Analisis de Rodiozonato

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

135

J. Electroanal. Chem., 213 (1986) 135-147


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

ELECTROCATALYTIC INFLUENCE OF UNDERPOTENTIAL Tl, Pb AND Bi


MONOLAVERS ON THE ELECTROCHEMICAL BEHAVIOUR OF
RHODIZONIC ACID AND TETRAHYDROXV-1,4-BENZOQUINONE ON Pt
IN ACID SOLUTIONS

G. KOKKINIDIS l, D. SAZOU and I. MOUMTZIS


Laborat o/ Physical Chemistry, Department of Chemrstry, CJniversr<v of Thessalonrkr,
Thessaloniki (Greece)

(Received 20th February 1986; in revised form 13th June 1986)

ABSTRACT

The electrochemical behaviour of rhodixonic acid and tetrahydroxy-1.4benzoquinone on bare Pt and


Pt surfaces covered by heavy metal monolayers deposited at underpotentials was studied in aqueous 0.5
M HClO, solutions. It was found that Tl, Pb and Bi monolayers catalyse markedly the oxidation of
rhodizonic acid and tetrahydroxy-1.4-benoquinone. The same underpotential layers improve the reversi-
bility of the redox system tetrahydroxy-l,4-benzoquinone/hexahydroxybenxene. The enhancement of the
overall oxidation and reduction processes has been interpreted in terms of the change of the reaction
mechanism from an “inner sphere” mechanism on bare platinum to an “outer sphere” one on the Pt
surfaces covered by underpotential layers. The two-electron oxidation of tetrahydroxy-1,4-benzoquinone
to rhodizonic acid is followed by a rapid pseudo-first-order hydration reaction, the kinetics of which were
studied by ring-disc experiments.

INTRODUCTION

Recently it was found that under-potential deposited (upd) layers of Pb, Tl and Bi
on Pt improve markedly the reversibility of some redox systems, such as quinhydrone
and pyrocatechol/o-benzoquinone [l], adrenaline/adrenalinequinone [2], o-benzo-
quinone dioxime/o-dinitrosobenzene [3] and ascorbic acid/dehydroascorbic acid
[4]. The standard potentials of these systems are very similar to the potential range
of the platinum oxide formation, and there are different views about the catalytic
role of the adatoms. Sakamoto and Takamura attributed the enhanced catalytic
activity of the Pt electrode for the oxidation of adrenaline [2] and ascorbic acid [4]
to the catalytic role of M,, (M = Pb, Tl and Bi) in providing increased coverage of

* To whom correspondence should be addressed.

OO22-0728/86/$03.50 0 1986 Elsevier Sequoia S.A.


136

OH radicals which take part in the acceleration of the subsequent chemical


reactions, such as the hydroxylation of adrenaline and its oxidation product, or the
hydration of dehydroascorbic acid. On the other hand, one of the present authors
[1,5] has interpreted the catalytic phenomena in terms of the change of the reaction
mechanism from an “inner sphere” mechanism involving adsorbed intermediates on
bare Pt to an “outer sphere” one without complications from the adsorption of the
reacting molecules on the Pt/M,, surfaces. Thus, the interest arose to study the
influence of the upd on analogous redox systems with standard potentials suffi-
ciently less positive so that adsorption of OH radicals both on Pt and on the
Pt/M(upd) modified surfaces does not come in.
Such values of standard potentials have the redox couples of tetrahydroxy-1,4-
benzoquinone (THQ) with its reduction product hexahydroxybenzene (HHB) and
with its oxidation product rhodizonic acid (RDZ) [6].

:;eH _(2c-+ZH+)_ Hy#~ :*c-+ZH+)_ yg

OH 0
(HHB) (THO) (RDZ)

The system HHB/THQ exhibits reversible behaviour at a dropping mercury


electrode in aqueous acid solutions, while the system THQ/RDZ is more com-
plicated, owing to the hydration reaction of RDZ [7-91.

