Вы находитесь на странице: 1из 9

Catalysis Today 202 (2013) 105–113

Contents lists available at SciVerse ScienceDirect

Catalysis Today
journal homepage: www.elsevier.com/locate/cattod

Water dissociation on well-defined platinum surfaces:


The electrochemical perspective
Maria J.T.C. van der Niet a , Nuria Garcia-Araez a,b , Javier Hernández c , Juan M. Feliu c , Marc T.M. Koper a,∗
a
Leiden Institute of Chemistry, Leiden University, Einsteinweg 55, P.O. Box 9502, 2300 RA Leiden, The Netherlands
b
FOM Institute for Atomic and Molecular Physics (AMOLF),P.O. Box 41883, 1009 DB Amsterdam, The Netherlands
c
Departamento de Química Fisica, Universidad de Alicante, Ap. 99. E-03080 Alicante, Spain

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

Article history: This paper discusses three important discrepancies in the current interpretation of the role of water
Received 11 January 2012 dissociation on the blank cyclic voltammetry of well-defined single-crystalline stepped platinum sur-
Received in revised form 28 March 2012 faces. First, for H adsorption both H-terrace and H-step contributions have been identified, whereas for
Accepted 23 April 2012
OH adsorption only OH-terrace has been identified. Second, different shapes (broad vs. sharp) of the H-
Available online 20 June 2012
terrace and H-step voltammetric peaks imply different lateral interactions between hydrogen adatoms
at terraces and steps, i.e. repulsive vs. attractive interactions. Third, the H-step peak exhibits an unusual
Keywords:
pH-dependent shift of 50 mVNHE /pH unit. We propose here a model that can explain all these obser-
Electrocatalysis
Water dissociation
vations. In the model, the H-step peak is not due to only ad- and desorption of hydrogen, but to the
Platinum replacement of H with O and/or OH. The O:OH ratio in the step varies with step geometry, step den-
sity and medium. In alkaline media relatively more OH is adsorbed in (or on) the step than in acidic
media, under which conditions more O is adsorbed in (or on) the step. This would explain the anomalous
pH dependence and would provide a possible explanation for the higher catalytic activity of alkaline
media for electro-oxidation reactions. Although the model certainly still contains speculative elements,
we believe it provides the most consistent interpretation of platinum single-crystal electrochemistry cur-
rently available, and presents an important and significant improvement over previous interpretations.
In situ spectroscopic data are ultimately needed to confirm or disprove some of the assumptions of the
model.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction border potentials depending on surface structure, electrolyte


cation and anion, voltammetric scan rate, and scan direction, is
The interface between a platinum electrode and an aqueous well accepted and discussed in many textbooks [2]. Much of our
electrolyte solution is arguably the most frequently studied current understanding of the blank cyclic voltammetry of platinum
electrode–electrolyte combination in modern physical elec- comes from the many detailed experiments that have been carried
trochemistry. The reactivity and dissociation of water at out with single-crystalline platinum electrodes since the seminal
the platinum–aqueous electrolyte interface, in particular the work of Clavilier [3–8]. These measurements have formed the
effect of platinum surface structure, is of immense impor- basis for linking the platinum surface structure and surface species
tance to electrocatalysis [1]. The electrochemical signature of formed to the electrochemical response.
the platinum–aqueous electrolyte interface is its blank cyclic The first modeling attempts to compute current–voltage rela-
voltammetry (CV), i.e. the current–potential curve obtained under tionships based on a statistical–mechanical description of the
potential cycling in presence of supporting electrolyte only. Its platinum–solution interface date back to Armand and Rosinberg
interpretation in terms of hydrogen adsorption or hydrogen [9]. This work was later taken further by others, and has led, for
underpotential deposition (HUPD ) (at potentials below ∼0.35 V instance, to the conclusion that the adsorption of hydrogen at
vs. RHE, reversible hydrogen electrode), double-layer region (at the Pt(1 1 1)/electrolyte interface is well approximated by a mean-
potentials between 0.35 and 0.6 VRHE ), and OH adsorption and/or field (or “Frumkin”) isotherm, implying relatively weak repulsive
oxide formation (at potentials above 0.6 VRHE ), with some of the interactions between the adsorbed UPD hydrogen [10–12]. More-
over, the hydrogen adsorption energy and the interaction energy
between adsorbed hydrogens are very similar to those at the
Pt(1 1 1)/ultra high vacuum (UHV) interface, as obtained either by
∗ Corresponding author.
experiment or by first-principles density functional theory (DFT)
E-mail address: m.koper@chem.leidenuniv.nl (M.T.M. Koper).

0920-5861/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cattod.2012.04.059
106 M.J.T.C. van der Niet et al. / Catalysis Today 202 (2013) 105–113

