You are on page 1of 18

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/229757523

Oxygen reduction reaction on smooth single crystal electrodes

Chapter · December 2010


DOI: 10.1002/9780470974001.f205035

CITATIONS READS

14 129

1 author:

Philip N Ross
University of California, Berkeley
410 PUBLICATIONS   29,408 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Surface chemistry of silicon anodes in lithium-ion batteries View project

Electrochemistry NMR View project

All content following this page was uploaded by Philip N Ross on 24 October 2017.

The user has requested enhancement of the downloaded file.


Oxygen reduction reaction on smooth single crystal
electrodes
P. N. Ross Jr
Volume 2, Part 5, pp 465–480

in

Handbook of Fuel Cells – Fundamentals, Technology and Applications


(ISBN: 0-471-49926-9)

Edited by

Wolf Vielstich
Arnold Lamm
Hubert A. Gasteiger

 John Wiley & Sons, Ltd, Chichester, 2003


Chapter 31
Oxygen reduction reaction on smooth single crystal
electrodes

P. N. Ross Jr
University of California, Berkeley, CA, USA

1 INTRODUCTION effects). In real systems, all of these factors will, in gen-


eral, operate simultaneously, and separating these effects
and assessing their relative importance in the catalytic
In spite of many attempts in the last decade by researchers activity and reaction mechanism is a very challenging
and fuel cell developers to create a nonPt catalyst for problem.
low temperature (<200 ◦ C) air cathodes, Pt remains the The situation in alkaline electrolyte is rather different
catalyst of choice at least for acid-based fuel cells. There- from acid electrolyte. Here there are many opportunities for
fore, much of the art and science of catalyst develop- nonPt catalysts for the ORR.[2, 3] Materials that have little
ment for the oxygen reduction reaction (ORR) rely on or no measurable activity for oxygen reduction in acid have
both the fundamental understanding of the reaction at substantial, even commercially practical, levels of activity
the Pt-electrolyte interface[1–4] and the optimization of in alkaline solution.[3] Although this fact has been known
the catalytic properties of the Pt surface.[5, 6] In prin- for some time, the fundamental reason for this remarkable
cipal, fundamental understanding can be used to create pH effect has never been fully explained. I present here,
tailormade surfaces with improved catalytic activity. The for the first time, a simple thermodynamic rationale based
optimization of the most promising catalysts must incor- on the energetics of the O2 /O2 − (superoxide radical anion)
porate both noncatalytic and catalytic factors. Examples redox couple.
of noncatalytic factors are partial replacement of nonsur- In this chapter, I will summarize recent progress in
face Pt atoms, i.e., ‘buried’ Pt atoms, with an inexpen- understanding the role of surface structure in the kinetics
sive metal, or the maximization of the Pt surface area of the ORR on well-defined metal surfaces in both acid
while exposing the most active microstructures. These fac- and alkaline electrolyte. In acid, the focus will be on the
tors will minimize the amount (and cost) of catalysts low index Pt single crystals, and relationship between the
required for a given level of activity. The most impor- Pt surface structure on the kinetics of the ORR in order to
tant catalytic factor is the modification of the intrinsic create a fundamental link between the specific activity of
activity of Pt surface atoms, usually by making bimetal- Pt (rate per unit area) and the particle size (for various
lic surfaces. The enhancement of catalytic properties by nanoparticle shapes) of supported catalysts. In alkaline
the second metal may occur through the change of the electrolyte, the focus is on the low index single crystal
local bonding geometry (structure effects), the distribution surfaces of Au in addition to Pt. Au is discussed as a
of active sites (ensemble effects) or directly by modify- prototype electrode material for a nonPt catalyst with weak
ing the reactivity of platinum surface atoms (electronic adsorption of O2 − .

Handbook of Fuel Cells – Fundamentals, Technology and Applications, Edited by Wolf Vielstich, Hubert A. Gasteiger, Arnold Lamm.
Volume 2: Electrocatalysis.  2003 John Wiley & Sons, Ltd. ISBN: 0-471-49926-9.
466 Part 5: The oxygen reduction/evolution reaction

2 OXYGEN REDUCTION REACTION via an (H2 O2 )ad intermediate, and I will use that pathway
PATHWAY/MECHANISM for the discussion here. The surface peroxide intermediate
may or may not be further reduced to water depending on
In order to discuss the structure sensitivity of the ORR on the relative values of k4 and k5 . In either case, the rate-
metal surfaces and to attribute effects to specific surface determining step appears to be the addition of the first
electron to O2,ad to form the adsorbed superoxide radical
chemistry, it is convenient to formulate the discussion in
terms of a reaction pathway/mechanism. Oxygen reduction anion, žO2,ad −
reaction is a multi-electron reaction that may include a
O2,ad + e− −−−→ žO2,ad − (1)
number of elementary steps involving different reaction
intermediates. In his excellent recent review, Adzic[3] has
The rate expression is then,[2, 12]
summarized the many proposed paths and intermediates
and discussed their merits and deficiencies with respect to
i = nF KcO2 (1 − ad )x exp(−βF E/RT )
various electrode materials and the available experimental
data. For the purpose here of discussing the reaction on × exp(−γrad )/RT (2)
different metals at different pH in a common framework, a
more generalized and simplified scheme is more useful than where n is the number of electrons, K is the chemical
the more complex schemes discussed by Adzic. The overall rate constant, cO2 is the concentration of O2 in the solu-
scheme valid for both acid and base and all the electrode tion, ad is the total surface coverage by all adsorbed
materials discussed here is the following: species, x is either 1 or 2 depending on the site require-
ments of the adsorbates, i is the observed current, E is
k1
the applied potential, β and γ are the symmetry factors
(assumed to be 1/2) and r (the so-called Frumkin or Temkin
k2 k3 parameter) is a parameter characterizing the rate of change
O2 O2, ad H2O2, ad H2O
of the free energy of adsorption of the rate-determining
k4
k5 intermediate with the surface coverage by (all) adsorbing
species. Assuming that the coverage of ORR intermediates
H2O2 is low under reaction conditions,[12] for most cases only two
adsorbed species need to be considered, OHad and specif-
Based on this reaction scheme, O2 can be electrochemi- ically adsorbing anions like (bi)sulfate or halides. In the
cally reduced either directly to water with the rate constant following sections, we will use this reaction pathway and
k1 without intermediate formation of H2 O2,ad (so-called rate expression to analyze the effects of various factors on
‘direct’ 4e− reduction) or to H2 O2,ad with the rate con- the kinetics of the ORR on Pt(hkl) surfaces.
stant k2 (‘series’ 2e− reduction). The adsorbed peroxide One particular limiting case is of special interest, that
can be electrochemically reduced to water with the rate where the free energy of adsorption of the rate-determining
constant k3 (‘series’ 4e− pathway), catalytically (chemi- intermediate, žO2,ad − , is unaffected by the co-adsorption
cally) decomposed on the electrode surface (k4 ) or desorbed of the spectator anions like (bi)sulfate or halides, r = 0.
into the bulk of the solution (k5 ). The direct 4e− reduc- Under this assumption, the second exponential term is a
tion requires dissociation of oxygen prior to the transfer potential independent constant and the so-called Tafel slope,
of the first electron.[7] The dissociation energy of O2 is (∂ log i/∂E)−1 , is −2 × 2.3 RT /F mV/decade of current.
quite large (498.3 kJ mol−1 ), meaning that dissociation is
energetically unfavorable unless the M–O bond is very
strong, e.g., >250 kJ mol−1 . Metals with very strong M–O 3 ORR AT Pt(hkl) SURFACES IN ACIDIC
bond strengths have poor activity for the ORR[5] because ELECTROLYTE
the reaction stops with the formation of surface OH, i.e.,
reaction becomes desorption rate limiting. The more ener- Interest in the role of the local symmetry of Pt surface atoms
getically favored path is the superoxo/peroxo path, with in electrocatalysis has risen steadily over the last three
transfer of the electron to the oxygen molecule. The dis- decades. Ever since the pioneering work of Will,[13] one
sociation energy of O2 − and/or O2 2− is much lower than of the major themes in electrochemical research has been
that of O2 (by 98.7 kJ mol−1[8] ), resulting in a more facile the study of the relationship between the electrochemical
reaction path without a strong M–O bond. reactivity and surface structure, a functionality generally
In recent studies on Pt surfaces from our laboratory,[9–11] termed ‘structural sensitivity’.[4] The recent advance in the
we have interpreted the results assuming the series pathway in-situ methods of surface X-ray scattering (SXS)[14] and
Oxygen reduction reaction on smooth single crystal electrodes 467

scanning tunneling microscopy (STM)[15] have made it


100
possible to establish the true surface structure of Pt single
25 °C

Ir (µA)
crystal surfaces in electrolyte as a function of potential. 50
This advance enables one to determine with confidence the
true relationship between the kinetics of electrochemical 0.0
reactions and structure of the surface.[4]
It is now well established that the structure sensitivity of
the ORR on Pt(hkl) surfaces depends critically on the nature 0.5