(RDZ.2 H,O)

The anodic behaviour of RDZ has so far not been studied either on mercury or
on other solid electrodes.
In this paper we describe the catalytic effect of the upd of Pb, Tl and Bi on Pt on
the oxidation of RDZ and the reduction and oxidation processes of THQ in acid
solutions. The study was carried out employing cyclic voltammetric, rotating disc
and ring-disc techniques.

EXPERIMENTAL

The experimental procedure and the equipment used have been described previ-
ously [1,3]. Disc and ring-disc measurements were performed using Tacussel
rotating electrodes, type EDI-Pt and EAD 1OCKKl Pt-Pt. Pretreatment of the
electrodes consisted of mechanical polishing with diamond paste followed by
electrochemical activation by applying a continuous sweep between hydrogen and
137

just before oxygen evolution in aqueous 0.5 M perchloric acid solution. The
roughness factor of the electrodes was about 1.15, which was estimated by integra-
tion of the hydrogen adsorption-desorption current-potential curves, assuming that
210 PC cm-’ of hydrogen is adsorbed on a perfectly smooth Pt electrode. A Pt sheet
with a roughness factor of about 1.35 was also used for the cyclic voltammetric
experiments. The electrode potentials, E,, given in this paper are referred to a SHE.
All measurements were carried out at T = 298 K except where stated otherwise.
The supporting electrolyte was prepared from triply distilled water and HClO,
(Merck, suprapure). Sodium rhodizonate (puriss p.a.) and THQ (puriss p.a) were
from Fluka and were used without further purification. The other reagents used
were of suprapure or GR grade quality (Merck). Perchloric salts were prepared from
the respective oxides or carbonates and perchloric acid. Before the beginning of
each experiment, the solution was deaerated with purified nitrogen. During the
measurements a nitrogen stream was passed above the solution.

RESULTS AND DISCUSSION

Oxidation of RDZ .2 H,O on Pt and Pt surfaces covered by Tl, Pb and Bi upd layers

The effect of covering a Pt disc with a Bi(upd) layer on the polarization


behaviour of dihydrated rhodizonic acid (RDZ * 2 H,O) in 0.5 M HClO, is shown
in Fig. 1. In aqueous solutions RDZ is dihydrated [7,10]. At pH < 1, this form
predominates in the solution of sodium rhodizonate since the acid-base pKs of
RDZ .2 H,O have values of 3.9 and 4.1 [7].

.O-

‘i

-fl
5 1-E
1
pl a5Jg.0
‘6 2
.o- 0.5
3’
‘*oI; ( , ,

-0.2 0.6 1 .o 1.4


Q/V

Fig. 1. Current density-potential curves for RDZ.2 Hz0 (lo-’ M sodium rhodizonate) oxidation on a
Pt rotating-disc electrode in 0.5 M HClO4 in the absence ( -) and presence (- - -) of 5 X lo-’
M Bi(ClO4)s. ldE/drJ =lO mV s-l. Rotation frequency/ Hz: (1) 12.5; (2) 18.3; (3) 25; (4) 35; (5) 50;
(6) 75. Inset: Limiting diffusion current density id vs. w1j2 (w = 2rrf).
138

On bare Pt, RDZ ~2 H,O undergoes oxidation only. Up to the potential of


hydrogen evolution no reduction wave appears. As shown by Fleury et al. [7-91,
RDZ .2 H,O - which exhibits two adjacent cu-dioxo structures, one of the 0x0
groups being hydrated in each of these o-dioxo structures - is electrochemically
reduced at much more negative potentials than the non-hydrated RDZ without
preceding dehydration.
As can be seen from Fig. 1, the upd layer of Bi increases the electrocatalytic
activity of the Pt electrode markedly in the case of the RDZ * 2 H,O oxidation. The
oxidation wave is shifted towards more negative potentials (250-300 mV, depending
on the rotation frequency), indicating catalytic action at the charge-transfer con-
trolled potential region of the oxidation process. Independently of the presence of
Bi3+ in solution, the limiting current density increases linearly with the square root
of w = 2rf, indicating limiting diffusion control both on Pt and on Pt/Bi(upd).
From the experimental slope did/d&l2 = 7.5 x low5 A cm-2 of the linear plot (Fig.
1, inset), the diffusion coefficient D = 4.8 X lop6 cm2 s-l [8], the kinematic viscos-
ity Y = 8.929 x lop3 cm2 s-l and cRr,Z.2 u10 = lO-‘j mol cmp3, the number of
electrons consumed per RDZ * 2 H,O molecule was evaluated by means of the
corrected Gregory-Riddiford equation [ll]; it was found that n = 2.08. This value is
in agreement with electrolysis experiments carried out on a Pt net with and without
Bi(upd) present on the plateau of the oxidation wave. The only product that was
isolated from the electrolysed solution was triquinoyl octahydrate. This suggests the
following overall oxidation reaction:

