Anda di halaman 1dari 12

PERGAMON

Electrochimica Acta 44 (1999) 18051816

Platinum electrodeposition on graphite: electrochemical


study and STM imaging
F. Gloaguen b, *, J.M. Leger b, C. Lamy b, A. Marmann a, U. Stimming a, 1,
R. Vogel a
a

Institut fur Energieverfahrenstechnik (IEV), Forschungszentrum Julich GmbH (KFA), 52425 Julich, Germany
Equipe Electrocatalyse, Universite de Poitiers, UMR CNRS 6503, 40 Avenue du Recteur Pineau, 86022 Poitiers, France

Received 8 May 1998; received in revised form 15 July 1998

Abstract
The electrodeposition of Pt on a thermally oxidized HOPG surface was performed by single potential
perturbation in dilute chloroplatinic acid solutions. The quantitative analysis of the current versus time transient
responses, on the time scale of seconds, indicated a low saturation density of nuclei of 2  106 cm2. The
characterization by STM revealed a heterogeneous distribution of deposited Pt on the substrate surface. Most of the
deposits were composed of agglomerates of spherical nanoparticles. Local particle densities exceeding 1010 cm2
were observed, which is several orders of magnitude higher than the saturation coverage of nuclei evaluated from
the analysis of the transient at t>0.3 s. From the electrochemical and the structural investigations, it appears that
Pt electrodeposition on graphite involves several faradaic steps on the time scale. Firstly, at t < 0.3 s, a large
number of nanosized clusters are possibly formed. In the course of the electrodeposition process, these very small
clusters may be mobile and assemble in an accidental distribution of Pt agglomerates. Secondly, at t>0.3 s, it is
very likely that the more rapid growth of a small number (2  106 cm2) of Pt particles gives a transient response,
which is adequately described by the model of Scharifker and Hills for a diusion controlled growth. As a result, it
seems that cluster diusion contributes signicantly to the structural evolution of platinum electrodeposits on
graphite. # 1999 Elsevier Science Ltd. All rights reserved.
Keywords: Pt electrodeposition; Oxidized HOPG surface; Nucleation process; Scanning tunneling microscopy

1. Introduction
Carbon-supported platinum particles are widely used
as electrocatalysts, e.g. for fuel oxidation and oxygen
reduction in low-temperature fuel cells. Considerable
eorts were made in order to correlate the catalytic activity of such electrodes with their morphology, usually
by characterizing the catalyst particles within rather
complex composite electrodes. The conclusions derived

* Corresponding author. Fax: +33-549-453-580; E-mail:


gloaguen@cri.univ-poitiers.fr
1
Present address: TU Munchen, Physik Departement E19,
85748 Garching, Germany.

from such investigations remained inconsistent [1, 2],


which may be to some extent due to the structural
complexity of the investigated systems. Recent
advances in the understanding of the catalytic properties of nm-scale particles were derived from investigations of model electrodes with reduced structural
complexity [35]. These advances showed clearly that
the degree of structural denition of model electrodes
is a crucial point regarding the correlation of structure
and catalytic properties.
It has been shown that well-dened distributions of
nm-scale silver clusters can be prepared on graphite
substrates by electrochemical deposition [6, 7]. It has
been recently pointed out [8] that narrow particle size
distribution on graphite substrates can be obtained, by

0013-4686/99/$ - see front matter # 1999 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 1 3 - 4 6 8 6 ( 9 8 ) 0 0 3 3 2 - 6

