You are on page 1of 12

Research Article

pubs.acs.org/acscatalysis

Defects do Catalysis: CO Monolayer Oxidation and Oxygen


Reduction Reaction on Hollow PtNi/C Nanoparticles
Laetitia Dubau,*,†,‡ Jaysen Nelayah,§ Simona Moldovan,∥ Ovidiu Ersen,∥ Pierre Bordet,⊥,# Jakub Drnec,∇
Tristan Asset,†,‡ Raphael̈ Chattot,†,‡ and Frédéric Maillard*,†,‡

Université Grenoble Alpes, LEPMI, F-38000 Grenoble, France

CNRS, LEPMI, F-38000 Grenoble, France
§
Laboratoire Matériaux et Phénomènes Quantiques (MPQ), UMR 7162, CNRS & Université Paris-Diderot, Bâtiment Condorcet, 4
rue Elsa Morante, F-75205 Paris Cedex 13, France
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.


Institut de Physique et Chimie des Matériaux de Strasbourg (IPCMS), UMR 7504, CNRS-Université de Strasbourg (UdS), 23 rue
du Lœss, Cedex 2 Strasbourg, France

Université Grenoble Alpes, Institut Néel, F-38000 Grenoble, France
#
CNRS, Institut Néel, F-38000 Grenoble, France
Downloaded via CINVESTAV on January 8, 2020 at 07:28:11 (UTC).


European Synchrotron Radiation Facility, ID 31 Beamline, BP 220, F-38043 Grenoble Cedex, France
*
S Supporting Information

ABSTRACT: The catalytic performance of extended and nanometer-


sized surfaces strongly depends on the amount and the nature of structural
defects that they exhibit. However, whereas the effect of steps or adatoms
may be unraveled with single crystals (“surface science approach”),
implementing reproducibly in a controlled manner structural defects on
nanomaterials remains hardly feasible. A case that deserves particular
attention is that of bimetallic nanomaterials, which are used to catalyze the
oxygen reduction reaction (ORR) in proton exchange membrane fuel cells
(PEMFC). Point defects (vacancies), planar defects (dislocations and
grain boundaries), and bulk defects (voids, pores) are likely to be
generated in alloy or core@shell nanomaterials based on Pt and a
transition metal due to the high lattice mismatch between the two
elements. Here, we report the morphological and structural trajectories of
hollow PtNi/C nanoparticles during thermal annealing under vacuum, N2,
H2, or air atmosphere by in situ transmission electron microscopy and synchrotron X-ray diffraction. We evidence atmosphere-
dependent restructuring kinetics, which enabled us to synthesize a set of catalysts with identical chemical compositions and
elemental distributions but different morphologies, crystallite sizes, and lattice strain. By combining the results of Rietveld and
pair-distribution function analyses and electrochemical measurements, we demonstrate that the structurally disordered areas
located at the interface between individual crystallites are highly active for two reactions of interest for PEMFC devices: the
electrochemical COads oxidation and the ORR. These results shed fundamental light on the effect of structural defects on the
catalytic performance of bimetallic nanomaterials and should aid in the rational design of more efficient ORR electrocatalysts.
KEYWORDS: proton exchange membrane fuel cell, oxygen reduction reaction, platinum−nickel alloy, hollow nanoparticles,
structural defects, grain boundary, vacancy

■ INTRODUCTION
In both gas-phase and liquid catalysis, the reactivity of extended
There are evidences that open, low-density Pt(110) and
Pt(100) facets are more active than the close-packed Pt(111)
and nanometer-sized surfaces strongly depends on their surface in COads stripping experiments.5 The influence of
crystallographic orientation, the average coordination number structural defects is well-established for this reaction: on
of the atoms, and the amount and type of structural defects that stepped Pt[n(111)x(111)] single crystals, the onset and the
they exhibit.1−4 In that respect, the structure sensitivity of peak potential of COads electrooxidation shift toward negative
platinum (Pt) and platinum-based alloys (Pt-alloy) for the potentials with an increase in the step fraction.6,7 This indicates
electrochemical oxidation of COads is particularly illustrative.
This two-electron reaction can be performed either in the Received: April 18, 2016
absence of CO in solution (“COads stripping” experiments) or Revised: June 2, 2016
in its presence (“bulk CO” electrooxidation experiments). Published: June 7, 2016

© 2016 American Chemical Society 4673 DOI: 10.1021/acscatal.6b01106


ACS Catal. 2016, 6, 4673−4684
ACS Catalysis Research Article

that (i) steps are the “active sites” where water is dissociated functional theory (DFT) calculations, Calle-Vallejo et al.3
into oxygen-containing species (referred to as OHads in what showed that cavities modify the average coordination number
follows), which are necessary to complete the COads electro- of Pt atoms and enhance their catalytic activity for the ORR by
oxidation, and that (ii) adsorbed CO molecules diffuse rapidly a factor close to 3 relative to a cavity-free Pt surface. This is in
to these active sites. Note also that the structure sensitivity is line with the DFT findings of Callejas-Tovar et al.40 that
maintained in the presence of CO in the electrolyte.8 subsurface vacancies compress the Pt lattice parameter and
At the nanometer scale, and in contrast to what could be reduce the affinity for oxygenated species, thereby enhancing
expected on the basis of single-crystal studies, the onset and the the ORR kinetics on Pt-based catalysts.
peak potential in COads stripping voltammograms shift toward Here, hollow nanoparticles composed of a PtNi shell
positive potential when the fraction of highly undercoordinated surrounding a hollow core were synthesized by a combination
atoms increases (that is, when the size of the Pt nanocrystallites of galvanic replacement and nanoscale Kirkendall effects (a
decreases).9−12 This is related to the decrease of the COads vacancy-mediated diffusion mechanism in binary alloys where
diffusion coefficient and of the rate of the COads + OHads one species diffuses faster than the other). Using high-
recombination on Pt nanocrystallites with size less than 3−4 resolution transmission electron microscopy (HR-TEM), we
nm.9,13 Strikingly, the “negative” particle size effect levels off show that the PtNi shell is highly disordered and composed of
upon nanoparticle aggregation. The kinetics of the CO nanocrystallites interconnected by structurally disordered areas
electrooxidation9−11,14 and of more complex molecules, (“grain boundaries”). Thermal treatments under different
including formic acid,15 methanol,14−16 and ethylene glycol,17,18 atmospheres (vacuum, N2, H2, or air) were used to heal
is significantly enhanced on “multi-grained” nanostructures these structural defects in a controlled manner, resulting
formed by agglomerated Pt/C nanoparticles. The enhancement eventually in the collapsing of the hollow nanoparticles. This
in catalytic activity has been ascribed to the structurally provided us with a set of catalysts featuring identical chemical
disordered areas formed at the interface between crystal grains compositions and elemental distributions but different
(also referred to as “grain boundaries” in what follows). crystallite sizes, lattice strains, and morphologies, which was
Interestingly, the same holds true when a more oxophilic metal, used to shed light on the structure sensitivity of the
such as Ru or Sn, is alloyed to Pt: “multi-grained” Pt-alloy electrochemical COads oxidation and the ORR.