or
H2 O + e− + ∗ter  Had,ter + OH− (2)
where “ter” stands for a (1 1 1) terrace site. In elegant thermody-
namic work, Jerkiewicz has shown that from the H-ter feature one
can extract the H adsorption energy on Pt(1 1 1), and that this value
agrees well with the value for H adsorption at the Pt(1 1 1)-vacuum
interface [10]. Moreover, the shape of the H-ter feature is mod-
eled well by a Frumkin-type isotherm implying weak repulsive
interactions between the adsorbing hydrogens [10,16].
The feature termed OH-ter is less well understood, but is typi-
cally associated with the adsorption of OH from water dissociation
[17,18]:
H2 O + ∗ter  OHad,ter + H+ + e− (3)
or
OH− + ∗ter  OHad,ter + e− (4)
From DFT calculations of OH co-adsorbed with water on Pt(1 1 1)
and Rh(1 1 1), [19] it is suggested that OHad,ter would bind to a sin-
gle surface atom with the O H bond parallel to the metal surface in
order to accommodate for hydrogen bond formation from a neigh-
boring water molecule. Such a parallel adsorption mode would
make this species invisible for IR spectroscopy. The observed pH
dependence clearly agrees with reaction (3), and also the observed
Fig. 1. Cyclic voltammetry profiles of (a) Pt(1 1 1), (b) Pt(5 5 3), and (c) Pt(5 3 3) in
0.1 M HClO4 solution and 0.05 M NaOH taken at 50 mV s−1 . reversibility of the OH-ter feature strongly suggests that no further
oxidation than to OH has taken place. From the charge associated
with this feature in perchloric acid solution (1 1 0 ␮C cm−2 ), one
would estimate an OHad,ter surface coverage of ∼0.46. In alkaline
calculations [13]. This result implies that apparently water has little solution, the corresponding charge and coverage are 160 ␮C cm−2
influence on the adsorption properties of hydrogen on Pt(1 1 1). and ∼0.66. Although the OH-ter features in acidic and alkaline
The issue under specific consideration in this article is illustrated solution occur in exactly the same potential region, the shapes of
in Fig. 1, which displays the voltammetry of Pt(1 1 1) (a) and the the features are different. In acidic solution, the feature is always
stepped Pt(5 5 3) (b) and Pt(5 3 3) (c) surfaces in both acidic (per- associated with a sharp spike at 0.8 VRHE . No such sharp feature
chloric) and alkaline solution. Crystalographically, the Pt(n,n,n − 2) is observed in alkaline solution. The origin of the sharp feature
surface is a surface consisting of n-atom wide terraces of (1 1 1) was discussed by Berna et al. [20], who suggested that the broad
orientation and steps of (1 1 1) orientation, or n =n − 1 atom wide and sharp peaks are associated with two different kinds of sur-
terraces of (1 1 1) orientation and steps of (1 1 0) orientation. Elec- face water, namely ice-like water and isolated water molecules,
trochemically, the steps behave like (1 1 0)-type sites as they show the exact molecular origin of which has so far remained elusive.
the same adsorption/desorption process in the same potential Alternatively, it has been suggested that the sharp peak is associ-
range as Pt(1 1 0). Therefore, we shall consider the surface as having ated with a disorder-order phase transition in the OHad,ter adlayer,
(1 1 0) steps in the rest of this work. The Pt(n + 1, n − 1, n − 1) surface as these sharp voltammetric features are typically related to phase
consists of n atom wide terraces with (1 0 0) steps. Perchloric acid transitions in the adlayer [11,16]. The phase transition would then
is chosen as the acidic electrolyte as the influence of co-adsorbing be absent in alkaline media.
anions should be absent or minimal in this electrolyte. The voltam- When comparing Fig. 1a to Fig. 1b, we observe that the intro-
metry of Pt(1 1 1) in perchloric acid and sodium hydroxide (Fig. 1a) duction of steps of (1 1 0) orientation into the (1 1 1) surface leads
consists of two main peaks: a hydrogen adsorption–desorption fea- to the appearance of only one additional voltammetric feature in
ture (to be referred to as “H-ter” from now on) at 0.05–0.35 VRHE , both acidic and alkaline solution (steps of (1 0 0) orientation have a
and a hydroxyl (OH) adsorption–desorption feature (to be referred very similar effect, but at more positive potential, see Fig. 1c). Inter-
to as “OH-ter” from now on) between 0.6 and 0.85 VRHE . Note that estingly, the feature occurs at 0.125 VRHE in 0.10 M HClO4 solution,
the fact that both features occur at the same potential on the RHE but at a higher potential (∼0.26 VRHE ) in 0.05 M NaOH solution. This
scale implies a 60 mV/pH unit shift on the NHE scale, i.e. a surface shift in peak potential with pH on the RHE scale is “anomalous”
reaction involving 1 proton being exchanged with the solution per since there is no such peak shift for the H-ter feature (vide supra).
transferred electron. This is the usual and expected pH dependence Traditionally, the step-related feature is ascribed to the adsorption-
for a coupled proton–electron transfer reaction. The adsorbed desorption of hydrogen at the step, and we will therefore refer to
hydrogen formed in the H-ter feature has appeared very difficult it as “H-step”:
to be observed by Fourier transform infrared (FTIR) spectroscopy H3 O+ + e− + ∗step  Had,step + H2 O (5)
although the hydrogen formed at more negative potentials, which
is involved in the hydrogen evolution reaction, has been obserbed but we emphasize that this reaction can NOT explain the observed
by FTIR to bind atop to the Pt surface atoms [14,15]. Osawa et al. “anomalous” pH dependence, as it would predict the potential of
[15] have suggested H-ter (or HUPD ) to be adsorbed in hollow sites. the H-step feature NOT to shift with pH on the RHE scale, whereas
In any case, the observed pH dependence suggests a reaction of the in reality the peak potential shifts significantly with pH (see Fig. 1b
type: and c). Moreover, as we explained in previous papers [21,22], the
sharpness of this feature implies attractive interactions between
adsorbed hydrogens, if indeed these are the only species adsorb-
H3 O+ + e− + ∗ter  Had,ter + H2 O (1) ing on the step. Attractive interactions seem somewhat unlikely,
M.J.T.C. van der Niet et al. / Catalysis Today 202 (2013) 105–113 107

especially if these same interactions were concluded to be repul- atmosphere and transported to the electrochemical cell under a
sive on the terrace. Therefore, the anomalous pH dependence and protective droplet of de-aerated water. The electrode was brought
the peak sharpness strongly suggest that reaction (5) does not rep- in contact with the solution under electrochemical control. We
resent accurately what is really happening at the step. note that the flame-annealing procedure is well established in
Another anomaly related to the H-step feature that seems to the electrochemistry of single crystals and has been shown in
have been overlooked by many in the literature, is that it does the pioneering work of Clavilier [3–8] to yield highly-ordered
not have an “OH-step” counterpart, i.e. there is no feature in the surfaces comparable in quality to those prepared in ultra high vac-
blank voltammograms of Fig. 1b and c that is associated with OH uum. We also note that from the extensive knowledge available
adsorption in step sites. If there are clear features related to H and about the the reconstruction of single-crystal platinum surfaces
OH adsorption on the terraces, and to H adsorption on the steps, in an electrochemical environment, to be cited below, there is
then why is there no feature ascribable to OH adsorption on the no reason to believe that stepped platinum surfaces with suffi-
steps? This issue, which seems to have been overlooked in a signif- ciently wide terraces reconstruct. We also note that there is no
icant part of the literature, is especially pressing as the OH adsorbed evidence for the reconstruction of steps in Pt(5 5 3) and Pt(5 3 3)
in step and defect sites is believed by many (including ourselves) surfaces (i.e., the surfaces used in Fig. 1b and c) in UHV (see ref.
to be the active oxygen-donating species in many technologically [25] and references therein). It is known that extensive potential
important electrocatalytic oxidation reactions [23]. cycling may lead to (small) changes in voltammetry and presum-
In this paper we propose a model that is capable of reconcil- ably surface structure, but this extensive cycling was avoided in
ing many of the observed anomalies, including peak sharpness, the experiments to be described below. Therefore, for the rest of
anomalous pH dependence, and the absence of the OH-step fea- the paper, we will assume that the surfaces have their nominal
ture. Admittedly, the model is partially based on assumptions structure.
that need further testing and confirmation, primarily by detailed All electrochemical measurements were performed with a
spectroscopic and theoretical investigations. However, the main potentiostat (Autolab, PGSTAT 30 or PGSTAT 12) controlled by a
advance offered by the model in this paper is that it represents computer. Mathematical treatment of the data has been done with
the first attempt at a fully consistent interpretation of the single- the Wavemetrics IGOR PRO© 6.12A software package. Since the
crystal voltammetry of stepped single-crystal platinum electrodes, electrodes have been cut at an angle ˛ from the (1 1 1) plane, the
which have served as model catalysts in many electrocatalysis stud- geometric area (which can be calculated from the measured diam-
ies. Such a detailed interpretation and discussion does not exist eter) differs from the area available for adsorption, as defined by
in the current literature, but it is crucial in any fully consistent Clavilier’s hard-sphere model [26,27], through
interpretation of the catalytic properties of these electrodes. We
Ageom
also note that UHV modeling experiments of the co-adsorption Aact = (6)
cos˛
of oxygen and water on stepped platinum surfaces, as carried out
recently in Leiden, support many of the assumptions made in the The resulting transferred charges have been corrected using this
model [24]. An additional attraction of the model is that, from a equation, but the reported current densities have not.
mathematical point-of-view, it is analytically solvable (albeit in a
mean-field approximation for the adsorption on the terraces) with 3. Results
a clear physical meaning of all the physical parameters entering the
model. 3.1. Pt(S)[(n − 1)(1 1 1) × (1 1 0)]

