You are on page 1of 11

580 Langmuir 1993, 9, 580-590

Effect of Anions the Underpotential Deposition of Cu


on on
kPt(lll) and Pt(100) Surfaces
Nenad Markovic* and P. N. Ross
Materials Sciences Division, Lawrence Berkeley Laboratory, University of California,
Berkeley, California 94720
Received July 27, 1992. In Final Form: November 16, 1992

The effect of chloride and (bi)sulfate anions in the supporting acid electrolyte on the chemistry of
underpotential deposition of Cu on Pt(lll) and Pt(100) single crystal surfaces was studied using a
combination of electrochemical and nonelectrochemical techniques. The presence of these anions above
a threshold concentration caused a splitting in the voltammetry peaks for Cu UPD on both Pt(lll) and
Pt(100), with die second or split-off peak at a lower underpotential. This splitting was attributed to
competition between the Cu adatoms and the adsorbed anions and the increase in thermodynamic driving
force needed to displace the anions from the Pt substrate in order to form a Cu monolayer. The magnitude
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

of the splitting appeared to be proportional to the relative strengths of the anion-Pt bonding, being larger
for chloride than (bi)sulfate on either surface and being larger for the (100) surface than for the (111)
surface for either anion. By the use of ex situ AES and LEED, we determined that in nearly “Cl-free”
supporting electrolyte Cu appeared to be deposited at underpotential in metallic islands (or “patches”)
having the Pt lattice constant (pseudomorphic growth). In the presence of Cl" the Cu was deposited at
Downloaded via CINVESTAV on January 21, 2020 at 23:07:53 (UTC).

underpotentials into a Cu-Cl adlattice. At the Nernst potential, however, in both “Cl-free” and Cl-
containing electrolyte, Cu formed a uniform metallic monolayer having the Pt lattice constant, i.e. a
pseudomorphic monolayer.

Introduction proposed for these effects, although several mechanisms


have been suggested.
The underpotential deposition (UPD) of a metal, defined Schmidt and Wuthrich6 reported that halide ions formed
as the deposition of a metal onto a dissimilar metal a compound with the polycrystalline platinum that
substrate at a potential that is anodic of the Nernst
strongly inhibited the UPD of metal monolayers. On the
potential for bulk deposition, has been a subject of intense other hand Kolb et al.7 and some other authors8’9 argued
theoretical and experimental interest. Breiter et al.1 that halide interacted with the adsorbate more strongly
provide the first evidence that copper atoms adsorbed on than with the substrate and thus influenced the UPD state
platinum in the underpotential region interfere with of the admetal. More recent studies using single crystal
hydrogen adsorption. Since then a number of authors surfaces have demonstrated that concurrent anion ad-
have focused their attention to the interaction between
sorption at the electrode solution interface plays an
the platinum substrate and copper adatoms. Several
important role in the ordering and growth of the UPD
electrochemical techniques have been used for studying metal. In the work by Kolb et al.10 *it was shown that
the equilibrium and dynamic properties of copper dep- anions can have a pronounced influence on the copper
osition. In a rotating ring-disk electrode study, Tindall
underpotential deposition of Pt(lll), e.g. two adsorption/
and Bruckenstein* reported that in a sulfuric acid solution
12

desorption peaks hardly separated in Cl- free solution (0.05


the deposition of two monolayers of copper was necessary M H2SO4) become clearly separated in the presence (~10~5
before bulk deposition was observed. Using the thin layer M) of Cl-. Using a UHV surface analysis apparatus for
method, Schultze3 reported that the nature of adsorbed sample preparation and surface characterization, Andri-
copper on platinum depended on the coverage and that cacos and Ross11 found that copper UPD on clean well-
the copper UPD in sulfuric acid obeyed a Temkin isotherm. ordered platinum surfaces exhibited multiple adsorption/
Important observations on the role of anions on copper desorption peaks from an 0.3 M HF solution containing
UPD phenomena were discussed in the radiotracer work Cl- ion as an impurity. With the same technique Stickney
of Horanyi.4’6 He found that the potential dependence of et al.12 has demonstrated that UPD of copper on platinum
the adsorption of Cl" or HSO4" was modified in the presence
single crystals pretreated with I2 vapor can produce several
of an Cu2+ ion and concluded that the electrosorption of ordered Cu-I overlayers. Specifically, the Pt(lll) (3 x 3)
Cu2+ ions, i.e. Cu UPD, induced the adsorption of acid structure was observed for partially covered surface by
anions. The effect of specific adsorption of anions on UPD
has since been discussed by a number of other research
copper adatoms, and a (10 x 10) structure was observed
at full coverage. Recently in situ scanning tunneling
groups.3"15 However, no consistent interpretation has been
microscopy (STM) data for UPD of Cu on Pt(lll) in
sulfuric acid demonstrate13 that the structure of the first
*
Permanent address: Institut of Electrochemistry, ICTM, Uni- Cu adlayer is (Vs x V3)B30°. It was suggested that the
versity of Belgrade, Belgrade, Yugoslavia.
(1) Breiter, M.; Knorr, C.; Volkl, W. J. Elektrochem. 1955, 59, 681. (10) Kolb, D.; Al Jaaf-Golze, K.; Zei, M. Dechema-Monographien;
(2) Tindall, G.; Bruckenstein, S. Anal. Chem. 1968, 40, 1637. Verlag Chemie: Weinheim, 1986; Vol. 12, p 53.
(3) Schultze, J. Ber.Bunsen-Ges, Phys. Chem. 1970, 74, 705. (11) Andricacos, P.; Ross, P. J. Electroanal. Chem. 1984,167, 301.
(4) Horanyi, G.; Vertes, G. J. Electroanal. Chem. 1973, 45, 63. (12) Stickney, J.; Rosasco, S.; Hubbard, A. J. Electrochem. Soc. 1984,
(5) Horanyi, G. J. Electroanal. Chem. 1974, 55, 45. 131, 260.
(6) Schmidt, E.; Wflthrich, N. J. Electroanal. Chem. 1970, 28, 349. (13) Sashikata, K.; Furuya, N.; Itaya, K. J. Electroanal. Chem. 1991,
(7) Kolb, D.; Przasnyski, M.; Gerischer, H. Elektrokhimiya 1977,13, 316, 361.
700. (14) Leung, L.-W.; Gregg, T.; Goodman, D. Chem. Phys. Lett. 1992,
(8) Schultze, J.; Vetter, K. Electrochim. Acta 1975, 19, 915. 188, 467.
(9) Juttner, K.; Lorenz, W.; Staikov, G.; Budevski, E. Electrochim. (15) Aberdam, D.; Durand, R.; Faure, R.; El Omar, F. Surf. Sci. 1985,
Acta 1978, 23, 741. 162, 782.