Id (mA)
of the supporting electrolyte. Within the limited scope of
this report, it will not be possible to review all of these
1.0
results. Rather I will show, using representative examples,
the kind of information that can be used to improve our
understanding of the role of the symmetry of Pt surface 1.5
atoms in the O2 reduction reaction. The most outstanding 100
conclusions from this body of work is: (i) that the structural Pt(111)

Ir (mA)
Pt(100) 60 °C
sensitivity is dramatically affected by the adsorption of 50
anions from the supporting electrolyte, e.g., HSO4(ad) , Clad , Pt(110)
and Brad , and (ii) that the structure sensitivity in the solution 0.0
containing bisulfate or halide anion is determined entirely
by the (1–ad ) coverage dependent pre-exponential term.
0.5
Two representative sets of polarization curves for Pt(hkl)
Id (mA)

in 0.05 M H2 SO4 (pH = 1) are shown in Figure 1 for two


different temperatures. The polarization curves clearly show 1.0
that the ORR in H2 SO4 increases in the sequence Pt(111) <
Pt(100) < Pt(111). An exceptionally large deactivation is
observed at the Pt(111) surface. A full description of the 1.5
0.0 0.2 0.4 0.6 0.8 1.0
polarization curves for the ORR on Pt(111) in pure sulfuric
E (V)
acid solution is given in our previous publications.[10, 16]
The exceptionally large deactivation is due the strong Figure 1. (a) Oxygen reduction on Pt(hkl) in 0.05 M H2 SO4
adsorption of the (bi)sulfate anion from the symmetry match at 298 K (20 mV s−1 , 900 rpm); lower part, disk currents; upper
between the fcc (111) face (see insert of Figure 1) and the part, ring currents. (b) Oxygen reduction on Pt(hkl) at 333 K
C3v geometry of the oxygen atoms of the sulfate anion.[17] 250 mV s−1 , 900 rpm): lower part, disk currents; upper part, ring
currents. (From Ref. [10].)
The fact that the activity of all three low-index platinum
planes is significantly higher in HClO4 [4] does, however,
suggest that adsorption of (bi)sulfate anions affects the The family of polarization curves for the ORR on Pt(hkl)
kinetics of the ORR on all three surfaces, but is particularly in 0.05 M H2 SO4 + 10−3 M Cl− are shown in Figure 2.
strong on Pt(111). It should be noted that, although bisul- The polarization curves for the ORR in pure 0.05 M H2 SO4
fate adsorption onto Pt(hkl) surfaces inhibits the reduction solutions is shown for one rotation rate as a reference. In
of molecular O2 , most likely by blocking the initial adsorp- the presence of Cl− , the catalytic activity increases in the
tion of O2 , it does not affect the pathway of the reaction, order (100) < (110) < (111).[9] Figure 2a shows that the
since no H2 O2 is detected on the ring electrode for any of ORR on the Pt(111) surface in 0.05 M H2 SO4 + 10−3 M
the surfaces in the kinetically controlled potential region. Cl− is strongly inhibited, with an activity in a solution
Tafel plots of the mass-transfer corrected kinetic currents containing Cl− being orders of magnitude lower than in
showed[10] that the Tafel slope has the ‘ideally’ temper- 0.1 M HClO4 ,[4, 9] and about the same inhibition as in Cl−
ature dependent value of −2 × 2.3 (RT /F ) on all three free 0.05 M H2 SO4 . If one were to examine just the ORR
surfaces. The activation energy was also independent of polarization curve alone, one might conclude that the (111)
crystal face, having the value of 42 kJ mol−1 (extrapolated surface is hardly affected by the presence of Cl− in solution
to the reversible potential). Since the energetic terms in the due to the already strong adsorption of (bi)sulfate. But in
rate expression (2) are all independent of crystal face, the fact the base voltammetry in O2 free electrolyte makes it
structure sensitivity of the ORR in sulfuric acid must arise clear that Cl− displaces (bi)sulfate anions entirely from the
entirely from the pre-exponential term, (1 − ad ), i.e., a surface throughout the potential region shown.[18] Unlike
purely site-blocking effect of the adsorbed (bi)sulfate anion. Pt(111), the ORR on Pt(110) (Figure 2b) and Pt(100)
468 Part 5: The oxygen reduction/evolution reaction

2Ir /N
xH2 O2 = (3)
Id + Ir /N
50
Ir (µA)

is almost constant and is close to 25% (note in this relation,


0
N denotes the collection efficiency of the ring-disk configu-
ration, Id and Ir are the disk and ring currents, respectively).
Id (mA cm−2)

−2
At potentials below 0.3 V, the surface coverage with Clad is
−4 reduced substantially, but the O2 reduction currents sharply
−6
Pt(111) decrease to diffusion limiting currents for ca. 2e− per O2
molecule at the negative potential limit. In the same poten-
(a) tial region, the O2 reduction currents on the disk electrode
are accompanied quantitatively by the H2 O2 oxidation cur-
Id (mA cm−2) Ir (µA), H2O2 (%)

rents on the ring electrode, reaching maximum peroxide


2 production of ca. 90% at 0.075 V. In Cl− -free solution,
0
however, a 3e− reduction is observed at the same potential
limit.[16] This may suggest that Clad is co-adsorbed with
−2 Hupd on the (100) surface, and thus the concentration of
empty Pt-pair sites required for the breaking of the O–O
−4 bond is significantly reduced so that the ORR proceeds
Pt(110) entirely through the 2e− peroxide pathway. The amount
−6
of peroxide formed on the Pt(110) surface is substantially
smaller (Figure 2b) ca. 2.5%, and no increase in peroxide
(b) formation was observed in the Hupd region.
The re-ordering of the structure sensitivity of the ORR
0.05 M H2SO4 + 10−3 M Cl− in the presence of Cl− versus (bi)sulfate can be explained
Id (mA cm−2) Ir (µA), H2O2 (%)

0.05 M H2SO4 qualitatively by analyzing the relative strengths of the


50 Pt–Clad interaction, the specific type of sites for Clad , and
the effect of Clad at these sites on the rate of O–O bond
0 breaking and/or bond making. In particular, due to the
strong interaction of Clad with Pt(100), the ORR between
−2
0.3 and 0.75 V is inhibited and the number of platinum sites
−4 required for adsorption of O2 and the breaking of O–O bond
Pt(100) are reduced, which is in accord with the peroxide formation
−6
in the same potential region. In the case of Pt(110), although
0.2 0.4 0.6 0.8 the interaction of Clad with Pt(110) is as strong as with
(c) E (V vs. RHE) (100) sites, Clad is most likely adsorbed in the ‘troughs’
of the surface (insert of Figure 1), thus leaving the top
Figure 2. Ring and Disk currents for the ORR in 0.05 M sites available for O2 adsorption and the cleavage of the
H2 SO4 + 10−3 M Cl− on (a) Pt(111), (b) Pt(110) and (c) Pt(100) O–O bond. Apparently, the Pt(111)–Clad system has the
at different rotation rates (full lines). The respective polarization same effect on the ORR as (bi)sulfate anions, which is
curves on Pt(hkl) in 0.05 M H2 SO4 are shown as a reference
(dotted line). The gray circles represent the fraction of peroxide
somewhat surprising given the differences in ionic radii
formed during the ORR; 50 mV s−1 , 293 K. (From Ref. [27].) of these anions, e.g., Cl− is about 75% that of sulfate or
bisulfate.[20]
Very similar results are observed for the Pt(hkl)/Br− sys-
tem as for Cl− , but in fact this system is even better under-
(Figure 2c) in Cl− containing solution is accompanied by stood, since the Brad coverages have been measured quan-
peroxide formation, implying that within the same potential titatively using a combination of SXS[21] and the rotating
limits H2 O2 is formed as an intermediate on the Pt(110) and ring-disk electrode (RRDE) current-shielding method.[18]
Pt(100) electrodes covered with Clad , but not on Pt(111) Examples are shown in Figures 3 and 4 for Pt(111) and
covered by (bi)sulfate. On Pt(100) the fraction of H2 O2 (100), respectively. Figure 3(a) shows cyclic voltammo-
(current efficiency) produced between 0.25 < E < 0.6 V grams for the Pt(111) surface in pure 0.1 M HClO4 solution
during the ORR, calculated from Figure 2(c) using[19] and in 0.1 M HClO4 + 10−4 M Br− under a rotation rate of
Oxygen reduction reaction on smooth single crystal electrodes 469

