You are on page 1of 11

This article was originally published in a journal published by

Elsevier, and the attached copy is provided by Elsevier for the


authors benefit and for the benefit of the authors institution, for
non-commercial research and educational use including without
limitation use in instruction at your institution, sending it to specific
colleagues that you know, and providing a copy to your institutions
administrator.
All other uses, reproduction and distribution, including without
limitation commercial reprints, selling or licensing copies or access,
or posting on open internet sites, your personal or institutions
website or repository, are prohibited. For exceptions, permission
may be sought for such use through Elseviers permissions site at:
http://www.elsevier.com/locate/permissionusematerial

Electrochimica Acta 52 (2007) 23762385

py

In situ FTIR studies of the catalytic oxidation of ethanol on Pt(1 1 1)


modified by bi-dimensional osmium nanoislands

co

V. Pacheco Santos, V. Del Colle 1 , R. Batista de Lima, G. Tremiliosi-Filho


Universidade de Sao Paulo, Instituto de Qumica de Sao Carlos-USP, CP 780, 13560-970 Sao Carlos, SP, Brazil
Received 13 March 2006; received in revised form 22 August 2006; accepted 22 August 2006
Available online 10 October 2006

Abstract

pe

rs

on

al

This paper presents the study of ethanol electrooxidation on Pt(1 1 1) electrode modified by different coverage degrees of a submonolayer of
osmium nanoislands, which were obtained by spontaneous deposition. The ethanol oxidation reaction was extensively studied by employing in
situ FTIR. Collections of spectra of the ethanol adsorption and oxidation processes were acquired over a series of positive potential steps, in order
to determine the intermediate species and the main products that are formed. It was shown that the increase in the catalytic activity of Pt(1 1 1)
after osmium deposition for ethanol oxidation is greater than that observed on nonmodified Pt(1 1 1). It was also demonstrated that the mechanistic
pathway for this reaction depends directly on the degree of osmium coverage. Thus, for low osmium coverage ( Os 0.28), the formation of CO
as an intermediate is favored, and hence the full oxidation of adsorbed ethanol to CO2 is increased, additionally, the formation of acetaldehyde is
also observed in low degrees of osmium coverage. For intermediate osmium coverage (0.28 < Os 0.40), the oxidation of ethanol to acetaldehyde
and then to acetic acid is favored, although on Pt(1 1 1) the formation of acetaldehyde is promoted. For higher degrees of osmium coverage
( Os > 0.51), the catalytic activity of the electrode for ethanol oxidation decreases. For an almost complete osmium layer ( Os = 0.92), obtained by
electrodeposition at 50 mV, catalytic activity for ethanol oxidation shows the lowest value. In addition, the direct oxidation of ethanol to acetic acid
at lower potentials is observed.
2006 Elsevier Ltd. All rights reserved.
Keywords: Ethanol electrooxidation; In situ FTIR; Platinum single-crystal electrode; Electroless deposition; Osmium nanoislands

r's

1. Introduction

Au

th
o

In the last two or three decades, much attention has been


devoted to the study of the electrooxidation of small organic
molecules, due to their possible utilization as fuels in direct
organic fuel cell (DOFCs) devices [13]. In this period, a considerable number of spectroscopic, electrochemical and mass
analysis methods have been employed in order to determine the
reaction intermediates and to elucidate the manner in which the
organic molecules adsorb on metal surfaces [4,5]. Among these
methods, the most employed has been FTIR spectroscopy, due to
the fact that it can provide considerable information and valuable
insight into the electrocatalytic processes that are of interest in
Corresponding author. Tel.: +55 16 33739933; fax: +55 16 33739952.
E-mail address: germano@iqsc.usp.br (G. Tremiliosi-Filho).
1 Present address: Departamento de Qumica, Fundac
a o Universidade Estadual de Alagoas, Av. Governador Luiz Cavalcanti S/N, 57312-000 Arapiraca
(AL), Brazil.
0013-4686/$ see front matter 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.electacta.2006.08.044

electrochemical surface science [614]. FTIR spectroscopy provides a way to understand the nature of the interactions between
organic molecules and the electrode surface. Additionally, this
technique supplies a molecular-level description of the reactions
and allows a correlation of the catalytic activity with the crystallographic structure of the electrode surface. Thus, it is possible to
elucidate the mechanistic pathways of various surface reactions
and to analyze the electrooxidation kinetics of several organic
molecules [15].
Basically, a dual path mechanism has been proposed for
ethanol electrooxidation. The first path is via cleavage of the
C C bond, which preferentially produces adsorbed CHx species
[12] and adsorbed carbon monoxide [9,12,14]. The second
path is via oxidation of adsorbed ethanol to form acetaldehyde
[9,14,16] and acetic acid [14,17].
Most studies of ethanol electrooxidation have been carried
out on polycrystalline platinum electrodes. This has been done
mainly to quantify the reaction intermediates, products formed
and/or kinetic parameters [68,11,12,18]. However, only a few

