Sie sind auf Seite 1von 12

Electrochimica Acta 47 (2002) 3663 /3674

www.elsevier.com/locate/electacta

Electrocatalysis of methanol oxidation


T. Iwasita
Instituto de Quı́mica de São Carlos, Universidade de São Paulo, Cx.P. 780, 13560-970 São Carlos, SP, Brazil

Received 10 January 2002; received in revised form 10 February 2002

Abstract

Most relevant aspects of methanol electrooxidation on Pt electrodes, including voltametric as well as spectroscopic data on the
system are presented and discussed. Parallel reaction pathways have been demonstrated, producing CO2 and HCOOH or HCHO as
soluble products, to an extent that depends on several parameters of the systems like methanol concentration, electrode roughness,
and time of electrolysis. Several issues like the intermediates of the reaction and the yields of oxidation products remain yet unclear.
A status of-the-art on these and other aspects of methanol oxidation is presented, which may be useful for future investigations. This
is the aim of the present contribution and referenced papers were accordingly selected. # 2002 Published by Elsevier Science Ltd.

Keywords: Methanol; Fuel cell; Electrocatalysis; Electrooxidation; FTIR

1. The methanol system

From the different small organic molecules (HCOOH,


H2CO, CH3OH), methanol is the one being most
intensively investigated at present. Methanol oxidation Both of these pathways require a catalyst, which
reaction has been the subject of a large number of should be able to (a) dissociate the C /H bond and (b)
studies in the past [1,2]. Early works [3 /6] already facilitate the reaction of the resulting residue with some
showed up a complex reaction mechanism, the electro- O-containing species to form CO2 (or HCOOH). On a
catalysis of methanol oxidation being the most difficult pure Pt electrode, which is known to be the best catalyst
task in the realization of a direct acid methanol fuel cell for breaking the C /H bond, complete oxidation takes
(DMFC). Since methanol oxidation reaction has been place via two processes occurring in separate potential
reviewed several times, e.g. [7 /10], in the present paper regions:
we intend just to present the most relevant aspects of the
/ The first process, involving adsorption of methanol
electroacatalysis of methanol oxidation, thus giving a
molecules, requires several neighboring places at the
status of-the-art on the system.
surface. Since methanol is not able to displace
The thermodynamic potential for methanol oxidation
adsorbed H atoms, adsorption can only begin at
to CO2, lies very close to the equilibrium potential of
potentials where enough Pt sites become free from H,
hydrogen: i.e. near 0.2 V versus RHE for a polycrystalline Pt
electrode.
/ The second process requires dissociation of water,
CH3 OHH2 O 0 CO2 6 H 6e E  0:02 V
which is the oxygen donor of the reaction. On pure Pt
electrode, a strong interaction of water with the
catalyst surface is only possible at potentials above
However, compared with hydrogen oxidation, this 0.4 /0.45 V versus RHE.
reaction is by several orders of magnitude slower. As
early suggested by Breiter [5], the total oxidation process Thus, on a pure Pt catalyst methanol oxidation to
consists of a pattern of parallel reactions which can, in CO2 cannot begin below, say 0.45 V. However, the
principle, be formulated as follows: adsorbate layer does not exhibit a good reactivity below
0013-4686/02/$ - see front matter # 2002 Published by Elsevier Science Ltd.
PII: S 0 0 1 3 - 4 6 8 6 ( 0 2 ) 0 0 3 3 6 - 5
3664 T. Iwasita / Electrochimica Acta 47 (2002) 3663 /3674

approximately 0.7 V, i.e. a high rate of oxidation at pure recording of mass (m /e/44) corresponding to CO2 is
Pt occurs at potentials without technological interest. given. Thus the current peak can only be due to
Faradaic processes occurring during methanol adsorp-
tion.
2. Methanol adsorption

It was suggested that methanol adsorption takes place


in several steps, forming different species due to
dissociation of the molecule: 3. The nature of the adsorbed methanol species