0743-7463/93/2409-0580$04.00/0 © 1993 American Chemical Society


Deposition of Cu on Pt Langmuir, Vol. 9, No. 2, 1993 581

appearance of the widely spaced structure of (V3 x


V 3)f?30° is due to the presence of coadsorbed anions like
HSO4' and SO42". Although an ordered structure was
observed, a discrepancy between the coverage of copper
adatoms estimated from STM versus coulometry was
observed; e.g. the charge under the first UPD peak was
nearly twice that expected for an ordered (V3 x V 3)f?30°
structure (160 pC/cm2). The authors attributed the charge
imbalance to the desorption of bi(sulfate) anions during
the UPD process. Leung et al.14 reported a discrepancy
between the coverage of Cu on Pt(lll) in HCIO4 deter-
mined ex situ by AES and that calculated from coulometry
which they attributed to perchlorate anion adsorption.
However, Aberdam et al.15 have reported that the Cu
coverage from the Cu/Pt AES peak ratio correlated well
with the coulometric measurement of the amount of Cu
deposited on Pt(110) in perchloric acid. An interesting
observation about coadsorption of copper with different
halide anions on Pt(lll) was recently discussed by White
and Abruna.16 They found that a partial charge transfer
from adsorbed copper to the coadsorbate occurs in the
following manner S2~ > I- >Br~ > Cl". Further evidence
for charge transfer from the UPD metal to the coadsorbed
anion was provided by X-ray adsorption data.16
The variety of different structures observed for Cu UPD
on Pt surfaces, as well as the dramatic differences in
voltammetry in different electrolytes, suggests there is a
strong effect of the anions present in the electrolyte in
which Cu UPD occurs. The purpose of the present work
was to understand this effect more completely by adding
varying amounts of different anions to either perchloric
or hydrofluoric acid supporting electrolyte and observing
the systematic changes in the voltammetry. In the case
of HF supporting electrolyte, these changes in voltammetry
could be correlated to the changes in the surface structure Figure 1. Cyclic voltammetry of Pt(lll) and Pt(100) single
observed on emersed electrodes with LEED. crystals in different acid electrolytes of the same pH (=1). 50
mV/s.
Experimental Section
annealed Pt(lll) and Pt(100) obtained in three different
The preparation and pretreatment of platinum single crystal acid electrolytes are shown in Figure 1. The general shapes
electrodes were fully described in ref 17. Most of the results
of these curves are well established,18 but the interpretation
reported here were obtained with the flame technique for the
preparation oflargesinglecrystals.18 Results with UHV prepared
of the processes which occur within hydrogen adsorption
surfaces18 are noted specifically. The emersion procedure for ex and oxide formation potential regions is still contro-
situ surface analysis was described in detail in ref 19. Calibration versal.20-22 For our purposes here we are focusing on the
factors for estimating the coverages of Cl and Cu from Cl/Pt, effects of acid anions on the voltammetric features in these
Cl/Cu, and Cu/Pt AES peak ratios were given in ref 17. The two potential regions with both Pt(lll) and Pt(100)
surface coverages (in monolayers (ML)) were based on the surfaces. Figure 1 shows that hydrogen adsorption/
absolute atomic densities of the Pt(lll) and (100) surfaces (1.505 desorption at Pt(lll) in the potential region -0.05 < £ <
X 1016 and 1.303 x 1016 atoms/cm2, respectively) assuming one
0.25 is not affected by the nature of the anions, indicating
adatom per Pt atom at 1 ML.
The solutions were prepared using J.T. Baker reagent grade
that in this potential region on this surface there is no
acids and pyrolitically triply distilled water. All potentials are coupling between hydrogen adsorption and anion de-
given against the reversible hydrogen electrode in the same sorption. The curves for the Pt(100) surface show,
electrolyte (RHE). The physical surface areas for the flame- however, that within the hydrogen region -0.05 < £ < 0.4
annealed Pt(lll) and Pt(100) crystals were 0.568 and 0.572 cm2, V anion adsorption/desorption on this surface is concurrent
respectively. The contact area of the electrolyte drop on the with hydrogen desorption/adsorption. In contrast to
crystal surfaces in the thin layer cell was approximately 0.35- relatively symmetrical peaks for Pt(100) obtained in
0.45 cm2. perchloric and sulfuric acid, that in HF exhibits asym-
The voltammograms were recorded with sweep rates of 50 and metry, which has been attributed to the relatively high
5 mV/s.
concentration (~5 x 10-6 M) of Cl- impurities present in
0.3 M HF.17-18 The different levels of Cl- impurities in
Results
HCIO4 versus HF also have a significant effect on the shape
Voltammetry ofPt(lll) andPt(100) in Different
1. and the position of both the anomalous features on Pt-
Acid Electrolytes. Cyclic voltammograms for flame- (111) as well as the “true” oxide formation peak at ~0.85
V. The sharp oxide peak at ~0.85 V obtained in 0.1 M
(16) White, J.; Abruna, H. J. Phys. Chem. 1990, 94, 894. HCIO4 is shifted anodically in HF due to the higher
(17) Markovic, N.; Ross, P. J. Electroanal. Chem. 1992, 330, 499.
(18) Markovic, N.; Hanson, M.; McDougal, G.; Yeager, E. J. Electroanal.
Chem. 1986,214, 5S5. (20) Markovic, N.; Marinkovic, N.; Adzic, R. J. Electroanal. Chem.
(19) Ross, P.; Wagner, P. In Advances in Electrochemistry and 1988 241 309.
Electrochemical Engineering; Gerischer, H., Tobias, C. W., Eds.; Wiley: (21) Clavilier, J.; Rodea, A.; Achi, K. El. J. Chim. Phys. 1991,88,1291.
New York, 1984; Vol. 13, p 69. (22) Wagner, F.; Ross, P. J. Electroanal. Chem. 1988,280, 301.
582 Langmuir, Vol. 9, No. 2,1993 Markovic and Ross

