345–348, 352 Bimetallic nanocrystals, 51, 60, 169–214
Ag seed, 98, 107, 108, 120, 122, 123, 130, 148 Bimetallic tripods, 53
Alcohol, 54, 195–202, 314, 318, 384–385, 422, Binary alloys, 491
440–442, 466, 477–506, 518, 520, 533, Binding energy, 69, 73, 196, 282, 286–291,
536, 539 293, 294, 303, 332, 348, 353, 537
Alkaline electrocatalysis, 32, 56, 73, 74, 171, Binding sites, 369, 392, 393, 526, 528, 530
197, 201, 206, 283, 296, 314, 334, 478, Bio-alcohols, 478, 506
479, 482, 497, 500, 502, 503 Biocompatibility, 552, 553, 563, 564
Alkaline fuel cell (AFC), 293, 409, 410, 484, Biofuel cell, 502, 504
492, 518 Biomass, 195, 200, 436, 477–479, 505
Alloy nanoparticles, 36, 51, 52, 55, 57, 65, Bio-refinery, 478, 506
73–77, 96, 98, 101, 108, 114, 178, 201, Brust-Schiffrin reaction, 125
206, 259, 262, 283, 287, 293 Butler-Volmer equation, 4, 6, 325
Altervalent, 314, 315, 317, 319, 324, 341, 355,
356, 358, 360
Amphiphilic, 379 C
Anatase, 315, 316, 332, 333, 339, 344, 348, Cage-bell structure, 94, 104–116, 122, 148, 149
350, 352–357, 360, 388 Capping agent, 32, 33, 37, 39–42, 44–47, 52,
Anion exchange membrane (AEM), 478, 479, 54, 55, 58, 61, 64, 68–70, 75, 77, 126,
487, 491, 502, 504, 506, 518 172, 176, 189
Anisotropic growth, 120, 136 Carbon black, 13, 15, 21, 64, 72, 171, 173,
Anode, 93, 110, 147, 152, 169, 171, 193, 196, 205, 297, 441, 497, 526, 527, 531,
198, 203, 206, 214, 253, 367, 368, 372, 533, 556, 563

© Springer International Publishing Switzerland 2016 577


K.I. Ozoemena, S. Chen (eds.), Nanomaterials for Fuel Cell Catalysis,
Nanostructure Science and Technology, DOI 10.1007/978-3-319-29930-3
578 Index