V. Pacheco Santos et al. / Electrochimica Acta 52 (2007) 23762385

on

al

co

py

monolayer of osmium nanoislands were deposited on Pt(1 1 1)


electrode.
There is a well-known connection between the structure and
the reactivity of each single-crystal face of platinum. Thus, the
electrocatalysis of ethanol oxidation is known to depend on the
surface geometry of platinum [14,17,20]. The alteration of low
Miller index crystallographic planes, from Pt(1 1 1) to Pt(1 0 0)
or Pt(1 1 0), causes a significant change in the work function and
in the coordination properties of the surface. It is very attractive
to study the structural effect involved in the enhancement of
surface reactivity on well-ordered surfaces such as Pt(1 0 0) [40]
and Pt(1 1 1). This last surface will be considered in this paper.
Additionally, fundamental studies of the catalytic decomposition
of organic molecules on well-ordered platinum surfaces are very
important due to the fact that in practical applications the metal
catalysts are dispersed on substrates as crystallite nanoparticles
[41,42].
As previously described, there is conclusive experimental
evidence that ethanol oxidation is enhanced after osmium addition on platinum single-crystal surfaces, and that this enhancement depends on their surface geometry [31,32,40]. These
observations create a template for investigating the oxidation
process of ethanol on platinum surfaces that are modified with
osmium. Furthermore, the Pt(1 1 1)/Os system shows higher catalytic activity for ethanol oxidation at lower potentials than
Pt(1 0 0)/Os [31] and FTIR spectroscopy can certainly provide a
molecular-level perspective, as an aid to propose the mechanistic
pathways of ethanol oxidation and the mechanistic differences
among each of these surfaces. Therefore, the aim of this work
is to study the activity of Pt(1 1 1), modified by spontaneous
osmium deposition, for ethanol oxidation and to identify the
intermediates and products of this reaction in series of potential
steps from 0.1 to 1.0 V, using in situ FTIR spectroscopy.

Au

th
o

r's

pe

rs

studies have detailed FTIR studies of this reaction on singlecrystal surfaces [9,14,19,20]. Some aspects, such as the reaction
mechanism and surface reactivity, can be more precisely investigated by working with well-defined surfaces.
Ethanol is a very attractive liquid fuel for DOFC devices
[21,22]. Its low toxicity added to its high availability are important characteristics for use as a fuel in DOFC systems. A very
efficient catalyst for alkaline direct alcohol fuel cell based in a
mixture of Fe, Co and Ni was claimed by Bianchini et al. [23].
However, this interesting device operate in alkaline medium and
does not need platinum as catalyst. When platinum is used as a
catalytic material in acid medium, occurs a significant surface
poisoning that is detrimental for the catalysis. Then, further studies are necessary to assure that ethanol can be considered as an
ideal candidate for such applications. The poisoning is caused
by dissociative ethanol chemisorption, which leads to the formation of strongly adsorbed and partially oxidized products, such
as CO, CHOH (enol type species) and/or CH3 CHO, depending
mainly on the bulk ethanol concentration [8] and the electrode
surface structure [17,20].
To improve the electrocatalytic activity of platinum, the addition of a second metal (e.g. Ru [2428], Os [2932], Sn [24,33]
and/or Rh [34]) has been widely used to form a bimetallic surface
catalyst. The well-known reason for the use of these promoters is
explained on the basis of the bifunctional mechanism [35] and/or
the electronic effect [28]. Another important issue in ethanol oxidation is to find metallic promoters that are capable of breaking
the C C bond of the molecule [16,36].
Apparently, the PtRu system represents a promising alternative for ethanol oxidation [2427]. In situ FTIR spectroscopy
suggests that this catalyst can break the ethanol C C bond at
lower potentials [27]. However, studies involving 13 C-labelled
ethanol, ascertained that Ru stimulates the formation of CO2
(originating from the CH2 OH group) at low potentials on Pt.
In this light, it seems that Ru is not a good modifier for ethanol
oxidation [36].
On the other hand, Wieckowski and co-workers [29], while
comparing PtRu and PtOs electrodes, demonstrated that
Os deposits show better catalytic activity than Ru deposits
for methanol oxidation at potentials of 0.60.7 V, although
Ru is more active for this reaction at potentials lower than
0.6 V. According to these authors, the Ru deposits present
excessively higher oxidation states in the potential range of
0.60.7 V, which is known to be prejudicial to the catalytic
activity of the electrode. Moreover, the Os deposits present a
small anodic wave in this potential range (ca. 0.65 V), related
to OsO2 formation [37,38]. This enhances significantly the
catalytic activity of PtOs electrodes for methanol oxidation,
as well as avoiding the complete formation of Ru oxides in
their highest oxidation state in a PtRuOs ternary electrode
[39].
A few studies demonstrate that osmium tends to improve
the catalytic activity of the electrode for ethanol electrooxidation [31,32,40]. Osmium, spontaneously deposited on
the surface of platinum single-crystal electrodes, forms
nanoislands with diameters ranging from 1 to 4 nm with
monoatomic thickness [32]. In the present study, similar sub-

2377

2. Experimental
2.1. Chemicals
All solutions were prepared using MilliQ purified water
(>18 M). A solution of 0.1 M HClO4 (Merck Grade Suprapur)
was prepared as the base electrolyte. The osmium deposits were
obtained from a solution of 1.0 mM H2 OsCl6 (Alfa Aesar) in
0.1 M HClO4 . The reactant solution for the ethanol oxidation
study was 0.5 M ethanol (J.T. Baker) in 0.1 M HClO4 . In order
to deaerate the solutions, argon (99.99%) or nitrogen (99.999%)
was used.
2.2. Electrodes and cells
Commercial Pt(1 1 1) cylinder of ca. 5.0 mm in diameter and
6.3 mm in height, supplied by MaTek, were used as working
electrodes. The electrochemical experiments were performed in
a conventional glass cell. The FTIR experiments were carried
out in a glass cell fitted with 60 prismatic CaF2 or flat ZnSe windows in the thin electrolyte layer pattern [13,14]. All potentials
in this study are referred to the reversible hydrogen electrode
(RHE).