1806

F. Gloaguen et al. / Electrochimica Acta 44 (1998) 18051816

a short time potential pulse (t < 100 ms); the metal


loading of the resulting electrodes was however very
low: equivalent to less than 0.1 monolayer of Pt. We
investigated the possibilities to prepare electrocatalytic
model electrodes, with a metal loading equivalent to
1.0 monolayer of Pt, at least, by potential pulse over
larger times (t>1.0 s). The surface morphology was
characterized with scanning tunneling microscopy
(STM) and cyclic voltammetry (CV), i.e. by hydrogen
adsorption coulometry.
The inuence of deposition parameters on the surface morphology can, in principle, be derived from
monitoring the metal deposition reaction `in-situ' with
an electrochemical STM. An inuence of the STM tip
on the local kinetics of galvanic reactions has, however, to be anticipated [912]. Since for platinum as
well as for palladium deposition on graphite substrates
the tip inuence has been found to be strongly
inhibitive [1315], this route was not considered for the
present investigation. Instead, platinum deposits were
prepared in a conventional electrochemical cell and the
deposit morphology was investigated afterwards under
`ex-situ' conditions.
The electrodeposition of Pt on highly oriented pyrolytic graphite (HOPG) at small overpotentials has been
studied in detail [16]. The authors concluded that the
formation of platinum nuclei occurs at defects on the
graphite surface. Similar results were obtained for Pd
deposition on HOPG [15]. A comparative structural
ex-situ characterization of Ag-clusters on HOPG with
STM and non-contact atomic force microscopy (NCAFM) [6] indicated, however, that electrodeposited
clusters on the HOPG basal plane may escape detection with an STM and can only be characterized with
NC-AFM. Nevertheless, a preferential deposition at
substrate defects was also supported by the latter
technique [7].
It was reported that the defect density of HOPG
surfaces can be increased by thermal oxidation [17]
and such defective HOPG surfaces were used as substrates for the investigation of lead electrodeposition [10]. Similar graphite surfaces with defect
densities about 2 orders of magnitude higher than that
of a freshly cleaved HOPG surface were used as substrates for platinum electrodeposition at a high overpotential. The prepared deposits were characterized
with electrochemical techniques and scanning tunneling
microscopy.

2. Experimental
Square shaped HOPG samples of 1 cm2 geometric
area were used as substrates. The surfaces were prepared by cleaving with an adhesive tape. Prior to ex-

periments, the freshly cleaved surface was treated with


a Bunsen ame for 2 s.
HClO4 (Suprapur Merck) and H2PtCl6 (Alfa) were
used as received. The solutions were prepared with
ultra pure water (Milli-Q system). Electrochemical
equipment included a potentiostat (PAR 362, EG&G),
a waveform generator (PAR 175, EG&G), a numerical
oscilloscope (Nicolet 2090-III) and an XY recorder
(HP).
The electrochemical experiments were carried out at
room temperature in a standard three-compartment
glass cell. The contact between the graphite working
electrode (WE) and the electrolyte was performed by
the hanging meniscus technique. The counter electrode
(CE) was a Pt wire. The CE and WE compartments
were separated by a glass frit. A reversible hydrogen
electrode (RHE), consisting of a Pt wire immersed in
H2-saturated 0.1 M HClO4, was connected to the WE
compartment by a Luggin capillary. All potentials are
quoted to the RHE scale.
Before each deposition experiment, several voltammograms (0.05 to 1.2 V at 20 mV s1) were recorded
in Ar purged 0.1 M HClO4, both to ensure that the
HOPG surface is free of Pt and to control the reproducibility of the surface area of the electrode. The platinum electrodeposition was performed by a potential
step in Ar purged and unstirred 0.1 M HClO4+x mM
H2PtCl6 (x = 0.2, 2, 10). The potential perturbation
(0.75 to 0.1 V) was applied immediately (i.e. at most
30 s after that the graphite surface was exposed to the
platinum plating solution) or after over-night contact
with H2PtCl6 solution (the so-called pre-impregnation
process). It has been recently pointed out [8], that
platinum particles are `spontaneously' formed at
freshly cleaved HOPG surface exposed to solutions
containing PtCl2
6 ions; electroless deposition of Pt is
however dramatically reduced on an oxidized graphite
surface [8], as used in this work. The amount of deposited platinum, i.e. the loading W (mg cm2), was estimated from the integrated charge measured on the It
transient response. After deposition, the electrodes
were removed from the deposition solution, thoroughly
rinsed with ultra pure water and transferred to a similar cell containing pure 0.1 M HClO4. The real surface
area, Ar (cm2), of deposited platinum was estimated
using the H adsorptiondesorption coulometry. All
current and charge densities are quoted in terms of mA
cm2 and mC cm2 of apparent geometric area, Ag, of
the substrate electrode.
The rinsed sample was nally transferred to the
STM chamber. The structural characterization was
performed either with a Burleigh Instructional STM in
air, or with a modied Delta Phi STM in 0.1 M
HClO4.

F. Gloaguen et al. / Electrochimica Acta 44 (1998) 18051816

3. Results
3.1. Substrate characterization
STM images of dierently pretreated HOPG surfaces are presented in Fig. 1. Fig. 1(a) represents the
freshly cleaved surface. Typically 1 monoatomic step is
found per m2. Fig. 1(b) to (d) show HOPG surfaces
after various times of oxidation in a Bunsen burner
ame. At very short oxidation time (<0.5 s), Fig. 1(b),
a hexagonal pattern was found to cover the surface
with a periodicity of about 20 nm. The bright features
on top of the hexagonal pattern are probably nuclei of

1807

surface oxidation. At longer oxidation times characteristic patterns of monoatomic steps were obtained.
About 5 and 10 atomic layers are exposed to the surface after 2 (Fig. 1c) and 5 s (Fig. 1d) of oxidation, respectively. On the at regions of the oxidized surfaces
the atomic lattice of graphite could be observed.
Graphite samples, which were oxidized for 2 s in the
ame were used as substrates for further experiments.
The number of atoms in defect sites for these samples
is estimated to be 2 orders of magnitude higher than
on the cleaved HOPG surface. The regular surface
morphology is supposed to develop while the sample
cools down in air, but no further attention was paid to

Fig. 1. STM images of freshly cleaved HOPG surface (A) and of graphite surfaces which were ame treated for <0.5 s (B), 2 s (C)
and 5 s (D).