First we return to the HUPD region of Fig. 1b, already discussed


2. Experimental briefly in the Introduction, which shows the CVs of Pt(5 5 3) in
acidic and alkaline solution. In acidic media (0.10 M HClO4 , dashed
Electrochemical experiments were conducted at room temper- line) a step induced feature appears at 0.125 VRHE . This feature is
ature in a two compartment electrochemical cell using a large area completely reversible. In alkaline media (0.05 M NaOH, solid line)
platinum counter electrode and a reversible hydrogen electrode the feature has an identical shape, but has shifted to 0.26 VRHE on
(RHE) as a reference electrode, separated from the main com- the positive-going sweep. Moreover, the feature is less reversible
partment through a Haber-Luggin capillary. After initial boiling than in acidic media, with the negative-going sweep peak at
in sulphuric acid, the cell was boiled several times in water from 0.25 VRHE . This lack of reversibility in alkaline media is in agreement
an Elga Purelab Ultra system or a Millipore Milli-Q gradient A10 with impedance spectroscopy measurements [22], which show a
system (both 18.2 M cm resistance) before each experiment. non-zero charge-transfer resistance in the Hupd region in alkaline
0.05 M NaOH-solutions were prepared using NaOH pellets solution.
(Sigma–Aldrich, 99.998 %). HClO4 (Suprapur, 70 %) and KClO4 (pro Measurements comparing different step densities of steps with
analysis) were obtained from Merck. The phosphate buffers with the (1 1 0) geometry in acidic media (0.5 M H2 SO4 and 0.1 M HClO4 )
various pH’s were prepared with a constant phosphate concentra- showed that 1 electron is transferred per step platinum atom.
tion of 0.10 M using a combination of ortho-phosporic acid (Merck, Also on terrace sites a transfer of 1 electron per platinum atom
Suprapur, 85 %), anhydrous NaH2 PO4 (Merck, Suprapur, 99.99 %), is assumed, using a hydrogen coverage of 0.67 ML at the lower
anhydrous Na2 HPO4 (Merck, Suprapur, 99.99 %), and NaOH pel- potential limit that coincides with the beginning of hydrogen evolu-
lets. After each measurement the exact pH was determined with a tion, resulting in a reasonable double layer charging of 51 ␮C cm−2
pH-meter (Radiometer analytical, PHM 220) employing a GK2401C [26,27].
electrode head. All solutions where de-aerated by bubbling argon Fig. 2a shows the oxidative sweeps of a series of
(BIP plus, 6.6, Air Products) through them. A blanket of argon was Pt(S)[(n − 1)(1 1 1) × (1 1 0)] electrodes with n = 14, 6, 5, 4, and
kept over the solution throughout the measurement. 3 [Pt(7 7 6), Pt(3 3 2), Pt(5 5 3), Pt(2 2 1), and Pt(3 3 1)] in 0.05 M
All electrodes were prepared from single crystal platinum beads NaOH. We observe the step induced feature at 0.26 VRHE for all
oriented, cut and polished down to 0.25 ␮m with diamond paste surfaces. The magnitude of the feature increases with decreasing
as described elsewhere [8]. Before each experiment the elec- terrace width. For terraces ≤5 atoms a shoulder appears on the
trode was flame annealed in an air–gas flame and cooled down low potential side. On Pt(3 3 1) the peak at 0.26 V has diminished
in an argon–hydrogen (UltraPure Plus, 6.0, Air Products) reductive and the two shoulders are observed primarily.
108 M.J.T.C. van der Niet et al. / Catalysis Today 202 (2013) 105–113

Fig. 2. The oxidative sweeps of the (a) Pt(S)[(n − 1)(1 1 1) × (1 1 0)] and (b) Pt(S)[n(1 1 1) × (1 0 0)] surfaces in 0.05 M NaOH.

According to the hard sphere model developed by Clavilier et al. for other surfaces. When multiple scans are taken the asym-
[26,27] the charge transferred per (1 1 0) step atom is given by metry increases slightly for successive scans for all surfaces,
2e suggesting slow but small structural changes during extensive
Qs,(1 1 0) = √ , (7) cycling.
(n − 2/3) 3d2
where e is the electron charge and d the diameter of a plat-
inum atom. In Fig. 3a the experimentally measured charge density
(referred to the area of the projection of the stepped surface on
(1 1 1) plane, i.e. Eq. (6)) is plotted against 1/(n − 2/3). On the dashed
line 1 electron is transferred per step platinum atom (i.e. a slope
of 241 ␮C cm−2 ). All data points are observed to cluster around
this line, except perhaps for the shortest terraces, suggesting that
the charge involved in the step-related feature is not significantly
dependent on pH, and corresponds to ca. 1 electron per Pt step
atom.

3.2. Pt(S)[n(1 1 1) × (1 0 0)]

Next, we discuss the platinum surfaces with (1 0 0) steps. Fig. 1c


shows the CV of Pt(5 3 3) in acidic and alkaline solution as an exam-
ple. For this step geometry the step induced feature is located at
higher potentials than for (1 1 0) steps. In acidic solution it is located
at 0.27 VRHE and reversible. In alkaline media, however, the fea-
ture becomes irreversible with its oxidative sweep peak potential
at 0.42 VRHE and its reductive sweep peak potential at 0.36 VRHE .
The slowness of the adsorption and desorption involved in the
Hter and Hstep features in alkaline media were recently studied
by impedance spectroscopy [22]. The shape of the peak is also
different compared to the CV in acidic solution (for details see
[22]).
Measurements in acidic media (0.5 M H2 SO4 and 0.1 M HClO4 )
on single crystals with varying (1 0 0) step density show that 1
electron is transferred per platinum step atom in the peak at
0.27 VRHE [28]. Fig. 2b shows the oxidative sweeps of a series
of Pt(S)[n(1 1 1) × (1 0 0)] electrodes with n = 14, 9, 5, 4, and 2
[Pt(15, 13, 13), Pt(5 4 4),Pt(3 2 2), Pt(5 3 3), and Pt(3 1 1)] in 0.05 M
NaOH. We observe the step induced feature at 0.42 VRHE for
all surfaces. The peak intensity increases with decreasing ter- Fig. 3. The charge density transferred in the step-related feature for the (a)
race width. For some surfaces the peak is more asymmetric than Pt(S)[(n − 1)(1 1 1) × (1 1 0)], and (b) Pt(S)[n(1 1 1) × (1 0 0)] surfaces in 0.05 M NaOH.
M.J.T.C. van der Niet et al. / Catalysis Today 202 (2013) 105–113 109

same trend. This slope of 10 mVRHE /pH unit has also been found for
the (1 1 0) and (1 0 0)-related peaks on a polycrystalline platinum
surface [29]. As a final remark, we note that the total transferred
charge in the peak in the phosphate buffers is the same for all pH
values (not shown).