Figure 2. Cu UPD on Pt(lll) in different supporting electrolytes of pH = 1 containing mM levels of Cu2+ ion. 50 mV/s.
concentration of Cl'.12 The most anodic peak for oxide titatively the charge associated with these two pseudoca-
formation on Pt(lll) is obtained, however, in the elec- pacitive processes. Figure 2 also shows that in an
trolyte containing sulfuric acid anions, probably due to electrolyte containing relatively high (>10“® M) concen-
the strong adsorption of 3-fold coordinated (bi)sulfate on tration of Cl~ impurities, such as the 0.3 M HF,17'18 the Cu
the (111) sites.18 In contrast to this, the oxide formation UPD peaks become nearly symmetric and are split. The
peak on Pt(100) in 0.3 M HF appears at the potential second and more intriguing characteristic is that the
which is ~0.15 V more anodic than the potential for the hydrogen adsorption/desorption pseudocapacitances in the
oxide peaks in H2SO4. The relative peak positions for 0-0.4 V region are altered in an unusual way by Cu UPD.
oxide formation on Pt(100) in chloride versus (bi)sulfate Since hydrogen does not adsorb on bulk copper, we would
electrolyte suggest that chloride is more strongly bound expect that the adsorption of hydrogen would be blocked
on this surface than (bi)sulfate, which is the opposite to on the sites occupied by copper adatoms, as observed in
the relative strengths of bonding to the (111) surface. These the gas-phase adsorption studies.23 Contrary to this
anion effects on Pt(lll) and Pt(100) voltammetry in Cu- expectation, Figures 2 and 3 show that the most significant
free electrolyte have a related role in the voltammetry of effect of increasing amounts of UPD Cu on both the (111)
Cu UPD on these surfaces. and (100) surfaces was a redistribution of charge, rather
2. Voltammetry of Cu UPD—Low Concentrations than a reduction in total charge, in the 0-0.4 V region.
(<0.1 mM). The presentation of the voltammetry results Even with an apparent coverage by Cu of ca. 0.1 ML (at
is divided into two sections according to the concentration 10~5 M), the total charge in this region was more than 90%
of Cu ions in the solution. This division is based on the of the charge in Cu-free electrolyte. This redistribution
important observation that below ca. 0.1 mM Cu ion in of charge appeared to be due to a complex interaction
pH = 1 supporting electrolytes there was still a measurable between adsorbed hydrogen, UPD Cu, and adsorbed
pseudocapacitance for hydrogen adsorption/desorption anions.
even at very slow (5 mV/s) sweep rates. Above this On the (111) surface, the redistribution of charge did
concentration, sufficient Cu was deposited underpoten- not appear to depend on the nature of the anion to the
tially to completely suppress hydrogen adsorption. same extent as on the (100) surface. For the (100) surface,
The effect of addition of relatively small amounts of multiple peaks were eliminated in favor of a central single
copper to different acid electrolytes on the voltammetry peak. In 0.3 M HF, which contains ca. 5 X 10"5 M Cl" as
of Pt(lll) and Pt(100) is shown in Figures 2 and 3. There an impurity,1718'24 the central peak was much sharper on
are two characteristics of the voltammograms that rep- the cathodic sweep them on the anodic sweep. This
resent the effect of Cu UPD at these concentrations. The characteristic of the voltammetry of Pt(100) in 0.3 M HF
first characteristic is that the stripping of Cu occurs within + 5 x 10"6 Cu2+ is similar to the change produced by adding
a narrow potential range while the deposition occurs over HC1 to Cu-free electrolyte.17-18 The asymmetry in peak
a broad potential range without forming a distinct peak. shape was strongly dependent on the Cl~ concentration,
For Cu UPD on Pt(lll) in either 0.1 M HCIO4 or 0.05 M becoming larger at lower concentrations and disappearing
H2SO4, only one stripping peak is obtained. The main at higher concentrations. These observations suggested
difference between these two electrolytes is that the copper
(23) Paffett, M.; Campbell, C.; Taylor, T.; Srinivaaan, S. Surf. Sci.
stripping peak in perchloric acid is superimposed on the 1985,154, 284.
anomalous features making it difficult to separate quan- (24) Ross, P. J. Chim. Phys. 1991, 88,1353.
Deposition of Cu on Pt Langmuir, Vol. 9, No. 2,1993 583

J_I_I_I_U_I_I_I_1__I_I_I_I_L_
0 0.4 0.8 0 0.4 0 0.4 0.8
E/V
Figure 3. Same as Figure 2 with Pt(100).
that the asymmetry in the peak shape in the 0-0.4 V region
can be attributed to desorption of chloride anions during
hydrogen adsorption and diffusion of these anions away
from the electrode surface on the cathodic sweep, such
that on the anodic sweep hydrogen is desorbed without
coupled readsorption of chloride, i.e. the readsorption of
chloride is slowed by diffusion from the bulk electrolyte
and the charge is spread over a wide potential range, but
desorption is rapid and coupled to the charge for hydrogen
adsorption. At the lower Cl' concentrations, the central
peak was also much sharper in the presence of Cu than
in Cu-free electrolyte, suggesting the underpotential
deposition of small amounts of Cu enhances Cl adsorption.
3. Voltammetry of Cu UPD—High Concentrations
(>0.1 mM). Figures 4 and 5 show the voltammetry curves
for Pt(lll) and Pt(100) in 0.1 M HCIO4 with varying
amounts of Cu2'*' ions above 0.1 mM. At these concen-
trations, hydrogen adsorption was completely suppressed
indicating that Cu formed a uniform monolayer (at least)
atUPDpotentialsonbothPt(lll) andPt(100). It should
be noted that the surface coverage (by coulometry) by
copper was still strongly dependent on the scan rate applied
(the surface coverage by copper increased at lower scan
rate) even at these higher concentrations, indicating
deposition of copper is a relatively slow process. The
maximum amount of Cu deposited at underpotentials was
determined by integration of current potential curves
(anodic sweep) in Figures 4 and 5 and is summarized in
Table I. The charge values for Cu on Pt(lll) and Pt(100)
also indicate that for >1 mM Cu2+ essentially a full
monolayer of copper was deposited underpotentially from
HCIO4 electrolyte.
A comparison of Figures 4 and 5 shows that Cu
deposition started at a more (~ 0.1V) anodic potential on
the (100) versus the (111) surface, but otherwise the Figure 4. Cu UPD on Pt(lll) in HCIO4 supporting electrolyte
with different levels of Cu2+ in solution. 5 mV/s.
voltammetry curves for the two surfaces in HCIO4 shared
several characteristics. First, on the cathodic sweep metal relatively broad deposition peaks at all Cu2+ ion concen-
deposition onto the initially bare Pt surface produced trations (with the exception of Cu on Pt(100) from 1 X
584 Langmuir, Vol. 9, No. 2, 1993 Markovic and Ross