900 rpm Pt(100)

20 µA
10 µA 0

0.5
0.4

θBr (ML)
0.4 0.3
0.2 0.1 M HClO4
0.3
0.1 +10−4 M Br−
θBr (ML)

(a) 0.0

0.2 60 2500
Pt(111)

Ir (µA)
30
0.1 M HCIO4
0.1
0.1 M HCIO4 + 8 × 10−3 M Br 0 ω (rpm)
H ads/des on Pt(111)/Br 400
0.5 900
(a) 0 Pt(111)/Br
1600
1.0 2500
Id (mA)

1.5
150 2500
1600 2.0
Ir (µA)

100 900
400 −0.4 −0.2 0.0 0.2 0.4 0.6
50
(b) E (V vs. SCE)
rpm
0
400 Figure 4. (a) Cyclic voltammetry of Pt(100) in the RRDE
assembly in oxygen-free electrolyte: ( ) 0.1 M HClO4 +
−Id (mA)

1.0
2500 1 × 10−4 M Br− and (- - - -) 0.1 M HClO4 at a sweep rate of
50 mV s−1 ; (-Ž-Ž-) Brad adsorption isotherm; (b) Disk (Id ) and
ring (Ir ) currents during oxygen reduction on Pt(111) in 0.1 M
HClO4 (- - - -) and in 0.1 M HClO4 + 8 × 10−4 M Br− ( )
2.0 at a sweep rate of 50 mV s−1 (ring potential = −0.95 V).
−0.2 0 0.2 0.4 0.6
(b) E (V)
the Brad isotherm reveals that at the negative potential limit,
Figure 3. (a) Cyclic voltammetry of Pt(111) in the RRDE the Brad coverage goes to zero and the saturation coverage
assembly in oxygen-free electrolyte: ( ) 0.1 M HClO4 + by Hupd is identical to that in the Br− free electrolyte. To
8 × 10−4 M Br− and (- - - -) 0.1 M HClO4 at a sweep rate of confirm this, the hydrogen adsorption/desorption currents
50 mV s−1 ; ( . . . . ) deconvolution of the contribution from hydro- were deconvoluted from the total disk currents (the total
gen adsorption/desorption; (-·-·-·) Brad adsorption isotherm; (b)
Disk (Id ) and ring (Ir ) currents during oxygen reduction on Pt(111) disk charge Qtd ≈ 265 µC cm−2 ) by subtracting the bromide
in 0.1 M HClO4 (- - - -) and in 0.1 M HClO4 + 8 × 10−4 M Br− flux measured with the ring-shielding experiments (details
( ) at a sweep rate of 50 mV s−1 (ring potential = −0.95 V). of this deconvolution were reported previously[18] ). We
(From Ref. [18b].) found that at 0.55 V the charge associated with adsorption
−2
of bromide, QBr d , is ≈100 µC cm , equivalent to a cover-
900 rpm. The bottom section of Figure 3(a) shows the age ≈0.42 ML Brad per Pt for fully discharged atoms. This
adsorption isotherm for Br− on the Pt(111) surface. For is close to saturation coverage based on the van der Waals
coverages above about 0.33 ML (1 ML = 1 Brad per Pt), radius of Br atoms and the surface density of Pt atoms. The
−2
the Brad adlayer forms a series of ordered structures that difference between Qtd and QBr d amounts to 170 µC cm or
−2
are interesting in themselves but not relevant to the present to ≈160 µC cm if corrected by a double-layer capacitance
topic. The interested reader is referred to other publications of ≈20 µF cm−2 . This is essentially equal to the pseudoca-
specifically about these structures.[21] A close inspection of pacitance in the 0–0.3 V region of the Pt(111) surface in
470 Part 5: The oxygen reduction/evolution reaction

Br− - free solution, thus this charge is attributed to Hupd . high vacuum (UHV) annealed crystal is immersed in elec-
The resulting potentiodynamic curve for Hupd on Pt(111) in trolyte. However, the Brad coverage has been measured
solution containing Br anions is shown in Figure 3a (dot- using the current-shielding RRDE method, and the adsorp-
ted line), and demonstrates the reduced potential region of tion isotherm along with the base voltammetry is shown
stability of Hupd on Pt(111) in the presence of Br− due to in Figure 4(a). As with Pt(111), the potential region for
competitive adsorption. Hupd is dramatically reduced by competitive adsorption of
Figure 3(b) summarizes a family of polarization curves Brad , but unlike the (111) surface Brad is present on the
for the ORR on Pt(111) in 0.1 M HClO4 + 10−4 M Br− (100) surface at and negative to the H2 electrode potential
along with a representative polarization curve recorded in (−0.244 V versus SCE). The ORR polarization curves are
a solution free of Br− anions. In all experiments, the ring shown in Figure 4(b). There is no observable rate of O2
was potentiostated at 0.95 V, where oxidation of peroxide reduction until the coverage by Brad is about 90% of the
arriving at the ring is under diffusion control. In pure per- saturation coverage, i.e., 10% of the Brad is desorbed. Brad
chloric acid, starting at ca. 0.6 V and sweeping the disk causes the half-wave potential to be shifted negatively by
potential negatively to 0.0 V, the ring currents negligible, about 0.5 V, about the same as on the (111) surface, and
implying that even on the surface covered with reversible the reaction proceeds entirely through the production of
adsorbed OHad (the so-called ‘butterfly region’) oxygen peroxide intermediate in solution. Unlike the effect of Clad ,
reduction proceeds without the production of peroxide in however, Brad results in virtually no structure sensitivity for
solution. The appearance of peroxide oxidation currents on the ORR, with the reaction proceeding with nearly identical
the ring electrode begins at potentials negative of 0.0 V polarization curves on both the (111) and (100) surfaces.
saturated calomel electrode (SCE), and parallels with the
adsorption of hydrogen on Pt(111). At the negative poten-
tial limit the limiting current corresponds to an exactly 4 IMPLICATIONS OF Pt(hkl) RESULTS
two-electron reduction of O2 . Figure 3(b) shows that the FOR FUEL CELLS
ORR on the Pt(111) surface is strongly inhibited by Brad ,
with an activity orders of magnitude lower than in pure The structure sensitivity of the ORR in sulfuric acid found
perchloric acid. Figure 3(b) shows that the O2 reduction at using Pt(hkl) electrodes provides some new insight into the
the Pt(111)–Br interface starts at ca. 0.3 V (θBr ≈ 0.4 ML ‘particle size effects’ reported for the ORR using supported
and θOH ≈ 0.0 ML in Figure 1(a)) and is under com- Pt catalysts. Unfortunately, the observations of structure
bined kinetic-diffusion control in the wide potential region, sensitivity of the ORR are not at present as useful as
0.1 < E < 0.3 V. Figure 3(b) also shows that peroxide was they might be because the experimental conditions, e.g.,
detected on the ring electrode in the same potential region, temperature and electrolyte, used in most ORR kinetic mea-
indicating that the H2 O2 is formed as an intermediate on surements with supported catalysts are not those employed
the electrode covered with θBr ≈ 0.4–0.35 ML. At poten- in the RRDE measurements with single crystals. There is
tials more negative than E = 0.1 V the surface coverage only one out of the many particle size effect studies that
with Br decreases substantially, e.g, from θBr ≈ 0.35 ML have employed exactly the same conditions as those used
to θBr ≈ 0.25 ML, and the O2 reduction currents increase here for single crystals: the study by Peukart et al.,[23] who
to a diffusion limiting current for 3e− per O2 molecule. reported results in dilute sulfuric acid at ambient tempera-
In the same potential region, the O2 reduction currents are ture. Fortunately, this was a very carefully done study using
accompanied quantitatively by the H2 O2 oxidation currents well-characterized supported Pt catalysts with (average)
on the ring electrode. It is clear that Brad simultaneously crystallite sizes ranging from 11.7 nm down to 0.78 nm.
suppress the initial adsorption of the O2 molecule and the As shown by Kinoshita,[6] the crystallite size dependence
formation of pairs of platinum sites needed for the breaking in the data of Peukart et al., are essentially perfectly fit
of the O–O bond. As in the case of Pt(111) in 0.1 M HClO4 , by a model which assumes: (i) that only the (100) face is
however, below ca. −0.05 V the diffusion limiting currents active for the ORR; and (ii) that the Pt particles are perfect
fall to a value for the 2e− reduction, reflecting that in this cubo-octohedra. Below about 6 nm, the fraction of the sur-
potential region there is an additional inhibiting effect of face having (100) facets drops rapidly, and actually goes to
Hupd on the kinetics of the ORR. zero below 1.8 nm; particles below 2 nm are formed almost
SXS analysis indicated that Brad does not form an ordered entirely from (111) facets and their intersections (edges and
structure on Pt(100) at any potential. This is probably corners).[6] Our ORR data from single crystal Pt in sulfu-
due to the fact that Pt(100) is intrinsically an atomi- ric acid fully support this model, since we have observed
cally rougher surface than (111) due to lifting of the that the (100) face is more active by two orders of mag-
‘hex’ reconstruction[22] when the flame annealed or ultra nitude versus the (111) face (in the model calculations, a
Oxygen reduction reaction on smooth single crystal electrodes 471