nanoparticles remain intrinsically more active than isolated Pt-
alloy nanoparticles for the electrooxidation of COads,19−22 RESULTS AND DISCUSSION
methanol,20,23 and ethanol,18,24 confirming that grain bounda-
ries play a key, yet underestimated, role in electrocatalysis. Synthesis and Characterization of Hollow PtNi/C
Structural defects also influence the rate of electroreduction Nanoparticles. Hollow PtNi nanoparticles supported on
reactions. On vicinal Pt surfaces, the kinetics of the oxygen high surface area carbon with a Ni fraction of 15 at. % were
reduction reaction (ORR) increase with an increase in the step synthesized via a one-pot method first introduced by Bae et
density, irrespective of the step site symmetry.25−30 This effect al.46 During the synthesis, Ni/C or Ni-B/C nanoparticles are
was recently rationalized by Bandarenka et al.,30 who postulated believed to form first47 and then act as sacrificial templates for
that the OHads binding energy on (111) terraces is weakened the deposition of Pt2+ ions via galvanic replacement (a simple
when the step density increases. Similarly to what was reported and spontaneous electrochemical process where the oxidation
for electrooxidation reactions, the catalytic activity for the ORR of Ni atoms provides electrons for the deposition of Pt2+
is depreciated when going from bulk Pt surfaces to nano- atoms). The continuous corrosion of Ni atoms progressively
particles (the specific activity decreases by a factor of 4 going creates a gradient in chemical potential between the core (rich
from bulk Pt to 2 nm sized crystallites31,32). However, even if in Ni) and the external surface (where Ni atoms are leached
the beneficial effect of steps and kinks on the ORR rate has out) and acts as a strong driving force for Ni atoms to diffuse
been documented for nanocatalysts,33,34 the impact of grain outward. This in turn drives a flux of vacancies in the opposite
boundaries, vacancies, and cavities remains understudied. direction, which coalesce in the center of the material
Recently, Wang et al.35 loaded carbon nanotubes with different (“nanoscale Kirkendall effect”) and ultimately results into
Pt mass fractions (in the range of 10−93 wt %). The authors hollow PtNi/C nanoparticles with a cavity similar in size/shape
reported aggregation of the Pt/C nanoparticles at mass fraction to that of the initial Ni or Ni-B/C nanoparticles.48−53 The
greater than 20 wt % concomitant with a positive shift of the scenario described above is supported by the high angle annular
half-wave potential in ORR curves (enhanced catalytic activity). dark field in scanning transmission electron microscopy
However, these results should be taken with caution, since the (HAADF-STEM) images and the X-ray energy dispersive
Pt loadings (the mass of Pt per geometrical surface area) were spectroscopy (X-EDS) chemical maps displayed in Figure 1. In
not identical in the different experiments. Using the same particular, the presence of a central void in the as-synthesized
approach, Nesselberger et al.36−38 evidenced an increased PtNi/C nanoparticles and the homogeneous distribution of Pt
specific activity for the ORR on highly loaded Pt/C and Ni atoms in the outer shell are clear evidences that the
nanoparticles. These findings were rationalized by considering nanoscale Kirkendall effect operated during the synthesis.
the overlap of the electrical double layers between adjacent Pt Figure 1 also shows the morphological changes of hollow
nanoparticles and the associated change of the electrochemical PtNi/C nanoparticles upon thermal annealing at T = 400 °C
potential at the compact layer,36,39 but the possible link under N2, H2, or air atmosphere. Some chemical and structural
between enhanced ORR electrocatalytic activity and grain parameters of the as-synthesized hollow and thermally annealed
boundaries was not considered. In addition to grain boundaries, PtNi/C nanoparticles, in particular the chemical composition
recent ab initio calculations2,3,40 and experiments from Adzic’s determined by X-EDS on hundreds of nanoparticles (low-
group41 and from the authors42−45 underlined the beneficial magnification TEM images) or on individual nanoparticles
effect of point (vacancies) and bulk defects (cavities; i.e., a (high-magnification STEM-XEDS images), are displayed in
cluster of vacancies) on the ORR kinetics. By means of density Table 1.
4674 DOI: 10.1021/acscatal.6b01106
ACS Catal. 2016, 6, 4673−4684
ACS Catalysis Research Article

size, indicating that the Pt atoms strained to accommodate the


lattice of the Ni-rich/C sacrificial template.
Post-synthesis thermal treatments were performed to control
the concentration of these structural defects, following the
approach described by Wang et al.54 for Pt3Co/C nanoparticles.
The authors reported that (i) surface atoms relaxed at 300 < T
< 400 °C and (ii) Pt and Co atoms segregate in the topmost
and second surface layers at T > 400 °C. Similar results by
other groups validated the thermal-annealing approach for
various Pt-based alloys.54−58 Figure 4 illustrates the morpho-
logical changes of hollow PtNi/C nanoparticles upon annealing
under vacuum.
Minor changes in morphology were noticed for T < 400 °C.
A nanoscale secondary void (indicated by arrows in Figure 4d)
Figure 1. HAADF-STEM images, line scan analysis, and elemental formed at T ≥ 400 °C, concomitant with a change in size and
distribution in the hollow PtNi/C nanoparticles treated at 400 °C shape of the central cavity. This void then interdiffused to the
under different atmospheres. A HAADF image is displayed in (a) and outer surface at 500 °C and annihilated at 600 °C, in agreement
STEM images in (b)−(d). In the elemental maps (middle) and in the with the Kirkendall theory.48,49 The hollow PtNi/C nanostruc-
line scans (right), the Pt atoms are represented in red and the Ni tures collapsed between 600 and 700 °C, resulting in the
atoms in green. The X-EDS signals recorded at each point of the line formation of solid PtNi/C nanoparticles. During this process,
scan (blue line) are displayed on the right-hand side. the number of individual grains was divided roughly by a factor
of 3, in agreement with the ex situ HR-TEM images displayed
Morphological Changes of Hollow PtNi/C Nano- in Figure 2.
particles during Thermal Annealing. Conventional TEM The restructuring kinetics were found to be strongly
images of the as-synthesized and thermally annealed PtNi/C dependent on the size of the central void: the hollow PtNi/C
nanoparticles are displayed in Figure 2a−d. They show that the nanoparticles featuring the smallest central void collapsed at
fresh PtNi/C nanoparticles are polydisperse with respect to size temperatures lower than those with larger voids (this is clearly
(10−15 nm) and shape (spherical to ellipsoidal). Details on the visible if one compares Figure 4 and Figure S1 in the
average inner and outer diameters of the fresh hollow PtNi/C Supporting Information, where another series of STEM-
nanoparticles may be found in Figure S3 of ref 44. HAADF images is displayed). This suggests that the
The HR-TEM images of Figure 2 indicate that the PtNi shell restructuring of hollow nanoparticles is thermodynamically
of the as-synthesized hollow nanoparticles is composed of driven (as the passage from two to one surface minimizes the
individual nanocrystallites ca. 2−3 nm in size featuring different total surface energy of the nanoparticle) but is kinetically
crystallographic orientations and being interconnected by limited by the amount of vacancies contained in the central
structurally disordered regions. This polycrystalline nanostruc- void (that is, the size of the hollow core). Keeping the same
ture suggests that multiple Pt nucleation centers formed on the idea in mind, we emphasize that the kinetic Monte Carlo
initial Ni or Ni−B/C nanoparticles grew and then contacted simulations of Gusak et al.59 showed that the time to collapse
each other. It is important to stress that the structural disorder for a hollow nanoparticle depends also on the temperature, the
is not limited to the grain boundaries: the PtNi lattice is also radius of the collapsed nanoparticle, the surface tension of the
distorted in the individual crystallites, therefore prohibiting any metal(s) considered, their atomic volume, and their inter-
clear indexation of the lattice planes present (Figure 3). The diffusion coefficients.
interplanar distance measured for different crystallites was on In addition, we noticed a significant influence of the gas
average 2.18 ± 0.06 Å, suggesting (111) planes in a face- atmosphere on the restructuring kinetics. As shown by Figure 5,
centered cubic (fcc) PtNi structure (Figure 3d,e). This distance the rearrangement of PtNi atoms initiated at 210 °C in the
is shorter than that in pure solid Pt nanocrystallites of the same presence of oxygen and was complete at 280 < T < 350 °C.

Table 1. Chemical and Structural Parameters of the Electrocatalysts Evaluated in This Worka
X-EDS XRD
lattice lattice param
at. comp (%) from analyses on hundreds of at. comp (%) from analyses on individual param contraction vs Pt/C
nanoparticles (low magnification) nanoparticles (high magnification) (nm) (%)
solid Pt/C Pt100Ni0 Pt100Ni0 0.3933 0
bulk Ni (PDF card no. 0.3524 10.4
00-004-0850)
fresh hollow PtNi/C Pt86±2Ni14±2 Pt88±2Ni12±2 0.3869 1.6
hollow PtNi/C after thermal Pt87±2Ni13±2 Pt88±5Ni12±5 0.3908 0.6
annealing under N2
hollow PtNi/C after thermal Pt87±2Ni13±2 Pt90±1Ni10±1 0.3886 1.2
annealing under H2
hollow PtNi/C after thermal Pt88±2Ni12±2 Pt89±1Ni11±1 0.3902 0.8
annealing under air

a
The atomic composition (at. comp) was determined by X-EDS on hundreds of nanoparticles (low-magnification TEM images) or on individual
nanoparticles (high-magnification STEM-XEDS images).

