Sei sulla pagina 1di 17

JOURNALOF

ELSEVIER

Journal of Electroanalytical Chemistry 382 (1995) 111-127

The use of polarography and cyclic voltammetry for the study of redox
systems with adsorption of the reactants. Heterogeneous vs. surface path
E. Laviron
Laboratoire de Synth~se et d'ElectrosynthOse Organomdtalliques (Unitd de Recherche associ~e au CNRS 1685), Facult~ des Sciences, 6 Bd. Gabriel
21000 Dijon, France

Received 16 June 1994; in revised form 21 July 1994

Abstract

The use of polarography and linear-sweep voltammetry (LSV) for the study of a redox reaction O + ne ~ R when both O and
R can be adsorbed (Langmuir isotherm) is examined, on the basis of a rigorous theoretical treatment presented earlier for a
rotating disk electrode (r.d.e.) (E. Laviron, J. Electroanal. Chem., 124 (1981) 19 and J. Electroanal. Chem., 140 (1982) 247).
In aqueous medium on a mercury electrode, the reaction practically always occurs via the adsorbed species (surface redox
reaction). However, two cases can be distinguished, according to whether the rate of desorption of the product of the reaction (in
polarography) or of the adsorbed reactant (in LSV) is large or small when compared with the duration of the measurement (r in
polarography, R T / n F v in voltammetry). In the first case, the reaction appears as heterogeneous, with an apparent rate constant
khm, which is much larger than the normal constant kh, and which can be determined by using the classical theories for a
heterogeneous reaction. In the second case, the reaction has a "surface" character, and the electrochemical surface rate constant
k s can be determined by using the appropriate theories. The domain for each reaction can be represented by using adsorption
diagrams log~- or logv vs. log(bobR) 1/2 (b o, bR; adsorption coefficients). The advantages of using polarography and cyclic
voltammetry rather than r.d.e, voitammetry for the study of the above systems are discussed; they are theoretical (non-steady-state
nature of the methods) as well as experimental (use of the dropping mercury electrode).
Keywords: Polarography; Cyclic voltammetry; Adsorption

I. Introduction

A d s o r p t i o n o f the species p a r t i c i p a t i n g in an elect r o c h e m i c a l r e a c t i o n is e x t r e m e l y f r e q u e n t , e s p e c i a l l y


in t h e case o f o r g a n i c c o m p o u n d s [1,2]. This is p a r t i c u larly true in a q u e o u s solutions, b u t a d s o r p t i o n in n o n a q u e o u s solvents, a l t h o u g h w e a k e r t h a n in w a t e r , can
still b e a p p r e c i a b l e [3-7]. T h e g e n e r a l s c h e m e for a
simple r e a c t i o n is given in Fig. 1, in which Oad s a n d
Rad s a r e the a d s o r b e d forms, Oso I a n d Rso I t h e u n a d s o r b e d forms n e a r t h e surface, 0 a n d R t h e forms in
t h e b u l k s o l u t i o n away f r o m the e l e c t r o d e . T h e surface
c o n c e n t r a t i o n s a r e F o a n d FR, t h e c o n c e n t r a t i o n s in
solution Co(X,t) a n d CR(X,t) , C a n d c b e i n g t h e i r
value for x ~ o0. T h e surface a n d h e t e r o g e n e o u s elect r o c h e m i c a l r a t e c o n s t a n t s a r e respectively k s (in s - l )
a n d k h (in c m s - l ) . In theory, a possibility o f d i r e c t
e l e c t r o n e x c h a n g e b e t w e e n the a d s o r b e d a n d n o n - a d s o r b e d systems is possible; however, we shall not consider it h e r e , b e c a u s e it can be n e g l e c t e d in most
p r a c t i c a l cases [8-10].
0022-0728/95/$09.50 1995 Elsevier Science S.A. All rights reserved
SSDI 0022-0728(94)03684-5

C o n s i d e r e d in its generality, the m a t h e m a t i c a l t r e a t m e n t of t h e p r o b l e m (solution o f t h e diffusion e q u a tions with a d e q u a t e initial a n d b o u n d a r y c o n d i t i o n s ) is


very complex, as far as t r a n s i e n t e l e c t r o c h e m i c a l m e t h o d s a r e c o n c e r n e d . A full discussion of the diverse

/
/
/
/
~//
O/
b~ ,/ /
~ /
~/
//
//
/

F0

c o(O,t)

Oads ~

ks(s4)

Rads~
FR

c o (x,t)

Oso I ~

"~

kh(Cm s l )

RsoI ~
CR(O,t )

o --~)
Co(X
0

transport

~
C R(x,t)

Fig. 1. The reaction scheme.

R
C~, (x .-),at)

E. Laviron / Journal of Electroanalytical Chemistry 382 (1995) 111-127

112

works concerning these methods can be found in Ref.


[11]. The mathematical difficulties are illustrated, for
example, by the theories for faradaic impedance (in
particular by the works of Delahay et al., Barker,
Sluyters-Redbach, Sluyters et al.; cf. the 30 references
quoted in Ref. [11]) or by Wopschall and Shain's classical treatment in cyclic voltammetry [12], which gives
complex results, although it is restricted to the case of
a reversible reaction. Analytical solutions can be obtained only in a few limiting cases, such as that of a
linear isotherm with an immobile electrode, or for
Frumkin or Langmuir isotherms when both O and R
are adsorbed strongly (surface system) [11], or when
only one of them is adsorbed, the other being present
in the solution in high concentration [13,14].
However, we showed in two previous papers [15,16]
that the problem can be solved simply in the case of
voltammetry on a rotating disk electrode (r.d.e.), which
is a steady-state method. Easy-to-interpret analytical
solutions are obtained, when the rate of the adsorption
reaction itself is assumed to be so fast that it is not
rate-controlling, and when a Langmuir isotherm is
obeyed. Both conditions are usually fulfilled in the case
of mercury electrodes, on which the adsorption rates
are usually so high that they cannot be measured by
existing methods (see Ref. [17] and more especially the
discussion by Delahay [18a]), while langmuirian conditions can be attained at low surface concentrations
[11]. On solid electrodes, by contrast, adsorption can
be slow and often involves the formation of chemical
bonds between the molecules and the surface
[2,18b,19,20], and difficulties arise owing to the heterogeneous nature of the surface [18c,19].
Transient methods, such as polarography and
linear-potential-sweep voltammetry (LSV) are much
more advantageous to use, firstly because the nonsteady nature of the measurements allows the reactions
to be studied in a much larger range of adsorption (see
below), and also because of practical reasons (reproducibility and ease of use of dropping or pseudo-stationary mercury electrodes, compared with the difficulty of obtaining good mercury films on the r.d.e.,
larger span of "measurement" times, etc.). We examine in this paper, on the basis of our previous results
for the r.d.e., how polarography and LSV can be used
to study the problem.
We shall consider the case of a reduction; the results can easily be transposed to that of an oxidation
when needed.
In general, the reduction or oxidation current for an
organic compound is given by an equation of the form
(reduction currents are taken as positive)
i = nFAk~

[~:o e(-anF/RTXE-Ee)

sCRe((l-a)nF/RTXE Ee)]

(1)

which holds either for a heterogeneous reaction


( w i t h k e = k h , ~ o = CO, ~R = CR, and E e is the equilib-

rium potential) or for a surface reaction when a Langmuir isotherm is obeyed [11,21-26] (with k e = ks, ~o =
Fo, ~:R = FR, and Ee is the surface equilibrium potential). Theoretically a can have any value; in practice,
however, three cases can occur, i.e. n = 1, a = 0.5, or
n = 2 with a = 0.25 or 0.75 [23-26].

2. Rotating-disk electrode: summary of the main results


2.1. Generalization of the theory for any value of a
In our previous study [15,16], we defined the problem for any value of a, but the equations were solved
only in the particular case when ot = 0.5. In view of Eq.
(1), we need to generalize the results for any value of
OL

For a simple heterogeneous reaction without adsorption, we have the equation (Eq. (7) of Ref. [15])
X

3'[(1 - X ) O -'~ --XO 1-a ]

(2)

in which X is the dimensionless concentration of the


reduced form near the electrode
C

CR(O,t)//C T

cS//CT

(3)

i.e. the concentration near the surface, which is


constant here, divided by the total concentration c T of
O and R in the solution (here, in the case of a
reduction, the concentration of O). The parameter 0 is
defined by

Co( O,t )
0 - CR(o,t------~

cs
cS

exp[(nF/RT)(E-Ee) ]

(4)

and 3' is the dimensionless electrochemical rate


constant
y = kh~$D -1

(5)

where 8 is the thickness of the diffusion layer, and


D is the diffusion coefficient.
The dimensionless current I is given by
i

I = nFAcTDS- 1

(6)

When the reactants can be adsorbed as shown in


Fig. 1, the total current, which is the sum of the surface
current i a and the heterogeneous current i h, is given
by Eq. (21) of Ref. [15], when a Langmuir isotherm is
obeyed. The transfer coefficients are assumed to be the
same for both reactions; although they could be different in theory, they have been found to be equal in the

E. Laviron /Journal of Electroanalytical Chemistry 382 (1995) 111-127

diverse experimental cases which we have examined


[23-26].