CH3 OH 0 C H2 OHHe 0 C HOHHe Establishing the nature of the adsorbed species


x xx

0 C OHHe 0 C OHe   formed during the adsorption of small organic mole-


xxx x cules is a difficult task. The issue was approached in a
where x stands for a Pt site [6]. It was suggested that number of studies over many years [10]. Except from a
formaldehyde and formic acid could be formed from the study using gas chromatography [5] earlier papers were
intermediates CH2OH and CHOH, respectively [6]. mainly based in data of charge measurements during
If a cyclic voltammogram is started after contacting a adsorption of methanol and oxidation of the adsorbed
polycrystalline Pt electrode with a methanol containing residue [2,4,12]. The use of analytical methods for the in
solution at a potential of 0.05 V or less, methanol situ, ex situ and on line analysis of the electrode surface
adsorption can be observed as soon as hydrogen cover- began in the decade of 1980 [7]. Different adsorbed
age decreases to a certain extent. The dissociation species were suggested on the basis of data from infrared
process gives rise to a current peak in the H-region spectroscopy [13,14], thermal-desorption MS [15] and
(Fig. 1), which can be observed only during the first DEMS [7].
potential scan, i.e. when the surface is free from organic Infrared spectra obtained during methanol adsorption
residues. The experiment in this figure was performed at 0.35 V on polycrystalline Pt show well characterized
using the DEMS technique [11]. Briefly, in this techni- bands for linearly adsorbed CO at approximately 2040
que the electrode is a porous Pt layer on a PTFE cm 1 together with other bands in the 1200/1300 cm 1
membrane, lying on a porous plate at the entrance of a region, which have been interpreted in terms of the C /
mass spectrometer. This setup allows the entrance of OH stretching of an hydrogenated species (as, e.g. COH
any volatile product in the MS in fractions of a second [14] or HCOH). These bands were also observed in
after being produced. During the experiment in Fig. 1, spectra obtained on single crystal Pt(100) and Pt(111)
the recording of mass signals did not show any volatile (see below).
product from methanol oxidation. As an example, the Another approach to establish the nature of the
adsorbate was made by thermal desorption mass
spectrometry, performed on electrodes transferred into
UHV [15]. These results also confirmed the presence of
hydrogenated species after adsorption of methanol.
Moreover, these data have shown that the ratio between
the amount of CO and (hydrogenated) species depends
on methanol concentration [15]. It should be noticed
that large discrepancies of results among different
groups may lie, at least in part, in the experimental
approaches. The necessity of thoroughly eliminating
traces of oxygen from the solution and from the gas
atmosphere above the solution must be emphasized,
since adsorbed oxygen can interact with organic resi-
dues. Recall that O2 reduction on a Pt surface produces
adsorbed peroxide species, which can act as an oxidizing
agent for species present at the surface. Other sources of
discrepancies may originate in the fact that the reaction
is surface sensitive and, consequently, results for poly-
Fig. 1. First potential scan for a porous polycrystalline Pt electrode in crystalline Pt may depend on the method of surface
0.1 M CH3OH/0.05 M H2SO4 solution (upper part) and simulta-
neouly recording of mass intensity for CO2 production (lower part); 10
pretreatment, electrode roughness and the time scale of
mV s 1. Dashed lines: current and MS signals in supporting the experiments, since adsorbed species can undergo
electrolyte. slow transformations.
T. Iwasita / Electrochimica Acta 47 (2002) 3663 /3674 3665

4. Methanol oxidation products Childers [22]. This could be the reason why formalde-
hyde remained almost disregarded in the methanol fuel
The oxidation products of CH3OH are well known cell literature. Korzeniewski and Childers determined
since early works of Pavela [16] and Schlatter [17]. These formaldehyde yields fluorometrically after applying
authors used long-term electrolysis at potentials between different constant potentials on a smooth polycrystalline
0.5 and 0.6 V versus RHE and found CO2, HCHO, Pt electrode, during 5 min in a micro cell. They report
HCOOH and HCOOCH3. The latter product, methyl for formaldehyde a yield of 38% under following
formate, originates in a reaction: conditions: 0.25 V versus Ag/AgCl (ca. 0.48 V vs.
RHE), 15 mM CH3OH/0.1 M HClO4. The yield
HCOOH CH3 OH HCOOCH3 H2 O
decays at higher potentials [22]. On porous Pt electrodes,
The yields of oxidation products depend on methanol Wang et al. found at 0.65 V versus RHE 50% of HCHO,
concentration, temperature, electrode roughness and 34% of HCOOH and only 16% of CO2. It is worth
time of electrolysis [18,19]. noting that Ota et al. also found relatively high yields of
The study of the products of methanol oxidation HCHO on platinized Pt electrodes at 0.6 V versus RHE
during a potential scan was the first goal of on-line mass [19].
spectrometry, DEMS [20]. In Fig. 2, the potentiody-
namic formation of CO2 and methyl formate on a Pt
electrode was followed during the potential scan by 5. Structural dependence of methanol oxidation
recording the corresponding ion currents in the MS: (m /
e /44) and (m /e /60), respectively. No mass signals for 5.1. Results of cyclic voltammetry
HCHO were observed, but a weak ion current for
methylal (CH2(OCH3)2), indicated its formation via As shown by the cyclic voltammograms in Fig. 3 for
reaction of HCHO with CH3OH [7]. However, there the three low index surfaces of platinum, adsorption and
must be some problem with the volatility of formalde- oxidation of methanol present a strong sensitivity to the
hyde or its hydration product in aqueous solution (gem surface structure [23]. The Pt surfaces in these experi-
diol, CH2(OH)2), which makes somehow difficult its ments were contacted with methanol at 0.05 V (a
direct detection using the DEMS technique [21]. These potential where methanol adsorption is negligible) and
difficulties extend to other modern analytical methods then the first potential sweep was recorded. From the
like in situ FTIR as pointed out by Korzeniewski and three surfaces, Pt(100) is the only one presenting a well
defined current peak at 0.35 V for methanol dissociative
adsorption. This process is superimposed on the current
for H-desorption (compare with Fig. 1). Almost no
activity towards methanol oxidation is observed until
the potential reaches approximately 0.72 V. Contrasting
with this result, no indication of dissociative adsorption
of methanol parallel to hydrogen desorption is observed
at Pt(111). On the other hand, at Pt(110) the lowering of
current in the H region may indicate that the methanol
adsorption begins already at the initial potential of the
voltammogram (0.05 V). This behavior is expected from
the low potential for H desorption in this surface. For a
comparison, the first part of the potential scan is
magnified in Fig. 4. Here, one sees that a CV for
Pt(111) shows up the highest oxidation current in the
potential range between 0.45 and 0.65 V, followed by
Pt(110) and Pt(100). However, before establishing a
ranking for the activity of the three surfaces, infrared
results should be taken into consideration, as we shall
do in the coming section.

