Você está na página 1de 9

Journal of The Electrochemical Society, 162 (14) H929-H937 (2015)

0013-4651/2015/162(14)/H929/9/$33.00 The Electrochemical Society

H929

Electrochemical and Physicochemical Characterizations of


Gold-Based Nanomaterials: Correlation between Surface
Composition and Electrocatalytic Activity
Yaovi Holade, Karine Servat, Julie Rousseau, Christine Canaff, Suzie Poulin,
Teko W. Napporn, and K. Boniface Kokoh,z

Universite de Poitiers, IC2MP, UMR 7285 - CNRS, Equipe


SAMCat, 4, rue Michel Brunet, B27, TSA 51106,
86073 Poitiers cedex 09, France
Seminal papers have highlighted the promoting effect of gold when associated with other metals. There is still a lack of information
about the origin of the exceptional electrochemical performances of gold-based nanostructures, known so far as one of the most active
electrocatalysts for glucose-based energy conversion devices. In this report, we examined the correlation between the electrocatalytic
properties and the surface composition on a line-up of Au-based nanostructures: gold-palladium, gold-platinum binaries, and goldpalladium-platinum ternaries from the so-called Bromide Anion Exchange (BAE) method. BAE enables to obtain carbon supported
nanocrystals having high Miller indices. These truncated octahedron nanocrystals present twins as well as (110) facets. Both
X-ray photoelectron spectroscopy (XPS) and electrochemical characterizations have shown that the surface of the multimetallic
nanomaterials contains less atoms of gold. The most exposed platinum and palladium enable improving glucose electrooxidation
reaction kinetics at low electrode potentials. Additionally, XPS measurements have shown that the binding energy (BE) of either Pd
or Pt shifts towards low values when associated with Au, indicating strong electronic interactions between the different elements
in the multimetallic nanomaterials. These properties have led to improved catalytic performances when tested for CO stripping
experiments and glucose (potential fuel) electrooxidation reaction in alkaline medium.
2015 The Electrochemical Society. [DOI: 10.1149/2.0601514jes] All rights reserved.
Manuscript submitted August 24, 2015; revised manuscript received September 16, 2015. Published October 13, 2015.

The design of cutting-edge nanomaterials to be used as advanced


electrode materials in energy conversion and storage technologies has
focused the attention of researchers in the last decade. For a metal having a face-centered cubic (fcc) crystallographic symmetry (Au, Pt, Pd,
Ag. . . ), the equilibrium shape of its nanocrystal is predicted by fundamental thermodynamic considerations to be a cubo-octahedral.14
Thus, it is exclusively composed of (111) and (100) facets. Most of
the prepared supported metal catalysts have this morphology. Various crystalline structures with high Miller indices (hkl) have been
reported only for unsupported nanoparticles.57 Zhang and coworkers
reported the preparation of 24 high-index (720) facets of Au concave
nanocubes using a combination of silver and chloride ions through a
seed-mediated synthetic method.8 The group of Xia has shown the possibility of obtaining nanocrystals Pt(510), Pt(720) and Pt(830) using
a kinetically-controlled method with tunable electrocatalytic properties towards the well-known oxygen reduction reaction (ORR).5,9 By
using a simple route based on seed growth approach, they obtained
Pd concave nanocubes bounded by high index Pd(730) facets, where
Pd nanocubes were used as seeds for the reduction of a Pd precursor in aqueous solution.10 The majority of these investigations has
concerned monometallic materials even if some endeavors have been
undertaken for Au-based core-shell bimetallic materials having high
index (hkl) facets.11,12
The unsupported nanoparticles serve often as models in catalysis.
For the design of realistic applications like fuel cells (FCs) technologies, an electrical conducting support is necessary to first disperse
these nanomaterials and then, to enhance the current densities, while
reducing concomitantly the metal content in the electrode catalyst.
Unfortunately, this support gives rise to an additional constraint. It
restricts the controllability of nanoparticles shape and finally does
not enable obtaining nanoparticles with desired crystallographic orientations. Highly dispersed catalysts provide a maximum available
surface area for electrocatalytic reactions. On the other hand, the decrease of particles size especially in bimetallic materials down to few
nanometers is undoubtedly accompanied by variation of numerous
physicochemical properties, thus affecting the catalytic properties.24
It should be noted that the entire reactions in heterogeneous catalysis are not size-depending. But, it is unanimously accepted that the
Electrochemical Society Student Member.
Electrochemical Society Active Member.
z
E-mail: boniface.kokoh@univ-poitiers.fr

presence of different elements in a nanomaterial composition has notable effect on their electronic properties. Seminal works have shown
that bulk gold-based electrode materials exhibit good electrocatalytic
activity towards carbohydrates electrooxidation.13,14 The effective development of efficient carbohydrate-based direct alkaline fuel cells
(DAFCs) offers many benefits among of which the electrochemical synthesis of added-value chemical from selective electrooxidation of carbohydrates (abundant, renewable and non-toxic organic
compounds from biomass, an extensive and endlessly renewable resource). Thus, various synthetic methods have emerged to prepare
carbon-supported Au, Pt and Pd nanoparticles for their electrooxidation, especially glucose. Fundamental understanding of the correlation
between their surface properties and electrocatalytic ones for the optimization of performances in DAFCs is still missing.
In previous reports, the so-called Bromide Anion Exchange,
BAE15 method has been revised to prepare various compositions
of Au-Pd, Au-Pt and Au-Pt-Pd nanomaterials supported on Vulcan
XC 72R carbon.16,17 Their good electrocatalytic properties have been
pointed out, particularly as anode catalysts for glucose electrooxidation in hybrid biofuel cell either in conditions mimicking physiological
ones18,19 or human serum.20 The most fascinating aspect of this simple, fast and convenient synthetic method is an approach free from
organic molecule as surfactant or capping agent. The present work
aims at scrutinizing the direct correlation between the physicochemical and electrochemical properties of these nanomaterials. Herein,
nanostructures were first characterized by high-resolution transmission electron microscopy (HRTEM) together with energy dispersive
X-ray (EDX). Then, their surface and electronic properties were examined by X-ray photoelectron spectroscopy (XPS) measurements.
Carbon monoxide (CO) and glucose were used as surface probing
molecule and fuel, respectively, to evaluate the catalytic activity of
these electrode materials.
Experimental
Preparation of the nanomaterials from a surfactant-free method:
Bromide Anion Exchange (BAE). Multimetallic nanostructures AuPt, Au-Pd and Au-Pt-Pd were synthesized from the revised Bromide Anion Exchange, BAE method.16 The entire optimization
is reported elsewhere.17 The singularity of this approach relies on
its softness and simplicity of implementation without using an organic molecule as surfactant or capping agent. Conveniently, BAE
allows an effective preparation of well-dispersed nanoparticles with a