a
0 0

4x
OH OH
Ho 0
OH - (2e-+2H+) OH H,O
OH OH - hydration products (3)
HO 0
OH OH
0 0

Typical Tafel-type plots of log i/(id - i) vs. E on Pt and Pt/Bi(upd) at w = 35


Hz are shown in Fig. 2. Slopes of about 138 mV decade-’ on Pt and about 33 mV
decade-’ on Pt/Bi(upd) are obtained, indicating the transformation from a totally
irreversible oxidation process to an almost two-electron reversible reaction. The
improvement of the reversibility concerns only the charge transfer and has nothing
to do with the reverse reaction. The primary oxidation product, which is probably
the dihydrated triquinoyl, is converted very rapidly to the non-reducible final
product. This can be deduced from ring-disc measurements, since no cathodic
current is detected when the ring electrode is at reducing potentials for rotation
rates up to 10000 ‘pm. Cyclic voltammetry would not help, since the capacitive
current due to the Pt oxide formation is too large in this potential region to detect
cathodic peaks at scan rates higher then 0.5 V s-l.
The reaction order with respect to RDZ a2 H,O was determined on bare Pt by
plotting logi vs. loge,, 2 H,O at different constant potentials. An average value of
dlogi/dlogc RDZ.2 n,o = 0.62 was obtained. The reaction order on Pt/Bi(upd)
cannot be determined in the same way, since the slope of the Tafel plots increases at
139

0.8

-0.8

I i i 8 ,

0.70 0.74 0.78 0.82 0.86 0.90

Q/V

Fig. 2. TafeI plots for RX-2 Hz0 oxidation on Pt and Pt/Bi(upd~ in 0.5 M HCIO, obtained from the
rotating disc data of Fig. 1. Rotation frequency f = 35 Hz.

RDZ -2 H,O concentrations higher than 10e3 mol 1-r. This presumably is due to
the competitive adsorption of the organic molecules which, as we shall see below, is
responsible for the pure reversibility of the oxidation process on bare platinum.
In order to determine the reaction order on Pt/Bi(upd), the following equation
has been used [12,13]:

i = k[RDZ * 2 H,O]”
i
%$
1 m

where k is the rate constant, m is the reaction order and [RDZ +2 H,O] is the
concentration of RDZ a2 H,O in the solution. By plotting logi vs. log (i, - Q/id at
different rotation rates and different potentials, the slope M can be determined.
Such plots, corresponding to the polarization curves of Fig. 1, are shown in Fig. 3.
They give m = 0.68 (average value) on Pt and m = 0.96 on Pt/Bi(upd). The value
m = 0.68 on Pt is in excellent agreement with that calculated from the Tafel plots,
while m= 0.96 on Pt/Bi(upd) is very close to unity.
The Tafel slopes and reaction orders provide strong evidence that the upd of Bi
hinders the adsorption of the reacting molecules, thus leading to the change of the
reaction mechanism from an “inner sphere” mechanism, involving adsorbed inter-
mediates on bare platinum, to an “outer sphere” one without complications from
140

Pt/Bi (upd)

-0.41 , , , , , 1
-1.4 -1.2 -1.0 -0.8 -0.6 -0.4 -0.2

log id - i
id

Fig. 3. Reaction order plots for RDZ.2 Hz0 oxidation on Pt and Pt/Bi(upd) in 0.5 M HC104. Data
obtained from Fig. 1.