2378

V. Pacheco Santos et al. / Electrochimica Acta 52 (2007) 23762385

2.3. Instrumentation
The electrochemical measurements were performed by using
a Potentiostat-Galvanostat Autolab, model PGSTAT 30. In situ
FTIR spectroscopy experiments were performed with a Nicolet
Nexus 670 spectrometer and an MCT detector.

py

2.4. Experimental procedure

submonolayer level by spontaneous deposition (Eo.c = 0.79 V)


at different deposition times between 2 and 150 s, and an almost
complete osmium layer was electrodeposited at 50 mV. After
each deposition, and before ethanol oxidation, the electrode
was firstly polarized at 50 mV for 5 s and cycled four times
in the hydrogen/adsorptiondesorption region (50900 mV) at
50 mV s1 in the base electrolyte saturated with N2 . The spontaneous osmium deposition times utilized in this work were related
to the respective osmium coverage degrees, according to the
method described in a previous study, in which the Os values
were estimated from the decrease of the charges of hydrogen
adsorption/desorption peaks in the voltammograms obtained in
HClO4 solution [40].
For the ethanol oxidation studies, the electrode was immersed
in a 0.5 M C2 H5 OH + 0.1 M HClO4 solution. During this process, the electrode was maintained polarized at 0.05 V for 5 s in
order to avoid strong ethanol adsorption on platinum. After this,
the voltammetric study of ethanol electrooxidation (first positive
scan) was recorded for each PtOs system, in the potential range
of 0.050.9 V at 20 mV s1 .
In another experiment, FTIR spectra of ethanol oxidation
were obtained using a CaF2 prismatic window. In this way, a
reference FTIR spectrum was acquired at 0.05 V. Then, other
spectra were recorded after applying successive potential steps
of 50 mV in the positive direction from 0.1 to 1.0 V versus RHE.
For this, spectra were computed from an average of 256 interferograms and the spectral resolution was set at 8 cm1 . Reflectance
spectra were calculated as the ratio (R/R0 ) of the sample (R) and
the reference (R0 ) spectra. Positive and negative growing bands
represent loss and gain of species at the sampling potentials,
respectively. When extension of the low wave number limit was
necessary (below 1000 cm1 ), the CaF2 window was replaced
by ZnSe.
After finishing each osmium deposition and respective
ethanol oxidation FTIR studies, the electrode was cycled
between 0.05 and 1.65 V at 200 mV s1 in a fresh electrolyte.
This procedure favors the oxidation and desorption of the
osmium deposits, which can be confirmed by the return to the
characteristic voltammetric profile of the platinum electrode in
HClO4 solution. Hydrogen flame annealing of the electrode was
also necessary, in order to regenerate the original single-crystal
surface due to the surface disordering observed after removal of
the osmium deposits [31,32].

Au

th
o

r's

pe

rs

on

al

co

Before each osmium deposition, the correct surface orientation of Pt(1 1 1) electrode was confirmed by cyclic voltammetry (always in a fresh HClO4 solution) using a meniscus
configuration. The voltammetric profiles also demonstrate the
cleanliness of the systems studied [43] (see solid line in Fig. 1).
Following this, osmium deposition on Pt(1 1 1) was performed
from a fresh solution of 1.0 mM H2 OsCl6 + 0.1 M HClO4 . The
osmium deposits were obtained on the electrode surface at a

Fig. 1. Cyclic voltammograms at 50 mV s1 in 0.1 M HClO4 of nonmodified


() and modified (
) Pt(1 1 1) with different osmium coverage degrees.

3. Results
3.1. Voltammetric studies
3.1.1. Characterization of osmium deposits
The characterization of osmium deposits on Pt(1 1 1) was carried out by cyclic voltammetry in 0.1 M HClO4 . The voltammetric profiles of nonmodified Pt(1 1 1) and of the same electrode
with different degrees of osmium coverage, obtained by spontaneous deposition, as well as with an almost complete osmium
layer obtained by electrolysis at 50 mV, are shown in Fig. 1.
As can be seen in Fig. 1, an increase of the anodic and
cathodic currents in the potential range of ca. 0.350.6 V

Au

th
o

co
al

r's

pe

rs

3.1.2. Ethanol electrooxidation on Pt(1 1 1)/Os


The voltammetric profiles (first positive scan) of ethanol oxidation on nonmodified (solid lines) and modified Pt(1 1 1) by
different coverages of spontaneously deposited osmium and on
the electrode covered with an almost complete osmium layer
obtained by electrodeposition, are presented in Fig. 2. After
osmium modifications the voltammetric profile for ethanol oxidation changes. A wide peak arises after osmium modification
and it is enhanced with the increase of the osmium coverage
degree, reaching a maximum value at Os = 0.51. In the potential range of 0.050.5 V, which is shown in detail in the inset in
Fig. 2, it is possible to observe that the highest current density
values for ethanol oxidation are obtained when the osmium coverage degree is 0.51. As will be emphasized later in Section 3.3,
this happens as a consequence of the formation of acetaldehyde
on Pt(1 1 1)/Os.
Voltammograms of ethanol electrooxidation on Pt(1 1 1) with
an almost complete osmium layer ( Os = 0.92) and for Os = 0.28
show similar anodic waves. This apparent strange behavior will
be discussed further in Section 5.

on

is observed in the voltammetric profile of Pt(1 1 1) after


osmium deposition. This extra current on the capacitive
one is attributed to the faradaic oxidation/reduction processes (Os + 2H2 O  OsO2 + 4H+ + 4e ) of osmium deposits
[32,37,40]. Such currents are enhanced as the degree of osmium
coverage increases, whereas the hydrogen adsorption/desorption
peaks in the potential range of 0.050.35 V diminish. In this
work, the osmium coverage degree on Pt(1 1 1) was estimated
from the charge decrease in the hydrogen peaks as the deposition
time increased, similarly to the method described in previous
papers [31,32] and utilized in a recent study for Pt(1 0 0) [40]. In
addition to the gradual increasing of the anodic/cathodic currents
above 0.35 V after osmium has been deposited, the lowering
of the characteristic peak of Pt(1 1 1) in HClO4 solution in the
potential range of 0.50.8 V (attributed to the formation of (OH)
species on the electrode [4447]) is also observed.
The voltammetric characterization of the osmium deposits
on Pt(1 1 1) infer that the highest value of osmium coverage
obtained by spontaneous deposition, from an H2 OsCl6 /HClO4
solution, is ca. 0.51. However, in order to obtain additional information about the catalytic effect of greater osmium coverage on
the ethanol oxidation reaction, the results for the coverage degree
of 0.92 (obtained by electrolysis at 50 mV) were also included
in this work.

