You are on page 1of 12

24 7

J. Electroanal. Chem ., 329 (1992) 247-258


Elsevier Sequoia S .A., Lausanne
JEC 05050

Superficial structure of Au(110) at the electrochemical


interface *

A . Hamelin
Laboratoire d Electrochimie Interfaciale du CNRS, 1 place A . Briand, 92195 Meudon (France)

(Received 3 December 1991 ; in revised form 30 January 1992)

Abstract

Electrochemical characterization of the Au(110) face in aqueous solutions using cyclic voltammetry
and differential capacity-potential measurements was found to be consistent with in-situ structural
observations by surface X-ray scattering and scanning tunneling microscopy . The structural changes of
the Au(110) surface in the double-layer range of potential and their kinetics make determination of the
potential of zero charge of the 1 x 1 structure difficult ; the observation of the potential of zero charge
for a reconstructed structure is ambiguous .

INTRODUCTION

For Au(100), (110) and (111)-aqueous solution interfaces, reconstruction and


lifting of the reconstruction of the top layer(s) of surface atoms were suggested to
explain the secondary effects observed on differential capacity-potential (C(E))
curves, and cyclic voltammograms (CV) in the double-layer range of potential [1,2] .
Indeed, at first in our laboratory, the hysteresis between positive- and negative-
going sweeps of C(E) curves and CVs, and the different values of the potential of
zero charge (pzc) observed for one face, were suggested as originating from a
change in superficial structure of the gold face with imposed potential . To be able
to draw such conclusions, the chemical conditions at the interface have to be
strictly controlled.
Other authors reported electrochemical results for these gold faces, along with
electron diffraction data after emersion of the electrodes, which were interpreted

Dedicated to Professor Roger Parsons on the occasion of his retirement from the University of
Southampton and in recognition of his contributions to electrochemistry .

0022-0728/92/$05 .00 © 1992 - Elsevier Sequoia S.A . All rights reserved


248

by reconstruction and lifting of the reconstruction [3,4]. However, for Au(110) the
reconstructed structure was observed to be stable over the entire double-layer
region of potential [5] .
Ultrahigh vacuum observations by electron diffraction have demonstrated that
these faces are spontaneously reconstructed at room temperature [6] ; non-recon-
structed structures are observed for these three faces at higher temperatures . The
transition to a non-reconstructed structure occurs at a lower temperature (350°C)
for Au(110) than for the two other gold faces .
When reviewing [7] what is known about the superficial structure of gold faces
at electrochemical interfaces, comparison with the results of in-situ techniques,
particularly surface X-ray scattering (SXRS) and scanning tunneling microscopy
(STM) was emphasized . (Scattering of X-rays from a truncated lattice in vacuum is
reviewed in ref . 8 ; the case of gold Au(110) is analyzed in ref . 9 . The application of
SXRS to electrochemical interfaces is possible [10,11] . Early works using STM are
reviewed in refs . 12 and 13 ; STM has been applied to electrochemical systems
[141.)
In the past, in-situ SXRS has been applied to electrochemical gold systems to
study surface reconstruction [11] . STM has already been used to investigate gold
faces at electrochemical interfaces [15,16], but only a few papers give atomic
resolution images [17-19] .
To date, for Au(100), SXRS results [11] and STM observations [17] have
demonstrated that reconstructed and (1 x 1) structures exist at the electrochemical
interfaces in the ranges of potential suggested by CV and C(E) curves . Similarly
for Au(111), X-ray results [20] and STM observations [19] are roughly in agreement
with conventional electrochemical data [7] .
In this paper attention will be focused on electrochemical results obtained at
the Au(110)-aqueous solution interfaces, and their relationship with in-situ struc-
tural probe observations using STM [18] and SXRS [211 .

ELECTROCHEMICAL RESULTS

For Au(110) in aqueous solutions, the electrochemical results already published


[1,2,22,23] and obtained recently, are not fully understood, for either nearly
non-adsorbing ions in solution or for adsorbing anions.