1808

F. Gloaguen et al. / Electrochimica Acta 44 (1998) 18051816

the oxidation process itself; the goal being to obtain


reproducible graphite surfaces.
Cyclic voltammetry in 0.1 M HClO4 was applied to
monitor the eect of the surface oxidation on the
double layer characteristics of the HOPG electrodes.
An increase of the double layer capacity from 12 to 18
F cm2, estimated from linear sweep voltammetry at
0.1 V s1, was observed to result from the ame treatment, but the voltammetric curve remained as featureless as for a freshly pealed surface. This indicates that
no electroactive surface groups are formed by the oxidation process.
3.2. Detailed analysis of the I versus t transient
responses
Typical current versus time transient responses
recorded for a potential step from the open circuit
potential (0.75 V) to 0.1 V are shown in Fig. 2. The
deposition was performed from Ar purged and
unstirred 0.1 M HClO4+x mM H2PtCl6 (x = 2, 10).
In both cases, the shapes of the curves are rather
similar. At very short time (t < 0.3 s), a sharp peak is
observed. In the succeeding part of the transient, the
current increases with time and goes through a maximum value, IM, at the time tM. After the fall of the
current, a plateau is nally reached.
The quantity of electricity corresponding to the
initial portion of the transient is about 900 mC cm2,
which is roughly equivalent to the charge required for
the deposition of 1.0 monolayer of Pt from Pt(IV)
ions. This does not necessarily mean that one monolayer of Pt is actually formed. This charge is however
far in excess of that required to recharge the double
layer of the electrode. A similar observation was made
in the case of Pt deposition on glassy carbon [18] from
a K2PtCl4 solution.
At t>0.3 s, the transient responses have the shape
expected for a 3D nucleation process with diusion
control. According to the theory of Scharifker and
Hills [18], the rise in current corresponds to an increase
in the electroactive area. This increase in area, limited
by spherical diusion around each nucleus, is due to
(i) an increase in the nucleus size and/or (ii) an increase
in the number of nuclei. The spherical diusion zones
then overlap and the mass transfer becomes linear to a
planar surface. This change in diusion regime leads
to a decrease of the current with time.
In multiple nucleation, two limiting cases have to be
considered: high or low nucleation rate [18]. At a high
nucleation rate, all the nuclei are immediately created
and their number remains constant during the growth
process, i.e. instantaneous nucleation occurs. On the
other hand, at low nucleation rate, new nuclei are continuously formed during the whole deposition process,
i.e. progressive nucleation occurs. In order to charac-

Fig. 2. Current versus time transient responses recorded at


HOPG electrodes (Ag=1 cm2). The potential is stepped from
0.75 to 0.1 V in Ar purged and unstirred 0.1 M HClO4+2
mM H2PtCl6 (a) and 0.1 M HClO4+10 mM H2PtCl6 (b).

terize the nucleation process, the transient responses


are plotted in reduced variables [18], (I/IM)2 versus
t/tM (Fig. 3). For a rather dilute solution, i.e. 2 mM
H2PtCl6, a good correlation with the theoretical curve
for progressive nucleation is obtained (Fig. 3a). For a
more concentrated solution, i.e. 10 mM H2PtCl6, the
nucleation seems to become instantaneous (Fig. 3b).
When instantaneous nucleation occurs, the density
of nuclei, N (cm2), is estimated from [18]:
IM 0:6382zFDckN 1=2

F. Gloaguen et al. / Electrochimica Acta 44 (1998) 18051816

1809

I 2M tM 0:1629DzFc2

Since c = 105 mol cm3, IM=2.42  103 A cm2


and tM=1.55 s, the calculation yields N = 1.4  106
cm2 and D = 3.74  106 cm2 s1. On the other
hand, when progressive nucleation occurs, the product
aN0 is calculated by
IM 0:4615zFcD3=4 k 0 aN0 1=2

where the parameter a is the rate constant of nucleation at steady state and No the density of active sites.
The new material constant k 0 for Pt is
k 0 4k=3

while the diusion coecient, D, is now estimated


from
I 2M tM 0:2598DzFc2

Since c = 2  106 mol cm3, IM=3.34  104 A cm2


and tM=4.75 s, calculations yield aN0=6.87  105 s1
cm2 with D = 3.40  106 cm2 s1. The density of
nuclei at saturation, Ns (cm2), is nally given by:
Ns aN0 =2k 0 D1=2 1:9  106 cm2