the adsorption of the reacting molecules, which favours faster electron exchange at
the electrode-solution interface.
Different behaviour is detected on Pt/Tl(upd) and Pt/Pb(upd). As seen in Fig. 4,
the polarization behaviour of RDZ .2 H,O on Pt/Tl(upd) is characterized by the
appearance of two well-separated oxidation waves. (The increase of the anodic
current density at the end of the potential sweep is due to the onset of the reaction
Tl+ + T13+ + 2 e-.) Similar behaviour is observed on Pt/Pb(upd). However, it is
hard to accept that these waves represent two steps in the reaction mechanism. The
relative height of the first wave decreases at high rotation frequencies and at high
concentrations of RDZ * 2 H,O. This behaviour may be interpreted in terms of the
assumption that in both waves an overall two-electron process is operating. In the
first wave the oxidation occurs on Pt/M(upd) surfaces without complications from
the adsorption of the reacting molecules, while in the second wave the oxidation
proceeds with an overvoltage as on bare platinum owing to the adsorption of the
reacting molecules following anodic desorption of the upd layers. This behaviour is
141

2.0 -

1.5 -

;
E

i l.O-
‘G

0.5 -

0 I 1
0.6 1.o 1.4
EH/V

Fig, 4. Current density-potential curves for RDZ’2 Hz0 (lo-’ M sodium rhodizonate) oxidation on a
Pt rotating-disc electrode in 0.5 M HClO, in the presence of lo-’ M TlClO,. Id E/dr 1= 10 mV s-l
Rotation frequency f Hz: (1) 12.5; (2) 18.3; (3) 25; (4) 35; (5) 50; (6) 75.

to be expected since desorption of Tlfupd) and Pb(upd) layers occurs at more


negative potentials than that of Bi [14-171.

Oxidation-reduction processes of THQ on Pt and Pt surfaces covered by Tl, Pb and Bi


upd layers

In contrast to RDZ * 2 H,O, THQ undergoes both reduction and oxidation in


acid solutions. Figure 5 illustrates the current density-potential Curves for the
reduction and oxidation of THQ on a Pt rotating-disc electrode with and without
Tl(upd) present. The Tl(upd) layer lowers the overpotential of the reduction and the
first oxidation wave. Similar behaviour is observed on Pt/Pb(upd) and Pt/Bi(upd).
This happens when the concentration of Pb2’ and Bi3+ in the solution is lower than
5 x 1W5 M, because at higher concentrations complexation of THQ and its
reduction as well as oxidation products with these ions alter the polarization
behaviour of TI-IQ. The limiting current densities of the cathodic and the two
anodic waves both on Pt and Pt/Tl(upd) increase linearly with 3/Z, indicating
diffusion-controlled reduction and oxidation processes.
The reduction of THQ is a single two-electron process producing HHB. The
oxidation of THQ is characterized by two polarization waves of equal height. In the
first wave THQ is oxidized to the non-hydrated RDZ which, before its hydration,
undergoes in the second wave further oxidation to the final product. Indeed, E,,, of
the second anodic wave of THQ is * 120 mV more positive than E,,, of the
142

0 0.1 0.8 1.2


EW/V

Fig. 3. Current density-potential curves for THQ (5 X IO-’ M) reduction and oxidation on a Pt
rotating-disc electrode in 0.5 M HClO, in the absence (I) and presence (2) of lo-’ M TlClO,
IdE/dtj =I0 mV s-l; rotation frequency f = SOHz.

