16
Physical Chemistry Chemical Physics This paper is published as part of a PCCP Themed Issue on: Electrocatalysis: theory and experiment at the interface Guest Editor: Andrea Russell Editorial Electrocatalysis: theory and experiment at the interface Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b808799g Papers Structure of the water/platinum interface––a first principles simulation under bias potential Minoru Otani, Ikutaro Hamada, Osamu Sugino, Yoshitada Morikawa, Yasuharu Okamoto and Tamio Ikeshoji, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b803541e A first-principles study of molecular oxygen dissociation at an electrode surface: a comparison of potential variation and coadsorption effects Sally A. Wasileski and Michael J. Janik, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b803157f Electrochemical and FTIRS characterisation of NO adlayers on cyanide-modified Pt(111) electrodes: the mechanism of nitric oxide electroreduction on Pt Angel Cuesta and María Escudero, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b717396b Electrodeposited noble metal particles in polyelectrolyte multilayer matrix as electrocatalyst for oxygen reduction studied using SECM Yan Shen, Markus Träuble and Gunther Wittstock, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b802688b CoPt nanoparticles and their catalytic properties in electrooxidation of CO and CH 3 OH studied by in situ FTIRS Qing-Song Chen, Shi-Gang Sun, Zhi-You Zhou, Yan-Xin Chen and Shi-Bin Deng, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b802047g Kinetic studies of adsorbed CO electrochemical oxidation on Pt(335) at full and sub-saturation coverages Prachak Inkaew and Carol Korzeniewski, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b804507k Formic acid electrooxidation on Pd in acidic solutions studied by surface-enhanced infrared absorption spectroscopy Hiroto Miyake, Tatsuhiro Okada, Gabor Samjeské and Masatoshi Osawa, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b805955a Voltammetric surface dealloying of Pt bimetallic nanoparticles: an experimental and DFT computational analysis Peter Strasser, Shirlaine Koh and Jeff Greeley, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b803717e Published on 27 May 2008. Downloaded by TU Berlin - Universitaetsbibl on 13/04/2015 13:24:20. View Article Online / Journal Homepage / Table of Contents for this issue

Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

Physical Chemistry Chemical Physics

This paper is published as part of a PCCP Themed Issue on:

Electrocatalysis: theory and experiment at the interface

Guest Editor: Andrea Russell

Editorial

Electrocatalysis: theory and experiment at the interface Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b808799g

Papers

Structure of the water/platinum interface––a first principles simulation under bias potential Minoru Otani, Ikutaro Hamada, Osamu Sugino, Yoshitada Morikawa, Yasuharu Okamoto and Tamio Ikeshoji, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b803541e

A first-principles study of molecular oxygen dissociation at an electrode surface: a comparison of potential variation and coadsorption effects Sally A. Wasileski and Michael J. Janik, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b803157f

Electrochemical and FTIRS characterisation of NO adlayers on cyanide-modified Pt(111) electrodes: the mechanism of nitric oxide electroreduction on Pt Angel Cuesta and María Escudero, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b717396b

Electrodeposited noble metal particles in polyelectrolyte multilayer matrix as electrocatalyst for oxygen reduction studied using SECM Yan Shen, Markus Träuble and Gunther Wittstock, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b802688b

CoPt nanoparticles and their catalytic properties in electrooxidation of CO and CH3OH studied by in situ FTIRS Qing-Song Chen, Shi-Gang Sun, Zhi-You Zhou, Yan-Xin Chen and Shi-Bin Deng, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b802047g

Kinetic studies of adsorbed CO electrochemical oxidation on Pt(335) at full and sub-saturation coverages Prachak Inkaew and Carol Korzeniewski, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b804507k

Formic acid electrooxidation on Pd in acidic solutions studied by surface-enhanced infrared absorption spectroscopy Hiroto Miyake, Tatsuhiro Okada, Gabor Samjeské and Masatoshi Osawa, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b805955a

Voltammetric surface dealloying of Pt bimetallic nanoparticles: an experimental and DFT computational analysis Peter Strasser, Shirlaine Koh and Jeff Greeley, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b803717e

Publ

ishe

d on

27

May

200

8. D

ownl

oade

d by

TU

Ber

lin -

Uni

vers

itaet

sbib

l on

13/0

4/20

15 1

3:24

:20.

View Article Online / Journal Homepage / Table of Contents for this issue

Page 2: Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

Unique activity of Pd monomers: hydrogen evolution at AuPd(111) surface alloys Y. Pluntke, L. A. Kibler and D. M. Kolb, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b802915f

Shape-dependent electrocatalysis: methanol and formic acid electrooxidation on preferentially oriented Pt nanoparticles J. Solla-Gullón, F. J. Vidal-Iglesias, A. López-Cudero, E. Garnier, J. M. Feliu and A. Aldaz, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b802703j

The role of adsorbed hydroxyl species in the electrocatalytic carbon monoxide oxidation reaction on platinum Anthony R. Kucernak and Gregory J. Offer, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b802816h

An unexpected enhancement in methanol electro-oxidation on an ensemble of Pt(111) nanofacets: a case of nanoscale single crystal ensemble electrocatalysis Ceren Susut, George B. Chapman, Gabor Samjeské, Masatoshi Osawa and YuYe Tong, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b802708k

Surface Pourbaix diagrams and oxygen reduction activity of Pt, Ag and Ni(111) surfaces studied by DFT Heine A. Hansen, Jan Rossmeisl and Jens K. Nørskov, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b803956a

Electrocatalytic mechanism and kinetics of SOMs oxidation on ordered PtPb and PtBi intermetallic compounds: DEMS and FTIRS study Hongsen Wang, Laif Alden, F. J. DiSalvo and Héctor D. Abruña, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b801473f

Combinatorial screening of PtTiMe ternary alloys for oxygen electroreduction Ting He and Eric Kreidler, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b802818b

Electrochemical study on the adsorption of carbon oxides and oxidation of their adsorption products on platinum group metals and alloys Hanna Siwek, Mariusz Lukaszewski and Andrzej Czerwinski, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b718286b

Ethanol electrooxidation onto stepped surfaces modified by Ru deposition: electrochemical and spectroscopic studies V. Del Colle, A. Berná, G. Tremiliosi-Filho, E. Herrero and J. M. Feliu, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b802683a

The effects of the specific adsorption of anion on the reactivity of the Ru(0001) surface towards CO adsorption and oxidation: in situ FTIRS studies J. M. Jin, W. F. Lin and P. A. Christensen, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b802701c

Experimental and numerical model study of the limiting current in a channel flow cell with a circular electrode J. Fuhrmann, H. Zhao, E. Holzbecher, H. Langmach, M. Chojak, R. Halseid, Z. Jusys and J. Behm, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b802812p

Using layer-by-layer assembly of polyaniline fibers in the fast preparation of high performance fuel cell nanostructured membrane electrodes Marc Michel, Frank Ettingshausen, Frieder Scheiba, André Wolz and Christina Roth, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b802813n

Stripping voltammetry of carbon monoxide oxidation on stepped platinum single-crystal electrodes in alkaline solution Gonzalo García and Marc T. M. Koper, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b803503m

PtxRu1-x/Ru(0001) surface alloys—formation and atom distribution H. E. Hoster, A. Bergbreiter, P. M. Erne, T. Hager, H. Rauscher and R. J. Behm, Phys. Chem. Chem. Phys., 2008 DOI: 10.1039/b802169d

Publ

ishe

d on

27

May

200

8. D

ownl

oade

d by

TU

Ber

lin -

Uni

vers

itaet

sbib

l on

13/0

4/20

15 1

3:24

:20.

View Article Online

Page 3: Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

Voltammetric surface dealloying of Pt bimetallic nanoparticles:

an experimental and DFT computational analysis

Peter Strasser,w*ac Shirlaine Koha and Jeff Greeleyb

Received 4th March 2008, Accepted 15th May 2008

First published as an Advance Article on the web 27th May 2008

DOI: 10.1039/b803717e

Voltammetric dealloying of bimetallic platinum–copper (Pt–Cu) alloys has been shown to be an

effective strategy to modify the surface electrocatalytic reactivity of Pt bimetallic nanoparticles

(S. Koh and P. Strasser, J. Am. Chem. Soc., 2007, 129, 12624). Using cyclic voltammetry and

structural XRD studies, we systematically characterize the Pt–Cu precursor compounds as well as

the early stages of the selective Cu surface dissolution (dealloying) process for Pt25Cu75, Pt50Cu50,

and Pt75Cu25 alloy nanoparticles annealed at both low and high temperature. We also assess the

impact of the synthesis conditions on the electrocatalytic reactivity for the oxygen reduction

reaction (ORR). To gain atomistic insight into the observed voltammetric profiles, we compare

our experimental results with periodic DFT calculations of trends in the thermodynamics of

surface Cu dissolution potentials from highly stepped and kinked Pt(854) single crystal surfaces.

The modeling suggests a dependence of the electrochemical Cu dissolution potentials on the

detailed atomic environment (coordination number, nature of coordinating atoms) of the

bimetallic Pt–Cu surfaces. The DFT-predicted shifts in electrochemical Cu dissolution potentials

are shown to qualitatively account for the observed voltammetric profiles during Cu dealloying.

Our study suggests that metal-specific energetics have to be taken into account to explain the

detailed dealloying behavior of bimetallic surfaces.

1. Introduction

Over 2000 years ago, pre-Columbian goldsmiths of the high

cultures of Andean South America developed simple surface

processing technologies in order to change the properties of

metal alloys.2 They were interested in changing the appearance

of Au-containing metal alloys to make them appear as if they

consisted of pure gold. To achieve this, they invented two

techniques: surface metal galvanic displacement and surface

metal depletion gilding.2,3

Surface metal galvanic displacement consisted of treating

thin polished Cu metal sheets in neutralized heated baths of

Au salts, which were prepared from metallic Au by dissolution

in corrosive mineral mixtures.2 The Cu sheets were uniformly

coated with a thin layer of Au and appear as if they consisted

of pure Au: an affordable strategy for impoverished royal

families to dress in gold, which, in the then cultural attitude,

represented descent from the sun.

The Surface metal depletion technology started with bime-

tallic ingots of a less noble (Cu) and a more noble metal (Au),

known as ‘tumbaga’, which were repeatedly subjected to cold-

treatments, annealing, and pickling cycles in order to enrich

the surface with copper and then deplete surface copper by

chemical dissolution. In this process, the surface of the bi-

metallic ingot became more and more enriched in Au until it

displayed a golden surface.

Both ancient metallurgic strategies for altering surface

properties have found their way into modern electrochemical

metal surface processing.3–7 While visual appearance was the

target property of interest then, the surface electronic struc-

ture, and with it, the surface catalytic properties, are the focus

of the modifications today.

In order to change the surface electrocatalytic properties of

Pt metal alloys, a modern galvanic displacement method has

been developed and utilized over the past decade,8–12 which

came to be known as the Pt Monolayer Catalyst9 concept.

Following the ancient goldsmiths’ strategy, Pt monolayers

were deposited onto non-Pt substrates using galvanic displace-

ment of layers of surface Cu atoms (Fig. 1a). The resulting

catalysts have shown improved intrinsic surface electrocataly-

tic activity for the electroreduction of oxygen at much reduced

Pt content; they represent one of the most promising new

catalyst classes for fuel cell cathodes. Vukmiovic et al.8 re-

viewed the Pt monolayer concept in detail and pointed out its

structural and catalytic characteristics.