Carbon nanotubes (CNTs), 171, 173, D


200, 206, 208, 296, 440, 526, 556, d-band center, 48, 49, 69, 74, 101, 133, 134,
563, 568 147, 170, 207, 243, 283, 286, 293,
Catalyst, 1, 42, 93, 169, 243, 253, 281–303, 295, 303, 490
314, 367, 409, 436, 478, 518, 551 Dealloying, 264–267, 269, 272, 287
Catalyst loading, 255, 411, 464, 527, 558 Decontamination, 33, 37, 39–42
Catalytic activity, 2, 42, 96, 169, 254, 281–303, Defect, 31, 107, 256, 262, 263, 266, 271,
313, 382, 421, 437, 481, 526, 567 297–300, 395, 526
Cathode, 169, 171, 182, 195, 206, 207, 214 Dehydrogenation, 197, 203, 396, 397, 399,
Cavitation, 414–419, 426, 429 438, 439, 482, 483, 491, 503, 504
C-C bond cleavage, 64, 439, 466, 478, Dendritic, 94, 101, 103, 116–128, 130, 171,
481, 489 189–192, 196, 211, 421, 422, 427
Cellular electron acceptors, 557 Density functional theory (DFT), 65, 70,
Ceria (CeO2), 313, 317, 319, 330, 334, 336, 71, 101, 193, 230, 287, 290, 297,
339, 354, 360, 425, 428, 429, 482, 330, 445, 491, 526
487, 506, 533 Desorption, 17, 33, 34, 39, 170, 203, 255,
Cetyltrimethylammonium bromide (CTAB), 284, 294, 310, 312, 318, 320, 321,
37, 42, 47, 59, 64, 117, 173, 175, 423 334, 345, 346, 348, 350, 352, 356,
Characterization, 70, 94, 99, 107, 135, 395–397
137, 149, 154, 175, 178, 189, Diffusion, 2, 3, 6, 8, 9, 12–16, 18, 21, 27, 67,
339–342, 348 68, 174, 203, 207, 323, 340, 342
Charge separation, 385 Diffusion-limited current, 447–449, 455
Chemisorption, 55, 58, 145, 243, 286, 340, 347, Direct formic acid fuel cells (DFAFCs), 44,
369, 397, 438, 458, 479 169, 202
Chromatographic, 480 Direct methanol fuel cell (DMFC), 105, 142,
Citric acid (CA), 182, 186, 188, 441 145, 146, 169, 195, 202
CO adsorption, 17, 21, 37, 228, 233 Disproportionation, 310, 312, 314, 334, 335,
Co-chemical reduction, 171–181, 213 359
Collection efficiency, 22, 207, 452, 453, 455 Dissociative adsorption, 201, 230, 232, 236,
Composition, 51–53, 57, 58, 65, 75–77, 96, 240, 315–317, 320, 322, 324, 338,
103, 104, 114, 147, 171–173, 176, 178, 358, 438, 480
189, 192, 196, 197, 204, 205, 207–209, Dissolution, 107, 117, 119, 140, 170, 171,
214, 359 181, 182, 207, 256, 265, 287, 419,
Concave, 45–48, 52, 56, 58–60, 62–64, 429, 446, 462, 500, 523, 526
66–68, 180, 182, 183, 212, 495 Doping, 388, 393
Conducting polymer, 173, 200, 493, 531–533 Dual path mechanism, 44
Conjugation, 532 Dual-purpose supports, 538–539
Convection, 2, 4, 10, 21 Durability test, 77, 212, 213, 538
Coordinate solvent, 378
Core-sandwich-shell, 53
Corrosion, 170, 171, 207, 284, 368, 419, 490, E
500, 521, 523, 524, 531, 533, 534, Electrocatalysis, 15–16, 31–77, 116, 169–214,
536–538 309–361
CO tolerance, 54, 60, 195, 203, 241, 314, 347, Electrochemical surface area (ESA), 48, 191,
355, 373 207, 257, 284
Crystallographic orientation, 192, 203 Electrochemistry, 1–27, 54, 286, 334, 347,
Crystal Orbital Hamiltonian Population 355, 416, 418–420, 424
(COHP) formalism, 243 Electrode, 2, 32, 145, 172, 226, 258, 285,
Cubic, 32, 34, 35, 37–40, 43, 45–49, 52–55, 309–361, 367, 410, 438, 480, 518, 551
57, 59, 61, 63, 66, 67, 70, 72, 73, 180, Electrodeposition, 200, 419, 427, 428, 492, 497
182, 184–186, 189, 193, 203 Electrolysis, 9, 409, 410, 418, 461, 493, 518
Cubooctahedral, 45, 72, 255 Electron density, 145, 186, 294, 297, 299
Cytochrome, 557, 561, 562, 564 Electronic conductivity, 179, 338, 536
Index 579