factor of 100 between the two faces is sufficient to fit the would little or no crystallite size effect for the ORR in a
data[6] ). By inference then, the crystallite size effect for nonadsorbing acid, such as, e.g., trifluoromethane sulfonic
the ORR on supported Pt in sulfuric acid at ambient tem- acid or Nafion .
perature would be due to the structure sensitive adsorption The effects of Cl− on the ORR with a Pt/carbon fuel
of (bi)sulfate anions, this adsorption effectively eliminating cell catalyst have been reported in a recent work from our
the contribution of the (111) facets to the overall rate. laboratory.[27] Figure 5(a) shows the current versus potential
Kinoshita demonstrated that the ORR data for supported curve for the ORR on Pt/Vulcan in 0.5 M HClO4 with and
catalysts in hot, concentrated H3 PO4 (180 ◦ C, 97–98% without chloride. In accordance with the ORR on Pt(hkl),
acid) reported in three different studies were also fit by the onset for the ORR is shifted to higher overpotentials
this model.[6] Since the physical basis for the crystallite with increasing chloride concentration in the solution. Com-
size effect in sulfuric acid is anion adsorption, it would pared to the ORR in the absence of Cl− the activity at
be a reasonable to suggest that the same physical basis constant potential decreases roughly one order of magni-
applies to this crystallite size effect in H3 PO4 , i.e., struc- tude for each order of magnitude increase in chloride anion
ture sensitive anion (phosphate) adsorption. There are other concentration. In addition, considerable H2 O2 currents on
indications that this is the case. Anion adsorption in dilute the ring electrode (Figure 5b) are observed in solutions con-
phosphoric has very similar structure sensitivity as sulfate taining Cl− , e.g., with increasing Cl− concentration, the
adsorption.[24] Sattler and Ross[25] reported a similar crys- fraction of peroxide increases up to 10% at cCl− = 10−3 M
tallite size dependence of the ORR on supported Pt in dilute and 14% at cCl− = 10−2 M around 0.6 V, respectively. The
phosphoric acid at ambient temperature as that found in hot, effect of Cl− on the ORR on supported catalysts can be fully
concentrated acid with the same catalysts. But it is unclear explained by applying the single crystal results to the rela-
whether similar adsorption chemistry would exist in the tionship between particle size and the different surface sites
extreme conditions of hot, concentrated phosphoric acid. (crystal faces, edges and corners).[6] Based on a high reso-
With differences in activity between crystal faces of only lution transmission electron microscopy (HRTEM) analysis
a factor of five or less for oxygen reduction in perchloric of the catalyst used in this study, the structure of typical
acid,[26] the particle models of Kinoshita indicate that there Pt particle (mean particle size ca. 3.7 ± 1 nm) illustrated in

(111)

(100)

20

0.90
16
0.85
E (V vs. RHE)

xH2O2 (%)

0.80 12

0.75
8
0.70
4
0.65

0.60 0
0.1 1 10 100 0.0 0.2 0.4 0.6 0.8
(a) ik (mA cm−2) (b) E (V vs. RHE)

Figure 5. (a) Tafel plots for the ORR on a Pt/carbon black catalyst in 0.5 M HClO4 (open circles) and after addition of 10−4 M Cl−
(black diamonds), 10−3 M Cl− (gray triangles) and 10−2 M Cl− . (b) Current efficiency for peroxide formation during ORR in the
presence of Cl− (concentrations as in a). Top shows transmission electron micrograph of a typical particle in the catalyst and a model
of the microstructure. (From Ref. [27].)
472 Part 5: The oxygen reduction/evolution reaction

Figure 5 show clearly that the cubo-octahedral Pt nanopar- membrane electrode assembly (MEA) during preparation
ticles close to the edge of the carbon support is facetted, or due to contamination of the humidified fuel cell feed
exposing (111) and (100) crystal facets and edges and cor- streams, e.g., salt in the air feed, will strongly affect the
ner sites with (110) symmetry. As on the Pt(111) crystal, at MEA performance. Based on the data in Figure 5, chlo-
the (111) facets, Clad blocks O2 adsorption without enhanc- ride contamination on the order of 4 ppm would result in
ing the formation of H2 O2 . In addition to inhibiting the a fuel cell voltage loss of ca. 50 mV. These results indi-
adsorption of O2 , the breaking of the O–O bond at the (100) cate in general the high purity conditions required for MEA
facets and (110) corner/edge sites and not as facile and the preparation and the humidified feed streams for PEMFCs.
production of peroxide on a supported Pt catalysts, ca. 15%, In addition to the kinetic effect of Cl− adsorbed on the
is close to the average value of H2 O2 produced on Pt(100) supported metal catalyst particles, its enhancement of the
(ca. 25%) and Pt(110) (ca. 2.5%) single crystal surfaces. H2 O2 production will be deleterious to the stability of per-
As shown in the aforementioned sections, even trace fluorinated membranes and ionomers in the catalyst layers
amounts of chloride (i.e., 10−4 M Cl− or ca. 4 ppm) dras- (e.g., Nafion membranes and ionomers), which are known
to degrade in the presence of peroxide radicals.
tically change both the activity and the reaction path-
way of the ORR on Pt catalysts. Although these results
were obtained in liquid electrolyte, similar effects can 5 ORR ON Pt(hkl) IN ALKALINE
be expected on electrodes in a solid-polymer electrolyte ELECTROLYTE
as present in a polymer electrolyte membrane fuel cell
(PEMFC). In this case, chloride impurities on the level The polarization curves for the ORR on Pt(111) at
of a couple of ppm, either due to incorporation into the 293 K and 333 K (1600 rpm, 50 mV s−1 ) are illustrated in

80

40
40

i (µA cm−2)
30 293 K
Iring (µA)

0
20

10 −40
333 K
(c) 0 Pt(111)
0 −80
0.0 0.2 0.4 0.6 0.8 1.0
−1 (a) E (V vs. RHE)
Pt(111), 0.1 M KOH
1600 rpm, 50 m V s−1
1.0
−2
i (mA cm−2)

−3
E (V vs. RHE)

−4 293 K 0.9

333 K
−5
313 K
333 K
293 K
−6
0.0 0.2 0.4 0.6 0.8 1.0
E (V vs. RHE) 0.8
0.1 1 10 100
(b) (d) ik (mA cm−2)

Figure 6. (a) Base voltammetry (50 mV s−1 ) of Pt(111) at 276 K (solid line) and 333 K (dashed line) in 0.1 M KOH. (b) ORR polarization
curves (50 mV s−1 , 1600 rpm) on Pt(111) at 293 K (dashed line) and 333 K (solid line). (c) Ring currents for the detection of peroxide
formed during the ORR on Pt(111) at 293 K (dashed line) and 333 K (solid line). (d) Tafel plots for the ORR on Pt(111) at 293 K (white
triangles), 313 K (gray squares) and 333 K (gray circles) deduced from the polarization curves at 1600 rpm.
Oxygen reduction reaction on smooth single crystal electrodes 473