Herein, we report an experimental and DFT computational

study of a modern voltammetric depletion gilding method

designed to deliberately change the electrocatalytic surface

reactivity of Pt alloy catalysts for the electroreduction of

oxygen (ORR).1,13–16 We selectively dissolve Cu atoms (vol-

tammetric dealloying) from the surface of Pt–Cu bimetallic

compounds until a multi-layer Pt-rich shell region has formed

(Fig. 1b). While the dealloying of a less noble component from

aDepartment of Chemical and Biomolecular Engineering, Universityof Houston, Houston, Texas 77204, USA. E-mail: [email protected]

bCenter for Nanoscale Materials, Argonne National Laboratory,Argonne, IL 60439, USA

c Institut fur Chemie, Technische Universitat Berlin, 10624 Berlin,Germany. E-mail: [email protected] Secondary contact address: , , ,

3670 | Phys. Chem. Chem. Phys., 2008, 10, 3670–3683 This journal is �c the Owner Societies 2008

PAPER www.rsc.org/pccp | Physical Chemistry Chemical Physics

Publ

ishe

d on

27

May

200

8. D

ownl

oade

d by

TU

Ber

lin -

Uni

vers

itaet

sbib

l on

13/0

4/20

15 1

3:24

:20.

View Article Online

Page 4: Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

bimetallic alloys has been investigated on flat surfaces, such as

alloy single crystals, we focus on the initial stages of Cu

dealloying from nanometer-sized alloy particle ensembles.

Metal dealloying from alloy nanoparticles is a poorly under-

stood research area, yet it has tremendous importance for the

understanding of high-surface area alloy catalyst degradation

in electrocatalysis.

We found Cu dealloying from Pt–Cu bimetallic alloy nano-

particle precursors to result in very active ORR catalysts.1,17

We investigate the voltammetric profiles of Cu dissolution

during the early stages of the dealloying process and correlate

the voltammetry with structural alloy phase characteristics at

various Cu contents and synthesis conditions. We combine our

experimental results with computational DFT predictions of

trends in the shift in thermodynamic Cu dissolution potentials

in Pt–Cu bimetallic surfaces. In our calculations, we consider

metal-specific energetics and assess their impact on the Cu

dissolution potentials.

2. Experimental methods

Preparation of the alloy catalysts

Carbon supported Pt–Cu alloy nanoparticle electrocatalysts of

stoichiometries Pt25Cu75, Pt50Cu50, and Pt75Cu25 were synthe-

sized via a liquid salt precursor impregnation method reported

earlier. Measured amounts of a Cu nitrate (Cu(NO3)2 3H2O,

Aldrich) were dissolved in de-ionized water (418.2 MO, Milli-

Qs gradient system, Millipore Inc.). The precursor solutions

were then added to a weighted amount of commercial carbon-

supported Pt nanoparticle electrocatalyst (Tanaka Kikinzoku

Inc.) with a Pt weight loading of about 30 wt%. The mixtures

were then ultrasonicated with a sonifier horn (Branson 150,

level 2) for 1 min and then frozen in liquid nitrogen for 5 min.

Subsequently, the frozen samples were freeze-dried (Labcon-

co) for 24 h and annealed in flowing hydrogen/Argon mixtures

in a flow furnace (Lindberg Blue) to a temperature of 600 and

950 1C. The annealed powders were stored in a Nalgenes

desiccator, with continuous nitrogen gas flow, for future use.

The resulting Pt loadings of the Pt–Cu catalysts ranged from

22 wt% to about 30 wt%.

Electrochemical dealloying and ORR testing

A custom-made, three-compartment electrochemical cell was

employed for all voltammetric measurements. The working

electrode was a commercial glassy carbon rotating disk elec-

trode of 5 mm diameter (PINE Instruments). A catalyst ink

was prepared by mixing 20 mg catalyst powder with 16 ml

water (Millipore Gradient), 4 ml isopropanol (IPA) and 80 mlof a 5 wt% Nafions solution (Sigma, #274704) , followed by

sonication for 15 min at level 2 (Branson 150). A 10 ml aliquotwas dispensed onto the 5 mm working electrode disk resulting

in a typical geometric Pt loading of about 10 mg Pt/cm2. The

applied absolute amount of Pt of the catalyst thin films on the

RDE varies with stoichiometry of the electrocatalysts and

decreased in the order Pt 4 Pt75Cu25 4 Pt50Cu50 4 Pt25Cu75.

The reference electrode was a mercury–mercury sulfate

electrode. All electrode potentials were subsequently con-

verted into and are reported with respect to the reversible

hydrogen electrode (RHE) scale. The counter electrode was a

piece of platinum gauze to ensure large surface area. A

commercial rotator from Pine Instrument was used to conduct

the rotating disk experiment. The electrolyte used was 0.1 M

HClO4, prepared by diluting 70% redistilled HClO4 (Sigma

#311421) with de-ionized water. The disk potential was con-

trolled with a potentiostat, BiStat (Princeton Applied Re-

search, Ametek). All measurements were conducted at room

temperature. At the beginning of electrochemical measure-

ments, electrocatalysts were immersed into the electrolyte

under potential control and held at 0.05 V/RHE until the

measurements commenced. Dealloying was conducted in a de-

aerated N2 atmosphere electrolyte. The initial three cyclic

voltammetric cycles (CVs) were recorded at a scan rate of

100 mV s�1 between 0.05 V and 1.2 V. Thereafter, dealloying

was continued for 200 CV cycles over the same potential range

at a scan rate of 500 mV s�1, until a stable CV was obtained.

Subsequently, another slow CV was recorded at 100 mV s�1.

The platinum electrochemical surface area (Pt-ECSA) of the

catalyst was determined from that latter CV via the mean

integral charge of the hydrogen adsorption and desorption

areas after double-layer correction, using 210 mC cm2Pt as the

conversion factor. ORR measurements were conducted by

linear sweep voltammetry (LSV) from 0.05 V to the open

circuit potential (around 1.0 V) at the scan rate of 5 mV s�1. Pt

surface area normalized activities were established at 900

mV/[RHE], at room temperature by correcting the measured

current for mass-transport effects18–20 and dividing the result-

ing current by the Pt-ECSA. The electrochemical behavior

Fig. 1 Two distinct ancient pre-Columbian metallurgical techniques

for modifying the optical surface properties of noble metal/non-noble

metal bimetallics serve as the basis of two of the most promising

synthetic strategies towards more active oxygen reduction reaction

(ORR) electrocatalysts. (a) The ancient method of ‘Galvanic displace-

ment’ has been used to create Pt monolayer nanoparticles. (b) ancient

‘depletion gilding’ is the basis of a voltammetric dealloying technique

to modify the surface catalytic properties of bimetallic nanoparticles.

This journal is �c the Owner Societies 2008 Phys. Chem. Chem. Phys., 2008, 10, 3670–3683 | 3671

Publ

ishe

d on

27

May

200

8. D

ownl

oade

d by

TU

Ber

lin -

Uni

vers

itaet

sbib

l on

13/0

4/20

15 1

3:24

:20.

View Article Online

Page 5: Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

(CV and LSV) of the Pt–Cu catalysts was compared to a

30 wt% reference platinum electrocatalyst supported on a high

surface area carbon support.

X-Ray diffraction measurements

Laboratory-source XRD (Siemens D5000 y/2y Diffracto-

meter) was used to characterize the electrocatalysts structu-

rally. The diffractrometer is equipped with a Braun Position

Sensitive Detector (PSD) with an angular range of 81. The Cu

Ka source operates at a potential of 35 kV and current of

30 mA. 2y diffraction angles ranged from 20–701, scanned with

step size of 0.021 per step and holding time of 10–30 s per step.

Advanced X-ray Solution (X-ray commander, Bruker AXS)

software was used to control the diffractometer from a desk-

top computer.

The XRD sample holder was a custom-made 3 � 3 cm

plexiglass with a 1 cm width � 2 cm length � 1 mm depth well

in the center that hold the powder catalyst samples. The

catalyst powder was poured into the well and carefully flat-

tened to form a smooth surface, flushed with the surface of the

plexiglass.

All XRD diffraction patterns were analyzed using Jade

(MDI). Each peak profile of each reflection was obtained by

a non-linear least square fit of the Ka2 and background

corrected data. Instrumental broadening was determined using

an alumina standard under identical measurement conditions.

Particle sizes were estimated using the observed line broad-

ening and Scherrer equation at several fundamental reflec-

tions. Compositions of Pt–Cu alloys obeyed Vegard’s law21,22

which states that the composition of an alloy is directly

proportional to the lattice constant of the alloy with respect

to those of its components.

3. Theoretical and DFT computational models and

methods

The DACAPO code23 was used for all total energy calcula-

tions in this study. The Pt(854) surface was modeled with a

40-atom unit cell periodically repeated in a super cell geometry

with at least 17 A of vacuum between any two successive metal

slabs; the density functional theory-optimized Pt lattice con-

stant of 4.02 A is in reasonable agreement with the experi-

mental value of 3.91 A.24 The unit cell geometry is such that

either three or four metal layers (depending upon the location

on the slab surface and the number of surface atoms that have

dissolved) perpendicular to the (111) terraces are employed;

the top layer of Pt atoms is substituted with Cu, and Cu atoms

are sequentially removed to simulate Cu dissolution (see also

section 4.4). The calculations are fully converged with respect

to the number of metal layers included; removal of the bottom

layer of Pt atoms, and subsequent recalculation of the differ-

ential binding energies of surface Cu atoms at high coverages

(40.69 ML), shows average differential BE changes of less

than 0.02 eV (the actual differential BE changes calculated

were 0.011, �0.006, 0.024, and 0.037 eV for dissolution from

slabs with 1, 0.92, 0.85, and 0.77 ML of adsorbed Cu atoms,

respectively); these values are well within the accuracy range

associated with standard DFT calculations. The top (Cu) layer

of the slabs is relaxed until the total force on all atoms is less

than 0.04 eV A�1 in any Cartesian direction. Interslab dipole

corrections are not included in the reported results.13,25 Ionic

cores were described by ultrasoft pseudopotentials,25 and the

Kohn–Sham one-electron valence states were expanded in a

basis of plane waves with kinetic energy below 340 eV; a

density cutoff of 500 eV was used. The surface Brillouin zone

was sampled with a Chadi-Cohen 6(1�1) k point grid. The

convergence of the total energy with respect to the cutoff

energies and the k point set was confirmed. The exchange–

correlation energy and potential were described by the general-

ized gradient approximation (GGA-RPBE).23 The self-consis-

tent RPBE density was determined by iterative

diagonalization of the Kohn–Sham Hamiltonian, Fermi

population of the Kohn–Sham states (kBT ¼ 0.1 eV), and

Pulay mixing of the resulting electronic density.26 All total

energies were extrapolated to kBT ¼ 0 eV.

