Sie sind auf Seite 1von 20

143

J Electroanal. Chem., 241 (1988) 143-162


Elsevier Sequoia S.A., Lausanne - Printed in The Netherlands

MECHANISMS OF FORMIC ACID, METHANOL, AND CARBON


MONOXIDE ELECTROOXIDATION AT PLATINUM AS EXAMINED BY
SINGLE POTENTIAL ALTERATION INFRARED SPECTROSCOPY

DENNIS S. CORRIGAN and MICHAEL J. WEAVER


Department of Chemistry. Purdue Umuersrty, Wesi Lafayefre, IN 47907 (U.S.A.)

(Received 20th July 1987; in revised form 23rd September 1987)

ABSTRACT

Surface infrared spectra have been obtained as a function of potential and time during the
voltammetric oxidation of formic acid and methanol on polycrystalline platinum in order to probe the
possible role of adsorbed carbon monoxide in the electrooxidation mechanisms. The procedure mvolves
obtaining a sequence of single-beam infrared spectra using an FTIR spectrometer during potential sweep
or following potential-step perturbations, and referencing these to a spectrum obtained at the initial
potential or after complete oxidation had occurred. Using this single potential alteration infrared
@PAIR) techmque, individual spectra for the C-O stretch of adsorbed carbon monoxide, +o. as well as
the O-C-O stretch for the CO, product could be obtained under voltammetric conditions in as little as
3.5 s, and repetitively every 7 s. Such SPAIR spectra obtained during either potential sweep or step
excursions Indicate that the onset of both formic acid and methanol oxidation coincides with the
oxldattve removal of adsorbed carbon monoxide. The electrooxidation of CO irreversibly adsorbed from
solution carbon monoxide was also examined using this approach. The integrated absorbance of the Y,-o
band was generally found to be proportional to the CO coverage (determined from either the CO, band
intensity or the voltammetric charge) during electrooxidation. The dependence of the vco peak frequency
upon coverage during potentiostatic oxidation, however, is sensitive to the timescale over whtch the
process proceeds. The adsorption kinetics of CO formed from forrmc acid and methanol were evaluated
from the time dependence of the I+-o band intensity following suitable potential step sequences to
potentials where CO electrooxidation is suitably slow. These data suggest that while adsorbed CO may
act as an adsorbed intermediate for methanol oxidation, it probably acts as a chemisorbed poison for
formic acid oxidation.

INTRODUCTION

The application of in-situ infrared reflection-absorption spectroscopy (IRRAS)


to the identification of surface species formed during electrooxidation of small
organic molecules, especially formic acid and methanol, on noble metal electrodes
has received considerable attention in recent years [l--5]. Although it is widely
believed that more than one type of adsorbed intermediate is responsible for the
144

observed electrocatalytic behavior [6], the identity and reactivity of these species
continues to be a topic of much debate.
The early infrared studies of Beden et al. [l] employing electrochemically
modulated infrared spectroscopy (EMIRS) indicated that a predominant adsorbate
formed on platinum from methanol and formic acid is linearly bound carbon
monoxide. Left unanswered, however, is the possible importance of CO as an
adsorbed intermediate and/or poison in the electrooxidation mechanisms. To gain
such information it is desirable to obtain infrared spectra for conditions that
coincide with those for which the electrochemical kinetic measurements are made.
Potential modulation infrared techniques such as EMIRS suffer in this regard since
they are limited to the observation of reversible potential-induced spectral changes,
resulting from the need to modulate the potential back and forth over the range of
interest so as to minimize the bulk-phase spectral interference. Instead, in order to
gain information on the possible role of adsorbed species in the irreversible
electrooxidation pathways it is necessary to employ in-situ infrared spectral tech-
niques that can sense irreversible potential-induced changes in the surface composi-
tion while the electrode reaction is proceeding.
Two variants of IRRAS possess the characteristics to make them suitable for this
purpose. The first, polarization-modulated IRRAS (PM-IRRAS), achieves the nec-
essary subtraction of bulk-phase spectral interferences, instrument drift, etc. (at
least in part) by rapid alteration the polarization of the incident beam. Although less
sensitive than potential-difference methods, PM-IRRAS has the important virtue of
enabling absolute spectra to be obtained at a given electrode potential. Very
recently, Kunimatsu has utilized PM-IRRAS to examine the potential-dependent
CO coverage in relation to the steady-state electrooxidation rate of formic acid
and methanol on platinum [5b]. While this method is in principle suitable for
examining such potential-induced irreversible changes in the surface composition,
the time taken (- minutes) to record each spectrum (at least when a dispersive
spectrometer is employed) clearly limits its mechanistic utility.
Given that such irreversible electrode processes can be examined conveniently by
sweeping or stepping the potential from an initial value where no reaction occurs,
potential-difference infrared (PDIR) spectral methods should also be applicable
since the bulk-phase spectral interference can in principle be removed by referenc-
ing spectra observed at potentials where the reaction occurs to that obtained
previously at the initial potential. This is not feasible using a dispersive spectrometer
since the referencing procedure needs to be made repeatedly during the slow
wavelength scan necessary to acquire each spectrum. While it is also conventional to
alter the potential repeatedly when recording PDIR spectra using a Fourier trans-
form instrument, this modulation occurs over a longer timescale than that (5 1 s)
required for each interferometer scan, and is occasioned solely by the desire to
minimize the deleterious effects of temporal instrument drift (e.g. source, detector
fluctuations).
Potential-difference FTIR spectra can indeed be obtained for adsorbed CO and
other systems displaying strong infrared absorption by applying a single potential
145