Nearly non-adsorbing media

In neutral solutions (aqueous potassium hexafluorophosphate or sodium fluo-


ride), although for Au(110) the contribution of the diffuse part of the double layer
to the C(E) curves is observed clearly for positive-going sweeps (Fig . 1), oxide
begins to form at low positive charge densities, shown by the dispersion of the
C(E) curves with the frequency of the alternating signal (Fig . 4 of ref. 22) and by
CVs (Fig . 5 of ref . 22) . As seen in Fig . 1, the C(E) curves do not have the regular
shape observed for other gold faces in potassium hexafluorophosphate solutions



249

40 ( / ' 1'

l1 l. ~
1
4
1/

//
11\ /.
20

-0 .5 0 0' .5
E/V (SCE)
Fig . 1 . C(E) curves for Au(110) in KPF6 solutions, positive-going sweeps, frequency of alternating signal
10 Hz, 8 mV s - ', 22±2°C: - 5 .3, --- 10.8, 21 .4 and -- - 30 .0 mM [221 .

[22] . From these curves it seems that there are two C(E) minima which deepen
with dilution of the solution : one appears clearly, giving the values of the pzc
discussed in ref. 22, and the other, albeit less clear-cut, minimum appears about
0.2 V more positive (Fig. 1).
In perchloric acid solutions the formation of oxide occurs at more positive
potentials than in neutral solutions, so the double layer can be studied over a
range of positive charge densities (Fig . 2(a)) . The dual contribution of the diffuse
part of the double layer is shown for Au(110) in 0 .01 M perchloric acid in Fig . 2b ;
for higher concentrations of this acid the contribution of the diffuse layer is clearly
asymmetric [23] . These C(E) curves were recorded for positive-going sweeps .
Observation of the negative-going sweeps and comparison with the positive-going
sweeps is instructive . For 0 .01 and 0 .1 M perchloric acid, under rigorously clean
conditions, for a well prepared Au(110) face (flame treated and cooled in ultrapure
water just before being put in contact with the solution), the currents for charging
and discharging the double layer were found to be either identical (Fig . 2(a)) or not
identical (Fig . 3) . When for 0.01 M perchloric acid the CV was identical for
charging and discharging the double layer, the C(E) curve showed only a slight
hysteresis between 0 .2 and 0 .6 V ; however, there seemed to be a dual contribution
of the diffuse part of the double layer (Fig . 4(a)) . When these two currents were
found not to be identical, the C(E) curve showed a significant hysteresis (Fig .
4(b)) . Results given in Figs . 2 and 3 were obtained in a quiescent solution as well as
in a stirred solution . These results were obtained with the same sample (110)
oriented . All these observations suggest that there are several steps in the change
25 0

2
j/ l
zA.cm
5

100

10
0 0
E/V(SCE)

-300

50

40

30

20

Fig. 2 . (a) CV for Au(110) in 0 .01 M perchloric acid 23± 1°C, 80 mV s - ' ; no change when the negative
limit is set at -0 .10 or -0 .15 V vs . SCE. CV of the double layer region enlarged x40 for different
positive limits : ( ) 1 .45 V ; (---) 1 .1 V ; ( ) CV after four short cycles . Although
experimentally RHE was used, plots are drawn vs . SCE . (b) C(E) curves for Au(110) in 0 .01 M
perchloric acid solution, positive-going sweep, 1±0 .5°C, 20 Hz, 5 mV s -t . A rather sigmoid shape is
observed at slightly positive charge densities [23].

in the superficial structure of the Au(110) face at the electrochemical interface


with imposed potential .
In 0 .1 M perchloric acid, narrowing the range of potential on the positive side,
to avoid interference from any preliminary step of oxidation of Au(110), does not
eliminate the dual contribution of the diffuse layer (Fig . 4) .
In 0 .1 M perchloric acid, when the potential was held at -0 .3 V vs . a saturated
calomel electrode (SCE) or varied from -0 .35 to -0 .20 V for 5 min before

25 1

Fig. 3 . CVs for Au(110) in 0 .1 M ( ) and 0.01 M (---)perchloric acid solutions; SO mV s - t,


25±2°C after holding the potential at -0 .3 V or varying the potential from -0 .35 to -0 .20 V
for 5 min .