The approach for modeling the thermodynamic aspects of

metal dissolution on surfaces at standard ion concentrations

has been described previously;27 briefly, the raw, DFT-derived

total energies are corrected for electrode potential effects; the

latter effects are estimated by referencing the potential to the

standard hydrogen electrode. The dissolution potential of bulk

copper,24 U1, is taken as an empirical reference point, and all

shifts in dissolution potential due to changes in the surface

chemical environment are referenced to this value. The result-

ing expression for the total free energy of the Cu atoms on the

Pt(854) substrate, neglecting surface entropy effects, is:

GðNCu;UÞ ¼ Eslab;NCu� Eclean � ðEbulk;Cu þ 2ðU��UÞÞNCu;

ð1Þ

where NCu is the total number of Cu atoms per unit cell on the

Pt(854) surface, Eslab;NCuis the total, DFT-derived energy of

the slab with adsorbed Cu atoms, Eclean is the Pt(854) energy

with no adsorbed Cu, and Ebulk,Cu is the cohesive energy of

bulk Cu. Implicit in this expression is the assignment of a value

of 0 to the free energy of the clean Pt slab (with no adsorbed

Cu atoms). In all cases, Eslab;NCurefers to the most stable

arrangement of Cu atoms at the indicated value of NCu; test

calculations demonstrate that these most stable arrangements

are associated with removal of Cu atoms from the Cu kinks

(removal of Cu atoms from terraces or step edges is less

energetically favorable). We note that no corrections for

potential-induced electric field effects are included in the

calculated free energies; such effects have been shown to have

a negligible impact on the thermodynamics in computational

studies of the oxygen reduction reaction.28 Finally, in eqn (1),

it is assumed that two electrons are transferred to the substrate

for every copper ion that dissolves; this choice of this charge

state is motivated by the fact that Cu21 ions dissolve at lower

potentials than Cu1 ions.24,29

4. Results

4.1 ORR electrocatalysis of dealloyed Pt–Cu bimetallic alloy

nanoparticles

Six Pt–Cu bimetallic precursor alloy particle catalysts—three

different stoichiometries (Pt25Cu75, Pt50Cu50, Pt75Cu25) at

two different annealing temperatures (950 and 600 1C)

3672 | Phys. Chem. Chem. Phys., 2008, 10, 3670–3683 This journal is �c the Owner Societies 2008

Publ

ishe

d on

27

May

200

8. D

ownl

oade

d by

TU

Ber

lin -

Uni

vers

itaet

sbib

l on

13/0

4/20

15 1

3:24

:20.

View Article Online

Page 6: Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

(see Table 1)—were electrochemically dealloyed and subse-

quently subjected to slow anodic potential sweeps in oxygen

saturated electrolyte in order to characterize their intrinsic

oxygen reduction reaction (ORR) activity. Fig. 2 shows the

anodic potential sweeps of the dealloyed catalysts in oxygen-

saturated electrolyte. The flat portions of the sweep curves for

potentials below 0.8 V/RHE indicate a diffusion-limited re-

gime. For increasing Cu content of the precursor, the steep

portions of the curves are significantly shifted to higher

electrode potentials compared to the pure Pt standard catalyst

(dashed curve), indicative of an enhanced electrocatalytic

reaction rate for ORR (reduced overpotential) on dealloyed

Pt–Cu surfaces. Combined with the Pt amounts of the alloy

catalysts in the RDE film, a kinetic analysis resulted in a 4–5

fold increase in the Pt-mass normalized ORR reactivity for the

Cu-richest low temperature catalyst compared to the estab-

lished 0.10 A per mgPt of pure Pt at 0.9 V/RHE30 (see

Table 1). This dramatic increase is in accord with previously

reported RDE measurements of dealloyed Cu-rich Pt–Cu

nanoparticle catalysts.1 With increasing Pt content of the

precursor material, the observed Pt mass activity dropped to

values reported earlier for Pt rich alloys such as Pt75Ni25 or

Pt75Co25.31–35 Catalysts prepared at high temperatures gener-

ally showed reduced Pt mass based ORR activity.

Table 1 and the inset of Fig. 2 also report the measured

values of the Pt-ECSA and the Pt-ECSA normalized ORR

activities of the three dealloyed catalysts at 0.9 V/RHE. The

Pt25Cu75 catalyst prepared at 950 1C exhibited an almost 4 fold

intrinsic activity increase over pure Pt, in line with earlier

reports of this catalyst class.1,15,17,36 Low temperature

Pt25Cu75 and Pt50Cu50 catalysts showed slightly smaller activ-

ities, while both Pt75Cu25 and the high temperature Pt50Cu50only showed a 1–2� increase in specific activity, similar to

other Pt rich catalysts.30–32,37–49 The electrochemical surface

areas of the dealloyed catalysts indicate that none of the

catalysts showed a significantly increased surface area com-

pared to a standard carbon-supported Pt with about 77 m2).

To the contrary, the higher annealing temperature led to a

serious decrease of Pt-ECSA due to particle sintering.

4.2 Structural characteristics of Cu rich Pt–Cu bimetallic

alloy nanoparticles

To gain insight into the compositional and structural char-

acteristics of the Pt–Cu alloy precursor nanoparticle electro-

catalysts, we performed X-ray diffraction studies of all six

alloys prior to dealloying.

Pt25Cu75. Fig. 3a and b show the measured profiles of the

Pt25Cu75 particles over the 2y range of the fundamental (111)

and (200) peaks of Pt and Cu annealed at low and high

temperatures, respectively. The broad peaks around 2y ¼41.91 and 2y ¼ 48.51 envelop the (111) and (200) reflections

of fcc Pt–Cu bimetallic alloy phases.50,51 The absence of

superlattice peaks at lower angles confirmed the presence of

disordered fcc Pt–Cu alloys. Given the alloy nanoparticle

ensemble, the broad alloy phase peaks indicate the presence

of multiple Pt–Cu bimetallic fcc phases. In fact, as shown in

Fig. 3a, the broad (111) peak was fit using two distinct peaks at

2y ¼ 42.31 and 2y ¼ 40.91, representing two distinct Pt–Cu

alloy phases. Furthermore, the low temperature sample shows

sharp peaks around 2y ¼ 43.31 and 2y ¼ 50.51, representing

the (111) and (200) reflections of a face-centered cubic (fcc) Cu

phase. Vegard’s law, which establishes a linear relationship

between the unit cell parameter of a bimetallic alloy and its

composition, provides an excellent model to predict trends in

the composition of crystalline Pt–Cu bimetallic alloy phases

based on their peak position.21,22,52 The XRD peak analysis of

Table 1 Comparison of nominal and final compositions, Pt mass based catalytic ORR activities (at 900 mV/RHE), surface area normalized ORRactivities (at 900 mV/RHE), and surface areas of Pt–Cu alloy nanoparticle electrocatalysts

Norminalcomposition (at.%)

Annealilngtemperature

Specificactivity

Pt massactivity

Surface area Composition afterelectrochemical dealloying (at.%)

Catalyst Pt Cu 1C mA cm�2Pt A mg�1Pt m2 g�1Pt Pt Cu

1 25 75 600 643.77 0.53 88.51 79 212 25 75 950 787.68 0.35 45.16 80 203 50 50 600 595.96 0.51 85.83 — —4 50 50 950 427.61 0.11 26.80 — —5 75 25 600 359.71 0.27 74.77 86 146 75 25 950 233.85 0.03 13.01 81 19

Fig. 2 Sweep voltammetry of dealloyed carbon-supported Pt–Cu

nanoparticle electrocatalysts of various Pt to Cu ratios compared to

a standard Pt catalysts. Conditions: 5 mV s�1, 0.1 m HClO4, 25 C,

1600 rpm. Stoichiometries in the legend indicate the compositions of

the precursor alloys. Inset: surface area normalized current density

(specific activity) of de-alloyed Pt–Cu alloy nanoparticle catalysts at

900 mV/RHE. Dealloyed Cu rich Pt–Cu catalysts exhibit a clear

activity advantage over pure Pt catalysts (see also Table 1). Cu richer

precursors result in more active dealloyed catalysts.

This journal is �c the Owner Societies 2008 Phys. Chem. Chem. Phys., 2008, 10, 3670–3683 | 3673

Publ

ishe

d on

27

May

200

8. D

ownl

oade

d by

TU

Ber

lin -

Uni

vers

itaet

sbib

l on

13/0

4/20

15 1

3:24

:20.

View Article Online

Page 7: Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

the Pt25Cu75 particles thus revealed that low-temperature

annealing (600 1C) resulted in incomplete alloying of Pt and

Cu atoms, with excess Cu atoms forming large pure Cu

particles. Coexisting with the pure Cu phase are multiple

Pt–Cu alloy phases with varying compositions, some of which

must consist of a Pt content of 4 25 atomic % in order to

meet the overall stoichiometry of Pt25Cu75. The high-tempera-

ture catalyst in Fig. 3b, however, shows no sign of excess Cu.

The fundamental fcc (111) reflection exhibits a much smaller

full peak width at half maximum (FWHM). A peak fit of the

(111) reflection still required at least two different individual

peaks (Fig. 3b), suggesting the presence of more than a single

alloy phase. Yet, overall, annealing at 950 1C clearly leads to

an improved alloying between Pt and Cu atoms.

Pt50Cu50. Fig. 3c and d show the structural XRD profiles of

the Pt50Cu50 alloys annealed at 600 1C and 950 1C, respec-

tively. As with Pt25Cu75, broad reflections for the low tem-

perature catalyst indicated small particles and/or multiple

alloy phases. The low temperature profile was successfully fit

with two distinct (111) and (200) peak couples, suggesting the

presence of at least two bimetallic alloy phases. The high

temperature material profile, on the other hand, suggested

the presence of a rhombohedral Pt50Cu50 Hongshiite

phase53–56 (PDF file 00-042-1326, space group: R�3m, unit cell

parameters a, b ¼ 1.070 nm, c ¼ 1.319 nm). Consistent with

our measurements, this phase exhibits an intense (006) reflec-

tion at 2y ¼ 41.011 and a (404) reflection at 2y ¼ 47.91. Within

our probed angle range, the phase is uniquely identified by its

characteristic (205) and (241) reflections at 2y ¼ 39.21 and

2y ¼ 52.61, respectively. These latter two peaks are located at

angles outside the range between the fundamental (111) and

(200) reflections of pure Pt and pure Cu phases; hence, they

cannot be assigned to any disordered Pt–Cu bimetallic alloy

phase which would show its (111) and (200) reflection in

between those of the two pure metals.

Pt75Cu25. Finally, Fig. 3e and f depict the diffraction profile

of the Pt75Cu25 alloy nanoparticle electrocatalyst annealed at

low and high temperatures, respectively. As expected, the low

temperature profile exhibits broader peaks associated with

smaller alloy nanoparticles.50 However, both profiles can be

fit very well with one single Voigt-type peak, suggesting a

relatively homogeneous alloying in the Pt-rich composition

range. No sign of excess pure Cu was seen in either the low or

the high temperature alloy samples.

4.3 Cyclic voltammetric behavior during Pt–Cu dealloying

This section focuses on both the early stage and the final stage

of the catalyst dealloying process. Catalyst dealloying is

typically performed prior to ORR testing in deaerated electro-

lyte by voltammetric treatment of the bimetallic precursors.

Fig. 4a–f reports the initial cyclic voltammetric cycles of all six

Pt–Cu catalysts of interest.

Pt25Cu75. The first cyclic voltammograms (CVs) of the

Pt25Cu75/600 1C catalyst (Fig. 4a) show no hydrogen adsorp-

tion between 0.05 and 0.3 V, suggesting a very Cu-rich

precursor surface. A strong anodic current onset is observed

at around 0.25 V, forming a very broad anodic multi-peak

band stretching from 0.25 to 0.85 V with three individual peak

maxima at 0.4, 0.58, and 0.7 V. This broad anodic current

peak represents surface Cu dissolution from the as-prepared

precursor bimetallic surface. On the cathodic portion of the

first scan, a number of reduction peaks were observed at 0.6,

0.4 and 0.25 V, the latter two of which are likely associated

with redeposition of some Cu-ions from solution onto the

interface. On the anodic portion of the second and third CV,

the broad peak band gave way to a two-peak profile at 0.35

and 0.7 V. On the cathodic direction, the reduction peak

currents gradually decreased, and the initial three-peak profile

turned into a single reduction peak around 0.65 V. The third

CV shows some oxidation of underpotentially deposited hy-

drogen in the 0.1–0.3 V range, indicating the onset of the

exposure of Pt surface atoms, even at these very low poten-

tials. The voltammetric cycling was then continued until a

time-stable CV was obtained. The final recorded cyclic vol-

tammogram of the Pt25Cu75 is shown in Fig. 4a as well (solid

line). The final profile strongly resembles that of a pure Pt

catalyst surface with a well-developed, underpotentially-

deposited (upd) hydrogen ad/desorption region and the

formation of oxygenated surface species at anodic potentials

Fig. 3 X-Ray diffraction profiles of carbon supported Pt–Cu alloy nanoparticle precursor catalysts of varying stoichiometries and annealing

temperatures. All annealing time were 7 h. (a) Pt25Cu75 600 1C; (b) Pt25Cu75 950 1C; (c) Pt50Cu50 600 1C; (d) Pt50Cu50 950 1C; (e) Pt75Cu25 600 1C;

(f) Pt75Cu25 950 1C.