E/V

Figure 5. Cu UPD on Pt(100) in same solutions as in Figure 4. Figure 6. Effect of additions of Cl" (as HC1) to 0.1 M HCIO4 on
Table I. Coulometry of Cu UPD in 0.1 M HCIO4
Cu UPD on Pt(lll) and (100). 5 mV/s.

concentration Pt(lll) Pt(100) effect on both the position and the shape of copper UPD
of Cu2+/M Q/mC cm-2 Q/n C cm-2 peaks. For Pt(lll) in the presence of ca. 1 x 10"6 MCI",
1 x 10"6 35 30 the deposition peaks at 0.6 and 0.47 V were changed slightly
1 x 10"5 125 103 and a small shoulder appeared at ca. 0.58 V on the cathodic
1 X 10"4 400 360 side of the stripping peak for the UPD monolayer.
X 10"3 460“ 413“
However, with higher concentrations of Cl" (>5 X10"5 M)
1


The ideal charge for a monolayer of zero valent copper atoms deposition (designated A' and B') as well as stripping peaks
having the Pt lattice constant is 480(111) and 420(100) nC/cm2. (A and B) became sharp and well resolved. On Pt(lll)
two desorption processes at 0.7 and 0.58 V which were not
10”3 M Cu2+), but only a single stripping peak appeared clearly resolved in the solution with<l X 10~6 MCI" became
on the anodic sweep (provided no bulk Cu is deposited). clearly separated in the presence of a higher concentration
Second, when scanning into the region negative of the of chloride anions. It is interesting that on the (111) surface
Cu/Cu2+ Nernst potentials bulk Cu was deposited only Cl" caused the deposition peaks to shift in opposite
when the Cu2+ concentration was higher than ~lx 10~4 directions. The sharp peak at 0.63 V shifted to higher
M. For these concentrations, the shape of the voltam- bonding energy but the shoulder at 0.58 V was shifted to
mogram in the overpotential region was exactly that lower bonding energy. This suggests that Cl" has two
expected for diffusion limited reversible deposition of a different effects on the UPD process. For the initial
metal.25 Third, the sharp anodic stripping peak at 0.6- deposition of Cu onto the Pt(lll) surface, Cl~ appears to
0.7 V shifted anodically with an increase of Cu2+ concen- have a catalytic effect, causing the deposition of sub-
tration. Finally a very sharp peak at ~0.65 V (111) or monolayer Cu to be much more reversible, and also
0.67 V (100) was superimposed on the narrow stripping produces a more strongly bonded state. However, com-
maxima at 0.65 (111) or 0.69 V (100) when the potential pletion of the deposition to form a monolayer is apparently
was swept through the overpotential region. We hypoth- impeded by the presence of Cl~, which is consistent with
esized that this splitting was caused by desorption of the stabilizing role of Cl~ in the submonolayer state.10
specifically adsorbed impurity anions (e.g. Cl") concurrent The effect of Cl" on the UPD of copper on Pt(100) was
with Cu deposition. different than on Pt(lll), probably due to the difference
To confirm that multiple voltammetric peaks arise from in the free enthalpy of Cu and Cl~ adsorption on these two
an effect of anion (Cl~) on the underpotential deposition, surfaces. On Pt(100) no splitting of the Cu UPD peaks
small amounts of HC1 were added to the HCIO4 solution, was immediately apparent. Both the deposition arid the
as shown in Figure 6. The addition of HC1 had a significant stripping peaks were shifted cathodically with an increase
of Cl~ concentration. Another interesting difference was
(25) Andricacos, P.; Ross, P. J. Electrochem. Soc. 1984, 131, 1353, that the UPD peaks on Pt(100) were sharper in nominally
1531. “chloride-free solution” (0.1 M HCIO4) than in the solution
Deposition of Cu on Pt Langmuir, Vol. 9, No. 2, 1993 585

Table II. Coulometry of Cu UPD in 0.1 M HClOi as a


Function of Added HC1,1 mM Cu2+
Pt(lll) Pt(100)
Cl'/M Qa Qb Qr/(MC/cm2) Qa Qb QT/((tC/cm2)
0 460 0 466 390 30 420
1 X 10'6 310 120 430 300 110 410
5 X 10'5 275 140 415 280 130 410