I/µA

i
C/ IsF.cm

C µF.cm2
SO

0
40

30

20 .05 0 0.5
20 ' '
D 0 .5 E/V(SCE)
(a) E/V(SCE) (b)
Fig. 4 . (a) C(E) curves for Au(110) in 0 .01 M perchloric acid solution, positive- and negative-going
sweeps, 20 Hz, 5 mV s - t, 25±2°C ; . . . corresponding CV at 50 mV s-t . (b) C(E) curves for
Au(110) in .10 M perchloric acid for different positive limits, 20 Hz, 5 mV s - t .
25 2

100

5a

100

so

E/V (SCE)

Fig. 5 . C(E) curves for Au(110) in 0 .1 M NaCI: (a) positive and negative sweeps, 20 Hz, 5 mV s - ' ; (b)
positive-going sweep for 8, 12, 20, 80 and 180 Hz, 5 mV s - ' .

recording the CV, the contribution of the diffuse part of the double layer was more
pronounced on the positive-going sweep but the potential of the minimum was not
shifted (Fig. 3) .

Adsorbing halide solutions

In solutions containing an anion which adsorbs at gold electrodes (for instance


halides) only a slight hysteresis between the positive- and negative-going sweeps
exists, but one of the capacity peaks is clearly frequency dependent (Fig . 38 of ref.
2 and Fig . 8 of ref. 7 for iodide and bromide respectively) . For Au(110) in aqueous
0 .1 M NaCl solution, differential C(E) curves are given in Fig . 5 . The slight
hysteresis between both sweeps of the C(E) curves and the strong changes in
amplitude of the capacity peak at -0 .1 to -0 .2 V vs. SCE with the frequency of
the alternating signal used for measurements, are consistent with observations in
bromide and iodide solutions, and suggest some changes in surface structure at
about -0.1 V vs . SCE .
All the electrochemical results described above suggest that there are complex
surface reconstructions at the Au(110) face in contact with aqueous perchloric
solutions when the imposed potential is varied . The reconstruction would be
different in the case of strong anionic adsorption . Two clear-cut examples are

25 3

given below to show the consistency of these results with in-situ structural probe
observations .

COMPARISON WITH IN-SITU SCANNING TUNNELING MICROSCOPY OBSERVATIONS

STM as an atomic-resolution structural probe is providing opportunities for


exploring real space atomic distribution . It is a local structural technique allowing
elucidation of individual components of complex structures provided that true
atomic resolution is achieved .
In the STM experiment the tungsten tip is brought close enough to the gold face
for an electron tunneling current to flow . As explained in ref. 14, this tunneling
current comes from the overlap of the "atmosphere" of electrons that surrounds
the tip with the "atmosphere" of electrons above the sample . The tip scans over
the gold surface, gray-scale images are obtained "in which the height, z, is
displayed by a shade of gray (typically black for lowest and white for highest) at
each x, y position". Detailed atomic arrangements, as well as the symmetry of the
face and its reconstruction, can be determined from such images .
An initial in-situ STM study has been reported recently for Au(110) in 0 .1 M
perchloric acid [19], carried out at Purdue University, in collaboration with the
present author . The surface preparation of this gold face was the same as that of
the Au(110) faces used to obtain the electrochemical results described above ; it
was undertaken at LEI-CNRS . The gold face (covered with clean water) was