3674 | Phys. Chem. Chem. Phys., 2008, 10, 3670–3683 This journal is �c the Owner Societies 2008

Publ

ishe

d on

27

May

200

8. D

ownl

oade

d by

TU

Ber

lin -

Uni

vers

itaet

sbib

l on

13/0

4/20

15 1

3:24

:20.

View Article Online

Page 8: Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

40.80 V. The final CV suggests that dealloying led to an

almost pure Pt catalyst surface by gradual removal of Cu

atoms from the alloy particles.

The first three CVs of the Pt25Cu75/950 1C (Fig. 4b) ex-

hibited a distinctly different peak profile upon dealloying.

While the onset potential of the Cu dissolution again occurred

around 0.25 V with a current peak maximum at 0.4 V, there

were no Cu intermediate dissolution current peaks in the

0.4–0.8 V range. Another current peak maximum was obser-

vable at 0.8 V on the first scan. The first cathodic sweep was

similar to that of the low temperature material with a pro-

nounced peak around 0.65 V and smaller ones at 0.4 and

0.25 V. On the second and third cycles, the Cu dissolution

decreased in peak current and shifted slightly more cathodi-

cally compared to the first cycle. The final CV again resembled

a pure Pt catalyst with a small residual peak at around 0.7 V,

likely associated with continued Cu dissolution (see Discus-

sion). A clear, yet smaller amount of underpotentially depos-

ited hydrogen was observed to ad- and desorb in the 0.05–

0.4 V range.

Pt50Cu50. The first 3 CVs of the low and high temperature

Pt50Cu50 precursors are shown in Fig. 4c and d, respectively.

The behavior and profiles of the two catalysts are very similar

to each other and share most of their features. The anodic

sweep of the first scan immediately exhibits upd-hydrogen

oxidation followed by a relatively small Cu dissolution peak at

0.4 V. Similar to later stages of the Pt25Cu75 behavior, a

second asymmetric Cu dissolution peak appears at 0.7 V

during the first CV. Already on the second CV, the dissolution

peak near 0.4 V disappeared completely, indicating a limited

amount of Cu in the surface region. During the second and

third CV, the broad peak at 0.7 V was replaced by a sigmoid-

ally shaped voltammetric feature which gradually decreased in

intensity and finally disappeared completely. No significant

redeposition peaks were discernible on the cathodic portion of

any of the initial CVs. The final CV of Pt50Cu50 again strongly

resembled that of pure Pt, suggesting no residual Cu atoms in

the surface. A stark difference between the low and high

temperature material consisted in the magnitude of the H

upd integrals of the initial and the final time-stable CV. The

upd hydrogen integral, typically associated with the surface

area of Pt surfaces, was much larger for the Pt50Cu50/600 1C

material than for the Pt50Cu50/950 1C catalyst, in agreement

with the notion that high annealing temperature resulted in

larger alloy particles with reduced surface area.

Pt75Cu25. Fig. 4e and f report the initial 3 CVs of the

Pt75Cu25 precursor. Unlike the other catalysts, no Cu dissolu-

tion peak around 0.4 V was observable. Slight Cu dissolution

was indicated for the low temperature material by a peak at

around 0.7 V, which rapidly dropped in intensity during the

second and third cycles. As in the other two cases, the final CV

was in large part identical to that of a pure Pt surface, with the

high temperature profile showing a drastically reduced H upd

integral, suggesting large dealloyed nanoparticles.

4.4 Trends in the thermodynamics of Cu dissolution on Pt–Cu

alloy surfaces

The working surface of the Pt–Cu alloy nanoparticles is clearly

complex, involving both strong compositional profiles near the

surface and a variety of effects resulting from undercoordi-

nated surface sites. To gain insight into the detailed, atomic-

level energetic factors influencing the dissolution of Cu in these

bimetallic systems, we perform periodic density functional

theory calculations. We consider the dissolution of a Cu

monolayer from a kinked Pt(854) surface; this model provides

atomic-level insight into the effect of defects and other rough

surface features on the thermodynamics of Cu dissolution

(a detailed discussion of the relationship between the model

and the experimental systems is presented in section 5.2).

Table 2 gives the calculated binding energies of Cu atoms to

the Pt(854) surface as a function of Cu coverage. These data,

corrected for potential effects, are used to plot the surface free

energy of different numbers of Cu atoms on the Pt(854)

substrate at different potentials (Fig. 5(a)–(c)—the potential

scale, U, is referenced to the experimentally-measured equili-

brium potential for Cu21 dissolution from bulk Cu, 0.34 V vs.

Fig. 4 Initial three cyclic voltammetric profiles during the dealloying process of Pt–Cu catalysts from Fig. 3. (a) Pt25Cu75/600 1C; (b) Pt25Cu75/

950 1C; (c) Pt50Cu50/600 1C; (d) Pt50Cu50/950 1C; (e) Pt75Cu25/600 1C; (f) Pt75Cu25/950 1C. The solid line displays the cyclic voltammogram at the

end of the dealloying process. Conditions: 10 mV s�1, deaerated 0.1 M HClO4, 25 C.

This journal is �c the Owner Societies 2008 Phys. Chem. Chem. Phys., 2008, 10, 3670–3683 | 3675

Publ

ishe

d on

27

May

200

8. D

ownl

oade

d by

TU

Ber

lin -

Uni

vers

itaet

sbib

l on

13/0

4/20

15 1

3:24

:20.

View Article Online

Page 9: Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

SHE). The maximum number of Cu atoms considered, 13, is

equivalent to a full monolayer of Cu. If kinetic effects are

absent, then at a given value of the potential, Cu dissolution

will proceed until further dissolution would require an increase

in the free energy of the surface. Hence, at a value of UE 0.34

V, at most a single Cu atom is thermodynamically favored to

dissolve. This corresponds to a situation wherein the most

easily removed Cu atoms (those at the Cu-rich kink

sites—Fig. 5d) are primarily in contact with other Cu atoms;

hence, the dissolution potential is little shifted from the

dissolution potential of bulk Cu. At higher potentials, more

Cu atoms begin to dissolve. At B0.45 V, a large plateau in the

free energy vs. NCu plot develops; this plateau implies that a

significant number of Cu atoms will dissolve at or near a single

potential. This single dissolution potential, corresponding to

dissolution of Cu atoms from the regions between kinks

(Fig. 5e), is analogous to what might be observed in dissolu-

tion from flat regions of the Pt–Cu nanoparticles. Finally, at

potentials greater than 0.5 V, dissolution of Cu atoms that are

strongly bound to Pt step and kink edges (Fig. 5f) becomes

thermodynamically favorable. These results are summarized in

Fig. 5g, an isotherm of Cu dissolution from the Pt(854)

surface. As in earlier subpanels of Fig. 5, the number of Cu

atoms remaining on the surface at given potentials is deter-

mined by the criterion that further dissolution requires an

increase in free energy. We note that, although this isotherm is

determined on a bulk fcc Pt crystal, the qualitative trends

should remain valid for fcc Pt–Cu bulk alloys of other

compositions. We further note that, while this isotherm does

not include the effects of surface entropy on the dissolution

Table 2 Differential binding energy (B.E.) of Cu atoms to the Pt(854)surface. For a given value of NCu, the differential B.E. is defined asENCu

� ENCu�1 � Ebulk;Cu. Computational details are provided in theMethods section in the text (section 3)

NCu Differential Cu binding energy/eV

0 0.011 �0.142 �0.153 �0.134 �0.205 �0.216 �0.207 �0.228 �0.179 �0.3310 �0.4611 �0.4512 �0.6213 —

Fig. 5 Density functional theory analysis of Cu dissolution from the surface of a Pt(854) substrate. (a)–(c) Top views of the Pt(854) surface with

13, 6, and 3 Cu atoms per unit cell, respectively, adsorbed on the surface. The yellow lines in (a) denote the surface unit cell. Cu atoms are shaded

ochre, and Pt atoms are shaded silver-gray. (d)–(f) Free energy per unit cell (eV) for different Cu coverages at potentials of 0.34, 0.45, and 0.52 V vs.

SHE, respectively; for all potential values, the equilibrium potential for bulk Cu dissolution is taken from standard experimental tables. (g)

Isotherm of the most thermodynamically stable Cu coverage as a function of the electrode potential.

3676 | Phys. Chem. Chem. Phys., 2008, 10, 3670–3683 This journal is �c the Owner Societies 2008

Publ

ishe

d on

27

May

200

8. D

ownl

oade

d by

TU

Ber

lin -

Uni

vers

itaet

sbib

l on

13/0

4/20

15 1

3:24

:20.

View Article Online

Page 10: Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

potentials, it directly accounts for site-specific changes in

adsorption energy due to both changes in Cu coverage and

the structural heterogeneity of the Pt(854) substrate.

5. Discussion

The selective dissolution of a less noble component from a

bimetallic alloy nanoparticle is commonly referred to as

‘‘metal dealloying’’;4,57–60 at the atomic level, it continues to

remain a poorly understood process. The onset potential of

selective dissolution from the alloy is conventionally referred

to as the ‘‘critical dissolution potential’’. Dealloying is often

associated with the formation of porous structures provided

that the initial alloy composition was below the so-called

parting limit, that is the maximum noble metal content for

which continuous dissolution (dealloying) occurs.59 However,

the underlying physics controlling the dissolution potential as

function of alloy characteristics, such as composition, surface

roughness and others, is unclear. Unlike previous reports on

dealloying on flat surfaces, our experimental results focus on

the Cu dissolution in its early stages from Pt–Cu alloy

nanoparticle surfaces, that is, strongly curved and very rough

(stepped, kinked) surfaces. Our computational results address

trends in the thermodynamics of the Cu dissolution potential

from first principles, introducing quantum mechanical DFT

calculations of metal- and site-specific binding energies into

the discussion of metal dealloying. The importance of metal

specific energetics for deeper insight in the physics of the

critical dissolution potential on stepped and kinked surfaces

will now be discussed, and compared and contrasted to our

experimental results.

5.1 Synthesis, structure, and ORR activity

Synthesis. We have used an impregnation/freeze drying/

annealing synthesis method for the preparation of carbon-

supported Pt–Cu bimetallic alloy nanoparticle electrocata-

lysts. The nature of this synthesis accounts for the some of

the unusual multi-phase structural features discussed below.

As described in an earlier report,1,14 the synthesis does not

involve any liquid-phase reduction, precipitation, filtration, or

other steps where incomplete reaction or materials transfer

could occur. As a result, it yields very controlled (average)

metal stoichiometries for the final catalysts. The preparation

employs carbon-supported Pt nanoparticle precursors which

are impregnated with Cu salts.61 Thereby, the pre-existing

metallic Pt particles serve as centers for alloy formation; their

high dispersion (small mean diameter) keeps the resulting alloy

mean particle size small compared to co-impregnation techni-

ques in the absence of structure-directing agents. During the

annealing process, Cu ions are reduced to the metallic state

and diffuse into the preexisting Pt metal particle lattices,

gradually forming uniformly alloyed Pt–Cu alloy lattices.