2379

py

V. Pacheco Santos et al. / Electrochimica Acta 52 (2007) 23762385

3.2. In situ FTIR spectroscopy studies using a 60 prismatic


CaF2 window: detection of CO, CO2 and acetic acid
3.2.1. General information
The CaF2 window is appropriate to detect CO, CO2 and acetic
acid. The collections of FTIR spectra for ethanol electrooxidation over the series of positive potential steps on Pt(1 1 1) with
different osmium coverage degrees, show that the bands of the
main intermediates and products of this reaction (adsorbed CO
linearly bonded (2040 cm1 ) [14], CO2 (2340 cm1 ) [6,7,10,14]
and acetic acid (1290 cm1 ) [6,14,50]) increase significantly

Fig. 2. First positive anodic sweep of ethanol oxidation (0.5 M C2 H5 OH + 0.1 M


) Pt(1 1 1)
HClO4 ) at 20 mV s1 , on nonmodified () and modified (
with different osmium coverage degrees.

after osmium surface modification. In addition, other bands


arise in the spectra obtained with a ZnSe window. These
include mainly the band at 933 cm1 (attributed to the C C O
group of the acetaldehyde [14,48,49]). Additionally, the band at
1290 cm1 (related to the O H bond of COOH group of the
acetic acid [6,14,50]) and the band at 1710 cm1 (attributed
to the carbonyl group of both acetaldehyde and acetic acid
[6,10,14]) are also observed when ZnSe is used as window.
Another intense band at 1100 cm1 (attributed to the presence of perchlorate anion in the thin electrolyte layer [14]) is
also observed. The last two bands will not be considered in
this work.

2380

V. Pacheco Santos et al. / Electrochimica Acta 52 (2007) 23762385

co

al

on

Au

th
o

r's

pe

rs

3.2.2. Spectra analysis


Collections of spectra were acquired for nonmodified
Pt(1 1 1) and Pt(1 1 1) modified by different osmium coverage
degrees, as can be seen in Fig. 3ad. The wave number range of
2500 and 1200 cm1 selected in this figure includes the bands
associated to adsorbed CO, CO2 , carbonyl group (C O) and carboxyl group (O H bond of the COOH group) in order to clarify
all the features of the spectra when a CaF2 window is used.
However, only the band intensities of CO, CO2 and carboxyl
group are quantitatively analyzed in the sequence. The intensity of the adsorbed COlinear and CO2 bands (Ib ), taken from the
series of spectra in Fig. 3, are plotted in Fig. 4 as a function of the
potential for nonmodified and modified Pt(1 1 1). An important
difference between Pt(1 0 0) [40] and Pt(1 1 1), nonmodified or
modified by osmium addition, is the total absence of adsorbed
bridge-bonded CO on Pt(1 1 1).
In Fig. 4a, it can be observed that the adsorption of COlinear ,
which commences above 0.3 V on nonmodified Pt(1 1 1), is
shifted to a potential ca. 100 mV lower on Pt(1 1 1)/Os electrodes. Furthermore, the intensity of the adsorbed COlinear band
increases significantly on Pt(1 1 1) after osmium deposition, with
the highest increase at Os = 0.28. For higher osmium coverage degrees on Pt(1 1 1) ( Os = 0.40 and 0.51), the quantity of
adsorbed linear-bonded CO diminishes and a very low value is
reached at Os = 0.92. Except for the highest osmium coverage

obtained on Pt(1 1 1) ( Os = 0.92), in which the low amount of


adsorbed COlinear is rapidly consumed in the potential range
of 0.40.6 V, adsorbed linear-bonded CO reaches a maximum
value at 0.45 V on Pt(1 1 1)/Os and is desorbed from the surface
in the potential range of 0.450.8 V.
Also in Fig. 4b, it is observed that the formation of CO2
on Pt(1 1 1)/Os starts at 0.45 V. The onset potential of CO2
formation on Pt(1 1 1) does not change as a function of the
osmium coverage. The amount of CO2 is significantly enhanced
on Pt(1 1 1) after osmium modification, principally in the potential range of 0.450.6 V. As observed for adsorbed COlinear , the
maximum value of the CO2 band is found for Pt(1 1 1) with a low
osmium coverage ( Os = 0.28). For higher osmium coverages
( Os = 0.40 and 0.51), the amount of CO2 diminishes slightly.
On Pt(1 1 1)/Os, the CO2 band intensity reaches a maximum
value at ca. 0.8 V, except for Pt(1 1 1) with Os = 0.92, where
the lower quantity of CO2 is observed. Moreover, the very low
amount of adsorbed COlinear formed on the Pt(1 1 1) surface with
an almost complete osmium layer suggests that the CO2 comes
from another reaction intermediate, such as acetic acid. This will
be discussed further in Section 4.
On the surfaces studied, the most significant enhancement
of the catalytic activity of Pt(1 1 1) for the complete oxidation
of adsorbed ethanol to CO2 occurs at low osmium coverage
degrees. This can be related to their greater capacity to cleave
the ethanol C C bond at lower potentials than those observed
for nonmodified electrode. This can lead to the preferential formation of adsorbed CO as an intermediate and hence to CO2
as the main product of the ethanol oxidation reaction. However, the voltammetric studies of the ethanol electrooxidation on
Pt(1 1 1)/Os indicate that ethanol electrooxidation is improved
as a consequence of the increase of osmium coverage degree to
ca. 0.400.51. This suggests that the reaction might follow other
mechanistic pathways for higher degrees of osmium coverage,
which will be empathized in the next section.

py

As was observed in a previous study [40], the consumption


of ethanol from the thin electrolyte layer, within the potential
range studied, is confirmed by the positive growth of the C H
bands at 2980 and 2900 cm1 . These two bands correspond to
CH3 and CH2 groups from ethanol, respectively [7,12,14]. The
ClO4 band (1100 cm1 ) grows as the potential increases, due to
the entering of this species in the thin electrolyte layer, in order
to compensate the loss of charge during the positive potential
steps.