Downloaded on 2015-10-29 to IP 131.111.164.128 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

H930

Journal of The Electrochemical Society, 162 (14) H929-H937 (2015)

good chemical yield (>90%), exhibiting high electrochemical active


surface area and excellent catalytic properties. For catalysts preparation, hexachloroplatinic (IV) acid hexahydrate (H2 PtCl6 6H2 O,
ACS reagent, 37.50% Pt basis), tetrachloroauric (III) acid trihydrate
(HAuCl4 3H2 O, ACS reagent, 99.9%), potassium tetrachloropalladate (II) (K2 PdCl4 , 99%), sodium borohydride (NaBH4 , 99%), potassium bromide (KBr, 99%) and L-ascorbic acid (AA, 99%) were
purchased from Sigma-Aldrich and used as-received without further
purification. Contrariwise, Vulcan XC 72R carbon black used as support for nanoparticules dispersion was supplied from Cabot and thermally pre-treated in order to boost the electrocatalytic properties of
the nanoparticles as recently highlighted.21 The reducing agent NaBH4
was used for gold-platinum nanomaterials (Au-Pt/C). To avoid hydrogen (present in NaBH4 ) insertion into the palladium lattice network,22
AA was used for materials containing palladium (Au-Pd/C or Au-PtPd/C). The targeted metal loading was 20 wt.%. All the solutions were
prepared with Millipore Milli-Q water (MQ, 18.2 M cm at 20 C).
Typically, Aux Pty Ptz nanoparticles supported on Vulcan XC 72R
(Aux Pty Ptz /C, where x, y and z are the corresponding atomic
percentages) were synthesized using HAuCl4 3H2 O (for Au),
H2 PtCl6 6H2 O (for Pt) and K2 PdCl4 (for Pd) as metal precursors.
These metal salts, taken in the proportion corresponding to the desired atomic composition of the nanoparticles and total metal weight
of 20 mg are dissolved in 100 mL MQ water (thermostated at 25 C,
using a magnetic stirrer). Then, KBr was added to the solution (
= n(KBr) /n(metals) = 1.5) under vigorous stirring. After complete
solution homogenization, 80 mg Vulcan was added under constant
ultrasonic homogenization for 45 min. Then, a freshly prepared reducing agent solution was added dropwise to reduce the Au3+ , Pt4+
and Pd2+ species to their metallic state. As aforementioned, either
NaBH4 solution (0.1 mol L1 , n(NaBH4) /n(metals) = 15) or AA (0.1
mol L1 , n(AA) /n(metals) = 7) was used as reducing agent. The reaction continued for 2 hours at 40 C under vigorous stirring. Finally,
the obtained metal nanoparticles supported on Vulcan (Aux Pty Ptz /C)
were filtered on Buchner system, washed roughly three times with
MQ water and dried in an oven at 40 C for 12 hours.
Methods for physicochemical characterizations of the
nanomaterials. The metal loading was determined from differential and thermogravimetric analyses (DTA-TGA). Experiments
were conducted on TA Instruments SDT Q600 apparatus by thermally
heating a few milligrams of the sample (contained in an alumina
crucible) under air flow of 100 mL min1 from 25 to 900 C with a
5 C min1 temperature rate. The elementary analysis of the samples
was carried out by inductively coupled plasma optical emission
spectrometry (ICP-OES) using a spectrometer Optima 2000 DV
provided by Perkin Elmer after their mineralization in a reactor
containing 10 mg of the catalyst power dissolved in aqua regia. The
morphologies, particles dispersion on the support and the size of
the obtained materials were analyzed using transmission electron
microscopy (TEM) on a TEM/STEM JEOL 2100 UHR (200 kV)
equipped with a LaB6 filament while energy dispersive X-ray
(EDX, JED Series AnalysisProgram provided by JEOL) allowed to
determine their chemical elementary composition and homogeneity.
High-resolution (HRTEM) analyses were conducted on the same
microscope. The nanoparticules support, Vulcan XC 72R carbon
being electronically conducting, materials are free of nuisances like
charging.4 Thus, samples for TEM observations were easily prepared
by their dispersion in ethanol and then deposited onto copper grid. To
probe and characterize the oxidation state of the surface and electronic
interactions between the elements in the different as-synthesized
nanomaterials, X-ray photoelectron spectroscopy (XPS) was employed. Its suitability to study electrocatalysts has been reviewed by
Corcoran.23 XPS measurements were performed in a high vacuum
chamber (pressure 2 109 Torr) on a Kratos Axis Ultra DLD
spectrometer equipped with a monochromatic radiation source Al
Mono (AlK : 1486.6 eV) operating at 150 W (15 kV and 10 mA). The
survey spectra were performed with a step of 1 eV (transition energy:
160 eV). Based on the collected basic information, high-resolution