7 !

i = ia + ih = nFAD3-1c s
= nFAks(FO~7 -~ - F R r / ' - - )

+ n F A G ( c S O - " - cSO 1- ~ )

(8)

in which b o and b R are the adsorption coefficients


of O and R.
Assuming that the adsorption equilibrium is always
established and that a Langmuir isotherm is obeyed,
and proceeding as in Ref. [15], we easily obtain:
I = X = y M [ ( 1 - X ) 0 - " + x O 1-~]

(9)

with
or

b_lr~_l/Z+rZ,,(l_x)+r_2(l_,O

(10)

The dimensionless parameters or, b and r are defined by ( F m maximum surface concentration, assumed
to be equal for O and R):

ksrm

or = - khC T

(11)

b = C T ( b o b R ) 1/2

(12)

r = ( b o / b R ) '/2

(13)

Comparison of Eq. (9) with Eqs. (2) and (6) shows


that the global reaction is equivalent to a heterogeneous process with an apparent dimensionless rate
constant My, i.e. in view of Eq. (5), an apparent rate
constant khM defined by:
k hM

Mk h

(14)

Since M is always larger than one, the apparent


reversibility of the reaction is always increased. Generally speaking, M varies along the wave, since it depends on the current I = X (Eq. (10)), and also depends on c T through or and b; this point will be
discussed below in Section 2.4.
2.2. The three situations

Our previous study [15,16] led to the conclusion that


three types of situations can be distinguished (Fig. 2).
In the first situation (Fig. 2, case I), there is no
adsorption, and only the heterogeneous reaction (rate
constant k h) takes place. As shown in Ref. [15] (see
also below), this case should be very seldom encountered, if ever, in aqueous medium, but can exist in
non-aqueous solutions.
In the second situation, part (Fig. 2, case IIa) or all
(Fig. 2, case lib) of the reaction takes place on the

"

froz

film

b I

11

= (bo/bR)O

m=l+

(7)

with [11,15,27]

113

kh

III

khM = Mk h

Fig. 2. Schematic representation of the reaction scheme for the three


situations on the r.d.e. (---,) mass transport; ( ~ ) electrochemical
reaction.

surface. However, because the desorption rate of R is


large enough [15,16], the molecules of O are transported towards the electrode, they are reduced, and
then are transported away from it, which is similar to
what occurs for a normal heterogeneous reaction. As
shown above, the global reaction is indeed equivalent
to a heterogeneous process, with the apparent heterogeneous rate constant khM.
The third situation arises when the desorption rate
of the molecules of R becomes so slow that we have an
immobile ("frozen") film on the electrode [16] (Fig. 2,
case III). On a rotating-disk electrode, the electrochemical reaction is again heterogeneous in the sense
that the molecules of O are reduced without being
adsorbed, in the presence of the film; however, the film
can have an influence on the rate (autoinhibition effects [16,27]), so that the rate constant can be different
from k h, and its value cannot be predicted.
2.3. The value o f k s / k h

Quantitative calculations, using the equations which


we have derived, require the knowledge of the ratio
k h / k s. Brown and Anson [28] proposed the formula
k J k h = 6 108 cm - l

(15)

calculated on the basis of the theory of Marcus,


assuming that the reorganization energies were equal
for the surface and heterogeneous processes. More
rigorous calculations made by Mohilner [29] yield a
value of about 2 x 109 cm -~ for the ratio k s / k h [15].
Another approach, based on a concentration effect (a
volume concentration corresponding to the layer of
adsorbed molecules is calculated, and it is assumed
that the rate constant is k h) yields for the ratio about
2 x 10 7 c m - 1 [15]. Experimental values of k S, obtained
recently by indirect methods [21,24-26], do indeed
have the order of magnitude predicted by these theories. The ratio k s / k h should in any case be very large,
and the differences mentioned above do not change
the theory and the equations derived below; in particular, the limits between cases II and III do not depend

114

E. Laviron /Journal of Electroanalytical Chemistry 382 (1995) 111-127


~

H (case I1 )
75

ff 2s

A
-25
-15

-10

p
-s

0.01 0.5 0.99


0

S (case II1 )

/D

found easily by noticing that 1 in Eq. (17) then becomes negligible, so that:

l o g m = l o g ( k s / k h ) + l o g F m + log(bobR) L/2
+ (1/2 - a)logr

(19)

When, in contrast, we always have

b l r " - ' / 2 < < r 2 " ( 1 - X ) +r-2(1-~)X

(20)

Eq. (10) yields


5

0.5 log (bo b ~ ,'cm ~ tool -I )

10

15

\ 0 . 5 log ( b o b ~ )i

ks/'m
M = k h C x [ r 2 , ( 1 - - X ) + r-2'l-'~)X]

(21)

'0.5 log (b~ b R )/

Fig. 3. Variations of logM with log(bobR) 1/2 (Eq. 3; see text for the
complete discussion) for ks/kh=6lO 8 cm 1, D = 4 1 0 6 cm2
S - 1, F m = 5 10 -I mol c m - 2 ' r = 1 and X = 0.5: (
) CT 1 0 - 7
mol c m -3, ( - - - - - )
c T = 10 - 6 mol cm 3. The limit between zones
II and III was calculated for LSV (Eq. (76)) with c = 0.05 V s- ~.
=

on it. We shall use Eq. (8) in what follows when the


ratio k J k h is needed.

This corresponds to branch C, which is a horizontal


asymptote, since M no longer depends on bob R. In
this region, M d e p e n d s on the concentration and on
the potential (through X); because of these features,
the product M k h can no longer be considered as a
useful a p p a r e n t constant. We shall therefore establish
the conditions for which Eqs. (17)-(19) are valid.
T h e intersection of the straight lines C and D (cf.
Eqs. (19) and (21)) easily gives (Fig. 3)

log(bobR)l/2 = -logc. v + (1/2 - a)logr


2.4. The variations of M

with (bobR)1/2

- log[ r2"(1 - X ) + r 2(,-"'X ]

A n example of curve l o g M vs. log(bobR) 1/2 is shown


in Fig. 3 for two values of c v and for r = 1. All the
curves are similar; only the height of the plateau on the
right differs from one curve to the other when ca-, r or
a changes (see below). T h e y are m a d e up of three
branches A, B and C. Branch A is obtained when b (or
b o b R) becomes small e n o u g h (no adsorption, case I of
Fig. 2).
W h e n b (or b o b R) increases, two situations can
arise. T h e term r2~(] -- X) + r - Z ( 1 - ~ ) X in Eq. (10) varies
monotonically from r 2 a to r - 2 0 - a )
when X varies
from 0 to 1, i.e. when the potential varies from + ~ to
-~. We can thus find conditions (branch B) such that,
whatever the potential,

(22)

For a value log(bobR)~/2 equal to:


log(b o b R))/2 = log( bob R ) y 2 _ 0.7

(23)

the curve l o g M will be 0.09 log unit below the


straight line D (Fig. 3). We shall take this value as the
limit for the validity of Eq. (17).
Since the last term of Eq. (22) varies monotonically
with X, it suffices, to calculate the limit, to choose
whichever of the values X = 0 or X = 1 that yields the
smaller value for that term, to ensure that M remains
constant along the wave. W h e n X = 0, Eq. (23) becomes
log( bob R) ~/2 = _ 1OgCv _ 0.7 - (0.5 + a ) l o g r

(24)

whereas for X = 1
b lr'~ 1/2 >> r 2 ~ ( 1 - X) + r

2(1 a)X

(16)
log( b o b R ) ]/2 = _ log CV -- 0.7 + (1.5 - a ) l o g r

(25)

Eq. (10) then reduces to

M = m = 1 + o'br 1/2-~
= 1 + ( k s / k h ) F m ( b o b R ) l / 2 r 1/2-~

(17)

T h e reversibility coefficient m then d e p e n d s neither


on X (which means that it is i n d e p e n d e n t of the
potential E), nor on the analytical concentration c T.
T h e constant khM, now written
khm =

mk h

(18)

is thus a true a p p a r e n t constant, since its value does


not vary with E or c m.
In a large range of log(bobR) 1/2 values, logm varies
linearly (Fig. 3). T h e equation of the straight line D is

Eq. (24) must then be retained when r > 1, and Eq.


(25) when r < 1. Examples of limits, which do not
differ much when a changes, are shown in Fig. 4.
Actually, for large values of (bobR) t/2, the plateau
no longer has a physical meaning, since the film becomes frozen (case III, Fig. 2). However, this has no
bearing on the above derivation of log(bobR) 1. T h e
transition between cases II and III is studied below in
Sections 3 and 4. This transition, calculated from Eqs.
(76) or (78), in which z = l
and v = 0 . 0 5 V s -1, is
shown in Fig. 3.
We must here c o m m e n t on the values f o u n d experimentally for b o / b R. Few data, if any, concerning
aliphatic c o m p o u n d s are available, but in the case of

E. Laviron/ Journal of ElectroanalyticalChemistry382 (1995) 111-127

115

From Eqs. (10) and (27) we obtain


p = 1 -M-'
.---,

The reaction path becomes independent of E and


c T in the region, defined above, where M = rn (Eq. 17).
The values of log(bobR )1/2 for which p is equal to
0.01, 0.5 and 0.99 when, for example, r -- 1 are respectively - 2.52, - 0.52 and + 1.47; they are shown in Fig.
3.

%
E
2L.
/

log C T increases
by 1 unit

.g

"-

-2

-I

3. Heterogeneous vs. surface process in polarography

log b---9--
bR
Fig. 4. Limiting value log(bobR)~/2 for which Eqs. (17)-(19) are
valid, as a function of log(b o / b R ) . D = 4 X 10-6 cm 2 s-1, /'in = 5 X
10 10 mol cm-2: (
) a=0.5; (------)
a=0.75; ( ..... )
a = 0.25. The limits are shifted downward by one unit when logc T
increases by one unit.

aromatic compounds this ratio is usually not very different from unity, which is probably due to the fact
that the structure of the molecule, and in particular its
aromaticity, does not change very much when it is
reduced or oxidized. A large difference between b o
and b R should cause an adsorption wave or peak to
appear separately from the diffusion wave or peak
[27,30] when the reaction is reversible, since in the case
of a Langmuir isotherm, the standard surface potential
E ' is related to the "normal" standard potential E
by [11,31,32] (cf. Eq. 8)
E '= E - (2.3RT/nF)log(bo/bR)

(28)

(26)

Pre-waves have indeed often been observed [11], but


they are due to the influence of the interactions between the adsorbed molecules [11,33] rather than to a
large value of b o / b R. Ratios found experimentally for
redox pairs on mercury are: methylene b l u e /
leucomethylene blue, 0.1 [33]; riboflavine and rosinduline G G / r e d u c e d form, 1 [33,34]; nitrosobenzene/
phenylhydroxylamine, 0.14 [35]; nitro g r o u p / d i h y droxylamino group for 4-nitropyridine [24], 4-nitropyridine-N-oxide [25] and 4-nitroacetophenone [26],
= 1. Exceptions can be found, however; for example,
when the reduction occurs in the vicinity of the limits
of the adsorption domain [36].