Figure 6b. In short, the region of mixed kinetic-diffusion ca. 42 and 37 kJ mol−1 , respectively were similar to that on
control (0.8 V < E < 1 V) is followed by the diffusion Pt(111). Based on the polarization curves and the Tafel plots
limited current densities below ca. 0.8 V. At E < 0.4 V, at the different temperatures, the ORR activity increases
i.e., in the Hupd region, a deviation of the diffusion in the order P(110)(1 × 2) < Pt(100) < Pt(111). However,
limited currents to lower values can be observed. This in a previous study in KOH at room temperature, also
deviation can be quantitatively related to the formation of from this laboratory,[11] a slightly different order was found,
HO2 − as demonstrated by the HO2 − detection on the ring namely Pt(111) > Pt(110) > Pt(100). The discrepancy can
electrode,[4, 11] see Figure 6c. A more detailed discussion be explained by the differences in preparation and in surface
of the ORR in the Hupd region on Pt(hkl) can be found structure of the Pt(110) crystal. In the present study, we
in our previous papers.[4, 11] At lower overpotentials, i.e., carefully prepared the reconstructed Pt(110)(1 × 2) surface,
in the mixed kinetic-diffusion region, no peroxide can be versus the unreconstructed (1 × 1) surface used in the
observed. A comparison of the ORR polarization curves previous study.[11] Therefore, the lower activity in the
at 293 and 333 K indicates that as expected considerably present case can be ascribed to the fact that the very ‘open’
higher reaction rates are observed at 333 K, pointing Pt(110)(1 × 2) surface adsorbs OH more strongly than
to the fact that the ORR in KOH is a significantly (1 × 1) surface in the previous study, i.e., the adsorption
activated process. The apparent activation energy, H = , sites for molecular oxygen are significantly blocked by
is 47 kJ mol−1 determined at constant overpotential, η = adsorbed OH (see discussion below).
0.35 V in an Arrhenius analysis, (values summarized in
Table 1) according to a procedure described in Ref. [28].
A similar value was previously found in 0.05 M H2 SO4 6 ORR ON Au(hkl) IN ALKALINE
extrapolated to the reversible potential,[10] as discussed ELECTROLYTE
in the previous Section. The log i/E-relationships (so-
called Tafel plots) at 293, 313, and 333 K are plotted The study of Au(hkl) surfaces is complicated by the fact
in Figure 6d. The log i/E-relationships at the different that the flame annealed and/or the UHV annealed surfaces
temperatures were fitted by straight lines in order to of all three low index crystals of Au are reconstructed and
determine the Tafel slopes. Decreasing Tafel slopes from do not have the expected (1 × 1) bulk truncation structures.
ca. 86 to ca. 57 mV dec−1 were found when going from Fortunately, the stability of these surface reconstructions
293 to 333 K, as summarized in Table 1. In contrast to has now been well-studied by our group in alkaline
Pt(hkl) in sulfuric acid solution,[10] the Tafel slopes are electrolyte using in-situ SXS.[29] Figure 8 highlights the
not proportional to RT /F . Figure 7 illustrates the Tafel first sweep after emersion at ca. 0.1 V of the flame annealed
plots and the ring currents for HO2 − formed during the Au(100) crystal at 293 and 333 K, respectively. Directly
ORR on Pt(100) and Pt(110), respectively. In contrast to after flame annealing, the Au(100) surface is hexagonally
Pt(111), some HO2 − is detected in the potential region reconstructed forming a (5 × 27) unit cell.[30] Coming from
above 0.4 V, and there is generally a decrease of peroxide the negative emersion potential, upon sweeping through
formation by increasing temperature, an increase in activity the OHad region, this so-called (hex) reconstruction is
by increasing temperatures, and again decreasing Tafel lifted due to adsorption of hydroxyl species, i.e., positive
slopes when going from 293 to 333 K. It is noteworthy, of the peak at ca. 1.025 V at 293 K the Au surface has
that on Pt(100) and Pt(110) apparent activation energies of the unreconstructed (1 × 1) surface. At 333 K, due to the

Table 1. Summary of the Tafel slopes (Pt(hkl) and Au(hkl) and the apparent energies of activation (Pt(hkl) in 0.1 M
KOH.

293 K [mV dec−1 ] 313 K [mV dec−1 ] 333 K [mV dec−1 ] H = [kJ mol−1 ]b
lcd/hcda lcd/hcda lcd/hcda

Pt(111) –/86 –/72 –/57 47


Pt(100) –/112 –/78 –/60 42
Pt(110)(1 × 2) 92/190 80/120 –/83 37
Au(111) 60/185 – 72/150 –
Au(100) 47/119 – 50/85 –
Au(110) 44/127 – 37/86 –
a Tafel slopes determined at low (lcd) and high (hcd) current densities.
b Determined at E = 0.35 V.
474 Part 5: The oxygen reduction/evolution reaction

60 5

50 4
40
Iring (µA)

Iring (µA)
3
30
2
20
1
10

0 0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2
(b) E (V vs. RHE) (d) E (V vs. RHE)

333 K
313 K 0.9
293 K
0.9

E (V vs. RHE)
E (V vs. RHE)

0.8

0.8

0.7

0.7
0.1 1 10 100 0.1 1 10
(a) ik (mA cm−2) (c) ik (mA cm−2)

Figure 7. (a) Tafel plots for the ORR on Pt(100) at 293 K (white triangles), 313 K (gray squares) and 333 K (gray circles) deduced
from the polarization curves at 1600 rpm. (b) Ring currents for the detection of peroxide formed during the ORR on Pt(100) at 293 K
(white triangles), 313 K (gray squares) and 333 K (gray circles). (c) Tafel plots for the ORR on Pt(110)(1 × 2) at 293 K (white triangles),
313 K (gray squares) and 333 K (gray circles) deduced from the polarization curves at 1600 rpm. (d) Ring currents for the Figure 5:
magnification of the first voltammetric sweep after immersion at ca. 0.1 V of a flame-annealed Au(100) crystal in 0.1 M KOH solution
at 293 K (solid line) and 333 K (dashed line). 50 mV s−1 .

adsorption of OHad species at more negative potentials, the 80


lifting of the reconstruction is also shifted to more negative 333 K
potentials. The hex reconstruction reforms, albeit slowly,
60
upon holding the potential < ca. 0.4 V.[29] There is a similar Au(100)
lifting of the reconstructed surfaces of Au(111) and (110), 1st positive sweep
i (µA cm−2)

formed during flame annealing, when the potential is 40


scanned above ca. 1.1 V. For the results summarized here,
we used only the current potential curves recorded on 20
the cathodic sweep from 1.1 V, whereby all the surfaces
are known to have the (1 × 1) unreconstructed surface
0
structures throughout the potential region of interest.
Figure 9 summarizes the representative ORR polariza- 293 K
tion curves on Au(hkl) at 293 and 333 K (50 mV s−1 , −20
0.6 0.7 0.8 0.9 1.0 1.1
2500 rpm, lower panels in Figure 6) along with the cor-
responding ring currents for the oxidation of peroxide Figure 8. First voltammetry for a flame annealed Au(100) single
formed during the ORR (upper panels). The polarization crystal after immersion in 0.1 M KOH at 0.6 V. 50 mV s−1 , at 298
curves on Au(hkl) at 293 K are in agreement with previous and 333 K as shown.
results reported in the literature, whereas, to our knowl-
edge, it is the first time that ORR polarization curves on of the i –E curves on Au(hkl) are qualitatively similar
Au(hkl) at elevated temperature are reported. The shape to those described above for Pt(hkl). The region of mixed
Oxygen reduction reaction on smooth single crystal electrodes 475

120 Ring Ring Ring 120


I (µA)

I (µA)
80 80

40 40

0 0
0 0

−2 −2
i (mA cm−2)

i (mA cm−2)
293 K
293 K 293 K
−4 −4

333 K
333 K
−6 2500 rpm, 50 mV s−1 −6
333 K

0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2
(a) E (V vs. RHE) (b) E (V vs. RHE) (c) E (V vs. RHE)

Figure 9. ORR polarization curves (lower panel) along with the ring currents for peroxide detection (upper panel) at 293 K (dashed
lines) and 333 K (solid lines) on (a) Au(111), (b) Au(110) and (c) Au(100). 0.1 M KOH, 50 mV s−1 , 2500 rpm.