excursion only during the spectral acquisition [7]. We have dubbed this approach
single potential alteration infrared spectroscopy @PAIRS) [7]. The procedure
involves acquiring a given number of interferometer scans first at the initial base
potential and then at one or more sample potentials either after a potential step(s)
or during a potential sweep, the base spectrum being subtracted from (or ratioed
against) each of the sample spectra. In suitable cases, as little as five interferometer
scans (consuming ca. 3 s) are required at each potential to yield satisfactory
signal-to-noise, so that time-resolved surface infrared spectra can be obtained
simultaneously with the acquisition of conventional potential-sweep or -step voltam-
mograms [7].
In this paper we present such corresponding SPAIRS and voltammetric data for
the electrooxidation of carbon monoxide, formic acid, and methanol on polycrystal-
line platinum. Attention is focussed on employing SPAIRS to evaluate the kinetics
of adsorption and electrooxidation of CO formed from formic acid or methanol in
relation to those of the overall reactions in order to shed light on the possible
mechanistic role(s) of this adsorbate in these much-studied processes.

EXPERIMENTAL

Most experimental details of the surface infrared measurements are given in ref.
8. The infrared spectrometer used was a Bruker-IBM IR 98-4A Fourier transform
instrument, with a liquid nitrogen-cooled InSb detector (Infrared Associates). The
spectral resolution was * 4 cm -. The incident light was p-polarized by means of a
KRS-5 wire-grid polarizer (Harrick). The polycrystalline platinum surface was
pretreated immediately prior to use by mechanical polishing, dipping in hot chromic
acid, followed by rinsing and potential cycling in 0.1 M HClO, at 0.1 V s-l
between -0.25 and 1.4 V vs. SCE for 10 min. An electrochemical thin-layer
configuration, necessary for the spectral measurements, was formed by pressing the
electrode up against the CaF, optical window. A PAR Model 173/179 potentiostat
was employed, controlled by a PAR Model 175 with the resulting cyclic voltammo-
grams being plotted on a Houston Instruments X-Y recorder.
Formic acid (Fisher), 13C formic acid (Aldrich), methanol (Burdick and Jackson),
and perchloric acid (G.F. Smith) were reagent grade and used as supplied. All
electrode potentials are quoted with respect to the saturated calomel electrode
(SCE), and all measurements were made at room temperature, 23 f 1C.

RESULTS AND DISCUSSION

Electrooxidation of adsorbed CO

The electrooxidation of irreversibly adsorbed carbon monoxide at platinum


provides a convenient parent system with which to characterize combined
SPAIRS-voltammetric measurements with a view towards examining the role of
adsorbed CO in other electrooxidation processes. Figure 1A consists of a sequence
146

B 2062

326

362
v\,

500
5
IAR
,=7XIcr3

I 1.5X10-30.1~.

u/cm-l 185 I 50 v/cm- I 200

Fig. 1. Single potential alteration infrared @PAIR) spectra obtained for elcctrooxldation of Irreversibly
adsorbed CO at platinum during a positive-going potential sweep at 2 mV s- from -0.2 V vs. SCE m
0.1 M HClO,. Carbon monoxide was adsorbed at -0.2 V and removed from solution by nitrogen
purging prior to formation of the thin layer. (A) Spectra, displayed as AR/R (i.e. relative reflectance) vs.
v/cm-, obtained by acquirmg a sequence of single-beam spectra at the average potenttab indicated
(mV vs. SCE), each generated by the coaddttion of ten interferometer scans and ratioing these to the
corresponding spectrum obtained at the initial potential (see text). (B) Absolute SPAIR spectra in the
Y,-o region, 2000-2150 cm-, generated from the absorbance spectra for each of the sample potentrals by
subtracting the corresponding single-beam spectrum taken at 0.38 V (at which the adsorbed CO has
undergone complete oxidative removal).

of SPAIR spectra obtained during a positive-going potential sweep at 2 mV s-


from -0.2 V vs. SCE in 0.1 M HClO,. The carbon monoxide was adsorbed with
the potential held at -0.2 V by first saturating the electrolyte with CO, which was
then removed from the solution by nitrogen purging prior to forming the electro-
chemical thin layer. The spectral data were obtained by acquiring a sequence of
147

I 0.01 mA

O-

0 400 800
E/mV vs SCE
Fig. 2. Cychc voltammogram for electrooxtdation of irreverstbly adsorbed carbon monoxide at platinum
in 0.1 M HClO, (- ) obtamed stmultaneously with the spectral data of Fig. 1. The sweep rate was 2
mV ss and the solution was purged with nitrogen prior to formation of the spectral thm layer. Anodic
current is plotted upwards; the electrode area 1s 0.71 cm. The dashed trace was obtained in 0.1 M
HClO, alone under identical condittons.