All the equations given above can be applied directly in polarography in case II of Fig. 2, since the
general process (diffusion-reduction-diffusion) is the
same.
However, polarography can also be used in case III
(strong adsorption, slow desorption of R), because of
its non-steady-state nature. The molecules of O diffuse
towards the electrode, they are adsorbed and reduced;
the molecules of R remain attached to the surface (Fig.
5); this is true whatever the isotherm, since the only
condition is that the concentrations co(O,t) and CR(O,t) ,
which are in equilibrium with the surface concentrations, remain much smaller than the analytical concen-

Oads

III
kS

2.5. The reaction path


The reaction path is defined as the percentage of
the reaction going via the surface path, i.e. as the ratio
of the surface current i a to the total current i = i a q- i h.
It is given by [15]
or

P = ' + b - l r c ~ - l / 2 + r 2 ~ ( 1 - X ) +r-2(1-a)X

(27)

R f

ads

Fig. 5. Reaction scheme for strong adsorption in polarography (cf.


case III of Fig. 2): ( ~ ) diffusion; ( - ~ ) electrochemical reaction
(reduction).

E. Laviron /Journal of Electroanalytical Chemistry 382 (1995) 111-127

116

tration ca-, so that the transport of the molecules of O


is diffusion-controlled. This process can go on as long
as the electrode is not completely covered by a film of
O and R, or R. The time t m needed to reach full
coverage is given by Koryta's equation [27,37]

t m = 1.82F2/Dc 2

(29)

The theory for this type of situation has been published previously for a reversible reaction [27,32] and in
the general case of any degree of reversibility [38].
It is thus of interest to define the conditions for
which the reaction practically appears as a purely heterogeneous or as a purely surface process.

Frumkin isotherms become linear, in order to avoid


the influence of the interactions between adsorbed
molecules.
We have also:

Co(O~,t ) -~ c~

(38)

CR(O%t) ~ 0

(39)

The current i is given by

+D[ ac(x't) ]
dt
ax
x =o

dF

nFA

3.1. Reversible reaction

[ acR(x't) ]

dFR
-

dt

ax

(40)
x=0

The problem, which can be studied by solving the


diffusion equations

The problem can easily be solved by using the


Laplace transformation [39,42]. We obtain for the concentrations near the electrode

aC0

Co(0,t ) = CT0(1 + 0)-111

at
aC R
--

at

a2Co
O -ax 2

--

exp(A2)erfc(A)]

(41)

(30)
cR(0,t ) = CT(1 + 0)-111 -- exp(A2)erfc(h)]

02CR
- D - -

(31)

ax 2

for a planar, immobile electrode, is similar to the


problem of the adsorption of an electroinactive substance, which has been treated by Delahay and Trachtenberg [39], who showed that the results can practically
be applied directly to the dropping mercury electrode.
We shall consider the case of a reduction; transposition to the oxidation is obvious.
The initial conditions

by

and we get for the dimensionless fluxes (flux divided


CTD1/2~"- 1/2t- 1/2) at the surface

q~o = Y- 1D

x=0

F(A)
1+0

(44)

F ( A ) = 7rl/2h2exp(A2)erfc(A)

CR(X,0) = 0

(33)

A = KR(1 +

express that there is no adsorption at time t = 0; the


adsorption process will take place as soon as the electrolysis begins, as is the case in polarography, since the
drop does not exist at t = 0.
The boundary conditions are

cTD1/2
y = ~1/2tl/2

O 1 / 2 t l / 2 ( 1 + O)

KoO/KR )

(45)

O1/2t1/2(1 + O)

= bRFm(1 +

_ Eo)]

~0Oeq

Fo = boFmco(O,t ) = Koco(O,t )
F R = bRFmCR(O,t ) = KRCR(O,t)

(46)

(47)

dro
r20[1 - F(A)]
-dt
1 + r20

= y-1 --

)CCR((00,'tt ) = exp [ ~--~(E


nF

r20)

We can also easily derive from Eqs. (36),(37) and


(41),(42) the equivalent fluxes corresponding to the
variations of F o and F R

(34)

(no adsorption at time t = 0)

(43)

1+0

in which

(32)

aCo

= 1

(ac R ]
q~R=y-ID~-~X ]x= o

Co(X,0) = c v

--~-x )x=odt+Dfo(~x )x=odt

(42)

(48)

(35)

q~R~q = y - i d F a _ 1 - F ( A )
dt
1 + r20

(36)

In view of Eqs. (40), (43), (44), (48) and (49), the


dimensionless current I is given by

(37)

We use here a linear isotherm, which will not limit


the validity of our conclusions; we have indeed shown
previously [11,33,40,41] that the experiments must be
conducted at low coverages, when the Langmuir or the

7rl/2tl/2i

= nFAD1/2c T

(49)

1 -F(A)

F(A)

1 + r20

+ 1+ 0

(50)

This is the general equation of the wave for any


degree of adsorption.

E. Laviron /Journal of Electroanalytical Chemistry 382 (1995) 111-127

For strong adsorption, if b o is not too different


from bR, )t ----)0, F ( M ~ 0, and
1
I ~ - 1+r20

(51)

1+~7

Eq. (51) has been derived earlier [27,32].


For weak adsorption, h ~ 0% F(A) ~ 1, and
1
I -, - 1+0

(52)

which is indeed the equation of a normal "heterogeneous" polarogram [43].


We can use the equations derived above to express
that, as soon as they are formed, the molecules of R
leave the electrode (weak adsorption, cases I or II, Fig.
2), or remain adsorbed (strong adsorption, case III,
Fig. 5). Two criteria can be used.
(1) We can directly compare the flux of R to that of
O, i.e. form the quantity

be chosen arbitrarily. However, it is more advantageous to use the second (Eq. (54)); generally speaking,
both x' and x vary along the wave, since they depend
on the potential via 0, which appears in the equations
directly, and indirectly through F(A). This means that
the "surface" character of the current varies along the
wave. Calculations show that the variation is smaller if
the second criterion is chosen.
The ratio x defines the percentage of molecules of
R remaining adsorbed. If, for example, x = 0.01, 1% of
the molecules will remain adsorbed, i.e. 99% will diffuse away from the surface; we can speak of weak
adsorption. If, in contrast, x = 0.99, the situation is
reversed (strong adsorption).
In general, as mentioned above, x varies along the
wave, since h depends on the potential (cf. Eqs. (46)
and (54)). Let us first, however, examine the case when
b o = bR; then A, which becomes
DW2tl/2

=)to
= 1 + 0 - OF()t)

(53)

(2) We can compare the amount of R which remains


adsorbed to the total amount which is formed
1

Req

x =
~Req

-~- ~0R

1+

(54)

(1 + r20)F(h)
(1 + 0)[1 - F ( ) t ) ]

For weak adsorption, F()t) --) 1; then x' ~ 1, so that


the flux of R equals the flux of O (no molecule of R
remains adsorbed), and x --) 0 (~Req << q~Req+ q~R; all
the molecules of R diffuse away from the electrode).
When the adsorption is strong, F()t)--)0, x'--)0 (the
flux of R is much smaller than the flux of O; R remains
adsorbed), and x - ) 1 (which expresses the same thing,
since ~R -* 0). Therefore either of the two criteria can

117

b R IVrn

(55)

is independent of the potential, as well as x, which


is equal to 1 - F ( ) t o ) , so that
F()to) = 1 -x

(56)

For a given value of x, the value of )to can thus be


determined from tables [44] of the function F. For
example, when x = 0.5, )to = 0.43; when x = 0.01, )t o
= 7.3; when x = 0.99, h o = 0.0175, etc. Eq. (55) can be
written
logb R = log(Dl/2/Fm) - log)t o + 0.51ogt

(57)

We shall rewrite it in a more general form, analogous to that of other equations derived below. Since

bR = ( bobR) W2( bR/bo ) 1/2

(58)

Eq. (57) becomes


logt = 21ogh o - 21og(D1/e/Fm)
+ l o g ( b o / b R ) + 2log( bo/b R)1/2

0.5 log[( b o b R )v2/cm 3mol_l]= log[b / cm3mol.1]= log[bR / cm 3 mol-1]


Fig. 6. Polarographic adsorption diagram for a reversible reduction
for D = 4 1 0 - 6 cm 2 s -1 and F m = 5 1 0 -1 mol cm - 2 (Eqs. (57)
or (59)). T h e value of x is shown on each curve.