Electron transfer, 1–4, 6, 8, 9, 57, 144, 171, Graphene, 13, 57, 60, 64, 171, 174, 179,
191, 201, 206, 295, 314, 338, 419, 180, 191, 200, 201
449, 450, 455, 503, 504, 518, 553, Graphene oxide (GO), 15, 55, 57, 179, 180,
558–565, 567, 570 491, 533
Electrooxidation, 32, 103, 179, 384–385, 437, Graphene quantum dots (GQDs), 294, 296–299
477–506
Electrostatic, 114–116, 443, 562
Elemental mapping, 95, 150, 176, 177, H
180, 181 H2PdCl4, 56, 64, 174
Energy density, 200, 409, 418, 436, 477, 481 H2PtCl4
Epitaxial growth, 143, 145, 146, 172, 184–189 Hammet constant, 294, 295
Etchant, 107 Heterodimer, 129, 130, 132, 134, 150,
E-TEK Pt catalyst, 52, 111, 532 151, 269
Ethanol, 39, 93–154, 200, 225, 384, 435–466, High-angle annular dark-field scanning TEM
477, 517 (HAADF-STEM), 95, 98, 142, 176,
Ethanol oxidation, 61–65, 201 177, 179, 181, 185, 190, 193, 194, 271,
Ethylene glycol, 39, 58, 173, 178, 227, 229, 341–343
421, 440–443, 477, 478, 484–486, 488, High index facets, 45, 47, 50, 52, 59, 62–67,
491, 493, 503, 504, 520, 527 180, 494–498
Evaluation, 1–27, 31, 33, 154, 332 Hydrodynamic method, 21
Exoelectrogenesis, 561, 562 Hydrogenation, 272, 333, 477, 567
Extended Hückel molecular orbital Hydrogen bond, 126, 373, 374, 376, 378
(MO) theory, 243 Hydrogen evolution reaction (HER), 20–21,
Extended X-ray absorption fine structure 225, 325, 326, 329, 330, 334, 358
(EXAFS), 174 Hydrogen peroxide, 12, 22, 25, 206, 444, 445,
466, 566, 567
Hydrogen production and storage, 368, 369,
F 397, 398
Face-centered cubic (fcc), 31, 45, 107, 132, Hydrogen storage, 368, 390–397
143, 170 Hydrophilicity, 375, 523, 526
Fatty acid, 124 Hydrophobicity, 500, 523
FeCo, 487, 491, 495, 502, 504, 527, 529 Hydrothermal, 57, 58, 68, 178–180, 297, 300
FeCu, 487, 493 Hyper-d-metal, 313, 339
Fermi level, 74, 134, 145, 170, 242–244, 283, Hypo-d-oxide, 313, 314, 316–319, 321, 324,
286, 330–331, 536 327, 329, 330, 335, 336, 338–343,
First-principles study, 291 347–350, 353–355, 357–360
Formaldehyde, 57, 68, 178, 179, 311, 318,
319, 321
Formic acid, 44–54, 63, 77, 172–174, 183, I
193, 202–206, 321 Infrared (IR), 458, 483
Frank-vander Merwe mode, 184 In Situ Infrared Spectroscopy, 457–460
Functionality, 371, 378, 394, 397–399 Interbonding, 313, 330–334, 342, 359
Interface, 2, 101, 116, 117, 132, 134,
144, 145, 176, 186, 238, 243, 283,
G 332, 342, 354, 425, 561
Galvanic replacement, 50, 58, 105, 171, 172, Interfacial engineering, 281–303
181–184, 192–194, 210, 212, 295 Interfacial vacancy defects, 96
Geometric and electronic effects, 101 Iron porphyrin, 383
Glycerol, 57, 173, 477, 478, 484–486, 489, Isosteric heat of adsorption, 394
491, 493, 494, 503, 520, 527, 533
Grain size, 422, 428
Gram-negative bacteria, 560, 562 J
Gram-positive bacteria, 560, 562 J-coupling, 230
580 Index