In both cases, instantaneous and progressive nucleation, the calculated values for the diusion coecient
are similar and in good agreement with values reported
in literature, i.e. D = 4.5  106 cm2 s1, measured in
the course of PtCl2
6 reduction on glassy carbon [19].
The density of nuclei, N, calculated in the case of instantaneous nucleation (Eq. (1)), is close to the density
at saturation, Ns, calculated in the case of progressive
nucleation (Eq. (7)), as expected.
3.3. Evaluation of the Pt deposit characteristics from
electrochemical measurements

Fig. 3. (I/IM)2 versus t/tM analysis of transient responses


shown in Fig. 2. The solid and dotted lines correspond to
data calculated for progressive and instantaneous nucleation,
respectively [18].

The charge density, QPt (mC cm2), was calculated


from integration of the current versus time transient
response. Assuming that this quantity of charge is
mainly due to the Faradaic reaction

PtCl2
6 4e 4Pt 6Cl

the Pt loading, W (mg cm2), is


where z = 4, F is the Faraday constant (96500 C
mol1), c (mol cm3) is the PtCl2
6 concentration in solution, D is the diusion coecient and k is a material
constant. The material constant k for Pt is calculated
from
k 8pcM=r1=2

2
1

where M = 195.1 g mol is the molecular mass and


r = 21.4 g cm3 is the density of Pt. The diusion
coecient, D (cm2 s1), is estimated according to

W QPt M =4F

9
1

where M = 195.1 g mol is the atomic weight of Pt.


The contribution of parasitic processes, such as double
layer charging, partial reduction of Pt4+ to Pt3+ or
Pt2+ and hydrogen evolution at the deposited Pt
particles, are considered negligible under the given
experimental conditions.
The real surface area of the dispersed platinum, Ar
(cm2), was estimated from cyclic voltammetry in Ar
purged 0.1 M HClO4 solution. On the voltammograms

1810

F. Gloaguen et al. / Electrochimica Acta 44 (1998) 18051816

(Fig. 4), integration in a potential range from 0.05 to


0.45 V, and correction for the double layer charging,
allow estimation of the quantity of electricity, QH
(mC), due to the hydrogen adsorption or desorption.
On the other hand, according to the reaction

For a more concentrated solution (cf. sample #6 in


Table 1) an increase in both the loading and the surface area is consistent with a change in the nucleation
process from progressive to instantaneous.

Had H e

3.4. Average particle size and distribution. Comparison


with STM measurements

10

and assuming one Had per Pt surface atom, a theoretical quantity of electricity Q0H=210 mC per real cm2 of
Pt is obtained [20]. Accordingly, the Pt surface area,
Ar (cm2), is estimated from
Ar QH =Q0H

11

Since both the Pt loading and the Pt surface area are


available from electrochemical data, the specic catalyst area, S (m2 g1), is nally calculated according to:
S 100Ar =Ag W

12

The Pt loading W and the surface area Ar are dependent on PtCl2


6 concentration as well as on the length
of time, ts (s), during which the potential was stepped.
The results are summarized in Table 1. The code numbers appearing in the rst row are used to identify the
deposits in mentioned below.
Under the same concentration, length of deposition
time and potential step amplitude (cf. samples #2 and
#4 in Table 1), previous pre-impregnation increases the
loading by a factor 1.5 and the specic area by a factor
2. This has already been observed [19] and may be due
to a slow chemisorption of the PtCl2
6 anions on the
carbon surface. This can also explain a dierent shape
for the transient response (not shown here).