containing added levels of Cl", the opposite of what we


observed with Pt(l 11). The shift in potential for the initial
deposition of Cu was also in opposite directions on the
two surfaces, i.e. increasing Cl~ caused a cathodic shift on
(100) versus an anodic shift on (111). These voltammetric
features on Pt(100) indicate a strong competitive inter-
action between Cl" and Cu throughout the UPD process,
with no evidence of the catalytic effect seen for the initial
deposition of Cu on the (111) surface. The difference
appears to be due to the stronger interaction of Cl" with
the Pt(100) surface. The Cl" adsorption isotherms (from
AES data) we reported in a previous study17 indicated
that at a given Cl' ion concentration the coverage of Cl on
Pt(100) was higher than on Pt(lll) at all potentials,
consistent with the conclusion of stronger interaction. It
should also be noted that the Cl' adsorption on Pt(100)
is at saturation coverage17 at the potential where Cu
deposition begins.
A careful analysis of the coulometry in the UPD potential
region on Pt(100) reveals a dramatic splitting of the
deposition due to the presence of Cl~ that is not imme-
diately obvious from the voltammetry itself. The reason
for this is more evident by examining the charges in Table EN
II. Even at a level of 1 x 10'5 M Cl", the charge under the Figure 7. Comparisonof Cu UPD on Pt(lll) and Pt(100) in
main UPD peak near 0.6 V was reduced by nearly 30%, HCIO4 and HF (which contains Cl" as an impurity) of the same
and the charge to complete the monolayer occurred at a pH showing the similarity when Cl' is added (as HC1) to the
potential that was very close, e.g. a few tens of mV, to the HCIO4. 5 mV/s.
Nernst potential. Thus, on the (100) surface, one also
sees a splitting of the UPD process due to the presence of from the Pt surface, while in HF nearly 0.5 ML remained
Cl~ ion, but the splitting (in potential) is much larger than on the surface. These comparisons indicate that there
it is on the (111) surface and the deposition in the second are qualitative similarities to the effects of these two
peak (B/B0 is distributed over a ~0.2 V potential region. adsorbing anions on Cu UPD but also some quantitative
Andricacos and Ross11 reported essentially identical differences that have both kinetic and thermodynamic
voltammetry in 0.3 M HF, although it was not known at consequences.
the time that the multiple states on both surfaces were 3. LEED/AES Analysis of Emersed Electrodes. As
caused by the presence of Cl" as an impurity in the HF. we discussed in our previous paper,17 emersion experiments
As shown by Figure 7, essentially identical voltammetry from electrolytes containing nonvolatile solutes are com-
was obtained in HCIO4 as in HF when HC1 was added to plicated by the residue left from the adhering film of bulk
the HCIO4 to the level of 5 x 10"5 M. The much larger electrolyte. In our apparatus, the concentration of non-
splitting in potential of the UPD process on the (100) volatile solute must be below 1 mM in order for the residue
surface versus the (111) surface appears to be due to the to have a minimal effect on ex situ surface analysis.
stronger adsorption of Cl either on the bare Pt(100) surface Emersion from supporting electrolytes like 0.05 M H2SO4
and/or on a Pt(100) surface with a partial monolayer of or 0.1 M HCIO4 produces the undesirable complication of
Cu. acid hydrate residues26 which we preferred to avoid for
The addition of sulfuric acid to perchloric acid produced the purposes of this study. Thus our emersion experiments
qualitatively similar effects on the voltammetry as the were done from HF electrolyte, containing either HC1 as
addition of HC1 did, but much larger quantities of the an impurity or with HC1 added, since both are volatile
(bi)sulfate anion were needed to produce these effects. As solutes. In previous studies of Cl adsorption on Pt in this
shown by comparison of the voltammetry curves for Cu laboratory,24 AES analyses of electrodes emersed from 0.3
UPD on Pt(lll) in Figure 8, the presence of (bi)sulfate M HF, containing HC1 as an impurity, were used to obtain
anion also appears to catalyze the Cu deposition process, Cl adsorption isotherms. To extend that study and make
producing symmetric anodic and cathodic current-po- use of that data in a meaningful way in this study, we
tential curves, and to split the single stripping peak in conducted a series of AES analyses of electrodes emersed
HCIO4. These comparative effects of (bi)sulfate ion from Cu containing HF electrolyte at various concentra-
appeared at a level of ca. 10 mM H2SO4, more than 2 tions of Cu2+ and HF and at various emersion potentials.
orders of magnitude higher than the concentration of Cl' The objectives of these analyses were (i) to establish that
needed to produce similar effects. Another interesting the presence of Cu2+ in solution does not effect the coverage
comparative feature between (bi)sulfate and chloride is of Cl on Pt at potentials above ca. 0.7 V (where there is
that the entire UPD process took place in a much narrower no UPD Cu), (ii) to measure the maximum concentration
potential region, so that on the anodic sweep at 0.62 V in
sulfuric acid essentially all (>90 %) of the Cu was stripped (26) Zei, M.; Weick, D. Surf. Set. 1989, 208, 425.
586 Langmuir, Vol. 9, No. 2, 1993 Markovic and Ross

E/V
l i . I I _I_i_
200 400 600 eoo 1000
Electron Energy (eV)

Figure 9. Voltammogram for Cu UPD on Pt(lll) in “nearly


Cl-free" electrolyte showing the potential of emersion (insert)
and the corresponding LEED pattern (134 eV) AES spectrum
for emersed electrode.
Table III. Summary of AES Peak Ratios and
Corresponding Coverages for Emersed Pt Single Crystal
Electrodes, 0.3 M HF + mM Cu2+ 1

AES ratio coverage (ML)6


potential0
E/V (V) Cl/Pt Cu/Pt Cl/Cu Cl Cu
Figure 8. Comparison of Cu UPD on Pt(lll) at high Cu2+ ion Pt(lll) 0.35 3.8 1.4 2.7 0.25 (±0.1) 0.50 (±0.1)
concentration in different supporting electrolytes at the same 0.1 7.3 2.6 2.8 0.45 1.0
pH (=1). Pt(100) 0.35 4.5 1.6 2.9 0.30 0.55
0 7.5 2.5 3.0 0.50 1.0

of Cu2+ in solution that produced a Cu-free surface emersed 0


Versus Cu/Cu2+ Nernst potential. 6 Calculated as if only Cl or
at potentials above ca. 0.7 V, and (iii) to determine the Cu wereon the Pt surface (see ref 17).
maximum concentration of HF supporting electrolyte such
that emersion of Pt at 0.7 V produced no Cl AES signal. at potentials near the Nernst potential for Cu/Cu2+ and
The first objective was easily confirmed, as expected. The at a potential between the two voltammetry peaks. The
maximum Cu2+ concentration that would produce the coverages were calculated from the AES ratios using
correct “blank" result (emersion above 0.7 V without an calibration factors determined independently17 by ad-
AES signal for Cu) was ca. 1 mM, just high enough to do sorption of CL and by the vapor-phase deposition of Cu.
emersions from the electrolyte in Figure 7. However, the No attempt was made to correct the AES signals for the
dilution of HF needed to produce a truly Cl-free emersed effect one adsorbate had on the Auger electron emission
electrode was not practical, as it would require a re- from the other. The results are summarized in Table III.
examination of Cu UPD in the other supporting electro- In the case of Cu, the AES coverages indicated the
lytes in the unusual pH range of 3-4. A dilution to 0.1 M formation of a monolayer at the Nernst potential and
HF appeared to be a practical compromise, producing approximately a half-monolayer at the potential in the
“nearly Cl-free" emersed surfaces. middle of the two UPD peaks in Figure 7. The coverage
The LEED/AES analysis of Pt(lll) emersed from 0.1 by Cl increased monotonically with increasing Cu coverage.
M HF containing 5 X 10~5 M Cu2+ is shown in Figure 9, The simplest interpretation of the coverages for the
electrodes emersed near the Nernst potential is that Cl is
along with the corresponding voltammetry prior to emer-
on top of the Cu monolayer. It that were the case, the Cl
sion at 0.2 V. Coulometry indicates a total charge of UPD
Cu of £200 jiC/cm2, or a coverage of ca. 0.4 ML, while the coverages are in excellent agreement with the saturation
AES Cu/Pt peak ratio suggests a coverage of 0.2-0.3 ML. coverages of Cl adsorbed on Cu(lll)28* and CuflOO).2811
While a Cl AES peak at 181 eV is clearly seen in Figure To determine more about the structural arrangement of
the Cu and Cl atoms on the surface as a function of
9, the Cl coverage was actually quite small, less than 0.1
ML (Cl has the highest AES sensitivity factor of any potential, LEED analyses were done on these same
emersed electrodes.
element!). The LEED pattern was (1X 1), indicating that
from nearly Cl-free solution Cu was deposited as metallic The results for the AES/LEED experiments of the Pt-
islands (or “patches") having the Pt lattice constant (111) electrode related to the voltammetry in Figure 7 are
shown in Figures 10 and 11. The electrode was emersed
(pseudomorphic growth), as occurs in the vapor-phase on the cathodic sweep about 0.1 V positive to the Nernst
deposition of Cu on Pt(lll).27
Other emersion experiments were done in 0.3 M HF + potential for bulk deposition, where according to the AES
mM Cu2+ in order to examine the coverages of Cu and Cl peak ratio there was the formation of a metallic Cu
that may be associated with the multiple peaks seen in the monolayer. The LEED pattern is shown in detail in Figure
11. The pattern has three sets of spots, the fundamental
voltammetry of Figure 7. AES measurements were made
(28) (a) Goddard, P.; Lambert, R. Surf. Sci. 1977,67,180. (b) Weatphal,
(27) Yeates, R.; Somorjai, G. Surf. Sci. 1983, 134, 729. D.; Goldman, A. Surf. Sci. 1983, 131, 92.
Deposition of Cu on Pt Langmuir, Vol. 9, No. 2,1993 587