(a) fbI

Fig. 6. STM images of reconstructed Au(110) in 0 .1 M perchloric acid at -0 .3 V vs . SCE : (a) top view
image showing (1 x 2) and (1 x 3) regions ; (b) height-shaded close-up image (30' from surface normal) of
largely (I x2) domains [191,
254

placed in the bottom of a small Teflon vessel ; four drops of solution were enough
to fill this cell . When oxygen was not removed from the solution, it was rapidly
exhausted at the level of the gold face (from observation of the CVs) in such a
mechanically isolated set-up .
Atomic-resolution images were obtained at positive charge densities where a
(1 X 1) structure is observed (Fig . 6 of ref. 19) and at negative charge densities
where reconstructed structures are found : (1 x 2) and (1 x 3) structures and a
mixture of symmetries . At -0 .3 to -0 .4 V vs . SCE stacked sets of parallel ribbon
segments located along the [110] direction are discerned . Each ribbon consists of
three parallel rows of gold atoms, the interatomic spacing along the [110] direction
being that of unreconstructed Au(110) (2 .9 ± 0 .2 A) (Fig . 6) . The spacing between
the ribbons is 4 .0 X n A where n = 2 or 3 . These observations are described in
more detail in ref . 19 . They suggest that some surface relaxation occurs for the top
and underlying atoms, producing corrugations in the z direction, i .e . perpendicular
to the surface, and a mixture of subtly different (1 x 2) structures appears to be
present . The lifting of this reconstruction occurs within a few seconds, upon
altering the potential to about 0 .2 V vs . SCE.
The reconstruction described above and its kinetics, markedly faster than for
Au(100), are in agreement with the conventional electrochemical results for
Au(110) in perchloric acid described above .

COMPARISON WITH IN-SITU SURFACE X-RAY SCATTERING DATA

Since the gold surface has a small number of atoms, about 10 1' atoms cm - Z,
surface X-ray diffraction measurements require a high-brightness X-ray source .
Measurements at the Au(110) electrode surface were carried out at beam line
X22B (A = 1 .53 A) at the National Synchrotron Light Source at Brookhaven
National Laboratory, in collaboration with the present author .
The diffraction from a crystalline surface yields three-dimensional Bragg peaks
and two-dimensional diffraction features . (X-ray diffraction is conveniently defined
by the scattering vector Q = k' - k, the difference of the incident and scattered
wave vectors .) The two-dimensional diffraction features are referred to as crystal
truncation rods (CTRs) [9] or non-specular reflectivity [24] . The positions of these
rods-within the surface plane-are analogous to low energy electron diffraction
(LEED) patterns . Measurements of CTRs provide detailed information on the
atomic positions and surface roughness . The scattered X-ray intensity is sensitive
to sub-surface distortions as well as to the top atom layer .
The Au(110) surface normal (optical axis) was aligned accurately to within 0 .05°
of the (110) planes [21] . Despite the considerable penetration depth of X-rays
(approximately 1 mm), a thin layer cell has to be used to reduce the bulk water
scattering . In the cell, a 10 Am-thick electrolyte layer is contained by a 6 µm-thick
propylene window . For the Au(110) measurements the surface was in contact with
0 .1 M NaCl electrolyte [21]. A computer-controlled four-circle diffractometer was
used to access the scattering in reciprocal space, i .e . the Fourier transform space of
255

real space ; Miller indices of the reciprocal space are H, K, L . The intensity profile
of scattered X-rays was recorded, either along paths in reciprocal space at various
imposed potentials, or as a function of the potential imposed at the Au(110) face
at fixed positions in reciprocal space . Further details are given in ref. 21 .
In the current set of measurements, scans were carried out along the L axis in
reciprocal space-normal to the "dense atomic rails" of the (1 x 1) (110) face-
with a grazing incidence angle of 1 .5° . The following results were obtained for
Au(110) in 0 .1 M NaCl :
at 0 V vs . SCE only integer reflections are observed corresponding to the (1 x 1)
structure ; when the potential is shifted to -0.3 V, third-order reflections appear
corresponding to a (I x 3) structure (Fig . 1 of ref. 21) ;
the potential dependence of the intensity at (0 .1, 0 .1, 5/3)-when varying the
imposed potential linearly with time in both directions-clearly displays a rapid
change at about -0.1 V, corresponding to a transition between the (1 x 1) and
(1 x 3) states (Fig . 2 of ref. 21).
A more detailed description is given in ref . 21 . These variations are consistent
with the variations in differential capacity as a function of imposed potential
shown in Fig . 5, and are consistent with the interpretation of the frequency
dispersive peak of Fig . 5(b) being due to a "fast" reconstruction of Au(110) .

