You are on page 1of 18

Solid State Ionics 180 (2009) 746–763

Contents lists available at ScienceDirect

Solid State Ionics


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / s s i

Ceramic materials as supports for low-temperature fuel cell catalysts


E. Antolini a,b,⁎, E.R. Gonzalez b
a
Scuola di Scienza dei Materiali, Via 25 aprile 22, 16016 Cogoleto, Genova, Italy
b
Instituto de Química de São Carlos, USP, C. P. 780, São Carlos, SP 13560-970, Brazil

a r t i c l e i n f o a b s t r a c t

Article history: The performance and durability of low-temperature fuel cells seriously depend on catalyst support materials.
Received 26 November 2008 Catalysts supported on high surface area carbons are widely used in low temperature fuel cells. However, the
Received in revised form 4 March 2009 corrosion of carbonaceous catalyst-support materials such as carbon black has been recognized as one of the
Accepted 11 March 2009
causes of performance degradation of low-temperature fuel cells, in particular under repeated start-stop
cycles or high-potential conditions. To improve the stability of the carbon support, materials with a higher
Keywords:
Nanocomposites
graphitic character such as carbon nanotubes and carbon nanofibers have been tested in fuel cell conditions.
Transition metal oxides These nanostructured carbons show a several-fold lower intrinsic corrosion rate, however, do not prevent
Electrocatalyst carbon oxidation, but rather simply decrease the rate. Due their high stability in fuel cell environment,
Fuel cells ceramic materials (oxides and carbides) have been investigated as carbon-substitute supports for fuel cell
catalysts. Moreover, the higher specific electrocatalytic activity of some ceramic supported metals than
unsupported and carbon supported ones, suggests the possibility of a synergistic effect by supporting metal
catalyst on ceramic supports. This paper presents an overview of ceramic materials tested as a support for
fuel cell catalysts, with particular attention addressed to the electrochemical activity and stability of the
supported catalysts.
© 2009 Elsevier B.V. All rights reserved.

1. Introduction nanosized electrocatalyst particles in low temperature fuel cells


because of its large surface area, high electrical conductivity, and
Fuel cell technology offers an attractive combination of highly pore structures [4]. However, porous electrically conductive carbon
efficient fuel utilisation and environmentally-friendly operations [1]. blacks do not always exhibit adequate resistance to corrosion caused
Fuel flexibility is an important characteristic of the fuel cells: the ideal by electrochemical oxidation in the fuel cell. The problem of the
fuel for optimum fuel cell operation is hydrogen, although various electrode stability, particularly of cathode stability, is well known.
materials, such as low molecular weight alcohols, that can undergo an Universally, oxygen reduction is a slow reaction, and good catalysts
oxidation reaction, can be used as fuels for a fuel cell system [2]. In have been very difficult to find. As reported by Couper et al. [5], the
order to obtain sufficient electrical output from a fuel cell, the problems of finding a good electrocatalyst (and a stable, inert
electrochemical reaction must occur at a sufficient rate. In low substrate) are compounded by the instability of most materials to
temperature fuel cells (such as polymer electrolyte membrane fuel corrosion close to the reversible potential for the O2/H2O couple.
cells (PEMFCs) and phosphoric acid fuel cells (PAFCs), operating at Corrosion can be accelerated by the presence of trace hydrogen
temperature b200 °C), this is achieved using platinum-based catalysts peroxide, and the lifetimes of potential candidate catalysts can be
[3]. It has to be promptly pointed out, however, that the acid reduced by poisoning and sintering. Substrate corrosion is as difficult
environment in the PEMFCs is different from that of PAFCs. The to avoid as dissolution of the catalyst. Catalyst support corrosion has
PEMFCs operate at less than 100 °C, as compared with the PAFCs, been predominantly observed in phosphoric acid fuel cells. Indeed,
which operate at twice this temperature. Furthermore, the anions of thermodynamically, carbon as a material is unstable at the cathode
the perfluorinated sulfonic acid polymer are only weakly adsorbed on potential in hot phosphoric acid present in PAFCs. Carbon is lost from
Pt, in contrast to the phosphoric anions, which are strongly adsorbed. the system through oxidation leading to significant losses of carbon
The importance of the metal support in fuel cell catalysis is well over a short period of time. The stability of carbon support affects the
recognized. Typically, the support provides a physical surface for loss of platinum surface area following both platinum particle
dispersion of small metal particles, which is necessary for achieving sintering and platinum release from the carbon support [6–8].
high surface area. Carbon is a common choice for supporting Kangasniemi et al. [9] reported that the surface of carbon Vulcan
XC-72 is oxidized under conditions simulating the cathode environ-
⁎ Corresponding author. Scuola di Scienza dei Materiali, Via 25 aprile 22, 16016
ment of a PEMFC (0.8-1.2 V vs. RHE in 1 M H2SO4 at 65 °C). During
Cogoleto, Genova, Italy. single PEMFC durability testing, Borup et al. [10] found that carbon
E-mail address: ermantol@libero.it (E. Antolini). corrosion of the cathode catalyst layer increases with increasing

0167-2738/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.ssi.2009.03.007
E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763 747

potential and decreasing humidity. Wang et al. [11] investigated the However, these alternative materials do not prevent carbon
durability of Pt/C catalysts with different carbon black supports under oxidation, but rather simply decrease the rate. For this reason non-
simulated PEMFC conditions. They found that carbon support with carbon materials have been investigated as catalyst support. Con-
high surface area in PEMFC electrode is susceptible to corrosive ducting oxides are emerging candidates for oxidation resistant
conditions. If the support is oxidized to CO2/CO, Pt may be lost from catalyst supports. In addition to their high stability in fuel cell
the support, so the more the carbon support is oxidized, the more Pt is environment, unlike carbon, which does not help electrocatalytic
lost. If the support is partially oxidized to surface oxide, it may activities, but serves only as a mechanical support, many metal oxide
accelerate the increase of Pt particle size, because the presence of supports can act as co-catalysts. Indeed, it is well known that metal
surface oxides may weaken the platinum–support interaction, leading oxides such as RuO2, SnO2 and WO3 enhance the catalytic activity of
to a lower resistance to surface migration of Pt particles. Accordingly, platinum for methanol and ethanol oxidation [19–24].
corrosion of the cathode catalyst support may affect the durability of The main requirements of a suitable fuel cell catalyst support are
the PEMFC. On the other hand, the anode catalyst could also be the following:
exposed to much more oxidative conditions than the cathode during
• high surface area, to obtain high metal dispersion;
the cell voltage reversal caused by fuel starvation [12]. Summarizing,
• suitable porosity, to boost gas flow;
high-surface-area carbon supports in PAFC and PEMFC electrodes are
• high electrical conductivity;
susceptible to corrosive conditions, which include high water content,
• high stability under fuel cell conditions.
low pH, high temperature, high potential, and high oxygen concen-
tration. In addition, Pt catalysts seem to accelerate the rate of carbon It has to be remarked that, frequently, in the literature the
corrosion [13]. Two different reaction pathways for carbon oxidation difference between catalyst-metal oxide dispersion and metal oxide
have been distinguished, reversible oxidation to oxygenated surface supported catalyst is not adequately taken into account. In the former
carbon species and irreversible oxidation to CO2. If the carbon is partly case the metal oxide only acts as a cocatalyst, not as a support, and the
oxidized to oxygenated surface carbon species, the conductivity at the size of metal and metal oxide particles are not correlated, while in the
contact between the carbon support and the catalyst nanoparticles latter metal particle size depends on, and is considerably lower than
may decrease. If the carbon is oxidized to CO2, catalyst nanoparticles the metal oxide particle size.
may be lost from the support, resulting in a significant decrease of the This paper presents an overview of some metal oxides and
catalyst active surface area, which in turn reduces the performance carbides, which could substitute carbon as fuel cell catalyst support.
and operation lifetime of the PEMFC. Typically, CO2 formation is Particular attention has been addressed to the electrochemical activity
generalized as given by the following equation [14]: and stability of the supported catalysts.

2. Inorganic metal oxides


C + 2H2 O Y CO2 + 4H
+ −
+ 4e : ð1Þ

Inorganic metal oxides have been studied to determine whether


The standard potential for the electrochemical oxidation of carbon they can serve as good corrosion-resistant supports. Unfortunately,
to carbon dioxide is 0.207 V vs. RHE at 25 °C [15]. Therefore, under replacing carbon with traditional metal-oxides is difficult, due to their
typical PEMFC cathode operation conditions, carbon corrosion is not electrical-insulating properties at temperatures below 200 °C. How-
only thermodynamically feasible due to the high potentials (0.6– ever, sub-stoichiometric metal oxides, such as reduced oxidation state
1.2 V) and high O2 concentrations, but also kinetically enhanced by the titania (e.g., Ti4O7 and Ebonex), doped metal oxides, such as doped
elevated temperatures (50–90 °C). TiO2 and SnO2, and nanostructured metal oxides, such as TiO2
Therefore, various alternatives of electrocatalyst supports are nanotubes and WO3 nanorods, have been proposed as electrically
being searched. Due their high surface area and high amount of conductive support materials with high corrosion-resistant properties.
mesopores, which allow high metal dispersion and good reactant flux,
ordered mesoporous carbons and carbon gels have been receiving 2.1. Ti-based oxides
attention as fuel cell catalyst supports. Catalysts supported on these
carbons showed higher catalytic activity than the same catalysts Among the inorganic metal oxides, titanium dioxide, TiO2, is used
supported on carbon black. Their thermal stability, however, is almost in many applications that depend on its photoelectrochemical and
the same to that of carbon blacks [4]. In view of their high thermal catalytic properties and its excellent resistance to corrosion in various
stability, boron-doped diamonds (BDDs) have been tested as catalyst electrolytic media [25,26]. Titania exists in three main crystallographic
supports. The results indicated that the anchoring of metal atoms on forms e.g., anatase, rutile, and brookite. Each structure exhibits
BDD surface in a stable way has to be improved. The use of Nafion as different physical properties, which lead to their different applica-
the binder seems to enhance the stability of BDD supported catalysts tions. It is generally accepted that anatase titania is more efficient as
[4]. Recently, nanostructured carbon materials with graphitic struc- photocatalyst than rutile titania. Regarding the application in fuel
ture, such as carbon nanotubes (CNTs) and carbon nanofibers (CNFs) cells, Gustavsson et al [27] investigated thin film of Pt and
were investigated as catalyst supports in fuel cells [4]. The higher stoichiometric TiO2 in a polymer electrolyte electrochemical cell.
catalytic activity of Pt and Pt–M catalysts supported on CNTs and CNFs Individual thin films of Pt and TiO2 were deposited directly on Nafion
than that of the same catalysts supported on carbon blacks was membranes by thermal evaporation. Oxygen reduction reaction
ascribed to their unique structure and properties such as high surface polarization plots showed that the presence of a thin TiO2 layer
area, good electronic conductivity and chemical stability [4]. Tests between the platinum and the Nafion increases the performance
carried out in PEM fuel cell conditions indicated that these materials compared to a Pt film deposited directly on Nafion. They attributed
can be more durable and can outlast the lifetime of conventional this improvement to a better dispersion of Pt on TiO2 compared to on
Vulcan XC-72 [16–18]. By comparing carbon nanotubes and mesopor- Nafion and in addition, substantial proton conduction through the
ous carbons, taking into account of the cost of the materials, the thin TiO2 layer.
complexity of the synthesis methods, and the versatility in pore size Stoichiometric rutile is an insulator with a band gap of 3.06 eV at
and pore distribution tailoring, the mesoporous carbons seem to have room temperature. Bulk transport measurements show that for very
more changes to substitute carbon blacks as fuel cell catalyst slightly nonstoichiometric TiO2, with a carrier concentration in the
substrates. The stability in fuel cell conditions of mesoporous carbons range below 2.0 1020 cm − 3, the low-temperature activation energy
is lower than that of CNTs, but can be increased by graphitization [4]. for conduction is about 0.028 eV [26]. n-type dopants for TiO2, as
748 E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763