5.2. Results of infrared spectroscopy

Infrared spectra during methanol adsorption and


Fig. 2. DEMS experiment: current and mass signals (ion current) for
volatile products during methanol oxidation at porous polycrystalline oxidation are shown in Fig. 5, for Pt(111), Pt(100) and
Pt, surface roughness: ca. 50; 0.1 M CH3OH/1 M HClO4; 20 mV s 1. Pt(110) [23]. Bands at approximately 2060 and 1850
[7]. cm 1 are assigned to linearly (COL) and bridge (COB)
3666 T. Iwasita / Electrochimica Acta 47 (2002) 3663 /3674

Fig. 4. Comparison of current for methanol oxidation on Pt(hkl ).


Conditions as in Fig. 3.

of Pt(111) and Pt(110) highlights some interesting


features of the mechanism of methanol oxidation. We
let at first Pt(100) out of consideration because of the
complication of presenting two forms of adsorbed CO
(on top and bridge), thus making it difficult to estimate
relative coverages on the basis of band intensities [24]. In
the case of Pt(111) the intensity of the band for COB is
Fig. 3. First potential scan for methanol oxidation on Pt(hkl ); 1.0 M so small that we shall simply neglect it in the following
CH3OH in 1.0 M HClO4; sweep rate: 50 mV s 1 [21]. discussion.
Measurable amounts of CO are observed at Pt(110)
bonded CO, respectively. The band at 1260 cm 1 was already at 0.1 V. The CO signal rapidly grows with
assigned to some H containing intermediate, possibly potential and reaches a maximum at around 0.3 V. At
COH [14]. Other H-containing adsorbate may be Pt(111), the initial adsorption can be extrapolated to
responsible for the feature at 2950 cm1 in the spectra around 0.2 V, judging from the band for linear bonded
for Pt(100). This band, which can be assigned to the C / CO. This feature grows somewhat slower than that for
H stretching of a CH2 group appears already at 0.2 V Pt(110). Also, a weak band for bridge bonded carbon
and becomes better defined at more positive potentials. monoxide is apparent at potentials above 0.35 V. But
All spectra exhibit a band at 2341 cm 1, due to the the most important observation here, is that the
oxidation product CO2. Although not shown here, a dissociative adsorption of methanol at Pt(111) takes
weak band at 1230 cm 1 has been observed on Pt(111) place at potentials within the H region. Therefore we
at potentials above 0.45 V, which was assigned to the can conclude that the lack of methanol adsorption in the
C /O /C stretching of methyl formate [23]. H-region during the potential scan at 50 mV s 1 (Fig. 3)
Surface sensitive effects become more evident when is due to a kinetic limitation of the adsorption process at
analyzing the integrated band intensities for CO and Pt(111) which is not observed on the other two surfaces.
CO2 for all three surfaces shown in Fig. 6a and b. In In Fig. 6b is presented the integrated band intensity
comparing these results with those of the cyclic voltam- for the CO2 observed in the potential region below 0.8
mograms, one has to be aware that the time scale for the V. We can state that in this region CO2 production
spectra is much larger than for the CV. Thus, the proceeds with almost the same intensity at both Pt(111)
voltammetric curves in Fig. 4 were taken at 50 mV s 1, and Pt(110) and is much lower at Pt(100). Both former
i.e. 0.02 s mV 1). On the other hand, spectra collection surfaces have almost the same catalytic activity for CO2
requires about 70 s at each potential and considerable production, at least at potentials below approximately
adsorption can take place during this time. The behavior 0.7 /0.75 V. However, the relative oxidation currents
T. Iwasita / Electrochimica Acta 47 (2002) 3663 /3674 3667

Fig. 5. Infrared spectra in 1.0 M CH3OH in 0.1 M HClO4 at Pt(100), Pt(111) and Pt(110). Reference spectra collected at 50 mV; sample spectra
collected at the indicated potentials [21].

observed in the voltammograms of Fig. 4 for Pt(110) slow progress along many years can be easily under-
and Pt(111) indicate a higher activity of the latter stood in the light of the difficulties created by the
towards methanol oxidation. We can thus conclude, existence of parallel reaction paths with yields depend-
that in the interval of potentials between, say, 0.4 and ing on potential, time, surface structure, etc. At present,
0.6 V, the potenciodynamic current at Pt(111), origi- only two global processes are distinguishable as already
nates to a large extent from the dissociative adsorption said in the Introduction namely, adsorption of methanol
of methanol and/or from the parallel pathway men- molecules and oxidation of adsorbed residues. Analizing
tioned in the Introduction, ending in other products the IR results presented above it is possible to follow the
than CO2. Our present knowledge of the yields of pathway leading to CO2 via formation of adsorbed CO
HCOOH and/or HCHO at single-crystal electrodes is and to identify for this pathway, whether adsorption of
insufficient for establishing the extent to which these methanol or removal of CO is the rate determining
parallel pathways contribute to the current. Also Pt(100) process.
presents evidences for high yields of other products via a Due to the fact that the interface components of the
parallel reaction: the pronounced increase of current at respective Pt surfaces depend on potential and admitting
0.7 V (Fig. 4), contrasts with the moderate variation of
the necessity of oxygen donor species at the surface, for
the CO2 band intensity in Fig. 6b.
oxidation of the adsorbed residues, discussing the total
oxidation is only meaningful at potentials above 0.4 V
or higher. However, what the adsorption process con-
6. On the mechanism of methanol oxidation cerns, we should start the discussion by considering the
methanol adsorption at potentials below 0.4 V. It is well
The mechanism of methanol oxidation is an issue known that methanol cannot displace adsorbed hydro-
which can be considered to be at its very beginning. The gen from the Pt surface. This is a well known phenom-
3668 T. Iwasita / Electrochimica Acta 47 (2002) 3663 /3674