Fig. 3. Collections of in situ FTIR spectra (256 scans at 8 cm1 resolution, obtained with a CaF2 window) of the series of positive potential steps for ethanol oxidation
(0.5 M C2 H5 OH + 0.1 M HClO4 ) on: (a) nonmodified Pt(1 1 1) and on Pt(1 1 1) modified with different osmium coverage degrees: (b) Os = 0.28, (c) Os = 0.51 and
(d) Os = 0.92.

2381

r's

pe

rs

on

al

co

py

V. Pacheco Santos et al. / Electrochimica Acta 52 (2007) 23762385

Fig. 4. Band intensity of: (a) adsorbed COlinear at 2040 cm1 and (b) CO2 at 2340 cm1 as a function of the potential, from FTIR spectra of ethanol oxidation on
nonmodified Pt(1 1 1) and on Pt(1 1 1) modified with different osmium coverage degrees (data obtained from Fig. 3).

Au

th
o

The influence of osmium coverage on the amount of acetic


acid formed during the ethanol oxidation on Pt(1 1 1)/Os has
been studied in Fig. 5. In this figure, the intensities of the
band at 1290 cm1 are plotted as a function of the potential for
Pt(1 1 1)/Os electrodes. It can be clearly observed that the acetic
acid band is significantly enhanced after osmium modification,
even with low values of osmium coverage (0.19 and 0.28), in
comparison to the nonmodified electrode. On Pt(1 1 1)/Os, the
acetic acid band arises in the FTIR spectra of ethanol oxidation
at around 0.35 V. This band, however, becomes higher only at
potentials above 0.6 V.
3.3. In situ FTIR spectroscopy studies using a ZnSe
window: detection of acetaldehyde and acetic acid
3.3.1. General information
Additional information about the probable intermediates and
products of ethanol oxidation, which were formed during the

series of positive potential steps on Pt(1 1 1)/Os electrode, was


obtained using FTIR spectra at wave numbers below 1000 cm1 .
This was achieved using a ZnSe window.
3.3.2. Spectra analysis
The analyses of the acetaldehyde band at 933 cm1 and acetic
acid band at 1290 cm1 are done in the sequence. A selected
group of spectra obtained with a ZnSe window, for nonmodified Pt(1 1 1) and for each Pt(1 1 1)/Os electrode, is shown in
Fig. 6. The influence of osmium coverage on the band intensities of acetaldehyde and acetic acid formed during the ethanol
oxidation on Pt(1 1 1)/Os is shown in Fig. 7. In this figure the
intensities of the bands at 933 and 1290 cm1 are plotted as a
function of the potential for Pt(1 1 1)/Os electrodes. According
to Fig. 7a, the formation of acetaldehyde starts at ca. 0.3 V on
Pt(1 1 1)/Os, and it is significantly enhanced after osmium modification, even with low osmium coverage, in comparison to the
value obtained on the nonmodified electrode. The highest inten-

2382

V. Pacheco Santos et al. / Electrochimica Acta 52 (2007) 23762385

on

al

co

py

of IR windows does not affect the experimental results. Additionally, this similarity of behavior allows to compare directly
the band intensities obtained with the two different windows
employed in this work. Similarly to observed in Fig. 5, the band
intensity of acetic acid is significantly enhanced after osmium
modification, even with low values of osmium coverage (0.19
and 0.28), when compared to the nonmodified electrode. On
Pt(1 1 1)/Os, the acetic acid band arises in the FTIR spectra of
ethanol oxidation at around 0.35 V. This band, however, becomes
higher only at potentials above 0.6 V, suggesting that most of
acetic acid comes from the oxidation of acetaldehyde produced
in the potential range of 0.30.6 V. Furthermore, according to
Fig. 7b, the highest intensity of the acetic acid band above 0.6 V
is reached with osmium coverage degrees of 0.40 and 0.51,
coinciding with the coverages in which the highest intensity
of acetaldehyde band is observed in Fig. 7a.
The acetic acid band is higher at lower potentials (up to 0.6 V)
on the highest osmium coverage studied, when compared with
other osmium coverages and with the nonmodified electrode.
However, the direct oxidation of adsorbed ethanol to acetic acid
at lower potentials on Pt(1 1 1) with Os = 0.92 is not evident,
due to the strong presence of acetaldehyde on this surface in the
potential range of 0.30.6 V.

th
o

r's

pe

rs

4. Discussion

Fig. 5. Band intensity of acetic acid (1290 cm1 ) as a function of the potential, from FTIR spectra for ethanol oxidation on nonmodified Pt(1 1 1) and on
Pt(1 1 1) modified with different osmium coverage degrees (data obtained from
Fig. 3); () nonmodified electrodes and () modified electrodes.