13a, was a reasonably sharp c(2 X 2), but with a relatively


high background intensity. We attribute this background
intensity to disordered coadsorbed Cl, and the c(2 x 2) to
the UPD Cu. An ideal c(2 X 2)-Cu lattice would have a
coverage of 0.5. Flashing the crystal in UHV to 900 K and
repeating the AES analysis after cooling showed (Figure
12b) that the Cl was completely desorbed, and the Cu/Pt
AES ratio decreased from 1.55 to 1.25, indicating the Cu
coverage decreased to 0.3-0.4 ML. The LEED pattern
after flashing, shown in Figure 13b, was p(2 X 2), which
corresponds to an ideal coverage of 0.25 ML. It appears
that in flashing the crystal to 900 K, not only is the Cl
desorbed, but some Cu is also lost from the surface, either
by diffusion into the bulk23’27 or evaporation (possibly as
CuCl). The p(2 X 2) probably represents the formation
of a surface alloy, as observed by Yeates and Somorjai.27
Emersion at the Nernst potential also produced a c(2 X
Figure 10. AES spectrum for Pt(100) emersed on the cathodic 2) LEED pattern, and a Cu/Pt AES ratio of 2.5, consistent
sweep (insert). Conditions for Cu UPD were the same as in Figure with the coulometric indication of a full monolayer of Cu
7. (Table III). The Cl/Cu AES peak ratio was essentially
the same as for the partial monolayer. In this case, flashing
beams for (lll)-l x 1, superlattice beams as triplets near the emersed electrode to 900 K desorbed the Cl and
the (V3 x V3)i230 positions, and a set of weaker triplets reduced the Cu/Pt AES peak ratio only slightly (5 %), but
which are double diffraction spots. We interpret this changed the LEED pattern from c(2 x 2) to (1 x 1). Thus,
pattern based on the physical model of a pseudomorphic we interpret the original (100)-c(2 x 2) LEED pattern
Cu monolayer having the Pt(lll)-1 X 1 lattice. On the from the electrode emersed at the Nernst potential as a
top of this Cu monolayer there is an adsorbed layer of Cl (100)-(1 X 1) pseudomorphic monolayer of Cu with an
which is associated with the LEED spots near the V3 adsorbed layer of Cl in a (100)-c(2 x 2)-Cl adlattice, i.e.
positions. The Cl superlattice is more easily visualized by the analogous structure to that observed on the Pt(lll)
surface emersed at the Nernst potential.
referencing it to the (V3 X V3)f?30 unit mesh, as done
in Figure lib in reciprocal space and 11c in real space. It is not possible to determine definitively the real space
The Cl superlattice is compressed by ca. 20% relative to structure of an adlayer based solely on the symmetry of
the LEED pattern. There are other (than the ones
the V3 structure and is rhombic. As shown in Figure 11c,
presented above) real space structures that could have
the nearest coincident cell with the substrate is (V3 x 15) the same symmetries as the LEED patterns in Figures 11
rect, and there are three domains of these coincident cells and 13, but we were able to eliminate these other structures
each rotated by 120° with respect to the other. The Cl-Cl based on additional experiments which are presented in
near-neighbor distance is 3.83 A, and the theoretical Cl detail elsewhere.30* Briefly summarizing, these experi-
coverage is 7/15 or 0.47. The AES peak ratio is consistent ments used a combination of AES chemical shifts and
with this theoretical coverage. The Cl superlattice prim- thermal desorption spectroscopy to eliminate adlayer
itive unit mesh is actually nearly identical to that reported structures derived from the CuCl (zinc blend) lattice in
by Goddard and Lambert28* for the high coverage state of favor of the Cl-adsorbed on Cu monolayers structures just
Cl on Cu(lll) in vacuum. The principal difference described.
between the two structures is the relationship to the
substrate, where in the case of the pseudomorphic Cu Discussion
monolayer the substrate unit mesh is expanded by ca. 8 %
relative to the bulk Cu unit mesh, producing a different The cyclic voltammetry of Pt single crystals we have
coincidence relation with the Cl adlayer. In both cases, reported here clearly indicate the presence of specifically
the Cl-Cl bond distance is nearly the same as it is in the adsorbed anions like Cl" and HSO4' (SO42') in the
CuCl zincblend structure.29 Experiments to determine supporting electrolyte has a dramatic effect on the
the structure of the Cu adlayer after just the first stage underpotential deposition of Cu. The sensitivity is
of deposition, by emersion at 0.35 V positive of the Nernst especially great for Cl', where concentrations as small as
potential, were not as successful. The LEED patterns 10'5 M have a pronounced effect. The complexity of the
were poorly ordered with a high diffuse background voltammetry for Cu UPD in the presence of these anions
intensity and some poorly structured intensity around the points to complex interactions between Cu and the
adsorbed anions which are difficult to unravel without
(\/3 X V3)JZ30 positions of reciprocal space. However, additional information, e.g. coverages of the different
the patterns did appear to be more indicative of two
superstructures corresponding to coadsorbed Cu and Cl, species as a function of potential.
as opposed to the Cu island growth observed with nearly There were two significant pieces of information pro-
Cl-free electrolyte (Figure 9). vided by the AES/LEED results that help to resolve the
The results of LEED/ AES analysis of Pt(100) emersed complexity of the voltammetry. The first is that in a
electrodes are shown in Figures 12 and 13. For emersion supporting electrolyte in which the concentration of
at 0.35 V positive of the Nernst potential, where the total specifically adsorbing anions is below a critical level, the
charge indicates the coverage of Cu was 0.67 ML, the Cu/
growth of the Cu monolayer on Pt by electrodeposition is
Pt AES peak ratio suggests the coverage was somewhat very similar to the growth during vapor-phase deposition
lower, 0.5-0.6 ML. The LEED pattern, shown in Figure
(30) (a) Markovic, N.; Ross, P. Submitted to J. Vac. Sci. Technol. A.
(b) Rosa, P. In Structure of the Electrified Interface, Volume II of
(29) Wyckoff, R. Crystal Structures, 2nd ed.; Interscience: New York Frontiers in Electrochemistry; Lipkowski, J., Ross, P., Eds.; VCH
and London, 1963. Publishers, Inc.: New York and Wienheim, in press.
588 Langmuir, Vol. 9, No, 2, 1993 Markovic and Ross