Depending on the desired average Pt to Cu ratio, different

amounts of Cu atoms have to diffuse into the preexisting Pt

particles. This is where the pronounced sensitivity of this

method to annealing conditions arises and accounts for the

observed multiphase structure for Cu-rich stoichiometries.

Preparation of Pt25Cu75 requires about 10 times as many Cu

atoms (compared to preparation of Pt3Cu) to inter diffuse with

Pt particles in order to form a uniformly alloyed material.

The synthesis method used here differs from all-liquid-

precursor impregnation/liquid-phase reduction methods. In

these, often assisted by surfactants,62,63 metal ions are pre-

mixed to achieve atomic proximity and then impregnated onto

high surface area supports, followed by reduction and alloy-

ing. While the surfactants may provide a means to control the

resulting alloy particle size, they are generally difficult to

remove afterwards without compromising the particle disper-

sion. Co-impregnation/ reduction of two or more metal salts in

the absence of surfactants generally results in large alloy

particles.

Alloy uniformity and phase structure. Alloy nanoparticle

ensembles are generally not characterized by discrete proper-

ties but exhibit distributions of particle size and composi-

tion.64–69 For the Pt25Cu75 catalyst, low temperature annealing

results in small, yet incompletely alloyed, Pt–Cu particles, as

observed in Fig. 3a. Incomplete alloying was obvious from the

presence of excess Cu. The broad (111) reflection of the alloy

phases (Fig. 3a) also suggested the presence of multiple Pt–Cu

phases with different Pt-to-Cu atomic ratios. Annealing at

higher temperatures (Fig. 3b) resulted in more uniform Pt–Cu

alloys, yet also larger average particles.

In the Pt50Cu50 catalyst (Fig. 3c and d), no Cu fcc reflections

are observed, indicating that all Cu has diffused into Pt. While

the low temperature phase structure still suggests the presence

of multiple Pt–Cu alloy phases, the high temperature profile

shows the rarely reported single rhombohedral Hongshiite

Pt–Cu phase.53,55,56 The small peak at 2y ¼ 39.21 is a uniquely

characteristic reflection of this ordered Pt–Cu alloy phase. In

contrast to all other alloy phase reflections in Fig. 3, this peak

cannot arise from a disordered Pt–Cu fcc phase, because

insertion of Cu atoms into Pt fcc lattices invariably leads to

lattice-contracted Pt–Cu alloy unit cells with (111) reflections

at larger 2y compared to the pure Pt (111), at 2y ¼ 39.71.

In the Pt75Cu25 catalyst (Fig. 3e and f), even at an annealing

temperature of 600 1C, the Cu atoms form uniformly alloyed

disordered fcc Pt–Cu alloy phases. Again, with increasing

annealing temperature, the mean alloy particle size increases,

as seen by a comparison of the FWHM of the (111) reflections

in Fig. 3e and f.

ORR activity of Pt–Cu alloys. In earlier reports,1,36,70,71 we

described the significant increase in the ORR activity achieved

using dealloyed Pt25Cu75 bimetallic nanoparticle electrocata-

lysts. Prepared at 600 and 800 1C, the dealloyed Pt25Cu75catalyst precursors were found to exhibit a 4–6 fold activity

advantage over a standard carbon supported 45 wt% Pt

nanoparticle catalyst in terms of Pt mass-based (0.4–0.6 A

mg�1Pt) and Pt surface area-based activity (700–900 mAcm�2Pt).

1 Similar activity increases were recently also realized

in single MEAs under realistic fuel cell conditions.17

The current results provide insight into the relationship

between the ORR reactivity and the Cu content of the

electrocatalyst precursor alloy. The compositional data given

in the last column of Table 1 evidences that Cu dealloying does

occur during voltammetric treatment of the Cu-rich Pt–Cu

This journal is �c the Owner Societies 2008 Phys. Chem. Chem. Phys., 2008, 10, 3670–3683 | 3677

Publ

ishe

d on

27

May

200

8. D

ownl

oade

d by

TU

Ber

lin -

Uni

vers

itaet

sbib

l on

13/0

4/20

15 1

3:24

:20.

View Article Online

Page 11: Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

precursors (catalysts 1 and 2). The Pt75Cu25 catalyst precursor

showed better corrosion stability and suffered less relative

Cu loss.

Annealed at 600 1C, the dealloyed active catalysts seem to

maintain their high Pt-mass and specific activity, at least up to

a Pt atomic ratio of 50% (see Table 1, Catalysts 1 and 3).

Annealed at 950 1C, alloy particle growth generally reduced

the ESCA which, in turn, reduced the Pt mass activity. The

alloy phase uniformity, however, improved and the individual

phase compositions approached that of the overall stoichio-

metry. The Pt25Cu75/950 1C precursor characterized by the

largest particle size and by uniform alloying exhibited the

highest intrinsic activity after Cu dealloying; this is in agree-

ment with recent MEA measurements of dealloyed Pt25Cu75catalysts.17 This suggests that a highly dispersed, uniformly

alloyed Pt25Cu75 precursor might result in a very favorable

dealloyed catalyst from both a Pt-mass and an intrinsic

activity point of view.

The opposite relation between annealing temperature and

intrinsic activity was observed for Pt50Cu50. This analysis

shows that the alloy phase structure characteristics (ordered,

disordered, multi phase or single phase, etc.) significantly

impact the catalytic activity of the resulting dealloyed

catalysts.

The Pt-rich Pt75Cu25 precursors showed a familiar activity

increase of about 2–2.5 � over pure Pt at low annealing

temperatures, in line with many previous reports on Pt alloy

ORR catalysts with about 75 at.% Pt.35,38,45,72,73 Pt-rich alloys

annealed at high temperature (large particles) did not show

any activity advantage at all.

To put the observed ECSA-normalized activity values of

dealloyed Pt–Cu bimetallic catalysts in perspective with the

corresponding activities of extended flat surfaces, we point out

that flat Pt and Pt alloy surface exhibit much larger ESCA

normalized (specific) ORR activity.30,31,33,74,75 Pt-mass based

ORR activities are meaningless for extended surfaces. Flat

pure Pt shows a specific ORR activity of about 1500 mAcm�2Pt,

30,74 while extended Pt alloy surfaces show the familiar

2–2.5 fold enhancement (2000–4000 mA cm�2Pt).30,74

Selective metal dissolution from flat Pt alloy surfaces for the

purpose of improved electrocatalysis was reported as early as

1988 by Paffett et al.76 The authors investigated the impact of

Cr surface dissolution from flat Cr-rich Pt–Cr precursors on

the observed ORR activity. They observed significant surface

roughening due to Cr dissolution, which led to increases in the

absolute (geometric area-normalized) ORR currents. When

normalized to the real surface area, however, the ORR

activities of Pt–Cr was comparable to that of pure Pt. The

authors concluded that the addition of Cr to a Pt surface, and

the associated surface roughening, does not provide a sub-

stantial benefit for ORR activity. Other reports where de-

alloyed flat metal surfaces were proposed for use as high

surface area interfaces for catalysis and sensing include the

work by Ding et al.57 as well as Liu et al.77 Both studies

emphasize the advantage of nanoporosity derived from flat

surfaces for achieving improved catalyst dispersion.

As outlined earlier1,17 and confirmed in the present study,

the mechanistic origin of the ORR activity improvements of

carbon-supported dealloyed Pt–Cu nanoparticle catalysts

compared to carbon-supported Pt catalysts is not based on

an increased surface area. We believe that the increased

reactivity of dealloyed Cu-rich Pt–Cu nanoparticle catalysts

lies primarily in the modified geometric and electronic proper-

ties of the Pt-enriched nanoparticle shell. This is in analogy to

other core–shell type catalyst concepts, such as the ‘‘Pt mono-

layer’’ particle catalysts reported by Adzic et al.9 or the

‘‘Pt-skin’’ catalyst concept introduced by Markovic and Sta-

menkovic.74,75,78 The ‘‘Pt monolayer’’ and ‘‘Pt skin’’ catalysts

are characterized by an immediate neighborhood of Pt surface

atoms and non-Pt atoms in the second layer, resulting in

changes in the Pt–Pt distance on the surface (geometric effects)

as well as changes in the electronic structure due to charge

transfer (ligand effects). Both effects are convoluted in these

catalyst structures. In contrast, tuning the thickness of the Pt

enriched shell of dealloyed Pt–Cu particles should, in princi-

ple, allow a separation between geometric and ligand effects by

isolating geometric effects for thicker shells.

Recently, the hypothesized core-shell structure of our active

catalysts was confirmed using anomalous small angle X-ray

scattering experiments.50 Compressive strain in the Pt-en-

riched particle shell appears to play a critical role in the

activity enhancement.

5.2 Experimental and theoretical Cu dissolution potentials

Pt25Cu75. The absence of H upd peaks during the first CV in

Fig. 4a and b for the two Pt25Cu75 precursors indicates an

essentially pure Cu surface and is in agreement with the

tendency of Cu to segregate to the surface of Pt–Cu alloys

under thermal treatment.79 The onset potential of the broad

anodic faradic process in Fig. 4a and b in the 0.30–0.35 V

range is associated with pure Cu dissolution according to

Cu(s) - Cu21 þ 2e� E1 ¼ þ0.34 V/NHE. (2)

Formation of Cu(I) species is not favored under the chosen

conditions.29 We suspect that the anodic process represents the

dissolution of Cu surface atoms embedded in a Cu environ-

ment (Cu bulk atoms) and denote these Cu surface atoms as

Cu@Cu. Considering the alloy phase structure discussion

above, multi layers of pure Cu are indeed present in the

Pt25Cu75/600 1C sample, both in the form of pure Cu particles

as well as in the form of surface-segregated Cu layers on top of

the Pt–Cu alloy particles for both the Pt25Cu75/600 1C and the

Pt25Cu75/950 1C precursor. The behavior of the Cu@Cu

atoms is qualitatively similar to that of the first one or two

Cu atoms to dissolve from the monolayer on the Pt(854)

surface in our theoretical calculations (see discussion below);

as noted above, those atoms are primarily in contact with

other Cu atoms and are calculated to dissolve at essentially the

bulk Cu dissolution potential.

The broad voltammetric Cu dissolution feature stretching

from 0.4–0.8 V, with peaks at 0.6 and 0.75 V for the

Pt25Cu75/600 1C catalyst, indicates the presence of other Cu

surface atoms with distinctly more anodic dissolution poten-

tials. We suspect that, depending on their chemical environ-

ment at the atomic scale, their individual corrosion potentials

vary over a certain range. Indeed, a direct correlation between

the Cu surface atomic environment (local composition,

3678 | Phys. Chem. Chem. Phys., 2008, 10, 3670–3683 This journal is �c the Owner Societies 2008

Publ

ishe

d on

27

May

200

8. D

ownl

oade

d by

TU

Ber

lin -

Uni

vers

itaet

sbib

l on

13/0

4/20

15 1

3:24

:20.

View Article Online

Page 12: Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

coordination number, and nature of the coordinated neigh-

boring atoms) of each individual alloy particle and its corre-

sponding local Cu corrosion potential is hypothesized and

further supported by our DFT results in Fig. 5.