(59)

This equation (in which log(bo/b R) = 0) allows us


to define an adsorption diagram log(bobR)t/2/logt (or
log~', r being the drop-time) in which three zones can
be defined, i.e. a zone of weak adsorption H in which
the reaction appears as heterogeneous, an intermediate zone I, and a zone of strong adsorption S, in which
the reaction is of a surface nature. Such a diagram is
shown in Fig. 6, for representative values of D and F m
equal to 4 10 -6 cm 2 s -1 and 5 10 - 1 0 mol c m - 2
respectively.
The graphs for b o ~ b R can be obtained as follows.
One easily gets from Eqs. (44), (49), (50) and (54)

F()t) =

[1 - x I ( 1 - r2)] (1 - x )
1 - x ( 1 - r 2)

(60)

E. Laviron /Journal of Electroanalytical Chemistry 382 (1995) 111-127

118

0.01

For a totally irreversible reduction, the half-wave


potential for a heterogeneous reaction is given by [45]

2.3RT
E1/2-E - --logO.886(khmzl/2D-1/2)
anF

/
/

(62)

in which khm is the experimental constant which is


actually measured (cf. Eq. 17).
For surface process we have, if the adsorption obeys
a Langmuir isotherm [38],

2.3RT
El~ 2 - E '
I
5

-2

0.5 log [( b o b a )I/2 / cm 3 mol-I

anF

(63)

or, in view of Eq. (26)

2.3RT

log( 1.197ksz )

_~]

El~z-E- ~nF lg[l'197ksr(b/bR)

log(b R / m 3mol-l)for b o / b R = 0. I
Fig. 7. P o l a r o g r a p h i c a d s o r p t i o n d i a g r a m for a reversible r e d u c t i o n
( x = 0.5) for diverse v a l u e s of b o / b R (shown on each curve). T h e
abscissae can be c h o s e n as log(bobR) 1/2, which is valid for all the
curves, or I o g b a , which m u s t be d e f i n e d for each curve (an e x a m p l e
is given for b o / b R = 0.1).

(64)

The transition (x = 0.5) between the heterogeneous


(for large values of z) and the surface reactions (for
small values of r ) takes place when Eqs. (62) and (64)
are equal, which gives
logz = - 2log(ks~kin) - 1og(1.830)
+ 2 a log( b o / b R )

(65)

or, in view of Eqs. (18) and (19)


We can thus calculate F(A) for given values of r 2, I
and x, and get the corresponding value of A from
tables. Then we calculate 0

I(1 - x )

+ 2 1 o g ( b o / b R)1/2

(66)

with

F(A)
0

logr = A + (0.5 + a ) l o g ( b o / b R )

(61)

A = 21og(0.74FmD-1/2)

(67)

For oxidation, we have


and deduce successively b R from Eq. (46) and b o
from the ratio r 2 -----b o / b g . The calculations were carried out for D = 4 10 -6 cm 2 S -1, /'m ---~5 X 10 -1 mo1
cm -2 and 0.01 < r 2 < 100.
The results show that, for a given value of x, the
limit does not change much along the wave (0.2
log(bobR) 1/z units when one passes from Ej/4 to
E3/4), which means that it will practically still be a
straight line. We have shown in Fig. 7 the limits obtained for x = 0.5 and I = 0.5 (E1/2). They are shifted
to the right when r > 1, to the left when r < 1, by
about 0.5 log units for a tenfold increase or decrease of
r. These graphs are valid for reduction; for oxidation,
the values of b o and b R must be exchanged.

3.2. Totally irreversible reaction


Resolution of the diffusion equations is very complex in this case. Therefore we have used another
method, based on the comparison of the half-wave
potentials, for large values of z, when the reaction will
tend to be heterogeneous, since there is more time for
the reduced molecules to desorb, with those for smaller
values, when the reaction will be of a surface nature.

log r = A - ( 1.5 - a ) log(b o / b R) + 2log( b o / b R ) 1/2

(68)
Eqs. (66) and (67) can also be written
log~" = A + ( a - 0 . 5 ) l o g ( b o / b R ) + 21ogb~

(69)

with b i = b o for a reduction, and b i = b R for an


oxidation.

4. Heterogeneous vs. surface process in cyclic voltammetry


Cyclic voltammetry can be applied directly to the
study of case II of Fig. 2, using the classical theories
[43,46] for a heterogeneous process whose rate constant is khm (cf. Section 2).
Moreover, because of its non-steady nature, it can,
like polarography, be used to study the surface reaction
(cf. case III, Fig. 2), which is not possible on the r.d.e.
On a pseudo-stationary Hg electrode, the potential
scan is started at a time t I (delay time) after the
formation of the drop; during this time, a certain
amount of O has adsorbed at the surface. When the
sweep rate is large enough, the current due to the

E. Laviron/ Journal of Electroanalytical Chemistry 382 (1995) 111-127

119

In this equation, c T is in principle the analytical


concentration (its meaning will be discussed later), and
v is the sweep rate.
For a surface reaction when the adsorption obeys a
Langmuir or a Henry isotherm [11,27]
ips = 0 . 2 5 ( n F ) 2 ( R T ) - 1 A F T v

(71)

where F T is the surface concentration at the start of


the sweep.
Let

ips
--

(72)

iph

We shall designate by u z the corresponding sweep


rate.
We get
L'z = 3 . 2 4 ( R T / n F ) DcZF~r2z 2

R~ads
Fig. 8. Reaction scheme for a strong adsorption in linear potential
sweep voltammetry (cf. case III of Fig. 2). ( ~ ) diffusion; (-I~)
electrochemical reaction (reduction).

adsorbed molecules will be much larger than that due


to the molecules brought by diffusion to the electrode.
In other words, the number of molecules of O diffusing
towards the electrode, or of R diffusing away from the
electrode, will be much smaller than that of the adsorbed molecules (Fig. 8). The reaction will be of a
purely surface nature; the theory for this case has been
published previously [11,27,32,47]. Solutions having formally the same mathematical expressions have also
been obtained for the case when only one of the two
forms O and R is adsorbed, the other being present in
the solution in high concentration [13,14].
It is thus of interest to determine the conditions in
which the reaction appears as a heterogeneous (weak
adsorption) or as a surface process (strong adsorption).

4.1. Reversible reaction

In Eq. (71), F T is the surface concentration Fo(t t)


of O at the start of the potential scan, i.e. at time t]
after the beginning of the drop formation (or more
generally the beginning of the adsorption process).
Since the adsorption is at equilibrium, the concentration of O near the surface in the case of a Henry
isotherm is

Fo( t 1) = boFmco( O,t 1)

iph = 0.45( h E ) 3 / 2 ( R T ) -I/2ACTD1/2 u 1/2

(70)

(74)

At first sight, we can thus equate c T and F T in Eqs.


(70) and (71) to Co(0,t 1) and F(t]) respectively. Strictly,
however, Eq. (70) should not be used, since its derivation [43,46] implies that the solution is homogeneous at
the start of the sweep, whereas it is not, because of the
adsorption [39].
However, (a) for a weak adsorption, co(0,t l) will
have returned to a value not too different from c T at
time t t [39], and Eq. (70) will be approximately valid at
slow sweep rates, whereas at high sweep rates, Eq. (71)
will be applicable, since, as explained above, only the
adsorbed molecules are reduced, and (b) for a strong
adsorption, the current due to the adsorbed molecules
will predominate, even at slow sweep rates.
If we substitute, in Eq. (73), c T by co(O,t 0 and
F T = Fo(t 1) by their values taken from Eq. (74), we
obtain
logcz = B + 21ogz - 21ogb o

A rigorous analytical solution cannot be obtained.


We shall assume that the " h e t e r o g e n e o u s " and the
surface currents are independent; at slow sweep rates
only the heterogeneous current will indeed be observed, whereas at high sweep rates only the surface
current will be seen.
For a heterogeneous process, the p e a k current is
given by [43,46]

(73)

(75)

or, since b o = (bo/bR)l/Z(bobu) 1/2


logv z = B + 21ogz - 21ogr - 21og(bobR) 1/2

(76)

For oxidation, we get


logv z = B + 21ogz - 21ogb R

(77)

or

logv z = B + 21ogz + 21ogr - 21og(bobR) ]/2

(78)

E. Laviron/Journal of Electroanalytical Chemistry 382 (1995) 111-127

120

I
2

I
3

o b t a i n e d a r e s o m e w h a t different. This type of result is


not u n e x p e c t e d ; it is found, for e x a m p l e , in the transition b e t w e e n reversible a n d kinetic d o m a i n s o f kinetic
d i a g r a m s [48,49], w h e r e t h e limit d e p e n d s on w h e t h e r
p e a k heights a r e c o m p a r e d , or w h e t h e r i n t e r s e c t i o n of
a s y m p t o t e s is c o n s i d e r e d . In the p r e s e n t case, c o m p a r i son of the p e a k heights is useful to e v a l u a t e t h e interm e d i a t e region, b u t it is p r e f e r a b l e to use t h e intersection of the a s y m p t o t e s to establish the a d s o r p t i o n diagrams, since the m e a s u r e m e n t s of t h e c o n s t a n t s a n d
t h e diagnosis o f t h e m e c h a n i s m s a r e b a s e d on the study
of the a s y m p t o t e s [23,24-26,47].

log[( bo bR)1/2/ cm3tool-']= log[bo / cm3 mol-']= Iog[bR/ cm3 mol"]


Fig. 9. LSV adsorption diagram for a 2e reversible reduction or
oxidation for bo = b R (Eqs. (76)-(79)); the value of z is shown on
each curve.