4675 DOI: 10.1021/acscatal.6b01106


ACS Catal. 2016, 6, 4673−4684
ACS Catalysis Research Article

Figure 2. Conventional TEM and aberration corrected HR-TEM images of the as-synthesized and thermally annealed PtNi/C nanoparticles: (a−d)
assemblies and (e−h) HR-TEM images of the as-synthesized PtNi/C nanoparticles and after thermal annealing at 400 °C under different
atmospheres. Zoom-in images of the fine nanostructure of the PtNi shell are shown in (i−l). The boundaries between the individual PtNi
nanocrystallites are highlighted in color.

This result is ascribed to the decrease of the surface energy in the surface and near-surface region increased upon thermal
upon chemisorption and the enhanced diffusivity of Pt atoms annealing under hydrogen, most impressively after thermal
under hydrogen or oxygen coverage. As shown on Pt single annealing under air. We rationalized this by considering that,
crystals, the motions of Pt surface atoms are much faster under due to the acid leaching step performed in the last step of the
a gas atmosphere,60,61 and facilitate the restructuring of the bulk synthesis, the fresh hollow PtNi/C nanoparticles feature a
of the nanoparticles.62 Convincing evidence of enhanced “skeleton” structure composed of a 2−3 monolayer thick Pt
restructuring of Pt nanoparticles under gas atmosphere may shell covering the mother PtNi alloy. Therefore, thermal
also be found in the study by Cabie et al.63 The authors showed annealing at T = 400 °C led to a more homogeneous
that exposure of cubic Pt nanoparticles to an O2 or H2 distribution of Pt and Ni atoms, resulting in an increased Ni/Pt
atmosphere promotes the development of (001) or (111) ratio at the surface (Figure S4A) with (Ni/Pt)air > (Ni/Pt)H2 >
facets, respectively. The shape changes were found to be
(Ni/Pt)fresh. Moreover, we found that thermal annealing under
reversible and were also observed for Cu,64 Rh,65 and Pd66
air atmosphere facilitated the formation of Pt and Ni oxides
nanoparticles.
Interestingly, slight segregation of Ni or Pt atoms to the (Figure S4B,C), in agreement with former reports of Menning
surface and subsurface layers was observed upon thermal et al.71 and Ahmadi et al.56,70 on the PtNi system and of Xin et
annealing under air or H2 atmosphere by electron energy loss al.72 on PtCo nanoparticles, respectively. However, it is
spectroscopy (EELS) elemental maps (Figure S2 in the important to note that the electrocatalytic activity of the
Supporting Information) or local X-EDS analyses (Figure S3 PtNi/C nanoparticles is only mildly affected by chemical
in the Supporting Information). A slight enrichment in Pt in changes in the surface and near-surface layers. Indeed, in acidic
the surface and near-surface layers was observed upon thermal electrolyte and especially under ORR conditions, Ni atoms are
annealing under H2 by local X-EDS analyses (Figure S3), in leached out from the first 2−3 monolayers. This is clearly
agreement with the findings of refs 67−69. Similarly, after visible in Figure 1 if one compares the X-EDS line-scan profiles
thermal annealing under an air atmosphere, the surface of the for Pt and Ni atoms. Convincing evidence was also provided by
PtNi/C nanoparticles was slightly enriched in Ni, in agreement Durst et al.69 with the help of aberration-corrected scanning
with refs 56 and 70. Note, however, that surface enrichments transmission electron microscopy (HR-STEM) and EELS
detected by X-EDS remained within the incertitude bar of the elemental mapping that Co atoms are removed from the
analyses (±1−2 at. %; see Table 1) and that X-EDS and EELS second and third atomic layers of Pt3Co/C nanoparticles upon
cannot provide quantitative insights into the composition of the exposure to an acidic environment. The change in structure was
surface and near-surface layers (these are bulk techniques). To fast and independent of the configuration of the starting Pt3Co/
better quantify the surface composition resulting from thermal C nanoparticles: Pt3Co/C alloy or Pt skin nanostructure (Pt
annealing, X-ray photoelectron (XPS) measurements were skin refers to a configuration in which the outermost surface
performed. The results displayed in Figure S4 showed that the layer is Pt rich, while the second layer is Ni rich). In other
Ni content and the oxidation state of Pt and Ni atoms located words, the surface and near-surface composition of the fresh or
4676 DOI: 10.1021/acscatal.6b01106
ACS Catal. 2016, 6, 4673−4684
ACS Catalysis Research Article

Figure 3. Lattice distortions in the shell of the as-synthesized hollow PtNi/C nanoparticles: (a) HR-TEM image, (b, c) zoom-in images of the PtNi
shell. The white boxes with dotted line indicate the regions over which the interplanar distances displayed in (d, e) have been measured. These
distances are expressed in angströms.

thermally annealed PtNi/C electrocatalysts under operando atmosphere-dependent healing of point (vacancies), surface
conditions is assumed to be Pt rich. planar (grain boundaries), and bulk defects (voids, cavities).
Structural Changes of Hollow PtNi/C Nanoparticles To separate the effects of crystallite size and microstrain
during Thermal Annealing under Different Atmos- (inhomogeneous interplanar distance) on the broadening of
pheres. To gain insights into the structural changes occurring the XRD peaks, the synchrotron XRD patterns were refined
on the hollow PtNi/C nanoparticles during thermal annealing with the Rietveld method after accounting for the instrumental
under different atmospheres, in situ laboratory X-ray diffraction broadening (Figure 6). Complementary to Rietveld refinement,
(XRD) experiments were performed (the in situ XRD spectra an atomic pair distribution function (PDF) analysis of the
taken at different holding temperatures are displayed in Figures synchrotron data was carried out. The measured and fitted
S5−S7 in the Supporting Information), and the thermally XRD patterns may be found in Figures S8 and S9 in the
annealed samples were then analyzed ex situ with synchrotron Supporting Information, and the parameters used for the
XRD radiation. A pronounced narrowing of the XRD peaks Rietveld and PDF analyses are given in Tables S1 and S2 in the
Supporting Information, respectively.
(signaling the growth of the PtNi crystallites) was observed
The results of the Rietveld analysis indicated that the
during thermal annealing under H2 or air (Figure 6), thereby
coherent domain size varied by a factor of 3 during thermal
confirming the electron microscopy findings obtained under
annealing (Figure 6b) but always remained between 3.3 and
the same experimental conditions. A contrario, the XRD peaks 10.1 nm, i.e., much smaller than the diameter of the hollow
remained broad for the fresh hollow PtNi/C nanoparticles nanoparticles. This confirms the HR-TEM images showing that
before/after thermal annealing under N2. Along with these the hollow PtNi/C nanoparticles are formed by several
changes, the XRD patterns systematically shifted toward smaller crystallites (grains), which have no structural coherence
2θ angles upon thermal annealing: i.e., the PtNi lattice relaxed, between them. In all of the samples, the PtNi nanocrystallites
whatever the gas atmosphere. Similarly to what was found for featured anisotropic shapes and were found elongated along the
the changes in morphology, the relaxation of the PtNi lattice [111] direction. The anisotropy (difference between the
was gas dependent, being maximum under N2, less pronounced longest and shortest directions) was close to ∼6−7 nm,
under air, and minimum under H2. Since the Ni content and independent of the average crystallite size. Finally, the Rietveld
the elemental distribution remained identical in the fresh and analysis revealed that microstrain significantly contributes to
the heat-treated PtNi/C nanoparticles, this result is ascribed to the broadening of the XRD peaks for the fresh and the N2-heat-
4677 DOI: 10.1021/acscatal.6b01106
ACS Catal. 2016, 6, 4673−4684
ACS Catalysis Research Article

Figure 4. In situ snapshots of the structural changes occurring on hollow PtNi/C nanoparticles during thermal annealing under vacuum: (a, c)
aberration-corrected HAADF-STEM and (b, d) bright-field-STEM (BF-STEM) images. The heating rate was 20 °C min−1. Thirty minute steps were
performed at each temperature of interest. Note that the scale bar is different in the sixth column (700 °C) of (c, d) relative to the other columns. In
these images, the magnification was changed to illustrate that the collapsed PtNi nanoparticles remained polycrystalline upon thermal annealing.
Arrows indicate the formation, interdiffusion, and annihilation of a nanoscale secondary void at T ≥ 400 °C.