oxidation wave of RDZ - 2 H,O on Pt/Tl(upd) (Fig. 1) owing to the inductive effect
of the hydrated carbonyl groups.
The upd effect of Tl, Pb and Bi on the reversibility of the HHB/THQ and
THQ/RDZ redox couples on Pt was studied under potent~~yn~~ conditions.
Figure 6 presents the cyclic vol~~o~ams of THQ, in the absence and presence of
10v3 M Tl’ in the solution, as a function of the potential scan rate. Similar cyclic
voltammograms were also recorded in the presence of 5 x 1O-5 it4 Pb*+ and Bi3+
ions. The cathodic peak of THQ corresponds to an anodic peak of equal height
upon scan reversal, while the anodic peak has no corresponding cathodic peak at
scan rates lower than 0.2 V sei. This is obviously due to the rapid hydration of
RDZ. The anodic and cathodic current peaks of THQ on Pt and Pt/Tl(upd)
electrodes vary linearly with the square root of the scan rate (Fig. 6, inset),
indicating single effusion-controlled type reduction and o~dation processes.
As shown in Fig. 6, the upd of Tl markedly improves the reversibility of the
THQ/HHB redox system and the oxidation of THQ. This is readily verified by
comparing the peak potentials on Pt~~~upd) with those on bare Pt. Furthermore,
while the peak potentials on bare Pt depend linearly on logu, on Pt/Tl{upd) they
remain constant at all potential scan rates. This is evident from the E,-logu plots of
the reduction and oxidation wave of THQ (Fig. 6, inset). Also, on Pt/Tl(upd) the
difference between the peak potential and the half-peak potential for both the
cathodic and anodic branches of the cyclic voltammograms for the system
THQ/HHB and the oxidation of THQ is =c30 mV at all potential scan rates. This
143

1.5 I

I ’ 1

1.2

-0.3

7
6 -0.6
a:
E 0.6

1 I I I I
0 0.4 0.8

%I/”

Fig. 6. Cyclic voltammograms of THQ (10m3 M) on a Pt sheet in 0.5 M HCIO, in the absence (A) and
presence (B) of lo-’ M TlClO,. Scan rate 0 mV s-‘: (1) 50: (2) 100; (3) 200; (4) 300: (5) 400. Insets:
Plot of I vs. u*‘* for THQ reduction (C) and oxidation (C’) in the absence (1) and presence (2) of
TlCIO,. {lot of Ep vs. logu for THQ reduction (D) and oxidation (D’) in the absence (1) and presence
(2) of TlClO,.
144

30 0.3

6
--.
disc
5 ‘\ ring
--\ \ 0.2
4
---_ ‘4
1 1 \\
---\\\I
L--_
.\I \2
1 \
---_ 0.1.2
L\
\
s
"ring' 0.50 v
f

4
E -0.1
\
2

-0.2 1 ' '-4


0 0.4 0.8 1.2

Q/V

Fig. 7. Ring-disc measurements for THQ (5 x 10e4 M) reduction and oxidation on Pt in 0.5 M HCIO,
in the presence of 10m3 M ‘lIC10,. Id E/dt 1= 10 mV s- ‘. Rotation frequency / Hz: (1) 12.5; (2) 25; (3)
35; (4) 50; (5) 75; (6) 100.

value is in good agreement with the theoretical value of 28.2 mV for a two-electron
reversible process calculated from the equation of Nicholson and Shain [18].
The hydration kinetics of RDZ have been studied by Moiroux et al. [9] from
cyclic voltammetric experiments of THQ on mercury. They found that between pH
2 and 3, THQ is oxidized yielding RDZ*- followed by a pseudo-first-order
hydration reaction (rate constant k = (8 f 3) X lo5 [H’]’ s-l). k could not be
determined at pH -z 2 because the anodic current of THQ was masked by the
oxidation of mercury. Cyclic voltammetry on Pt/Tl(upd) does not help either, since
very high scan rates are required for this determination and at these scan rates
capacitive currents become too large.
The rotating ring-disc electrode is ideally suited for the study of EC processes
[19]. The oxidation of THQ to RDZ follows this type of mechanism. The improve-
ment of the reversibility of the oxidation wave of THQ by upd monolayers allows
ring-disc measurements on upd-modified Pt surfaces. Ring-disc experiments for
THQ oxidation and reduction on Pt/Tl(upd) are shown in Fig. 7. The ring potential
was kept at 0.45 and 0.50 V (vs. SHE), respectively, where the reduction of RDZ
(the oxidation product of THQ) and the oxidation of HHB (the reduction product
145