4.2.1. Comparison of the peak currents


F o r a totally irreversible h e t e r o g e n e o u s process, t h e
p e a k c u r r e n t for a r e d u c t i o n is given by [43,46]

ioh = 0.496(nF)3/Z( R T ) -l/eal/ZAcvD1/Zvl/2

(80)

F o r a surface r e d u c t i o n w h e n the a d s o r p t i o n obeys a


L a n g m u i r i s o t h e r m [27,47]

In t h e s e e q u a t i o n s
B = 0.51 + l o g ( D R T / n F ) - 21ogF m

(79)

T r a n s i t i o n f r o m t h e w e a k to the strong a d s o r p t i o n
r e g i o n will be o b t a i n e d for z = 1. T h e i n t e r m e d i a t e
r e g i o n c o r r e s p o n d s to c o u p l e s of values o f z such as
0.1 a n d 10, 0.05 a n d 20, etc. A d s o r p t i o n d i a g r a m s a r e
shown in Figs. 9 a n d 10 for b o = b R a n d b o = 0.1b R.
T h e t r a n s i t i o n for z = l
a n d v = 0 . 0 5 0 V s -1 is
shown in Fig. 3.

ips = e - I ( n F ) Z ( R T ) - l a A v F T

(81)

P r o c e e d i n g as in t h e reversible case, we o b t a i n
logv z = c - l o g a n - l o g ( b o / b R )
- 2 l o g ( b o b R ) t/2
= c - l o g a n - 21ogb o

(82)

for r e d u c t i o n , a n d

4.2. Totally irreversible reaction

logv z = c - log(1 - a ) n + l o g ( b o / b R )

In this case, two m e t h o d s can b e u s e d to establish


the a d s o r p t i o n d i a g r a m ; c o m p a r i s o n of t h e p e a k
heights, or i n t e r s e c t i o n of the asymptotes. T h e limits

- 2log( bob R) 1/2


= c - log(1 - a ) n - 21ogb R

(83)

for oxidation, with

6i

_~~_2

..........

c = 0.26 + 21ogz + l o g ( D R T / F ) - 21ogFm

(84)

4.2.2. Intersection of the asymptotes


F o r a totally irreversible h e t e r o g e n e o u s r e d u c t i o n ,
the c a t h o d i c p e a k p o t e n t i a l can b e e x p r e s s e d u n d e r the
form [46]

2.3RT
Epc - E -

- -

1 . 2 3 ( r r D a n F v / R T ) 1/2
log

anF

khm

(85)

F o r a surface r e d u c t i o n w h e n t h e a d s o r p t i o n obeys a
Langmuir isotherm

[47]

2
i

i
2

3
4
5
6
7
l o g [ ( b o b R )1/2 / c m 3 mol-l]( reduction or oxidation
i
3

i
4

i
5

i
6

F i g . 10. L S V a d s o r p t i o n

diagram

( 7 6 ) - ( 7 9 ) ) f o r b o = 0.1 a n d z = 1.

i
i
7
8
log bR ( oxidation
I

2.3RT
Epc - E '

anF

anFv
log--

(86)

RTk s

In view of Eq. (26), we can write

log bo ( r e d u c t i o n

for a 2e reversible reaction (Eqs.

- -

Epc - E -

2.3RT
anFv(bo)
a n F lg R-----~s

~
(87)

E. Laviron / Journal of Electroanalytical Chemistry 382 (1995) 111-127


200
H I

/ ~16""/;8

>

E
L~A

100 ~

121

G /

k~d
v

Eo,
E

2i
~

log ~.s

--"~

:--2--

*lR

I \\

o-"

log ~.

X V/ *R }tR

-2

a: ~

---~

-3
-4

k~2
-100

Fig. 11. Examples of theoretical variations of gpa and Ep c as a


function of logv for a heterogeneous reaction () [37] and for a
surface reaction (o) [38] at 20C, for n = 2, a = 0.75, b o / b R = 0.4,
when the uptake of the second electron is rate-controlling [23]. The
slopes indicated on the asymptotes are expressed in mV per logv
unit; 3. = FDV/RTk2m, As = n F v / R T k s.

Intersection of the two asymptotes defined by Eqs.


(85) and (87) (Fig. 11) gives
logv z = 0.68 + log(DRT/F) - l o g a n
+ 2log( ks/khm )
- 2alog(bo/bR)

(88)

I
6

I
7

I
I {I

log (b o bR)l/2 / crn 3 mol i

from reversible to irreversible behavior occurs w h e n


the horizontal asymptote ( E = Epc for a reversible reduction) intersects the oblique a s y m p t o t e defined by
Eqs. (85) or (86).
For a h e t e r o g e n e o u s reduction w h e n the reaction is
reversible [43], E p c = E - 1.1(RT/nF)(cf. Fig. 11).
T h e intersection with the oblique asymptote (Eq. 85) is
thus given by

DF
(89)

which can also be written


logv z = G - l o g a n + (0.5 - a ) l o g ( b o / b R )
- 21ogb o

I
5

Fig. 12. LSV adsorption diagram for a 2e totally irreversible reaction


for b o = bR, c~= 0.75 and z = 1 (Eqs. (89)-(192)). For the explanation
of the R / I R zones, see Section 4.4. P: equivalent limit for polarography.

logv z = G - l o g a n - (0.5 + a)iog(bo/bR)


2log( b o b R ) I / 2

I
4

logv = 0.963 - 21og(1.237r 1/2)

or, in view of Eqs. (18) and (19)

I
3

(90)

log-~-~ - l o g a n + 21Ogkhm

(94)

T h e R - I R d i a g r a m is shown in Fig. 14 (in order to


facilitate c o m p a r i s o n with polarography, lognv is studied instead of logv).
For a surface reduction, the intersection occurs w h e n
Eoc = E ' [11,27,32,47], whence, from Eq. (86)

RT

F o r oxidation we have

logv = log--if- - l o g a n + logk S

logv z = G - l o g ( l - a)n + (1.5 - a)log(bo/bR)


- 2 1 o g ( b o b R ) 1/2

(91)

(95)

T h e R - I R d i a g r a m is shown in Fig. 15. Since the


transition from the horizontal to the oblique asymptote

or

logv z = G - l o g ( l - a)n + (0.5 - a)log(bo/bR)


- 2logb R

(92)
3

In these equations

G = 0.68 + l o g ( D g T / F )

- 21ogF m

(93)

W e have shown in Figs. 12 and 13 examples of


adsorption diagrams (by definition, z = 1 in the p r e s e n t
case).

4.3. Reversibility / irreversibility (R-IR) diagrams


Eqs. (85) and (86) allow us to define the d o m a i n s of
reversibility/irreversibility in the plane logv/lOgkhm
or l o g v / l o g k S. W e shall consider that the transition

~lR

-4

I
t

I
4

IN

I
5

N. . . . .

L
6

~C

_ _'kll

CR

'

10

log [ ( b o bR) t'2 / cm 3 m o l - i ]

Fig. 13. LSV adsorption diagram for a 2e totally irreversible reaction


for b o = 0.1bR, o~= 0.75 and z = 1 (Eqs. (89)-(92).

E. Laviron /Journal of Electroanalytical Chemistry 382 (1995) 111-127

122

5
-~" 4
3
=

LSV /

R
"7

e.,-

0
-|

--~ -2
-3
-4
-4

-3

-2

~1

oxidation, a must be changed to 1 - a in Eqs. (94) and


(95), which yields the same limit (logv = - 0 . 7 5 ) for the
heterogeneous reaction and logv = - 0 . 2 0 for the surface reaction. We have also represented on the right of
the graph the cathodic and anodic asymptotes for Ep c
and EPa [23] (cf. Fig. 11). For the reduction, for example, when the figurative point is below C in Fig. 12, the
reaction is heterogeneous. At slow sweep rates (below
B) it is reversible; when v increases, it becomes irreversible at B. At C, a direct transition to the irreversible surface reaction occurs.
Another example is given in Fig. 13 for the same
conditions, except that b o / b R = 0 . 1 . We have now
logm =5.60; for khm = - 2 . 2 ,
the R - I R limit in the
heterogeneous region is still obtained for logv = - 0 . 7 5 ,
but the limits in the surface region are now respectively
logv = - 0 . 8 0 and - 0 . 3 2 for the reduction and the
oxidation.

l o g ( k h / c m s q)
Fig. 14. The R / I R diagram log(m,)= f(logk h) for heterogeneous
reduction at 20C. The value of a is indicated on the curves. The
limits are the same for oxidation For LSV, the difference between
the curve for a = 0.5 and that for a = 0.25 or 0.75 is only 0.06 lognV
units P: equivalent limit for polarography.

is gradual, an intermediate zone which is not shown in


Figs. 14 and 15 can be defined, based on the difference
(e.g. 2 mV) between the curve and its asymptotes (see,
for example, Ref. [50] for a heterogeneous process, and
Ref. [47] for a surface reaction).
In the case of oxidation, a should be changed to
1 - a in Eqs. (94) and (95).

5. Discussion

An LSV sweep rate Veq equivalent to the polarographic drop time can be defined [51] by the equation

Ueq

RT
nFr

(96)

We shall therefore discuss the results for polarography and cyclic voltammetry together.

4.4. Reversibility /irrecersibility in the adsorption diagrams


Domains of reversibility and irreversibility can be
defined in the adsorption diagrams (Figs. 12 and 13).
However, the R - I R limits in the weak and strong
adsorption regions are not independent; their relationship depends on the value of log(bobR) 1/2.
Let us consider, for example, a reduction for which
log(bobR )1/2 = 6, b o / b R = 1, a = 0.75 and n = 2. From
Eq. (19) (cf. Fig. 3), assuming that F m = 5 10- t0 mol
cm -2, we find logm = 5.48.
Let us now choose for 1Ogkhm the value - 2 . 2 . We
can calculate, from Eq. (94), with D = 4 10 - 6 c m 2
S-1 (cf. also Fig. 11) that the R - I R limit is obtained for
logv = - 0 . 7 5 (Fig. 12). Using Eq. (18), with logm =
5.48 and IOgkhm = --2.2, we obtain logk h = -7.68. Eq.
(19) then gives logk s = 1.1; for this value of logk s, Eq.
(86) (cf. also Fig. 15) gives for the R - I R transition
logv = 0.68. The results are shown in Fig. 12. For the

0.5

-7

3
2
t

IR

0
-t

_o

-2
-3
-4
-4

I
-3

-2

-1

log(ks / s q)
Fig. 15. R / I R diagram log(nv)= f(logk h) for a surface reduction at
20C. The value of a is indicated on the curves. For an oxidation, the
limits for a = 0.5 and 0.75 must be exchanged. P: equivalent limit for
polarography.