treated hollow PtNi/C samples and is reduced upon thermal hollow nanoparticles relative to solid nanoparticles with
annealing under H2 or air (Figure 6d), during which the identical crystallite size (Figure S10).
average crystallite size increased markedly. Finally, we emphasize that no direct relation exists between
The results of the PDF analysis were remarkably similar the PtNi lattice relaxation and the microstrain broadening. This
(note, however, that anisotropic size effects are not taken into is because the microstrain represents the local deviation from
account by this analysis, therefore explaining why it systemati- the average lattice parameter (i.e., Δa/a, where a is the lattice
cally yields smaller crystallite sizes). The PDF refinements parameter) caused by the structural defects. An illustration of
showed that the thermally annealed PtNi/C nanoparticles this may be found in Figure 6c,d: the PtNi/C nanoparticles
maintained an fcc structure but that the lattice parameter which were thermally annealed under an H2 atmosphere feature
relaxed significantly under the different gas atmospheres. We a more compressed lattice constant (by 1.2%) in comparison to
also noticed that the values of the atomic displacement those thermally annealed under air (0.8%), but both catalysts
parameter Uiso (determined by PDF refinement) and of the have similar microstrain. The same holds true for the PtNi/C
nanoparticles in the fresh state and after thermal annealing
microstrain (determined by Rietveld analysis) were linearly
under N2 (similar microstrains but different lattice constants).
correlated. Since Uiso reflects the distribution of interatomic
Structure−Catalytic Activity Relationships. Motivated
distances from their average values and thus incorporates the by these very different structural characteristics, we investigated
effects of microstrain, this result confirms the variations of the the catalytic activity of the fresh and thermally annealed PtNi/C
microstrain for the different samples. nanoparticles for two reactions of interest for PEMFC devices:
As illustrated in Figure S10 in the Supporting Information, a the electrochemical COads oxidation and the oxygen reduction
large part of the microstrain arises from the nanometric reaction. Figure 7 shows a clear relationship between the
dimension of the metal crystallites. This should be correlated to concentration of structural defects and the catalytic activity of
the 4υmγ/d term in the Gibbs−Thomson equation,73 which the PtNi/C nanoparticles for the COads electrooxidation. Two
quantifies the increase of the chemical potential for nanoma- oxidation prepeaks at 0.66 < E < 0.70 V vs RHE dominate the
terials relative to the bulk (υm is the molar volume, γ is the reactivity of the fresh and the N2-annealed hollow PtNi/C
surface tension, and d is the crystallite size). However, due to nanostructures. In contrast, a peak located at 0.80 < E < 0.82 V
the large concentration of structural defects in the hollow PtNi/ vs RHE predominates in the COads stripping voltammograms
C nanocrystallites, the microstrain is more pronounced in performed on the solid Pt/C and the solid PtNi/C nano-
4678 DOI: 10.1021/acscatal.6b01106
ACS Catal. 2016, 6, 4673−4684
ACS Catalysis Research Article

Figure 5. In situ monitoring of the structural changes occurring on hollow PtNi/C nanoparticles during thermal annealing under oxygen at
atmospheric pressure: aberration-corrected (a, c) HAADF and (b, d) STEM images. The heating rate was 30 °C min−1.

Figure 6. Structural parameters of the fresh and the heat-treated


hollow PtNi/C nanoparticles and the reference Pt/C material used in
this study: (a) synchrotron XRD spectra after linear heating from
room temperature to 400 °C at a rate of 2 °C min−1; (b) average
crystallite size; (c) lattice parameter; (d) microstrain estimated by Figure 7. Electrocatalytic activity of the hollow and solid PtNi/C and
Rietveld analysis of the XRD spectral pattern. the reference solid Pt/C nanoparticles for the electrochemical COads
oxidation: iR-corrected (a) base and (b) COads stripping voltammo-
grams measured on the hollow and the solid PtNi/C nanoparticles and
particles (which were thermally annealed in H2 or in air). The the reference Pt/C material. Conditions: Ar-saturated 0.1 M HClO4; v
contribution of the COads electrooxidation prepeaks to the total = 0.020 V s−1; 25 ± 1 °C; no rotation of the electrode; Pt loading 20
electrooxidation charge was maximum for the fresh and the μgPt cm−2geo.
heat-treated under N2 hollow PtNi/C nanoparticles (featuring a
high concentration of grain boundaries) and minimum for the A second structure−activity relationship appears if the
PtNi/C nanoparticles thermally annealed under H2 or air and catalysts are sorted into two distinct groupsthe hollow
the reference solid Pt/C material. PtNi/C nanoparticles (fresh and thermally annealed under N2)
4679 DOI: 10.1021/acscatal.6b01106
ACS Catal. 2016, 6, 4673−4684
ACS Catalysis Research Article

Figure 8. Electrocatalytic activity of the hollow and solid PtNi/C and the reference Pt/C nanoparticles for the oxygen reduction reaction. (a) Linear
sweep voltammograms. (c) Tafel plots of ohmic and mass transport corrected linear sweep voltammograms displayed in (a). (b) and (d) are the
specific and mass activities (MA0.95) for the ORR determined at E = 0.95 V vs RHE. Conditions: O2-saturated 0.1 M HClO4: potential sweep rate
0.005 V s−1; ω = 1600 rpm; positive-going potential sweep from 0.20 to 1.05 V vs RHE; temperature 25 ± 1 °C; Pt loading 20 μgPt cm−2geo. The
error bars are the standard deviation of at least three independent measurements.

and the solid PtNi/C nanoparticles (thermally annealed under


air or H2)and compared pairwise. Within both groups, the
PtNi/C nanocatalysts featuring relaxed lattice parameters better
catalyze the COads monolayer electrooxidation, as deduced
from the shift of the onset and the peak potential to lower
potential value (Figure 7b). These results can easily be
understood in view of the higher affinity of relaxed PtNi
lattices for OHads species,74,75 which are required to complete
the electrooxidation of the COads monolayer. From a more
global perspective, the results indicate that the CO ads
monolayer electrooxidation is a fast and effective way to
check the affinity of bimetallic Pt-alloy nanocatalysts for
oxygenated species.
An opposite structure−catalytic activity relationship was Figure 9. Dependence of the specific activity for the ORR measured at
found for the ORR (Figure 8). In agreement with the E = 0.95 V vs RHE on the lattice parameter and the nanostructure/
literature,75,76 the electrocatalysts featuring the more com- concentration of grain boundaries and atomic vacancies of/in the
pressed PtNi lattice were catalytically more active for the ORR. nanoparticles.
This may again be easily rationalized in the frame of the d band
theory,77−79 which states that a compression of the Pt lattice structural defects, such as grain boundaries and vacancies, also
results in a downshift of its average energy with respect to the contribute to the catalytic enhancement. We rationalize our
Fermi level. This downshift in turn weakens the binding energy findings in the framework of the recent DFT “coordination-
of ORR intermediates (OHads and OOHads) and accelerates the activity” plots of Calle-Vallejo et al.,3 which evidenced that the
ORR kinetics. presence of subsurface cavities on a Pt(111) surface locally
Strikingly, for nanocatalysts featuring similar lattice constants, increases the mean coordination number of the surface atoms
the ORR rate was strongly enhanced in the presence of grain and enhances their ORR activity. Our results show for the first
boundaries and vacancies (Figure 9). Since (i) no Ni was lost time that grain boundaries and vacancies are beneficial defects
and (ii) the chemical composition of the topmost and near- for ORR catalysis. We argue that the lattice distortions and
surface layers remains Pt rich under operando conditions different atomic packing densities in (or in close proximity to)
(acidic electrolyte), the better electrocatalytic activity of the these structural defects modify the electronic structure of Pt
hollow PtNi/C nanocatalysts can only be related to their atoms and optimize the binding of ORR intermediates. In this
unique structural arrangement: i.e., to their hollow nanostruc- framework, the synchrotron XRD results presented in this
ture. In our last paper,44 we ascribed the enhanced ORR study unequivocally show that microstrain provides catalytic
kinetics on hollow PtNi/C nanoparticles relative to solid PtNi/ sites with compressed lattice constants, leading to ORR activity
C nanoparticles to a combination of strain and ensemble effects ascending toward the apex of the Sabatier volcano plot (PtNi
and to their porous architecture. The present study shows that catalysts with Ni fraction less than 25 at. % are located on the
4680 DOI: 10.1021/acscatal.6b01106
ACS Catal. 2016, 6, 4673−4684
ACS Catalysis Research Article

strong-binding side of the volcano plot).55,75,76 Hence, the Synchrotron X-ray Diffraction Measurements. The
combination of a hollow nanostructure and of a Pt-alloy/C synchrotron XRD measurements were performed at the ID31
nanocatalyst is beneficial to ORR catalysis: first because the beamline at the European Synchrotron Radiation Facility
hollow-induced large microstrain allows accessing a large set of (ESRF) in Grenoble, France. Radiation from one undulator was
Pt−Pt bond distances and second because microstrain is stable monochromatized with a multilayer monochromator (band-
unless a high-temperature annealing procedure is employed width 0.3%) to the final energy of 60.0 keV (λ = 0.207 Å). The
(Figures 4 and 5). This is in contrast with (but complementary X-ray beam was focused with two transfocators to the final size
to) conventional Pt-alloy/C catalysts for which strain/ligand of 5 μm × 20 μm (vertical × horizontal relative to the plane of
effects are progressively vanishing due to dissolution of the the accelerator ring), and the flux at the sample position was 5
alloying element in the harsh operating conditions of a × 1012 photons s−1. The data were collected using a
PEMFC, leading ultimately to depreciated ORR ki- PerkinElmer 2D detector and azimuthally integrated using
netics.42,76,80−89 the pyFAI software package.