4. The model

Our attempts to explain the seemingly paradoxical observations


outlined in the Introduction and partly illustrated in the foregoing
Results section will be based on three model assumptions:

1. The sharpness of the H-step feature is due to the co-adsorption


of H and OH, and the oxidative current associated with the fea-
ture is due to the replacement of Had by OHad , and vice versa for
the reductive current. Note that this co-adsorption of H and OH
happens only on the steps; it does not happen on the Pt(1 1 1)
terrace, where, as mentioned, the adsorption of H-ter and OH-
ter are well separated on the potential scale. Also note that
the effective co-existence of H and OH on the step site only
occurs in the narrow potential region of the step. This implies
that water is not stable in the step and spontaneously dissoci-
ates, an effect driven by the strong interaction of OH and/or O
with the (1 1 0) and (1 0 0) step sites, compared to the weaker
adsorption of OH on Pt(1 1 1). This effect is similar to the “half-
Fig. 4. The pH dependence of the step-induced feature for (a) Pt(5 5 3) and (b)
Pt(5 3 3) surfaces in various phosphate buffers, HClO4 and NaOH.
dissociated” water layer on a Ru(0 0 0 1) surface in UHV, now
well accepted in the surface science literature [35]. We showed
previously that such a “replacement” process, i.e. replacement of
In Fig. 3b the experimentally found charges are plotted vs.
H-step by OH-step or vice versa, may lead to a sharp voltammet-
1/(n − 1/3), as suggested by the hard sphere model of Rodes et al.
ric feature with apparent attractive lateral interactions even if
[28]:
only repulsive interactions exist between the adsorbates them-
2e selves [30]. This model explained very well many aspects of
Qs,(1 0 0) = √ . (8)
(n − 1/3) 3d2 the voltammetric features observed for a Pt(1 0 0) electrode in
bromide-containing solution, in which adsorbed hydrogen is
The dashed line represents 1 electron transferred per platinum step
replaced by adsorbed bromide in a very sharp current peak. The
atom (i.e. a slope of 241 ␮C cm−2 ). The measured charges also clus-
further consequences of this model assumption will be discussed
ter around the line corresponding to 1 electron per platinum atom.
throughout the paper. A further argument in favor of the for-
For narrow terraces (high step densities) the charge transferred
mation of OHad in the H-step feature is the fact that the peak
in the step-related peak appears to level off somewhat; the exact
potential shifts negatively upon the presence of Li+ in the solu-
interpretation of this effect is unknown but we suspect that such
tion, a cation supposed to have a strong interaction with OH,
narrow terraces, approaching Pt(1 1 0) and Pt(1 0 0), are not very
whereas Li+ has no influence on the H-ter feature [31].
stable and may reconstruct
2. In order to explain both the pH dependence of, and the charge
associated with the H-step feature (see Section 3), we rule out a
3.3. pH-dependence simple replacement reaction of the type:

Fig. 1 shows that the positions of both step-induced fea- Had + H2 O  OHad + 2H+ + 2e− (9)
tures changes with pH. Fig. 4 shows the position of the peaks
in the oxidative sweep as a function of pH. It is clear from First of all, reaction (9) would predict a charge of 2 electrons
Fig. 2 that the peak position does not shift with terrace width. per step site atom (if the maximum hydrogen and hydroxide
Therefore, two model surfaces are used in Fig. 4: Pt(5 5 3) for coverage is unity), which is in disagreement with our data and
the Pt(S)[(n − 1)(1 1 1) × (1 1 0)] series (a) and Pt(5 3 3) for the existing experiments of Rodes and Clavilier [26–28]. Second,
Pt(S)[n(1 1 1) × (1 0 0)] series (b). A series of phosphate buffers with reaction (9) predicts the “usual” pH dependence of 60 mV/pH
constant molarity was used for this purpose. In order to exclude unit, which is in disagreement with the experimentally observed
anion effects we have also included the peak positions in 0.1 M value of ca. 50 mV/pH unit reported here. From a purely modelis-
HClO4 and a mixture of 0.1 M KClO4 and 10−3 HClO4 (pH 3) and, tic point-of-view, the first discrepancy can in principle be “fixed”
0.05 M NaOH. For both surfaces the peak shifts linearly to higher by assuming stoichiometric coefficients different from 1 and/or a
potentials with increasing pH with 10 mVRHE /pH unit, which cor- maximum H coverage in the step smaller than 1, and the second
responds to a shift of 50 mV/pH unit on the normal hydrogen discrepancy by assuming that one of the surface-bonded species
electrode (NHE) scale. We emphasize again that this kind of pH is able to store charge such that the ratio of protons over elec-
dependence is very unusual (see Section 1), and contains a key trons exchanged during the reaction is different from 1. Purely
piece of information that has so far remained undiscussed in the formally, one could write a reaction of the type:
literature. We will suggest a possible explanation in the next sec- Had + xH2 O  xOHı+ + [x + 1]H+ + [(1 + ı)x + 1]e− (10)
ad
tion. The peak positions in HClO4 and KClO4 are identical to what
would be expected based on a fit through the phosphate buffer where x = i/j, with i the stoichiometric number of hydroxides
data. The NaOH data are not completely in line with the phosphate formed and j the stoichiometric number of adsorbed hydrogens
buffers, showing a slightly larger shift per pH unit, but do show the involved in the reaction. In this reaction, ı should be interpreted
110 M.J.T.C. van der Niet et al. / Catalysis Today 202 (2013) 105–113