E. Laviron / Journal of Electroanalytical Chemistry 382 (1995) 111-127

5.1. Range of experimentally measurable values of khm


and k~
A priori, apart from solid electrodes, either a
pseudo-stationary dropping mercury electrode or a stationary (hanging drop or coated mercury electrode) can
be used in cyclic voltammetry.
The dropping electrode presents several advantages:
reproducibility, clean surface, possibility of carrying out
rapidly a large number of measurements. It also has a
distinct advantage over the stationary electrode, in that
the time at which the adsorption begins (i.e. the birth
of the drop) is well defined, which ensures that the
conditions at the surface are well known and reproducible. This allows, in particular, experiments to be
carried out using any concentration in solution, since
one does not have to wait for the equilibrium between
the surface concentration of the adsorbed species and
the analytical concentration to be established.
By contrast, stationary electrodes suffer from several
drawbacks. They are less easy to reproduce, they do
not permit many measurements to be made in a short
time, and, more particularly, the time at which the
adsorption begins is not well defined. In these conditions, one should wait until the adsorption equilibrium
between the surface concentration and the analytical
concentration is established; meanwhile, adsorption of
impurities can take place. Moreover, platinum, gold or
silver mercury-plated electrodes are unsatisfactory, in
particular because of the dissolution of the substrate in
the mercury [52-54 and references therein], whereas
mercury deposits on carbon lack homogeneity because
of their poor adherence to the surface [53]; iridiumbased electrodes seem more promising [52-54].
Although mercury electrodes are much easier to use
because of the reasons e n u m e r a t e d above, work on
well-defined solid electrodes in clean solutions is a
priori possible. However, if chemisorption occurs (cf.
Introduction), the theories developed in the present
p a p e r for weak adsorption (case II) may not be applicable. It may nevertheless be interesting to use such
electrodes to study case III.
In polarography, drop times of 0.1 to 20 s can be
envisaged. In cyclic voltammetry, the area of the electrode must be as small as possible in order to minimize
the ohmic drop effects, i.e. a capillary with a small rate
flow must be used, and the delay time t~ must be as
short as possible. The diameter of the capillary cannot
be decreased too much, because the high back-pressure which then appears must be surmounted, e.g. by
heating the mercury [55], which introduces experimental difficulties. In practice, a mercury flow of about
0.15 mg s -~ and a delay time of about 10 ms can be
used [23-26,35]. U n d e r these conditions, the voltammograms for small concentrations are free from ohmic
drop effects for l~ < 2000 V s-1 [23-26,35].

123

In view of the above considerations, we can define


the range of experimentally measurable values of khm
and k~ (Figs. 14 and 15), taking into account the fact
that part of the oblique asymptote (for fast sweep
rates) or of the horizontal asymptote (for slow sweep
rates) must be obtained; we have allowed for that
about one logv unit. We find that measurable values
are 10 -4 to 0.45 cm s -~ for khm (a comparable upper
limit has already been defined for heterogeneous constants determined by cyclic voltammetry with the same
upper limit for the scan rate [50]), and 2.5 s ~ to
5 )< 103 s - I for k~. If E or E ' values can be known,
for example by extrapolation [24-26], much smaller
rate constants can be determined (e.g. as small as
about 10 -6 cm s -1 for a heterogeneous process; see
Refs. [24,25]).
According to Eq. (15), surface rate constants should
be of the order of 10~-109 s -1 since k h values are
usually in the range 0.1 to 10 cm s - 1 [24-26], which has
been confirmed experimentally by using indirect methods [21,24-26,35].
On the other hand, even for weak adsorption, khm
values are in the 104-106 cm s - l range (cf. Eq. 17). As
can be deduced from Figs. 14 and 15, such values are
quite out of reach of LSV; it is even doubtful whether
they can ever be measured by direct electrochemical
methods. However, in aqueous medium, the reduction
of organic compounds is often made up of a succession
of electron and proton uptakes, which can be described
by using square schemes or their derivatives (fence,
ladder, cubic schemes, etc.) [57-60]. When the protonations are fast, i.e. can be considered to be at equilibrium, which has indeed been shown to hold when
nitrogen or oxygen atoms are involved [21,24-26,61],
the theory shows that the global process is equivalent
to a simple reaction, with an apparent electrochemical
rate constant which becomes much smaller than the
constants k~ or khm for the monoelectronic elementary
reactions, so that its value falls in the measurable
range. We have investigated diverse reactions using
these properties [21,24-26,35].
When b o 4: bR, it must be noticed that the results
for a reversible reduction in polarography (Fig. 7) are
the reverse of those for an irreversible reaction (Eq.
(66)), or for LSV (Eqs. (66), (68), (75), (77), (82), (83),
(89), (91)); for bo/b R = 0.1, for example, in the case of
a reduction, the limit between the zones of weak and
strong adsorption is shifted towards larger log~" (or
smaller log(bobR) 1/2) values (Fig. 7), whereas the reverse is true in the other cases. This is a result of the
different nature of the process. For a reversible reduction in polarography, a decrease of b o, i.e. an increase
of b R for a given value of bob R, results in an increase
of the adsorption of R; less molecules leave the electrode. The adsorption zone expands. For a totally
irreversible reduction in polarography, in contrast, a

E. Laviron/Journal of Electroanalytical Chemistry 382 (1995) 111-127

124

= 0.01, we o b t a i n 1.08, 0.61 a n d 0.34. F o r o x i d a t i o n ,


t h e e x p o n e n t is 0 . 2 5 - a / 2 a n d t h e r e s u l t s a r e rev e r s e d . T h e ratio o f t h e p e a k s is n o t very d i f f e r e n t
f r o m u n i t y in every case.

4
3

oo

>

ez0

l
0

S~

-I
-2

-3
-4

6
9 7
I

log[( b 0 b R ) 1 2 / c m 3 mol.i]
Fig. 16. Limits of the adsorption diagram (x = 0.5 or z = 1) obtained
from the diverse methods for b o / b n = l , D = 4 1 0 -6 cm 2 s - l ,
/ ' m = 5 1 0 - l tool cm 2: curve (1), Eq. (82), Red, irreversible,
n = 2, a = 0.75, and Eq. (83), Ox, irreversible, n = 2, a = 0.25; curve
(2), Eqs. (76) and (78), Red, Ox, reversible, n = 2; curve (3), Eq. (89),
Red, irreversible, n = 2, a = 0.75, and Eq. (91), Ox, irreversible,
n = 2, a = 0.25, and Eqs. (76) and (78), Red, Ox, n = 1; curve (4), Eq.
(82), Red, irreversible, n = 2, a = 0.25, and Eq. (83), Ox, irreversible,
n = 2, a = 0.75, and Eqs. (82) and (83), Red, Ox, n = 1, a = 0.5; curve
(5), Eq. (89), Red, irreversible, n = 2, a = 0.25, and Eq. (91), Ox,
irreversible, n = 2, a = 0.75, and Eqs. (89) and (91), Red, Ox, irreversible, n = 1, a = 0.5; curve (6), Eq. (59), Red, Ox, reversible, n = 1;
curve (7), Eq. (59), Red, Ox, reversible, n = 2; curve (8), Eqs. (66)
and (68), Red, Ox, irreversible, n = 1, a = any value; curve (9), Eqs.
(66) and (68), Red, Ox, irreversible, n = 2, a = any value.

d e c r e a s e of b o d i m i n i s h e s t h e s u r f a c e n a t u r e of t h e
process, since R n o l o n g e r plays a role in t h e r e a c t i o n .
I n L S V ( r e v e r s i b l e o r i r r e v e r s i b l e case) a d e c r e a s e of
b o also d i m i n i s h e s t h e s u r f a c e n a t u r e of t h e r e d u c t i o n ,
since t h e q u a n t i t y of m o l e c u l e s a d s o r b e d b e f o r e t h e
start of t h e s w e e p d e c r e a s e s .

5.3. The width of the transition region


A c c o r d i n g to t h e t h e o r e t i c a l results, t h e i n t e r m e d i ate r e g i o n b e t w e e n t h e p u r e l y h e t e r o g e n e o u s a n d
p u r e l y s u r f a c e r e a c t i o n s e x t e n d s o v e r a large r a n g e o f
l o g z o r l o g v (cf. Figs. 6 a n d 9). I n practice, h o w e v e r ,
t h e t r a n s i t i o n r e g i o n c a n b e m u c h n a r r o w e r . L e t us, for
e x a m p l e , c o n s i d e r Fig. 11, in w h i c h t h e d i f f e r e n c e
b e t w e e n k s a n d khm is large e n o u g h t h a t t h e c u r v e s
i n t e r s e c t in t h e r e g i o n of t h e o b l i q u e a s y m p t o t e s . A t
t h e i n t e r s e c t i o n , b y d e f i n i t i o n , t h e p e a k p o t e n t i a l s for
the surface and heterogeneous redox processes are
e q u a l . W h e n v is e i t h e r i n c r e a s e d or d e c r e a s e d , t h e
p e a k p o t e n t i a l is d e t e r m i n e d by o n e of t h e two reactions, a n d t h e v a r i a t i o n s closely follow t h e a s y m p t o t e s ,
as is c o n f i r m e d e x p e r i m e n t a l l y [25,26,35]. T h e r e is n o
difficulty in c h a r a c t e r i z i n g t h e a s y m p t o t e s , w h o s e i n t e r s e c t i o n s with E o r E ' give khm a n d k s respectively
( 2 1 , 2 4 - 2 6 ) (in t h e p r e s e n t case, t h e c o n s t a n t s a r e relative to t h e s e c o n d e l e c t r o n u p t a k e ; cf. Fig. 11 a n d Ref.
[23]). If t h e d i f f e r e n c e b e t w e e n k s a n d khm is too
small, t h e s i t u a t i o n b e c o m e s m o r e c o m p l e x (Fig. 17),
because the intersection of the curves can occur out of
t h e a s y m p t o t i c r e g i o n . I n Fig. 17, for e x a m p l e , t h e
a n o d i c a s y m p t o t e for t h e h e t e r o g e n e o u s p r o c e s s (slope,
58 m V ) c a n n o t b e d e t e r m i n e d e x p e r i m e n t a l l y , since t h e
c u r v e s for t h e h e t e r o g e n e o u s a n d s u r f a c e p r o c e s s coincide in this r e g i o n ; e v e n t h e c a t h o d i c a s y m p t o t e (slope,
- 3 8 m V ) will b e difficult to c h a r a c t e r i z e . A m o r e
c a r e f u l study, u s i n g t h e w h o l e curve, m u s t b e c a r r i e d
o u t [35,61,62].