single-beam spectra, each generated by coadding ten interferometer scans and


ratioing each of these to the corresponding spectrum obtained at the initial potential
just prior to the onset of the potential sweep. Since the acquisition of ten scans
requires about 7 s, and a further 7 s is consumed by computer commands associated
with data manipulation, repetitive spectra can be recorded every 14 s. (Satisfactory
spectra were also obtained by employing as few as five scans per spectrum, see
below.) Consequently, the potential is altered by only 28 mV during the acquisition
of each spectrum in Fig. lA, the various potentials labelled on this figure being the
average value during each spectrum. The linear-sweep voltammogram obtained
during these SPAIR measurements is shown in Fig. 2. The sharp anodic peak
corresponds to the electrooxidation of the adsorbed CO. (The slightly greater
current relative to the blank electrolyte seen at more positive potentials in Fig. 2 is
probably due to surface structural changes since solution CO is entirely absent
under these conditions.) The complete disappearance of the anodic current peaks
due to hydrogen desorption upon CO adsorption at -0.2 V (Fig. 2) indicates that
essentially a complete CO monolayer is formed under these conditions.
Two features of Fig. 1A are of interest (cf. ref. 7). Firstly, a negative-going band
at 2343 cm-, due to the C-O stretch of CO,, is produced at potentials > 0.30 V,
beyond the onset of CO oxidation. Secondly, the bipolar band around 2060-2070
cm- seen at the least positive potentials gives way to a positive-going unipolar
band at potentials > 0.40 V, where CO oxidation is complete (Fig. 2). This unipolar
feature corresponds to an absolute co band for CO adsorbed at the initial potential
since no adsorbate remains at the sample potentials under these conditions. This
circumstance also enables absolute Yco spectra to be extracted for potentials prior
to CO oxidation by ratioing individual single-beam spectra taken during the
Fig. 3. Relative integrated absorbance of the +-o band, A,(CO). obtained from SPAIR spectra during
electrooxidatlon of adsorbed CO (as in Fig. 1) plotted against the adsorbate coverage 8. The latter was
obtained from the relative intensities of the 2343 cm- CO, band (see text). Data obtained for potential
sweep rates of 1, 2, and 5 mV s-l are plotted as open squares, triangles and circles, respectively. The
closed diamonds refer to SPAIRS data collected via single potential steps (see text for details).

voltammogram at potentials prior to complete CO oxidation with a spectrum


obtained after CO removal. Such absolute SPAIR spectra in the vco region are
shown in Fig. 1B. This display mode has the obvious advantage of enabling the peak
frequency, bandshape, and integrated intensity of the vco band to be extracted more
readily than from the difference spectra in Fig. 1A.
Data such as those in Fig. 1 can also be employed to ascertain the relation
between the co band intensity and the adsorbate coverage, 8, during CO electro-
oxidation. This is achieved by noting that the integrated intensity of the CO, feature
at 2343 cm- is necessarily proportional to the quantity of CO oxidized, and
therefore to (1 - 0). The resulting plot of the integrated intensity of the vco band,
A,(CO), against the adsorbate coverage determined in this manner is shown in Fig.
3. Included in this plot are data from SPAIR spectra obtained for potential sweep
rates of 1, 2, and 5 mV s- (open symbols), and also for single potential steps
(closed symbols). The latter were acquired in the same manner as the potential-sweep
data, except that spectra were obtained sequentially every 11 s (i.e. 5 interferometer
scans per spectrum) following potential steps from -0.2 V to values (0.25 to 0.4 V)
where CO electrooxidation proceeded at a suitable rate. The CO, band intensity
remained essentially constant upon completion of the potential-step CO oxidation
as signalled by the cessation of faradaic current, indicating that essentially no CO,
is lost from the thin-layer cavity under these conditions (timescale G 3 min).
Inspection of Fig. 3 reveals that the A,--8 plot is linear within experimental error
throughout the entire coverage range. This simple behavior has an obvious utility in
enabling relative 6 values to be estimated from the vco intensities (see below). A
149

Fig. 4. Comparison between the fractional coverage, 8. for irreversibly adsorbed CO as a function of the
electrode potential during a positive-going sweep at 1 mV s- as obtained from the coulombic charge for
CO electrooxidation (O , left-hand axis) and from the SPAIRS CO, band intensity (0). Corresponding
estimates of 0 obtained from the vc, band intensity are plotted as open squares (see text).

similar result has been reported recently by Kunimatsu [5a], utilizing the extent of
hydrogen adsorption determined by cyclic voltammetry as a means of estimating the
CO coverages. The present method of determining 0 is more direct. Interestingly,
the observation of linear A,-8 plots for electrochemical CO adsorption contrasts the
markedly non-linear or even peaked plots that are commonly seen for CO adsorp-
tion on metals from the gas phase [9], including platinum [lo].
An additional measure of the CO coverage as a function of potential during the
linear sweep voltammetry is provided in the present case by the coulombic charge
contained under the current-potential peak itself (Fig. 2). Figure 4 is a comparison
between the fractional CO coverage as a function of the electrode potential during a
positive-going sweep at 1 mV s- as obtained from the coulombic charge (filled
circles) and from the CO, band intensity (open circles). The close agreement seen
between the points confirms the quantitative validity of this spectrophotometric
method for evaluating the CO coverage. Also included in Fig. 4 are corresponding
estimates of 0 from the vco band intensity (open squares). Similar agreement
between these three methods for evaluating the time-dependent CO coverage during
electrooxidation was obtained for different sweep rates and under potential-step
conditions.
These simultaneous SPAIRS-voltammetric measurements also enable the CO,
band intensities to be calibrated on the basis of the total quantity of CO oxidized.
By integrating the coulombic charge under voltammograms such as Fig. 2, and given
150

that a two-electron oxidation is involved, we deduce that the surface concentration