DISCUSSION AND CONCLUSION

Although surface purity and surface order are paramount concerns for all
studies of gold surfaces, it is clear that experimental conditions are different for
the three techniques described above . The cleanliness required to give meaningful
results in conventional glass cells is not achieved in STM and SXRS cells . It is
difficult to say how much this matters . However, these differences do not preclude
comparisons of the results obtained in the three sets of experiments . The consis-
tency of the results is understandable because SXRS and STM are probably less
sensitive than CVs and C(E) curves to solution purity and cell cleanliness .
With X-ray scattering techniques, the surface lattice constants can be measured
with a precision of one part in 10000 . STM measurements may be done on a large
scale or on a small scale ; individual atomic resolution was obtained for gold faces .
Traditional electrochemical results cannot claim to give any such information on
surface structures directly . In-situ SXRS is an averaging technique (the results are
an average of scattered X-rays from a chosen section of the electrode of about 10 14
atoms) and preferentially probes well-ordered regions, while in-situ STM can
describe a chosen section of the electrode surface, ordered or disordered, with
atomic resolution . (STM images are a two-dimensional representation of a three-
dimensional structure, which could make determination of the superficial structure
difficult .) In contrast, CVs and C(E) curves are a response of the whole area of
the electrode (including the edge which is not sharp down to the atomic level) .
Furthermore, STM and SXRS are directly related to the electrode surface struc-
ture itself while CVs and C(E) curves are sensitive to the electrode surface
256

structure as well as to the solution electrochemical double-layer region .


The STM results and C(E) curves were found to be consistent for Au(100) [17]
and Au(111) [18] in perchloric acid . For Au(110) the "faster" kinetics of the
reconstruction observed by STM (faster than for Au(100) for instance) is in
agreement with observed CVs and C(E) curves of this face . Inspection of Figs . 2-4
suggests that there are intermediate structures existing at this gold face before
reaching the potentials at which the (1 x 1) structure exists . From STM observa-
tions for potentials more positive than 0 .1 V vs . SCE, the (1 X 1) structure would
exist at the (110) gold face, while at negative potentials (for instance -0.3 V) the
surface is reconstructed . The reconstruction would be more complete when the
potential is held at the more negative values imposed in this work . As said above,
C(E) curves obtained in perchloric acid suggest more complicated structural
changes (Fig . 4) . In the first study of STM observations of structural changes of
Au(110) in 0 .1 M perchloric acid [19], small irregular domains of (1 x 2) and
(1 x 3) structures, i .e . a mixture of symmetries, were observed simultaneously at
the Au(110) surface ; a real-time dynamical study of the different steps of the
`deconstruction' or "reconstruction" would be welcome .
In the 1980s, we suspected that the "sigmoid" shape of the contribution of the
diffuse part of the double layer to the C(E) curves (Fig . 2) originated from
definitive faceting of the (110) gold face during flame treatment . From STM
observations, it may be concluded that this is not the case ; a (1 x 1) structure is
always observed at sufficiently positive charges . The symmetry or asymmetry of the
observed CVs, in the double layer region (Figs . 2(a) and 3) is not related to the
purity of the solution (the results are independent of the stirring of the solution) ; it
may be connected to the last treatment of surface before contact with the solution .
Ex-situ Au(110) faces were found to exhibit surface disorder, steps [25] and (111)
microfacets when reconstructed [26] . These facts could explain these two types of
CVs and C(E) curves .
For the case of strong ionic adsorption potassium, SXRS data [21] show only
one transition from the (1 X 3) structure to the (1 x 1) structure . This is also the
case in neutral solutions for other halides (NaF, NaBr, LiCI, CsCI) [21]. Slight
adsorption of the cation could be responsible for the (1 x 3) structure, from the
discussion of ref. 26 . Indeed, slight adsorption of the (NaCI) cation was observed at
the Au(110) face when there was nearly no adsorption of the anion [22] .
For Au(100) and (111), the systematic shifts in potential dependence of the
in-situ X-ray scattered intensity with the nature of the anion [11,20], correspond to
the systematic shifts in C(E) curves on the potential axis (Fig . 35 of ref. 2) . Similar
results are obtained for Au(110) [21] . For the Au(110) face, the changes in
structure of the top layer(s) with imposed potential in 0 .1 M NaCI, observed by
SXRS, at about -0 .1 V vs . SCE correspond to the frequency dispersive peak
observed on the C(E) curves (Fig . 5(b)) . This conclusion may be extended to the
frequency dispersive peak observed for Au(110) in KBr solutions (Fig . 8 of ref. 7) .
Determination of the potential of zero charge (pzc) for Au(110) remains
puzzling . The pzc discussed in refs . 22 and 23 for positive-going sweeps is probably
257