K Microwave heating synthesis, 196


K2PdCl4, 172, 184, 186 Migration, 2, 3, 115, 130, 132, 134, 256, 314,
K2PtCl4, 32, 58, 117, 122, 126, 172, 175, 315, 330, 367, 373, 395
178, 179, 188, 189, 191, 192 Mixed valence network, 315
Kinetics, 4–9, 19, 27, 170, 178, 183, 192, 193, Monolayer, 13, 51, 72, 101, 145, 171, 172, 176,
207, 325, 346 193–195, 209, 210, 213, 230, 269, 293,
Korringa constant, 231 316, 322, 323, 355, 359, 391
Koutecky-Levich equation, 11 Monomer, 238
Multimetallic, 94, 95, 104, 270
Multiply twinned, 57, 58, 96, 98, 99,
L 107, 120, 122, 124
Latent storage, 309–361
Lattice mismatch, 48, 98, 101, 145, 269, 490
LiBH4, 395, 396 N
Ligand-mediated activity, 293–299 Nafion, 152, 233, 323, 346, 372–374,
397, 464, 532, 533, 558
Nanoaggregate, 421
M Nanoclusters, 192, 338, 340, 342, 396, 526
Macroporous, 522 Nanocomplexes, 426
Magneli phases, 315, 324, 335–337, 358 Nanoconfinement, 395–397
Marcus-Hush theory, 5 Nanocrystal, 45, 117, 169–214, 257, 332, 397,
Mass activity, 51, 53, 68, 71, 73–77, 422, 495
100, 192–194, 196, 201, 202, Nanocubes (NCs), 40–42, 46–48, 51–54,
208, 209, 211, 212 57–60, 63, 64, 66–72, 74, 75, 178–186,
Mass transport, 2–4, 6, 8–10, 13, 16, 198, 199, 212
18, 27, 196, 272, 381, 383, 416, Nanoelectrochemistry, 13
418–420, 429, 451, 452, 500, 502, Nanofiber, 174, 175, 199, 200, 526
553, 558, 564 Nanoflowers, 56, 60, 62, 64
Mechanism, 23, 50, 66, 70, 73, 96, 126, 152, Nanomaterials, 1–27, 70, 73, 74, 77, 93–154,
170, 172, 190, 195, 197, 201, 203, 214, 176, 196, 200, 207
315, 323, 330, 347, 354 Nanoparticle, 12–15, 31–77, 94, 171,
Mediator, 502, 553, 561–563, 565 225–244, 281–303, 323, 387, 411,
Mesoporous, 60, 171, 196, 258, 369, 372, 398, 440, 479, 520, 555
520, 522, 536 Nanoporous, 269, 272, 273, 395, 490
Metabolism, 557, 562 Nanorod, 48, 49, 71, 135, 136, 138–140, 174,
Metal-adsorbate bonding, 230, 232 182, 196, 212, 300
Metal carbides, 531, 534–536 Nanowire, 102, 173, 181, 182, 186, 190, 193,
Metal-ligand interfacial bonding, 294 194, 196, 210–212
Metal nitride, 101, 536–538 Ni/NiO, 489
Metal-organic framework (MOF), 367–399 Non platinum catalysts, 381
Metal oxide, 105, 171, 284, 299–303, 317, 388, Nuclear magnetic resonance (NMR),
425, 426, 441, 489, 499–502, 531, 225–244, 505
533–534, 556 Nuclear spin-lattice (T1) and spin-spin (T2)
Methanol, 44, 93–154, 169, 225, 293, 377, relaxation, 230
409, 436, 478, 517, 555
MgH2, 395
Microbial fuel cell (MFC), 410, 551–571 O
Microbial secretion, 561 Octahedron, 31, 54, 117, 185, 198, 261,
Microelectrode, 10, 18 335, 356
Microemulsion, 61, 105, 142, 174, 440 Oleic acid, 40, 60, 75, 95, 124, 140
Microkinetic model, 288 Oleylamine (OAm), 40, 54, 59, 60, 75, 95, 96,
Microorganisms, 551 98, 103, 107, 121, 122, 124–130, 136,
Microporous, 369, 372, 383, 384, 398, 522 140, 176, 190
Index 581