of the CO monolayer, I, is 2.0 (50.2) x 10e9 mol cm-2; the corresponding
integrated CO, band intensity is 0.069 cm-.
Another interesting aspect of the present data concerns the variation of the vco
peak frequency, $o, with potential and time during electrooxidation. Over the
potential region -0.2 V to 0.25 V, prior to the onset of oxidation, r~:o is seen to
increase up to ca. 0 V and remain essentially constant at more positive potentials
(Fig. 1B). A similar observation has been made previously using PM-IRRAS [3b].
More significant changes in v:. occur, however, during electrooxidation. For this
purpose it is advantageous to examine time-dependent absolute SPAIR spectra
for vco following potential steps rather than during sweeps since the possible effects
of potential and coverage upon v:o can then be separated. Plots of v:. as a
function of 8 for potential steps from -0.2 V to 0.25 V (open squares), 0.30 V
(triangles), 0.35 V (diamonds), and 0.40 V (circles) are shown in Fig. 5. At the lowest
overpotentials, v:o is observed to decrease substantially, by as much as 20 cm-, as
the CO undergoes oxidation. Similar frequency decreases resulting from partial
oxidation of the CO layer have been noted previously [3,5a]. However, Fig. 5 shows
that the extent of the vp co-8 dependence is diminished progressively at higher
overpotentials, so that at 0.35 and 0.4 V, v& remains virtually independent of 8.
Differences were also seen in the vco bandshape depending on the timescale of
the oxidation process. For rapid oxidation (i.e. at 0.35 or 0.4 V), the vco
bandwidth remained virtually unchanged, the full width at half height being ca.
15 cm-. However, when oxidation proceeded slowly, the decreases in vFo were
accompanied by significant increases in the bandwidth (cf. refs. 3b, 5a).
A possible origin of these varying v:~- 0 dependencies is that the timescale over
which the potentiostatic oxidation occurs is diminished substantially as the over-
potential is increased. For example, only 20 s are required to decrease 8 from 1.0 to
0.2 for oxidation at 0.4 V, whereas the same process requires 180 s at 0.25 V. In
order to check this possibility, a series of double potential-step perturbations were
applied in which the potential was altered first from -0.2 V to 0.35 V so as to
oxidize a certain fraction of the adsorbate, and then back to 0.25 V whereupon
SPAIR spectra were recorded. The v:o values obtained in this fashion at 0.25 V are
plotted (filled squares) against 0 in Fig. 5. In contrast to the ZJ:~ values obtained by
potentiostatic oxidation at 0.25 V, these latter v:o values display little or no
dependence upon 8. It is therefore apparent that the substantial decreases in v:o
with decreasing coverage when electrooxidation proceeds at lower overpotentials are
associated with the extended periods of time over which this process takes place
under these conditions.
These vp co-B variations are presumably associated with relaxation effects within
the CO layer as it undergoes oxidation. The oxidation mechanism is considered to
involve reaction at the interface between islands of adsorbed CO and the chemical
surface oxidant, either adsorbed water or metal oxide [ll]. Provided that the CO
islands do not rearrange during oxidation, v:. is not expected to vary greatly with
decreasing coverage [9], thereby accounting for the observed behavior under condi-
151

2075

T
E
::
0
Y

2065

2055
0.1 0.3 I
0.5 07 0.9
8

Fig. 5. Plots of the +o frequency, v:o, as a function of surface coverage, 8, during potential-step
electrooxtdation of Irreversibly adsorbed CO. Open symbols correspond to values of Y,- extracted from
absolute SPAIR spectra (see text) obtained as a function of time following steps from -0.2 V to 0.25
V (o), 0.30 V (A),0.35 V (0). and 0.40 V vs. SCE (0). CO was adsorbed initially at - 0 2 V. The data
represented as the closed squares were extracted from absolute SPAN spectra obtained at 0.25 V
where the coverage was decreased rapidly vta parttal electrooxtdation at 0.4 V for varying time
increments before stepping back to 0.25 V (see text for details).

tions of rapid oxidation. When the removal of the adsorbed layer is slow, on the
other hand, the CO islands may well rearrange and even randomize. Under these
circumstances the extent of dipole-dipole coupling between nearby CO molecules
will decrease, and $.o should thereby decrease substantially, in harmony with the
observed behaviour. Further detailed examination of this interesting relaxation
phenomenon is clearly worthwhile, and is planned.

Electrooxidation of formic acid and methanol

The foregoing results demonstrate that SPAIRS can be employed to monitor the
quantity and nature of adsorbed CO present while electrode reactions are proceed-
ing under conventional voltammetric conditions. This prompts the use of the
152