according to the voltammograms of these surfaces


measured in HClO4 (Fig. 7). Thus, e.g. for Pt(110) at
0.2 V, uH /0.2 and the band intensity for CO is large
(about 80% of its maximum value). On the other hand,
for the same uH at Pt(111), (E /0.29 V), the band
intensity is much lower. In fact, it grows slowly even
beyond 0.4 V, where the surface is totally free from
adsorbed hydrogen. It can thus be stated that the
differences in methanol adsorption rate for these two
surfaces originates in some property inherent to the
surface itself and not in the availability of H-free sites:
adsorption of methanol is intrinsically a slow process at
Pt(111).
We proceed now to discuss the potential region above
0.4 V, where CO2 is produced. Considering that
independent of the respective CO coverage, the rate of
CO2 formation is the same for both surfaces, we can
state that the reaction rate for CO2 formation is the
same for both surfaces Pt(111) and Pt(110). The most
simple interpretation of this result is that the oxidation
of the adsorbed residue is the rate determining process.
Or, with other words, at all potentials above 0.4 V, there
is enough adsorbed methanol on both surfaces as to
reach the maximum rate of oxidation at the respective
potential. Also Gasteiger et al. [25] suggested that
oxidative removal of CO is the rds in this potential
region. Following a somewhat different approach,
Christensen et al. [26] arrived to the same conclusion
for polycrystalline Pt. Summarizing, oxidation of ad-
Fig. 6. Integrated band intensities from IR spectra at Pt(111) and sorbed CO and not adsorption of methanol is likely to
Pt(100) and Pt(110) in 1.0 M CH3OH in 0.1 M HClO4. (a) Linear be the rate determining process in the potential region
(COL) and bridge (COB) bonded carbon monoxide; (b) carbon dioxide. between, say 0.5 and 0.70 V.
It has been shown that interaction of water with the
enon and is now documented by the data in Fig. 6a, Pt surface increases as the potential is made more
where the structure and potential dependence of metha- positive [27] and competition of methanol with water
nol adsorption can be analyzed from the band intensity for adsorption sites becomes important. Therefore, at
for adsorbed CO. high potentials (above, say, 0.7 V) methanol adsorption
The above data were obtained by applying potential becomes again rds [28] and for this reason the reaction
steps of approximately 70 s of duration, i.e. they are rate passes through a maximum and then decays.
neither stationary nor dynamic, but since the procedure For the pathway analyzed here, leading to CO2
for collecting spectra was the same for all surfaces, it formation via adsorbed CO, we can state that CO is
should be valid to compare with the band intensities in
order to have an idea of the rate of methanol adsorption
in the low potential region. Thus, it can be stated that
adsorption of methanol occurs at a higher rate at
Pt(110) than at Pt(111) (Fig. 6a). While COL coverage
at Pt(110) rapidly increases reaching a saturation value
between 0.3 and 0.4 V, the band for COL at Pt(111)
grows more slowly and does not present a real max-
imum but a sudden falling of intensity at a potential of
0.6 V. Additionally, in the whole range of potentials the
CO band intensity is substantially higher on Pt(110)
than on Pt(111). More about the methanol adsorption
Fig. 7. Potentiodynamic current /potential response for between 0.05
process can be learnt by analyzing in parallel the band and 0.5 V vs. RHE for Pt(111) and Pt(100) in 0.1 M HClO4; sweep rate
intensity for CO in Fig. 6a and the hydrogen coverage 50 mV s 1.
T. Iwasita / Electrochimica Acta 47 (2002) 3663 /3674 3669