vanadium, niobium, and tantalum, can be substitutionally incorpo- the Ebonex® supported PtCo has higher catalytic activity compared to
rated within the rutile lattice to occupy at least 50% of the cation sites. unsupported PtCo samples. They assumed that the observed improved
Nb doping gives rise to shallow donor states 0.02–0.03 eV below the catalytic activity of PtCo/Ebonex® is due to formation of surface
conduction-band minimum and therefore at essentially the same oxides and electronic interactions between the metallic components
energy as the donor states introduced by oxygen deficiency [28]. On of the catalyst and the supportive material, in which Co plays a
the basis of their high chemical stability, various electrically predominant role. It has to be remarked, however, that no comparison
conductive forms of the titanium–oxygen system have been tested with carbon supported PtCo was made!
as a fuel cell catalyst support. Their preparation method and physical Ioroi et al. [41] prepared Ti4O7 supported Pt catalysts (2.5 and 5 wt.%)
characteristics as well as the catalytic activity and stability in fuel cell and compared their activity for both the hydrogen oxidation reaction and
conditions of related supported catalysts are reported in the following the oxygen reduction reaction to that of Vulcan XC-72 supported Pt
sections. catalyst under PEMFC operation at 80 °C. Single-phase sub-stoichiometric
titanium oxide powder was prepared by the reduction of TiO2 powder at
2.1.1. The homologous series TinO2n−1, 4 ≤ n ≤ 10 high temperature (6 h at 1050 °C under flowing H2). Platinum was loaded
The titanium–oxygen system contains a sub-stoichiometric com- on the Ti4O7 by a conventional impregnation method, by reducing the Pt
position of the general formula TinO2n−1 (Magneli phases), where n is precursor at 250 °C in a flowing mixed H2/N2 atmosphere. Some of the
a number between 4 and 10 [29]. The phases consist of two- deposited Pt particles were aggregated, especially in 5 wt.% Pt/Ti4O7,
dimensional slabs of the TiO2 rutile structure with a thickness of n- because the surface area of the Ti4O7 prepared in this study was rather
layers of Ti atoms separated by oxygen deficient layers [29,30]. Among small (0.95 m2 g − 1). Tests on Pt-free Ti4O7 and Vulcan XC-72 supports
this series of distinct oxides, Ti4O7 exhibits the highest electrical showed that the onset potential of the corrosion current during the
conductivity exceeding 103 S cm − 1 at room temperature [31]. positive direction sweep of cyclic voltammetry (CV) was much higher for
Moreover, these oxides show no peaks or waves between 0.0 and Ti4O7 than for Vulcan XC-72, which indicates that Ti4O7 would be quite
2.0 V vs. RHE in 1 M sulfuric acid at 25 °C [31], indicating highly oxidation resistant under the actual PEMFC operating conditions. PEMFC
oxidation-resistant characteristics. According to these conductive and polarization curves using 5 wt.% Pt/Ti4O7 and conventional 20 wt.% Pt/C
oxidation-resistant properties, TinO2n−1 phases have attracted atten- cathodes indicated that the mass activity of Pt/Ti4O7 is lower than that of
tion as a possible support for fuel cell catalysts [32]. Commercially Pt/C. Compared to the Pt/C cathode, ca. a 100 mV voltage loss was
available products known as Ebonex® are supplied by Atraverda Ltd. observed at 500 mA cm − 2 for the 5 wt.% Pt/Ti4O7 cathode. This was
(Sheffield, UK). Ebonex® is an electrically conductive ceramic ascribed to the smaller electrochemically active area of Pt/Ti4O7 relative
consisting of several suboxides of titanium dioxide, mainly, Ti4O7 to Pt/C, because the Pt particle size deposited on Ti4O7 is larger than that
and Ti5O9. Ebonex® may have a surface morphology (microrough and of the Pt/C catalyst. However, the specific activities of both catalysts,
porous) which give it particularly good properties as a substrate for based on the electrochemically active surface area, were very close to
metal layers [33–35]. The structure and properties of Ebonex® each other, indicating that the activity for the oxygen reduction reaction
ceramic electrodes, together with their electrochemical behaviour of Pt catalyst particles on both Ti4O7 and Vulcan XC-72 is near the same.
and applications were excellently reviewed by Smith et al. [36]. Accordingly, the Ti4O7 support would be applicable to PEMFC electro-
Farndon et al. [37] demonstrated that metal deposition on Ebonex® catalysts without degrading the intrinsic catalytic activity of Pt, so that Pt/
substrate occurred by a mechanism parallel to that observed at metal Ti4O7 would be a potentially durable catalyst material for PEMFCs. The
or carbon substrates: nucleation of hemispherical Pt centres was same authors [42] investigated the stability of 5 wt.% Pt/Ti4O7 and 30 wt.
followed by three dimensional growth under electron transfer control. % Pt/C cathodes in PEMFCs by polarization measurements to evaluate fuel
As an electrocatalyst support, Ebonex® has the potential of interacting cell performance loss and CV measurement to evaluate electrochemically
with the deposited metal so as to alter the activity of the catalyst [38]. active area loss after potential holdings (N1.0 V). The Pt/C catalyst was
Indeed, its chemical composition and hypo-d-electron character used after heat treatment at 1000 °C under H2 to obtain Pt dispersion
suggests an ability to interact with hyper-d-electron metals such as similar to that of 5 wt.% Pt/Ti4O7, and was denoted as 30% Pt/C-HTT.
Pt, Ni and Co. Farndon and Pletcher [39] investigated the stability of Fig. 1a [42] shows the obtained Ti4O7: although the starting material
platinized Ebonex® electrodes prepared by Pt electrodeposition. Pt/ (TiO2) is a fine powder (70 nm diameter, BET area: 20 m2 g− 1), sintering
Ebonex® electrodes with Pt loadings of 2 mg cm − 2 were tested in a occurs during a high-temperature reaction, so that the surface area of the
flow electrolysis cell (1 M NaOH) for N500 h without significant prepared Ti4O7 powder is considerably low (BET area: 2 m2 g− 1).
changes in properties. Dieckmann and Langer [38] studied the electro- Subsequently, Pt particles were deposited on Ti4O7 by the impregnation-
oxidation of simple aliphatic and aromatic alcohols and formaldeyde reduction method. As can be seen in the scanning electron microscopy
on Ebonex® and graphite supported platinum in acid and basic (SEM) image (Fig. 1b), Pt particles are uniformly dispersed on Ti4O7. The
electrolytes. The Pt/Ebonex® catalyst was prepared by suspending diameter of the Pt particles was about 10–20 nm, which is much larger
Ebonex® pellets in a H2PtCl6 solution (Pt loading 2 mg cm − 2), and than that of the actually used Pt/C catalyst, because the small surface area
then reducing Pt using the electroless ethanol reduction method. Pt of the Ti4O7 prepared in this study. Fig.1c shows the SEM image of 30% Pt/
surface area was more than twice higher on graphite than on C-HTT. Pt particles with a diameter of 10–20 nm are dispersed on Vulcan
Ebonex®. The catalytic activity of Pt/Ebonex® was lower than that XC-72 support, which is similar to the Pt particle size of Pt/Ti4O7. To
of Pt/C, even when surface areas were considered. Slavcheva et al. [40] compare the catalyst stability, Pt particle size is important because the
prepared Ebonex® supported Pt and PtCo catalysts for the oxygen stability of Pt particles is strongly affected by the particle size. Smaller
evolution reaction in alkaline media by impregnation of the metal particles are more unstable, especially under high potentials or potential
precursors on the surface of the support, followed by reduction with sweep conditions. The high-potential holding test showed significantly
sodium borohydride. The specific surface area of the Ebonex® is very greater stability of Pt/Ti4O7 than the conventional Pt/C-HTT catalyst. The
low (2–3 m2 g − 1), and, as a consequence, the unsupported Pt and effects of potential holding on the CV and cell performance for the Pt/
PtCo nanoparticles prepared by NaBH4 reduction possess much higher Ti4O7 cathode were much smaller than for the Pt/C-HTT cathode.
specific surface area (SSA, 38 and 40 m2 g − 1, respectively), than that
of Ebonex® supported catalysts (from 5 to 25 m2 g − 1, depending on 2.1.2. Metal-doped TiO2 (M–TiO2, M = Ru, Nb)
the metal/support ratio). The PtCo/Ebonex® catalyst, selected for the It has been shown that mixtures of NbO2 and TiO2 sintered at
electrochemical testing, contained 50 wt.% Ebonex® and presented a 1000 °C lead to the formation of an electrically conducting material
specific surface areas of 10.37 m2 g − 1. The results of galvanostatic [28,43]. This high temperature synthesis method shows promising
polarisation experiments showed that, notwithstanding its lower SSA, results, but it leads to a low surface area material that requires long
E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763 749

presented higher current per mass of Pt compared to the E-TEK


PtRu/C at 0.40 V. According to the authors, this result suggests that the
anatase titania support enhances the catalytic activity of PtRu in ways
that carbon cannot.
Park and Seol [47] prepared a Nb–TiO2 supported Pt cathode
catalyst for polymer electrolyte membrane fuel cells. Nb-doped TiO2
nanoparticles were obtained by hydrothermal synthesis at 120 °C
using Ti(IV) isopropoxide and Nb(V) ethoxide as metal precursors. To
activate Nb donor dopant in TiO2, the nanoparticles were annealed at
400 °C for 2 h under pure H2. Nb–TiO2 presented a TiO2 rutile-
dominant crystal structure, and its electrical conductivity was about
0.1 S cm − 1. The Nb–TiO2 supported Pt (40 wt.%) catalyst, prepared by
the borohydride reduction method, exhibited a good dispersion of Pt
catalysts (~3 nm) on the Nb-TiO2 nanostructure supports (~10 nm).
The Pt/Nb–TiO2 showed a higher activity for oxygen reduction than
that of a carbon supported Pt prepared by the same route. According
to the authors, the enhanced catalytic activity of Pt/Nb–TiO2 in
electrochemical half cell measurement may be mainly due to the good
dispersion of Pt nanoparticles on Nb–TiO2 nanosized supports. In
addition, from XANES spectra of Pt L edge obtained with the
supported catalysts, the improved catalytic activity of Pt/Nb–TiO2
for oxygen reduction may be caused by an interaction between oxide
support and metal catalyst.
Recently, Ru-doped TiO2 was prepared by a sol–gel method, then Pt
atoms were supported on this material by an impregnation reduction
method, and finally the supported catalysts was electrochemically
tested in fuel cell conditions by Haas et al. [48]. The composition of
RuxTi1−xO2 was in the range from xRu = 0.17 to xRu = 0.73. XRD
analysis carried out on the mixed oxide indicated that the system
exhibits a change from solid solution above 0.5 mole fraction of
ruthenium to a mix of two phases below 0.5 mol fraction of
ruthenium. Ru–TiO2 particle size from SEM images was estimated to
vary between 200 and 300 nm. The conduction of these oxides can be
compared to semi conductors and varied in the range of 0.1 b σ b 100 S
cm − 1. Regarding the Ru–TiO2 supported catalyst the Pt loading varied
from 23 to 40 wt.%. The Pt crystallite size was in the range 3.3–5.4 nm.
The electrochemical characterization of the catalysts was done by
cyclic voltammetry in half cell configurations in 0.5 M phosphoric acid
electrolyte and showed that the catalysts have an active platinum
surface area comparable to commercial carbon supported catalysts. In
particular, the active surface area of the Pt supported on RuTiO with
xRu = 0.71 was 1) nearly twice as large as the other oxide compositions
and 2) larger than pure Pt on Vulcan at the same Pt loading.
Chen et al. [46] compared the activity and stability as supports for
Fig. 1. SEM images of (a) prepared Ti4O7 support, (b) 5% Pt/Ti4O7, and (c) 30% Pt/XC-72-HTT. unitized regenerative fuel cell electrocatalysts of Ebonex®, Ti4O7 and
Reprinted from Ref. 42, copyright 2008, with permission from The Electrochemical Society. Nb0.1Ti0.9O2, made, respectively, by reducing and doping rutile
titanium dioxide. The synthesis of both Ti4O7 and Nb0.1Ti0.9O2 started
from the same precursor, ultrafine rutile TiO2, having a BET surface
synthesis times. Because TiO2 undergoes a phase transition from area of 110 m2 g − 1. Ti4O7 was obtained by hydrogen reduction of TiO2
anatase to the less catalytically active rutile near 700 °C [44], the high- at 1050 °C, while Nb0.1Ti0.9O2 was prepared by heating a mixture of
temperature synthesis may also reduce some catalytic promotion of TiO2 and NbO2. Both ceramic samples had an electrical conductivity
methanol oxidation by the support. On this basis, to increase the similar to Ebonex®. The XRD pattern for Nb0.1Ti0.9O2 showed a pure
surface area and avoid the phase transition, a low-temperature rutile microcrystalline phase. BET surface area measurements gave
synthesis route for Nb0.1Ti0.9O2 via a surfactant templating method relatively low surface areas for all three supports, 2 and 1.4 m2 g − 1 for
was proposed by Garcia et al. [45]. The synthesis method is based on the synthesized Ti4O7 and Nb0.1Ti0.9O2, respectively, and 1 m2 g − 1 for
using octadecylamine as a template, and niobium(V)ethoxide and Ebonex®. Supported Pt4Ru4Ir1 catalysts were prepared forms by
titanium(IV)butoxide as metal precursors. BET surface area of the reducing the corresponding metal salts onto Ebonex® and synthe-
resulting Nb0.1Ti0.9O2 was 136 m2 g − 1. This surface area is much sized Ti4O7 and Nb0.1Ti0.9O2, using the borohydride method. These
higher than the value of 1.4 m2 g − 1 that was measured for the same supported catalysts were then tested in the gas diffusion half-cell for
compound synthesized at high temperature using the ceramic both oxygen reduction and evolution. The Ti4O7-supported catalyst
method [46]. An Nb0.1Ti0.9O2 supported PtRu catalyst was prepared showed catalytic activity very similar to that of the Ebonex®
by a colloidal method. Nb0.1Ti0.9O2 was added to an H2PtCl6/RuCl3 supported catalyst for both reactions, while the catalyst supported
mixture in tetrahydrofuran. Then, the metal precursors were reduced on the Nb0.1Ti0.9O2 gave higher currents at all applied voltages.
with LiBH4. Tests in direct methanol fuel cells (DMFCs) were Because the difference increased as the catalysts were progressively
performed at 70 °C for the PtRu/Nb0.1Ti0.9O 2 catalyst and a oxidized, the authors attributed the lower current at a given bias to a
commercial PtRu/C catalyst from E-TEK. The PtRu/Nb0.1Ti0.9O2 greater internal IR drop in the Ebonex®- and Ti4O7-supported
750 E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763