approach to probe the presence of adsorbed CO as a function of potential and/or


time during the voltammetric oxidation of small organic molecules, such as formic
acid and methanol.
Two variants of these experiments were performed. The first involved acquiring
SPAIR spectra during positive-going potential sweeps (typically 1, 2, and 5 mV s-)
in 0.1 M HClO, containing formic acid or methanol (usually 25 mM). Such typical
SPAIR spectra, for the electrooxidation of 25 mM formic acid at a sweep rate of 2
mV s-i from an initial potential of -0.2 V, are shown in Fig. 6A. Each spectrum
was obtained by employing only five interferometer scans, consuming about 3.5 s.
The potentials selected for inclusion in Fig. 6A span the region. ca. 0.2 to 0.55 V,
where the voltammetric oxidation occurs, as illustrated in Fig. 7A. Corresponding
absolute SPAIR spectra in the vco region are shown in Fig. 6B. The form of the
SPAIR spectra for formic acid oxidation (Fig. 6) are roughly comparable to those
obtained for the oxidation of carbon monoxide (Fig. 1). Thus the disappearance of
the negative-going component of the bipolar vco band, indicating oxidation of the
adsorbed CO, occurs in the same potential region where the 2343 cm- band is
rapidly gaining intensity, corresponding to an escalating rate of overall electrooxida-
tion.
Although the band intensities for the adsorbed CO in Fig 6 are similar to those
in Fig. 1, inferring CO coverages that approach unity (cf. ref. 5a), the CO, band in
the former attains considerably greater intensities. This is of course expected since
in the former case CO, will be produced from formic acid in the thin-layer solution
in addition to that from initially adsorbed CO. Indeed, the time-dependent intensity
of the CO, band can be related quantitatively to the extent of the overall reaction
from the known proportionality constant between the band absorptivity and the
amount of CO, produced (see above). Given that formic acid oxidation to CO,
involves two electrons, the derivative of the CO, band intensity with respect to time
yields the faradaic current density, if, as a function of potential, E, during the
linear sweep voltammogram. A typical log if-E plot determined in this manner for
a 5 mV s-i sweep is illustrated as the dashed curve in Fig. 8. The corresponding log
if-E plot obtained simultaneously from the thin-layer voltammogram itself (e.g.
Fig. 7A) is shown for comparison as the solid curve in Fig. 8. These two curves in
Fig. 8 are in tolerable agreement at the least positive potentials, although the former
exhibits significantly smaller currents than the latter at higher potentials.
This disparity between the infrared and electrochemical oxidation rate-potential
behaviour is most likely due to the non-uniformity of the thin layer. Thus the
infrared beam will sample preferentially regions of the solution layer that are
sufficiently thin to yield a significant transmittance through the aqueous solvent. As
a consequence, the effective solution reservoir for the infrared-active regions of
the electrode surface will tend to be thinner, and hence contain on average less
formic acid per unit area, than the entire surface which is presumably sensed
electrochemically. Reactant depletion within the infrared-active regions should
therefore occur more rapidly than in the remaining portions of the thin layer,
therefore accounting for the earlier decline (i.e. at less positive potentials) of the
153

3 2061

358
-

424
t- ---+

534 358 2q56


-

380

I
y=3xKT2

25 I00
2 3
v/cm- 185 150
II 5XlG3

u/cm-
au

2oc 3

Fig. 6. SPAIR spectra obtained dunng a positive-gomg potential sweep at 2 mV s- from -0.2 V vs.
SCE m 0.025 M HCOOH+O.l M HCIO,, after adsorption at -0.2 V for 20 min. (A) Relative
reflectance spectra, obtained by acquiring a sequence of smgle-beam spectra (each using 5 interferometer
scans) at the potentials indicated (mV vs. SCE), and ratioing these to the spectrum obtamed at the initial
potential (see text). (B) Some corresponding absolute SPAIR absorbance spectra in the vco region,
generated by subtractmg the single-beam spectrum obtained at 0.53 V (i.e. followmg CO electrooxida-
tion; see text).

reaction rate as sensed with the infrared versus the electrochemical probe (Fig. 8).
These data are, nonetheless, not inconsistent with the quantitative agreement noted
above between the infrared and electrochemical probes for the electrooxidation of
irreversibly adsorbed CO since no solution species are involved in the latter case.
As noted above, comparisons between the potential-dependent intensities of the
adsorbed CO and solution CO, bands are of particular interest since they provide
direct information on the reactivity of the former in relation to that of the overall
0 400 800
E/mVvs.SCE

o- #
1
0
I I

400
I 1

800
I r J
E/mV vs. SCE
Fig. 7. Anodic-cathodic cyclic voltammograms for electrooxidation of forrmc acid (A) and methanol (B)
at platinum; (A) was obtained simultaneously with the acquition of the spectral data in Fig. 6. The
solution conditions were 0.025 M HCOOH and 0.025 M CH,OH m 0.1 M HCIO,, respectively; the
sweep rate was 2 mV s-l and the electrode area was 0.71 cm2,

reaction. Such a comparison is shown in Fig. 9 for the oxidation of 25 mM formic


acid in 0.1 M HClO,, extracted from SPAIR spectra such as in Fig. 6. The squares,
triangles and circles correspond to sweep rates of 1, 2, and 5 mV s-l; the filled and
open points refer to the integrated absorbances of the adsorbed CO and solution
CO, bands. respectively. Figure 10 consists of the plots obtained from correspond-
ing SPAIR spectra for the oxidation of 25 mM methanol in 0.1 M HClO,. A
representative thin-layer voltammogram is shown in Fig. 7B. Comparable results
were obtained for both formic acid and methanol using other reactant concentra-
tions in the range lo-100 mM.
Provided that the uco absorbance is again approximately proportional to the CO
coverage for these systems (cf. ref. 5a), we can deduce from Figs. 9 and 10 that the
electrooxidation of diffusing formic acid or methanol commences only when the
adsorbed CO layer itself begins to react and the coverage starts declining. A related
observation has been made recently by Kunimatsu and Kita [5b] by employing
PM-IRRAS to measure the +o band intensity at progressively more positive
potentials in relation to the steady-state currents for formic acid and methanol
155

oxidation. The present measurements, however, have the advantage of being ob-
tained more rapidly and under actual voltammetric conditions.
The second variant of time-dependent SPAIR measurements for formic acid and
methanol oxidation involved potential-step conditions. Figure 11 contains plots of
the intensities of the adsorbed CO (filled symbols) and solution CO, bands (open
symbols) for 25 mM formic acid in 0.1 M HClO, as a function of time following
potential steps from -0.2 V to 0.25 V (squares), 0.3 V (triangles), and 0.4 V
(circles). The results in Fig. 11 show that the relation between the co band intensity
(and presumably the CO coverage) and the extent of overall reaction, as given by
the CO, band intensity, is essentially independent of the timescale over which the
reaction proceeds.