Onset potential, 47, 55, 61, 65, 193, Poly(vinyl pyrrolidone) (PVP), 39–42, 54, 55,
196–198, 200, 202, 208, 211, 212, 57, 58, 60, 61, 63, 178–180, 182, 186,
235, 285, 294, 298, 383, 384, 438, 189, 192, 421, 422, 426, 427
439, 466, 487, 493, 498 Pore volume, 390–392, 399
Organic linkers, 369, 371, 372, 390, 392, 393, Porphyrin, 383, 386, 567, 568
397, 398 Potential cycling, 40, 49, 53, 66, 117, 209, 233,
Organic pollutant, 555 255, 265, 266, 269, 270
OrganoMetallic Fuel Cell (OMFC), 479, 486, Potential of zero charge, 237
503–505 Potentiodynamic, 309, 312, 314, 315,
Overlayer, 51, 145, 490 318–321, 323, 333, 336–338, 340,
Overpotential, 8, 13, 15, 16, 21, 44, 206, 253, 355, 356, 359, 360
322, 323, 424, 426, 427, 439, 446, 447, Power density, 169, 372, 383, 384, 466,
449, 453, 479, 489, 490, 493, 503, 553, 490, 491, 493, 504, 520, 553, 555,
566, 567 557, 558, 570
Oxidation potential, 117, 196, 199, 203, 258, Precursor, 15, 21, 32, 57, 60, 171–174, 176,
264, 270, 385, 438, 567 182, 183, 186, 187, 189, 192
Oxidative etching, 49 Primary oxide, 309–361, 482
Oxide, 15, 34, 99, 171, 228, 284, 309–361, 371, Proton conduction, 372–376, 378–380, 397
423–429, 436, 479, 518, 555 Proton exchange membrane fuel cell
Oxidorectase, 560 (PEMFC), 169, 410, 436, 486, 518
Oxophilic, 54, 93, 114, 233, 260, 330, 438, Pt/C, 21, 23, 45, 46, 51, 52, 55, 57–60,
466, 489 63–65, 68, 70, 72, 74–76, 99, 100,
Oxygen evolution reaction (OER), 310, 314, 117, 124, 144, 179, 182, 191, 198,
318, 321, 324, 330, 335, 355, 358–360 199, 201, 206, 208–213, 228, 253–256,
259, 260, 269–272, 285, 294, 312, 319,
321, 323, 328, 333, 335, 336, 338, 340,
P 348, 355, 356, 358, 447, 448, 450, 451,
Particle size, 16, 34, 53, 57, 70, 72–74, 76, 93, 453, 455, 464, 466, 487, 489, 491, 497,
99, 101, 108, 114, 119, 122, 191, 209, 537, 538
227, 253–257, 259–264, 267, 270, 339, Pt3Co, 60, 260, 265, 286
340, 426, 427, 439, 440, 442, 443, 495, PtCu, 51, 59, 264, 287, 288
498, 527, 536 Pt3Ni, 74, 75, 77, 261, 272, 274, 284–286
Pathway, 15, 22, 44, 66, 133, 184, 198, 201, Pt–O binding, 255, 258
203, 206, 213 Pt-skin, 74, 77, 261, 267–269, 285, 286
Pd-Fe, 293 PtSn, 438–442, 459, 460
Perovskite oxide (ABO3), 300, 303 Pulse voltammetry, 45, 46
Peroxide, 70, 284, 286, 445, 446, 453, 455, 566
Photocatalyst, 368, 385–388
Pluronic, 101, 191, 192 Q
Poisoning intermediate, 44, 195, 487, 499 Quantum yield, 387
Poly(amidoamine) dendrimer (G6-OH),
172, 208
Polyaniline (PANI), 174, 175, 200, 384, R
493, 532 Reaction kinetics, 44, 117, 129, 173, 184,
Polybenzimidazole (PBI), 372, 530 187, 197, 206, 373, 395, 447
Polycrystalline, 45, 66, 198, 272, 310, 311, Reaction polarization, 310, 312–314, 318,
337, 338, 348, 493, 499 323, 334, 335, 359, 360
Polyhedral, 31, 52, 63, 176, 196, 272, 371, 377, Reflux, 112, 440, 527
392, 495 Renewable energy, 169, 368, 502, 517, 518
Polymer electrolyte membrane (PEM), 147, Rotating ring-disk electrode (RRDE), 10–12,
281, 367, 368, 372–380, 397, 398 22, 24, 207, 294, 382, 444, 450–455
Polyol, 57, 64, 101, 173, 181, 491 Ruthenium oxides, 115, 479
Polyvalent, 315 Rutile, 315, 335, 356
582 Index