that of a reconstructed surface, although the non-uniform distribution of electronic


charge at reconstructed gold faces makes the pzc determination as ambiguous as
for polycrystalline electrodes [17] . The minimum observed at 0 .4 V (Fig. 4(a)) does
not correspond to a pzc because it does not deepen with dilution of the solution
(Fig . 1 of ref. 23). Up to now, the question "what is the pzc of the (1 x 1) structure
of Au(110)?" has not been answered .
When the potential was held in the negative charge density range in 0 .1 M
perchloric acid, the pzc value (supposed to be that of the reconstructed structure)
was not altered, but the contribution of the diffuse part of the double layer was
more pronounced . In the same way, when the Au(110) electrode in perchloric acid
is "groomed" (i .e . the potential is varied in a small range of very negative charge
densities) the variation in intensity of the scattered X-rays is steeper [21) .
Although for a set of gold faces, with crystallographic orientations on the three
main zones of the unit projected stereographic triangle, the parallelism of varia-
tions in density of broken bonds and the pzc is qualitatively satisfactory [27] ; a
more quantitative relation between these two parameters requires consideration of
the reconstruction of the faces . However, in the case of Au(110) the average
number of broken bonds per surface atom is the same for the (1 x 1), (1 x 2) and
(1 x 3) structures [28] . As said above, the "fast" transition from the reconstructed
structure to the (1 x 1) structure occurs in the range of potential of the contribu-
tion of the diffuse part of the double layer-as observed in Figs . 1-4-and
therefore precise determination of the pzc of each structure possibly existing at the
surface of Au(110) is difficult to achieve .
When for Au(110) the variations in inner layer capacity as a function of charge
density (calculated from the Gouy-Chapman theory) were found to display two
maxima [23]-in contrast to silver faces and Au(210)-this was probably because
there are two (or more) types of superficial structure at the "supposed" low charge
density values examined . These different types of structure could be the origin of
the two maxima observed in Fig. 7 of ref. 23 for inner layer capacity-charge
density curves .

ACKNOWLEDGMENTS

The author benefited from improvements of her text and valuable discussions
and suggestions by M .J . Weaver and B .M . Ocko . The X-ray scattering measure-
ments were supported through the Division of Materials Research, US Depart-
ment of Energy under Contract No. DE-AC02-76CH00016 and were carried out
with B .M. Ocko, J. Wang, G . Helgesen and B . Schardt at Brookhaven National
Laboratories, Upton, NY, USA. The STM measurements were supported by the
National Science Foundation and the Office of Naval Research ; they were success-
fully carried out by X . Gao with M .J . Weaver in Purdue University, West
Lafayette, IN, USA.
258