The Rietveld refinements were carried out using FullProf.90
CONCLUSION The XRD pattern of a CeO2 sample was used to determine the
instrument resolution function. The refinements were based on
In conclusion, using thermal annealing under different gas the structure of a cubic close-packed metal (Fm3̅m, a ≈ 3.9 Å),
atmospheres, we synthesized a series of PtNi/C nanoparticles including a polynomial description of the background. The best
with identical chemical compositions and elemental distribu- agreements were obtained with a description of peak shape
tions but different nanostructures (hollow vs solid), crystallite using the TCH function,91 including an isotropic description of
sizes, and lattice strains. By combining synchrotron XRD, in microstrain and a uniaxial anisotropic description of particle
situ TEM, and electrochemical measurements, we provided sizes, with the [111] direction as main axis. The refined
experimental evidence that structural defects such as grain
parameters and agreement factor values are summarized in
boundaries and vacancies enhance the kinetics of two reactions
Table S1 in the Supporting Information and the refined plots in
of interest for PEMFC electrocatalysis: the electrochemical
Figure S9 in the Supporting Information. The PDFs were
COads monolayer oxidation and the ORR. Our results shed
obtained from the same patterns using PDFGetX3.92 The data
fundamental light on the effect of structural defects on the
were considered up to Qmax = 23 Å−1. The pattern from a
catalytic performance of bimetallic nanomaterials and should
capillary containing only the carbon phase used in the
ultimately aid in the rational design of electrocatalysts with
nanoparticle samples was subtracted from the data. The PDF
enhanced activity for the sluggish ORR, the reaction limiting
of the CeO2 standard was used to determine the instrumental
the electrical performance of PEMFC devices.


parameters (damping and broadening of PDF peaks due to
experimental resolution), which were then fixed during the
MATERIALS AND METHODS PDF refinements, carried out with PDFgui.93 The refinements
Reference Electrocatalyst. A Pt/Vulcan XC72 catalyst were carried out for 1 ≤ r ≤ 40 Å on the basis of the same
with a weight fraction (wt %) of 20% was purchased from E- metal structure as for the Rietveld refinements. The refined
TeK and used as reference material without any treatment. The parameters were a scale factor, the a cubic cell parameter, the δ
number-averaged Pt nanoparticle size was 2.9 ± 0.6 nm. parameter taking into account linear atomic correlations, the
Synthesis of Hollow PtNi/C Nanoparticles. The hollow isotropic overall atomic displacement parameter (adp) Uiso, and
PtNi/C nanoparticles were synthesized by mixing 0.154 g of the parameter D corresponding to the diameter of the isotropic
Pt(NH3)4Cl2.H2O (Alfa Aesar, Specpure) and 0.180 g of NiCl2 coherent domain size. The results are shown in Table S2 in the
(Fluka, >98.0%) with 0.3 g of Vulcan XC72R (Cabot), 10 mL Supporting Information, and the refinement plots are shown as
of ethanol, and 140 mL of deionized water (Millipore). An Figure S8 in the Supporting Information.
aqueous solution of NaBH4 (Aldrich 99.99%; 5.5 mmol, 0.22 X-ray Photoelectron Spectroscopy Measurements.
M) was then added at a rate of 5 mL min−1 and the mixture The XPS measurements were acquired in an ultra-high vacuum
stirred for 1 h with magnetic stirring at room temperature (20 chamber (10−10 mbar base pressure, 10−8 mbar during the
± 2 °C). The resulting mixture was filtered, thoroughly washed measurements) equipped with a hemispherical analyzer (VSW
with deionized water, and dried for 45 min at 110 °C. The H100) and monochromatic X-ray source (SPEC XR-50)
catalyst powder was then acid-treated for 22 h in a stirred 1 M operating at 270 W (13.5 kV, 20 mA). The hemispherical
H2SO4 solution at 20 °C. analyzer was working in the constant pass energy mode (44
Laboratory X-ray Diffraction Measurements. The eV). Analyses were carried out at an angle of 60° between the
synthesized and reference electrocatalysts were analyzed using sample surface and the analyzer axis. The XPS data signals were
a PANalytical X’Pert Pro MPD vertical goniometer/diffrac- taken in increments of 0.25 eV with dwelling times of 10 s. The
tometer equipped with a diffracted-beam monochromator using PtNi/C electrodes were prepared by depositing a nonquantified
Cu Kα mean radiation (λ = 0.15418 nm) operating at 45 kV amount of powder on carbon scotch, which was then glued
and 40 mA. The 2θ angle extended from 10 to 125° and varied onto a molybdenum block and introduced into the XPS
using a step size of 0.033°, accumulating data for 525 s. In situ chamber within 3 h before the measurements. The XP spectra
XRD spectra on the hollow PtNi/C nanoparticles were were analyzed using the CasaXPS software from CasaSoftware
recorded during thermal annealing from room temperature to Ltd. using the Shirley background corrections, asymmetric peak
400 °C at a heating rate of 2 °C min−1. The heating assembly shapes for Pt 4f doublets, and blend of Lorentzian and Gaussian
(Anton-Paar DHS 1100) consisted of a chemically resistant peak shape for Ni 2p doublets. The reference was a partially
ceramics heating plate and a homemade stainless-steel crucible oxidized Ni foil.
(sample holder). A domed-shape X-ray transparent window Conventional TEM Measurements and X-EDS Map-
made of graphite, which was fixed on the heating assembly, ping. The electrocatalysts were examined with a JEOL 2010
allowed control of the gas atmosphere. TEM instrument operated at 200 kV with a point to point
4681 DOI: 10.1021/acscatal.6b01106
ACS Catal. 2016, 6, 4673−4684
ACS Catalysis Research Article