kinetic-diffusion control (ca. 0.7 < E < 1 V) is followed seen from the polarization curve at 293 K in Figure 9(c).
at lower potentials by a region where the reaction is Interestingly, at 333 K the reaction takes place via a 4e−
mainly under mass-transport control. However, in the process in a much wider potential range, and only below
region of diffusion limited current densities, no clear cur- ca. 0.45 V are significant amounts of peroxide formed. In
rent plateaus for a 4e− process are observed as was the order to get further insight into the reaction intermediate,
case on Pt(hkl). Deviations from the diffusion limited cur- e.g., solution phase HO2 − , we studied the reduction of
rent for the 4e− reduction to OH− can be quantitatively oxygen in a solution containing a known amount of solution
related to the formation of peroxide species, as indicated peroxide, as summarized in Figure 10. The temperature-
by the ring currents in the upper panels. As in the case dependent ORR curves without solution phase peroxide
of Pt(hkl), less peroxide is formed by increasing the tem- are shown in Figure 10(a), the corresponding ring currents
perature to 333 K. It is clear that the ORR on Au(hkl) for the oxidation of intermediate peroxide species are
is a strongly structural sensitive process in the temper- shown in Figure 10(b), and, finally the oxidation/reduction
ature range between 293 K and 333 K, with the kinetics in 0.1 M KOH + 2 × 10−3 M HO2 − shown in Figure 10c.
increasing in the order Au(111) < Au(110)  Au(100). The results in Figures 10(a, c) for Au(100) are similar to
The Tafel slopes of the ORR on Au(hkl) at the differ- those reported previously by Adzic and co-workers.[31] This
ent temperatures are summarized in Table 1. The values figure clearly demonstrates that the ORR closely follows
at room temperature are consistent with previously pub- the reduction of HO2 − . The activity of Au(100) for HO2 −
lished Tafel slopes[31] with only the value for Au(111) reduction peaks at ca. 0.2–0.8 V and above this potential
deviating slightly. This discrepancy may be attributed to only negligible amounts of peroxide are detected on the
some arbitrariness in drawing the tangent through the data ring electrode in ORR. Below ca. 0.6 V, the activity for
points that appear to be a straight line. Nevertheless, as on HO2 − decreases, and in the ORR there is a corresponding
Pt(hkl), we observe decreasing Tafel slopes with increasing increase in HO2 − production and reduction in the diffusion-
temperature. limiting current. This behavior clearly points to a serial
Although well known in the literature,[31–33] I want to ORR mechanism through the formation of solution phase
emphasize the exceptional behavior of Au(100) (Figure 9c) HO2 − (2 + 2e− reaction, k2 , k3 pathway in the reaction
towards the ORR. Coming from the positive potential limit, scheme). As one can also see from the polarization curves
oxygen reduction occurs in essentially a 4e− process to in Figure 10, the temperature dependence in the kinetically
form OH− . At potentials negative of the so-called catalytic controlled potential regions, i.e., E > 0.8 V, is relatively
peak, a transition to an almost 2e− process is observed, as slight, much less than on Pt(hkl). In this narrow temperature
476 Part 5: The oxygen reduction/evolution reaction

4 (O2 − )ad + H2 O −−−→ (HO2 − )ad + OHad (5)


3 − −
(HO2 )ad −−−→ HO2 (6)
2
(HO2 − )ad −−−→ Oad + OHad
i (mA cm−2)

1
(7a)
− −
0 Oad + H2 O + 1e −−−→ OHad + 2OH (7b)
293 K
−1 − −
313 K OHad + 1e −−−→ OH (×2) (8)
−2 333 K
−3
For Pt surfaces, steps (4a, b) are generally considered
(c) −4 to be rate-determining, and are often written as a single
150 reaction in which (O2 )ad does not appear explicitly. But
such a reaction is not an elementary step, and writing
0.1 M KOH
100 2500 rpm such a reaction simply means that the elementary steps
Ir (µA)

50 m V s−1 (4a, b) proceed at the same rate. Likewise (7a, b) are


50 often combined to yield a single 1e− step in which Oad
does not appear explicitly, but again this would not be
(b) 0 an elementary step. With (4a) and (4b) rate-determining,
0 steps (5–8) are at equilibrium. Careful comparisons of
voltammetry on Pt surfaces with and without O2 in solution
have shown that at all potentials the coverage by OHad
−2
i (mA cm−2)

is unchanged by the presence of O2 , i.e., it is determined


entirely by the equilibrium constant of step (8). This implies
−4 that the equilibria for steps (7a, b) and (8) are shifted
strongly to the right hand side, and that the coverage by
Au(100)
−6 (HO2 − )ad in particular is low, as is the amount of HO2 −
in solution. The mechanism is very useful in establishing
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
the role of OHad in the structure sensitivity of the reaction
(a) E (V vs. RHE)
on Pt(hkl). Stronger adsorption of OHad causes the rate
Figure 10. (a) ORR polarization curves on Au(100), (b) ring cur- of (4a, b) to ‘back up’ due to the accumulation of
rents for the detection of peroxide and (c) oxidation and reduction OHad on the surface, blocking sites for electron transfer
of ca. 2 × 10−3 M HO2 − on Au(100) at 293 K (solid line), 313 K to O2 . The structure sensitivity then follows the relative
(dashed-dotted line) and 333 K (dashed line), respectively. 0.1 M strengths of adsorption of OHad and the OHad adsorption
KOH, 50 mV s−1 , 2500 rpm. isotherm, (110) > (100) > (111),[11, 34] with the reaction
rates then varying in reverse proportion, (111) > (110) >
range of 293–333 K, it was difficult to obtain an accurate (100). However, examination of the kinetic parameters for
measure of the activation energy, and thus no activation the ORR in Table 1 reveals that this analysis might be
energies are reported for Au(hkl). At potentials below 0.8 V, an oversimplification, and does not necessarily capture all
kinetics of HO2 − reduction are significantly increased, of the factors that contribute to the structure sensitivity.
nearly eliminating the catalytic peak at ca. 0.7 V and Although the apparent activation energies for all low-
producing 2 + 2e− reduction to much lower potentials. index Pt(hkl) surfaces are very similar, between 37 and
47 kJ mol−1 , the Tafel slopes are different on all three
surfaces. In line with the OHad site blocking effect discussed
7 MECHANISM OF THE ORR ON Pt AND
above, it should be noted that above a certain OHad
Au(hkl) IN ALKALINE ELECTROLYTE coverage not enough adsorption sites are present on the
surface, and, hence, a change in the rate-determining step
I will discuss the results on Pt(hkl) and Au(hkl) in terms of from reaction (4b), the first electron transfer, to the O2
the elementary steps derived from the generalized reaction adsorption step, reaction (4a), might be expected. That
scheme for oxygen reduction given in the previous section this may indeed be the case is seen in the results for
but now applied specifically to alkaline electrolyte: Pt(110) electrode, the least active surface, where the ORR
proceeds on a surface highly covered by oxygenated species
O2 −−−→ (O2 )ad (4a)
with a high Tafel slope, e.g., above 130 mV dec−1 (see
− −
(O2 )ad + 1e −−−→ (O2 )ad (4b) Table 1) which implies a chemical rate-determining step,
Oxygen reduction reaction on smooth single crystal electrodes 477