-0.5

Z-I.0
E

:
\
.z_z7
-:

-I .5

-2.c I-

I 8
300 400 500
E/mV vs SCE
Fig. 8. Plot of logarithm of the faradaic current density. log I, versus electrode potential, E. for
electrooxidation of formic acid at platmum, obtained for a 5 mV s- potential sweep. Electrolyte was
0.025 M HCOOH +O.l M HClO,. The current densities used to generate the dashed curve were obtained
by analyzing the derivative of the SPAIRS CO, band intensity (such as those in Fig. 6A) with respect to
time (see text). The solid curve is the log I -E plot obtained simultaneously from the voltammogram
-0-

I
200 300 400 500 600
E / mV vs. SCE

Fig. 9. Relative integrated absorbance of the rco band for adsorbed CO (filled symbols) as a functton of
electrode potential during a linear potential sweep from - 0.2 V for formrc acid oxidation in comparison
with the extent of overall reatton as determined by the integrated absorbance of the CO, band (open
symbols). Squares, triangles. and circles correspond to sweep rates of 1. 2, and 5 mV s-. Solution was
0.025 M HCOOH in 0.1 M HCIO,.

1.0

5uv bUU IVW


E / mV vs. SCE

Fig. 10. As for Fig. 9, but for methanol oxidatton. Solution was 0.025 M CH,OH in 0.1 M HCIO,
157

150 200

Fig. 11. Relative integrated absorbance of the vco band for adsorbed CO (filled symbols) in comparison
wtth the extent of overall reaction as determined by the mtegrated absorbance of the CO, band (open
symbols) as a function of time following potential steps from -0.2 V to 0.25 V (squares), 0.3 V
(triangles), and 0.4 V (circles).

The role of adsorbed carbon monoxide: Poison or reaction intermediate?

The clear synergy between the electrooxidation of adsorbed CO and that of


solution-phase formic acid or methanol demonstrated in Figs. 9-11 might be
interpreted as indicating that the carbon monoxide is acting merely as an electrode
poison, partial removal of which is required in order to release surface sites
necessary for the organic electrooxidation process to proceed. Such a conclusion was
reached by Kunimatsu and Kita [5b], and indeed much discussion has centered
around this notion [12] following the identification by infrared spectroscopy of
adsorbed CO for these systems [l].
It has been demonstrated that formic acid electrooxidation on platinum can
occur via a dual pathway mechanism featuring both weakly and strongly ad-
sorbed intermediates [6a]. The presence of the former pathways is associated with
an anodic voltammetric peak (peak I) appearing at around 0.3-0.4 V less positive
potentials than the major anodic feature seen in Fig. 7A (peak II). The prevalence
of peak I is known to decrease rapidly as the electrode is held at the initial potential,
especially at less positive values, as the strongly adsorbed intermediate is formed
[6c]. As evidenced by Fig. 7A, this condition applies to the SPAIRS-voltammetric
experiments described here, as expected since the relatively slow potential sweep
rates necessitated by the spectral data acquisition aided the complete cessation of
158

peak 1. For methanol oxidation at platinum, the form of the anodic voltammograms
is relatively insensitive to such factors [5b], suggesting that a single type of reaction
pathway suffices to account for the observed kinetics. The identification of adsorbed
CO as the (or at least a major) strongly chemisorbed species for both formic acid
and methanol systems under the conditions encountered here is firmly supported by
the earlier infrared evidence [1,5].
An important possibility not addressed by these considerations, however, is that
the adsorbed CO may act as a true reaction intermediate for these processes, in that
most or all of the CO, produced from the solution-phase reactant under voltammet-
ric conditions may be produced via transient dissociative adsorption to form CO.
For formic acid oxidation, therefore, we might envisage the following simplified
mechanism (ignoring detailed dissociation steps):

HCOOH + CO(ad) + H,O (1)


CO(ad)+HzO-2 e-+CO,+2H (2)
and for methanol oxidation:

CH,OH + CO(ad) + 4 H (3)


CO(ad) + H,O - 2 e- -j CO, + 2 H+ (4)
Such mechanisms are clearly distinct from those where the electrooxidative removal
of adsorbed CO merely provides surface sites at which another intermediate can
form. For the former pathways to be tenable, however, the chernisorption steps (1)
or (3) must be sufficiently rapid, at least at the potentials where steps (2) and (4)
occur, to account for the observed rates of the overall reactions. If adsorbed CO is
acting instead as a poison, formation and electrooxidative removal can occur at
rates which are markedly slower than for the preferred reaction pathway.
A distinction between these two possibilities can therefore be made by evaluating
the rates of CO adsorption from formic acid or methanol solutions in relation to
those for the overall voltammetric reaction. To this end, a number of time-resolved
SPAIRS measurements were made in which the growth of the uco band intensity
was monitored following potential steps to values where the equilibrium (or steady-
state) CO coverage is larger than at the initial value. Such data enable the CO
adsorption kinetics to be evaluated quantitatively if it is again assumed that the CO
coverage is proportional to the vco band intensity.
Two variants of these experiments were performed. The first involved stepping
from a relatively negative potential to a value where the CO is removed partly by
oxidation, the potential then being returned to a less positive value where upon the
~co intensity-time relation was monitored. In some cases the oxidative desorption
pulse also resulted in substantial reactant depletion in the thin layer. This difficulty,
however. could be circumvented by pulling the electrode back from the CaF,
window during this pulse, the thin layer being reformed immediately before the
return potential step was applied and the SPAIRS measurements initiated.
The second, simpler, variant of these experiments involved a single positive
159