/6

200

5.2. Coherence of the results


>

A t first sight, t h e diverse m e t h o d s u s e d for d e t e r m i n i n g the h e t e r o g e n e o u s a n d s u r f a c e r e a c t i o n z o n e s


( s o l u t i o n of t h e d i f f u s i o n e q u a t i o n s , i n t e r s e c t i o n of t h e
a s y m p t o t e s , c o m p a r i s o n o f t h e p e a k h e i g h t s ) s e e m to
yield r a t h e r d i f f e r e n t results. H o w e v e r , for b o = b R,
t h e limits a r e n o t very d i f f e r e n t (Fig. 16). E v e n w h e n
b o 4 : b R, t h e d i f f e r e n c e s r e m a i n small. W e can, for
e x a m p l e , c a l c u l a t e for a r e d u c t i o n t h e ratio ips/ioh
w h e n t h e a s y m p t o t e s i n t e r s e c t by s u b s t i t u t i n g t h e v a l u e
o f v d e r i v e d f r o m Eq. (89) i n t o Eqs. (80) a n d (81),
t a k i n g in a c c o u n t Eq. (74). W e o b t a i n

ips/iph = 0.61( b o / b R ) '~/2-25

(97)

W h e n b o / b R = 0.1, this r a t i o is t h u s 0.81, 0.61 a n d


0.46 for a = 0.25, 0.5 a n d 0.75 respectively; for b o / b R

#/
11158

/ 1

II i # . 4 ##
##,C/

HiS

100

Z
v

/ I

Eo,
>
E

o~

2
I

3I log ks
I

log k

-39~"-,~

-19 ~

k~d

K
~

-I00

Fig. 17. Example of theoretical variations of EPa and Epc as a


function of logv for a heterogeneous reaction () [37] and for a
surface reaction (o) [38] at 20C, for n = 2, a = 0.75, bo/b n = 0.2,
when the uptake of the second electron is rate-controlling [23]. The
slopes indicated on the asymptotes are expressed in mV per logv
unit; h = FDV/RTk~m, As = nFv/RTk~.

E. Laviron/ Journal of Electroanalytical Chemistry 382 (1995) 111-127


5.4. Definition of an average time of stay of the molecules
on the surface
L e t u s c o n s i d e r Eq. (55); w h e n x = 0.5, i.e. w h e n
h a l f o f t h e m o l e c u l e s o f R leave t h e e l e c t r o d e , a n d h a l f
r e m a i n a d s o r b e d , a o = 0.43. L e t us d e s i g n a t e by t~ t h e
t i m e d e f i n e d by

D
I
I|
It

--A

1324
I

----

0.43bR FmD - 1/2

>~l I
Ill

go

I I

L iI

(98)

or

-3

tseq

nFUseq

0.30b2 F2mD-1

(100)

F o r a n i r r e v e r s i b l e r e d u c t i o n in p o l a r o g r a p h y (Eqs.
(62), (64) a n d (19))
ts

0"55r 2 a - 1/~2~'or"2I r ar~,_.

ii I
I II
| iI

~--

~--'

,,

(99)

W e c a n c o n s i d e r t~ as a n a v e r a g e t i m e of stay of t h e
m o l e c u l e s o n t h e e l e c t r o d e . If t >> t~, t h e r e is e n o u g h
t i m e for t h e m o l e c u l e s to d e s o r b , so t h a t t h e r e a c t i o n
a p p e a r s as h o m o g e n e o u s (Fig. 2, case II; Figs. 6 a n d 7).
W h e n t << t~, in c o n t r a s t , t h e m o l e c u l e s do n o t h a v e
e n o u g h t i m e to d e s o r b , a n d t h e r e a c t i o n t a k e s p l a c e o n
t h e surface.
T h e s a m e r e l a t i o n s h i p , w i t h a slightly d i f f e r e n t n u m e r i c a l c o e f f i c i e n t , is o b t a i n e d in t h e o t h e r cases w h i c h
we h a v e t r e a t e d .
F o r a r e v e r s i b l e r e a c t i o n in L S V we get, for e x a m ple, f r o m Eqs. ( 7 0 ) - ( 7 2 ) a n d (74)

RT

II I
II I
-2

t~ = O.lSb2RF2D - t

1I

101/ |l

tsl/2

125

I
2

I
3

I J
4

I
5

14
[
7

13
J
8

I
I
LL
9

I
I0

Iog(bo ' cm ~ mol 1)( reduction ) or Iog(b R 'cm ~ moll)( oxidation )

Fig. 18.

log(bo/b R) as a function

of logb o (for reduction) or logb R

(for oxidation): lines (1)-(6), LSV, irreversible reaction, t, = 2000 V


s 1, z = 1 (line (1), Red, a = 0.75, n = 2; line (2) Ox, a = 0.25, n = 2;
line (3), Ox, a = 0.75, n = 2; line (4), Red, a = 0.25, n = 2; line (5),
Ox, a = 0 . 5 , n = l and Red, a = 0 . 5 , n = l ) ; lines (6)-(9), LSV,
irreversible reaction, c = 0.050 V s- l, z = 1 (line (6), Red, a = 0.75,
n = 2; line (7), Ox, a = 0.25, n = 2; line (8), Red and Ox, a = 0.75,
n = 2; line (9), Red, a = 0.25, n = 2); lines (10)-(12) LSV, reversible
reaction, n = 2, e = 2000 V s- 1, Red or Ox (line (10), z = 0.05; line
(11), z = 1; line (12), z = 20); lines (13)-(15) polarography, irreversible reaction, n = 1 or 2, Red or Ox, x = 1, z = 2.0 s (Eq, (61))
(line (13), a = 0.75; line (14), a = 0.25; line (15), a = 0.5). In region
A, the reaction appears as heterogeneous whatever the sweep rate.
In region B, the reaction always has a surface character, whatever c
or r. In regions C and E, the reaction has a surface or heterogeneous
character according to the sweep rate and the nature of reaction
(reduction, oxidation, reversibility). In D, it can appear as a heterogeneous or surface reaction, whatever the type of reaction.

(101)

F o r a n i r r e v e r s i b l e r e d u c t i o n in LSV, w h e n t h e p e a k
h e i g h t s a r e c o m p a r e d (Eqs. (80), (81) a n d (74))

tseq

RT
= 0.55ab2F2mD-1
nFv seq

(102)

I n t h e s a m e case, f r o m t h e i n t e r s e c t i o n of t h e
a s y m p t o t e s (Eqs. (85), (86) a n d (19))

ts~q

RT
- -

= 0.21r2~-lb2 F 2 D - 1

(103)

nFVseq
T h e t i m e d e f i n e d b y t h e s e e q u a t i o n s is o f t h e s a m e
o r d e r of m a g n i t u d e if r is n o t too d i f f e r e n t f r o m u n i t y .
T h e t i m e t s t h u s d e f i n e d , p a r t i c u l a r l y in t h e case o f
r e v e r s i b l e r e a c t i o n s , will b e u s e f u l for t h e d i s c u s s i o n of
chemical reactions accompanying the electron transfers
[63].

5.5. Experimental accessibility of the surface/heterogeneous domains


A s s h o w n by t h e a d s o r p t i o n d i a g r a m s (Figs. 6, 7, 9,
10, 12, 13 a n d 16), t h e n a t u r e of t h e r e a c t i o n ( h e t e r o g e n e o u s o r s u r f a c e ) is d e t e r m i n e d b y b o t h t h e v a l u e o f v
o r r a n d t h a t of t h e a d s o r p t i o n c o e f f i c i e n t s . If v < 2000

V s - 1 , for e x a m p l e , t h e r e a c t i o n will always a p p e a r as


heterogeneous
in t h e c o n d i t i o n s o f Fig. 12 if
log(bobR) 1/2 < 4.5, etc. T h e p r o b l e m c a n b e s t u d i e d by
c o n s i d e r i n g Eq. (69), in w h i c h z is e q u a l to its maxim u m v a l u e , 20 s, o r Eqs. (75), (77), (90) a n d (92), in
w h i c h v t a k e s e i t h e r its u p p e r o r its lower v a l u e s , i.e..
2000 V s -1 a n d 0.050 V s -1. G r a p h s giving log(bo/b R)
as a f u n c t i o n o f e i t h e r l o g b o (for r e d u c t i o n ) or l o g b R
(for o x i d a t i o n ) c a n t h u s b e o b t a i n e d (Fig. 18). T h e solid
l i n e s c o r r e s p o n d to x = 0.5 ( p o l a r o g r a p h y ) o r z = 1
(LSV), t h e d a s h e d lines o n t h e left to z = 0.05 ( 5 % of
s u r f a c e r e a c t i o n for lines 1 a n d 2), t h o s e o n t h e right to
x = 0.95 ( 9 5 % o f s u r f a c e r e a c t i o n ) for l i n e s 13 a n d 14.
A s c a n b e seen, t h e r e a c t i o n will always a p p e a r as
h e t e r o g e n e o u s w h e n l o g b o (for r e d u c t i o n ) or l o g b R
(for o x i d a t i o n ) is s m a l l e r t h a n 2.75, w h e r e a s it will
always b e o f a s u r f a c e n a t u r e for l o g b o o r l o g b R l a r g e r
t h a n 9. B e t w e e n t h e s e values, b o t h types of r e a c t i o n s
c a n a p p e a r , w h e n v (or z) varies. T h e limits i n d i c a t e d
a r e valid w i t h i n t h e r a n g e of v a l u e s of c a n d b o / b R
d e f i n e d . D a t a for o t h e r c o n d i t i o n s (e.g. if l a r g e r v a l u e s
of u are u s e d o n a n i r i d i u m - b a s e d m e r c u r y e l e c t r o d e ;
cf. S e c t i o n 5.1) c a n easily b e c a l c u l a t e d .