catalysts. To further investigate their electrochemical stability, they nanophase particles along the inner and outer wall of TONT. Average
examined Ebonex®- and Nb0.1Ti0.9O2-supported catalysts during a crystalline size of Pt and Co mixed nanophases was estimated by the
relatively long period of anodic polarization in the gas diffusion half magnified image of Fig. 2b, indicative of a well dispersed distribution.
cell, with continuous 0.5 M H2SO4 electrolyte flow. The Nb0.1Ti0.9O2- The crystalline planes proving the presence of crystalline Pt and Co,
supported catalyst showed no significant difference in current after together with anatase TONT support, were clearly observed in Fig. 2c,
the polarization period. Conversely, there was a distinct current drop confirming the crystallinity of the binary nanophase layer. SAED
for the Ebonex®-supported catalyst, especially for the oxygen patterns in Fig. 2d also demonstrated that the annealed TiO2
reduction reaction. The likely cause of this effect is partial oxidation nanotubes were a polycrystalline anatase phase, designated “A.” The
of Ebonex®, which creates a resistive layer through which current specific crystalline planes associated with Pt and Co nanophase
must pass. The loss in current showed that, although they are stable structure were also found. As-deposited PtNi/TONT showed poor ORR
for short periods of time, Ebonex® and Ti4O7 are still oxidized to activity, in particular at high potential, related to OH adsorption.
nonconductive TiO2 at the catalyst/support/electrolyte three-phase Annealed PtNi/TONT at 400 °C under hydrogen atmosphere displayed
interface after extensive polarization at the positive potentials of the a significant enhancement of the ORR, also in the presence of
oxygen electrode. On the other hand, Nb0.1Ti0.9O2 showed good methanol. To clarify the effects of the nanotubular TiO2 structure
stability under these conditions. and pore size, compact TiO2 films on Ti substrate were prepared and
compared to TiO2 nanotube arrays. A remarkable increase of ORR
2.1.3. Nanostructured TiO2 activity and a positive shift of onset potential were achieved using
Nanostructured TiO2 materials, with a typical dimension less than Pt70Co30/TONT. This phenomenon was attributable to the favorable
100 nm, have recently emerged. Such materials include spheroidal accessibility of the inner wall of TONT for diffusing oxygen and to the
nanocrystallite and nanoparticles together with (more recently synthe- high surface area of PtCo catalyst layers at the upper lip and inner wall
sized) elongated nanotubes, nanosheets, and nanofibers [49]. In view of of TONT [53]. However, it has to be promptly pointed out that in all
to their high stability, high surface area and moderate electrical these works a comparison of the catalytic activity of TONT supported
conductivity, titanium dioxide nanotubes (TONTs) have been investi- catalysts with that of a carbon supported catalysts was not made!
gated as fuel cell catalyst support [50–53]. The nanotube structure has a
variety of advantages as a support of electrocatalysts associated with the 2.2. Sn-based oxides
specific geometry. The geometry of the arrays affects the dispersion of
catalysts, the effective catalyst loading, and easy permeation of reactant 2.2.1. M-doped SnO2 (M = Sb, Ru)
species resulting from the highly porous structure, leading to an Tin dioxide, SnO2, belongs to the family of transition-metal dioxide
enhancement of catalytic activity [49]. In relation to methanol oxidation compounds with rutile structures. The mineral form of SnO2 is called
and oxygen reduction processes in DMFCs, Pd and PtRu, and PtNi and cassiterite, and this is the main ore of tin. SnO2 is usually regarded as
PtCo supported on TONTs were investigated for the methanol oxidation an oxygen-deficient n-type semiconductor. Hydrous forms of SnO2
reaction (MOR) [50,51] and the oxygen reduction reaction (ORR) have been described in the past as stannic acids, although such
[52,53], respectively. Wang et al. [50] investigated the electro-oxidation materials appear to be hydrated particles of SnO2 Although SnO2 is
of methanol in sulfuric acid solution using palladium well-dispersed on insoluble in water, it is an amphoteric oxide, although cassiterite ore
titanium nanotubes. TONTs were prepared by the alkaline hydrothermal has been described as difficult to dissolve in acids and alkalis.
method [54]. Pd/TONT catalysts were synthesized at 120 °C using the SnO2 supported Pt and Pd metals are interesting catalyst systems for
conventional reduction method with glycol. Pd dispersed on titania various chemical reactions, such as low temperature oxidation of CO and
nanotubes, which leads to high surface area substrates, showed higher methane [55–57], selective reduction of NO by hydrocarbons [58], and
catalytic activity for the MOR than that of pure Pd and Pd supported on crotonaldeyde hydrogenation [59], because of the synergistic effect. In
TiO2 nanoparticles. Macak et al. [51] tested nanotubular TiO2 matrices the same way, SnO2 has been proposed as a support material for fuel cell
(generated from mixed H2SO4/HF electrolytes using the self-organizing electrocatalysts because of its chemical properties: it adsorbs OH species
anodization process) as active support for Pt/Ru nanoparticles for at low potentials and/or induces the electronic effect with Pt catalysts.
methanol electrooxidation. The nanotubular TiO2 layers consisted of These properties promote the electrooxidation on Pt of CO [60] and low-
individual tubes of 100 nm diameter, 500 nm length and 15 nm wall molecular-weight alcohols, such as methanol [61] and ethanol [22]. For
thickness. This nanotubular TiO2 support provided a high surface area its use as a catalyst support in fuel cells, however, the electrical
and it significantly enhanced the electrocatalytic activity of Pt/Ru for conductance of SnO2 has to be improved. Indeed, undoped tin dioxide is
methanol oxidation (relative to the performance of Pt/Ru at the same a wide bandgap semiconductor (Eg ~ 3.6 eV) with electrical resistivity
loading but immobilized on a conventional compact TiO2 support). varying from 10 to 106 Ω cm, depending on the temperature and the
Annealed to anatase, the TiO2 nanotubular support exhibited even a stoichiometry of the oxide [62]. The electrical resistivity was found to
higher enhancement effect during electrooxidation of methanol than drastically decrease to 10 − 2–10 – 3 Ω cm by doping SnO2 with Sb5+ [63].
when used in the “as-formed” amorphous structure. A key effect of Santos et al. [64] electrodeposited platinum micro particles on
annealing seems to be a strong adhesion of the bimetallic Pt/Ru antimony doped SnO2 (ATO) thin films, and tested their activity for
nanoparticles to the nanotubular TiO2 matrix. Moreover, they showed methanol electro-oxidation. The tin oxide films were prepared through
that the electrocatalytic activity is further enhanced by UV-light the thermal decomposition of polymeric precursors. Antimony was
illumination. introduced in the precursor solution by directly adding Sb2O3 to the tin
Kang et al. [52,53] formed elongated TiO2 nanotube arrays on Ti solution. The Sb content was 2 mol%. The platinum particles were
substrate using an electrochemical anodization process. TONT arrays electrodeposited at a constant 0.2 V versus reversible hydrogen
with a length of approximately 1 μm and a diameter of 120 nm were electrode (RHE) potential. The crystallite size of Sb–SnO2 supported Pt
produced easily and reproducibly. A dual-gun sputtering technique was in the range 8.5–12 nm. The electroactivity of Sb–SnO2 supported Pt
with Pt, Ni and Co targets was used for deposition of nanoparticle PtNi was investigated by chronoamperometry (CA) and compared to the
(PtNi size 5–10 nm) [52] and PtCo (PtCo size 3–4 nm) [53] catalysts on activity of platinized platinum. Pt/ATO electrodes not only presented
the TONT. The crystalline and morphological properties of Pt70Co30 higher roughness factors than platinized Pt electrodes, but also had
catalyst layer on the TONT were investigated by HRTEM, as shown in larger intrinsic electrocatalytical activity for the electrooxidation of
Fig. 2 [53]. Fig. 2a shows the representative top surface morphology of methanol. According to the authors, these results suggest that tin oxide
Pt70Co30/TONT, revealing the nanotubular structure as a support for can be used as a good matrix for Pt (or Pt alloys) catalyst dispersion for
the deposition of the binary catalyst and the existence of metal applications in DMFC. The method applied for the preparation of Sb–
E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763 751

Fig. 2. HRTEM images of (a) Pt70Co30/TONT of the top surface nanotubular structure, (b) the distribution of binary alloy nanophases along the inner wall, (c) crystallographic analysis
of the nanophase alloy and anatase TiO2, and (d) SAED of Pt70Co30/TONT, indicating the polycrystalline nature of the individual components. Reprinted from Ref. 53, copyright 2008,
with permission from The Electrochemical Society.

SnO2 films can be useful to prepare small oxide particles as a conductive can be attributed to the effects of SnO2 adjacent to Pt. As the Pt/Sn
support for fuel cell catalysts. On this basis, Lee et al. [65] deposited Pt system is more effective for the EOR than for the MOR [66], the EOR
colloidal particles on ATO nanoparticles with various amounts of Pt activity of the Pt/ATO was enhanced over that of Pt/C to a much greater
loading (5, 10, 20, and 40 wt.%), and investigated their electrocatalytic degree than the MOR activity. The Pt/ATO sample exhibited higher
activities and stabilities for methanol (MOR) and ethanol (EOR) current density than the 5 wt.% Pt/SnO2 sample, again confirming the
oxidation reactions. The activities for the MOR and EOR were compared Sb-doping effect. Interestingly, the 5 wt.% Pt/SnO2 exhibited a higher
with those of Pt supported on carbon, prepared through the same current density than Pt/C, despite a much lower electrical conductivity
procedure using Vulcan XC-72. The ATO nanoparticles were synthesized of SnO2 than that of carbon, clearly demonstrating the effect of SnO2 as a
by the reflux method using SnCl4·5H2O and SbCl3 as precursors. promoting support. The Pt/ATO exhibited much higher stability than Pt/
Preparation of Pt/ATO and Pt/C nanoparticles was based on the polyol C, which can be attributed to a higher corrosion resistance of the ATO
method. The size of the ATO nanoparticles was 5.2 nm. The electrical support and the strong interaction between Pt redeposited after
conductivities of undoped and Sb-doped SnO2 were 0.0028 and 0.11 S electrochemical dissolution and the ATO support.
cm − 1, respectively. The specific surface areas of the ATO and the Vulcan Pang et al. [67] prepared Ru-doped SnO2 nanoparticles by chemical
XC-72R from a BET analysis were 99.7 and 239.6 m2 g − 1, respectively. precipitation and calcination at 550 °C. The structure and crystallite
Fig. 3a–f [65] shows the TEM images of ATO nanoparticles and of the size (20 nm) of SnO2 particles did not alter after Ru doping. Pt/Ru–
supported electrocatalysts. As shown in Fig. 3a and b, the ATO particles SnO2 was prepared by impregnation of Ru–SnO2 with H2PtCl6,
were uniform and spherical. In addition, the electron diffraction pattern followed by reduction of the Pt precursor with NaBH4. Pt particle
(inset of Fig. 4a) showed good crystallinity of the ATO particles. The size size was 4 nm. Safonova et al. [68] reported that the addition of Pd, Ru
of the ATO particles from TEM images was ~5 nm. Fig. 3c shows well- and Rh induce an increase in the resistivity of tin dioxide films.
distributed Pt particles on carbon support. The average particle size from Conversely, Pang et al. [67] found that the electrontransfer resistance
TEM was ~2.5 nm. It should be noted that all of the prepared (Rct) decreases with the introduction of Ru. The values of Rct on the
electrocatalysts were synthesized using the same Pt colloid; therefore, Pt/Ru-doped SnO2/graphite and Pt/SnO2/graphite electrodes were
the Pt particles should have the same average size in all of the prepared 4.92 and 6.28 Ω, respectively. This implies that Ru doping enhances
electrocatalysts. Fig. 3d, e and f shows TEM images of the Pt/ATO the conductance of SnO2. Ru-doping enhance not only the electrical
electrocatalysts for the 40, 10, and 5 wt.% Pt/ATO samples, respectively. conductivity, but also the activity and the stability of Pt/SnO2.
As the Pt loading decreased, the distribution of Pt particles on the ATO Compared with pure SnO2 particles, a certain amount of Ru can
support became remarkable. In the 40 wt.% Pt/ATO sample, Pt particles improve the electrocatalytic activity of the Pt/SnO2 catalyst. According
covered most of the ATO surface and were stacked on top of one another; to the authors, in the case of the Ru-doped SnO2 system the
whereas in the 5 wt.% Pt/ATO sample, the Pt particles were evenly interaction between RuOx and SnO2 promotes the formation of
distributed on the ATO surface. The activities of the Pt/ATO for both MOR hydroxyl species to catalyze the oxidation of poisonous species
and EOR were greater than those of the Pt/C. These enhanced activities (such as COads) adsorbed on the active sites of platinum nanoparticles,
752 E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763

Fig. 3. TEM images of (a) ATO, (b) magnified one of (a), (c) 40 wt.% Pt/C, (d) 40 wt.% Pt/ATO, (e) 10 wt.% Pt/ATO, and (f) 5 wt.% Pt/ATO. Reproduced from Ref. 65, copyright 2008,
with permission from Elsevier.

which results in the increased MOR activity of Ru-SnO2. Moreover, the 2.2.2. SnO2 nanowires
long-term cycle stability of the Pt/Ru-doped SnO2/graphite electrode Nanowires (NWs) are a class of one-dimensional nanomaterials
was higher than that of the Pt/SnO2/graphite electrode. However, with a high aspect ratio [70]. NWs can be made of various materials
comparison with Pt/C, which is fundamental for the evaluation of the and they have solid cores. The broader choice of various crystalline
suitability of a fuel cell catalyst support, was not made! materials and easier doping methods provide highly tunable proper-
Generally SnO2 has a surface area (b 100 m2 g − 1) lower than that of ties (e.g., electrical) of NWs [70]. While previous studies utilized SnO2
reference Vulcan XC-72. The synthesis of tin oxide with high surface area is in form of film or particles, Saha et al. [71,72] directly grew SnO2
a primary requirement for its use as catalyst support. Hagemeyer et al. [69] nanowires on fuel cell backings as catalyst support for fuel cells. NWs
investigated a variety of methods for the preparation of high surface area were grown on carbon-paper fuel cell backings form a 3D structure
tin oxide. BET surface areas N100 m2 g − 1 were achieved by a variety of that leads to a higher gas permeability. According to the authors, this
methods after calcination in the temperature range 300–500 °C. High NW-based 3D structure will take full advantage of combined factors
surface area tin oxides with excellent sintering resistance were including increased utilization of the noble metal catalyst, improved
synthesized by the hydrazine method. High surface areas N200 m2 g − 1 metal-support interaction, and enhanced mass transport in the
could also be achieved by a modified Pechini method, using aqueous electrodes for fuel cell applications. SnO2 NWs were grown directly
glyoxylic acid as dispersant for Sn(IV) acetate. on the fibers of a carbon paper by a thermal evaporation method.
E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763 753

deposition of Pt nanoparticles indicates a good electrical contact


between the SnO2 NWs and the carbon fibers. Higher electrocatalytic
activities of the Pt/SnO2 NW/carbon paper composite electrode for
both oxygen reduction reaction and methanol oxidation reaction have
been achieved in comparison with standard Pt/C electrode. In the
same way the electrocatalytic activities of SnO2 NW supported PtRu
composite electrodes for methanol oxidation were investigated and
higher mass and specific activities in methanol oxidation were
exhibited as compared to Pt–Ru catalysts deposited on glassy carbon
electrode [72]. The higher electrocatalytic activities of SnO2 NW
supported catalysts were attributed to the unique 3D structure and
electronic properties of SnO2 NWs, the strong interaction between Pt
catalyst particles and the SnO2 NW surface, the synergies resulting
from the combined properties of Pt nanoparticles and SnO2 NW
supports, and the low impurity level of SnO2 NWs compared to Vulcan
carbon XC-72. Such a hierarchical electrode with the combined
properties has potential applications for fuel cells. This approach also
provides a route to fabricate electrodes based on depositing noble
metal nanoparticles onto metal oxide NWs grown on the fuel cell
backings. One of the major concerns for the application of NWs to
catalyst supports in PEMFCs and DMFCs is whether there is strong
adhesion of these NWs to the carbon paper during actual fuel cell
operation. To address this concern, Saha et al. [71] carried out
preliminary studies by immersing the SnO2 NW/carbon paper
composite electrode in 0.1 M H2SO4 solution over a period of 2 months
at 50 °C. SEM examinations after the test showed that high-density
NWs are maintained on the surface of carbon fibers, suggesting strong
adhesion between SnO2 NWs and the carbon paper.