(c)
Figure 11. (a) LEED pattern (68 eV) for the emersed Pt(lll) electrode of Figure 10. (b) Schematic of the LEED pattern showing
the relation of hypothetical (v3 x v3)/?30 unit mesh (dashed) to the observed "rhombic” unit mesh, (c) Superposition of the
real-space rhombic unit cell onto the (1X1) substrate lattice showing the (\/3 X 15) rect. coincidence unit cell; a (V 3 x V3)fi30 unit
cell (dashed) is shown for comparison.

is formed with Cu atoms having the Pt lattice constant


and with Cl adsorbed on the Cu layer to saturation.
The structures we found for the Cl layers on these Cu
monolayers are also closely related to structures reported
for Cl adsorbed on Cu(lll) and Cu(100) surfaces in vacuum
at saturation coverage. In the case of the Cu monolayer
(100) surface, the Cl adlayer has the same (100)-c(2 X 2)
symmetry as Cl on Cu(100) at saturation in the gas-phase,
but because the Cu monolayer is ca. 8% expanded versus
the bulk Cu crystal lattice, the Cl-Cl distance is slightly
larger. Ehlers et al.32 reported a structure for Cl on a
Cu(100) electrode emersed from 0.1 mM HC1 that is
essentially identical to the (100)-c(2 X 2)-Cl structure
observed in vacuum. They also reported that the Cl
adlayer structure was insensitive to the emersion potential.
From these observations combined with our own, it appears
that at least for the (100) surface the pseudomorphic Cu
monolayer interacts with Cl in the same way as the hulk
Cu(100) surface. With respect to the (111) surface, that
conclusion cannot be reached as simply. The structure
Figure 12. AES spectrum for Pt(100) emersed on the anodic we observed here is very different from that reported by
sweep (insert). Cu UPD conditions were the same as those given
in Figure 7. Stickney and Ehlers33 for Cu(lll) emersed from 0.1 mM
HC1 but is essentially the same structure as that reported
by Goddard and Lambert28® for Cl on Cu(l 11) at saturation
on Pt at room temperature in UHV.27 In this process, Cu in vacuum. Stickney and Ehlers reported the formation
adatoms sit in the bulk lattice positions that Pt atoms of a Cu-Cl bilayer having a structure closely related to the
would occupy in the next atomic layer, a growth process (111) plane of the CuCl zincblend lattice. Although both
termed pseudomorphic. In the submonolayer state, the papers came from the same laboratory, there was no
Cu atoms are nucleated into islands, and the monolayer discussion in the later paper on Cu(100) of the difference
forms by expansion and coalescence of the islands. The in the nature of the structure between the (111) and (100)
second is that for deposition in the presence of CP in faces, e.g. a CuCl-like layer formed on Cu(100) only when
solution above a critical level, there is an attractive
interaction between Cu and Cl such that deposition of Cu (31) Durand, R.; Faure, R.; Aberdam, D.; Traore, S. Electrochim. Acta
1989, 34,1653. Stuve, E.; Borup, R.; Sauer, D. In Proc. Symp. on Appl.
draws Cl to the surface. This interaction also leads to a Surf. Anal, to Env./Mat. Inter., Baer, D., Clayton, C., Davis, G., Eds.;
submonolayer structure in which Cu and adsorbed Cl The Electrochemical Society; Pennington, NJ, 1991; PV-91-7.
(32) Ehlers, C.; Villegas, I.; Stickney, J. J, Electroanal. Chem. 1990,
coexist on thePt surface (most probably in a superlattice).
284, 403.
Then at the Nernst potential, an epitaxial Cu monolayer (33) Stickney, J.; Ehlers, C. J. Vac. Set. Technol. 1989, AT, 1801.
Deposition of Cu on Pt Langmuir, Vol. 9, No. 2, 1993 589