For the Pt25Cu75/950 1C material, only Cu@Cu bulk

dissolution around 0.35 V and the Cu dissolution around

0.75 V are present. This is consistent with the notion of a

more uniformly alloyed nanoparticle ensemble, which shows a

narrower particle composition distribution and therefore ex-

hibits more defined corrosion potentials. This is also in

excellent agreement with anodic polarization behavior re-

ported by Pugh et al.7,80 who monitored the dealloying of

arc-melted, flat Pt25Cu75 surfaces with very well controlled

stoichiometries. Pugh’s measurements (Fig. 2 in ref. 7) showed

Cu@Cu bulk dissolution around 0.35 V, followed by reduced

currents in the 0.4–0.6 V potential range (termed ‘‘Passiva-

tion’’ by Pugh); he identified the peak at 0.75 V as the actual

Cu dissolution peak from the Pt–Cu alloy. He did not,

however, provide atomic insight into the Cu dissolution

processes.

A classic single-crystal study by Markovic and Ross of the

faradic Cu dissolution and deposition processes on Pt(111)

and Pt(100)81–83 provides valuable experimental insight into

the atomistic processes underlying our measured Cu dissolu-

tion behavior. Remarkably, the observed qualitative voltam-

metric profiles on Pt(111) and Pt(100), as well as the Cu-

related peak potentials, are nearly identical to the measured

profiles on the Pt–Cu nanoparticles. The surprising correspon-

dence between the structurally complex, compositionally var-

ied nanoparticle surfaces and the structurally well-defined

Cu@Pt(111) and Cu@Pt(100) surfaces suggests that reason-

ably flat regions may exist on our nanoparticles; it may also be

true that occasional steps or kinks are present in the single

crystal samples. The correspondence also suggests that quali-

tatively useful insights into the nanoparticle dealloying process

can be obtained from our simple Cu@Pt(854) single

crystal model.

Markovic and Ross’s combined voltammetric and LEED

studies83 related the anodic peak around 0.35 V to Cu

multilayer (Cu@Cu bulk) dissolution, while asymmetric

anodic peaks around 0.75 V were unambiguously associated

with Cu monolayer upd and Cu monolayer stripping.

It follows that Cu atoms sitting on top of pure Pt subsurface

layers (denoted here as Cu@Pt) exhibit a significantly

more anodic dissolution potential than those sitting on pure

Cu surfaces. This observation is completely consistent

with our theoretical calculations (showing an increased

dissolution potential of Cu atoms on terrace-like Pt(111)

regions compared to bulk Cu—Fig. 5e and g) and lends

support to our hypothesis that the chemical environment has

a significant impact on the detailed metal dissolution profiles.

Markovic and Ross further observed that while the Cu@Pt

dissolution occurred in a fairly narrow potential range, its

formation on the cathodic scan occurs over a broad potential

range with multiple peaks at 0.6 and 0.4 V, in line with the

observed peak potentials in Fig. 4. Bulk Cu deposition started

at around 0.2 V on the single crystals (Fig. 4 and 5 in ref. 83),

also in close agreement with the observed cathodic peaks in

this study.

Pt50Cu50, Pt75Cu25. The small Cu@Cu bulk dissolution

peak present in the Pt50Cu50 alloys (Fig. 4c and d) indicates

that some bulk-like Cu surface atoms must be present, given

the absence of unalloyed pure Cu particles. However, the

presence of H upd between 0.05–0.4 V also suggests the

presence of Pt surface atoms. Segregated Cu therefore appears

to be present as islands unable to cover the entire particle

surface. Finally, the Cu@Pt dissolution peak around 0.7–

0.75 V becomes weaker with increasing Pt content and sharply

drops in intensity on the second potential scan, suggesting a

rapid depletion of Cu near the surface. As with the 75 at.% Cu

materials, continued cycling finally resulted in a pure Pt

profile. Finally, the Pt75Cu25 compositions (Fig. 4e and f)

seem to have very little surface Cu initially and show a well-

developed H upd peak, even on the first CV scan.

The critical Cu dissolution potential. Fig. 6a plots the anodic

polarization portions of the initial 3 CVs of the Pt25Cu75/

600 1C alloy particles in a log I–E plot in order to highlight the

quasi-constant Tafel Slope of the Cu dissolution onset be-

tween 0.2 and 0.4 V. We attempt to provide some insight into

the observed dissolution behavior of Pt–Cu alloy nanoparticle

Fig. 6 (a) A semi-logarithmic plot of the anodic sweeps of Fig. 4a,

Pt25Cu75/600 1C. A sharp onset of Cu dissolution currents are

discernible at a critical dealloying potential Ec. (b) Diagrammatic

illustration of how the critical dissolution potential of a Cu monolayer

depends on the composition of its subsurface layer. Cu monolayers on

Cu substrates behave bulk-like with dissolution potential E0. Cu

monolayers on Pt–Cu alloy layers exhibit a significantly more positive

dissolution (dealloying) potential Ecritical,2.82,83 Cu atoms on top of a

Pt–Cu alloy exhibit intermediate dissolution potentials Ecritical,1. The

critical potential of pure Pt is very high (also denoted as E0 here).

This journal is �c the Owner Societies 2008 Phys. Chem. Chem. Phys., 2008, 10, 3670–3683 | 3679

Publ

ishe

d on

27

May

200

8. D

ownl

oade

d by

TU

Ber

lin -

Uni

vers

itaet

sbib

l on

13/0

4/20

15 1

3:24

:20.

View Article Online

Page 13: Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

ensembles, such as that described in Fig. 6a, using simple,

schematic, anodic log I–E polarization schemes (Fig. 6b). Such

schemes were first discussed by Pickering5 and later by

Sieradzki84 in order to illustrate the observed polarization

behavior of flat bimetallic alloy surfaces.

The scheme in Fig. 6b shows the Butler Volmer (BV)

dissolution characteristics of bulk Cu and bulk Pt as the left-

most (solid) and right-most (dotted) diagonal line, respec-

tively, with their characteristic thermodynamic dissolution

onset potential E1 given at the bottom of the graph. Also

shown are cartoons of the surface atomic arrangements asso-

ciated with the two distinct dissolution potentials of pure Pt

and pure Cu. The cartoons are to emphasize the impact of the

nature of the coordinating atoms surrounding a surface atom

to be dissolved; they do not do justice to the effects of

coordination number (discussed in our theoretical treatment

below), which also impact the dissolution behavior. Cu atoms

on Cu atoms (Cu@Cu) result in the Cu bulk dissolution, while

Pt surface atoms on top of Pt atoms (Pt@Pt) lead to the Pt

dissolution at 1.19 V.24

In between the two pure metal log I–E curves, two addi-

tional Cu dissolution characteristics are depicted (dashed and

solid curves) in Fig. 6b for two different Pt–Cu alloy surfaces

with distinct stoichiometry. From dissolution studies on flat

surfaces,4,58–60,84 it is known that the bimetallic alloy stoichio-

metry below the parting limit correlates with a characteristic

critical dissolution potential (Ecitical,1 and Ecritical,2 in Fig. 6b)

above which major dissolution of the less noble metal compo-

nent commences (steep portion of curve). Cathodic of the

critical dissolution potentials no significant Cu dissolution

occurs (flat portion of polarization curve).

A huge body of literature has been dedicated to clarifying

the origin and the controlling factors of the critical dissolution

potential Ecritical and the associated dealloying pro-

cesses.3–6,57–60,84–93 There is even some controversy as to

how to best determine the true critical dissolution potential

of a bimetallic alloy.94 Experimental and theoretical methods

used to characterize and describe the processes around the

critical potential include polarization curves, cyclic voltamme-

try, in situ XRD, electron microscopy, Monte Carlo simula-

tions, as well as empirical free energy models.

Pickering5 argued that the critical potential arises from the

competition between the surface dissolution rate of the less

noble element and its volume diffusion rate to the surface. Past

the critical potential, he argues, pitting corrosion sets in and

inflicts large-scale damage to the alloy surface.

Sieradzki84 reproduced essential features of bimetallic

dealloying curves using kinetic Monte Carlo simulations in

2D and 3D. Based on percolation theoretical arguments, he

stressed the competition between the dissolution kinetics of the

less noble metal and the surface diffusion of the more noble

metal as the controlling factor for the critical dissolution

potential; it represents the balance point between roughening

and smoothing.

Quasi-atomistic approaches4,60 to model the critical dissolu-

tion potential were reported by Erlebacher and coworkers;

these analyses considered the thermodynamic impact of the

number of nearest neighbors on the activation energy and rate

of metal dissolution and diffusion. The kinetic models were

able to capture the evolution of nanoporosity length scales

during the bulk dealloying above the critical potential.

Other modeling approaches have considered the thermo-

dynamics of curved surfaces85 by taking into consideration the

Gibbs–Thomson relation or have identified the interfacial

energy of two metals as a lumped empirical thermodynamic

parameter to model the critical dissolution potential.

We stress that the modeling approaches reviewed above do

not consider alloy-specific energetics of intermetallic bond

making and bond breaking. Thus, while the models can

provide useful insight into the dealloying process and in the

evolution of nanoporosity,4 their general applicability to any

given bimetallic alloy system remains unclear. Indeed, Pugh

et al.7 pointed out in his discussion of dealloying in the Pt–Cu

system that the inclusion of metal- and site-specific binding

energies, that is, the thermodynamics on the atomic scale,

might be necessary in order to accurately capture the experi-

mental voltammetric dealloying behavior; below, we discuss

the first steps in such a metal- and site-specific theoretical

treatment.

Experimental in situ XRD investigations of the dealloying

processes at the atomic scale to the left and right of the critical

dissolution potential was realized by a number of excellent

studies by Renner et al.89–91 In a study on Cu dissolution from

Cu75Au25 single crystal surfaces, the authors were able to show

the formation of a three-monolayer thick Au-enriched passi-

vation layer with inverted stacking at potentials cathodic of

the dissolution potentials. Anodic of these potentials, the

authors found 12 monolayer thick Au-rich islands, which

inherited the inverted stacking of the passivation layer.91

While these studies clarified the dealloying mechanism from

Cu3Au, it is unclear whether these results are equally valid for

other bimetallic systems, such as Pt–Cu. A similar study

performed with Cu83Pd17 remained inconclusive whether or

not a similar Pd rich passivation layer was formed cathodic of

the Cu corrosion potential.89

The impact of metal specific binding energies. Our first-

principles DFT calculations explicitly incorporate metal-spe-

cific binding energies into the treatment of dissolution, thereby

permitting us to establish relationships between the detailed

local chemical environment (the coordination number and the

nature of the coordinating atoms) of the Cu surface atoms and

their thermodynamic dissolution potential. As discussed

above, although the structural model does not explicitly treat

the effects of bulk alloying on dissolution dynamics, it does

provide insight into the effect of undercoordinated surface

features on the dealloying process, and it is able to qualita-

tively reproduce certain features of the cyclic voltammograms

in Fig. 4.

Our analysis provides a direct link between the binding

energy of Cu atoms to the Pt(854) substrate and the dissolu-

tion potential. From eqn (1) above, it may be shown that the

potential at which surfaces with Ncu � 1 and Ncu copper

atoms, respectively, have equal free energies (i.e., the potential

at which at transition between these two states—a dissolution

event—will become thermodynamically favorable) is

U�U� ¼ � 1

2ðENCu

� ENCu�1 � Ebulk;CuÞ ð3Þ

3680 | Phys. Chem. Chem. Phys., 2008, 10, 3670–3683 This journal is �c the Owner Societies 2008

Publ

ishe

d on

27

May

200

8. D

ownl

oade

d by

TU

Ber

lin -

Uni

vers

itaet

sbib

l on

13/0

4/20

15 1

3:24

:20.