resolution of 0.19 nm. The X-EDS elemental maps were was denoised using standard principal component analysis
acquired using a JEOL 2100F microscope operated at 200 kV implemented in the hyperspectral data analysis toolbox
and equipped with a SDD Centurio retractable detector. The HyperSpy.95
X-EDS spectra were recorded on individual nanoparticles by Electrochemical Measurements. All the glassware
scanning the beam in a square region adjusted to the particle accessories used in this study were first cleaned by soaking in
size. The quantitative analyses were performed on Pt L and Ni a H2SO4/H2O2 mixture for at least 12 h and thoroughly
K lines using the K factor provided by the JEOL software. washing with ultrapure water. The 1 M H2SO4 solution used for
In Situ TEM Measurements. The in situ measurements acid leaching was prepared with Milli-Q water (Millipore, 18.2
under vacuum or O2 environment were performed using a MΩ cm, total organic compounds <3 ppb) and H2SO4 96 wt %
JEOL 2100F TEM/STEM electron microscope equipped with (Suprapur, Merck).
a Cs probe corrector. The images were acquired in the STEM The electrochemical measurements were conducted using an
mode using simultaneously both the HAADF (high angle Autolab PGSTAT302N instrument in a custom-made four-
annular dark field) and BF (bright field) detectors. For the in electrode electrochemical cell thermostated at 25 ± 1 °C. Fresh
situ measurements under vacuum, a Gatan heating holder was electrolyte solution (0.1 M HClO4) was daily prepared with
used, and the PtNi/C electrocatalysts were deposited on a Milli-Q water (Millipore, 18.2 MΩ cm, total organic
classical TEM grid covered by a carbon membrane. The sample compounds <3 ppb) and HClO4 96 wt % (Suprapur,
was heated from room temperature to 700 °C at a rate of ca. 20 Merck). The counter electrode was a glassy-carbon plate and
°C min−1, and held for 30 min at each temperature. For the in the reference electrode a commercial reversible hydrogen
situ measurements under oxygen, a Protochips Atmosphere electrode (Hydroflex, Gaskatel GmbH) connected to the cell
Gas environmental cell was used. The PtNi/C electrocatalysts via a Luggin capillary. A Pt wire connected to the reference
were deposited on the Si3N4 membrane in contact with the SiC electrode was used to filter the high-frequency electrical noise.
heating ceramic device (E chips) submitted to a pure oxygen More details on the dual-reference system used in this work can
environment at atmospheric pressure. The specimen was be found in ref 96.
heated from room temperature to 450 °C at a rate of ca. 30 To prepare the working electrodes, a suspension containing
°C min−1 and held for 30 min at each temperature.
10 mg of Pt/C from the fresh or the heat-treated hollow PtNi/
Aberration-Corrected HR-TEM and Spectrum-Imaging
C nanoparticles, 54 μL of 5 wt % Nafion solution (Electro-
EELS. Aberration-corrected HR-TEM images and two-dimen-
chem. Inc.), 1446 μL of isopropyl alcohol, and 3.6 mL (18.2
sional spatially resolved EELS elemental maps were acquired
MΩ cm) of deionized water (MQ-grade, Millipore) was made.
with a JEOL-ARM 200F (JEOL) transmission electron
microscope equipped with a cold-field emission gun operated After sonication for 15 min, 10 μL of the suspension was
at 200 kV and a CEOS aberration corrector of the objective pipetted onto a glassy-carbon disk that was then sintered for 5
lens.94 For the TEM studies, the PtNi nanoparticles were min at 110 °C to ensure evaporation of the Nafion solvents,
dispersed in dry conditions on holey film copper TEM grids yielding a loading of ca. 20 μgPt cm−2geo. Prior to any
(Agar Scientific Ltd.). The TEM investigations were under- electrochemical experiment, the working electrode was
taken on nanoparticles lying over the holes of the thin holey immersed into the deaerated electrolyte at E = 0.40 V vs
carbon films to eliminate the contribution of the latter to the RHE (Ar >99.999%, Messer). One hundred cyclic voltammo-
HR-TEM and EELS signals. grams were then recorded in Ar-saturated electrolyte between
For spatially resolved EELS, the microscope was operated in 0.05 and 1.23 V vs RHE with a potential sweep rate of 0.500 V
STEM mode with the probe size setting “5c” and convergence s−1 (electrochemical activation), three cyclic voltammograms at
and collection angles set at 32 and 100 mrad, respectively. EEL v = 20 mV s−1 under the same potential conditions, and one
spectra were acquired using a Gatan Quantum ER imaging filter COads stripping measurement to determine the electrochemi-
fitted with a fast CCD camera with a selected energy dispersion cally active surface area. In the last experiment, the CO
of 0.5 eV per pixel and a vertical binning of ×26. To analyze the saturation coverage was established by bubbling CO for 6 min
Pt and Ni distribution within the nanoparticles, core-loss and purging with Ar for 34 min, while the electrode potential
spectra over a large energy domain (∼300−1500 eV) was kept at E = 0.1 V vs RHE. The real surface area was
containing both Pt-N3 (518 eV) and Ni-L2,3 (855 eV) were estimated by assuming that the electrooxidation of a COads
acquired in the spectrum-image (SI) mode. In this mode, the monolayer requires 420 μC cm−2 of Pt. The electrocatalytic
EELS data are captured sequentially in space by scanning in a activity for the ORR was measured in an O2-saturated 0.1 M
controlled way the focused electrons at the surface of the HClO4 solution (20 min of purging by oxygen >99.99%,
sample and acquiring the loss signal over the energy range of Messer) by linearly sweeping the potential from 0.20 to 1.05 V
interest at each probe position. In the present study, the core- vs RHE at a potential sweep rate of 5 mV s−1 and at different
loss EEL spectra were acquired over square or rectangular areas rotational speeds (400, 900, 1600, and 2500 rpm). The ORR
enclosing the nanoparticles with an acquisition time of 0.7−1 s specific/mass activity was determined by normalizing the
per spectrum. To compensate for spatial drift of the sample current measured at E = 0.95 V vs RHE, after correction
during EELS acquisition, a drift correction was applied at the from the oxygen diffusion in the solution and the Ohmic drop,
end of each row of the SI. Spectral drifts in the core-loss spectra to the real surface area/mass of deposited Pt determined by
were monitored and corrected by acquiring, in parallel to each COad stripping voltammetry, respectively.


core-loss SI, the corresponding SI of the zero-loss peak (with an
acquisition time of 1 μs per spectrum) using the dual EELS ASSOCIATED CONTENT
capability of the imaging filter. Elemental maps were
constructed by analyzing the intensities under the Pt-N3 and *
S Supporting Information

Ni-L2,3 edges after background subtraction using a power-law The Supporting Information is available free of charge on the
model. Prior to the generation of the elemental maps, each SI ACS Publications website at DOI: 10.1021/acscatal.6b01106.
4682 DOI: 10.1021/acscatal.6b01106
ACS Catal. 2016, 6, 4673−4684
ACS Catalysis Research Article

In situ TEM images showing the morphological and (20) Gavrilov, A. N.; Savinova, E. R.; Simonov, P. A.; Zaikovskii, V. I.;
structural changes during thermal annealing, local X-EDS Cherepanova, S. V.; Tsirlina, G. A.; Parmon, V. N. Phys. Chem. Chem.
Phys. 2007, 9, 5476−5489.
analyses, XPS, laboratory, and synchrotron XRD spectra,
(21) Savinova, E. R.; Hahn, F.; Alonso-Vante, N. J. Phys. Chem. C
and results of the Rietveld refinements and PDF analysis 2008, 112, 18521−18530.
(PDF) (22) Kuznetsov, A. N.; Simonov, P. A.; Zaikovskii, V. I.; Parmon, V.


N.; Savinova, E. R. J. Solid State Electrochem. 2013, 17, 1903−1912.
(23) Wang, G.; Takeguchi, T.; Muhamad, E. N.; Yamanaka, T.; Ueda,
AUTHOR INFORMATION
W. Int. J. Hydrogen Energy 2011, 36, 3322−3332.
Corresponding Authors (24) Ma, Y.; Wang, H.; Ji, S.; Linkov, V.; Wang, R. J. Power Sources
*E-mail for L.D.: laetitia.dubau@lepmi.grenoble-inp.fr. 2014, 247, 142−150.
*E-mail for F.M.: frederic.maillard@lepmi.grenoble-inp.fr. (25) Macia, M. D.; Campina, J. M.; Herrero, E.; Feliu, J. M. J.
Electroanal. Chem. 2004, 564, 141−150.
Notes (26) Kuzume, A.; Herrero, E.; Feliu, J. M. J. Electroanal. Chem. 2007,
The authors declare no competing financial interest. 599, 333−343.


(27) Hitotsuyanagi, A.; Nakamura, M.; Hoshi, N. Electrochim. Acta
ACKNOWLEDGMENTS 2012, 82, 512−516.
(28) Gómez-Marín, A. M.; Feliu, J. M. J. Solid State Electrochem.
This work was performed within the framework of the Centre 2015, 19, 2831−2841.
of Excellence of Multifunctional Architectured Materials (29) Gómez-Marín, A. M.; Rizo, R.; Feliu, J. M. Catal. Sci. Technol.
“CEMAM” No. ANR-10-LABX-44-01. The authors acknowl- 2014, 4, 1685−1698.
edge the French CNRS and CEA-METSA network and (30) Bandarenka, A. S.; Hansen, H. A.; Rossmeisl, J.; Stephens, I. E.
financial support from the University of Grenoble-Alpes L. Phys. Chem. Chem. Phys. 2014, 16, 13625−13629.
through the AGIR program (Grant No. LL1492017G) and (31) Gloaguen, F.; Andolfatto, F.; Durand, R.; Ozil, P. J. Appl.
from the French National Research Agency through the Electrochem. 1994, 24, 863−869.
(32) Perez-Alonso, F. J.; McCarthy, D. N.; Nierhoff, A.; Hernandez-
HOLLOW project (Grant No. ANR-14-CE05-0003-01).