as in reaction (1).[35] However, a full explanation of the anomalous temperature dependence in the Tafel slope even
structure sensitivity of the ORR on Pt(hkl) in alkaline when α is not temperature dependent. Returning to the
electrolyte will await a more quantitative kinetic model than rate expression derived in the first Section, since the rate-
we can offer at this time. determining step (rds) is the same in alkaline as in acid,
Relative to the Pt surfaces, if we assume the same and neither the proton nor water appear in the rds, the rate
mechanism and rate-determining step for Au surfaces, the expression is exactly the same according to[2, 19, 35]
difference in activity could qualitatively be attributed to  
−αF E
the much weaker interaction of OHad and Oad on the Au i = nF KcO2 · (1 − OH ) · exp
RT
surface, causing an accumulation of HO2 − both on the  
surface and in solution. The base voltammetry shown in −γrOH OH
× exp (9)
Figures 6, 7 and 9 clearly shows that indeed there is a much RT
weaker adsorption energy for OHad and Oad on the Au(hkl)
There are no other anions besides OH− to consider, so
surfaces than for the Pt(hkl) surfaces, with the onset of
we need consider only the potential dependent coverage
OH formation shifted by more than +0.4 V for each crystal
of OHad . If one includes only the coverage dependence of
face. Au(100) is the exception in terms of ORR activity,
the pre-exponential term (i.e., rOH = 0), OH adsorption is
having ‘Pt-like’ activity in the potential region above 0.8 V,
assumed to follow a Langmuir isotherm; if rOH = 0, the
without the production of any peroxide in solution, i.e.,
adsorption proceeds according to a Frumkin isotherm with
4e− reduction. Taylor et al.[10] pointed out that the Pt-like
lateral interactions of the adsorbates. The Tafel slopes at
behavior is dependent on the electrode pre-history, and is constant temperature, say 293 K, expected from this model
observed only on the negative-going sweep from potentials should deviate from the 116 mV dec−1 towards smaller val-
above ca. 1.1 V. They postulated two possible explanations: ues depending the value of rOH (under both Langmuirian
(a) the (100) surface is un-reconstructed above 1.1 V and and Frumkin conditions). That means, in order to model
becomes reconstructed upon reducing the potential below decreasing Tafel slopes with increasing temperatures, the
ca. 0.6 V, this transition causing the catalytic peak in the (1–OH ) term and/or the Frumkin-term must overcom-
polarization curve around 0.7 V; (b) surface oxide produced pensate the Butler–Volmer term. Consequently, in order
at potentials above 1.1 V are actually catalytic for the ORR to fit the experimental Tafel slopes using the experimen-
and their reduction from the surface below ca. 0.6 V causes tal OHad adsorption isotherms (in the absence of O2 [17] ),
the catalytic peak. In a later study from this laboratory,[27] mathematically one has to include rOH values that increase
we found from in-situ SXS that there is no structural with the temperature. Recall that higher rOH values describe
transition on the Au(100) surface on the negative going stronger repulsive interactions of the adsorbates (=lower
sweep from 1.1 V in the potential region around 0.6–0.8 V, free energy of adsorption at given coverage), and increas-
i.e., the surface remains in the unreconstructed (1 × 1) ingly stronger interactions with increasing temperature (via
geometry. A catalytic effect of the (relatively) irreversible increased mobility) is consistent with chemical intuition.
oxide formed above 1.1 V is against chemical intuition, and Nonetheless, the effect of decreasing Tafel slopes with
furthermore at 333 K there is essentially no ‘oxide’ left increasing temperature still need further theoretical mod-
on the surface at 0.8–0.9 V (Figure 9b) but the region of eling in order to reach unambiguous conclusions.
4e− reduction is extended to even lower potentials. So an Finally, I conclude with a new discussion of the effect of
explanation for the unique behavior of the Au(100) for the pH on the kinetics of oxygen reduction. It has been known
ORR in alkaline solution remains elusive. for some time that there are certain electrode materials that
The temperature dependence of the Tafel slopes on are useful, even practical, catalysts for oxygen reduction
Pt(hkl) and Au(hkl) as summarized in Table 1 are worth in alkaline electrolyte, but not in acid electrolyte.[4] While
more detailed discussion. Most notably, independent of the in many cases the pH effect may be attributed simply
exact value of the Tafel slopes, one trend can clearly be to corrosion, e.g., transition metals are corroded in acid
observed: on both Pt(hkl) and Au(hkl) the Tafel slopes but not in alkaline solution, that is not case for the two
always decrease with increasing temperature. This result materials I will focus on here, carbon and Au. Both these
is in the contrast to Pt(hkl) in acid electrolyte, where materials are corrosion resistant in the 0–1 V potential
the Tafel slopes increase with the temperature, viz., from range in both strong acid and strong base solution even
116 mV dec−1 at 293 K to 132 mV dec−1 at 333 K as at elevated temperatures. The case of glassy carbon, as
expected.[4, 5] One explanation for decreasing Tafel slopes illuminated in the recent study by Yang and McCreery,[36] is
with increasing temperature is a temperature dependence particularly interesting in this regard. Figure 11 reproduces
of the apparent transfer coefficient, α, as discussed and two polarization curves from this study, the one at pH = 0
summarized by Adzic.[3] It is also possible to derive and the one at pH = 14. At each pH, different surface
478 Part 5: The oxygen reduction/evolution reaction

80 the standard equilibrium potential for (1c) to be −0.31 V


pH 14 versus SHE (pH = 0), which puts the half-wave potential
60
(E1/2 ) at −0.15 V versus SHE or −0.36 V versus Ag/AgCl.
As expected, the outer sphere reaction is pH independent,
Current (µA)

and E1/2 remains at the same potential versus the Ag/AgCl


40
reference E1/2 . What is unexpected is that E1/2 for the
inner sphere electron transfer also remains at the same
AC/IPA cleaned
20 potential, shifted positively from the outer sphere potential
Polished by ca. 0.15 V independent of pH. Yang and McCreery
0 suggest the inner sphere reaction is the same as above for
MP derivatized Pt and Au surfaces,
−20
0.0 −0.2 −0.4 −0.6
(b) Potential (V vs. Ag/AgCl) O2 −−−→ (O2 )ad (4a)
− −
70 (O2 )ad + 1e −−−→ (O2 )ad (4b)

60 pH 0
Note that these reaction steps are also independent of pH,
50 so that inner sphere transfer is also pH independent, as
Current (µA)

40 Polished
observed experimentally. The reactions following (4b) and
(4c), e.g., protonation to form HO2 − , are pH dependent,
30
but as long as (4b) or (4c) are rate-determining, the
20 AC/IPA cleaned E1/2 values will be pH independent. For glassy carbon,
the ca. 0.15 V shift in E1/2 is essentially just the so-
10
called Frumkin (double-layer) correction to the electron
MP modified
0 transfer rate between an inner sphere and an outer sphere
−10
reaction.[37] At, pH = 14, the standard potential for oxygen
0.2 0.0 −0.2 −0.4 −0.6 −0.8 reduction to peroxide is −0.144 V versus SHE or −0.35 V
(a) Potential (V vs. Ag/AgCl) versus Ag/AgCl, which means that both inner sphere
Figure 11. ORR polarization curves on a glassy carbon electrode
and outer sphere electron transfer of O2 to superoxide
in 1 M KOH (a) and 0.5 M H2 SO4 (b) with different surface mod- occurs close to the thermodynamic potential for 2e−
ifications. Curves recorded potentiodynamically at 200 mV s−1 . reduction of oxygen to peroxide, and, by definition at low
Note the potential scale uses a reference electrode that is pH overpotential (for 2e− ). This is shown more graphically
independent. (From Ref. [36] with permission.) in the modified Pourbaix diagram for hydrogen peroxide
in Figure 12. The conventional Pourbaix diagram[38] is
preparations were used to generate different fundamentally modified to show the equilibria for reaction (4c). Hence,
different behavior: preparations such as polishing to prepare at pH = 14 carbon surfaces reduce oxygen to peroxide
a ‘clean’ surface, and modification with a covalently bonded at low overpotentials, but at pH = 0 the overpotential
monolayer of methylphenyl groups. The methylphenyl is very large, increased by essentially 0 .059 × 14 or
groups are designed to act as a charge transfer layer that ca. 0.84 V because the thermodynamic potential has shifted
prevents the O2 molecule from interacting directly with the by 0.84 V positively but the E1/2 for the rate-determining
carbon surface, but allows outer sphere electron transfer to step (4b) has not. Because of the overlap in the equilibrium
occur to form the superoxide radical anion, potentials of peroxide and superoxide at pH = 14, nearly
any electronically conducting electrode material can be
O2 + 1e− −−−→ O2 − (4c) used to reduce oxygen to peroxide, no specific chemical
interaction of O2 (or O2 − ) is required. However, at pH = 0
The lifetime of solution phase O2 − is greater in the absence only surfaces with strong chemical interaction with O2 (or
of protons (pH = 14), and thus the oxidation of superoxide O2 − ) can reduce oxygen to peroxide at low overpotential,
on the reverse sweep is absent at low pH. Otherwise the and even stronger interaction to reduce oxygen to water
curves are virtually identical at both pH 0 and 14 for at low overpotential. These are fundamental facts of nature
both conditions of the surface. In agreement with other which lie at the heart of the difficulty in finding alternatives
studies in nonaqueous solvent, Yang and McCreery find to Pt as the catalyst for ORR in acidic electrolyte.
Oxygen reduction reaction on smooth single crystal electrodes 479

2.2
2.0 1′
1.8 2 H2O2 HO2−
1.6 0
−2
1.4 −4
1.2 b log(H2O2) = −6

1.0

0.8 −6 3 0
4 −4 −2
−) = −6−4
0.6 −2 log(HO2
(H2O2)
0.4 log =0
PO2 (HO2−)
E (V)

0.2
log =
PO2
−6
−6
0.0 5 −4
−2
−4
−2 0
−0.2
6 0
−0.4 (O2−)
log = a
PO2
−0.6
−0.8
−1.0
−1.2
−1.4
−1.6
−1.8
−2 −1 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
pH

Figure 12. Pourbaix diagram[38] for hydrogen peroxide with the equilibria for the superoxide/oxygen reaction (labeled 6 in the diagram)
added. Other notations for equilibria as in Ref. [38].