Fig. 12. Adsorptron kinetics of fornuc acid at 0 V (squares) and 0.1 V vs. SCE (circles) as determined
from the relative mtegrated absorbance of the ZJ,-- band as a function of time following various potential
steps. For the open symbols, the surface was initially cleaned via a potential pulse to 1.2 V for several
seconds wtth the electrode pulled away from the wmdow: the spectral data were than gathered by
reforming the thin layer immediately before the reverse potential step. The closed symbols refer to partial
electrooxidative desorptron before stepping back to the adsorption potential. The adsorbed CO was
oxidized at 0.35 V for varying time intervals and the potential was stepped back to the re-adsorption
value. Again, the squares and circles refer to adsorption at 0 and 0.1 V respectively. The closed diamonds
correspond to a readsorption potential of 0 V, with an mitral surface coverage close to 0.5.

potential step. This procedure was applicable to methanol (but not to formic acid),
taking advantage of the minor extent of CO adsorption that takes place in the
potential region where hydrogen adsorption occurs [5]. Stepping the potential from
this region (- 0.2 V was usually chosen) to various more positive values therefore
enabled the CO adsorption kinetics as a function of coverage to be monitored
readily.
A representative summary of the adsorption kinetic data so obtained for formic
acid and methanol is presented in Figs. 12 and 13, respectively. The former data
refer to two adsorption potentials, 0 V (squares, diamonds) and 0.1 V (circles)
following an oxidation pulse to 1.2 V; data gathered by reforming the thin layer
immediately before the reverse potential step are plotted as open symbols in Fig. 12.
The kinetic data for methanol solutions in Fig. 12 refer to adsorption potentials of 0
V (squares), 0.1 V (circles), 0.2 V (diamonds), and 0.3 V (triangles). The open
symbols refer to data gathered during single potential steps from -0.2 V, whereas
the closed symbols involve partial electrooxidative desorption before stepping back
to the adsorption potential (see figure captions for further details).
Fig. 13. Adsorption kinetics of methanol at 0 V (squares), 0.10 V (circles), 0.20 V (diamonds), and 0.30 V
vs. SCE (triangles) as determined from the relative integrated absorbance of the pco band as a function
of time following various potential steps. The open symbols refer to data gathered during single potential
steps from -0.20 V; the residual adsorption of CO at the initial potential for each of these curves 1s
represented by the open cross. The closed symbols refer to data obtained follovvmg the partial
electrooxidation desorption of adsorbed CO at 0.3 V for 1.5 mm before stepping back to the adsorption
potential.

Inspection of the kinetic data in Figs. 12 and 13 reveals that relatively rapid CO
adsorption occurs from formic acid and especially methanol, and increasingly so at
more positive adsorption potentials, > 0 V. Comparable results have been obtained
recently for formic acid adsorption on platinum using a radiotracer method [13].
The results for methanol adsorption are more clearcut in this regard, since the single
potential-step experiments feasible with this system enable a wider range of adsorp-
tion potentials to be examined. At least for adsorption potentials, 0.1 to 0.3 V,
within the double-layer region and close to the values, 0.4 to 0.65 V, where adsorbed
CO and solution methanol undergo oxidation (Fig. 9), the data in Fig. 12 indicate
that CO adsorbs sufficiently rapidly so as to attain high coverages, 8 = 0.75, even
within the time, 10 s, required for the first SPAIR spectrum to be acquired following
the potential step.
These initial adsorption rates extracted from Fig. 12 can be converted into
effective current densities by assuming that adsorption is followed by rapid two-
electron oxidation; this procedure yields values of at least 50-100 PA cm-*. These
values are indeed sufficient to account for the current densities actually observed
here for methanol electrooxidation (e.g. Fig. 7B). Moreover, the effective CO
coverages present in the potential region where methanol oxidation occurs voltam-
161