The specic surface area is a usual macroscopic


quantity for the characterization of dispersed catalysts,
however it is not very illustrative for the respective surface morphology. Assuming simple structural models
and correlating them with the specic surface area
allows to derive expectations for the surface morphology, which can be compared to the structural
characterization of the respective electrode surfaces.
A maximal density of nuclei, Ns12  106 cm2, was
derived from evaluation of the electrodeposition transients. Hence it is a rather straightforward model to
assume, that Ns equals the number of particles on the
electrode. A characteristic interparticle distance around
7 m should then be expected. For loadings from 4.7 to
17 mg cm2 the typical mass of a single particle is
between 2.4 and 8.5 1012 g for the dierent samples.
Assuming a hemispherical overall shape for the particles, typical radii in the range 0.37 to 0.58 mm are calculated. The surface area of a single hemispherical
particle is then between 0.9 and 2.1 mm2, leading to
1.7  102 to 4.2  102 cm2 total area of the deposit
when the particles are assumed to be smooth. Since the
measured surface area of the deposit is between 0.6
and 2 cm2, typical roughness factors between 35 and
120 are accordingly derived for the particles.
These considerable roughness factors may be modelled by assuming that the submicrometer scale particles are composed of spherical nm-scale Pt-clusters.
The characteristic cluster diameter, d (nm), is then
related to the specic area, S (m2 g1), by:
d 6000=Sr

13
3

where r = 21.4 g cm . The density of Pt particles, N 0


(cm2), is then calculated according to
N 0 1014 Ar =Ag pd 2

Fig. 4. Voltammograms recorded, at 20 mV s1, in Ar purged


0.1 M HClO4. The Pt loading is 10.8 mg cm2 (sample #5).
The current density is related to the geometric area.

14

The values of d and N 0 calculated for the various


samples are listed in Table 2. The characteristic cluster
diameters for the investigated electrodes range from 10
to 50 nm.
The characterization of the prepared electrodeposits
with STM revealed a heterogeneous distribution of Pt
on the substrate surface. Occasional feedback instabilities were detected throughout the measurements. Such
instabilities are generally accepted to result from
adparticles being swept away by the tip. The relevance
of this tip-sample interference is discussed later on.

F. Gloaguen et al. / Electrochimica Acta 44 (1998) 18051816

1811

Table 1
Electrochemical characteristics of Pt deposits on HOPG substrates (1 cm2 in geometric area). The potential was stepped from 0.75
V to Ed during ts seconds. Solutions were Ar purged and unstirred 0.1 M HClO4+x mM H2PtCl6. Other symbols are dened in
the text
Samplea

x (mM)
Ed (V)
ts (s)
W (mg cm2)
Ar (cm2)
S (m2 g1)

#1

#2

#3

#4b

#5b

#6

0.2
0.1
430
17.0
1.0
5.9

2.0
0.1
46
4.7
0.6
12.8

2.0
0.1
250
13.0
1.5
11.5

2.0
0.1
42
7.1
1.7
23.9

2.0
0.1
238
10.8
1.9
17.6

10
0.1
11
7.8
1.7
21.8

Experimental code number.bPre-impregnated over-night.

For deposits obtained from 0.2 and 2 mM solutions


extended areas were observed, which showed the
characteristic surface morphology of the oxidized
graphite surface without evidence for the presence of
electrodeposited platinum. On the other hand, regions
were observed where the typical characteristics of the
substrate were absent. These regions were composed of
agglomerates of spherical particles and are attributed
to the electrodeposited platinum.
The agglomerates were rather rough and attempts to
image them on the scale of several m resulted in irreversible changes of the deposit distribution as well as
of the tip due to the imaging. Reproducible imaging
was only possible on top of the agglomerates and on a
smaller scale. Fig. 5 shows examples of such agglomerates from dierent samples. Their presence is consistent with a rather low density of large and rough
particles on the surface. The agglomerates consist of
hills on the scale of 100 to 200 nm, the hills being composed of 5 to 30 nm diameter clusters. The size of
these clusters is in reasonable agreement with the predictions of Eq. (14), which resulted in characteristic
cluster diameters of 12 and 16 nm for the samples represented in Fig. 5(a) and (b), respectively. For Fig. 5(c)
a comparison with model expectations is not available
since the respective deposition was performed in the
regime of H2-evolution at 1 V, thus estimation of the

Pt loading from the transient charge was impossible. It


is evident that larger clusters are formed under conditions, where the reactant transport in the interfacial
region is improved due to the stirring eect of gas
evolution.
Smaller agglomerates with less overall roughness
were also observed. Imaging of such agglomerates was
rather reproducible, even on the mm-scale, but prolonged imaging was nevertheless found to alter the
morphology. An example for such an agglomerate on
sample 1 is shown in Fig. 6. In Fig. 6(a) three dierent
characteristic deposit morphologies are discerned,
namely an agglomerate of Pt clusters with sizes ranging
from 5 to 40 nm, individual clusters of 10 to 40 nm diameter and a ring-like structure, which supposedly
stems from step decoration of a hole in the graphite
top layer. The larger scale image, Fig. 6(b), shows that
the surrounding of the agglomerate is mainly bare
graphite. It should be noted that for such large scale
images it is not always clear, whether a small feature
on the surface represents an island of the substrate or
a deposit particle. This can, in principle, be claried by
zooming in on the structure of interest, as exemplied
by Fig. 7, but this procedure is time consuming and
thus inadequate for quantitative application. In addition, the structures were occasionally removed in the
course of zooming in.