8 CONCLUSIONS In acid electrolyte, the ORR polarization curves on Au


surfaces are shifted negatively [on a reference hydrogen
The structure sensitivity of the oxygen reduction reaction on electrode (RHE) scale] by hundreds of mV, and the
Pt and Au surfaces has been examined using single crystal reaction becomes structure insensitive. The large effect of
electrodes and modern methods of surface preparation and pH on the overpotential for ORR on Au is due to the
characterization. For Pt in sulfuric and phosphoric acid absence of chemisorption of either O2 or O2 − . The rate-
electrolytes, the structure sensitivity is dominated by the determining step on Au is inner sphere electron transfer
inhibiting effects of anion adsorption, which is strongly to O2 form O2 − , and this reaction step is independent
structure sensitive. In nonadsorbing acids like perchlorate of pH, i.e., the reaction occurs at a fixed potential on a
or triflate, the structure sensitivity is weak, probably due SCE scale, so the overpotential (on an RHE scale) changes
to the single-site nature of the rate-determining step, i.e., with pH.
electron transfer to form adsorbed superoxide radical anion.
Thus, a particle size effect is not expected for supported
Pt catalyst in a PEMFC cathode if the geometric change
in particle shape with size is the only factor contributing, ACKNOWLEDGEMENTS
i.e., absent electronic effects such as quantum confinement.
For Au in alkaline electrolyte, oxygen reduction is strongly This work was supported by the US Department of Energy,
structure sensitive, with the (100) surface the only Au the Office of Science, Basic Energy Sciences, Materials Sci-
surface with direct 4e− oxidation at potentials near 0.9 V. ences Division, and the Office of Advanced Transportation
There is no explanation available for the unique catalytic Technologies, Fuel Cell Systems, under contract DE-AC03-
activity of Au(100) for the ORR in alkaline electrolyte. 76SF00098.
480 Part 5: The oxygen reduction/evolution reaction

REFERENCES B. N. Grgur and P. N. Ross, J. Electroanal. Chem., 467, 157


(1999).
19. T. J. Schmidt, U. A. Paulus, H. A. Gasteiger, N. Alonso-
1. A. Damjanovic, ‘Modern Aspects of Electrochemistry’, J. O. Vante and R. J. Behm, J. Electrochem. Soc., 147, 2620
M. Bockris and B. E. Conway (Eds), Plenum Press, New (2000).
York, pp. 369 (1969).
20. J. A. Dean (Ed), ‘Lange’s Handbook of Chemistry’, 11th
2. M. R. Tarasevich, A. Sadkowski and E. Yeager, ‘Compre- edition, McGraw-Hill, New York, pp. 5–5 (1973).
hensive Treatise in Electrochemistry’, J. O. M. Bockris, B. E.
Conway, E. Yeager, S. U. M. Khan and R. E. White (Eds), 21. (a) C. A. Lucas, N. M. Markovic and P. N. Ross, Phys.
Plenum Press, New York, p. 301 (1983). Rev. B, 55, 7964 (1997); (b) N. M. Markovic, C. A. Lucas,
H. A. Gasteiger and P. N. Ross, Surf. Sci., 372, 239 (1997).
3. R. R. Adzic, ‘Electrocatalysis’, J. Lipkowski and P. N. Ross
(Eds), Wiley-VCH, New York, pp. 197–242 (1998). 22. P. N. Ross, ‘Structure of Electrified Interfaces’, J. Lipkowski
and P. Ross (Eds), New York and Wienheim, pp. 35–65
4. N. M. Markovic and P. N. Ross Jr, ‘Interfacial Electrochem-
(1993).
istry–Theory, Experiments and Applications’, A. Wieckowski
(Ed), Marcel Dekker, New York, pp. 821–841 (1999). 23. M. Peukart, T. Yoneda, R. Dalla Betta and M. Boudart, J.
Electrochem. Soc., 133, 944 (1986).
5. A. J. Appleby, Catal. Rev., 4, 221 (1970).
24. P. Olivera, M. Patrito and H. Sellers, ‘Interfacial Electro-
6. K. Kinoshita, ‘Electrochemical Oxygen Technology’, John
chemistry–Theory, Experiments and Applications’, A. Wiec-
Wiley & Sons, New York (1992).
kowski (Ed), Marcel Dekker, New York, pp. 821–841
7. E. Yeager, M. Razaq, D. Gervasio, A. Razaq and D. Tryk, (1999).
‘Structural Effects in Electrocatalysis and Oxygen Electro-
chemistry’, D. Scherson, D. Tryk, M. Daroux and X. Xing 25. P. Sattler and P. N. Ross, Ultramicroscopy, 20, 21 (1986).
(Eds), The Electrochemical Society, Pennington NJ, PV92- 26. N. M. Markovic, H. A. Gasteiger and P. N. Ross, J. Elec-
11, p. 440 (1992). trochem. Soc., 144, 1591 (1997).
8. R. J. Celotta, R. A. Bennett, J. L. Hall, M. W. Siegel and 27. N. M. Markovic, T. J. Schmidt, V. Stamenkovic and P. N.
J. Levine, Phys. Rev. A, 6, 631 (1972). Ross, Fuel Cells, 1, 105 (2001).
9. V. Stamenkovic, N. M. Markovic and P. N. Ross Jr, J. Elec- 28. U. A. Paulus, T. J. Schmidt, H. A. Gasteiger and R. J. Behm,
troanal. Chem., 500, 44 (2000). J. Electroanal. Chem., 495, 134 (2001).
10. B. N. Grgur, N. M. Markovic and P. N. Ross Jr, Can. J. 29. I. M. Tidswell, N. M. Markovic, C. A. Lucas and P. N. Ross
Chem., 75, 1465 (1997). Jr, Phys. Rev. B, 47, 16 542 (1993).
11. N. M. Markovic, H. A. Gasteiger, B. N. Grgur and P. N. 30. D. M. Kolb, Prog. Surf. Sci., 51, 109 (1996).
Ross, J. Electroanal. Chem., 467, 157 (1999).
31. (a) R. R. Adzic and N. M. Markovic, J. Electroanal. Chem.,
12. M. R. Tarasevich, Electrochimiya, 9, 578 (1973). 138, 443 (1982); (b) R. R. Adzic, N. M. Markovic and
13. F. Will, J. Electrochem. Soc., 112, 451 (1965). V. B. Vesovic, J. Electroanal. Chem., 165, 105 (1984); (c)
N. M. Markovic, R. R. Adzic and V. B. Vesovic, J. Elec-
14. (a) B. Ocko, O. Magnusen, J. Wang and R. Adzic, ‘Nanoscale
troanal. Chem., 165, 121 (1984).
Probes of the Solid/Liquid Interface’, A. Gewirth and
H. Siegenthaler (Eds), Kluwer Academic Publishers, Nether- 32. E. J. Taylor, N. R. K. Vilambi and A. Gelb, J. Electrochem.
lands, pp. 103–119 (1995); (b) M. Toney, ‘Synchrotron Tech- Soc., 136, 1939 (1989).
niques in Interfacial Electrochemistry’, C. Melendres and 33. N. M. Markovic, I. M. Tidswell and P. N. Ross, Langmuir,
A. Tadjaddine (Eds), Kluwer Academic Publishers, Nether- 10, 1 (1994).
lands, pp. 109–125 (1994).
34. T. J. Schmidt, N. M. Markovic and P. N. Ross, J. Phys.
15. (a) A. Gewirth and B. Nicce, Chem. Rev., 97, 1129 (1997); Chem. B, 105, 12 082 (2001).
(b) K. Itaya, Prog. Surf. Sci., 58, 121 (1998).
35. D. B. Shepa, M. V. Vojnovic and A. Damjanovic, Electro-
16. N. M. Markovic, H. A. Gasteiger and P. N. Ross, J. Phys.
chim. Acta, 26, 781 (1981) and references therein.
Chem., 99, 3411 (1995).
36. H.-H. Yang and R. L. McCreery, J. Electrochem. Soc., 147,
17. (a) J. O. Bockris, M. G. Gamboa-Aldeco and M. Sklarczyk,
3420 (2000).
J. Electroanal. Chem., 339, 355 (1992); P. Faguy, N. Mark-
ovic, R. Adzic and E. Yeager, J. Electroanal. Chem., 289, 245 37. W. R. Fawcett, ‘Electrocatalysis’, J. Lipkowski and P. N.
(1990). Ross (Eds), Wiley-VCH, New York, pp. 323–372 (1998).
18. (a) H. A. Gasteiger, N. M. Markovic and P. N. Ross, Lang- 38. M. Pourbaix, ‘Atlas of Electrochemical Equilibria in Aqueous
muir, 12, 1414 (1996); (b) N. M. Markovic, H. A. Gasteiger, Solutions’, Pergamon Press, London, pp. 108 (1966).

View publication stats

You might also like