basis of an explanation or rationalization for most of the


Pt (100) W features of the Cu UPD voltammetry. Let us consider
each of the characteristic features in the voltammetry
associated with the presence of adsorbing anions and
consider how that feature may be related to the coverages
and structures just described.
(1) The deposition of a small amount of Cu from
supporting electrolyte-containing chloride or sulfate anions
caused a redistribution in the states of adsorbed hydrogen
similar to the redistribution caused by increasing the
concentration of these anions to Cu-free electrolyte.
The explanation for this characteristic was presented
in detail previously17 and is summarized here for com-
pleteness. As revealed by the AES coverages, the initial
deposition of Cu onto Pt from these electrolytes induces
anion adsorption onto the Pt surface in the vicinity of the
Cu adatom. This enhanced anion adsorption around Cu
adatoms we have attributed to a lowering of the local work
function18 at Pt atoms neighboring the Cu. In the case of
HC104 supporting electrolyte, the similarity of the effect
of Cu in all three electrolytes in this work (Figure 2)
suggests that the anions adsorbing from the HC104 solution
were the (bi)sulfate and chloride anions present as
impurities in the concentrated perchloric acid used to
prepare the solution.
(2) For the Pt(lll) surface, the addition of Cl" to
perchloric acid supporting electrolyte enhanced the re-
versibility of the UPD process and split the UPD process
into two distinct stages, separated in potential by ca. 0.12
V.
The catalytic effect of chloride can be rationalized by
the attractive interaction between Cu and adsorbed Cl
indicated by the AES data. The first (more anodic) UPD
peak appears to correspond to the formation of a Cu-Cl
adlayer having a more favorable thermodynamic potential
than Cu adatoms alone and thus lowering the activation
energy for the initial deposition. At the same time, this
stabilization by coadsorbed Cl would require an increase
in the thermodynamic driving force to decompose this
adlayer and form the Cu monolayer. The second (more
cathodic) UPD peak thus would correspond to displace-
ment of Cl from the Pt sites and replacement by Cu to
form the pseudomorphic Cu monolayer. Cl is adsorbed
onto the Cu monolayer to saturation.
(3) For the Pt(100) surface, the addition of Cl" also split
the UPD process into (at least) two stages, but the second
(more cathodic stage) was distributed over a 0.2-V potential
region and the monolayer was not formed until the
--1__1-1- -1--.-i-1-L 1
potential was very close to the Nernst potential. The
0 200 400 600 800 1000 reversibility was not enhanced significantly, and there was
a distinct cathodic shift in the potential for the onset of
Electron energy (eV) Cu deposition.
Figure LEED pattern (62 eV) and AES spectrum for the
13. (a) The AES derived Cl adsorption isotherms for the (100)
emersed Pt(100) electrode of Figure 12. (b) LEED pattern and and (111) Pt surfaces are very different and indicative of
AES spectrum after flashing (15 K/s) to 900 K.
a much stronger Pt-Cl bonding on the (100) versus the
(111) surface. As on the (111) surface, the first deposition
polarized into the potential region for Cu dissolution, peak for Pt(100) appears to correspond to the formation
whereas a CuCl-like layer formed on (111) at all potentials. of a Cu-Cl adlayer, perhaps with the Cu adatoms in a c(2
The structure we observed for Cl on the emersed (lll)-l X 2) adlattice and Cl randomly occupying the vacant Pt
X1 Cu monolayer was essentially the same as that observed sites. The thermodynamic stability of a Cu-Cl adlayer is
for the saturation coverage of Cl on Cu(lll) in the gas- likely to depend on a delicate balance of chemical
phase, having the same Cl adlattice primitive unit mesh interactions between the Pt surface, Cu adatoms, and
(rhombic) and Cl-Cl near neighbor distances (3.86 A), but adsorbed Cl. The stronger binding of Cl to Pt(100) may
because of the different lattice spacing of the monolayer reduce the stability of the Cu-Cl adlayer on this surface
there is a different coincidence lattice versus the bulk relative to the comparable layer on the (111) surface,
crystal. producing the observed cathodic shift for the first stage
The independently determined coverages and structures of deposition. At the same time, the stronger Pt-Cl
of Cu and Cl as a function of potential can be used as the interaction on the (100) surface would require a larger
590 Langmuir, Vol. 9, No. 2,1993 Markovic and Ross

shift in thermodynamic driving force to displace Cl from be more similar to Cu UPD on Pt in Cl~ solutions than to
the Pt sites, accounting for the much larger splitting in Cu UPD on iodine pretreated Pt.
potential between the two stages on the (100) surface than
on the (111) surface. The random structure of the adsorbed Conclusions
Cl following the first stage of deposition may explain why
the final stage of deposition to form the monolayer was (1) The presence of chloride and (bi)sulfate anions in
distributed over a relatively broad potential region. a supporting acid electrolyte had a significant effect on
This study has shown dramatic effects of the supporting the chemistry of underpotential deposition of Cu on Pt.
electrolyte anion on the chemistry of Cu UPD on Pt, with The effects were observed above a threshold concentration
the effects being largest for Cl~. It is of interest to compare of these anions in either HF or HCIO4 supporting
the present results for Cu UPD on Pt single crystals with electrolyte, ca. 1 x 10-6 M for chloride and 1 x 10~3 for
earlier studies by Hubbard and co-workers12 on the effect (bi)sulfate.
of other halide anions, e.g. I- and Br-, on Cu and Ag UPD (2) The presence of these anions above the threshold
on Pt(lll). In the case of T, the iodine was not present concentration caused a splitting in the voltammetry peaks
as an anion in the electrolyte12 but was pre-adsorbed from for Cu UPD on both Pt(lll) and Pt(100), with the second
the gas-phase prior to immersion. It is known from both or split-off peak at a lower underpotential. We attribute
ex situ LEED12 and in situ STM34 that upon immersion this splitting to competition between the Cu adatoms and
at a potential anodic to Cu UPD the Pt(lll) surface is the adsorbed anions and the increase in thermodynamic
covered by a compact iodine adlayer having the coincidence driving force needed to displace the anions from the Pt
lattice (v 7 X V7)Iil9.1. In our case with trace quantities substrate in order to form a Cu monolayer.
of Cl' in HF, ex situ LEED shows that at the same potential (3) The magnitude of the splitting appeared to be
the Pt surface has only a very low coverage by halide atoms, proportional to the relative strengths of the anion-Pt
e.g. <0.1 ML. Thus, the initial condition of the Pt surface bonding, being larger for chloride them (bi)sulfate on either
for the initial deposition of Cu is very different in the two surface and being larger for the (100) surface than for the
systems. In the presence of iodine, the initial deposition (111) surface for either anion.
requires anion displacement, whereas in our case the initial (4) By the use of ex situ AES and LEED, we determined
Cu coexists with the Cl already present on the Pt surface, that chloride had an effect on the submonolayer Cu
and there is some evidence that more Cl is actually
structure, but not on the Cu monolayer. In nearly “Cl-
adsorbed onto Pt atoms neighboring Cu adatoms.17 free” supporting electrolyte, Cu appeared to be deposited
Another major difference between the two systems is the as metallic islands (or upatches”) having the Pt lattice
change in halide/Pt ratio as a function of potential, i.e. as constant (pseudomorphic growth), but in the presence of
Cu is deposited. In the case of iodine, the I/Pt ratio Cl~ the Cu was deposited into a Cu-Cl adlattice. At the
remained constant, whereas in our case the Cl/Pt ratio Nerast potential, however, Cu formed a uniform metallic
increased monotonically with Cu coverage; i.e. Cl is drawn
to the surface by Cu deposition. One aspect of the Cu monolayer having the Pt lattice constant (pseudomorphic
UPD process that is very similar in both systems is the monolayer) in both “Cl-free” and Cl-containing electrolyte.
halide-halide geometry, where the I-I distance and the
Cl-Cl distances are very close to comparable distances in Acknowledgment. The authors acknowledge the
the Cul and CuCl (both compounds have the zincblend valuable assistance of our colleague Michel Van Hove in
the interpretation of the LEED patterns. This work was
crystal structure) bulk crystals. Although Cu UPD on Pt
from solutions containing Br has not been studied, analogy supported by the Assistant Secretary for Conservation
with Ag surfaces would suggest that the chemistry would and Renewable Energy, Office of Transportation Tech-
nologies, Electric and Hybrid Propulsion Division of the
(34) Yan, S.-L.; Vitus, C.; Schardt, B. J. Am. Chem. Soc. 1990,112, U.S. Department of Energy under Contract No. DE-AC03-
3677. 76SF00098.

You might also like