Table 2
Geometrical characteristics of the Pt deposits estimated from the electrochemical measurements. Characteristic cluster diameter d
and particle density N 0 are calculated assuming smooth spherical clusters
Samplea

d (nm)
N 0 (1010 cm2)

#1

#2

#3

#4b

#5b

#6

47.5
1.4

21.9
4.0

24.4
8.0

11.7
39.6

15.9
23.9

12.9
32.5

1812

F. Gloaguen et al. / Electrochimica Acta 44 (1998) 18051816

Fig. 5. Morphology of large Pt agglomerates formed under dierent deposition conditions. (A) Deposition at 0.1 V for 42 s from 2
mM solution (sample #4), (B) deposition at 0.1 V for 238 s from 2 mM solution (sample #5) and (C) deposition at 1 V for 5 s
from 10 mM solution. The height scale (black to white) corresponds to 16, 55 and 70 nm for A, B and C, respectively.

Fig. 6. Pt agglomerate formed by deposition at 0.1 V for 430 s from 0.2 mM solution (sample #1), imaged at dierent scales.

Fig. 7. High resolution image of isolated platinum particle on defective HOPG.

F. Gloaguen et al. / Electrochimica Acta 44 (1998) 18051816

1813

Fig. 8. Distribution of the Pt deposit after deposition at 0.1 V for 42 s from 2 mM solution (sample #4), imaged on dierent
scales.

Fig. 8 shows the deposit distribution on sample 4.


On a small scale, Fig. 8(a), a similar agglomerate as in
Fig. 6(a) is seen. Larger scale images, Fig. 8(b) and (c),
reveal that the distribution of agglomerates is more
homogeneous than on sample 1. Between 50 and 100
particles per mm2 can be distinguished. This gives an
extrapolated particle density between 5  109 and 1010
per cm2, more than 3 orders of magnitude above the
evaluated density of nucleation centers.
Qualitatively dierent results were obtained on
sample 8. On this sample, in the regions between platinum agglomerates, a considerable density of small isolated clusters was observed which were mainly
attached to small graphite islands. Examples are
shown in Fig. 9. The diameter of the clusters was
around 10 nm, their extrapolated density is of the
order of 1010 cm2. From size and density these isolated clusters are estimated to represent at most 10%
of the total deposit.
4. Discussion
STM is a nm-scale technique and the full exploitation of its capabilities regarding the structural characterization of adparticles on substrates is only possible
for samples, which satisfy several severe restrictions.
Firstly, the used substrate should be extremely at in
order to allow a unambiguous separation of adparticles and substrate features. Secondly, the adparticles
should be in the 150 nm diameter range in order to
allow their reasonable imaging. Thirdly, the adparticles
should be uniformly distributed with densities between
109 and 1011 cm2.
The goal of the present work was, to prepare model
electrodes which satisfy these restrictions by electrochemical deposition of Pt onto HOPG substrates with
a controlled defect density. Low Pt4+-concentrations

and high deposition overpotentials were considered as


the most promising conditions. For Ag-electrodeposition it has been shown, that dened nm-scale cluster
distributions on HOPG can be obtained by a similar
approach [6, 7]. For Pt this was also achieved, but for
a metal loading which is equivalent to 0.1 monolayer
of Pt, at best [8].
Evaluation of the transients recorded in the course
of the deposition indicated a rather low saturation coverage for nuclei formation of 2  106 cm2.
Consequently the monoatomic steps, which have a
high density on the oxidized substrate surface, are supposed to have a low probability to act as nucleation
centers. From the low density of nuclei one can derive
the expectation that very large and rough particles are
formed during deposition with typical interparticle
distances of several mm. STM images can thus not be
expected to reveal the deposit distribution in arepresentative way.
The structural characterization of the respective electrodes with STM showed, that indeed large agglomerates of platinum are present on the surface (Fig. 5),
aside from extended uncovered substrate areas (similar
to Fig. 1c). On the agglomerates clusters on the scale
of several nm are observed. Considering the high
roughness factors evaluated from the electrochemical
data one needs to assume that the agglomerates are
porous, e.g. they may rather be envisaged as stochastically packed piles of nm-scale clusters than as massive
crystallites with a rough surface. So far, a reasonable
qualitative agreement between the expectations derived
from the electrochemical data and the structural investigation with the STM is obtained. The characterization of Pt electrodeposits on glassy carbon with
TEM [21] and SEM [22] gave similar results, except
that the observed particle densities were signicantly
higher.