S Surface cleaning, 37–42, 49, 77, 259, 263,


Sabatier principle, 286 412, 416, 420
Seed-mediated growth, 94, 117–120, 124, Surface disordering, 42–43
130, 134, 135, 172, 184–192, 195, Surface-enhanced Raman scattering
203, 213 spectroscopy (SERS), 225
Segregation, 96, 192, 261, 267, 269, 271, Surface functionalization, 284, 293, 294, 300,
295, 442 526, 527
SEIRAS, 226, 233, 235, 236, 242, 244 Surface geometry, 254
Selectivity, 77, 105, 111, 152, 178, 182, Surface site, 32–45, 49, 71, 203, 234, 240,
272, 446, 457, 478, 479, 489, 492, 254, 258, 481, 482
494, 505, 506, 533 Surface/volume ratio, 174, 178, 297,
Self-assembly, 135, 176, 369 447, 558
Self-poisoning, 203 Surfactant, 21, 54–56, 59, 63, 65, 75, 94, 101,
Semiconductor, 130, 134–140, 142, 144, 112, 140, 192, 259, 263, 421–423,
145, 148 426–429, 440, 441, 443
Shape controlled nanoparticles, 59, 77 Synergy, 52, 96
Single crystal, 17, 31–34, 36, 40, 42, 47, 49,
50, 54, 59, 61, 62, 66, 68, 70, 71, 73,
74, 186, 203 T
Sodium borohydride (NaBH4), 42, 59, 60, T1 value, 230
135, 172, 173, 227, 388, 389, 398, Tafel plot, 11, 100, 299, 319, 326, 335, 336,
440, 500, 501 450, 455
Solid oxide fuel cells (SOFC), 195, 410, Tensile effect, 98, 101, 133
423–429, 518 Tetrahedral-octahedral, 37, 43, 45
Solvothermal, 44, 47, 52, 75, 300, 388, 491, Tetrahexahedral (THH), 45, 50, 52, 53, 55,
495, 538 56, 62–65
Sonochemistry, 414–416 Thermal activation, 383
Sonoelectrochemistry, 418–420, 428 Thermal annealing, 267, 268, 348
Specific activity, 46, 64, 68–72, 74–76, 100, Thermal decomposition, 126, 415, 441
132, 199, 201, 202, 207, 209, 211, Thermogravimetric analyses (TGA), 39,
254–256, 260, 269, 272, 294, 295, 298 55, 300
Spectroelectrochemistry, 249 TiO2, 21, 134, 313, 317, 319, 321, 322, 324,
Spherical, 12, 18, 32, 34, 35, 45, 47, 48, 52, 54, 330, 332–336, 338–342, 344, 345,
60, 61, 64, 67, 71, 72, 122, 124, 142, 347–350, 352, 353, 355, 356, 358, 360
176, 340, 421, 422, 427 Tolerance, 54, 57, 58, 60, 64, 105, 152,
Spillover, 309–361, 395, 482 196–198, 209, 439, 532, 537, 539
Square-wave potential, 63, 64, 496, 497 Transmission electron microscopy (TEM),
Stability, 34, 46, 47, 53–58, 60, 63–65, 69, 95, 185, 525
70, 77, 95, 101, 104, 107, 116, 133, Turnover number, 386
154, 171, 178, 179, 196–198, 200,
201, 204, 206, 207, 209, 211, 314,
324, 325, 328, 336, 338, 359, 360 U
Stellated heterogeneous nanostructure, 120 Ultrahigh vacuum (UHV), 262, 267,
Stoichiometry, 300 348, 349, 352
Stripping, 36, 37, 53, 72, 76, 199, 233, 235, Ultrasonication, 426–429
236, 356 Ultrasound, 411–414, 416, 418–420,
Strong metal-support interactions (SMSI), 422, 426, 428
313, 332, 487, 491 Under potentially deposited (UPD), 34, 36, 42,
Structure-property relationship, 254, 261 47, 48, 50, 70, 72, 76, 193, 255, 269,
Sulfonated polyether-ether ketones (SPEEK), 270, 318, 320, 340, 356
372 UV/ozone (UVO), 42, 43, 69
Sulfonation, 527 UV photo-oxidation (UVPO), 39, 55
Index 583

V W
Volcano curve, 193, 286, 288, 325, WOx, 340, 439
328, 329 Wormlike micelle networks, 173
Volmer-Weber mode, 124, 184
Voltammetry, 2, 6–15, 20, 21, 61, 62,
70, 99, 144, 208, 211, 285, 294, X
295, 323, 355–357, 449, 458, X-ray absorption near edge spectroscopy
493, 498 (XANES), 174, 439
Vulcan XC-72, 99, 128, 173, 335,
440, 441, 464, 482, 503, 525,
532, 533 Z
Zeolitic imidazolate framework (ZIF), 383,
384, 388

You might also like