REFERENCES

1 A. Hamelin, J . Electroanal. Chem ., 142 (1982) 299 .


2 A. Hamelin, in B .E. Conway, R .E. White and J .O .'M . Bockris (Eds.), Modern Aspects of Electro-
chemistry, Vol . 16, Plenum, New York, 1986, Chapter 1, p . 1 .
3 D.M . Kolb and J . Schneider, Electrochim . Acta, 31 (1986) 929 .
4 M.S . Zei, G . Lempfuhl and D .M. Kolb, Surf. Sci, 221 (1989) 23 .
5 R. Michaelis and D .M. Kolb, Surf. Sci . Lett ., 234 (1990) L281 .
6 G .A . Somorjai and M .A. van Hove, Progr . Surf . Sci ., 30 (1989) 201 .
7 A. Hamelin, J . Chim . Phys ., 88 (1991) 1453 .
8 R. Feidenhans'l, Surf. Sci. Rep., 10 (1989) 105 .
9 T .K. Robinson, Phys . Rev . Lett ., 50 (1983) 1145.
E . Vlieg, 1K. Robinson and K. Kern, Surf. Sci., 233 (1990) 248 .
10 M.F . Toney and O .R . Melroy, Surface X-ray scattering, in H .D . Abruna (Ed .), Electrochemical
Interfaces, VCH, Deerfield, FL, 1991, Chapter 2 .
11 B .M . Ocko, J . Wang, A . Davenport and H. Isaacs, Phys . Rev . Lett ., 65 (1990) 1466 .
12 C .F . Quate, Phys . Today, 39 (1986) 26 .
13 P .K . Hansma and J : Tersoff, J . Appl . Phys., 61 (1987) Rl .
14 R. Sonnenfeld, J. Scheneir and P.It Hansma, in R.E. White, J .O.'M . Bockris and B .E. Conway
(Eds .), Modern Aspects of Electrochemistry, Vol . 21, Plenum, New York, 1990, Chapter 1, p . 1 .
15 J . Wiechers, T . Twomey, D .M . Kolb and R .J . Behm, J. Electroanal . Chem., 248 (1988) 451 .
16 R.J . Nichols, O .M . Magnussen, J . Hotlos, T . Twomey, R.J . Behm and D .M. Kolb, J . Electroanal .
Chem., 290 (1990) 21 .
17 X. Gao, A . Hamelin and M .J. Weaver, Phys . Rev. Lett., 67 (1991) 618; J . Electroanal . Chem., 323
(1992) 361 .
18 X. Gao, A. Hamelin and M .J. Weaver, J . Chem . Phys., 95 (1991) 6993 .
19 X. Gao, A . Hamelin and M .J. Weaver, Phys . Rev . B, 44 (1991) 10983 .
20 J . Wang, B.M . Ocko, A .J . Davenport and H .S. Isaacs, submitted to Phys . Rev . B.
21 B .M . Ocko, G. Helgesen, J. Wang, B. Schardt and A . Hamelin, submitted to Phys. Rev . Lett .
22 A. Hamelin and L . Stoicoviciu, J. Electroanal . Chem ., 234 (1987) 93 .
23 A. Hamelin, L . Stoicoviciu and F . Silva, J . Electroanal . Chem ., 236 (1987) 283 .
24 D . Gibbs, B .M. Ocko, D .M . Zehner and S .G.J . Mochrie, Phys. Rev . B, 38 (1988) 7303 .
25 See for instance, D . Wolf, H . Jagodzinski and W. Moritz, Surf. Sci ., 77 (1978) 265 .
Y . Kuk, P .J. Silberman and F.M . Chua, 152 (1988) 449 .
26 T . Gritsch, D, Coulman, R .J . Behm and G . Ertl, Surf. Sci., 257 (1991) 297 .
27 R. de Levie, J . Electroanal . Chem ., 280 (1990) 179 .
28 K . Wardell, in R. Guidelli (Ed.), Electrified Interfaces in Physics, Chemistry and Biology, NATO
ASI, Series C, Vol . 355, Kluwer, Dordrecht, 1991 .

You might also like