metrically are deduced from Fig. 10 to be sufficiently small, 0 < 0.75, to correspond
to those where rapid CO adsorption is observed (Fig. 13). The observation that
facile methanol oxidation does not commence voltammetrically until potentials are
reached where the CO coverage falls below this point may well be connected with
the sluggish kinetics of CO readsorption observed at higher coverages (Fig. 13). We
therefore can deduce with some confidence that the observed kinetics of dissociative
chemisorption of methanol are consistent with a mechanism (steps 3, 4) of methanol
electrooxidation involving adsorbed CO as a reaction intermediate, at least under
the voltammetric conditions encountered here. It should be stressed, however, that
different conclusions may apply for other conditions, for example during rapid
potentiodynamic pulses.
A somewhat different conclusion, however, is reached on this basis for formic
acid electrooxidation. Figure 9 shows that, in contrast to methanol, rapid oxidation
of solution formic acid commences even at relatively high CO coverages, say
8 - 0.8-0.9. While measurable CO adsorption rates are observed at such high
coverages at 0.1 V (Fig. 12), these adsorption kinetics are insufficiently rapid to
account for the measured rates of formic acid oxidation. Admittedly, it is conceiva-
ble that more rapid adsorption-electrooxidation turnover of CO may occur at the
more positive potentials, ca. 0.3 to 0.45 V, of relevance to formic acid oxidation;
also a minority of active sites at which rapid CO adsorption occurs may in any
case remain even at high CO coverages. Nevertheless, on the basis of the present
data it appears that adsorbed CO is less likely to act as a major intermediate for the
electrooxidation of formic acid than for methanol on platinum.
Further evidence against the significant involvement of adsorbed CO in the
voltammetric oxidation of formic acid was provided by the following i3C isotope
experiment. The platinum electrode was dosed with 25 mM H13COOH in 0.1 M
HClO, and the solution replaced with the r2C isotope just before acquiring a
sequence of SPAIR spectra for formic acid oxidation using a potential sweep rate of
2 mV s-i. Only a single vco band in the absolute SPAIR spectra was observed at
a frequency, ca. 2020 cm-, consistent with the presence of adsorbed i3C0,
throughout the potential range prior to complete oxidation. The absence of the
corresponding CO band indicates that no significant quantity of fresh adsorbed
CO is formed under these conditions.
Some evidence supporting these findings is obtained from electrochemical data
on single-crystal platinum faces in that the voltammetric behaviour of methanol and
carbon monoxide oxidation is quite similar, especially on the Pt (100) and (111)
planes, yet distinct from those for formic acid oxidation [14]. Several possible
adsorbed intermediates, including COOH and COH [5,6], may be imagined to give
rise to formic acid oxidation by binding to sites vacated upon CO electrooxidation.

CONCLUDING REMARKS

We believe that the foregoing illustrates a significant mechanistic virtue of


Fourier transform infrared spectroscopy in enabling the formation and removal of
162

surface as well as solution-phase species to be monitored continuously during the


evolution of irreversible electrochemical processes, on a timescale down to a few
seconds. The simultaneous acquisition of voltammetric and infrared spectral infor-
mation in this manner enables the molecular-specific compositional information
obtained from the latter probe to have an especially direct relevance to the
electrochemical properties at hand. In this respect, such infrared methods involving
singular alterations in the electrode potential during the spectral data acquisition
offer a crucial advantage over instrumental approaches that require rapid repetitive
potential modulation.
Further applications of SPAIRS to mechanistic elucidation for a greater variety
of electrochemical processes are in progress in our laboratory.

ACKNOWLEDGEMENTS

This work is supported by the National Science Foundation. DSC acknowledges


a Purdue Chemistry Department fellowship from the Conoco Corporation.

REFERENCES

1 (a) B. Beden, C. Lamy. A. Bewick and K. Kunimatsu, J. Electroanal. Chem.. 121 (1981) 343; (b) B.
Beden, A. Bewick and C. Lamy, J. Electroanal. Chem., 148 (1983) 147; (c) B. Beden, A. Bewick and
C. Lamy, J. Electroanal. Chem., 150 (1983) 505.
2 F. Hahn, B. Beden and C. Lamy, J. Electroanal. Chem.. 204 (1986) 315.
3 (a) K. Kummatsu, W.G. Golden, H. Seki and M.R. Philpott, Lnngmmr, 1 (1985) 245: (b) K.
Kunimatsu. H. Seki, W.G. Golden, J.G. Gordon II and M.R. Philpott, Langmuir, 2 (1986) 464.
4 (a) K. Kunimatsu, J. Electroanal. Chem., 140 (1982) 205; (b) K. Kunimatsu, J. Electroanal. Chem.,
145 (1983) 219.
5 (a) K. Kunimatsu, J. Electroanal. Chem.. 213 (1986) 149; (b) K. Kunimatsu and H. Kita. J.
Electroanal. Chem., 218 (1987) 155.
6 (a) A. Capon and R. Parsons, J. Electroanal. Chem.. 44 (1973) 1; (b) A. Capon and R. Parsons, J.
Electroanal. Chem., 44 (1973) 239; (c) A. Capon and R. Parsons, J. Electroanal. Chem., 45 (1973) 205.
7 D.S. Corrigan, L.-W.H. Leung and M.J. Weaver, Anal. Chem., 59 (1987) 2252.
8 D.S. Corrigan and M.J. Weaver, J. Phys. Chem., 90 (1986) 5300.
9 P. Hollins and J. Pritchard, Prog. Surf. Sci.. 19 (1985) 275.
10 B.E. Hayden and A.M. Bradshaw, Surf. Sci., 125 (1983) 787.
11 (a) S. Gilman, J. Phys. Chem., 68 (1964) 70; (b) C. McCalIum and D. Pletcher. J. Electroanal. Chem.,
70 (1976) 277; (c) T.K. Gibbs, C. McCaIlum and D. Pletcher, Electrochim. Acta.. 22 (1977) 525.
12 For a short review, see A. Bewick in A. Fernando Silva (Ed.), Trends in Interfacial Electrochemistry,
Reidel, Dordrccht, 1986, p. 281.
13 E.K. Krauskopf, K. Chan and A. Wieckowski, J. Phys. Chem., 91 (1987) 2327.
14 C. Lamy, J.M. Leger. J. Clavilier and R. Parsons, J. Electroanal. Chem., 150 (1983) 71.

Das könnte Ihnen auch gefallen