2.3. WOx

Tungsten oxide is an n-type semiconductor with a reported


bandgap of about 2.6 to 2.8 eV [73]. Tungsten has many oxidation
states, from 2 to 6, and can exist in many forms, making it suitable for
electrochromic, photochromic, photocatalyst, and gas-sensor applica-
tions. The intrinsic electric conductivity in tungsten oxide arises from
its nonstoichiometric composition, giving rise to a donor level formed
by oxygen-vacancy defects in the lattice [73]. It has also been shown
that tungsten oxides exhibit considerable proton transfer because of
the formation of tungsten trioxide hydrates [74], which is an attractive
property for fuel cell catalyst supports. Tungsten oxide support
Fig. 4. SEM and TEM micrographs of SnO2 NWs grown on carbon fibers of carbon paper
by thermal evaporation method. (a) SEM image showing full coverage of SnO2 NWs on
material has been shown to be more thermally stable under
fibers of carbon paper. (Inset) Fibers of bare carbon paper. (b) TEM image showing electrochemical oxidation conditions than Vulcan XC-72R [75].
individual SnO2 NWs. (c) TEM images showing Pt nanoparticles electrochemically Firstly, Niedrach et al. [76,77] reported that Pt admixed with
deposited onto SnO2 NWs. (Inset) Pt nanoparticles deposited onto a single SnO2 NW. tungsten oxides (e.g., WO2, W2O5, and WO3) is less susceptible to
Reprinted from Ref. 71, copyright 2007, with permission from The Electrochemical
carbon monoxide poisoning during its catalytic oxidation in fuel cells.
Society.
Then, Kulesza and Faulkner [78,79] observed that the electrocatalytic
activity for methanol oxidation is enhanced when platinum is
dispersed on WOx. The improvement in the electrocatalytic activity
Before the process of metal nanoparticle deposition, the SnO2 NWs of platinum in the presence of WOx is related to the fact that WOx can
grown on carbon paper were pretreated chemically with 5.0 M HNO3 form tungsten bronzes, thereby making the dehydrogenation of
aqueous solution. This surface activation step produces oxide methanol more effective and also the oxophilic nature of the oxide
functional groups such as hydroxyl, carboxyl and carbonyl on the helps in removing the adsorbed intermediates during methanol
surface of the SnO2 NWs. Pt nanoparticles were electrochemically oxidation [80]. Indeed, Hads coming from CH3OH dissociation can
deposited onto the surface of SnO2 nanowires [71]. Fig. 4 [71] shows spillover to the WOx substrate according to the reactions [80]:
the SEM and TEM images of the SnO2 NWs grown on a commercially
available carbon paper backing used in fuel cell applications. The WOx + yPtH Y Hy WOx + yPt ð2Þ
carbon paper is made of small carbon fibers with a diameter between
5 and 10 μm (inset in Fig. 4a). As shown in Fig. 4a, a thin layer of high-
density SnO2 NWs completely covers the surface of the carbon fibers
Hy WOx Y WOx + ye − + yH
+
ð3Þ

in the carbon paper. The length of the SnO2 NWs is in the range of 20–
30 μm. Further TEM examination shows that the NWs have a straight- leaving free Pt sites for further CH3OH adsorption/dissociation and
line morphology with a diameter of about 20 nm (Fig. 4b). Fig. 4c leading to accelerated kinetics. Moreover, methanol electrooxidation
presents the TEM image of the Pt nanoparticles electrochemically proceeds via COads species, and the best COads tolerance with Pt–WOx
deposited on the SnO2 NWs. Pt nanoparticles with a size of 4–6 nm materials was achieved when the interaction between Pt and WOx
can be observed clearly on the surface of NWs. The successful materials is maximized [81].
754 E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763

To avoid misunderstanding, it has to be promptly pointed out that spheres. Tungsten carbide (W2C) microspheres were synthesized by
in most works dealing on the Pt–WOx system, the catalyst has not heating resorcinol-formaldehyde polymer as a carbon precursor and
been dispersed on tungsten oxide to the aim to substitute the carbon ammonium metatungstate salt as a tungsten precursor. The carbon in
support! Pt was dispersed on the WOx matrix mainly to aim to the tungsten carbide microspheres was selectively removed by calcining
increase the catalytic activity of Pt [23,24]. When used in fuel cells, to tungsten carbide at 500 °C in an oxygen atmosphere for 3 h to obtain
increase WO3 surface area, before Pt dispersion, tungsten oxide is WO3 microspheres. Platinum particles were supported on the WO3
supported on carbon [24,81] or, to increase Pt dispersion, WOx is microspheres and the carbon microspheres by the conventional
deposited on preformed Pt/C [82,83]. borohydride reduction method. By XRD analysis they found that the
Mcleod and Birss [84], for the first time, evaluated WOx/Pt films in average crystallite size is 6.5, 15 and 12 nm for Pt, WO3 and W2C,
terms not only of the suitability of WOx as a co-catalyst for methanol respectively. The pore-size distribution of WO3 showed a large number
oxidation, but also of the relevance of WOx as a support material for of pores of around 45 nm. The large pore-size was attributed to the inter-
fuel cell catalysts. Then, recently, Ganesan and Lee [85] and Cui et al. particle spaces between WO3 nanoparticles of ~15 nm in diameter. The
[86], in microsphere and mesoporous form, respectively, and Rajeswari BET surface area of the WO3 microspheres was about 18 m2 g − 1. WO3
et al. [87] and Maiyalagan and Wiswanathan [88], in nanorod form, supported Pt displayed higher activity for the electrochemical oxidation
tested WO3 as a carbon substitute support for DMFC catalysts. of methanol than a commercial Pt–Ru/C catalyst by a factor of 1.9 (per
Currently, DMFC electrocatalysts are supported on carbon, which is mass of Pt) even without Ru. The onset potential and peak potential of
electronically, but not ionically conductive. This necessitates the methanol oxidation on Pt/WO3 was shifted by 100 and 50 mV more
addition of the fuel cell electrolyte, typically a polymer such as negative than PtRu/C. This is probably due to better CO tolerance of the
Nafion™, into the PtM/C mixture to create a high surface area triple catalyst. The stability of the Pt/WO3-microsphere catalyst was tested by
phase boundary where the catalyst, fuel, and electrolyte meet. Any Pt repeating electrochemical reaction cycles in a 1 M H2SO4–1 M CH3OH
sites that are not in contact with the Nafion™ electrolyte cannot act solution. The specific activity increased initially, but stabilized after
catalytically and are therefore wasted. By using a support material such about 15 cycles. Once a steady-state is established, there was no sign of
as WOx, which is both ionically and electronically conductive, a deactivation during 100 consecutive reaction cycles.
significantly greater portion of the Pt catalyst should be utilized. A Cui et al. [86] synthesized mesoporous WO3 with ordered pore
problem for the use of WOx as a fuel cell catalyst support could be the structure by a template replicating route. The mesoporous WO3
dissolution of WO3 in the acidic medium [89]. Raghuveer and support materials were prepared using mesoporous silica with cubic
Viswanathan [24] reported that the stability of WO3 in the acid Ia3d symmetry as hard template and 12-phosphotungstic acid as
medium can be improved by transition metal ion substitution such as tungsten source, followed by the removal of the hard template using
Ti4+ into the oxide framework. The improvement in the stability could 2 M HF solution. Pt/WO3 was obtained by deposition of H2PtCl6 on the
be due to the alteration in the rate of hole capturing process of WO3 by mesoporous WO3, followed by reduction with NaBH4. The average
the Ti4+ at the interface, which is responsible for the dissolution of crystallite size of Pt and WO3 was found to be about 6.5 nm and
tungsten in acidic solution. McLeod and Birss [84] synthesized WOx 6.2 nm, respectively. The BET specific surface area of the prepared
from W(OC2H5)6 by two different methods using the sol–gel approach, mesoporous WO3 was 86 m2 g − 1, and the wall thickness (diameter of
with ethanol (Type 1) or water (Type 2) as the solvent. Through the use WO3 nanorods or nanoparticles) is 6.2 nm, close to the pore size of
of sweep rate experiments, they found that Type 1 films had both a silica template (6.51 nm). TEM images of 15 and 20 wt.% Pt/WO3, in
smaller total electroactive area and fewer electroactive surface sites Fig. 5 [86], parts a and b, respectively, show that Pt nanoparticles
than Type 2 films. This can be attributed to a higher film density, and (darker dots) are dispersed well in the pore structure of mesoporous
therefore, a lower porosity of Type 1 film. XRD analysis revealed that WO3. No very large metal particles/aggregates can be found. However,
the products of both syntheses were either amorphous, or poorly some metal Pt nanoparticles may exist outside of the mesopores of
crystallized. Methanol oxidation CVs for WOx/Pt films made from both WO3 when the Pt loading amount comes up to 20 wt.% as shown in the
Type 1 and Type 2 syntheses exhibited a low resistance, which is a selected area in Fig. 5b. Small Pt nanoparticles can be found existing
necessary property of a fuel cell catalyst support in order to maximize within the pores of the mesoporous WO3 support in a larger
the cell efficiency. The WOx/Pt films derived from both Type 1 and Type magnification image of Fig. 5c. The high-resolution TEM in Fig. 5d
2 sols clearly exhibited methanol oxidation activity. For Type 1 films, shows two kinds of crystal lattice stripes. The smaller one of
methanol oxidation currents are observed to increase with time, which d = 0.23 nm belongs to the Pt metal, and another one of
was matched with a decrease in CV charge and a loss of WOx, while the d = 0.27 nm should be WO3, which is an additional evidence to
methanol oxidation currents on the more porous Type 2 films are more prove the Pt particles existing in the pores of the support. A more
stable. Methanol oxidation currents for the Type 1 film were also direct evidence of Pt nanoparticles dispersed in the WO3 matrix is the
observed to decrease as the WOx:Pt ratio increased. Thus, it can be clear Pt signals easily detected by EDS in Fig. 6e obtained on these
concluded that the Type 1 WOx blocked Pt sites and that the increase in composites. Pt/WO3 showed high electro-catalytic activity toward
methanol oxidation currents with time is due to a physical loss of WOx methanol oxidation and good electrochemical stability. The activity
and a corresponding increase in the number of exposed Pt sites. for methanol oxidation of 20 wt.% Pt/WO3 was significantly higher
Regarding Type 2 films, above a WOx:Pt ratio of 3 it exhibited methanol than that of commercial 20 wt.% Pt/C catalysts, and comparable to or
oxidation currents that decayed only very slowly. In terms of a fuel cell even higher than that of 20 wt.% PtRu/C catalyst by ETEK in the
catalyst support, WOx produced via the Type 2 synthesis seems to be potential region 0.5–0.7 V and lower than 20 wt.% PtRu/C in the low
more promising than Type 1 WOx. The more porous structure of the potential region 0.3–0.4 V. The enhanced electro-catalytic activity for
Type 2 films allows methanol to reach the Pt sites, even as the amount methanol oxidation on Pt/WO3 was ascribed to the enhanced catalytic
of WOx on the electrode is increased. The WOx synthesised in this effect and the mesoporous structure of WO3 support.
manner presented some co-catalytic properties, as evidenced by the
stable methanol oxidation currents of Type 2 films with WOx:Pt molar 2.3.1. WO3 nanorods
ratios of 3:1 or higher. According to the authors, combined with the Recently, one-dimensional (1D) nanorods have attracted remark-
overall ease of the synthesis method, these advantages of Type 2 films able attention owing to their unique properties and potential for
suggest that this may be a valid approach to synthesis a DMFC various novel applications [90]. Nanorods have large specific surface
electrocatalyst, without the need for a carbon support. areas, so WO3 in a nanorod form could be an efficient support for FC
Ganesan and Lee [85] prepared tungsten trioxide microspheres of 2– catalysts. Rajeswari et al. [87] synthesized WO3 nanorods by pyrolysis
4 μm diameter by controlled oxidation of tungsten carbide micro- of tetrabutylammonium decatungstate ((C4H9)4N)4W10O32 at 450 °C.
E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763 755

Fig. 5. TEM images of mesostructured 15 wt.% Pt/WO3 (a, c), 20 wt.% Pt/WO3 (b); high-resolution TEM image (d) and EDS pattern (e) of 15 wt.% Pt/WO3. Reproduced from Ref. 86,
copyright 2008, with permission from the American Chemical Society.

In the precursor compound, inorganic metal oxide clusters (W10O432−) acid medium. Whereas for the bulk WO3, the peak current was
are charge balanced with the surfactant group tetrabutylammonium observed to decrease, showing a poor stability in acid medium. The
bromide (TBABr). TBABr restricts the irregular arrangement of the electrochemical activity for methanol oxidation of platinum loaded
metal oxide clusters by providing a steric environment around them. WO3 nanorods, and for comparison of a commercial 20 wt.% PtRu/C
This leads to the formation of 1D array of metal oxide clusters. During and Pt loaded on bulk WO3, was studied by CV in 1 M H2SO4/1 M
pyrolysis at 450 °C, removal of the surfactant species ((C4H9)4N+) and CH3OH. The catalyst activities were in the following order: Pt/WO3
decomposition of the W10O32 leads to the formation of pure, single nanorods NPtRu/C N Pt/WO3 bulk. The activities of Pt/WO3 nanorods
crystalline WO3 nanorods. The dimensions of the nanorods varied in was approximately six times higher than that of the Pt/WO3 bulk,
the ranges of 130–480 nm and 18–56 nm of length and width, clearly indicating that the nanostructure of the catalytic support plays
respectively. Pt nanoparticles were supported on WO3 (20 wt.% Pt) by an essential role. MOR activity of Pt loaded WO3 nanorods was slightly
the wet impregnation method. The size of the Pt nanoparticles was is higher or comparable to that of the commercial PtRu catalyst. This
in the range of 4–6 nm. The electrochemical stability of WO3 nanorods shows that replacement of the noble metal Ru in the commercial
and bulk WO3 was studied in 1 M H2SO4. The stability was evaluated catalyst is possible with WO3 nanorods. All these facts demonstrate
by performing uninterrupted cycling of the potential between −0.2 that WO3 nanorods are better supports for platinum than bulk WO3
and 1.0 V. For the WO3 nanorods almost no activity variation was and carbon and hence they constitute a better catalytic system for
observed even after several cycles indicating its stability in sulfuric methanol oxidation.
756 E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763

Fig. 6. SEM and reflected electron (COMPO) images of the S–ZrO2 and Pt/S–ZrO2 particles. Pt particles were whitened in COMPO image. (a) SEM image of S–ZrO2. (b) SEM image of
Pt/S–ZrO2. (c) COMPO image of Pt/S–ZrO2. Reprinted from Ref. 104, copyright 2007, with permission from The Electrochemical Society.