XPS spectra were collected at a step of 0.1 eV (transition energy: 20


eV). XPS peaks were fitted using Gaussian-Lorentzian (GL) profile
function and asymmetry determined from pure metal which was
obtained after reduction under hydrogen gas atmosphere. It should be
noted that a systematic subtraction of the background noise preceded
the normalization of the spectra and the measurement of binding
energies (BE) was corrected based on the energy of C 1s at 284.4 eV
(Vulcan XC 72R carbon). The Fermi level was also probed and XPS
spectra presented herein use C1s as reference. In addition, reference
materials (bulk Pt, Pd, Au) were also used to better analyze the XPS
spectra of the prepared nanomaterials. It should be noticed that the
Wagners table for the sensitivity factors was used for quantifications
and the CASA XPS software for data acquisition and processing.
Procedures for the electrochemical characterizations and tests.
Electrochemical experiments were conducted in a conventional threeelectrode cell using an analogical potentiostat EG&G PARC Model
362 from Princeton Applied Research. The reference electrode consists of a fresh home-made reversible hydrogen electrode (RHE), separated from the solution by a Luggin capillary tip. The working electrode consisted of a catalytic ink (3 L) deposited onto a well-polished
glassy carbon (GC) disk of 3 mm diameter (geometrical surface area:
0.071 cm2 ) through an abrasive ADL disk with alumina powders of 1,
0.3 and 0.05 m. A slab of GC (6.48 cm2 geometrical area) was used
as counter electrode. The catalytic ink was prepared as previously
reported16,17,21,22 and according to the procedure initiated by Wilson
and co-workers for FCs applications.24 The total metal loading in the
catalytic layer on the electrode is ca. 80 gmetal cm2 . Cyclic voltammograms (CVs) were recorded in a 0.1 M alkaline solution (NaOH,
97%; from Sigma-Aldrich) used as supporting electrolyte. Prior to
any test, the solution was completely deoxygenated with nitrogen
for 30 min. Nanocatalyst surfaces were probed by carbon monoxide
(CO, Air Liquide, from ultra-pure) stripping experiments. Basically,
to make CO stripping experiment, there are 3 steps: (i) Adsorption
of CO for about 5 min on the electrode at a fixed potential (at ECO
= 0.10 V vs. RHE for Au-Pt/C electrode materials and at 0.3 V vs.
RHE for those containing Pd, for avoiding hydrogen absorption at
low electrode potential); (ii) Under ECO potential control, nitrogen is
bubbled for roughly 30 min in order to remove completely the nonadsorbed CO in bulk solution; (iii) Finally, the cyclic voltammetry
is performed starting at ECO . In this case, only the remaining CO,
which is adsorbed at the electrode surface, is electrooxidized during
the positive scan. In the reverse scan, there is no reactive species which
will be oxidized again. To investigate the electrocatalytic activity of
the electrode materials, 10 mM glucose (D-(+)-glucose, 99.5%; from
Sigma-Aldrich) was used as fuel. All electrochemical measurements
were conducted at controlled temperature of 25 C.
Results and Discussion
The as-synthesized nanomaterials were successively characterized
by TGA, ICP and XRD. The targeted metal loading of 20 wt.% was
confirmed for all materials. Furthermore, the nominal composition in
multimetallic materials is verified from ICP analyses: Au90 Pd10 (for
Au90 Pd10 ), Au79 Pd21 (for Au80 Pd20 ), Au74 Pt26 (for Au80 Pt20 ), Au58 Pt42
(for Au60 Pt40 ), Au74 Pt10 Pd16 (for Au70 Pt15 Pd15 ), Au60 Pt16 Pd21 (for
Au60 Pt20 Pd20 ).
Nanomaterials characterization by high-resolution transmission
electron microscopy (HRTEM). Prior to HRTEM analyses, nanomaterials were examined in low resolution (TEM) in order to establish
histograms of particles size distribution as well as the determination of
the catalysts dispersion (meaning the exposed fraction of atoms).2527
Fig. 1 shows the TEM micrographs of the monometallic and multimetallic nanomaterials. Overall, particles are well dispersed on the
support with mean particles size between 3 and 5 nm, except for Pd
(roughly 9 nm). This size deviation is only due to the reducing agent
used for Pd since ascorbic acid promoted the formation of bigger
particles than NaBH4 .22 The obtained catalyst dispersion goes from

Downloaded on 2015-10-29 to IP 131.111.164.128 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Journal of The Electrochemical Society, 162 (14) H929-H937 (2015)

H931

Figure 1. TEM images of: (a) Au/C, (b) Pt/C, (c) Pd/C, (d) Au90 Pd10 /C, (e) Au80 Pt20 /C and (d) Au70 Pt15 Pd15 /C nanomaterials.

10 to 30%, which is in good agreement with the reported values for


electrocatalysts.26,27
HRTEM observations and analyses of the monometallic
nanomaterials. We next undertook a high-resolution TEM study
(HRTEM) to access the morphology as well as the crystallographic
orientation of the nanoparticles. Figs. 2a and 2b show HRTEM images of Au/C where nanoparticles tend to adopt a truncated octahedron
shape with different degrees of truncation. Nanoparticles have facets
oriented by crystallographic planes (111) and (200) in the case of Pt/C
(Fig. 2c) and Pd/C (Figs. 2d and 2e). Obtaining high Miller indices
such as (200) instead of (100) suggests that the BAE synthesis method
offers favorable thermodynamic conditions. Contrariwise, within the
same material, different forms may coexist as it is the case of gold
with two different forms thermodynamically. In Fig. 2a, the nanoparticle is a truncated octahedron with facets composed of (111) and
(200) planes while the other in Fig. 2b is a pseudo-cuboctahedron
whose facets are determined by the planes (100) and (110). Indeed,
a real cubo-octahedron has the facets (100) and (111).24 The change
in the crystallographic orientation could be due to the presence of
both chloride bromide ions. In addition, the particle in Fig. 2b has
a quasi-symmetry element which is a mirror plane as indicated by
the circles in the picture. Bromide ions alone mediate the formation
of nanocubes, meaning facets (100). Since the determined interplanar spacing of crystallographic plane is d(hkl) = 0.29, instead of d(hkl)
= 0.39 for (100), we concluded that the final shape was modified by
the presence of both ions.
Considering supported nanoparticles, the formation of facets (111)
and (100) is thermodynamically more favorable since the surface energy () associated with different crystallographic planes follows the
order: (111) < (100) < (110).3 Basically, under thermodynamic
equilibrium conditions, the shape of a crystal is predicted by the
Wulffs theorem, which states that the minimum energy is obtained for
a polyhedron whose center distances to the surfaces are proportional
to their surface energies.1,4,28 Thus, the polyhedron corresponding to
the more thermodynamic stable morphology for a nanoparticle with