an intermediate of the reaction and the role of surface metal. Many O-adsorbing metals can produce negative
poison usually ascribed to COad should be revised. effects, e.g. inhibit methanol adsorption or may be not
Inspection of Fig. 6 indicates that CO indeed accumu- sufficiently stable for long-term use, as required for a
lates on the surface at low potentials (see the result for fuel cell. At present, there is a general consensus that
Pt(110)), however, the reason for the lack of CO2 PtRu offers the most promising results. Methanol
formation lies in the inability of Pt to dissociate water oxidation on PtRu binary catalysts has been the matter
and not in the degree of surface blocking by CO. of a number of papers [25,35/54]. The catalytic effect
Otherwise it would be difficult to understand that two has been observed using different kinds of PtRu
surfaces covered with CO to different extent produce materials, such as PtRu alloys, PtRu electrodeposits,
CO2 at identical rates. Ru evaporated on Pt, Ru adsorbed on single-crystal
Concerning the nature of the oxygen donor, it was Pt(hkl ), etc. and on technical (carbon supported electro-
originally suggested by Gilman that it is an adsorbed des) as well [54,55].
OH species coming from water dissociation [29]. But, When discussing the reason for the catalytic effect of
according to Wieckowski et al. [30] the oxygen donor is PtRu, the bi-functional mechanism is often invoked [41].
simply some activated water molecule adsorbed on the The term was suggested to give emphasis on the joint
Pt surface. For oxidative stripping of CO adlayers on activities of both metals, Pt being the one adsorbing and
platinum, Koper et al. [31] have shown that the dissociating methanol and Ru, the one oxidizing the
dissociation of water is a necessary step in order to adsorbed residues at low potentials. This description of
harmonize experimental data with results of Monte the mechanism is based in the observation that at
Carlo simulations. Although the real nature of the potentials below 0.4 V, Pt is a good catalyst for
oxygen donor for the oxidation of COad was not methanol adsorption, but not for water dissociation
demonstrated via spectroscopic methods, we are in- while Ru is able to dissociate water but it cannot adsorb
clined to accept the more or less general consensus that methanol. However, establishing a role for each metal as
it is OHad formed through the dissociation of water. in the bi-functional mechanism is of limited use, since it
is well known that at high potentials Pt dissociates water
6.1. Supporting electrolyte effects on the rate of methanol and, as shown in refs. [43,53] at high temperatures (60,
oxidation 80 8C) Ru adsorbs methanol. Moreover, even for
conditions where methanol adsorption occurs only on
Perchloric and sulfuric acids are the commonly used Pt, CO can move on the surface and occupy sites on Ru
supporting electrolytes for studies of methanol electro- atoms. Altogether, several adsorbed species could be
oxidation. It is, however noteworthy, that methanol involved in the oxidation process at the PtRu catalyst,
oxidation can be affected by the anion of the supporting namely, Pt(CO)ad, Ru(CO)ad, Ru(OH)ad and Pt(OH)ad.
electrolyte used. This is clearly the case for Pt(111) and In a simplified manner, we shall describe the bi-
Pt(100) electrodes, where cyclic voltammograms of functional mechanism as follows [25].
methanol exhibits much larger currents in HClO4 than The first step of the reaction is adsorption of
in H2SO4 [32,33]. The difference was justified by the methanol:
specific adsorption of sulfate species at Pt electrodes,
CH3 OH(sol) 0 (CO)ad 4 H 4 e (1)
which partially hinder methanol oxidation. Kita et al.
[32] observed for Pt(111) a factor of ten higher current in (CO)ad represents an adsorbed CO species either on Pt
HClO4 than in H2SO4, while for Pt(100) the enhance- or on Ru. Then, both Pt and Ru dissociate water to
ment factor was about two. The different behavior can form adsorbed OH:
be rationalized in terms of a stronger specific anion
RuH2 O 0 (OH)ad H e (2)
adsorption at Pt(111) than at Pt(100) [34]. No difference
between HClO4 and H2SO4, was observed at Pt(110) PtH2 O 0 (OH)ad H e (3)
[32]. Finally, following a Langmuir /Hinschelwood me-
chanism adsorbed CO reacts with adsorbed OH to
give CO2:
7. Catalyst promoters for methanol oxidation
(CO)ad (OH)ad 0 CO2 H e (4)
Several binary and ternary catalysts were proposed For CO adlayers obtained via adsorption of dissolved
for methanol oxidation, most of them based in mod- CO on PtRu, Koper et al. [56] analyzed reaction (4) for
ifications of Pt with some other metal. This metal must all possible species mentioned above and found that an
fulfill the requirement of forming O-containing surface enhanced effect is only possible if the final oxidation
species at low potentials. Among others, Sn, Re, Ru, Ge step occurs between CO adsorbed at Pt and OH
and Mo, were suggested [7,25,35 /54]. There are, of adsorbed at Ru. Therefore, reaction (4) can be specifi-
course, several practical factors limiting the choice of the cally written as
3670 T. Iwasita / Electrochimica Acta 47 (2002) 3663 /3674

Pt(CO)ad  Ru(OH)ad 0 CO2 H e (5)

Moreover, as a necessary condition for the


Langmuir /Hinschelwood mechanism, CO must diffuse
on the surface to the places where the OHad partner is
formed. Koper et al. [56] suggested that the relative rate
of CO diffusion on the surface must be high.
Besides the promoter effect of Ru through reaction
(5), there are experimental evidences of additional
effects of Ru on the reaction. Thus, the potentiodynamic
curves (first sweep) of Fig. 8, obtained at PtRu electro-
deposits in the course of DEMS experiments, show a
negative shift of the dissociative adsorption of methanol
depending on the Ru content of the electrodes. Further
evidence of the earlier adsorption was obtained via
infrared spectroscopy as shown in Section 5.1.
The catalytic effect of PtRu materials is usually
presented in the form of current /time plots at constant
potential, as observed in Fig. 9, [46,50,51]. Similar
responses have been reported by other authors under
comparable conditions [43,47]. In order to measure the
data represented in this plot Pt(111)/Ru electrodes were
prepared by adsorption of Ru onto Pt(111) [46] and
alloys were cleaned by sputtering and heating in UHV
following the pre-treatment technique suggested by
Gasteiger et al. [56]. This is a reliable method for
producing smooth surfaces of identical composition as Fig. 9. Current /time curves for comparing the catalytic activity of
the bulk. The quality of having a low roughness factor Pt(111) and PtRu alloys, towards methanol oxidation in 0.5 M
can be easily checked via comparison of the current in CH3OH/0.1 M HClO4. Potential 0.5 V vs. RHE. Room temperature,
voltammograms in base electrolyte measured with the electrodes surfaces cleaned in UHV.
alloy with those obtained using Pt(111)/Ru electrodes
[46].
As observed in Fig. 9, when smooth PtRu materials
are polarized at constant potential in methanol solu-
tions, the current decays continuously indicating a
pronounced loss in activity. The current does not reach
a stationary state even after several hours. Two different
origins of this effect causing electrode deactivation have
been identified. One of these is reversible and seems to