Fernandez, P.; Strebel, C.; Stephens, I. E. L.; Nielsen, J. H.;
Chorkendorff, I. Angew. Chem., Int. Ed. 2012, 51, 4641−4643.
REFERENCES (33) Lee, S. W.; Chen, S.; Suntivich, J.; Sasaki, K.; Adzic, R. R.; Shao-
(1) Somorjai, G. A. Chem. Rev. 1996, 96, 1223−1235. Horn, Y. J. Phys. Chem. Lett. 2010, 1, 1316−1320.
(2) Calle-Vallejo, F.; Martinez, J. I.; Garcia-Lastra, J. M.; Sautet, P.; (34) Xia, B. Y.; Wu, H. B.; Wang, X.; Lou, X. W. Angew. Chem., Int.
Loffreda, D. Angew. Chem., Int. Ed. 2014, 53, 8316−8319. Ed. 2013, 52, 12337−12340.
(3) Calle-Vallejo, F.; Tymoczko, J.; Colic, V.; Vu, Q. H.; Pohl, M. D.; (35) Wang, S.; Jiang, S. P.; White, T. J.; Guo, J.; Wang, X. J. Phys.
Morgenstern, K.; Loffreda, D.; Sautet, P.; Schuhmann, W.; Chem. C 2009, 113, 18935−18945.
Bandarenka, A. S. Science 2015, 350, 185−189. (36) Nesselberger, M.; Roefzaad, M.; Fayçal Hamou, R.; Ulrich
(4) Quan, Z.; Wang, Y.; Fang, J. Acc. Chem. Res. 2013, 46, 191−202. Biedermann, P.; Schweinberger, F. F.; Kunz, S.; Schloegl, K.; Wiberg,
(5) Herrero, E.; Á lvarez, B.; Feliu, J. M.; Blais, S.; Radovic-Hrapovic, G. K. H.; Ashton, S.; Heiz, U.; Mayrhofer, K. J. J.; Arenz, M. Nat.
Z.; Jerkiewicz, G. J. Electroanal. Chem. 2004, 567, 139−149. Mater. 2013, 12, 919−924.
(6) Lebedeva, N. P.; Koper, M. T. M.; Herrero, E.; Feliu, J. M.; Van (37) Speder, J.; Zana, A.; Arenz, M. Catal. Today 2016, 262, 82−89.
Santen, R. A. J. Electroanal. Chem. 2000, 487, 37−44. (38) Speder, J.; Spanos, I.; Zana, A.; Kirkensgaard, J. J. K.;
(7) Lebedeva, N. P.; Koper, M. T. M.; Feliu, J. M.; van Santen, R. A. Mortensen, K.; Altmann, L.; Bäumer, M.; Arenz, M. Surf. Sci. 2015,
J. Phys. Chem. B 2002, 106, 12938−12947. 631, 278−284.
(8) Strmcnik, D. S.; Tripkovic, D. V.; van der Vliet, D.; Chang, K. C.; (39) Speder, J.; Altmann, L.; Bäumer, M.; Kirkensgaard, J. J. K.;
Komanicky, V.; You, H.; Karapetrov, G.; Greeley, J.; Stamenkovic, V. Mortensen, K.; Arenz, M. RSC Adv. 2014, 4, 14971−14978.
R.; Markovic, N. M. J. Am. Chem. Soc. 2008, 130, 15332−15339. (40) Callejas-Tovar, R.; Balbuena, P. B. J. Phys. Chem. C 2012, 116,
(9) Maillard, F.; Eikerling, M.; Cherstiouk, O. V.; Schreier, S.; 14414−14422.
Savinova, E.; Stimming, U. Faraday Discuss. 2004, 125, 357−377. (41) Wang, J. X.; Ma, C.; Choi, Y.; Su, D.; Zhu, Y.; Liu, P.; Si, R.;
(10) Maillard, F.; Schreier, S.; Hanzlik, M.; Savinova, E. R.; Weinkauf, Vukmirovic, M. B.; Zhang, Y.; Adzic, R. R. J. Am. Chem. Soc. 2011, 133,
S.; Stimming, U. Phys. Chem. Chem. Phys. 2005, 7, 385−393. 13551−13557.
(11) Maillard, F.; Savinova, E. R.; Stimming, U. J. Electroanal. Chem. (42) Lopez-Haro, M.; Dubau, L.; Guétaz, L.; Bayle-Guillemaud, P.;
2007, 599, 221−232. Chatenet, M.; André, J.; Caqué, N.; Rossinot, E.; Maillard, F. Appl.
(12) Mayrhofer, K. J. J.; Arenz, M.; Blizanac, B. B.; Stamenkovic, V. Catal., B 2014, 152−153, 300−308.
R.; Ross, P. N.; Markovic, N. M. Electrochim. Acta 2005, 50, 5144− (43) Dubau, L.; Lopez-Haro, M.; Durst, J.; Guetaz, L.; Bayle-
5154. Guillemaud, P.; Chatenet, M.; Maillard, F. J. Mater. Chem. A 2014, 2,
(13) Andreaus, B.; Maillard, F.; Kocylo, J.; Savinova, E. R.; Eikerling, 18497−18507.
M. J. Phys. Chem. B 2006, 110, 21028−21040. (44) Dubau, L.; Asset, T.; Chattot, R.; Bonnaud, C.; Vanpeene, V.;
(14) Cherstiouk, O. V.; Simonov, P. A.; Zaikovskii, V. I.; Savinova, E. Nelayah, J.; Maillard, F. ACS Catal. 2015, 5, 5333.
R. J. Electroanal. Chem. 2003, 554-555, 241−251. (45) Dubau, L.; Lopez-Haro, M.; Durst, J.; Maillard, F. Catal. Today
(15) Rhee, C. K.; Kim, B. J.; Ham, C.; Kim, Y. J.; Song, K.; Kwon, K. 2016, 262, 146−154.
Langmuir 2009, 25, 7140−7147. (46) Bae, S. J.; Yoo, S. J.; Lim, Y.; Kim, S.; Lim, Y.; Choi, J.; Nahm, K.
(16) Napolskii, K. S.; Barczuk, P. J.; Vassiliev, S. Y.; Veresov, A. G.; S.; Hwang, S. J.; Lim, T. H.; Kim, S. K.; Kim, P. J. Mater. Chem. 2012,
Tsirlina, G. A.; Kulesza, P. J. Electrochim. Acta 2007, 52, 7910−7919. 22, 8820.
(17) Lebedeva, N. P.; Kryukova, G. N.; Tsybulya, S. V.; Salanov, A. (47) Shan, A.; Chen, Z.; Li, B.; Chen, C.; Wang, R. J. Mater. Chem. A
N.; Savinova, E. R. Electrochim. Acta 1998, 44, 1431−1440. 2015, 3, 1031−1036.
(18) Sieben, J. M.; Duarte, M. M. E. Int. J. Hydrogen Energy 2011, 36, (48) Kirkendall, E.; Thomassen, L.; Uethegrove, C. Trans. AIME
3313−3321. 1939, 133, 186−203.
(19) Maillard, F.; Bonnefont, A.; Chatenet, M.; Guétaz, L.; Doisneau- (49) Kirkendall, E. O. Trans. AIME 1942, 147, 104−109.
Cottignies, B.; Roussel, H.; Stimming, U. Electrochim. Acta 2007, 53, (50) Smigelskas, A. D.; Kirkendall, E. O. Trans. AIME 1947, 171,
811−822. 130−142.