of THQ) are under diffusion control. The ring current was recorded as a function of
the disc potential.
The currents observed on the ring electrode when the disc potential is on the
plateau of the first anodic wave are much smaller compared to those obtained when
the disc potential is on the plateau of the reduction wave of the THQ. This is due to
the hydration of RDZ on its way from the disc to the ring electrode. As noted
above, the hydrated RDZ (RDZ * 2 H,O) is reduced at much more negative
potentials. No current is detected on the ring electrode when the disc electrode is at
potentials on the plateau of the second anodic wave, which is in agreement with
ring-disc experiments performed with RDZ - 2 H,O.
The rate constant of RDZ hydration (reaction 2) can be determined by means of
the equation

proposed by Albery and Bruckenstein [20]. In this equation, N, is the kinetic and N
the ordinary collection efficiency of the ring-disc electrode and k is the rate
constant. Figure 8 shows a plot of l/N, against l/w (w = 2?~f) for the system
THQ/RDZ at E = 0.7 V. The intercept agrees with the experimentally determined
value of N = 0.20 for the electrode from the ring-disc data of the system THQ/HHB
which is free of chemical complications. This value is very close to the value (0.21)
calculated according to Albery and Bruckenstein [21] from the geometric data of the
electrode (ri = 0.20, r2 = 0.22 and r3 = 0.24 cm). Values of N, corresponding to
rotation frequencies f between 70 and 35 Hz were taken. At f < 35 Hz, the ring
currents were too small to be determined accurately. At f > 70 Hz, parameter K,
defined [20] as K = (0.51)-‘/3(v/D)‘/6( k/w) ‘12, takes values lower than 3.5. As

120

80

Fig. 8. Plot of l/N, vs. l/o (w = 2nf) for the system THQ/RDZ on Pt/Tl(upd).
146

TABLE 1
RDZ hydration rate constants at different temperatures

T/K (v/D)“’ = 10-Z k/s-’


308 10.70 4.2 + 0.4
298 12.29 2.8*0.2
288 14.55 1.6*0.2
278 17.53 0.9*0.1

a Y is the kinematic viscosity of water. D values at different temperatures were calculated by means of
the Einstein-Stokes equation using the value of D = 4.8 X 10m6 cm* SC’ for T = 298 K.

suggested by Albery and Bruckenstein [20], values of K between 0.5 and 3.5 must be
avoided in order to be able to treat the results theoretically. From the slope
d(l/N,)/d(l/w) of different plots (one of them being presented in Fig. 8) from five
independent experiments at T = 298 K, the mean value of the hydration rate
constant was determined; it was found that k = (2.8 f 0.2) X lo* s-l *. This value
is more than one order of magnitude smaller than the value calculated for pH = 1
from the relation k = 8 X lo5 [H’]* proposed by Moiroux et al. [9] for pHs
between 2 and 3. This is not unexpected since it is reasonable to assume that this
relation does not hold at pH < 1 where RDZ undergoes hydration in the non-ionized
form, while at pH > 2 the hydration of the rhodizonate dianion predominates.
Values of k at different temperatures were also determined and are summarized
in Table 1. The analysis, taking into account the Arrhenius equation, resulted in a
value of the activation energy of RDZ hydration E, of = 9000 cal mall’.
The dependence of the kinetic collection efficiency on the potential at a constant
rotation rate may be used as an approximate indication of the effect of the situation
of the modified electrodes at different potentials on the kinetics of the hydration
reaction and the probable relation of the acceleration of this reaction with the
catalysis of THQ oxidation. Such effects caused by upd monolayers of Pb, Tl and Bi
on the hydration of dehydroascorbic acid and the hydroxylation of adrenaline
quinone following the oxidation of ascorbic acid and adrenaline were reported by
Sakamoto and Takamura [2,4]. The catalytic role of M(upd) adatoms on the anodic
oxidation of the above substances was attributed to the fact that M(upd) adatoms
increase the coverage of OH radicals which take part in the acceleration of the
subsequent chemical reactions. However, our results contradict this explanation. It
was found that the ratio i,/i, has constant values in the potential range between
0.6 and 0.8 V, where the oxidation of THQ to RDZ occurs on the Pt/M(upd) disc.
This suggests that the probable change in coverage of OH radicals on M(upd)
(M = Pb, T’l and Bi) does not affect in any way the rate of the hydration reaction
following the electron transfer. This behaviour and the fact that upd monolayers