as the charge flowing through the external circuit as a result focused on the closely-packed Pt(1 1 1) surface [33–35]. The general
of the formation of OHad , also known as the electrosorption consensus is that on Pt(1 1 1) water adsorbs molecularly at all cov-
valency. Note that ı could be either positive or negative. erages and temperatures (<180 K). Scanning tunneling microscopy
3. As an alternative to the somewhat artificial looking reaction (10), (STM) studies on an imperfect Pt(1 1 1) crystal show that water
one may assume that H and OH are still not the only species adsorbs preferentially on step sites, forming molecular chains [36].
involved in the H-step voltammetric feature. It is generally Temperature programmed desorption (TPD) studies show a stabi-
assumed that the Pt(1 1 1) surface is the only surface on which lization of water by the presence of step sites [25,37,38].
the formation of adsorbed OH and adsorbed O can be separated The stability of Oad as compared to OHad on a platinum sur-
on the potential scale [32]. On a different surface, or a step of face in ultra high vacuum can be studied by the co-adsorption of
(1 1 0) or (1 0 0) orientation, one could postulate: Oad (obtained by dissociative adsorption of O2 ) and H2 O. The co-
adsorption of H2 O and Oad on Pt(1 1 1) is known to produce OHad ,
Had + [x + y]H2 O  xOad + yOHad + [2x + y + 1]H+ which is incorporated in a hydrogen bonded network of H2 Oad and
+ [2x + y + 1]e− (11) OHad [39,40] via

The peak potential (on the RHE scale) corresponding to this reac- 2 H2 Oad + Oad → H2 Oad + 2 OHad . (13)
tion may in principle depend on pH if the ratio of Oad over OHad ,
i.e. x/y, would depend on pH in a non-trivial way. In a general For surface coverages of Oad lower than 0.25 ML, all Oad participates
sense, reaction (11) expresses the idea that the more stable water in the OH formation [41], so that there is no Oad left on the Pt(1 1 1)
dissociation product on a platinum surface other than (1 1 1) is surface. This clearly shows that H2 O is necessary to stabilize the
Oad , rather than OHad . As we will discuss in the next section, this formed OH species [41,42].
is in agreement with recent modeling experiments in UHV [24]. We have shown recently that on stepped surfaces (i.e. Pt(5 5 3)
and Pt(5 3 3)), OH is not as easily formed as on Pt(1 1 1), as O
The reaction equations above can be combined with reaction (1) appears to be more stable in step sites than OH [24], in contrast
from Section 1 to derive an equation for the hydrogen coverages to Pt(1 1 1), on which OH is more stable than O. When only the step
on the terrace and step sites as a function of potential. Details are sites are pre-covered with Oad , and subsequently left to interact
given in our earlier papers [22,30]. The resulting equation that can with water, additional high temperature peaks are observed in the
be used to model the “hydrogen region” is: TPD spectra, indicating OHad formation. However, upon also pre-
covering terrace sites with Oad , followed by water dosing, these
dter dstep high temperature peaks become smaller and an additional peak is
j = ter Fv + step Fv (12)
dE dE observed, characteristic of the desorption of OHad (by recombina-
where  ter is the hydrogen coverage on the terrace, and  ter is tion of 2 OHad ) from (1 1 1) terrace sites. Moreover, the amount
the corresponding charge (“electrosorption valency”) associated of exchange of O atoms between pre-adsorbed O atoms and H2 O
with hydrogen adsorption on the terrace, and where  step is the hardly increases even though over twice as many O adatoms are
hydrogen coverage on the step, and  step is the corresponding total present on the surface when the entire surface is covered with O.
charge (including the possible adsorption/desorption of OH) asso- These experiments indicate that Oterrace is more prone to OH forma-
ciated with hydrogen adsorption/desorption on the step. The model tion than Ostep , leaving unreacted O atoms at step sites [24,43,44].
assumes that the process occurring on the step can be modeled with This observation supports reaction (11) as the most realistic reac-
only a single variable (since it involves only a single voltammetric tion to take place in step sites. Our UHV experiments would also
peak), chosen to be the potential-dependent hydrogen coverage, suggest that (1 1 0) steps sites would produce somewhat more OHad
with any other coverage (OH or O) following from the expression than (1 0 0) step sites [24], that is, O is more stable, or less likely to
for the equilibrium constant corresponding to Eq. (10) or (11). F and react with water, in (1 0 0) step sites than in (1 1 0) step sites.
E in Eq. (12) have their usual meaning. As far as co-adsorption of H and H2 O is concerned, H2 O is sup-
Analytical expressions for  ter and  step as a function of potential posed to have a minor effect on the H adsorption properties on
(“isotherms”) are given in the Appendix of ref. [22]. The isotherm Pt(1 1 1) [13]. On the other hand, H pre-adsorbed on the stepped
for  ter is essentially the “mean-field” or Frumkin isotherm, which Pt(5 5 3) and Pt(5 3 3) surfaces can drastically alter the properties
has been shown to give a good approximation to the experimen- of co-adsorbed water in UHV [45,46].
tal voltammogram of Pt(1 1 1) as well as to exact Monte Carlo Table 1 gives an overview of adsorption energies of O, H, and
simulations [12,16]. This isotherm is characterized by a relatively OH obtained under UHV conditions or from DFT calculations along
weak repulsive lateral interaction (H H,ter , for the exact defi- with the maximum coverages of the species on (1 1 1) terrace sites
nition, see ref. [22]) between the adsorbed hydrogens on the and (1 0 0) and (1 1 0) step sites. Hydrogen adsorption shows two
terrace. The isotherm for  step follows from the exact solution for interesting features: it shows an increase in binding strength fol-
a one-dimensional lattice gas with an effective nearest-neighbor lowing (1 1 0)step < (1 1 1)terrace < (1 0 0)step (i.e. it binds weaker to
interaction energy (H H,step , defined and discussed in ref. [22]). (1 1 0) steps than to (1 1 1) terraces) [25] and the maximum surface
Fitting parameters of the model are the (effective) adsorption ener- coverage appears to be 0.5 at step sites, whereas it is 1 at terrace
gies of hydrogen on the terrace and step, ad,H-ter and ad,H-step , the sites [47,59].
(effective) lateral interaction energies H H,ter and H H,step , and
the (effective) electrosorption valencies  ter and  step . 6. Discussion

5. Water dissociation on well-defined platinum surfaces: In the blank CV of stepped platinum surfaces, the (1 1 0) step
the UHV perspective peak is located at lower potentials than the (1 0 0) step peak. This
indicates that either Had is more strongly bound to (1 0 0) steps than
This section briefly summarizes some of the relevant experi- to (1 1 0) steps and/or the co-adsorbant (OHad /Oad ) is more strongly
ments related to the adsorption of water, H, and OH on stepped bound to (1 1 0) steps. As shown in Table 1, UHV experiments
platinum surfaces in ultra high vacuum (UHV), for comparison with suggest that Had has a binding energy in the order of step(1 0 0) >
the electrochemical situation. There have been many studies on the terrace(1 1 1) > step(1 1 0) [25]. In this work we also observe a step
adsorption of water on platinum in UHV, though most of them have binding energy in the order of step(1 0 0) > step(1 1 0) for both alkaline
M.J.T.C. van der Niet et al. / Catalysis Today 202 (2013) 105–113 111