a fcc crystal symmetry is a truncated octahedron.1,3,4 The latter typestructure is composed of eight hexagonal faces and six square faces
where the crystalline orientation of the surface atoms is described by
Miller indices (111) and (100), respectively.
Furthermore, the presence of the carbon support undoubtedly influences the final shape of the nanoparticle. The crystal shape shown in
Figs. 2b and 2e which cannot be rigorously predicted by Wulffs theorem indicates that the BAE method offers better conditions than other
methods like the well-known water-in-oil one.29 It has been shown that
for the glucose electrooxidation reaction, the activity of nanoparticles
increases in the direction (111) < (110) < (100).30,31 Furthermore, the
presence of pseudo-cuboctahedron shape and particles with exposed
facets (200) could lead to specific catalytic performances.
HRTEM observations and EDX analyses of multimetallic
nanomaterials. HRTEM observations coupled with EDX analyses
were made on multimetallic catalysts to identifying possible preferred
orientations of nanoparticles while analyzing the homogeneity of catalysts. For the atomic quantification, energy levels of 9.712 keV (for
Au), 9.441 keV (for Pt) and 2.852 keV (for Pd) were considered. The
signal of carbon (ca. 0.25 keV) is due to Vulcan used as nanoparticles support and that of the copper (ca. 0.95 keV, 8.15 and 8.95 keV)
belongs to the grid used for microscopic observations.
HRTEM photographs in Fig. 3a show a particular shape of Au80 Pt20
nanoparticle with crystallographic facets (111), (200) and finally the
facet (110). A closer examination reveals the presence of a mirror
plane, which induces the formation of a twin (top micrograph of
Fig. 3a). The occurrence of twins in a crystal results from pooling two
grains along a crystallographic plane. This has the effect of forming a
nanocrystal made of two half-crystals, their structure being the mirror
reflection of each other by the seal twin. Recently, same phenomena
have been observed in the case of PtCo and PtNi nanoparticles: several rows of atoms along the junction with simulation experiments.32
Habrioux et al.29 had already observed the twins presence in the
nanocrystals Au70 Pt30 /C prepared by water-in-oil microemulsion. But
the shape shown in Fig. 3a has not yet been obtained experimentally

Downloaded on 2015-10-29 to IP 131.111.164.128 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

H932

Journal of The Electrochemical Society, 162 (14) H929-H937 (2015)

Figure 3. EDX spectra and HRTEM micrographs of (a) Au80 Pt20 /C, (b)
Au90 Pd10 /C and (c) Au70 Pt15 Pd15 /C multimetallic nanomaterials.
Figure 2. HRTEM micrographs of (a, b) Au/C, (c) Pt/C and (d, e) Pd/C
monometallic nanomaterials. The insert pictures in (a), (c) and (d) are related
to the Fourier transform of the corresponding images for the determination
of crystallographic facets. The zone axis [hkl] during the measurement is
indicated.

with supported metal nanoparticles. It constitutes an advance towards


shape-controlled supported nanoparticles preparation. This nanoparticle has large twins with lots of atomic rows supported by the
(110) plane. The formation of well-defined twins has been also observed with monometallic Au/C nanomaterials synthesized with Lascorbic acid as reducing agent (not shown herein). Furthermore,
the HRTEM image of Au90 Pd10 /C (Fig. 3b) shows a shape close to
that of an icosahedron (three-fold symmetry)2,33 or decahedron(fivefold symmetry),2,34 both having been obtained in the case of gold
nanomaterials using ascorbic acid as reducing agent and some capping
agents. The nanoparticle displays the index (111) and (200) facets, as
those of the aforementioned crystal morphology. The HRTEM image
of Au70 Pt15 Pd15 /C depicts the same facets (111) and (200). The origin
of the formation of the high Miller indices (200) instead of (100) ones
could be due to the fact that no organic molecule was used during the
synthesis and certainly the presence of both chloride and bromide ions
as previously explained. So, the particles can grow with any constraints
from organic molecules. Due to the different adsorption kinetics of
halides, nanocrystals with high Miller indices can be formed.
The EDX analyses have given a consistent composition for
Au78 Pt22 (Au80 Pt20 /C). In contrast, the composition Au90 Pd10 /C and
those of trimetallic catalysts are heterogeneous (trimetallics having
some gold islands). In the case of Au90 Pd10 /C, different compositions
as Au94 Pd6 and Au86 Pd14 were observed. For trimetallic catalysts,
the compositions with the three metals are Au48 Pt17 Pd35 (+ gold islands) for Au70 Pt15 Pd15 /C and Au30 Pt16 Pd54 for Au60 Pt20 Pd20 /C (+

gold islands). The more heterogeneous materials are those where Lascorbic acid was used as reducing agent. This heterogeneity could be
explained by a thermodynamic approach than that involving the metal
salts reduction kinetics, when L-ascorbic acid is used. Indeed, the reduction of Au3+ into Au+ is very fast; but kinetics of the second step
from Au+ to Au0 as well as the reduction kinetics of Pt4+ and Pd2+
species into Pt0 and Pd0 depend on the experimental conditions. The
presence of other metal species as Pt4+ and Pd2+ could substantially
catalyze the last reduction of Au+ to Au0 . Subsequently, homogeneous core@shell nanoparticles having Au as core (thus, less atoms at
the surface) is expected. As, it will be shown in sections Electrochemical evidence of synergistic effect in Au-Pd bimetallic nanostructures
and Surface state and electronic properties in the nanomaterials: Xray photoelectron spectroscopy (XPS) measurements, multimetallic
nanoparticles have less Au atoms at their surface. Such reduction kinetics pathway leading to the formation of core-shell structures cannot
explain the presence of gold islands. So, we believe that two processes
occur during the synthesis. A part of Au4+ is swiftly reduced into Au+
followed by the co-reduction of Au+ by Pt4+ and Pd2+ species. Then,
the repulsion effect between these first multimetallic nanoparticles
and the remaining Au4+ species leads definitely to the formation of
isolated gold nanoparticles (Au islands) because of the metal miscibility limit at nanoscale level. With NaBH4 , the reduction occurs
so quasi-instantaneously that no segregation effect can take place.
Consequently, the miscibility of the three metals would limit thermodynamics, which should lead to homogeneous core@shell (a core
composed of Au atoms and a shell made of Pd and/or Pd atoms)
structure for multimetallic nanomaterials.
Electrochemical activity of the nanomaterials. The electrocatalytic activity of some as-prepared nanomaterials has been
recently compared by their polarization curves towards glucose