View Article Online

Page 14: Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

where the term in parentheses may be interpreted as the

binding energy of a Cu atom to the Pt substrate and can vary

with both the coverage of Cu and the structure of the

underlying Pt surface; U1 is the experimentally determined

dissolution potential of bulk Cu.24 This relationship immedi-

ately establishes an atomistic interpretation of the thermody-

namic dissolution potential; namely, for Cu atoms that

are strongly bound to the Pt substrate and to the neighboring

Cu atoms (implying that the term in parenthesis is

very negative), the dissolution potential will be shifted to more

positive values.

Eqn (3) can be used to interpret the simple isotherm

presented in Fig. 5g. Cu atoms that dissolve at potentials near

the bulk Cu dissolution potential are bound primarily to other

Cu atoms (Fig. 5a). In contrast, the significant number of Cu

atoms that dissolve at B0.45 V (corresponding to dissolution

from Cu on terrace-like Pt(111) regions of the (854)

surface—Fig. 5b) are bound to more Pt atoms than to Cu

atoms; the higher dissolution potential implies that Cu–Pt

bonds are stronger than Cu–Cu bonds (eqn (3)). We note

that, although this shift is qualitatively consistent with our

experimental CVs (Fig. 4), the quantitative value of the

dissolution potential, 0.45 V, is lower than the high-potential

peak shown in many of the CVs; this modest quantitative

discrepancy may be due to kinetic barriers to Cu atom

dissolution (not accounted for in this analysis) associated with

rearrangement of water atoms or other related processes,

analogous to the modest barriers determined for proton

discharge on Pt(111) by Skullason et al.95 Alternatively, as

suggested by Leiva et al.,96,97 it is possible that small quantities

of oxide or other species contaminate the experimental Cu

surface in this potential range, thereby rendering dissolution

more difficult—although it is important to note that no

voltammetric evidence of such species has been found in this

or previous work.82,83 Finally, the relatively small number of

Cu atoms that are bound primarily to Pt steps and kinks show

the highest dissolution potentials (Fig. 5c) and, by eqn (3), the

strongest binding to the surface.

The theoretical isotherm in Fig. 5g can also be related to the

experimental dissolution curves presented in Fig. 4. Each point

in Fig. 5g represents the thermodynamic dissolution potential

of a chemically distinct type of Cu site on the Pt(854) surface.

If, as a first approximation, we neglect interactions between

the various sites, and if we consider only the thermodynamic

limit of dissolution, then the Cu atoms at each type of site

should follow Langmuir dissolution isotherms. Each site

would then contribute an approximately Gaussian-shaped

peak to the dissolution current, with the peaks centered at

the equilibrium dissolution potential of the corresponding site.

The net superposition of such peaks is shown in Fig. 7.

Although this theoretical curve is clearly a simple approxima-

tion to the measured voltammograms, and although, as dis-

cussed above, many complex structural and electronic features

are likely responsible for these curves and for the critical

potential, the curve nonetheless provides insight into some

of the voltammetric features, including the multilayer peak at

0.34 V and the shifted underpotential deposition peak at

higher potentials. This simple analysis therefore suggests that

some features of the curves may be related to simple, atomic-

scale changes in the binding energy of Cu atoms to the Pt/alloy

surfaces as a function of Cu atom coverage.

The simple thermodynamic results presented in Fig. 5 and 7

can also provide qualitative insight into certain features of the

Tafel-style dissolution plots in Fig. 6. The second and third

curves of that figure represent dissolution of Cu atoms in

different chemical environments, analogous to the different

sites shown in Fig. 5. In Fig. 7, we have treated the current

produced by dissolution from such sites by assuming a finite

number of Cu atoms at each site; as these atoms are depleted,

the current decreases, leading to the Gaussian-type peak

shapes. However, if the surface Cu atoms are continuously

replaced by Cu atoms from the bulk, as suggested by Pickering

in his analysis,5 there will be no decrease, and continuous

current will be produced from each type of site after the

thermodynamic potential is reached, as in Fig. 6.

6. Conclusions

(1) We have presented a strategy to modify the surface

catalytic properties of Pt bimetallics. The strategy is in prin-

ciple based on the ancient method of ‘surface depletion gilding’

for changing surface compositions of bimetallic compounds,

which dates back over 2000 years to pre-Columbian gold-

smiths in Mesoamerica. Our modern flavor of the ancient

method uses electrochemical voltammetric sweeps to selec-

tively dissolve less noble alloy components (dealloying), in our

case Cu, while the more noble components, herein Pt, remain

in the alloy.

(2) We have established detailed qualitative relationships

between the synthesis conditions of the Pt–Cu alloy nanopar-

ticle precursors, their structural characteristics and their vol-

tammetric profiles during the early stages of Cu dissolution.

(3) We have shown that dealloyed Pt–Cu alloy nanoparticle

eletrocatalysts prepared by partial Cu dissolution from Cu rich

Pt25Cu75 precursor alloys exhibited a 4–5 fold ORR activity

enhancement on a Pt-mass basis compared to state-of-the-art

Fig. 7 Theoretical plot of Cu dissolution current versus potential.

Langmuir isotherms are assumed for each distinct Cu site, and ion

concentrations are taken at their standard state values. The potential

scale is referenced to the experimentally-tabulated dissolution poten-

tial of bulk Cu, and the current is normalized by the highest calculated

value.

This journal is �c the Owner Societies 2008 Phys. Chem. Chem. Phys., 2008, 10, 3670–3683 | 3681

Publ

ishe

d on

27

May

200

8. D

ownl

oade

d by

TU

Ber

lin -

Uni

vers

itaet

sbib

l on

13/0

4/20

15 1

3:24

:20.

View Article Online

Page 15: Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

pure Pt nanoparticle catalysts. The most active catalyst for the

ORR on a ECSA-normalized metric (specific ORR activity)

was achieved after surface Cu dissolution from a uniformly

alloyed (950 1C annealing) Pt25Cu75 nanoparticle precursor

alloy in accordance with recent MEA studies.17

(4) DFT computations of the energetics of Cu adatom

dissolution from a (854) model surface yielded a metal- and

site-specific correlation between the local atomic environment

of Cu surface atoms and their detailed dissolution potential.

Cu surface atoms with multiple Cu coordinations dissolved

near the thermodynamic bulk Cu potential, whereas under-

coordinated Cu surface atoms or Cu surface atoms with

primarily Pt neighbors showed higher dissolution potentials.

(4) The combination of experimental data and DFT com-

putational results has corroborated our hypothesis that, in

addition to the previously suggested balance between metal

dissolution and surface diffusion kinetics, metal-specific inter-

actions on the atomic scale are important controlling factors in

the understanding of the onset potentials of Cu dissolution for

Pt–Cu bimetallics (critical dissolution potentials). First prin-

ciples computation of binding energies has revealed that the

local chemical environment of surface Cu atoms is a control-

ling factor for the dissolution onset potentials.

Acknowledgements

This project was supported by the Department of Energy,

Office of Basic Energy Sciences (BES), under grant LAB04-20

via a subcontract with the X-ray Laboratory for Advanced

Materials (XLAM) and the Stanford Synchrotron Radiation

Laboratory (SSRL). Further support was provided by the

National Science Foundation (NSF) under the award

#0729722. Acknowledgment is made to the Donors of the

American Chemical Society Petroleum Research Fund for

partial support of this research (grant #44165). Support by

the State of Texas through the Advanced Research Program

(ARP) during the 2008–2010 funding period is gratefully

acknowledged. Also, partial financial support from Houston

Area Research Center (HARC) is gratefully acknowledged.

Use of the Center for Nanoscale Materials was supported by

the US Department of Energy, Office of Science, Office of

Basic Energy Sciences, under contract No. DE-AC02-

06CH11357. We acknowledge computer time at the Labora-

tory Computing Resource Center (LCRC) at Argonne

National Laboratory, The National Energy Research Scien-

tific Computing Center (NERSC), and the Molecular Science

Computing Facility (MCSF) at Pacific Northwest National

Laboratory.

References

1. S. Koh and P. Strasser, J. Am. Chem. Soc., 2007, 129,12624–12625.

2. H. Lechtman, Sci. Am., 1984, 250, 56–63.3. A. J. Forty, Nature, 1979, 282, 597–598.4. J. Erlebacher, M. J. Aziz, A. Karma, N. Dimitrov and

K. Sieradzki, Nature, 2001, 410, 450–453.5. H. W. Pickering, Corrosion Sci., 1983, 23, 1107–1120.6. H. W. Pickering and C. Wagner, J. Electrochem. Soc., 1967, 114,

698.

7. D. V. Pugh, A. Dursun and S. G. Corcoran, J. Electrochem. Soc.,2005, 152, B455–B459.

8. M. B. Vukmiovic, J. Zhang, K. Sasaki, A. U. Nilekar, F. Uribe,M. Mavrikakis and R. R. Adzic, Electrochim. Acta, 2007, 52,2257–2263.

9. R. R. Adzic, J. Zhang, K. Sasaki, M. B. Vukmirovic, M. Shao,J. X. Wang, A. U. Nilekar, M. Mavrikakis, J. A. Valerio andF. Uribe, Top. Catal., 2007, 46, 249–262.

10. J. Zhang, M. B. Vukmiovic, Y. Xu, M. Mavrikakis andR. R. Adzic, Angew. Chem., Int. Ed., 2005, 44, 2132.

11. J. Zhang, Y. Mo, M. Vukmirovic, R. Klie, K. Sasaki andR. Adzic, J. Phys. Chem. B, 2004, 108, 10955–10964.

12. S. R. Brankovic, J. X. Wang and R. R. Adzic, Electrochem. Solid-State Lett., 2001, 4, A217–A220.

13. L. Bengtsson, Phy. Rev. B, 1999, 59, 12301–12304.14. R. Srivastava, P. Mani, N. Hahn and P. Strasser, Angew. Chem.,

Int. Ed., 2007, 46, 8988–8991.15. P. Mani, R. Srivastava and P. Strasser, ECS Trans., 2007, 11,

933–940.16. N. Hahn, S. Koh and P. Strasser, ECS Trans., 2006, 3, 139–140.17. P. Mani, S. Srivastava and P. Strasser, J. Phys. Chem. C, 2008,

112, 2770–2778.18. T. J. Schmidt and H. A. Gasteiger, Rotating thin-film method for

supported catalysts, in Handbook of Fuel Cells—Fundamentals,Technology, and Applications, ed. W Vielstich, A Lamm andH Gasteiger, Wiley, Chichester, 2003, ch. 22, vol. 2, p. 316.

19. T. J. Schmidt, H. A. Gasteiger and R. J. Behm, J. Electrochem.Soc., 1999, 146, 1296–1304.

20. T. J. Schmidt, H. A. Gasteiger, G. D. Stab, P. M. Urban,D. M. Kolb and R. J. Behm, J. Electrochem. Soc., 1998, 145,2354–2358.

21. J. Friedel, Philos. Mag., 1955, 46, 514.22. L. Vegard, Z. Phys., 1921, 5, 2–26.23. B. Hammer, L. B. Hansen and J. K. Nørskov, Phys. Rev. B:

Condens. Matter Mater. Phys., 1999, 59, 7413–7421.24. CRC Handbook of Chemistry and Physics, CRC Press, New York,

76th edn, 1996.25. D. Vanderbilt, Phys. Rev. B: Condens. Matter Mater. Phys., 1990,

41, 7892–7895.26. G. Kresse and J. Furthmuller, Comput. Mater. Sci., 1996, 6,

15–50.27. J. Greeley and J. K. Nørskov, Electrochim. Acta, 2007, 52,

5829–5836.28. J. K. Norskov, J. Rossmeisl, A. Logadottir, L. Lindqvist,

J. R. Kitchin, T. Bligaard and H. Jonsson, J. Phys. Chem. B,2004, 108, 17886–17892.

29. M. Pourbaix, Atlas of Electrochemical Equilibria in AqueousSolutions, Pergamon Press, Oxford, New York, 1966.

30. H. A. Gasteiger, S. S. Kocha, B. Sompalli and F. T. Wagner, Appl.Catal., B, 2005, 56, 9–35.

31. U. A. Paulus, A. Wokaun, G. G. Scherer, T. J. Schmidt,V. Stamenkovic, N. M. Markovic and P. N. Ross, Electrochim.Acta, 2002, 47, 3787–3798.