Maximum coverages and adsorption energies for H, OH and O on Pt(1 1 1) terraces, (1 1 0) steps, and (1 0 0) steps. Coverages are based on UHV modeling experiments; adsorption energies on DFT calculations. Binding energies

H2 O co-adsorption is necessary to stabilize the formed OH. The most stable structure is a OH/H2 O overlayer with a maximum total coverage of 0.67 ML and a 1:1 ratio of H2 O and OH, corresponding to 0.33 ML of OH. The

Estimated from TPD results and a calculation of the pumping speed of the molecular beam chamber for H2 . DFT calculations do show that it is energetically less favorable to cover the second half of the step sites with Had [60].
and acidic media. It is well known that the H-ter terrace peak on

−5.87 to −5.10 [58]


Pt(1 1 1) (the adsorption feature below 0.4 V in Fig. 1a) is located
at a position in agreement with the hydrogen binding strength
in ultra high vacuum [10]. As concerns possible co-adsorbants,

0.5 [58]
both O and OH bind stronger to (1 0 0) steps than to (1 1 0) steps,
whereas (remarkably) H2 O binds stronger to (1 1 0) steps [25].

O
This indicates that even though OH and O are likely candidates for
co-adsorption, they do not seem to dominate the processes that

unknown
determine the peak potential. Although H2 O appears to show the

−2.7 [62]
right energy trend, H2 O as the only species adsorbing in step sites
OH

at low potentails (other than hydrogen) does not provide us with


an oxygen donor needed at the steps in the surface to start electro-
oxidation reactions at potentials lower than 0.6 V [23,63,64]. These
−0.59 [61], −0.84 [25]

arguments render it very unlikely that water is the only species


adsorbed in the step sites at potentials between 0.4 and 0.6 V.
In alkaline media the H-step peaks are observed at higher poten-
tials compared to acidic media. The HClO4 and NaOH data in Fig. 4
(1 0 0)step

0.5 [59] e

exclude anion effects as the source of this anomalous pH depen-


dence. The peak shift toward more positive potentials on the RHE
H

scale indicates that Had is more stable in alkaline media than in


From background dosing. Higher coverage states can be reached by extended temperature cycling [53,57], O-beam irradiation [53], or exposure to NO2 [55].

acidic media and/or that O/OH is less stable in alkaline media. As


−5.66 to −4.9 [58]

stated in assumption 3 of the model, the anomalous pH depen-


dence can in principle be explained by assuming that the O:OH
ratio in the steps varies with pH. From TPD measurements in UHV
0.5 [58]

it is clear that OHad binds weaker to the surface than Oad [24].
The shift of the peak potential of the step-related features with pH
O

suggests that the adsorption energy of the co-adsorbant decreases


with increasing pH value. This could indicate that relatively more
unknown
−2.7 [62]

OH binds to steps in alkaline media, i.e. oxidation to Oad in steps is


less pronounced than in acidic media.
OH

We would like to stress at this point that the potential of zero


total charge (pztc) shifts with 60 mVNHE /pH unit (0 mVRHE ) on
Pt(1 1 1) [65], whereas on Pt(1 1 0) a shift of around 48 mVNHE /pH
−0.41 [61], −0.47 [25]

unit (12 mVRHE ) has been observed [65] (with a similar shift also
observed for polycrystalline platinum [66]). These values are very
similar to the peak shift we observe, suggesting that the processes
causing the pH dependent peak shift and the pztc shift are probably
0.5 [25] d
(1 1 0)step

related, if not the same.


In alkaline media the Hupd peaks become more irreversible, indi-
H

cating a slower adsorption process than in acidic media both on


for H are vs. the hydrogen molecule, for OH and O vs. the corresponding species in vacuum.

terraces and steps [22]. The peak for (1 0 0) steps is broader and
0.25 [53–56] c

appears more irreversible than for (1 1 0) steps. Under UHV condi-


−4.2 [42]

tions (1 0 0) steps are less susceptible to OH formation than (1 1 0)


steps, favoring Oad at step sites [24]. If H2 O oxidizes all the way
through to Oad , it may be harder to react back (to H2 O), making the
O

No absolute values are given, but a similar  H is found as on Pt(5 3 3).

(1 0 0) peak more irreversible. In a previous paper we showed, using


impedance spectroscopy, that on Pt(1 1 1) the H adsorption process
maximum coverage of this overlayer is 0.75 ML, but it is less stable.
−2.4 [42], −2.51[62]

at terraces and steps is significantly slower in alkaline media than in


0.33 [39,48–52] b

acidic media. This slow adsorption and desorption to both terrace


and step sites below 0.5 VRHE in alkaline media has been ascribed to
the fact that hydrogen needs to be abstracted from H2 O rather than
from H3 O+ (reactions (1) and (2)). For a detailed discussion, see ref.
OH

[22].
In our previous UHV modeling study [24], we found that the OH
formation on (1 1 1) terraces is less affected by (1 0 0) steps than by
−0.14 [42], −0.42 [61]

(1 1 0) steps, as (1 0 0) steps contain more unreactive Oad . It is not


easy to say whether this is reflected in the electrochemical blank
voltammetry. Comparing left and right panels of Fig. 2, one could
(1 1 1)terrace

1.00 [47]a

argue that the OHad formation peak corresponding to the (1 1 1)


terraces is indeed also less affected for the surfaces with (1 0 0)
H

steps, especially at high step density. This would then be due to


the low interaction between the Oad -covered (1 0 0) steps and the
terrace OHad , compared to the (1 1 0) steps, which contain relatively
 max [ML]
Eads [eV]