Downloaded on 2015-10-29 to IP 131.111.164.128 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Journal of The Electrochemical Society, 162 (14) H929-H937 (2015)

H933

Figure 4. Specific activities, evaluated at different electrode potentials from polarization curves on (a) Au-Pd/C, (b) Au-Pt/C and (c) Au-Pt-Pd/C electrode
materials: Polarization curves were recorded at 20 mV s1 in a 0.1 mol L1 NaOH + 10 mmol L1 glucose at 25 C. (d) Typical curves on Au80 Pt20 /C electrode
material in the absence (i), presence of 10 mmol L1 glucose (ii) and after correction with the electrolyte contribution (iii) at 20 mV s1 and 25 C.

electrooxidation reaction in alkaline medium.17 But, the present


studies aim at examining the origin of the noticed difference. Fig. 4
depicts the Volcano plots of the specific activity at different electrode
potentials. The specific activity is obtained by correcting the collected
current during glucose electrooxidation by the supporting electrolyte
contribution. Typical full curves at 20 mV s1 on Au80 Pt20 /C electrode
material in the absence, presence of 10 mmol L1 glucose and after
correction with the electrolyte contribution at 25 C are provided in
Fig. 4d. We especially focused on electrode potentials less than 0.7 V
vs. RHE to mimic their performances as anode materials in glucosebased FCs. Indeed, in alcohol FCs, the electrode potential is expected
to not exceed 0.7 V vs. RHE during operation. This enables to have
large cell voltage and subsequently a high output power. Among
all gold-palladium bimetallic nanomaterials, the most synergistic
effect is observed when 10 at.% of gold is replaced by palladium. At
0.4 V vs. RHE, Au90 Pd10 /C is almost 2-fold more efficient than the
monometallics. Table I shows Tafel plot tests; the exchanged current
density is very high on Au90 Pd10 /C compared to monometallics Au/C
and Pd/C. This enhancement can be primary explained by a possible
synergistic effect. We will see in section Surface state and electronic
properties in the nanomaterials: X-ray photoelectron spectroscopy

(XPS) measurements that XPS spectra give sound evidences on the


electronic effects synonymous of interactions between each element,
responsible for increasing glucose oxidation reaction kinetics at low
electrode potentials. Additionally, the presence of either facet (200)
or (110) may also positively contribute to this kinetic improvement
since same nanomaterials prepared from the water-in-oil method with
only (100) and (111) have lower performances.29
Concerning gold-platinum binaries (Fig. 4b), the entire compositions show better reaction kinetics than monometallics. Even if
Pt/C shows high specific activity for E > 0.5 V vs. RHE, chronoamperometry tests have shown that this catalyst is quickly deactivated
during the reaction. It is a common phenomenon in electrocatalysis
when considering organic molecule electrooxidation. We report in
Fig. 4c the Volcano plots for the optimized trimetallic nanomaterials
used as electrode materials. It clearly indicates the synergistic
effect between the three metals. This phenomenon is the desired
behavior for anode materials catalyzing organic molecules oxidation
in FCs. Compared with the literature (appropriate normalization), the
nanocatalysts developed here for the glucose electrooxidation exhibit
better performances in terms of the onset potential, current density
(more than 2-fold increase) and stability over time. To gain clear

Table I. Exchanged current density, j0 , of nanocatalysts obtained from the polarization curves performed at a scan rate of 2 mV s1 in presence
of 10 mM glucose in 0.1 M NaOH at 25 C.
Catalysts

j0 (A cm2 )

Pt/C

Pd/C

Au/C

52

64

261

62

801

Au90 Pd10 /C
Tafel plots for E < 0.6 V vs. RHE.
646
Tafel plots for E > 0.6 V vs. RHE.
1024

Au80 Pt20 /C

Au70 Pt15 Pd15 /C

238

109

961

230

Downloaded on 2015-10-29 to IP 131.111.164.128 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

H934

Journal of The Electrochemical Society, 162 (14) H929-H937 (2015)

Figure 5. CO stripping experiments performed at 20 mV s1 in 0.1 mol L1 NaOH electrolyte at 25 C. On Au-Pd/C electrode materials: (a) Recorded CVs and
(b) Surface atomic composition determined from the Rand and Woods method. (c) Au-Pt electrode materials steady-state CVs at 50 mV s1 . Background corrected
CO stripping curves at (d, f) Au-Pt/C, (e) Au-Pt-Pd/C electrode materials. In (f), the current was normalized using Pt active surface.

understanding about these improved catalytic properties, we have


performed series of deep investigations. CO stripping experiments
were first carried out on Au-Pd bimetallic nanomaterials. Then, XPS
analysis might provide complete evidences.
Electrochemical evidence of synergistic effect in Au-Pd bimetallic nanostructures. The nanomaterials surface has been electrochemically characterized by CO stripping experiments, as shown in
Fig. 5a for Au-Pd/C bimetallic catalysts. During the positive scan,
the adsorbed CO (COads ) is oxidized at ca. 0.8 V vs. RHE. Then, the

catalyst surface being free from COads , is oxidized at high electrode


potential. Finally, these oxides are reduced during the reverse scan.
Gold oxides are reduced at 1.09 V vs. RHE and those of palladium
are reduced at around 0.62 V vs. RHE. This figure contains two major findings. First of all, the oxidation peaks of COads shift towards
low electrode potentials for the bimetallic catalysts. The Au/C catalyst has been also tested towards CO stripping. It does not show
any significant catalytic activity, despite the theoretical predictions of
Nrskov35 and experimental results from bulk Au(111).36 But in bulk
solution (meaning a CO-saturated alkaline solution), it exhibits good

Downloaded on 2015-10-29 to IP 131.111.164.128 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Journal of The Electrochemical Society, 162 (14) H929-H937 (2015)


catalytic activity.3639 The additional constraint imposed by the CO
stripping experiments is the maintenance of the CO molecule on the
surface. It seems that this binding energy is so low that the adsorption
and retention on the electrode are difficult. Consequently, the weakly
adsorbed CO is lost during the dissolved CO removal under nitrogen atmosphere (which lasts roughly 2030 min). Subsequently, the
reaction kinetics improvement can be explained by electronic interactions between gold and palladium in bimetallics. Such experimental
observations confirm the noticed excellent catalytic activity towards
glucose electrooxidation.
For intermediate atomic compositions in Au-Pd systems, there
is only one peak where the oxide of bimetallic materials is reduced.
More importantly, it shifts towards higher potentials when gold content
increases in the bimetallic material. On other hand, only one phase,
that of alloy, is formed. To gain further insights about its composition,
we used the electrochemical method proposed by Rand and Woods40
to estimate the surface atomic composition of the Au-Pd electrodes
according to Equation 1.