1814

F. Gloaguen et al. / Electrochimica Acta 44 (1998) 18051816

Fig. 9. Distribution of Pt clusters after deposition at 0.1 V for 11 s from 10 mM solution (sample #8).

This seems to indicate that each agglomerate is the


result of the growth of a single nucleus and that the
density of nuclei at saturation equals the area density
of Pt agglomerates on the substrate. Several observations are, however, inconsistent with this conclusion.
These are namely the overall shape of the agglomerates
(Fig. 6) and the observed local densities of small
particles (Figs. 8 and 9).
The dimensions of the elongated agglomerate shown
in Fig. 6 (height:width:length 1 1:10:50), as an
example, are very far from hemispherical. In addition,
several small particles are resolved in the vicinity of
the agglomerate, but without direct contact to it. Such
a morphology can not be imagined as resulting from
the growth of a single nucleus. Further on, the local

densities of particles of the order of 1010 cm2, exemplied by Figs. 8 and 9, is dicult to correlate with
the determined saturation density of nuclei. Thus it is
concluded that the evaluation of electrochemical data
and the structural investigation, although not totally
contradictive, are not compatible in detail.
Regarding the electrochemical data, a source of
uncertainty may be the excess charge convoluted with
the double layer recharging. This charge is roughly
equivalent to that required for the deposition of 1.0
monolayer of Pt and was not considered for the
evaluation. Regarding the structural investigation, two
crucial questions need to be discussed. The rst is, how
the ex-situ surface structure is related to the in-situ
deposit morphology or, in other words, how the

F. Gloaguen et al. / Electrochimica Acta 44 (1998) 18051816

1815

Fig. 10. Contact mode AFM images of Pt clusters electrochemically deposited onto HOPG. The Pt loading is ca. 155 mg cm2.

process of emersion, rinsing and drying may aect the


deposit distribution. For electrodeposited silver it has
been shown, that this process leaves the deposit morphology unchanged [6, 7], which is surprising in view
of the high exchange current density of silver. Thus
it may be considered unlikely that the situation is
dierent for platinum.
The second question is, how the imaging process
may aect the deposit distribution. It has already been
stated that occasional feedback instabilities indicated
that particles were swept away by the tip. Hence more
particles are present on the surface than are imaged
with the STM. The discrepancy between particle density and evaluated density of nuclei may thus be more
pronounced than evident from the STM data, but this
is of minor concern for the present discussion.
A possible explanation for the apparent discrepancy
between electrodeposition kinetics and surface morphology is derived from recent investigations of cluster
formation on HOPG under UHV conditions [23, 24].
These made evident that the deposit distribution on
HOPG can only be rationalized by assuming a

signicant mobility of clusters on the surface at room


temperature.
Considering these observations and a recently
reported work [8] it seems reasonable to conclude that
a large amount of very small (11 nm) platinum clusters is formed during the initial spike of the electrodeposition transient, i.e. at t < 0.3 s. In the course of Pt
electrodeposition, these very small clusters may be
mobile and assemble in an accidental distribution of Pt
agglomerates. At t>0.3 s, the rapid growth of a small
number (2  106 cm2) of Pt particles give a transient
response, which is adequately described by the model
of Scharifker and Hills for a diusion controlled
growth.
An additional experimental result (Fig. 10) supports
the assumption of cluster mobility in an electrochemical environment: in the course of the imaging process
with contact mode AFM, no Pt clusters moved under
the tip, conversely to what was sometimes observed
with STM, even at very low tunneling currents (0.5
nA). This discrepancy between contact AFM and STM
may be explained by a poor electronic contact between
some Pt crystallites and the graphite surface yielding