more OHad .
TPD.
Table 1

As a final point we would like to discuss the total charge trans-


a

ferred in the step feature. If the feature is indeed due to the


112 M.J.T.C. van der Niet et al. / Catalysis Today 202 (2013) 105–113

subsitution of H by OH and/or O, one would intuitively expect a The model suggested here is primarily based on providing a
charge transfer of 2 or more electrons per platinum atom. This is consistent interpretation for the three observations/discrepancies
not what is observed as we find values very close to 1 (i.e. from mentioned above. To the extent that there is no direct spectroscopic
the slope of Fig. 3). In acidic media also a charge transfer of 1 elec- evidence for some of the model assumptions, the model is specu-
tron per platinum atom is found [26–28]. Table 1 provides us with lative. Some referees of this paper have considered this a severe
a clue for this apparent discrepancy: in UHV a maximum cover- weakness of the model suggested here. However, obtaining direct
age of 1 H atom per 2 platinum atoms has been suggested [59]. spectroscopic evidence for OH or O adsorption on Pt single-crystal
For oxygen the same maximum coverage is found [58]. For OH no electrodes under electrochemical conditions is highly challenging.
maximum coverage is known. A maximum hydrogen coverage on Current lack of this kind of information should not be a reason for
steps of ∼0.5 ML could possibly explain why both in acidic [26–28] not formulating a consistent interpretation of the electrochemistry
and alkaline media (this work) a charge transfer close to 1 electron of platinum, especially if it is clear that the current interpretation
per platinum atom is found, even if reactions (10) and (11) require present in the literature is flawed. Moreover, even a model based
more than 1 electron (2 if OH is the final product). Deviations from on direct spectroscopic evidence will need to provide a consistent
the “1 electron line” may be related to varying O:OH ratios, or to explanation for the above three observations.
the total amount of O containing species varying with step density. Finally, we point out that our model implies that, except for
The idea that Had desorption and OHad adsorption are not decou- Pt(1 1 1), oxygen-containing species are already adsorbed on plat-
pled processes on Pt(1 0 0) and Pt(1 1 0) is not new; see e.g. the inum in what is traditionally called “the hydrogen region”. These
discussion in ref. [32]. Marichev has recently discussed the evi- oxygen-containing species appear to be the active species for elec-
dence of OH adsorption in the hydrogen region of polycrystalline trooxidation. Also, OHad is less stable in steps than on terraces
platinum [67]. Furthermore, Alsabet et al. [68] have discussed the and Oad appears to be preferentially formed there. Although we
experimental evidence that on polycrystalline platinum, Pt O is believe our model is a step forward in a consistent interpretation of
formed without any noticeable formation of an OHad intermediate. the electrochemistry of water at well-defined platinum electrodes,
This would be in agreement with the idea that on step sites, O is an important aim for the future remains the quantification of the
more stable than OH, certainly in the oxide-formation region on amount of OHad and Oad at various surfaces as a function of poten-
polycrystalline platinum. tial and to establish which of the two species is the oxygen donor
in electrocatalytic oxidation reactions.
7. Conclusion
Acknowledgements
In this paper, we have discussed three important discrepancies
in the current interpretation of the effect of water dissociation on This work was supported financially by the Netherlands Orga-
the blank cyclic voltammetry of stepped platinum surfaces: nization for Scientific Research (NWO). JMF thanks the support
of Generalitat Valenciana (Spain) through PROMETEO/2009/45
• For H adsorption both H-ter and H-step contributions are identi- (Feder) project.
fied. For OH adsorption only OH-ter has been identified.
• Different shapes (broad vs. sharp) of the H-ter and H-step peaks
References
imply different lateral interactions between hydrogen adatoms
at terraces and steps, i.e. repulsive vs. attractive interactions. [1] M.T.M. Koper (Ed.), Fuel Cell Catalysis: A Surface Science Approach, John Wiley
• The H-step peak has a non-trivial pH-dependence of & Sons, Hoboken, 2009.
[2] A.J. Bard, L.R. Faulkner, Electrochemical Methods: Fundamentals and Applica-
50 mVNHE /pH unit. tions, 2nd ed, Wiley-Interscience, 2004.
[3] J. Clavilier, R. Faure, G. Guinet, R. Durand, Journal of Electroanalytical Chemistry
107 (1980) 205.
Previous interpretations of the blank voltammetry of stepped
[4] J. Clavilier, Journal of Electroanalytical Chemistry 107 (1980) 211.
platinum surfaces cannot explain these observations. We have pro- [5] J. Clavilier, R. Durand, G. Guinet, R. Faure, Journal of Electroanalytical Chemistry
posed here a model that can deal satisfactorily with these three 127 (1981) 281.
[6] J. Clavilier, D. Armand, B.L. Wu, Journal of Electroanalytical Chemistry 135
discrepancies, and in our opinion this represents an important
(1982) 159.
step in correctly understanding platinum electrochemistry. In our [7] J. Clavilier, D. Armand, Journal of Electroanalytical Chemistry 199 (1986) 187.
model, the H-step peak is not due to just ad- and desorption of [8] J. Clavilier, D. Armand, S.G. Sun, M. Petit, Journal of Electroanalytical Chemistry
hydrogen, but to the replacement of H with O and/or OH. This 205 (1986) 267.
[9] D. Armand, M.L. Rosinberg, Journal of Electroanalytical Chemistry 302 (1991)
replacement process can straightforwardly explain the sharpness 191.
of the voltammetric peak, following the reasoning of ref. [30]. The [10] G. Jerkiewicz, Progress in Surface Science 57 (1998) 137.
O:OH ratio presumably varies with step density and medium. In [11] M.T.M. Koper, J.J. Lukkien, Surface Science 498 (2002) 105.
[12] G.S. Karlberg, T.F. Jaramillo, E. Skúlason, J. Rossmeisl, T. Bligaard, J.K. Nørskov,
alkaline media it appears that relatively more OH is adsorbed in Physical Review Letters 99 (2007) 126101.
steps than in acidic media where more O is adsorbed, providing [13] M.T.M. Koper, Faraday Discussions 140 (2008) 11.
an explanation for the non-trivial pH dependance. We speculate [14] R.J. Nichols, A. Bewick, Journal of Electroanalytical Chemistry 243 (1988) 445.
[15] M. Osawa, M. Tsushima, H. Mogami, G. Samjeské, A. Yamakata, Journal of Phys-
that this effect could also be at the origin of the higher activity of ical Chemistry C 112 (2008) 4248.
platinum in alkaline media for oxidation reactions. [16] M.T.M. Koper, J.J. Lukkien, Journal of Electroanalytical Chemistry 485 (2000)
We find a charge transfer of close to 1 electron per step plat- 161.
[17] K. Al Jaaf-Golze, D.M. Kolb, D. Scherson, Journal of Electroanalytical Chemistry
inum atom in alkaline media. Previously, a charge transfer close
200 (1986) 353.
to 1 electron per platinum atom was also found in acidic media [18] M. Wakisaka, H. Suzuki, S. Mitsui, H. Uchida, M. Watanabe, Langmuir 25 (2009)
[26–28]. Interestingly, both H and O have a saturation coverage of 1897.
[19] P. Vassilev, M.T.M. Koper, R.A. van Santen, Chemical Physics Letters 359 (2002)
0.5 ML on step sites in UHV [25,58,59]. The model may explain the
337.
amount of charge transferred, if the maximum coverage of H in the [20] A. Berná, V. Climent, J.M. Feliu, Electrochemistry Communications 9 (2007)
step is close to 0.5 and the number of electrons transferred per H is 2789.
2 (or more), as would be needed to make OH/O. A variation in the [21] M.T.M. Koper, J.J. Lukkien, N.P. Lebedeva, J.M. Feliu, R.A. van Santen, Surface
Science 478 (2001) L339.
ratio and total amount of adsorbed oxygen species may explain the [22] K.J.P. Schouten, M.J.T.C. van der Niet, M.T.M. Koper, Physical Chemistry Chem-
variation of the amount of charge transferred with step density. ical Physics 12 (2010) 15217.
M.J.T.C. van der Niet et al. / Catalysis Today 202 (2013) 105–113 113