Fig. 10. Plot of the current density for methanol oxidation as function
Fig. 8. First potential scan in 1.0 M CH3OH/0.5 M H2SO4, starting of Ru coverage, taken from current /time curves at 0.5 V as in Fig. 9.
at 50 mV (RHE); scan rate 1 mV s 1. Porous electrodes prepared by Data for UHV prepared PtRu alloys, measured after 20 min; data for
depositing Pt and Ru at 0.2 V vs. RHE up to a total charge of 1200 Pt(111)/Ru formed by spontaneous adsorption, measured after 5 min.
mC, from solutions containing x mM RuCl3 and y mM H2Cl6 Pt in Solutions: 0.5 M H2SO4/0.5 M CH3OH (data form ref. [43]), 0.5 M
0.5 M H2SO4. The x /y ratios are indicated in each curve. CH3OH/0.1 M HClO4 [46].
T. Iwasita / Electrochimica Acta 47 (2002) 3663 /3674 3671

active as oxygen donors for CO oxidation. The electrode


activity is thus partially recovered by applying a
potential step towards more negative values, where the
ruthenium oxide species are reduced. The other factor
causing the decay of current is, apparently, a blockage
of the surface by some organic residue, which is slowly
formed and can only be oxidized at high anodic
potentials. No spectroscopic proofs are yet available
on the nature of such blocking species.
The results of i /t curves obtained at room tempera-
ture for the two types of materials (alloys and Pt(111)/
Fig. 11. Current densities for methanol oxidation at 0.4 V in 0.5 M
Ru) are collected in Fig. 10, where the current after 20
CH3OH/0.5 M H2SO4 vs. Ru surface composition of sputter cleaned
Pt /Ru alloys. Electrode immersion at 0.075 V for 3 min prior to min of polarization at 0.5 V versus RHE is plotted as a
stepping to the indicated potential. Dashed lines are arbitrarily drawn function of the Ru:Pt surface composition. It is note-
smooth curves to connect the experimental data points. Data from ref. worthy that in spite of the fact that the data for the
[43], published with permission.
alloys in this figure were measured in two different
laboratories [43,46], they nicely fit together. Also,
be caused by the oxidation of the Ru surface [50,51], methanol oxidation at PtRu alloys seems not to be the
forming oxides like RuO2 and RuO3, which are not same for both supporting electrolytes H2SO4 and HClO4

Fig. 12. Comparison of in situ FTIR spectra for Pt(111), Pt(111)/Ru 39% and PtRu alloy (85:15) in 0.5 M CH3OH/0.1 M HClO4. Potentials as
indicated on each spectrum; reference spectrum taken at 0.05 V (from [46] with added spectra at 0.55 V).
3672 T. Iwasita / Electrochimica Acta 47 (2002) 3663 /3674

used. Two main points can be extracted from this plot:


(i) PtRu alloys are better catalysts than Pt(111)/Ru, and
(ii) both materials present a wide maximum (between ca.
10 and 40% Ru for the alloys and ca. 15 and 50% for the
Pt(111)/Ru electrodes) at room temperature.
In terms of the bi-functional mechanism, these max-
ima indicate that a Ru percentage of approximately 10/
45% on the surface is enough to provide an efficient
oxidation of adsorbed methanol residues. Within this
range one has a possibility of studying the influence on
the reaction of other parameters (like e.g. methanol
concentration) and possibly, the kinetic of the reaction.
The upper limit of Ru coverage, is given by the necessity
of having enough Pt sites for adsorbing and dissociating
methanol. This limit is somewhat higher for adsorbed
Ru than for the alloy (observe the wider maximum).
This and the higher currents observed on the alloys
may be related to the fact that the latter present a
more homogeneous distribution of Ru atoms than the
Pt(111)/Ru electrode [57]. STM data of Pt(111)/Ru show
that Ru tends to form aggregates (islands) on the surface
of Pt [46,58]. Thus, for the same Ru percentage, wider
Ru patches can be found on the Pt(111)/Ru surface than
on the alloy. It is noteworthy that the maximum of the
j /uRu plot for the alloy is shifted to higher uRu values at Fig. 13. DEMS experiment on PtRu porous electrode in 1.0 M
CH3OH/0.5 M H2SO4; v : 20 mV s 1. Electrode prepared by
higher temperatures (Fig. 11) [43]. This effect probably
electrodepositing Pt and Ru during 7 min at 0.05 V, on a gold
originates from the fact that at higher temperatures (e.g. substrate. Solution composition: 10 mM RuCl3/10 mM H2Cl6Pt in
60 8C), Ru becomes active for adsorbing methanol and 0.5 M H2SO4 [45] published with permission.
oxidizing the residue (see the value of current for pure
Ru in Fig. 11). production and also a high CO coverage (near satura-
tion) at 80 8C.
It is noteworthy that, although no features for other
8. Spectroscopic results of methanol oxidation on PtRu intermediates or soluble products were observed under
materials the conditions for taking the spectra on the PtRu
materials shown in Fig. [12], DEMS experiments in 1.0
Infrared spectra using Pt(111), Pt(111)/Ru and a PtRu M CH3OH using PtRu electrodes do exhibit a potential
alloy [46] are presented in Fig. 12. PtRu alloy electrodes dependent mass signal for HCOOCH3 (Fig. 12) [45]. As
were UHV cleaned, as for the experiments of Fig. 9. The pointed out before, the yields of different products on a
bands observed on both Pt(111)Ru and PtRu (85:15) Pt electrode depend on several experimental parameters
alloy correspond to CO2 (2341 cm 1) and COL (2050 such as methanol concentration and electrode roughness
cm 1). At Pt(111) a band at approximately 1820 cm 1 [19]. This condition seems to extent to PtRu materials.
is due to bridge adsorbed (COB) (Fig. 13). So far, IR spectra on well prepared PtRu alloys in
The PtRu alloy electrode presents the highest produc- concentrated methanol solutions (above 0.5 M) have not
tion of CO2 at a given potential (note the differences in been published and the question on the parallel paths on
scale). It is also noteworthy, that although the CO2 band PtRu should be let open to further discussion.
markedly grows with potential and also the highest band
intensity for linear bonded CO. At PtRu the CO feature
remains approximately constant in the range of poten-
tials between 0.3 and 0.55 V, indicating that the 9. Concluding remarks
intermediate CO reaches a stationary coverage balan-
cing the rates of formation and oxidation. In the last years, considerable progress has been made
In a recent paper [53], methanol oxidation was studied in the knowledge of methanol electrooxidation reaction
on two PtRu alloys having 10 and 90% Ru, in the range and on its catalysis.
of temperatures between 25 and 80 8C. For the 90% Ru As for other small organic molecules, parallel path-
alloy, this study shows a negligible activity towards ways occur during methanol oxidation. In addition to
methanol oxidation at 25 8C but an appreciable CO2 CO2 which is the major product, formic acid (methyl
T. Iwasita / Electrochimica Acta 47 (2002) 3663 /3674 3673