Au

sity of the acetaldehyde band is observed for osmium coverages


over 0.40. The fact that the bands of adsorbed COlinear and CO2
diminish on Pt(1 1 1) with higher osmium coverages than 0.28
(see Fig. 4), suggests that acetaldehyde is the main intermediate of ethanol oxidation on this electrode with Os = 0.40, 0.51
and 0.92. The intensity of this band is not affected even with an
almost complete osmium layer. However, this is in agreement
with the higher current density of ethanol oxidation observed on
electrode with Os = 0.40 and 0.51 (see Fig. 2).
A comparison of Figs. 7b and 5, reveals that the acetic acid
band intensity profiles measured using CaF2 or ZnSe windows
are very similar. This feature strongly supports that the change

According to the FTIRS results presented in this work, and


in agreement with previous studies [12,14,40], at least three
mechanistic pathways seem to be involved in the ethanol electrooxidation reaction on osmium modified Pt(1 1 1). Thus, due
to the fact that different species such as adsorbed COlinear , CO2 ,
CH3 CHO and CH3 COOH are detected during the ethanol oxidation on these surfaces, a simplified qualitative mechanism can be
proposed (see Fig. 8) within the potential and osmium coverage
ranges considered in this work.
In pathway 1 (Fig. 8), the ethanol molecule undergoes dissociative chemisorption (by breaking the C C bond) to form
adsorbed COlinear and CHx fragments. Most of the CO2 formed
comes from the adsorbed COlinear oxidation. The other portion
of the CO2 (given by the fraction) comes from a certain fraction () of the total amount of the previously formed fragments
CHx , that is converted to adsorbed COlinear and finally to CO2 .
According to the proposed mechanism, the total amount of carbon dioxide formed in pathway 1 is given by (1 + ) CO2 . The
amount of CHx that remains unchanged and adsorbed on the
electrode surface can be specified by the fraction (1 ) [40].
The fraction, ranging from zero to unity, generically represents
the qualitative fractional amount of the CHx specie that initially
forms adsorbed COlinear and subsequently CO2 . It is not related
to stoichiometric amount.
On Pt(1 1 1) modified with low osmium coverages, even
though the acetaldehyde formation is favored at lower potentials, pathway 1 is enhanced, that means, the amount of COlinear
and CO2 is more significantly enhanced than on nonmodified
electrodes. This, suggests that osmium promotes the C C bond
cleavage and favors the oxidation of CHx species to adsorbed
CO. Taking into account that the low osmium coverage and the

2383

co

py

V. Pacheco Santos et al. / Electrochimica Acta 52 (2007) 23762385

Au

th
o

r's

pe

rs

on

al

Fig. 6. In situ FTIR spectra (256 scans at 8 cm1 resolution, obtained in the wave number range of acetaldehyde and acetic acid bands with a ZnSe flat window)
of the series of positive potential steps for ethanol oxidation (0.5 M C2 H5 OH + 0.1 M HClO4 ) on: (a) nonmodified Pt(1 1 1) and on Pt(1 1 1) modified with different
osmium coverage degrees: (b) Os = 0.28, (c) Os = 0.51 and (d) Os = 0.92.

Fig. 7. Band intensity of: (a) acetaldehyde (933 cm1 ) and (b) acetic acid (1290 cm1 ) as a function of the potential, from FTIR spectra for ethanol oxidation on
nonmodified Pt(1 1 1) and on Pt(1 1 1) modified with different osmium coverage degrees (data obtained from Fig. 6); () nonmodified electrodes and () modified
electrodes.

V. Pacheco Santos et al. / Electrochimica Acta 52 (2007) 23762385

co

py

2384

Fig. 8. Scheme of simplified mechanistic pathways proposed for ethanol oxidation on Pt(1 1 1)/Os.

al

promotes CO2 formation by such a pathway is not interesting


for practical applications, due to the high potential necessary to
form significant amounts of CO2 .

on

5. Conclusions

The FTIRS results of ethanol oxidation presented in this


work, show that adsorbed COlinear and CO2 bands increase on
Pt(1 1 1) electrodes after spontaneous osmium deposition. The
highest intensity of these bands at lower potentials is reached
with osmium coverages of ca. 0.28 on Pt(1 1 1). On this surface,
the preferential oxidation of ethanol to adsorbed COlinear and
CO2 , can be related to the better capability of the Pt(1 1 1)/Os
system to cleave the C C bond of the ethanol molecule, when
compared to nonmodified platinum electrode.
The analysis of the band at 933 cm1 suggests that the oxidation of ethanol to acetaldehyde is enhanced on Pt(1 1 1) with
osmium coverages above 0.40. However, even with lower values of osmium coverage, a significant formation of acetaldehyde
is also observed at lower potentials. The better catalytic activity of Pt(1 1 1) towards the oxidation of ethanol to acetaldehyde
can be related to the easiness of osmium oxide formation, as
was observed in the voltammetric study in this paper and also
discussed in a previous study [32].
On Pt(1 1 1) modified with Os = 0.92, the direct oxidation of
ethanol to acetic acid at lower potentials is suggested, but this is
not clear due to the high presence of acetaldehyde on the surface
in the potential range of 0.30.6 V.
The high presence of CO2 with Os = 0.92, in contrast to the
very low quantity of adsorbed COlinear , suggests that the CO2
can also be formed from the oxidation of acetic acid at higher
potentials on the Pt(1 1 1) surface.
The voltamograms for ethanol oxidation on electrode with
osmium coverage of 0.19, 0.28 and 0.92 seem very similar, but
the reasons of these similarities, in terms of current intensities,
are essentially different. On the low coverage, 0.19 and 0.28,
the voltammetric current at potentials between 0.3 and 0.5 V
is maintained due to the oxidative adsorption of CO and oxidation to CO2 . While, for the higher osmium coverage, 0.92,
the voltammetric current intensity at potentials between 0.2 and