[23] M.T.M. Koper, S.C.S. Lai, E. Herrero, in: M.T.M. Koper (Ed.), Fuel Cell Catalysis: [46] M.J.T.C. van der Niet, A. den Dunnen, M.T.M. Koper, L.B.F. Juurlink, Physical
A Surface Science Approach, John Wiley & Sons, Hoboken, 2009, pp. 159–207. Review Letters 107 (2011) 146103.
[24] M.J.T.C. van der Niet, A. den Dunnen, L.B.F. Juurlink, M.T.M. Koper, Angewandte [47] Ş.C. Bădescu, K. Jacobi, Y. Wang, K. Bedürftig, G. Ertl, P. Salo, T. Ala-Nissila, S.C.
Chemie International Edition 49 (2010) 6572. Ying, Physical Review B 68 (2003) 205401.
[25] M.J.T.C. van der Niet, A. den Dunnen, L.B.F. Juurlink, M.T.M. Koper, Journal of [48] J.R. Creighton, J.M. White, Surface Science 139 (1984) 43.
Chemical Physics 132 (2010) 174705. [49] G.S. Karlberg, F.E. Olsson, M. Persson, G. Wahnström, Journal of Chemical
[26] J. Clavilier, K. El Achi, A. Rodes, Journal of Electroanalytical Chemistry 272 (1989) Physics 119 (2003) 4865.
253. [50] C. Clay, S. Haq, A. Hodgson, Physical Review Letters 92 (2004) 046102.
[27] J. Clavilier, K. El Achi, A. Rodes, Chemical Physics 141 (1990) 1. [51] G. Held, C. Clay, S.D. Barrett, S. Haq, A. Hodgson, Journal of Chemical Physics
[28] A. Rodes, K. El Achi, M.A. Zamakhchari, J. Clavilier, Journal of Electroanalytical 123 (2005) 064711.
Chemistry 284 (1990) 245. [52] T. Schiros, L.Å. Näslund, K. Andersson, J. Gyllenpalm, G.S. Karlberg, M. Odelius,
[29] R. Gisbert, G. García, M.T.M. Koper, Electrochimica Acta 55 (2010) 7961. H. Ogasawara, L.G.M. Pettersson, A. Nilsson, Journal of Physical Chemistry C 111
[30] N. Garcia-Araez, J.J. Lukkien, M.T.M. Koper, J.M. Feliu, Journal of Electroanalyt- (2007) 15003.
ical Chemistry 588 (2006) 1. [53] J.L. Gland, Surface Science 93 (1980) 487.
[31] C. Stoffelsma, P. Rodriquez, G. García, N. Garcia-Araez, D. Strmcnik, N. Marković, [54] K. Schwaha, E. Bechtold, Surface Science 65 (1977) 277.
M.T.M. Koper, Journal of the American Chemical Society 132 (2010) 16127. [55] D.H. Parker, M.E. Bartram, B.E. Koel, Surface Science 217 (1989) 489.
[32] N.M. Marković, P.N. Ross, Surface Science Reports 45 (2002) 117. [56] U. Starke, K. Heinz, N. Materer, A. Wander, M. Michl, R. Döll, M.A. van Hove, G.A.
[33] P.A. Thiel, T.E. Madey, Surface Science Reports 7 (1987) 211. Somorjai, Journal of Vacuum Science and Technology A 10 (1992) 2521.
[34] M.A. Henderson, Surface Science Reports 46 (2002) 1. [57] G.N. Derry, P.N. Ross, Surface Science 140 (1984) 165.
[35] A. Hodgson, S. Haq, Surface Science Reports 64 (2009) 381. [58] P.J. Feibelman, S. Esch, T. Michely, Physical Review Letters 77 (1996) 2257.
[36] M. Morgenstern, T. Michely, G. Comsa, Physical Review Letters 77 (1996) 703. [59] A.T. Gee, B.E. Hayden, C. Mormiche, T.S. Nunney, Journal of Chemical Physics
[37] D.C. Skelton, R.G. Tobin, G.B. Fisher, D.K. Lambert, C.L. DiMaggio, Journal of 112 (2000) 7660.
Physical Chemistry B 104 (2000) 548. [60] R.A. Olsen, S.C. Badescu, S.C. Ying, E.J. Baerends, Journal of Chemical Physics 120
[38] M.L. Grecea, E.H.G. Backus, B. Riedmüller, A. Eichler, A.W. Kleyn, M. Bonn, Jour- (2004) 11852.
nal of Physical Chemistry B 108 (2004) 12575. [61] T. Vehvilä inen, P. Salo, T. Ala-Nissila, S.C. Ying, Physical Review B 80 (2009)
[39] A. Michaelides, P. Hu, Journal of the American Chemical Society 123 (2001) 035403.
4235. [62] M.T.M. Koper, Electrochimica Acta 56 (2011) 10645.
[40] A. Michaelides, P. Hu, Journal of Chemical Physics 114 (2001) 513. [63] N.P. Lebedeva, A. Rodes, J.M. Feliu, M.T.M. Koper, R.A. van Santen, Journal of
[41] A. Shavorskiy, M.J. Gladys, G. Held, Physical Chemistry Chemical Physics 10 Physical Chemistry B 106 (2002) 12938.
(2008) 6150. [64] G. García, M.T.M. Koper, Physical Chemistry Chemical Physics 11 (2009) 11437.
[42] G.S. Karlberg, Physical Review B 74 (2006) 153414. [65] V. Climent, R. Gómez, J.M. Orts, A. Aldaz, J.M. Feliu, The Electrochemical Soci-
[43] M.J.T.C. van der Niet, A. den Dunnen, L.B.F. Juurlink, M.T.M. Koper, Physical ety Proceedings: The Electrochemical Double Layer, vol. 97-17, Pennington, NJ,
Chemistry Chemical Physics 13 (2011) 1629. 1997, pp. 222–237.
[44] M.J.T.C. van der Niet, O.T. Berg, L.B.F. Juurlink, M.T.M. Koper, Journal of Physical [66] A.N. Frumkin, O.A. Petrii, Electrochimica Acta 20 (1975) 347.
Chemistry C 114 (2010) 18953. [67] V.A. Marichev, Electrochimica Acta 53 (2008) 7952.
[45] M.J.T.C. van der Niet, I. Dominicus, M.T.M. Koper, L.B.F. Juurlink, Physical Chem- [68] M. Alsabet, M. Grden, G. Jerkiewicz, Journal of Electroanalytical Chemistry 589
istry Chemical Physics 10 (2008) 7169. (2006) 120.

Вам также может понравиться