Au
EAuPd
peak Epeak
AuPd
Pd
Au
[1]
Epeak = Pd Epeak +Au Epeak Pd = 100
Au
EPd
peak Epeak
Pd and Au (Pd = 100-Au ) are the Pd and Au surface atomic perAu
AuPd
centages (surf.at.%) and E Pd
peak , E peak and E peak are the oxide reduction
peak potentials for Pd, Au and AuPd, respectively.
The obtained surface atomic compositions are represented in
Fig. 5b. It is worth mentioning that values determined from the steadystate CVs at 20 mV s1 are quite the same. It can be concluded that
the electrochemical surface of the electrode is mostly composed of
palladium. A such phenomenon has been observed for Au-M (M
= Pt, Pd).41 When gold is associated either with palladium or platinum, the latter elements go on the surface under potential cycling.
Since HRTEM observations coupled with EDX analyses have shown
that a real core-shell structure was not formed, the present result suggests that gold and palladium form an alloy phase with an enrichment
in Pd atoms at its surface. The surface composition of Au90 Pd10 /C (38
surf.at.% Au and 62 surf.at.% Pd) electrode material is particularly
interesting. For organic molecules (and particularly carbohydrates)
electrooxidation, the activity at low electrode potentials increases in
the order Pt > Pd > Au. But for the stability, the order is Pt < Pd
< Au. Finally, the surface composition of Au90 Pd10 /C electrode material explains well the improved performances previously noticed.
As displayed in Fig. 5c, the Au-Pt electrodes show two distinguished
potentials of oxide reduction as already reported.29 No surface quantification was thus done; the Rand and Woods method is no longer
valid because of the presence of two peaks.40 Furthermore, Pt and Au
oxides reduction peaks in the CVs (Fig. 5c) were not used to quantify
their relative surface composition because these peaks are not well
defined when the metal loading is less than 20 wt.% on Vulcan for
the present nanocatalysts compared to 40 wt.% elsewhere.15,29 Only
a small AuOx reduction peak can be seen in the CVs. Thus, as we
have only one common reduction peak (where both AuOx and PtOx
are reduced), it is quite difficult to find the suitable monolayer charge
value as reference. Figs. 5d and 5e show the background corrected CO
stripping results for Au-Pt/C and Au-Pt-Pd/C nanocatalysts, respectively. The current is normalized using the weight of metals which
are active for CO stripping. These figures highlight the enhancement
of the electrocatalytic activity for the multimetallic catalysts at low
electrode potential. It should be emphasized that the same trend is
observed when CO stripping curves are normalized using the electrochemical active surface (ECSA) of material of Pt (Fig. 5f). In this case,
the ECSA of Pt was determined from hydrogen region for Au-Pt/C
electrode materials. The ECSA cannot be precisely determined for
Au-Pd/C and Au-Pt-Pd/C electrodes, because they present common
potential window where the oxides are reduced and there is no reference value for the monolayer charge. As no activity of gold is herein
observed for CO stripping, we attribute this highly enhancement to the
existence of synergistic effect, with certainly a substantial contribution
of electronic interactions between elements on these nanoparticles.

H935

Surface state and electronic properties in the nanomaterials: Xray photoelectron spectroscopy (XPS) measurements. To gain further insights into how the different elements interact in multimetallic
materials, we performed XPS measurements. This surface technique
provides deep information on the chemical composition of the analyzed sample together with the possible charge transfer between the
different chemical elements. A survey spectrum is first recorded for
all materials to have an overview and a qualitative analysis. Then,
a high-resolution spectrum is recorded for the different core-levels
of interest. Monometallics are first characterized in order to serve as
benchmarks. The high-resolution XPS spectra of Au 4f, Pt 4f and
Pd 3d regions are displayed in Figs. 6a, 6b and 6c, respectively. The
observed doublets for the same band are related to the spin-orbit splitting (1/2). We underscore important current trends that the obtained
materials are not (Pt/C) or less oxidized (Au/C and Pd/C). Further
examination of decomposed results in Fig. 6c indicates that there are
two kinds of oxides for Pd 3d. The doublet of gold metal are situated
at 83.9 eV for Au 4f7/2 and 87.6 eV for Au 4f5/2 . Those of Pt 4f are
located at 71.1 eV (Pt 4f7/2 ) and 74.5 eV (Pt 4f5/2 ) in agreement with
the literature.12,23,42,42 The doublets Pd 3d5/2 and Pd 3d3/2 are situated
at 335.3 eV and 340.6 eV.12 The presence of the oxide is simply due
to the natural oxidation of noble metal when exposed to ambient air.
Currently, a thin protective layer covers the metal surface from deep
oxidation (immunity layer). From high-resolution XRD patterns (not
shown herein), no additional oxide peak was observed, which endorses
completely the conclusion that the oxide amount is very insignificant.
We further expand our knowledge about possible electronic
interactions in multimetallic nanomaterials by recording their XPS
spectra. The high-resolution spectra are shown in Figs. 6d (Au 4f) and
6e (Pd 3d) for Au90 Pd10 /C, 6f (Au 4f) and 6g (Pt 4f) for Au80 Pt20 /C,
and 6h (Au 4f), 6i (Pt 4f) and 6j (Pd 3d) for Au70 Pt15 Pd15 /C samples.
In the material Au80 Pt20 /C, Pt 4f7/2 is situated at 70.8 eV, while that of
4f7/2 level of PtO and PtOx (unknown real composition) are situated
at 71.6 eV and 74.7 eV (Fig. 6g), respectively. When comparing the
BE of Pt 4f7/2 between the monometallic Pt/C material (71.1 eV) and
the bimetallic Au80 Pt20 /C one (70.8 eV), it appears that there is a
shift of 0.3 eV towards low BE. Such downshift has been reported
and attributed to a strong electronic interaction between gold and
platinum.41,43 According to Pederson et al,43 an electronic transfer
from the sub-layer 5d10 (filled) of Au to that of Pt (5d9 , unfilled)
occurs in Au-Pt alloy structures. Furthermore, Au90 Pd10 /C system
compared to Pd/C shows ca. 0.3 eV downshift of the BE of the Pd 3d
(3d5/2 at 335.0 eV for Au90 Pd10 /C) and less than 0.2 eV (the data being
collected each 0.1 eV) for Pt 4f7/2 and Pd 3d5/2 in the Au70 Pt15 Pd15 /C
material. It should be particularly pointed out that Pd 3d spectrum
in Au90 Pd10 /C and Au70 Pt15 Pd15 /C materials includes a contribution
from Au 4d and Pt 4d. As these peaks are quite wide, the difference
of 0.2 eV could include the error resulting from the fitting process.
Furthermore, quantifications show a surface enrichment in Pt and Pd
atoms as pointed out above in Electrochemical evidence of synergistic
effect in Au-Pd bimetallic nanostructures section and summarized in
Table II. Indeed, the surface atomic composition of Au90 Pd10 /C is 65
surf.at.% Au and 35 surf.at.% Pd. This composition is different to
that evaluated from the electrochemical method: 38 surf.at.% Au and
62 surf.at.% Pd. But the value of 35 surf.at.% Pd is 3.5 times higher
than the nominal composition (10 at.%) that has been confirmed
by ICP analyses. This highlights a surface enrichment when we
compare the ICP and XPS values, as observed from electrochemical
characterizations. For Au80 Pt20 /C, proportions are 50 surf.at.% Au
and 50 surf.at.% Pt. Considering the trimetallic Au70 Pt15 Pd15 /C
catalyst, 25 surf.at.% Au, 35 surf.at.% Pt and 40 surf.at.% Pd were
obtained. In the both cases, Pd and Pt contents are at least 2-fold
higher than the nominal ones. Obviously, the electrochemical and
physical methods lead to the same conclusions about the structure
of the as-prepared nanomaterials from BAE method. It is therefore
evident that the multimetallic surface enrichment in platinum and/or
palladium at the expense of gold combined with the strong electronic
interactions between the different chemical elements is undoubtedly
responsible for the enhancement of their electrocatalytic properties.