Au

th
o

r's

pe

rs

small amount of osmium oxide formed on Pt(1 1 1)/Os at lower


potentials, an electronic effect may also play a role in the catalytic activity of Pt(1 1 1)/Os surfaces for ethanol oxidation at
lower potentials.
Pathway 2 (Fig. 8) involves the formation of acetaldehyde,
which is oxidized to acetic acid at higher potentials [40]. On
Pt(1 1 1), the increase in the intensity of the acetaldehyde band
for higher osmium coverages is accompanied by the diminishing
of adsorbed COlinear and CO2 bands. Possibly, the formation of
osmium oxide on the surface may have an important role in the
preferential oxidation of ethanol to acetaldehyde. Additionally,
this suggests an inhibition of the C C bond cleavage.
In agreement with a previous study [32], the average size
of osmium nanoislands (ca. 48 nm2 for Os 0.16) diminishes
with increasing osmium coverage to a value of ca. 14 nm2 for
Os 0.45. This is due to the fact that, for higher degrees of
osmium coverage, the increase in the number of smaller islands
is much more significant than that observed for larger islands
[32]. In this way, the higher number of small islands formed at
Os 0.45, which exhibit high perimeter values in comparison
to their area, enhance significantly the catalytic activity of the
electrode for the oxidation of ethanol to acetaldehyde and acetic
acid. This occurs due to the increase of the active sites for organic
molecule oxidation, which is well-known to occur at the edge
of the islands [31].
Pathway 3 (Fig. 8) consists of the direct oxidation of ethanol
to acetic acid at lower potentials, without forming acetaldehyde
as an intermediate [40]. The increased presence of acetic acid
in the potential range up to 0.6 V on Pt(1 1 1) with Os = 0.92,
suggests the direct oxidation of ethanol to acetic acid at lower
potentials. But this is not complete sure due to the strong presence of acetaldehyde on this surface in the potential range of
0.30.6 V.
The observation that on Pt(1 1 1)/Os the intensity of the
CO2 band does not diminish above 0.8 V, as well as the discrepant quantity of CO2 formed on Pt(1 1 1) with Os = 0.92
when compared with the very low amount of adsorbed COlinear
(as observed in Fig. 4), suggest that CO2 can also be formed at
higher potentials by the direct oxidation of acetic acid present
in the thin electrolyte layer. However, developing a system that

V. Pacheco Santos et al. / Electrochimica Acta 52 (2007) 23762385

[19]
[20]
[21]
[22]
[23]

on

[24]
[25]

pe

r's

Acknowledgements

References

th
o

The authors acknowledge the Brazilian funding bodies


FAPESP, CAPES, CNPq and FAPEAL for the financial support.

Au

[1] T. Iwasita, W. Vielstich, in: H. Gersischer, C.W. Tobias (Eds.), Advances in


Electrochemical Sciences and Engineering, vol. 1, VCH, Weinheim, 1990.
[2] R. Parsons, T. Vandernoot, J. Electroanal. Chem. 257 (1988) 9.
[3] B. Beden, C. Lamy, J.-M. Leger, in: J.O.M. Bockis, B.E. Conway, R.E.
White (Eds.), Modern Aspects of Electrochemistry, vol. 22, Plenum, New
York, 1992, p. 97.
[4] J. Willsau, J. Heitbaum, J. Electroanal. Chem. 194 (1985) 27.
[5] B. Beden, M.-C. Morin, F. Hahn, C. Lamy, J. Electroanal. Chem. 229 (1987)
353.
[6] T. Iwasita, W. Vielstich, J. Electroanal. Chem. 257 (1988) 319.
[7] T. Iwasita, B. Rasch, E. Cattaneo, W. Vielstich, Electrochim. Acta 34 (1989)
1073.
[8] J.M. Perez, B. Beden, F. Hahn, A. Aldaz, C. Lamy, J. Electroanal. Chem.
262 (1989) 251.
[9] S.-C. Chang, L.-W.H. Leung, M.J. Weaver, J. Phys. Chem. 94 (1990) 6013.

py

[17]
[18]

co

[16]