32. U. A. Paulus, A. Wokaun, G. G. Scherer, T. J. Schmidt,V. Stamenkovic, V. Radmilovic, N. M. Markovic andP. N. Ross, J. Phys. Chem. B, 2002, 106, 4181–4191.

33. V. Stamenkovic, T. J. Schmidt, P. N. Ross and N. M. Markovic,J. Phys. Chem. B, 2002, 106, 11970–11979.

34. V. Stamenkovic, T. J. Schmidt, P. N. Ross and N. M. Markovic,J. Electroanal. Chem., 2003, 554, 191–199.

35. S. Koh, M. F. Toney and P. Strasser, Electrochim. Acta, 2007, 52,2765–2774.

36. P. Strasser, S. Koh and C. Yu, ECS Trans., 2007, 11, 167–180.37. S. Mukerjee, S. Srinivasan, M. P. Soriaga and J. McBreen,

J. Phys. Chem., 1995, 99, 4577.38. S. Mukerjee, S. Srinivasan, M. P. Soriaga and J. McBreen,

J. Electrochem. Soc., 1995, 142, 1409–1422.39. S. Mukerjee and S. Srinivasan, J. Electroanal. Chem., 1993, 357,

201.40. N. M. Markovic and P. N. Ross, Surf. Sci. Rep., 2002, 45, 117.41. E. Antolini, J. e. R. C. Salgado and E. R. Gonzalez, J. Power

Sources, 2006, 155, 161–166.42. F. H. B. Lima, W. H. E. Lizcano-Valbuena, E. Teixeira-Neto,

F. C. Nart, E. R. Gonzalez and E. A. Ticianelli, Electrochim. Acta,2006, 52, 385–393.

3682 | Phys. Chem. Chem. Phys., 2008, 10, 3670–3683 This journal is �c the Owner Societies 2008

Publ

ishe

d on

27

May

200

8. D

ownl

oade

d by

TU

Ber

lin -

Uni

vers

itaet

sbib

l on

13/0

4/20

15 1

3:24

:20.

View Article Online

Page 16: Physical Chemistry Chemical Physics This paper is ......25 alloy nanoparticles annealed at both low and high temperature. We also assess the impact of the synthesis conditions on the

43. N. Travitsky, T. Ripenbein, D. Golodnitsky, Y. Rosenberg,L. Burshtein and E. Peled, J. Power Sources, 2006, 161, 782.

44. J. N. Soderberg, A. H. C. Sirk, S. A. Camphell and V. I. Birss,J. Electrochem. Soc., 2005, 152, A2017–A2022.

45. L. Xiong and A. Manthiram, J. Electrochem. Soc., 2005, 152,A697–A703.

46. D. Thompsett, Pt alloys as oxygen reduction catalysts, in Hand-book of Fuel Cells—Fundamentals, Technology and Applications,ed. W Vielstich, A Lamm and H. A. Gasteiger, Wiley, Chichester,2003, ch. 37, vol. 3, p. 467.

47. E. Antolini, R. R. Passos and E. A. Ticianelli, Electrochim. Acta,2002, 48, 263–270.

48. M.-k. Min, J. Cho, K. Cho and H. Kim, Electrochim. Acta, 2000,45, 4211–4217.

49. G. Tamizhmani and G. A. Capuano, J. Electrochem. Soc., 1994,41, 968.

50. C. Yu, S. Koh, J. Leisch, M. T. Toney and P. Strasser, FaradayDicuss., 2008, 140, in print.

51. C. W. B. Bezerra, L. Zhang, H. Liu, K. Lee, A. e. L. B. Marques,E. P. Marques, H. Wang and J. Zhang, J. Power Sources, 2007,173, 891–908.

52. A. Cebollada, R. F. C. Farrow and M. F. Toney, Structure andmagnetic properties of chemically ordered magnetic binary alloysin thin film form, in Magnetic Nanostructures, ed. H. S. Nalwa,American Scientific, Stevenson Ranch, 2002, pp. 93–123.

53. ICDD, Power Diffraction File (PDF) 2—data CD. In Interna-tional Center for Diffraction Data—Power Diffraction File (PDF)2 data CD http://www.icdd.com/: 2002.

54. Z. Peng, C. Chang and L. Ximen, Acta Geol. Sinica, 1978, 4,326–336.

55. Z. Yu, Bull. Inst. Geol., Chin. Acad. Sci., 1982, 4, 75–81.56. K. Ding, Scientia Geol. Sinica, 1980, 2, 168–171.57. Y. Ding, M. Chen and J. Erlebacher, J. Am. Chem. Soc., 2004,

126, 6876–6877.58. K. Sieradzki, N. Dimitrov, D. Movrin, C. McChall, N.

Vasijevic and J. Erlebacher, J. Electrochem. Soc., 2002, 149,B370–B377.

59. J. Erlebacher, J. Electrochem. Soc., 2004, 151, C614–C626.60. J. Rugolo, J. Erlebacher and K. Sieradzki, Nature, 2006, 5,

946–949.61. G. Ertl, H. Knozinger and J. Weitkamp, Handbook of Hetero-

geneous Catalysis, Wiley-VCH, Weinheim, 1997, vol. 1, p. 426.62. Z. Liu, S. Koh, C. Yu and P. Strasser, J. Electrochem. Soc., 2007,

154, B1192–B1199.63. Z. Liu, C. Yu, I. Rusakova, D. Huang and P. Strasser, Top.

Catal., 2007, DOI: 10.1007/s11244-008-9083-2, in press.64. J. Xie, D. L. Wood, K. L. More, P. Atanassov and R. L. Borup,

J. Electrochem. Soc., 2005, 152, A1011–A1020.65. Y. Shao-Horn, W. C. Sheng, S. Chen, P. J. Ferreira, E. F. Holby

and D. Morgan, Top. Catal., 2007, 46, 285–305.66. P. J. Ferreira, G. J. laO’, Y. Shao-Horn, D. Morgan,

R. Makharia, S. Kocha and H. Gasteiger, J. Electrochem. Soc.,2005, 152, A2256–2271.

67. D. L. Fedlheim and C. A. Foss, Metal Nanoparticles: SynthesisCharacterization and Applications, CRC – Taylor Francis Group,London, 2001.

68. R. G. Finke, Transition-Metal Nanoclusters: Solution-PhaseSynthesis, Characterization and Mechanism of Formation, inMetal Nanoparticles: Synthesis, Characterization, and Applica-

tions, ed. D. L. Feldheim and C. A. Foss, Jr, Marcel Dekker,Inc., New York, 2002, pp. 17–54.

69. G. Schmid, Nanoparticles: From Theory to Applications, Wiley-VCH, New York, 2004.

70. S. Koh, C. Yu and P. Strasser, ECS Trans., 2007, 11, 205–215.71. S. Koh, M. F. Toney and P. Strasser, Lattice-strained Pt shell

Nanoparticle Catalysts for the Electroreduction of Oxygen atPEMFC cathodes, in American Institute of Chemical Engineers(AIChE) annual meeting abstracts, San Francisco, 2006.

72. S. Mukerjee, S. Srinivasan, M. P. Soriaga and J. McBreen,J. Phys. Chem., 1995, 99, 4577–4589.

73. D. Thompsett, Catalysts for the Proton Exchange Membrane FuelCell, in Fuel Cell Technology Handbook, CRC PressLLC, London,New York, 2003, ch. 6.

74. V. Stamenkovic, B. S. Mun, M. Arenz, K. J. J. Mayerhofer,C. A. Lucas, G. Wang, P. N. Ross and N. Markovic, Nat. Mater.,2007, 6, 241.

75. V. R. Stamenkovic, B. Fowler, B. S. Mun, G. Wang, P. N. Ross,C. A. Lucas and N. M. Markovic, Science, 2007, 315, 493.

76. M. T. Paffet, J. G. Beery and S. Gottesfeld, J. Electrochem. Soc.,1988, 135, 1431–1436.

77. H. Liu, P. He, Z. Li and J. Li, Nanotechnology, 2006, 17.78. V. R. Stamenkovic, B. S. Mun, K. J. J. Mayrhofer, P. N. Ross and

N. M. Markovic, J. Am. Chem. Soc., 2006, 128, 8702–8988.79. A. V. Ruban, H. L. Skriver and J. K. Nørskov, Phys. Rev. B:

Condens. Matter Mater. Phys., 1999, 59, 15990–16000.80. D. V. Pugh, A. Dursun and S. G. Corcoran, J. Mater. Res., 2003,

18, 216.81. V. Stamenkovic and N. Markovic, Langmuir, 2001, 17,

2388–2394.82. N. Markovic, H. A. Gasteiger and P. N. Ross, Langmuir, 1995, 11,

4098–4108.83. N. Markovic and P. N. Ross, Langmuir, 1993, 9, 580–590.84. K. Sieradzki, R. R. Corderman, K. Shukla and R. C. Newman,

Philos. Mag. A, 1989, 59, 713–746.85. K. Sieradzki, J. Electrochem. Soc., 1993, 140, 2868–2872.86. R. Li and K. Sieradzki, Phys. Rev. Lett., 1992, 68, 1168–1172.87. K. Wagner, S. R. Brankovic, N. Dimitrov and K. Sieradzki,

J. Electrochem. Soc., 1997, 144, 3545–3555.88. R. C. Newman and K. Sieradzki, Science, 1994, 263, 1708–1709.89. F. U. Renner, Y. Grunder, P. F. Lyman and J. Zegenhagen, Thin

Solid Films, 2007, 515, 5574.90. F. U. Renner, A. Stierle, H. Dosch, D. M. Kolb and

J. Zegenhagen, Electrochem. Commun., 2007, 9, 1639.91. F. U. Renner, A. Stierle, H. Dosch, D. M. Kolb, T. L. Lee and

J. Zegenhagen, Nature, 2006, 430, 707–710.92. S. G. Corcoran, The Morphology of Alloy Corrosion, in Critical

Factors in Localized Corrosion III, eds. R. G. Kelley,G. S. Frankel, P. M. Natishan and R. C. Newman, Electroche-mical Society, Pennington, NJ, 2000, vol. 98-17, pp. 500–507.

93. J. R. Davis, Corrosion—Understanding the basics, ASM Interna-tional, Materials Park, Ohio, 2000.

94. A. Dursun, D. V. Pugh and A. G. Corcoran, J. Electrochem. Soc.,2005, 152, B65–B72.

95. E. Skulason, G. S. Karlberg, J. Rossmeisl, T. Bligaard, J. Greeley,H. Jonsson and J. K. Nørskov, Phys. Chem. Chem. Phys., 2007, 9,3241–3250.

96. E. Leiva, Electrochim. Acta, 1996, 41, 2185–2206.97. C. G. Sanchez and E. P. M. Leiva, Langmuir, 2001, 17, 2219–2227.

This journal is �c the Owner Societies 2008 Phys. Chem. Chem. Phys., 2008, 10, 3670–3683 | 3683

Publ

ishe

d on

27

May

200

8. D

ownl

oade

d by

TU

Ber

lin -

Uni

vers

itaet

sbib

l on

13/0

4/20

15 1

3:24

:20.

View Article Online