Maiyalagan and Wiswanathan [88] synthesized tungsten oxide materials control their electrical properties through the balance of
nanorods infiltrating a phosphotungstic acid (H3PW12O40) solution electron and proton transport.
into an Anodisc alumina membrane as a template, and platinum Long et al. [19] demonstrated that, for Pt–Ru nanoscale blacks in
nanoparticles were supported on the nanorods by an impregnation which the Pt–Ru chemical state is known and varied from Pt0Ru0 to
method. The diameter of the nanorods was found to be around Pt–RuOxHy to Pt–RuO2, Pt0Ru0 has orders of magnitude less activity for
200 nm. The size of the Pt nanoparticles was in the range of 3–4 nm. the oxidation of methanol than does a mixed-phase electrocatalyst
WO3 stability tests, carried out by repeated voltammetric cycles in 1 M containing Pt0 and RuOxHy. The importance of the proton-conducting
H2SO4, showed that WO3 nanorods have higher stability compared contribution of RuOxHy in Pt–Ru nanoscale electrocatalysts was
with bulk WO3. Pt/WO3 nanorods exhibited higher methanol demonstrated by the drastic loss of activity for methanol oxidation
oxidation activity than the commercial Pt/C catalyst by a factor of once the structural water is thermally removed from the hydrous
two. Platinum supported on such nanorods was found to be stable oxide to form RuO2. They also show that bulk, rather than near-
over several cycles in an electrochemical environment. surface, quantities of RuOxHy are required to achieve the highest
activity for methanol oxidation. The presence of RuOxHy in practical
2.4. RuO2·xH2O DMFC catalysts has important mechanistic implications for direct
methanol oxidation, because RuOxHy is mixed III/IV valent; conducts
Ruthenium dioxide, RuO2, belongs to the family of transition-metal both electrons and protons; and innately expresses Ru–OH speciation,
dioxide compounds with rutile structures. Pure RuO2 is a metallic as denoted by Eq. (4) [19]:
conductor due to a d-band conduction mechanism [91]. Single crystals
p Y Ru
IV 2− + − III
of anhydrous RuO2 exhibit conductivity on the order of 104 S cm − 1, Ru O + H + e OH: ð4Þ
despite an oxygen deficiency which is charge compensated by Ru3+
defects in the lattice. Unlike its anhydrous, crystalline parent rutile, RuOxHy also efficiently dissociates H2O, as that is one of the
hydrous ruthenium dioxide, RuO2·xH2O or RuOxHy, exhibits mixed mechanisms by which protons are transported in proton conductors.
electron-proton conductivity [92]. RuO2·xH2O has important techno- RuO2·xH2O was investigated as active support material for fuel cell
logical applications and opportunities in electrocatalysis [93] and catalysts by Lasch et al. [97] and by Chen et al. [98]. In both works,
charge storage [94]. The conductivity of a 25–40% dense pellet of RuO2·xH2O was prepared by oxidation of ruthenium chloride with
RuO2·yH2O (y ~ 1.3) is ~ 1 S cm − 1 [95]. This indicates that Ru(III) sodium hydroxide. The catalyst was deposited on the oxide support by
defects are present and exhibit electronic conductivity due to a colloidal method. First, Lasch et al. [97] investigated the activity of
semiconduction or electron hopping from Ru(IV) centers to Ru(III) anhydrous RuO2 and hydrous RuO2·xH2O in half-cell experiments by
defects. EXAFS analysis results for partially hydrated RuO 2 stationary current voltage measurements. Pt–Ru-catalysts deposited
(RuO2·0.29H2O) indicated that the structure is a disordered RuO2 on a Vulcan XC-72 carbon black were used for comparison. High
rutile-like structure, where the RuO6 octahedra are similar to those surface area ruthenium oxides (≥120 m2 g − 1) were used as catalyst
found in the crystal structure, while the three-dimensional network of supports for Pt- and PtRu-catalysts. The electric conductivity of these
octahedral chains does not extend as far as in RuO2 rutile [96]. materials was comparable to that of Vulcan XC-72. X-ray diffraction
Amorphous RuO2·2.32H2O had disordered RuO6 octahedra connected showed that hydrous RuO2·xH2O consists of amorphous phases which
in chains, but the data show no evidence of a three-dimensional are partly reduced to metallic ruthenium in the course of the catalyst
network of octahedra. The orientation of water within this structure is preparation. For that reason, a part of the material was annealed at
not discernible from the EXAFS data, but clearly the presence of 370 °C in oxygen. From the XRD spectra, it was found that the calcined
structural water disrupts the three-dimensional rutile structure [96]. material consists of crystalline anhydrous RuO2. No Ru-metal phases
The structures of RuOxHy materials correlate to their mixed con- are observed. However, anhydrous RuO2 is a very bad proton
ductivity and pseudocapacitance. The local structures of RuOxHy conductor and therefore did not show any contribution as an ‘‘active
E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763 757

support material’’. Deposition of noble metal catalysts on calcined ionomer in the catalyst layer might be able to be reduced using S–ZrO2 as
RuO2 resulted in electrochemical activities comparable to those the support.
obtained for the Vulcan XC-72 support. On the other hand, catalysts As it is well known, the proton conductivity of the proton exchange
supported on hydrous RuO2 were completely inactive. The formation membranes, such as Nafion or sulfonated poly(ether ether ketone)
of a metallic ruthenium phase either reduces drastically the proton (SPEEK) membranes, depend on the water content [105]. The unsatu-
transfer to the support material or it blocks the catalytic activity of the rated-humidification or non-humidification of reactant gases prior to
deposited Pt-particles. It is well known that the ideal Ru-content for entry into a fuel cell may lead to a decreased water content, and
electrochemical methanol oxidation should not be larger than 50% accordingly reduced proton conductivity of the membranes. The
[99], which in this case was much exceeded. Conversely, Chen et al. reduction in membrane thickness has facilitated water back diffusion
[98] found that Pt/RuO2·xH2O possess higher electrocatalytic activity from the cathode to the anode, accordingly resulting in not significantly
towards methanol oxidation with respect to PtRu black catalysts. It has impaired proton conductivity of the operating membrane. However, this
to be promptly pointed out, however, that the comparison with PtRu also increased the reactant crossover because of the thinner membrane,
black is not appropriate, due to its low surface area. which may reduce the cell performance and accelerate the membrane
degradation. To overcome these problems, Uchida et al. [106] proposed
2.5. Sulfated zirconia (S–ZrO2) the use of self-humidifying membranes with highly dispersed nan-
ometer-sized Pt and a hygroscopic oxide (TiO2). The Pt nanoparticles in
Zirconia, when modified with anions such as sulfate, forms a highly the membrane could suppress the reactant crossover by catalytic
acidic or super-acidic catalyst that exhibits superior catalytic activity in recombination of the crossover H2 and O2, and the hygroscopic oxide
many reactions. Sulfated zirconia (SO2− 4 /ZrO2, S–ZrO2) is a crystalline contributed to maintain the membrane water equilibrium. On this basis,
solid acid that has monoclinic and tetragonal phases [100]. Sulfated a different use of the Pt/S–ZrO2 catalyst than an electrode material was
zirconia is a strong superacid with a Hammett acid strength H0 of −16.03, proposed by Zhang et al. [107]. They prepared SPEEK/Pt/S–ZrO2 self-
whereas Nafion® has only about −12 [101]. Data on the stability of humidifying membranes based on a sulfonated poly(ether ether ketone)
sulfated samples vs. temperature in the presence of methanol and water hybrid with sulfated zirconia supported Pt catalyst. The single cell
indicated that S–ZrO2 retains a high activity at 300 °C [102]. S–ZrO2 with a performance employing the SPEEK/Pt/S–ZrO2 membrane was evaluated
high atomic ratio of S/Zr has high proton conductivity: the proton and the resistance of the operating membrane under wet or dry H2/O2
conductivity of the S–ZrO2 with S/Zr=0.35 was 5 10- 2 S cm − 1 at 60– operations was also investigated, and compared to that of the plain
150 °C [103]. Moreover, S–ZrO2 has a high hydrophilicity. These SPEEK membrane. A single cell employing the SPEEK/Pt/S–ZrO2
characterstics make S–ZrO2 a suitable support for fuel cell catalysts. In membrane exhibited higher cell OCV values and performance than the
particular, due to its high proton conductivity, it is expected that the plain SPEEK membrane cell under both dry and wet conditions.
content of the recast electrolyte ionomer in PEMFC electrodes could be Furthermore, the performance of the single cell with a SPEEK/Pt/S–
reduced using S–ZrO2 as catalyst support. The reduction of ionomer might ZrO2 membrane was only slightly influenced by the humidification
result in the improvement of the gas diffusion and water transportation. conditions compared to that of the plain SPEEK membrane. The higher
In addition, the increase of platinum utilization is also expected. Because OCV values and the lower resistance for the SPEEK/Pt/S–ZrO2
the perfluorosulfonated ionomer (PFSI) cannot soak into small pores in membrane under dry operation, as compared to of the plain SPEEK
the carbon support, platinum particles that are deposited in the small membrane, were attributed to the incorporation of catalytic, hygroscopic
pores may have little contact with a proton conductor. As a result, the and proton-conductive Pt/S–ZrO2 catalyst inside the self-humidifying
platinum utilization becomes low. S–ZrO2 has acid sites over the surface membrane. However, no comparison was made with Nafion membrane.
in even such small pores, so that the platinum on S–ZrO2 can donate and
accept a proton and the platinum utilization would increase. Moreover, 2.6. SiO2
considering its high hydrophilicity, the S–ZrO2 might improve the cell
performance under a low relative humidity. On this basis, Suzuki et al. Silicon dioxide, also known as silica, is an oxide of silicon with a
[104] tested sulphate zirconia as Pt support for use in PEMFC. They used chemical formula of SiO2 and has been known for its hardness since
commercially available S–ZrO2 powder (Wako Pure Chemical Co., ca. antiquity. SiO2 has a number of distinct crystalline forms in addition to
80 m2 g − 1) as a support. Pt was deposited on S–ZrO2 by using ultrasonic amorphous forms. With the exception of stishovite and fibrous silica,
spray pyrolysis. The Pt content of the Pt/S–ZrO2 was 53 wt.%. Fig. 6 [104] all of the crystalline forms involve tetrahedral SiO4 units linked
shows SEM images of the S–ZrO2 and the Pt/S–ZrO2 particles. The particle together by shared vertices in different arrangements.
size of the S–ZrO2 was approximately 50-100 nm (Fig. 6a). As shown in Firstly, as in the case of Pt/S–ZrO2, Zhu et al. [108] and Wang et al.
Fig. 6b, highly dispersed nanoparticles were observed on the Pt/S–ZrO2. [109] used silica supported Pt to prepare self-humidifying mem-
The composition image (Fig. 6c) shows that the nanoparticles were branes. Zhu et al. [108] prepared a self-humidifying membrane with a
platinum. Therefore, it was found that platinum was highly dispersed on sandwich structure with Nafion-impregnated porous PTFE composite
the S–ZrO2 surface as nanoparticles. Regarding the electrocatalytic as the central layer and nanosized SiO2 supported Pt catalyst
activity of Pt/S–ZrO2, in the presence of Nafion ionomer, PEMFC embedded into the Nafion as the two side layers. Wang et al. [109],
performance of the Pt/S–ZrO2 cathode was lower than that of the Pt/C. instead, dispersed Pt/SiO2 catalyst particles uniformly into the Nafion
Because the slope of the voltage-current curve for the Pt/S–ZrO2 was resin, and then a Pt/SiO2/Nafion/PTFE reinforced composite mem-
similar to that of Pt/C in the high current density range (1.5–2.5 A cm − 2), brane was prepared using a solution-cast method. The Pt/SiO2 catalyst
there was no large iR loss caused by low electric conductivity of S–ZrO2. was prepared from H2PtCl6 and silicon dioxide (particle size of 20 nm)
Conversely, in the absence of Nafion ionomer, the cell performance of Pt/ by an impregnation/reduction method (Pt loading 1 wt.%, [108]) of
S–ZrO2 cathode was higher than that of Pt/C. Because the lack of Nafion silicon oxide or via a microwave-assisted polyol process (Pt loading
ionomer decreased the proton conductivity in the catalyst layer, the cell 2 wt.%, [109]). In both cases, the self-humidifying membrane
performance decreased in both cathodes without Nafion ionomer. outperformed the plain Nafion/PTFE membrane. The self-humidifying
However, the decrease of the cell performance on the Pt/S–ZrO2 was composite membranes could minimize membrane conductivity loss
smaller than that on the Pt/C. For instance, although the voltage decrease under dry conditions. According to the authors, this is due to the
of Pt/C at 1 A cm − 2 was approximately 0.22 V, the decrease of Pt/S–ZrO2 presence of Pt/SiO2 in the membrane.
was 0.10 V. The ratio of the cell voltage decrease for Pt/S–ZrO2 was 16%, More recently, Seger et al. [110] tested silica supported Pt as
lower than the 33% for Pt/C. This result suggests that Pt/S–ZrO2 support cathode material in PEMFCs. They prepared platinum nanoparticles by
provided enough protons to obtain a large current and the content of the reduction of a Pt(IV) precursor with sodium borohydride on a
758 E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763

colloidal silica substrate. By controlling the ratio of platinum to silica, 3.1. Earlier works on metal carbides as fuel cell catalyst supports
they obtained a well dispersed catalyst surface with minimal
conductivity loss. At low Pt concentrations (Pt:SiO2 ratio of 1:1 and Firstly in the 1960s at General Electric, boron carbides were tested
2:1), Pt dispersion is quite uniform and form an interconnected as a catalyst support in phosphoric acid [115] and alkaline [116] fuel
particle-network. The current interruption method was employed to cells. Grubb and McKee [115] investigated boron carbide supported
determine any possible ohmic voltage losses arising from the presence platinum as the anode material in propane/oxygen PAFCs. A
of silica in the fuel cell. The Pt/SiO2 samples showed ohmic losses commercial boron carbide powder was impregnated with a Pt
similar to that of the commercial Pt-black sample. These results precursor, and then Pt was reduced with H2 at high temperature. Pt
confirm that the silica support does not contribute to any significant crystallite size ranged from 1.6 to 4.7 nm, and increased with
increase in the overall resistivity. The performance of Pt/SiO2 increasing Pt loading. They found that Pt on boron carbide is more
composite particles in H2-fuel cell were compared to that of a Pt- resistant to sintering than platinum black or platinum on graphite at
black catalyst. All these systems exhibit losses due to polarization and equal surface coverage. Tests in PAFCs showed that the slopes of the
−1
further efforts are necessary to overcome these effects. The normal- current density vs. RHE/Pt loading plot are 5 mA mgPt for platinum
−1
ized power density curves highlighted the beneficial aspects of Pt/ on boron carbide and 0.6 mA mgPt for platinum black diluted with
SiO2 catalysts, and the Pt/SiO2 sample at the 2:1 ratio exhibited the boron carbide. This improvement by a factor of 9 is partially due to a
highest power density values. higher platinum surface area for the supported Pt. Then, the current
density was plotted against the real Pt area per unit electrode area.
The slopes of these straight lines were 9.3 and 3.5 μA cm − 2 for boron
2.7. Indium tin oxide (ITO, Sn–In2O3)
carbide supported platinum and unsupported platinum black,
respectively. According to the authors, this factor of improvement
Indium tin oxide (ITO) is essentially formed by substitutional doping
could indicate an enhancement of the Pt activity by the support.
of In2O3 with Sn which replaces In3+ atoms in the cubic bixbyite
Platinum iridium alloy catalysts supported on boron carbide were
structure of indium oxide [111]. In the In2O3 crystalline structure, the
investigated as anode materials in ammonia/oxygen alkaline fuel cells
replacement of In3+ by Sn4+ affects the optical and electrical properties
at 100 –120 °C using 54% potassium hydroxide as electrolyte [116].
of the ITO films. The replacement produces free electrons that contribute
Platinum and iridium were deposited together on boron carbide and
to enhancing the conductivity (N103 S cm − 1) [112]. Chhina et al. [113]
graphite substrates by thermal decomposition of mixed solutions of a
studied the electrochemical stability of indium tin oxide supported
platinum salt, Pt(NH3)2(NO2)2, and iridium chloride, IrCl3. The
platinum. Pt was deposited on ITO by the reflux method, and both the
catalysts were reduced in flowing hydrogen at 125 °C for 3 h. Cell
thermal and electrochemical stability were compared with that of a
resistances were about 0.3–0.35 Ω for boron carbide anodes and 0.2–
commercially available 40 wt.% Pt/C catalyst (Hispec 4000). The
0.25 Ω for graphite anodes. Satisfactory electrodes were obtained
crystallite sizes calculated from XRD were 38 nm and 13 nm for ITO
using both boron carbide and graphite as the catalyst substrate. Boron
and Pt/ITO, respectively. The total currents in the cyclic voltammetry
carbide-based anodes were, however, generally less reproducible than
tests for Pt/ITO were much lower than those for Pt/Hispec 4000,
those prepared from graphite.
because the active surface area of the Pt particles on the ITO is much
Then, in the end of 1980 s, silicon and titanium carbides were
lower than that in Hispec 4000. The ITO support was more stable to
investigated as supports for H2/O2 PAFC catalysts [117,118]. SiC is
oxidation than Vulcan XC-72R carbon. The ITO supported catalyst had a
stable in hot phosphoric acid, but it is not electronically conducting.
lower electrochemically active surface area loss under an accelerated
Honji et al. [117] evaluated silicon carbide as a candidate catalyst
voltage cycling test, and lower mass loss in TGA experiments compared
support material, achieving electronic conduction by incorporation of
to the Pt/carbon catalyst.
carbon black in the catalyst layer during the fabrication of the
electrode. Platinum supported on silicon carbide was prepared by
3. Carbides reduction of chloroplatinic acid with methanol in the presence of
silicon carbide powder and a surface active agent. The platinum
Carbides are compounds composed of carbon and a less electro- content was in the range 5–23 wt.%. The average particle diameter of
negative element. Carbides can be generally classified by chemical the silicon carbide powder was about 0.3 μm and its specific surface
bonding type as follows: (i) salt-like, (ii) covalent compounds, area about 10 m2 g − 1. Pt particles were well dispersed on the SiC
(iii) interstitial compounds, and (iv) “intermediate” transition metal surface. Pt particle size was 4.3, 5.7, and 7.8 nm in 5, 10, and 23 wt.%
carbides. The carbides of silicon and boron are described as “covalent platinum, respectively. The potential-current density plots for oxygen
carbides”, although virtually all compounds of carbon exhibit some reduction on an electrode using Pt/SiC electrocatalysts depended on
covalent character. Silicon carbide has two similar crystalline forms, the platinum content, and 23 wt.% Pt/SiC shows the lowest
which are both related to the diamond structure. Boron carbide, B4C, polarization, that is about 0.7 V vs. RHE at 220 mA cm − 2 (iR
on the other hand has an unusual structure which includes ico- corrected), though the platinum particle size of 23 wt.% Pt/SiC was
sahedral boron units linked by carbon atoms. In this respect boron larger than those of other Pt loadings. This value was almost the same
carbide is similar to the boron rich borides. Both silicon carbide, SiC, as that of platinum supported on carbon. Titanium carbide was
(carborundum) and boron carbide, B4C are very hard materials and identified as a satisfactory electrical conductor and appeared to be
refractory. The carbides of the group 4, 5 and 6 transition metals (with stable in hot phosphoric acid at oxygen electrode potentials in a fuel
the exception of chromium) are often described as interstitial cell. However, commercially available titanium carbide has a relatively
compounds. These carbides are chemically quite inert, have metallic low surface area (b1 m2 g − 1) and therefore is not generally useful as
properties and are refractory. Some exhibit a range of stoichiometries, a catalyst support. Jalan and Frost [118] prepared titanium carbide
e.g. titanium carbide, TiC. Titanium carbide and tungsten carbide are supports having a crystallite size in the range of about 5–50 nm and
important industrially and are used to coat metals in cutting tools. The relatively large average surface areas of 25–125 m2 g − 1 by contacting
carbides of transition metals, especially tungsten carbides, exhibited titanium tetrachloride in the vapor phase with a gaseous unsaturated
Pt-like catalytic properties because near the Fermi level, the electronic hydrocarbon (ethylene, acetylene) and hydrogen at 500–1250 °C.
density of states of tungsten carbides resembles that of noble metal They found that the titanium carbide support has a desirable chain-
platinum [114]. This means that these carbides are expected to be used like structure which maintains electrical contact and has an open
as electrocatalysts by themselves or as catalyst supports, which might porosity providing for an oxygen diffusion path and thus providing an
promote catalytic activity through synergistic effects. improved electrode, particularly for a fuel cell. Pt was supported on TiC
E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763 759