Downloaded on 2015-10-29 to IP 131.111.164.128 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

H936

Journal of The Electrochemical Society, 162 (14) H929-H937 (2015)

Figure 6. High-resolution XPS spectra of monometallic (a-d), bimetallic (d-g) and trimetallic (h-j) nanostructures.

Table II. Atomic composition of multimetallic nanomaterials: bulk


(from ICP-OES analyses) and surface (from XPS measurements).
Nominal atomic
composition
ICP: Real atomic
composition
XPS: Surface
atomic
composition
(surf. at.%)

Au90 Pd10

Au80 Pt20

Au70 Pt15 Pd15

Au90 Pd10

Au74 Pt26

Au74 Pt10 Pd16

Au

65

50

25

Pd
Pt

35
-

50

40
35

Conclusions
This study was aiming to establish the correlation between the
physico(electro)chemical properties and catalytic performances of
gold-platinum, gold-palladium and gold-platinum-palladium nanostructures synthesized from a recently initiated soft and convenient
chemical method. We underscore important physicochemical trends
that were then substantiated by electrochemical ones. The transmission electron microscopy (TEM) analyses showed that these nanoparticles are finely and well dispersed on the support (2-10 nm). These
nanoparticles have good proportion of metal atoms at their surface (exposed ratio of atoms or dispersion), reaching 35%. The high-resolution
images HRTEM allowed identifying the presence of twins leading to
the formation of crystalline facets (110) in Au80 Pt20 /C nanomaterials.

Downloaded on 2015-10-29 to IP 131.111.164.128 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Journal of The Electrochemical Society, 162 (14) H929-H937 (2015)


EDX and electrochemical analyses revealed that the structural composition of the bimetallic catalysts was composed of alloying structures which have less Au atoms at their surface. Furthermore, X-ray
photoelectron spectroscopy (XPS) measurements stressed out a downshift of ca. 0.3 eV of the BE of Pt 4f (in Au80 Pt20 /C) and Pd 3d (in
Au90 Pd10 /C) electrons.This shift demonstrates the strong electronic
interaction between the different elements, which enables reinforcing
the electrocatalytic properties of the obtained nanocatalysts. These
XPS analyses have definitely confirmed platinum or palladium surface enrichment of the gold-based bi and tri-metallic catalysts as previously mentioned. In other words, the alloy phase contains less Au
atoms at their surface. We effectively linked the structural properties
of these nanomaterials prepared without surfactant to their electrocatalytic performances towards glucose and CO oxidation.
Acknowledgments
This work was supported by funding from the French National Research Agency (ANR) through the financial grants ChemBio-Energy
program 20122015.
References
1. C. R. Henry, Prog. Surf. Sci., 80, 92 (2005).
2. D. Alloyeau, C. Mottet, and C. Ricolleau, Nanoalloys : Synthesis, Structure and
Properties, p. 420, Springer London (2012).
3. M. Lahmani, C. Brechignac, and P. Houdy, Les nanosciences: 2. Nanomateriaux et
nanochimie, p. 732, Belin, Paris, France (2012).
4. A. Wieckowski, E. R. Savinova, and C. G. Vayenas, Catalysis and Electrocatalysis
at Nanoparticle Surfaces, p. 970, Marcel Dekker, Inc., New York, USA (2003).
5. H. Zhang, M. Jin, and Y. Xia, Angew. Chem. Int. Ed., 51, 7656 (2012).
6. A. B. Shah, S. T. Sivapalan, B. M. DeVetter, T. K. Yang, J. Wen, R. Bhargava,
C. J. Murphy, and J.-M. Zuo, Nano Lett., 13, 1840 (2013).
7. Y. Yu, Q. Zhang, X. Lu, and J. Y. Lee, J. Phys. Chem. C, 114, 11119 (2010).
8. J. Zhang, M. R. Langille, M. L. Personick, K. Zhang, S. Li, and C. A. Mirkin, J. Am.
Chem. Soc., 132, 14012 (2010).
9. T. Yu, D. Y. Kim, H. Zhang, and Y. Xia, Angew. Chem. Int. Ed., 50, 2773 (2011).
10. M. Jin, H. Zhang, Z. Xie, and Y. Xia, Angew. Chem. Int. Ed., 50, 7850 (2011).
11. C.-W. Yang, K. Chanda, P.-H. Lin, Y.-N. Wang, C.-W. Liao, and M. H. Huang, J. Am.
Chem. Soc., 133, 19993 (2011).
12. G.-R. Zhang, J. Wu, and B.-Q. Xu, J. Phys. Chem. C, 116, 20839 (2012).
13. A. J. Appleby and C. Van Drunen, J. Electrochem. Soc., 118, 95 (1971).
14. M. W. Hsiao, R. R. Adzic, and E. B. Yeager, J. Electrochem. Soc., 143, 759
(1996).
15. P. Tonda-Mikiela, T. W. Napporn, C. Morais, K. Servat, A. Chen, and K. B. Kokoh,
J. Electrochem. Soc., 159, H828 (2012).
16. Y. Holade, C. Morais, S. Arrii-Clacens, K. Servat, T. W. Napporn, and K. B. Kokoh,
Electrocatalysis, 4, 167 (2013).