4683 DOI: 10.1021/acscatal.6b01106


ACS Catal. 2016, 6, 4673−4684
ACS Catalysis Research Article

(51) Yin, Y. D.; Rioux, R. M.; Erdonmez, C. K.; Hughes, S.; Somorjai, (80) Dubau, L.; Maillard, F.; Chatenet, M.; André, J.; Rossinot, E.
G. A.; Alivisatos, A. P. Science 2004, 304, 711−714. Electrochim. Acta 2010, 56, 776−783.
(52) Mehrer, H. Diffusion in solids: fundamentals, methods, materials, (81) Dubau, L.; Durst, J.; Maillard, F.; Guétaz, L.; Chatenet, M.;
diffusion-controlled processes; Springer-Verlag: Berlin and Heidelberg, André, J.; Rossinot, E. Electrochim. Acta 2011, 56, 10658−10667.
Germany, 2007; p 651. (82) Oezaslan, M.; Heggen, M.; Strasser, P. J. Am. Chem. Soc. 2012,
(53) Fan, H. J.; Gosele, U.; Zacharias, M. Small 2007, 3, 1660−1671. 134, 514−524.
(54) Wang, C.; Wang, G.; Van Der Vliet, D.; Chang, K. C.; Markovic, (83) Dubau, L.; Lopez-Haro, M.; Castanheira, L.; Durst, J.; Chatenet,
N. M.; Stamenkovic, V. R. Phys. Chem. Chem. Phys. 2010, 12, 6933− M.; Bayle-Guillemaud, P.; Guétaz, L.; Caqué, N.; Rossinot, E.;
6939. Maillard, F. Appl. Catal., B 2013, 142−143, 801−808.
(55) Stamenkovic, V. R.; Mun, B. S.; Arenz, M.; Mayrhofer, K. J. J.; (84) Gan, L.; Heggen, M.; O’Malley, R.; Theobald, B.; Strasser, P.
Lucas, C. A.; Wang, G. F.; Ross, P. N.; Markovic, N. M. Nat. Mater. Nano Lett. 2013, 13, 1131−1138.
2007, 6, 241−247. (85) Gan, L.; Cui, C.; Rudi, S.; Strasser, P. Top. Catal. 2014, 57,
(56) Ahmadi, M.; Behafarid, F.; Cui, C.; Strasser, P.; Cuenya, B. R. 236−244.
ACS Nano 2013, 7, 9195−9204. (86) Baldizzone, C.; Gan, L.; Hodnik, N.; Keeley, G. P.; Kostka, A.;
(57) Chung, Y.-H.; Kim, S. J.; Chung, D. Y.; Lee, M. J.; Jang, J. H.; Heggen, M.; Strasser, P.; Mayrhofer, K. J. J. ACS Catal. 2015, 5, 5000−
Sung, Y.-E. Phys. Chem. Chem. Phys. 2014, 16, 13726−13732. 5007.
(58) Gan, L.; Heggen, M.; Cui, C.; Strasser, P. ACS Catal. 2016, 6, (87) Caldwell, K. M.; Ramaker, D. E.; Jia, Q.; Mukerjee, S.;
692−695. Ziegelbauer, J. M.; Kukreja, R. S.; Kongkanand, A. J. Phys. Chem. C
(59) Gusak, A. M.; Zaporozhets, T. V.; Tu, K. N.; Gosele, U. Philos. 2015, 119, 757−765.
Mag. 2005, 85, 4445−4464. (88) Jia, Q.; Caldwell, K.; Strickland, K.; Ziegelbauer, J. M.; Liu, Z.;
Yu, Z.; Ramaker, D. E.; Mukerjee, S. ACS Catal. 2015, 5, 176−186.
(60) Besenbacher, F.; Lorensen, H. T.; Helveg, S.; Laegsgaard, E.;
(89) Han, B.; Carlton, C. E.; Kongkanand, A.; Kukreja, R. S.;
Stensgaard, I.; Jacobsen, K. W.; Norskov, J. K.; Besenbacher, F. Nature
Theobald, B. R.; Gan, L.; O’Malley, R.; Strasser, P.; Wagner, F. T.;
1999, 398, 134−136.
Shao-Horn, Y. Energy Environ. Sci. 2015, 8, 258−266.
(61) Esch, S.; Hohage, M.; Michely, T.; Comsa, G. Phys. Rev. Lett.
(90) Rodríguez-Carvajal, J. Phys. B 1993, 192, 55−69.
1994, 72, 518−521. (91) Thompson, P.; Cox, D. E.; Hastings, J. B. J. Appl. Crystallogr.
(62) Dubau, L.; Castanheira, L.; Berthomé, G.; Maillard, F. 1987, 20, 79−83.
Electrochim. Acta 2013, 110, 273−281. (92) Juhas, P.; Davis, T.; Farrow, C. L.; Billinge, S. J. L. J. Appl.
(63) Cabie, M.; Giorgio, S.; Henry, C. R.; Axet, M. R.; Philippot, K.; Crystallogr. 2013, 46, 560−566.
Chaudret, B. J. Phys. Chem. C 2010, 114, 2160−2163. (93) Farrow, C. L.; Juhas, P.; Liu, J. W.; Bryndin, D.; Božin, E. S.;
(64) Hansen, P. L.; Wagner, J. B.; Helveg, S.; Rostrup-Nielsen, J. R.; Bloch, J.; Th, P.; Billinge, S. J. L. J. Phys.: Condens. Matter 2007, 19,
Clausen, B. S.; Topsøe, H. Science 2002, 295, 2053−2055. 335219.
(65) Nolte, P.; Stierle, A.; Jin-Phillipp, N. Y.; Kasper, N.; Schulli, T. (94) Ricolleau, C.; Nelayah, J.; Oikawa, T.; Kohno, Y.; Braidy, N.;
U.; Dosch, H. Science 2008, 321, 1654−1658. Wang, G.; Hue, F.; Florea, L.; Pierron Bohnes, V.; Alloyeau, D.
(66) Mittendorfer, F.; Seriani, N.; Dubay, O.; Kresse, G. Phys. Rev. B: Microscopy 2013, 62, 283−293.
Condens. Matter Mater. Phys. 2007, 76, 233413. (95) Detailed information about the hyperSpy analysis toolbox can
(67) Wang, C.; Chi, M.; Li, D.; Strmcnik, D.; Van Der Vliet, D.; be found at http://hyperspy.org/ (accessed June 3, 2016).
Wang, G.; Komanicky, V.; Chang, K. C.; Paulikas, A. P.; Tripkovic, D.; (96) Herrmann, C. C.; Perrault, G. G.; Pilla, A. A. Anal. Chem. 1968,
Pearson, J.; More, K. L.; Markovic, N. M.; Stamenkovic, V. R. J. Am. 40, 1173−1174.
Chem. Soc. 2011, 133, 14396−14403.
(68) Wang, C.; Chi, M.; Li, D.; van der Vliet, D.; Wang, G.; Lin, Q.;
Mitchell, J. F.; More, K. L.; Markovic, N. M.; Stamenkovic, V. R. ACS
Catal. 2011, 1, 1355−1359.
(69) Durst, J.; Lopez-Haro, M.; Dubau, L.; Chatenet, M.; Soldo-
Olivier, Y.; Guétaz, L.; Bayle-Guillemaud, P.; Maillard, F. J. Phys. Chem.
Lett. 2014, 5, 434−439.
(70) Cui, C.; Ahmadi, M.; Behafarid, F.; Gan, L.; Neumann, M.;
Heggen, M.; Cuenya, B. R.; Strasser, P. Faraday Discuss. 2013, 162,
91−112.
(71) Menning, C. A.; Chen, J. G. J. Power Sources 2010, 195, 3140−
3144.
(72) Xin, H. L.; Alayoglu, S.; Tao, R.; Genc, A.; Wang, C. M.;
Kovarik, L.; Stach, E. A.; Wang, L. W.; Salmeron, M.; Somorjai, G. A.;
Zheng, H. Nano Lett. 2014, 14, 3203−3207.
(73) Shao-Horn, Y.; Sheng, W.; Chen, S.; Ferreira, P.; Holby, E.;
Morgan, D. Top. Catal. 2007, 46, 285−305.
(74) Cui, C. H.; Gan, L.; Heggen, M.; Rudi, S.; Strasser, P. Nat.
Mater. 2013, 12, 765−771.
(75) Strasser, P.; Koh, S.; Anniyev, T.; Greeley, J.; More, K.; Yu, C.;
Liu, Z.; Kaya, S.; Nordlund, D.; Ogasawara, H.; Toney, M. F.; Nilsson,
A. Nat. Chem. 2010, 2, 454−460.
(76) Jia, Q.; Li, J.; Caldwell, K.; Ramaker, D. E.; Ziegelbauer, J. M.;
Kukreja, R. S.; Kongkanand, A.; Mukerjee, S. ACS Catal. 2016, 6,
928−938.
(77) Hammer, B.; Morikawa, Y.; Nørskov, J. K. Phys. Rev. Lett. 1996,
76, 2141−2144.
(78) Hammer, B.; Nørskov, J. K. Surf. Sci. 1995, 343, 211−220.
(79) Mavrikakis, M.; Hammer, B.; Nørskov, J. K. Phys. Rev. Lett.
1998, 81, 2819−2822.

4684 DOI: 10.1021/acscatal.6b01106


ACS Catal. 2016, 6, 4673−4684

You might also like