by impregnation/reduction methods, and the platinum surface area compared the stability of Pt supported on commercial WC
was in the range of 20–90 m2 g − 1. The supports were characterized (BET = 1.6 m2 g − 1) with that of home made Vulcan XC-72 supported
by improved electrochemical corrosion resistance and the catalytic Pt [125] and a commercial Pt/C [126]. In both works, the stability of
activity of Pt/TiC in fuel cells was similar to that of electrodes made Pt/WC was higher than that of Pt/C. In the first work, similar amounts
with carbon black supports. of Pt (40 wt.%) were dispersed onto tungsten oxide and carbon
supports. However, due to the density difference between C and WC,
3.2. Tungsten carbides as PEMFC/DMFC catalyst support materials 1.8 g cm − 3 and 15.6 g cm − 3, respectively, a comparison between
40 wt.% Pt on WC and 40 wt.% Pt on C results in a tendency towards the
More recently, the use of tungsten carbides as support for low observation of a lower stability of the WC, as the volumetric loading of
temperature fuel cells has received considerable attention. As the catalyst on WC will be higher than that on C for the same
electrocatalysts, they are known to be highly resistant to CO poisoning gravimetric loading. Therefore, an additional comparison was also
and stable in acidic and basic solutions, yet their electrocatalytic made with a smaller volumetric loading of the catalyst on the WC
activity for methanol oxidation is very low. However, this low activity support. In this comparison, 40 wt.% Pt was dispersed on C, and a
of tungsten carbide could be improved dramatically by adding a small similar volume fraction of Pt was dispersed on the WC support
amount of platinum to tungsten carbide. Indeed, the methanol material. Equal solid volume ratios result in 6 wt.% Pt on WC. In both
oxidation on tungsten carbide films, modified by low coverages of cases, Pt/WC was more stable than Pt/C. The electrochemical activity
Pt, shows a promoting effect of Pt for the dissociation of methoxy of the Pt/WC remained nearly constant over 100 accelerated oxidation
[119]. Analogously, the presence of tungsten carbide promotes the cycles, while the activity of Pt/C was almost completely lost after only
electrochemical activity for oxygen reduction on platinum: the approximately 20 oxidation cycles. The high stability of Pt/WC was
overpotential is significantly reduced on a tungsten carbide nano- ascribed to the surface oxidation of W during the first scan, changing
crystal modified Pt catalyst, showing a synergic effect to improve the the structure of the support from Pt supported on WC to Pt supported
ORR activity [120]. Generally, tungsten carbide exists in different on a WOx shell encapsulating a WC core. Tungsten oxide, by reacting
phases, of which the most important are tungsten monocarbide (WC) with hydrogen, can form two stable hydrogen tungsten bronzes,
and subcarbide (W2C). WC has the lowest electrical resistivity H0.18WO3 and H0.35WO3, or sub-stoichiometric tungsten oxides. Since
(conductivity = 105 S cm − 1 at 20 °C) of any interstitial carbides, and tungsten oxide exhibits fairly high electrochemical stability, its
therefore qualifies as the most metallic carbide. A major drawback formation is less deleterious to the electrochemical performance of
preventing WC from being widely used as a catalyst support, however, the cell than the oxidation of a carbon catalyst support, which leads to
is the low surface area of tungsten carbide produced by traditional irreversible platinum detachment and corresponding complete
metallurgical routes. WC can be made by direct carburisation of W activity loss. However, the initial activity of the Pt/WC supported
metal with carbon or graphite at 1400–2000 °C in hydrogen or catalyst is much lower than that of Pt/C at comparable volumetric
vacuum. The carbide formation process can also use tungsten oxide, catalyst loadings, indicating that higher surface area WC supports and
tungstic acid, or ammonium tungstates as the starting materials. better platinum dispersion techniques are still required in order to
Carburization at high temperatures leads to low specific surface area enhance the catalyst layer activity. Synthesis of tungsten carbide with
carbides. higher surface area was reported by Koc and Kodambaka [127]. They
As in the case of WOx, various papers report the use of tungsten prepared tungsten carbide powders from novel carbon coated
carbide dispersed on a high surface area carbon (WC/C) as catalyst precursors. The process has two steps in which the oxide powders
support: WC is synthesized on a carbon support and then Pt is were first coated with carbon by cracking of a hydrocarbon gas,
deposited to produce a Pt/WC/C [121–123]. In this review, we do not propylene (C3H6), secondly mixed with a substantial amount of
consider this type of support, because it is not a carbon substitute. carbon black, and finally treated at temperatures in the range of 600–
Zellner and Chen [119] investigated the stability of WC and W2C by the 1400 °C to synthesize WC. A preparation method consisted in
combination of electrochemical CV measurements in 0.5 M H2SO4 and depositing 7.6 wt.% pyrolytic carbon on tungsten oxides and mixing
XPS measurements. They found that the WC film is stable at anode these composite particles with carbon black. The initial surface area
potentials below 0.6 V, demonstrating the potential use of WC as an was about 32 m2 g − 1. Surface area increases with temperature to a
anodic electrocatalyst in the DMFC. In contrast, W2C does not have a maximum of about 50 m2 g − 1 at 900 °C and then linearly decreases as
stable region, suffering immediate oxidation to form WxOy species the temperature increases. During the reduction process, oxide
when exposed to air or in an electrochemical environment. Zhang et particles were reduced to tungsten and oxygen was removed in the
al. [124] compared the electrochemical activity and stability in PEMFC form of CO or CO2. The increase in surface area was ascribed to the
environment of WxCy supported Pt with that of carbon supported Pt. reduction of the oxides and the formation of fine W particles. At
The tungsten carbide support was prepared through the carbonization temperatures above 1000 °C, the decrease in the surface area was
method using Vulcan XC-72 and H2WO4 as precursors. Pt was explained by the carburization of tungsten and the formation of necks
deposited on the tungsten carbide and Vulcan XC-72 by the among some of the particles. To prepare high surface are WC, Chhina
formaldehyde reduction method. The Pt loading of the catalyst was et al. [126] used high surface area carbon as a template upon which
10 wt.%. XRD patterns showed the various diffraction peaks of WC, tungsten was dispersed. They synthesized WC on a high surface area
W2C, and W2C0.85, and obvious peaks of Pt. The initial electrochemi- carbon using three different routes: aqueous tungstate dispersion on
cally active surface areas (ESA) of the Pt/C and Pt/WxCy catalysts were carbon, an incipient wetness technique to disperse tungstate on
16.1 and 10.2 m2 g − 1, respectively. This difference has to be ascribed carbon, and DC magnetron sputtering of tungsten on carbon. The
to the higher surface area of Vulcan XC-72 than that of the calcined synthesis of high surface area WC using the aqueous reduction
tungsten carbide. The ESA loss of the Pt/WxCy during the stability test technique led to a low yield (10%) of W deposition, and the incipient
in 0.5 M H2SO4 at various potentials was much lower than that of Pt/C. wetness technique led to the formation of dense WC particles with
Furthermore, the Pt/C and Pt/WxCy catalysts were applied to single agglomerates of 5–9 μm. DC magnetron sputtering of W onto C,
cells both as anode and cathode materials to compare their instead, proved to be a promising technique for high surface area WC
performance degradation. Before stability test, the PEMFC cell with synthesis, leading to porous agglomerates that were not as dense as
Pt/C better performed than that with Pt/WxCy. After stability test, those obtained by either of the other two techniques. Ganesan and Lee
instead, the single cell prepared with the Pt/WxCy catalyst showed a [128] synthesized W2C microspheres by heating mixtures of resorci-
better performance than the cell with Pt/C, indicative of a high nol-formaldehyde polymer and ammonium metatungstate salt
oxidation resistance of the Pt/WxCy catalyst. Chhina et al. [125,126] (AMT). XRD analysis of the microsphere indicated W2C as the major
760 E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763