H937

17. Y. Holade, K. Servat, T. W. Napporn, and K. B. Kokoh, Electrochim. Acta, 162, 205
(2014).
18. A. Both Engel, Y. Holade, S. Tingry, A. Cherifi, D. Cornu, K. Servat, T. W. Napporn,
and K. B. Kokoh, J. Phys. Chem. C, 119, 16724 (2015).
19. Y. Holade, A. B. Engel, S. Tingry, K. Servat, T. W. Napporn, and K. B. Kokoh,
ChemElectroChem, 1, 1976 (2014).
20. Y. Holade, K. MacVittie, T. Conlon, N. Guz, K. Servat, T. W. Napporn, K. B. Kokoh,
and E. Katz, Electroanalysis, 26, 2445 (2014).
21. Y. Holade, C. Morais, K. Servat, T. W. Napporn, and K. B. Kokoh, Phys. Chem.
Chem. Phys., 16, 25609 (2014).
22. Y. Holade, T. W. Napporn, C. Morais, K. Servat, and K. B. Kokoh, ChemElectroChem,
2, 592 (2015).
23. C. J. Corcoran, H. Tavassol, M. A. Rigsby, P. S. Bagus, and A. Wieckowski, J. Power
Sources, 195, 7856 (2010).
24. M. S. Wilson, J. A. Valerio, and S. Gottesfeld, Electrochim. Acta, 40, 355
(1995).
25. J. R. Anderson and K. C. Pratt, Introduction to Characterization and Testing of Catalysts, p. 457, Academic Press Australia, Sydney (1985).
26. H. Bonnemann and G. A. Braun, Angew. Chem. Int. Ed., 35, 1992
(1996).
27. U. A. Paulus, A. Wokaun, G. G. Scherer, T. J. Schmidt, V. Stamenkovic,
N. M. Markovic, and P. N. Ross, Electrochim. Acta, 47, 3787 (2002).
28. S. Miracle-Sole, Scholarpedia, 8, 31266 (2013).
29. A. Habrioux, W. Vogel, M. Guinel, L. Guetaz, K. Servat, B. Kokoh, and
N. Alonso-Vante, Phys. Chem. Chem. Phys., 11, 3573 (2009).
30. J. Wang, J. Gong, Y. Xiong, J. Yang, Y. Gao, Y. Liu, X. Lu, and Z. Tang, Chem.
Commun., 47, 6894 (2011).
31. S. Hebie, L. Cornu, T. W. Napporn, J. Rousseau, and B. K. Kokoh, J. Phys. Chem.
C, 117, 9872 (2013).
32. L. Gan, C. Cui, M. Heggen, F. Dionigi, S. Rudi, and P. Strasser, Science, 346, 1502
(2014).
33. A. Sanchez-Iglesias, I. Pastoriza-Santos, J. Perez-Juste, B. Rodrguez-Gonzalez,
F. J. Garca de Abajo, and L. M. Liz-Marzan, Adv. Mater., 18, 2529 (2006).
34. C. L. Johnson, E. Snoeck, M. Ezcurdia, B. Rodriguez-Gonzalez, I. Pastoriza-Santos,
L. M. Liz-Marzan, and M. J. Hytch, Nat. Mater., 7, 120 (2008).
35. B. Hvolbk, T. V. W. Janssens, B. S. Clausen, H. Falsig, C. H. Christensen, and
J. K. Nrskov, Nano Today, 2, 14 (2007).
36. P. Rodriguez, Y. Kwon, and M. T. M. Koper, Nat. Chem., 4, 177
(2012).
37. H. Kita, H. Nakajima, and K. Hayashi, J. Electroanal. Chem. Interf. Electrochem.,
190, 141 (1985).
38. T. F. Jaramillo, S.-H. Baeck, B. R. Cuenya, and E. W. McFarland, J. Am. Chem. Soc.,
125, 7148 (2003).
39. P. Rodriguez, N. Garcia-Araez, A. Koverga, S. Frank, and M. T. M. Koper, Langmuir,
26, 12425 (2010).
40. D. A. J. Rand and R. Woods, J. Electroanal. Chem. Interf. Electrochem., 36, 57
(1972).
41. R. P. Doherty, J.-M. Krafft, C. Methivier, S. Casale, H. Remita, C. Louis, and
C. Thomas, J. Catal., 287, 102 (2012).
42. G. C. Allen, P. M. Tucker, A. Capon, and R. Parsons, J. Electroanal. Chem. Interf.
Electrochem., 50, 335 (1974).
43. M. . Pedersen, S. Helveg, A. Ruban, I. Stensgaard, E. Lgsgaard, J. K. Nrskov,
and F. Besenbacher, Surf. Sci, 426, 395 (1999).

Downloaded on 2015-10-29 to IP 131.111.164.128 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

Você também pode gostar