formate) and formaldehyde are formed, their yields [13] B. Beden, C. Lamy, A. Bewick, K. Kunimatsu, J. Electroanal.
Chem. 121 (1981) 343.
depending on experimental conditions such as surface
[14] T. Iwasita, F.C. Nart, J. Electroanal. Chem. 317 (1991) 291.
roughness and methanol concentration and time of [15] S. Wilhelm, T. Iwasita, W. Vielstich, J. Electroanal. Chem. 238
electrolysis. (1987) 383.
In spite of noticeable structural effects of the Pt [16] T.O. Pavela, Ann. Acad. Sci. Fenn. Ser. K A2 (1954) 1.
surface, which are reflected on the amount and rate of [17] M.J. Schlachter, in: G.Y. Yong (Ed.), Fuel Cells, Reinhold, New
York, 1963, p. 199.
surface coverage by CO, comparable ‘quasi-stationary’
[18] M. Breiter, Disc. Faraday Soc. 45 (1968) 79.
rates of CO2 production are observed at Pt(111) and [19] Y. Ota, M. Nakagawa, Takahashi, J. Electroanal. Chem. 179
Pt(110) in the potential region between approximately (1984) 179.
0.4 and 0.75 V. This result indicates that the rate [20] (a) O. Wolter, C. Giordano, J. Heitbaum, W. Vielstich, Proceed-
determining process is more likely the oxidation of the ings of Symposium on Electrocatalysis, The Electrochemical
adsorbed species and not the adsorption of methanol. Society, Pennington, 1982, p. 235;
(b) O. Wolter, J. Heitbaum, J. Electrochem. Soc. 132 (1985) 1635.
At room temperature PtRu alloys having a Ru [21] H. Wang, T. Löffler, H. Baltruschat, J. Appl. Electrochem. 31
content between 10 and 40% are, at present, the best (2001) 759.
catalysts for the reaction. This is probably due to a good [22] C. Korzeniewski, C. Childers, J. Phys. Chem. B 102 (1998) 489.
distribution of Pt and Ru atoms as compared with other [23] X.H. Xia, T. Iwasita, F. Ge, W. Vielstich, Electrochim. Acta 41
materials where the atoms tend to segregate forming (1996) 711.
[24] I. Villegas, M.J. Weaver, J. Chem. Phys. 101 (1994) 1648.
patches of pure Ru or Pt, thus physically separating the [25] H. Gasteiger, N. Markovic, P.N. Ross, E.J. Cairns, J. Phys.
bi-functional partners. Experimental data show that in Chem. 97 (1993) 12020.
addition to the bi-functional action of both metals, the [26] P.A. Christensen, A. Hamnett, G.L. Troughton, J. Eletroanal.
promoter effect of Ru may also lie in other effects, such Chem. 362 (1993) 207.
as lowering the potential for methanol adsorption. [27] X.H. Xia, T. Iwasita, J. Electroanal. Chem. 411 (1996) 95.
[28] T. Iwasita, X.H. Xia, H.-D. Liess, W. Vielstich, J. Phys. Chem. B
Some deactivation effects are observed during polar- 101 (1997) 7542.
ization at constant potential, the origin of these being [29] S. Gilman, J. Phys. Chem. 68 (1964) 70.
probably, the formation of inactive ruthenium oxides of [30] H. Kim, I. Rabelo de Moraes, G. Tremiliosi-Filho, R. Haasch, A.
and blockage of the surface by organic residues of Wieckowski, Surf. Sci. 474 (2001) L203.
unknown nature. [31] M.T.M. Koper, A.P.J. Jansen, R.A. van Santen, J.J. Lukkiens,
P.A.J. Hibers, J. Chem. Phys. 109 (1998) 5061.
[32] H. Kita, Y. Gao, H. Hattori, J. Electroanal. Chem. 373 (1994)
177.
[33] N. Markovic, P.N. Ross, J. Electroanal. Chem. 330 (1992) 499.
Acknowledgements [34] (a) F.C. Nart, T. Iwasita, M. Weber, Electrochim. Acta 39 (1994)
961;
Financial support from FAPESP and CNPq, Brazil (b) T. Iwasita, F.C. Nart, M. Weber, Electrochim. Acta 39 (1994)
and DFG, Germany, is gratefully acknowledged 2093.
[35] K.J. Cathro, J. Electrochem. Soc. 116 (1969) 1608.
[36] M.M.P. Jansen, J. Moolhuysen, J. Catal. 46 (1977) 289.
[37] O.A. Petrii, B.I. Podlovchenko, A.N. Frumkin, H. Lal, J.
Electroanal. Chem. 10 (1965) 253.
References [38] H. Binder, A. Köhling, G. Sandstede, in: G. Sandstede (Ed.),
From Electrocatalysis to Fuel Cells, University of Washington,
[1] W. Vielstich, Fuel Cells, Wiley Interscience, Bristol, 1965. Seatle, 1972.
[2] M.W. Breiter, Electrochemical Processes in Fuel Cells, Springer- [39] M. Watanabe, S. Motoo, J. Electroanal. Chem. 69 (1976) 429.
Verlag, Berlin, 1969. [40] T. Iwasita, F.C. Nart, W. Vielstich, Ber. Bunsenges Phys. Chem.
[3] O.A. Petrii, B.I. Podlovchenko, A.N. Frumkin, H. Lal, J. 94 (1990) 1030.
Electroanal. Chem. 10 (1965) 253. [41] M. Watanabe, S. Motoo, J. Electroanal. Chem. 60 (1975) 267.
[4] V.S. Bagotzki, Y. Vassileiv, Electrochim. Acta 12 (1967) 1323. [42] N. Gasteiger, Markovic, P.N. Ross, Electrochim. Acta 41 (1996)
[5] M. Breiter, Electrochim. Acta 12 (1967) 1213. 2587.
[6] V.S. Bagotzki, Y.B. Vassiliev, O.A. Kazova, J. Electroanal. [43] H.A. Gasteiger, N. Markovic, P.N. Ross, E.J. Cairns, J. Electro-
Chem. 81 (1977) 229. chem. Soc. 141 (1994) 1795.
[7] T. Iwasita, in: H. Gerischer, C. Tobias (Eds.), Advances in [44] N. Markovic, H.A. Gasteiger, P.N. Ross, L. Villegas, M.J.
Electrochemical Science and Engineering, vol. 1, Verlag Chemie, Weaver, Electrochim. Acta 40 (1995) 91.
1990, p. 127. [45] M. Krausa, W. Vielstich, J. Electroanal. Chem. 379 (1994) 307.
[8] R. Parsons, T. Van der Not, Electroanal. Chem. 257 (1988) 1. [46] T. Iwasita, H. Hoster, A. John-Anacker, W.F. Lin, W. Vielstich,
[9] S. Wasmus, A. Küver, J. Electroanal. Chem. 461 (1999) 14. Langmuir 16 (2000) 522.
[10] A. Hamnett, in: A. Wieckowski (Ed.), Interfacial Electrochem- [47] W. Chrzanowski, A. Wieckowski, Langmuir 14 (1998) 1967.
istry. Theory, Experimental and Applications, Marcel Dekker, [48] W. Chrzanowski, W.H. Kim, A. Wieckowski, Catal. Lett. 50
New York, 1999, p. 843. (1998) 69.
[11] B. Bittins-Cattaneo, E. Cataneo, P. Königshoven, W. Vielstich, [49] G. Tremiliosi-Filho, H. Kim, W. Chrzanowski, A. Wieckowski,
in: A.J. Bard (Ed.), Electroanalytical Chemistry a Series of B. Grzybowska, P. Kulesza, J. Electroanal. Chem. 467 (1999) 143.
Advances, Marcel Dekker, 1991, p. 181. [50] H. Hoster, T. Iwasita, H. Baumgärtner, W. Vielstich, Phys. Chem.
[12] T. Biegler, J. Phys. Chem. 72 (1968) 1571. Chem. Phys. 3 (2001) 337.
3674 T. Iwasita / Electrochimica Acta 47 (2002) 3663 /3674