phase, and WC and WC1−x as minor phases. The BET surface area of the Specific activities, calculated dividing mass activity by EAS (CO) were
W2C microspheres was about 176 m2 g − 1 (compared to 635 m2 g − 1 for 188 and 144 mA m − 2 for the Pt/WC and Pt/C catalysts, respectively.
carbon microspheres). Platinum particles were supported on these W2C Hara et al. [130] prepared a WC promoted with a small amount of
and Pt–Ru particles on carbon microspheres by the conventional Pt metal by a wet impregnation method. Pt metal loadings were fixed
borohydride reduction method. The crystallite sizes from XRD were at 10 wt.%. The catalyst was prepared using a WC (surface area of
6 nm for 7.5 wt.% Pt on W2C, 14 nm for 15 wt.% Pt on W2C, and 14 nm 25.6 m2 g − 1) support, obtained by carburization of W2N, which was
for 20 wt.% PtRu on carbon microspheres. By CV measurements, the synthesized from WO3. Pt particles were fairly aggregate with around
electrochemical surface area of platinum supported on W2C was 20 nm in size probably due to the low surface area of WC. The mass
several-times higher than those of Pt–Ru supported on carbon activity of Pt/WC was lower than that of a commercial Pt/C (E-TEK), due
microspheres and commercial (E-TEK) 20 wt.% PtRu(1:1)/C carbon. to its lower ESA. The specific activity of Pt/WC (about 10 A m − 2), instead,
This result indicates that the platinum particles were better dispersed was higher than that of the commercial Pt/C catalyst (2.5 A m − 2). The
on the W2C spheres than on the other supports. The electrochemical same research group [131] prepared new WCs with high surface area by
activity towards methanol oxidation of both 7.5 wt.% and 15 wt.% Pt the carburizing of W2N and WS2 precursors, which were prepared from
supported on W2C microspheres was higher than that of PtRu starting materials other than WO3, such as W(CO)6, WCl6 and the like. In
supported on carbon microspheres and the PtRu/C catalyst by E-TEK. particular, the BET surface areas of WCs derived from W(CO)6/S and
The mass activity of platinum supported on W2C microspheres (NH4)2WS4 were around 80 m2 g − 1, which is the highest value among
(7.5 wt.% Pt) was higher by factors of 2.6 and 2.4 than those of PtRu on α-WCs reported. This is probably due to a unique structural change
carbon microspheres and the commercial PtRu catalyst, respectively. forming small crystalline WC particles with a plate-like structure in the
These results clearly showed that Pt supported on W2C microspheres carburization of WS2 and W2N. WC supported Pt metal was prepared by a
provides much better utilization of platinum than in the PtRu/C wet impregnation method and was evaluated as an anode catalyst for the
catalysts and that ruthenium could be entirely replaced. According to electrochemical hydrogen oxidation reaction. The activity of a binary Pt–
the authors, there are several reasons why Pt/W2C-microsphere WC catalyst was greatly improved by choosing a favorable WC and by
catalyst performs better than the commercial catalyst even with the modifying the preparative method. As a result, it turned out that the
smaller platinum loadings and in the absence of ruthenium. The first mass activity of the Pt–WC catalyst system was superior to that of the
is the ability of W2C to stabilize the high dispersion of platinum commercial Pt/C catalyst. The interaction of Pt metal with WC is
relative to carbon supports as discussed above. Second, in the concluded to lead to remarkable improvement in terms of activity, and
presence of platinum, W2C is active in the electrochemical methanol there is a strong synergistic effect between them. The following
oxidation and water decomposition. The formation of hydroxy mechanism of H2 oxidation on Pt/WC was postulated. The hydrogen
groups from water activation is essential for the removal of metal- adsorption at WC is slower and the exchange current density is smaller as
poisoning CO from the surface. This is the main role played by compared to hydrogen at Pt. He concluded that the reaction H2ad → 2Had
ruthenium in conventional Pt–Ru catalysts and this role can be played is the rate-determining step for the adsorption of H2 at the WC anode. On
by W2C. Finally, W2C showed high CO resistance. The CO desorption the other hand, Pt metal is well known to readily dissociate the H2
temperature for pure platinum is 187 °C. It decreases to 147 °C, when molecule to Had which diffuse into the catalyst surface as a spillover
the platinum is supported on W2C. This feature would decrease the phenomenon. Therefore, it is postulated that Pt plays the role of
poisoning of the catalyst surface by CO and increase the activity for accelerating the process of dissociative adsorption of H2, the rate-
electrooxidation of methanol. The stability of the Pt/W2C-micro- determining step for the WC, and WC could take over the rest of steps in
sphere catalyst was tested by repeating electrochemical reaction the process of the hydrogen electro-oxidation.
cycles in a H2SO4/CH3OH solution. The specific activity increased Ganesan et al. [132] and Ham et al. [133] tested mesoporous
initially, but stabilized after around 30 cycles. The authors believe tungsten carbide as a support for Pt catalysts. Ganesan et al. [132]
that surface oxygen species formed on the catalyst are removed synthesized a mesoporous tungsten carbide by using ammonium
during this transient period. Once a steady state is established, there metatungstate as a tungsten precursor and resorcinol-formaldehyde
was no sign of deactivation during 100 consecutive reaction cycles. polymer as a carbon source in the presence of a surfactant. They
Jeon et al. [129] synthesized a WC supported platinum catalyst prepared two-types of WC powders by controlling the amount of
(20 wt.% Pt) for methanol oxidation by a conventional impregnation mesopores. The synthesis of both WC samples involved polyconden-
method and compared its MOR activity with that of a commercial (E- sation of resorcinol–formaldehyde in the presence of AMT, followed
TEK) Pt/C. Commercial WC powder was used as a support. Average by heat treatment at 900 °C. A difference in the preparation methods
particle sizes of Pt in Pt/WC and Pt/C were 7.5 and 2.9 nm, for the two WC samples was the presence or absence of a surfactant
respectively. Pt particles were dispersed on WC surface uniformly, (CTABr) during the polycondensation step. The preparation with
however, most of the WC surface was covered with Pt particles despite CTABr gave WC with ample mesoporosity (denoted as meso-WC)
the small mass ratio of Pt. This result was caused by the spherical whereas that without CTABr yielded WC nanoparticles with a small
structure and larger density of WC. The electrochemically active pore volume (denoted as part-WC). Average pore sizes were similar,
surface areas (EAS) calculated from the H+ desorption were 11.2 and 4.6 nm for part-WC and 4.3 nm for meso-WC. The BET surface areas
5.74 m2 g − 1 for the Pt/WC and Pt/C catalysts, respectively. EAS of Pt/ were also similar, 95 m2 g− 1 for part-WC and 76 m2 g− 1 for meso-WC.
WC was about twice that of Pt/C although the particle size of Pt in Pt/ The most significant difference between two samples was the pore
WC was 2.6 times of that Pt/C. EAS calculated from CO stripping were volume. The part-WC showed a very small volume of 0.086 cm3 g – 1,
4.42 and 6.30 m2 g − 1 for the Pt/WC and Pt/C catalysts, respectively. while that of meso-WC was ca. 3-fold higher at 0.24 cm3 g − 1. Thus the
In the case of Pt/C, the difference between EAS (H+) and EAS (CO) was surfactant CTABr introduced during the polycondensation reaction
within 10%, while EAS (H+) was 2.5 times of EAS (CO) in the case of increased the pore volume of the sample. The XRD pattern of the
Pt/WC. The large EAS (H+) of Pt/WC was explained by the spill-over meso-WC sample showed the presence of WC as the major phase, and
phenomenon of H+ from Pt to WC. Spill-over of H+ creates fresh Pt minor amounts of W2C and WC1−x. The HRTEM images of meso-WC
surface resulting in large EAS (H+). CO stripping showed another sample are shown in Fig. 7 [132]. The WC samples are a collection of WC
interesting result: the CO electro-oxidation peak potential shifted nanoparticles of ca. 20 nm, which appears highly crystalline as evidenced
from 0.80 V in Pt/C to 0.68 V in Pt/WC. The improved CO oxidation by clean lattice fringes. The spacing of the lattice fringes was about
activity originated from the formation of OH by water discharge on the 0.25 nm (Fig. 7a) corresponding to the interplanar spacing of (100)
WC surface. Mass activities for methanol oxidation at 0.5 V were 831 planes of simple hexagonal WC. Selected area electron diffraction (SAED)
−1
and 910 mA gPt for the Pt/WC and Pt/C catalysts, respectively. patterns also show that WC is well crystallized. Pt particles were
E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763 761

Fig. 7. (a) Lattice TEM image of meso-WC (inset: SAED pattern of WC). (b) HRTEM images of Pt particles in Pt (3.5 wt.%)/meso-WC. Scale bars represent 2 nm (a) and 20 nm (b).
Reproduced from Ref. 132, copyright 2007, with permission from Elsevier.

deposited on these WC materials by the conventional borohydride 4. Conclusions


reduction method. The HRTEM image (Fig. 7b) shows that platinum
particles are finely dispersed on WC materials. The average Pt particle Different metal oxides and carbides have been tested as carbon-
size was ca. 2 nm for 3.5 wt.% Pt and ca. 6 nm for 7.5 wt.% Pt both on part- substitute supports for low-temperature fuel cell catalysts. The
WC and meso-WC. Pt supported on these materials shows higher activity characteristics of these ceramic material and carbon Vulcan XC-72
for electrochemical oxidation of methanol than a commercial PtRu/C supports, and the catalytic activity and stability in fuel cell conditions of
catalyst by a factor of six (per mass of Pt) even without Ru. The relative supported catalysts are summarized in Table 1. Some materials
mesoporosity and phase of WC appear important as Pt/meso-WC showed encouraging results, while others presented unsuitable proper-
performs better than Pt/W2C-microsphere and Pt/part-WC. It is ties to be used as fuel cell catalyst supports. Generally, all the materials
interesting to note that the two less active tungsten carbide catalysts presented high stability in a fuel cell environment and acceptable
possess higher surface areas than that of Pt/meso-WC, but it is mostly electrical conductivity, but most of them possess a low surface area,
due to micropores. Thus for high electrocatalytic activity, the facile mass resulting in a low metal dispersion and, as a consequence, in a poor
transport of hydrated methanol to the reaction sites through mesoporous electrochemical activity. In various studies, to increase metal dispersion,
appear more important than the inaccessible high surface area. Ham et supported catalysts with a low metal/ceramic ratio (b10 wt.%) were
al. [133] prepared a mesoporous tungsten carbide of WC-phase by using prepared. This way to tailor the metal/ceramic composite, however, is
the same method of Ganesan et al. [132], then deposited Pt on the not suitable for use as a fuel cell support. Indeed, a low metal/ceramic
mesoporous WC by the borohydride method (7.5 wt.% Pt). The Pt/WC ratio results in the presence of a large amount of ceramic material in the
catalyst shows not only two times higher mass activity but also much catalyst layer of the electrode, increasing the electrode resistance.
improved resistance to CO poisoning for hydrogen electro-oxidation Regarding the electrochemical activity, ceramic supported catalysts have
than a commercial Pt/C catalyst. The high activity for hydrogen oxidation to be evaluated with care. Most papers are indeed very optimistic
was attributed to the active participation of WC in the electro-oxidation regarding the potential interest of ceramic materials, due to presumable
together with Pt and the high CO tolerance to facile formation of surface high activity of ceramic supported metals. In many works [40,46,50–
hydroxyls on WC that could react away CO strongly adsorbed on Pt. It has 53,67,98,110] comparison of the catalytic activity of oxide/carbide
to be remarked that in both these works using mesoporous carbon the supported metals with that of the commonly used carbon supported
Pt/WC weight ratio was b10 wt.%. A higher Pt loading in Pt/WC should metals was not made, so it is not possible establish the effectiveness of
be desirable to decrease the amount of WC, and, as a consequence, the these materials as carbon-substitutes for fuel cell catalysts. In some cases
mass and the thickness of the catalytic layer of the electrode, at a fixed Pt the authors overvalued their results, for example comparing the activity
content. of oxide supported catalysts with that of the homologous unsupported

Table 1
Characteristics of ceramic material and carbon Vulcan XC-72 supports, and catalytic activity and stability in fuel cell conditions of relative supported catalysts.

Material Specific surface Electrical conductivity Metal dispersion, Catalytic activity of Stability in fuel Refs
area m2 g − 1 S cm − 1 particle size (d) nm supported metal cell conditions
Vulcan XC-72 254 4 High d 2.9 High Low [4]
Ebonex, Ti4O7 1–3 1, 103 Low dTi4O7 10–20 Generally low Conflicting results [40–42,46]
M–TiO2 1–1.4 (high Tsynt) 136 (low Tsynt) 10 − 1–102 High dRuTio2 3.3–5.4 dSbTiO2 3.0 High High [43,45–48]
M–SnO2 ~ 43⁎ (RuSnO2) 10 − 1–103 High dSbSnO2 2.5 dRuSnO2 4 High High [64–66]
100 (SbSnO2)
WO3 18, 86 n.d. Medium/High d 6.5 High High [85–88]
d 3–6 (nanorods)
RuO2·xH2O 125 1 n.d Low n.d. [97,98]
S–ZrO2 80 n.d. n.d. Low High [104]
Sn–In2O3 ~ 22⁎ N103 Low d 13 Low High [113]
WC 1.6 (Tsynt N 900 °C), 30–80 (Tsynt b 900 °C) 105 Low/Medium d 6–20 Low (low SA) High [124–133]
176 (microsphere) d 2–6 (meso-WC) High (high SA)

⁎Calculated from particle size.