l For this determination, the value D = 4.8 x 10e6 cm* s- ’ was taken. This value, reported for RDZ.2
H20 [8], was also used for RDZ.
147

also improve the reversibility of the redox reactions of the system THQ/HHB
(occurring in the absence of chemical complications at sufficiently negative poten-
tials, where adsorption of OH radicals both on Pt and Pt/M(upd) does not come in)
support the view of one of the present authors [l], i.e. that the improvement of the
reversibility of redox reactions by upd monolayers may be interpreted in terms of
the change of the reaction mechanism from an “ inner sphere” mechanism involving
adsorbed intermediates on bare Pt to an “outer sphere” one without complications
from the adsorption of the reacting molecules on the Pt/M(upd) surfaces. The
mechanism proposed by Motoo and Watanabe [22] based on the bifunctional theory
in electrocatalysis (i.e. the origin of the catalytic effects caused by upd foreign metal
submonolayers is the increased oxidation rate of organic molecules by oxygen-con-
taining species co-adsorbed on the adatoms) can be regarded as a satisfactory
explanation for the catalysis of the oxidation of some organic fuels which on Pt
leads mainly to kinetic currents due to the production of self-poisoning inter-
mediates. However, it does not seem to be operative for faster electrode redox
reactions exhibiting diffusion-controlled currents on platinum.

REFERENCES

1 G. Kokkinidis, J. Electroanal. Chem., 172 (1984) 265.


2 M. Sakamoto and K. Takamura, Bioelectrochem. Bioenerg., 10 (1983) 251.
3 G. Kokkinidis and N. Argyropoulos, Electrochim. Acta, 30 (1985) 1611.
4 K. Takamura and M. Sakamoto, J. Electroanal. Chem., 113 (1980) 273.
5 G. Kokkinidis, J. Electroanal. Chem., 201 (1986) 217.
6 P.W. Preisler, L. Berger and E.S. Hill, J. Am. Chem. Sot., 69 (1947) 326; 70 (1948) 871.
7 M.B. Fleury and G. Molle, Electrochim. Acta, 20 (1975) 951.
8 J. Moiroux, D. Escourrou and M.B. Fleury, Electrochim. Acta, 25 (1980) 785.
9 J. Moiroux, D. Escourrou and M.B. Fleury, Bioelectrcchem. Bioenerg., 7 (1980) 333.
10 M.B. Fleury and G. Molle, CR. Acad. Sci. Ser. C, 273 (1972) 605.
11 D.P. Gregory and A.C. Riddiford, J. Chem. Sot., (1956) 3756.
12 R.W. Zurilla, R.K. Sen and E. Yeager, J. Electrochem. Sot., 125 (1978) 1103.
13 R.R. AdZif, N.M. MarkoviC and V.B. VesoviC, J. Electroanal. Chem., 165 (1984) 105.
14 N. Furuya and S. Motoo, J. Electroanal. Chem., 98 (1979) 195.
15 B.J. Bowles, Electrochim. Acta. 10 (1965) 717, 731.
16 S.H. Cadle and S. Bruckenstein, Anal. Chem., 44 (1972) 1993.
17 G. Kokkinidis and P.D. Jannakoudakis, J. Electroanal. Chem., 130 (1981) 153.
18 R.S. Nicholson and I. Shain, Anal. Chem., 36 (1964) 706.
19 W.J. Albery and M.L. Hitchman, Ring-Disc Electrodes, Clarendon Press, Oxford (1971).
20 W.J Albery and S. Bruckenstein, Trans. Faraday Sot., 62 (1966) 1946.
21 W.J. Albery and S. Bruckenstein, Trans. Faraday Sot., 62 (1966) 1920.
22 S. Motoo and M. Watanabe, J. Electroanal. Chem., 60 (1976) 275; 69 (1976) 429; 98 (1979) 203.

You might also like