[51] H. Hoster, T. Iwasita, H. Baumgärtner, W. Vielstich, J. Electro- Society, 52nd ISE Meeting Ext. Abstract 330, San Francisco,
chem. Soc. 148 (2001) A496. 2001.
[52] A. Kabbaby, R. Faure, R. Durand, B. Beden, F. Hahn, J.-M. [55] A.O. Neto, J. Peres, W.T. Naporn, E. Ticcianelli, E. González, J.
Leger, C. Lamy, J. Electroanal. Chem. 444 (1998) 41. Braz. Chem. Soc. 11 (2000) 39.
[53] D. Kardash, C. Korzeniewski, N. Markovic, J. Electroanal. [56] M.T.M. Koper, J.J. Lukkien, A.P.J. Jansen, R.A. van Santen, J.
Chem. 500 (2001) 518. Phys. Chem. B 105 (1999) 5522.
[54] W.H. Lizcano-Valbuena, E.C. Bortholin, A. Oliveira Neto, V.A. [57] H. Gasteiger, P.N. Ross, E.J. Cairns, Surf. Sci. 293 (1993) 67.
Paganin, E.R. González, 200th Meeting of the Electrochemical [58] E. Herrero, J. Feliu, A. Wieckowski, Langmuir 15 (1999) 493.

Das könnte Ihnen auch gefallen