762 E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763

catalysts, which obviously have low active surface area and, as a [16] T. Maiyalagan, B. Viswanathan, U.V. Varadaraju, Electrochem. Comm. 7 (2005)
905.
consequence, low mass activity [40,98,110]. Doped TiO2 and SnO2, WO3 [17] A. Kongkanand, S. Kuwubata, G. Girishkumar, P. Kamat, Langmuir 22 (2006) 2392.
and WC are the most promising materials as catalyst supports: the [18] X. Wang, W. Li, Z. Chen, M. Waje, Y. Yan, J. Power Sources 158 (2006) 154.
catalysts supported on these materials presented both high stability and [19] J.W. Long, R.M. Stroud, K.E. Swider-Lyons, D.R. Rolison, J. Phys. Chem. B 104 (2000)
9772.
high activity. In addition, doped TiO2 and SnO2, WO3 and WC exhibited a [20] Q. Lu, B. Yang, L. Zhuang, J. Lu, J. Phys. Chem. B 109 (2005) 1715.
promoting effect on the electrocatalytic activity of Pt or PtM alloys, [21] L. Jang, G. Sun, S. Sun, J. Liu, S. Tang, H. Li, B. Zhou, Q. Xin, Electrochim. Acta 50 (2005)
making the ceramic supported catalyst more active than the homo- 5384.
[22] L. Jang, L. Colmenares, Z. Jusys, G.Q. Sun, R.J. Behm, Electrochim. Acta 53 (2007)
logous carbon supported catalyst. WO3, RuO2·xH2O and S–ZrO2 present
377.
the interesting characteristics of proton conducting materials. This [23] K.-W. Park, K.-S. Ahn, Y.-C. Nah, J.-H. Choi, Y.-E. Sung, J. Phys. Chem. B 107 (2003)
property allows the reduction of Nafion ionomer content in the 4352.
[24] V. Raghuveer, B. Viswanathan, J. Power Sources 144 (2005) 1.
electrode catalyst layer, and increases the platinum utilization. Regard-
[25] U. Diebold, Surf. Sci., Rep. 48 (2003) 53.
ing the nanostructured ceramic oxides, few works have been performed [26] Y. Fovet, J.-L. Gal, F. Toumelin-Chemla, Talanta 53 (2001) 1053.
on catalysts supported on these materials for use in low-temperature [27] M. Gustavsson, H. Ekstrom, P. Hanarp, L. Eurenius, G. Lindbergh, E. Olsson, B.
fuel cells. For this reason, at this time, notwithstanding the encouraging Kasemo, J. Power Sources 163 (2007) 671.
[28] D. Morris, Y. Dou, J. Rebane, C.E.J. Mitchell, R.G. Egdell, Phys. Rev. B 61 (2000)
results, it is a risk to affirm that the electrochemical activity of catalysts 13445.
supported on nanostructured ceramic oxides is higher than that of [29] A.S. Andersson, B. Collen, U. Kuylenstierna, A. Magneli, Acta Chem. Scand. 11 (1957)
carbon supported catalysts. 1641.
[30] M. Marezio, P.D. Dernier, J. Solid State Chem. 3 (1971) 340.
Generally, the electrocatalytic activity of supported Pt and Pt–M [31] R.F. Bartholomew, D.R. Frankl, Phys. Rev. 187 (1969) 828.
alloy catalysts is referred as mass activity (MA) or specific activity [32] N.V. Krstajic, L.M. Vracar, V.R. Radmilovic, S.G. Neophytides, M. Labou, J.M. Jaksic,
(SA), i.e. it is normalized with respect to the mass (MA) or the surface R. Tunold, P. Falaras, M.M. Jaksic, Surf. Sci. 601 (2007) 1949.
[33] J.E. Graves, D. Pletcher, R.L. Clarke, F.C. Walsh, J. Appl. Electrochem. 21 (1991) 848.
area (SA) of the catalyst in the electrode. Regarding the catalysts [34] L. He, H.F. Franzen, D.C. Johnson, J. Appl. Electrochem. 26 (1996) 785.
supported on ceramic materials, their mass activity is generally lower [35] R.R. Miller-Folk, R.E. Noftle, D. Pletcher, J. Electroanal. Chem. 274 (1989) 257.
than that of the same catalysts supported on carbon, due to the lower [36] J.R. Smith, F.C. Walsh, R.L. Clarke, J. Appl. Electrochem. 28 (1998) 1021.
[37] E.E. Farndon, D. Pletcher, A. Saraby-Reintjes, Electrochim. Acta 42 (1997) 1269.
surface area of the ceramic substrates. The specific activity, instead, is
[38] J. Dieckmann, S.H. Langer, Electrochim. Acta 44 (1998) 437.
generally higher than that of the corresponding catalysts supported on [39] E.E. Farndon, D. Pletcher, Electrochim. Acta 42 (1997) 1281.
carbon, due to the synergistic effect between the catalyst and the [40] E. Slavcheva, V. Nikolova, T. Petkova, E. Lefterova, I. Dragieva, T. Vitanov, E.
Budevski, Electrochim. Acta 50 (2005) 5444.
ceramic support. Due to their high specific activity, ceramic materials
[41] T. Ioroi, Z. Siroma, N. Fujiwara, S. Yamazaki, K. Yasuda, Electrochem. Comm. 7 (2005)
show a potential interest as carbon substitute for FC catalysts. The 183.
mass activity, however, has practical implications because the cost of [42] T. Ioroi, H. Senoh, S. Yamazaki, Z. Siroma, N. Fujiwara, K. Yasuda, J. Electrochem.
an electrode depends on the amount of catalyst used. To achieve Soc. 155 (2008) B321.
[43] W. Rudorff, H.-H. Luginsland, Z. Anorg. Allg. Chem. 334 (1964) 125.
economic metal loadings (b1 mg cm − 2) fuel cell electrodes must be [44] J. Arbiol, J. Cerda, G. Dezanneau, A. Cirera, F. Peiro, A. Cornet, J.R. Morante, J. Appl.
prepared from metal nanoparticles (d b 5 nm), and low metal particle Phys. 92 (2002) 853.
size for catalysts with metal/support weight ratio N10 wt.% can be [45] B.L. García, R. Fuentes, J.W. Weidner, Electrochem. Solid State Lett. 10 (2007)
B108.
obtained only using high-surface-area materials as catalysts supports. [46] G.Y. Chen, S.R. Bare, T.E. Mallouk, J. Electrochem. Soc. 149 (2002) A1092.
Thus, to increase the mass activity of supported catalysts, research [47] K.-W. Park, K.-S. Seol, Electrochem. Comm. 9 (2007) 2256.
goals have to be addressed to the achievement of ceramic support [48] O.E. Haas, S.T. Briskeby, O.E. Kongstein, M. Tsypkin, R. Tunold, B.T. Børresen, J. New
Mater. Electrochem. Syst. 11 (2008) 9.
with high surface area. [49] D.V. Bavykin, J.M. Friedrich, F.C. Walsh, Adv. Mater. 18 (2006) 2807.
Summarizing, the results regarding the catalytic activity of some [50] M. Wang, D. Guo, H. Li, J. Solid State Chem. 178 (2005) 1996.
oxide and carbide supported catalysts are very promising, but they are [51] J.M. Macak, P.J. Barczuk, H. Tsuchiya, M.Z. Nowakowska, A. Ghicov, M. Chojak, S.
Bauer, S. Virtanen, P.J. Kulesza, P. Schmuki, Electrochem. Comm. 7 (2005) 1417.
scarce and carried out overall in acidic solution. Further tests, particularly
[52] S.H. Kang, T.-Y. Jeon, H.-S. Kim, Y.-E. Sung, W.H. Smyrl, J. Electrochem. Soc. 155 (208)
in a single fuel cell, have to be performed to confirm the good B1058.
performance of these ceramic materials as a support for fuel cell catalysts. [53] S.H. Kang, Y.-E. Sung, W.H. Smyrl, J. Electrochem. Soc. 155 (2008) B1128.
[54] T. Kasuga, M. Hiramatsu, A. Hoson, T. Sekino, K. Niihara, Langmuir 14 (1998) 3160.
[55] M.M. Gadgil, R. Sasikala, S.K. Kulshreshtha, J. Mol. Cat. 87 (1994) 297.
Acknowledgments [56] D.R. Schryer, B.T. Upchurch, J.D. Van Norman, K.G. Brown, J. Schryer, J. Catal.122 (1990)
193.
[57] K. Sekizawa, H. Widjaja, S. Maeda, Y. Ozawa, K. Eguchi, Catal. Today 59 (2000) 69.
The authors thank the Conselho Nacional de Desenvolvimento [58] D. Amalric-Popescu, F. Bozon-Verduraz, Catal. Lett. 64 (2000) 125.
Científico e Tecnológico (CNPq, Proc. 310151/2008-2) for financial [59] K. Liberková, R. Touroude, J. Mol. Catal. A 180 (2002) 221.
assistance to the project. [60] T. Okanishi, T. Matsui, T. Takeguchi, R. Kikuchi, K. Eguchi, Appl. Catal. A 298 (2006)
181.
[61] Z. Liu, B. Guo, L. Hong, T.H. Lim, Electrochem. Comm. 8 (2006) 83.
References [62] E.E. Kohnke, J. Phys. Chem. Solids 23 (1962) 1557.
[63] I. Saadeddin, B. Pecquenard, J.P. Manaud, R. Decourt, C. Labrugère, T. Buffeteau, G.
[1] P. Costamagna, S. Srinivasan, J. Power Sources 102 (2001) 242. Campet, Appl. Surf. Sci. 253 (2007) 5240.
[2] C. Lamy, A. Lima, V. LeRhun, F. Delime, C. Coutanceau, J.-M. Leger, J. Power Sources [64] A.L. Santos, D. Profeti, P. Olivi, Electrochim. Acta 50 (2005) 2615.
105 (2002) 283. [65] K.-S. Lee, I.-S. Park, Y.-H. Cho, D.-S. Jung, N. Jung, H.-Y. Park, Y.-E. Sung, J. Catal.
[3] E. Antolini, Mater. Chem. Phys. 78 (2003) 563. 258 (2008) 143.
[4] E. Antolini, Appl. Catal. B, in press, doi:10.1016/j.apcatb.2008.09.030. [66] W.J. Zhou, B. Zhou, W.Z. Li, Z.H. Zhou, S.Q. Song, G.Q. Sun, Q. Xin, S. Douvartzides,
[5] A.M. Couper, D. Pletcher, F.C. Walsh, Chem. Rev. 90 (1990) 837. M. Goula, P. Tsiakaras, J. Power Sources 126 (2004) 16.
[6] G.A. Gruver, J. Electrochem. Soc. 125 (1978) 1719. [67] H.L. Pang, X.H. Zhang, X.X. Zhong, B. Liu, X.G. Wei, Y.F. Kuang, J.H. Chen, J. Coll. Int.
[7] P. Stonehart, Carbon 22 (1984) 423. Sci. 319 (2008) 193.
[8] J. McBreen, H. Olender, S. Srinivasan, K. Kordesch, J. Appl. Electrochem.11 (1981) 787. [68] O.V. Safonova, G. Delabouglise, B. Chenevier, A.M. Gaskov, M. Labeau, Mater. Sci.
[9] K.H. Kangasniemi, D.A. Condit, T.D. Jarvi, J. Electrochem. Soc. 151 (2004) E125. Eng. C 21 (2002) 105.
[10] R.L. Borup, J.R. Davey, F.H. Garzon, D.L. Wood, M.A. Inbody, J. Power Sources 163 (2006) [69] A. Hagemeyer, Z. Hogan, M. Schlichter, B. Smaka, G. Streukens, H. Turner, A. Volpe
76. Jr., H. Weinberg, K. Yaccato, Appl. Catal. A 217 (2007) 139.
[11] J. Wang, G. Yin, Y. Shao, S. Zhang, Z. Wang, Y. Gao, J. Power Sources 171 (2007) 331. [70] Z.L. Wang, Adv. Mater. 15 (2003) 432.
[12] A. Taniguchi, T. Akita, K. Yasuda, Y. Miyazaki, J. Power Sources 130 (2004) 42. [71] M.S. Saha, R. Li, M. Cai, X. Sun, Electrochem. Solid State Lett. 10 (2007) B130.
[13] L.M. Roen, C.H. Paik, T.D. Jarvi, Electrochem. Sol. State Lett. 7 (2004) A19. [72] M.S. Saha, R. Li, X. Sun, Electrochem. Comm. 9 (2007) 2229.
[14] K. Kinoshita, Carbon: electrochemical and physicochemical properties, Wiley, [73] M.A. Butler, J. Appl. Phys. 48 (1977) 1914.
New York, 1988, p. 319. [74] H. Nakajima, I. Honma, Solid State Ion. 148 (2002) 607.
[15] M. Pourbaix, Atlas of Electrochemical Equilibria in Aqueous Solution, National [75] H. Chhina, S. Campbell, O. Kesler, J. Electrochem. Soc. 154 (2007) B533.
Association of Corrosion Engineers, Houston, TX, 1979. [76] L.W. Niedrach, L.B. Weinstock, Electrochem. Technol. 3 (1965) 270.
E. Antolini, E.R. Gonzalez / Solid State Ionics 180 (2009) 746–763 763

[77] L.W. Niedrach, H.I. Zeliger, J. Electrochem. Soc. 116 (1969) 152. [107] Y. Zhang, H. Zhang, Y. Zhai, X. Zhu, C. Bi, J. Power Sources 168 (2007) 323.
[78] P.J. Kulesza, L.R. Faulkner, J. Electroanal. Chem. 259 (1989) 81. [108] X. Zhu, H. Zhang, Y. Liang, Y. Zhang, B. Yia, Electrochem. Solid State Lett. 9 (2006)
[79] P.J. Kulesza, L.R. Faulkner, J. Electrochem. Soc. 136 (1989) 707. A49.
[80] A.C.C. Tseung, K.Y. Chen, Catal. Today 38 (1997) 439. [109] L. Wang, D.M. Xing, Y.H. Liu, Y.H. Cai, Z.-G. Shao, Y.F. Zhai, H.X. Zhonga, B.L. Yi, H.M.
[81] F. Maillard, E. Peyrelade, Y. Soldo-Olivier, M. Chatenet, E. Chaînet, R. Faure, Zhang, J. Power Sources 161 (2006) 61.
Electrochim. Acta 52 (2007) 1958. [110] B. Seger, A. Kongkanand, K. Vinodgopal, P.V. Kamat, J. Electroanal. Chem. 621 (2008)
[82] J. Shim, C.-R. Lee, N.-K. Lee, J.-S. Lee, E.J. Cairns, J. Power Sources 102 (2001) 172. 198.
[83] C. He, H.R. Kunz, J.M. Fenton, J. Electrochem. Soc. 150 (2003) A1017. [111] J.C.C. Fan, J.B. Goodenough, J. Appl. Phys. 48 (1977) 3524.
[84] E.J. McLeod, V.I. Birss, Electrochim. Acta 51 (2005) 684. [112] M. Yamaguchi, A. Ide-Ektessabi, H. Nomura, N. Yasui, Thin Solid Films 447-448 (2004)
[85] R. Ganesan, J.S. Lee, J. Power Sources 157 (2006) 217. 115.
[86] X. Cui, J. Shi, H. Chen, L. Zhang, L. Guo, J. Gao, J. Li, J. Phys. Chem. B 112 (2008) 12024. [113] H. Chhina, S. Campbell, O. Kesler, J. Power Sources 161 (2006) 893.
[87] J. Rajeswari, B. Viswanathan, T.K. Varadarajan, Mater. Chem. Phys. 106 (2007) 168. [114] R.B. Levy, M. Boudart, Science 181 (1973) 547.
[88] T. Maiyalagan, B. Viswanathan, J. Power Sources 175 (2008) 789. [115] W.T. Grubb, D.W. McKee, Nature 210 (1966) 192.
[89] B. Reichman, A.J. Bard, J. Electrochem. Soc. 126 (1979) 583. [116] D.W. McKee, A.J. Scarpellino, I.F. Danzig, M.S. Pak, J. Electrochem. Soc. 116 (1969)
[90] G.R. Patzke, F. Krumeich, R. Nesper, Angew. Chem. Int. Ed. 41 (2002) 2446. 562.
[91] J.B. Goodenough, Prog. Solid State Chem. 5 (1971) 145. [117] A. Honji, T. Marl, Y. Hishinuma, J. Electrochem. Soc. 135 (1988) 917.
[92] K.E. Swider, C.I. Merzbacher, P.L. Hagans, D.R. Rolison, Chem. Mater. 9 (1997) 1248. [118] V. Jalan, D.G. Frost, US Patent 4795684 (1989).
[93] S. Trasatti, Electrochim. Acta 36 (1991) 225. [119] M.B. Zellner, J.G. Chen, Catal. Today 99 (2005) 299.
[94] J.P. Zheng, P.J. Cygan, T.R. Jow, J. Electrochem. Soc. 142 (1995) 2699. [120] H. Meng, P.K. Shen, J. Phys. Chem. B 1098 (2005) 22705.
[95] J.M. Fletcher, W.E. Gardner, B.F. Greenfield, M.J. Holdoway, M.H. Rand, J. Chem. [121] R. Venkataraman, H.R. Kunz, J.M. Fenton, J. Electrochem. Soc. 150 (2003) A278.
Soc. A 3 (1968) 653. [122] L.G.R.A. Santos, K.S. Freitas, E.A. Ticianelli, J. Solid State Electrochem. 11 (2007) 1541.
[96] D.A. McKeown, P.L. Hagans, L.P.L. Carette, A.E. Russell, K.E. Swider, D.R. Rolison, J. [123] M.K. Jeon, K.R. Lee, W.S. Lee, H. Daimon, A. Nakahara, S.I. Woo, J. Power Sources
Phys. Chem. B 103 (1999) 4825. 185 (2008) 927.
[97] K. Lasch, G. Hayn, L. Jorissen, J. Garche, O. Besenhardt, J. Power Sources 105 (2002) [124] S. Zhang, H. Zhu, H. Yu, J. Hou, B. Yi, P. Ming, Chin. J. Catal. 28 (2007) 109.
305. [125] H. Chhina, S. Campbell, O. Kesler, J. Power Sources 164 (2007) 431.
[98] Z. Chen, X. Qiu, B. Lu, S. Zhang, W. Zhu, L. Chen, Electrochem. Comm. 7 (2005) 593. [126] H. Chhina, S. Campbell, O. Kesler, J. Power Sources 179 (2008) 50.
[99] H.A. Gasteiger, N. Markovic, P.N. Ross, E.J. Cairns, J. Electrochem. Soc. 141 (1994) [127] R. Koc, S.K. Kodambaka, J. Eur. Ceram. Soc. 20 (2000) 1859.
1795. [128] R. Ganesan, J.S. Lee, Angew. Chem. Int. Ed. 44 (2005) 6557.
[100] M.S. Hino, S. Kobayashi, K. Arata, J. Am. Chem. Soc. 101 (1979) 6439. [129] M.K. Jeon, H. Daimon, K.R. Lee, A. Nakahara, S.I. Woo, Electrochem. Comm. 9 (2007)
[101] G.D. Yadav, J.J. Nair, Microp. Mesop. Mater. 33 (1999) 1. 2692.
[102] M. Waqif, J. Bachelier, O. Saur, J.-C. Lavalley, J. Mol. Cat. 72 (1992) 127. [130] Y. Hara, N. Minami, H. Itagaki, Appl. Catal. A 323 (2007) 86.
[103] S. Hara, M. Miyayama, Solid State Ion. 168 (2004) 111. [131] Y. Hara, N. Minami, H. Matsumoto, H. Itagaki, Appl. Catal. A 332 (2007) 289.
[104] Y. Suzuki, A. Ishihara, S. Mitsushima, N. Kamiya, K. Ota, Electrochem. Solid State [132] R. Ganesan, D.J. Ham, J.S. Lee, Electrochem. Comm. 9 (2007) 2576.
Lett. 10 (2007) B105. [133] D.J. Ham, Y.K. Kim, S.H. Han, J.S. Lee, Catal. Today 132 (2008) 117.
[105] T.A. Zawodzinski, T.E. Springer, F. Uribe, S. Gottesfeld, Solid State Ion. 60 (1993) 199.
[106] H. Uchida, Y. Ueno, H. Hagihara, M. Watanabe, J. Electrochem. Soc. 150 (2003) A57.

You might also like