1816

F. Gloaguen et al. / Electrochimica Acta 44 (1998) 18051816

to a crash of the STM tip (the AFM technique is not


sensitive to variations of the electronic conduction of
the sample). A poor electronic contact between Pt and
graphite implies that some Pt crystallites are not
located at their nucleation sites, where Pt electroreduction takes place. According to this, the nal morphology of the Pt electrodeposit may not depend on
the density of nucleation sites only, but may result in
the surface diusion of Pt ad-particles formed in the
course of the electrodeposition process.
A more systematic investigation, including the eect
of the deposition overpotential and the Pt salt concentration, as well as the inuence of the substrate morphology and structure, should, however, be performed
in order to clarify the above assumptions. This is
beyond the purpose of this paper.
5. Conclusions
The electrodeposition of platinum on graphite substrates with a high density of monoatomic steps was
investigated with electrochemical techniques and exsitu STM. The quantitative evaluation of the electrodeposition current versus time, at t>0.3 s, was carried
out according to the nucleation-growth model of
Scharifker and Hills. This analysis was found to be inconsistent with the structural investigation of the electrode surfaces. A very heterogeneous deposit
distribution was observed for all investigated samples,
which cannot reasonably be correlated with a distribution of active nucleation sites as calculated from the
transient analysis at t>0.3 s. As a trend it is observed,
that the density of particles increases with increasing
concentration of the electroactive species in solution. A
qualitative understanding of the experimental results is
obtained under the following assumptions, (i) Pt clusters are formed at t < 0.3 s and (ii) the diusion of Pt
clusters on the graphite surface is an important step in
the course of the phase formation process.
Acknowledgements
This work was carried out with the support of
ADEME, the French Agency for Environment and
Energy Conservation, to which we are greatly
indebted. We also acknowledge the support of the
FrenchGerman Exchange Program PROCOPE
through the `Deutsche Akademischer Austauschdienst'
and the French `Ministere des Aaires Etrangeres'.

References
[1] S. Mukerjee, J. Appl. Electrochem. 20 (1990) 537.
[2] M.S. Wilson, F.H. Garzon, K.E. Sickafus, S. Gottesfeld,
J. Electrochem. Soc. 140 (1993) 2872.
[3] Y. Takasu, N. Ohashi, X.-G. Zhang, Y. Murakami, H.
Minagawa, S. Sato, K. Yahikozawa, Electrochim. Acta
41 (1996) 2595.
[4] P.C. Biswas, Y. Nodasaka, M. Enyo, J. Appl.
Electrochem. 26 (1996) 30.
[5] G. Tamizhmani, J.P. Dodelet, D. Guay, J. Electrochem.
Soc. 143 (1996) 18.
[6] J.V. Zoval, R.B. Stiger, P.R. Biernacki, R.M. Penner, J.
Phys. Chem. 100 (1996) 837.
[7] J.V. Zoval, P.R. Biernacki, R.M. Penner, Anal. Chem.
68 (1996) 1585.
[8] J.V. Zoval, J. Lee, S. Gore, R.M. Penner, J. Phys. Chem.
B 102 (1998) 1166.
[9] R.J. Nichols, D.M. Kolb, R.J. Behm, J. Electroanal.
Chem. 313 (1991) 109.
[10] S.A. Hendricks, Y.-T. Kim, A.J. Bard, J. Electrochem.
Soc. 139 (1992) 2818.
[11] U. Stimming, R. Vogel, D.M. Kolb, T. Will, J. Power
Sources 43 (1993) 169.
[12] N. Breuer, U. Stimming, R. Vogel, Electrochim. Acta 40
(1995) 1401.
[13] N. Breuer, U. Stimming, R. Vogel, Surf. Coatings
Technol. 67 (1994) 145.
[14] N. Breuer, U. Stimming, R. Vogel, in: A.A. Gewirth, H.
Siegenthaler (Eds.), Nanoscale Probes of the Liquid
Solid Interface, vol. 288, Kluwer, Dordrecht, 1995, p.
121.
[15] X.Q. Tong, M. Aindow, J.P.G. Farr, J. Electroanal.
Chem. 395 (1995) 117.
[16] J.L. Zubimendi, L. Vazquez, P. Ocon, J.M. Vara, W.E.
Triaca, R.C. Salvarezza, A.J. Arvia, J. Phys. Chem. 97
(1993) 5095.
[17] H. Chang, A.J. Bard, J. Am. Chem. Soc. 112 (1990)
4598.
[18] J. Lin-Cai, D. Pletcher, J. Electroanal. Chem. 149 (1983)
237.
[19] B. Scharifker, G. Hills, Electrochim. Acta 28 (1983) 879.
[20] K. Shimazu, D. Weisshaar, T. Kuwana, J. Electroanal.
Chem. 223 (1987) 223.
[21] P. Allongue, E. Souteyrand, J. Electroanal. Chem. 286
(1990) 217.
[22] H. Shimazu, K. Uosaki, H. Kita, Y. Nodasaka, J.
Electroanal. Chem. 256 (1988) 481.
[23] G.M. Francis, I.M. Goldby, L. Kuipers, B.
VonIssendor, R.E. Palmer, J. Chem. Soc. Dalton
Trans. 0 (1996) 665.
[24] G.M. Francis, L. Kuipers, J.R.A. Cleaver, R.E. Palmer,
J. Appl. Phys. 79 (1996) 2942.

Anda mungkin juga menyukai