B. Rasch, T. Iwasita, Electrochim. Acta 35 (1990) 989.


T. Iwasita, E. Pastor, Electrochim. Acta 39 (1994) 531.
J.F.E. Gootzen, W. Visscher, J.A.R. van Veen, Langmuir 12 (1996) 5076.
X.H. Xia, T. Iwasita, F. Ge, W. Vielstich, Electrochim. Acta 41 (1996) 711.
X.H. Xia, H.-D. Liess, T. Iwasita, J. Electroanal. Chem. 437 (1997) 233.
S.-G. Sun, in: J. Lipkowski, P.N. Ross (Eds.), Electrocatalysis, Wiley-VCH,
Weinheim, Germany, 1998, p. 243.
E. Mendez, J.L. Rodrguez, M.C. Arevalo, E. Pastor, Langmuir 18 (2002)
763.
D.J. Tarnowski, C. Korzeniewski, J. Phys. Chem. B 101 (1997) 253.
H. Hitmi, E.M. Belgsir, J.-M. Leger, C. Lamy, R.O. Lezna, Electrochim.
Acta 39 (1994) 407.
L.-W.H. Leung, S.-C. Chang, M.J. Weaver, J. Electroanal. Chem. 266
(1989) 317.
J. Shin, W.J. Tornquist, C. Korzeniewski, C.S. Hoaglund, Surf. Sci. 364
(1996) 122.
J. Wang, S. Wasmus, R.F. Savinell, J. Electrochem. Soc. 142 (1995) 4218.
N. Fujiwara, K.A. Friedrich, U. Stimming, J. Electroanal. Chem. 472 (1999)
120.
C. Bianchini, P. Bert, G. Giambastiani, S. Catanorchi, M. Piana, R. Psaro, S.
Santiccioli, A. Tampucci, F. Vizza, Proceedings of the International Symposium on Surface Imaging/Spectroscopy at the Solid/Liquid Interfaces,
Krakow, Poland, 2006, p. 11.
T. Frelink, W. Visscher, J.A.R.A. van Veen, Surf. Sci. 335 (1995) 353.
V.M. Schmidt, R. Ianniello, E. Pastor, S. Gonzalez, J. Phys. Chem. 100
(1996) 17901.
V. Del Colle, M.J. Giz, G. Tremiliosi-Filho, J. Braz. Chem. Soc. 14 (2003)
601.
J.P.I. Souza, F.J.B. Rabelo, I.R. de Moraes, F.C. Nart, J. Electroanal. Chem.
420 (1997) 17.
G. Tremiliosi-Filho, H. Kim, W. Chrzanowski, A. Wieckowski, B. Grzybowska, P. Kulesza, J. Electroanal. Chem. 467 (1999) 143.
A. Crown, I.R. Moraes, A. Wieckowski, J. Electroanal. Chem. 500 (2001)
333.
C.K. Rhee, M. Wakisaka, Y.V. Tolmachev, C.M. Johnston, R. Haasch, K.
Attenkofer, G.Q. Lu, H. You, A. Wieckowski, J. Electroanal. Chem. 554
(2003) 367.
V. Pacheco Santos, G. Tremiliosi-Filho, J. Electroanal. Chem. 554 (2003)
395.
V. Pacheco Santos, V. Del Colle, R.M. Bezerra, G. Tremiliosi-Filho, Electrochim. Acta 49 (2004) 1221.
Y. Morimoto, E.B. Yeager, J. Electroanal. Chem. 444 (1998) 95.
J.P.I. Souza, S.L. Queiroz, K. Bergamaski, E.R. Gonzalez, F.C. Nart, J.
Phys. Chem. B 106 (2002) 9825.
M. Watanabe, S. Motoo, J. Electroanal. Chem. 60 (1975) 267.
R. Ianniello, V.M. Schmidt, J.L. Rodrguez, E. Pastor, J. Electroanal. Chem.
471 (1999) 167.
N.A. Kolpakova, L.A. Shvets, A.G. Stromberg, Elektrokhim 28 (1992)
736.
O.A. Petrii, V.D. Kalinin, Russ. J. Electrochem. 35 (1999) 627.
R. Liu, H. Iddir, Q. Fan, G. Hou, A. Bo, K.L. Ley, E.S. Smotkin, Y.E.
Sung, H. Kim, S. Thomas, A. Wieckowski, J. Phys. Chem. B 104 (2000)
3518.
V. Pacheco Santos, V. Del Colle, R.B. de Lima, G. Tremiliosi-Filho, Langmuir 20 (2004) 11064.
P. Stonehart, J. Appl. Electrochem. 22 (1992) 995.
V. Radmilovic, H.A. Gasteiger, P.N. Ross Jr., J. Catal. 154 (1995) 98.
A. Rodes, J. Clavilier, J. Electroanal. Chem. 338 (1992) 317.
J. Clavilier, J. Electroanal. Chem. 107 (1980) 211.
K. Jaaf-Golze, D.M. Kolb, D. Scherson, J. Electroanal. Chem. 200 (1986)
353.
J.M. Orts, R. Gomez, J.M. Feliu, A. Aldaz, J. Clavilier, Electrochim. Acta
39 (1994) 1519.
T. Iwasita, X.H. Xia, J. Electroanal. Chem. 411 (1996) 95.
P. Gao, S.-C. Chang, Z. Zhou, M.J. Weaver, J. Electroanal. Chem. 272
(1989) 161.
J.L. Davis, M.A. Barteau, Surf. Sci. 235 (1990) 235.
J.L. Rodrguez, E. Pastor, X.H. Xia, T. Iwasita, Langmuir 16 (2000) 5479.

al

[10]
[11]
[12]
[13]
[14]
[15]

[26]
[27]
[28]

rs

0.5 V is maintained due to the formation of acetic acid, as can be


seen in the FTIR results presented in Figs. 5 and 7b. At higher
potentials, the discrepant amount of CO2 formed on coverage
of 0.92, in comparison of insignificant formation of adsorbed
CO, once more support the assumption that the formation of
CO2 by the route of oxidation of ethanol to acetic acid. This
supposition can explain the voltammetric profile of this surface
at high potentials. As it was concluded in the end of this section,
this proposed way for the formation of CO2 at higher potentials
on the osmium coverage of 0.92, does not give to this surface
status of a good catalyst, like those with 0.19 and 0.28 osmium
coverage, which follow preferentially the way to formation of
CO and CO2 at lower potentials.
Knowledge of the different mechanistic pathways followed
by the ethanol oxidation reaction on Pt(1 1 1)/Os electrode, as a
function of the osmium coverage degree, can provide important
information for application in direct alcohol fuel cell devices.
This is true as, depending on the composition, the intermediates and products formed during the organic molecule oxidation
can damage the catalytic activity of the electrode, as well as
diminish its useful life. In this sense, the most interesting feature of PtOs system, according to the FTIR results presented
in this work, is observed for low osmium coverages, due to the
more efficient oxidation of ethanol to CO2 . With the increase
of osmium coverage, the oxidation of ethanol to acetaldehyde
and acetic acid is favored on PtOs system. Though the catalytic
activity of the electrode is not damaged with higher values of
osmium coverage (up to ca. 0.5), as observed in the voltammetric studies, this path to produce acetaldehyde and acetic acid is
not an interesting feature of active surfaces for application in
fuel cell devices. This is due to the fact that few electrons are
involved in these reactions (only 2e per CH3 CHO molecule or
4e per CH3 COOH molecule, while the number of electrons for
the complete oxidation of ethanol to CO2 is 12). Thus, a higher
amount of ethanol would be necessary to obtain a given current
value, lowering the efficiency of the ethanol oxidation process.

[29]
[30]

[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]

[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]

2385

You might also like