126

E. Laviron /Journal of Electroanalytical Chemistry 382 (1995) 111-127

6. Conclusion
The rigorous mathematical theory established for
the rotating disk electrode [15,16] constitutes the necessary basis for understanding quantitatively the role of
adsorption of the reactants during the redox processes.
However, non-steady-state methods are much more
advantageous to use for quantitative experimental studies, for several reasons. The first are of a theoretical
nature; non-steady-state methods allow both the heterogeneous and the surface reactions to be studied,
whereas only the heterogeneous one can be examined
using steady-state methods such as r.d.e, voltammetry;
also, if a mercury electrode is used, adsorption is
generally of a physical nature, so that the theories
developed in this paper are directly applicable. Another reason, which is practical, is linked to the advantages of using a dropping mercury electrode: good
definition of the time at which the experiments start,
reproducibility, cleanliness, possibility of performing
many experiments in a short time.
Application of the methods described in the present
paper has allowed us to establish the detailed mechanism of the reduction of diverse compounds [21,2426,35,62] in aqueous medium, and in particular to
confirm the order of magnitude of the rate constants
k~ and khm predicted by theory. Generally speaking,
our studies also show that adsorption phenomena,
which had hitherto been considered as an obstacle to
the study of mechanism in aqueous media, can on the
contrary be quantitatively incorporated into the global
analysis of the redox processes. We are currently studying other types of systems on the basis of the results
developed in the present paper, and we are planning to
examine non-aqueous media.
Lastly we think that our results can contribute to a
better understanding of the influence of the adsorption
of the reactants on solid electrodes in certain cases, if,
for example, slow chemisorption does not complicate
the reaction.

Acknowledgements
We would like to express our thanks to Mrs.
Raveau-Fouquet and Mrs. Tilleul for their help in
preparing the manuscript.

References
[1] M. Gouy, Ann. Chim. et Phys., 8 (1906) 291; Ann. Chim. et
Phys., 9 (1906) 75.
[2] B.B. Damaskin, O.A. Petrii and V.V. Batrakov, Adsorption of
Organic Compounds on Electrodes, Plenum Press, New York,
1971.

[3] U. Gaunitz and W. Lorenz, Collect. Cz. Chem. Commun., 36


(1971) 796.
[4] V.D. Bezuglyi, L.A. Korshikov and V.B. Titova, Elektrokhimyia,
6 (1970) 1150.
[5] U. Palm and A. Alumaa, J. Electroanal. Chem., 90 (1978) 219.
[6] R.I. Kaganovich, B.B. Damaskin and M.K. Kaisheva, Elektrokhimyia, 6 (1970) 1359.
[7] R.V. Ivanova, L.N. Kuznetsova and B.B. Damaskin, Elektrokhimyia, 13 (1977) 1881.
[8] E. Laviron, J. Electroanal. Chem, 169 (1984) 23,29.
[9] E. Laviron, J. Electroanal. Chem., 208 (1986) 357.
[10] E. Laviron and L. Roullier, J. Electroanal. Chem., 186 (1985) 1.
[lll E. Laviron, in A.J. Bard (ed.), Electroanalytical Chemistry,
Marcel Dekker, New York, 1982, 19. 53.
[12] R.H. Wopschall and 1. Shain, Anal. Chem., 39 (1967) 1514.
[13] S. Srivanasan and E. Gileadi, Electrochim. Acta, 11 (1966) 321.
[14] H. Angerstein-Kozlowska, J. Kiinger and B.E. Conway, J. Electroanal. Chem., 75 (1977) 45 and 61.
[15] E. Laviron, J. Electroanal. Chem., 124 (1981) 19.
[16] E. Laviron, J. Electroanal. Chem., 140 (1982) 247.
[17] A.N. Frumkin and B.B. Damaskin, in J.M. Bockris and B.E.
Conway (eds.), Modern Aspects of Electrochemistry, Butterworth, London, 1964, 19. 149.
[18] P. Delahay, Double Layer and Electrode Kinetics, Interscience,
New York, 1965, (a) p. 117, (b) 19. 243, (c) 19. 240.
[19] J.O'M. Bockris and K.T. Jeng, J. Electroanal. Chem., 330 (1992)
541.
[20] A.T. Hubbard, Chem. Rev., 88 (1988) 633.
[21] E. Laviron and L. Roullier, J. Electroanal. Chem., 157 (1983) 7.
[22] E. Laviron, J. Electroanal. Chem., 164 (1984) 213.
[23] E. Laviron and L. Roullier, J. Electroanal. Chem., 288 (1990)
165.
[24] E. Laviron, R. Meunier-Prest, A. Vallat, L. Roullier and R.
Lacasse, J. Electroanal. Chem., 341 (1992) 227.
[25] R. Lacasse, R. Meunier-Prest and E. Laviron, J. Electroanal.
Chem, 359 (1993) 223.
[26] E. Laviron, R. Meunier-Prest and R. Lacasse, J. Electroanal.
Chem., 375 (1994) 263.
[27] E. Laviron, J. Electroanal. Chem., 52 (1974) 355.
[28] A.P.Brown and F.C. Anson, J. Electroanal. Chem., 92 (1978)
133.
[29] D.M. Mohilner, J. Phys. Chem., 73 (1969) 2652.
[30] R. Brdicka, Z. Elektrochem., 48 (1942) 278.
[31] E. Laviron, J. Electroanal. Chem., 97 (1979) 135.
[32] E. Laviron, Bull. Soc. Chim. Fr., (1967) 3717.
[33] E. Laviron, J. Electroanal. Chem., 63 (1975) 245.
[34] I.M. Kolthoff and J.J. Lingane, Polarogra19hy, Interscience, New
York, 1952, 19. 256.
[35] E. Laviron, A. Vallat and R. Meunier-Prest, J. Electroanal.
Chem., 379 (1994) 427.
[36] R. Meunier-Prest, Ch. Gaspard-G6hin and E. Laviron, unpublished results.
[37] J. Koryta, Collect. Cz. Chem. Commun., 18 (1953) 206.
[38] E. Laviron and R. Meunier-Prest, J. Electroanal. Chem., 375
(1994) 79.
[39] P. Delahay and I. Trachtenberg, J. Amer. Chem. Soc., 79 (1957)
2355.
[40] E. Laviron, J. Electroanal. Chem., 52 (1974) 395.
[41] E. Laviron and L. Roullier, J. Electroanal. Chem., 115 (1980)
65.
[42] H.S. Carlslaw and J.C. Jaeger, Conduction of Heat in Solids,
Clarendon Press, Oxford, 1959.
[43] P. Delahay, New Instrumental Methods in Electrochemistry,
lnterscience, New York, 1954, Oxford University Press, 1959, p.
55.
[44] M. Abramowitz and I.A. Stegun (eds.), Handbook of Mathematical Functions, Dover, New York, 1970, p. 310.

E. Laviron /Journal of Electroanalytical Chemistry 382 (1995) 111-127


[45] J. Koutecky, Collect. Cz. Chem. Commun., 18 (1953) 597.
[46] A.J. Bard and L.R. Faulkner, Electrochemical Methods, Wiley,
New York, 1980.
[47] E. Laviron,, J. Electroanal. Chem., 101 (1979) 19.
[48] M. Mastragostino, L. Nadjo and J.M. Sav6ant, Electrochim.
Acta, 13 (1968) 721.
[49] L. Nadjo and J.M. Sav6ant, J. Electroanal. Chem., 48 (1973)
113.
[50] C.P. Andrieux and J.M. Sav6ant, in C. Bernasconi (ed.), Techniques of Chemistry, Vol. 3, Wiley, New York, 1986, Part II,
Chapter VII.
[51] Z. Galus, Fundamentals of Electrochemical Analysis, Ellis Horwood, Chichester, 1976.
[52] S.P. Konnaves and W. Deng, J. Electroanal. Chem., 301 (1991)
77.
[53] S.P. Konnaves and W. Deng, Anal. Chem., 65 (1993) 375.
[54] S.P. Konnaves, W. Deng, P.R. Hallock, G.T.A. Kovacs and
C.W. Storment, Anal. Chem., 66 (1994) 418.

127

[55] J.W. Pons, J. Daschbach, S. Pons and M. Fleischman, J. Electroanal. Chem., 239 (1988) 427.
[56] A.J. Bard and H. Lund (eds.), Encyclopedia of Electrochemistry
of the Elements, New York, 1979.
[57] E. Laviron, J. Electroanal. Chem., 109 (1980) 57; J. Electroanal.
Chem., 124 (1981) 9; J. Electroanal. Chem., 146 (1983) 1; J.
Electroanal. Chem., 146 (1983) 15.
[58] E. Laviron and R. Meunier-Prest, J. Electroanal. Chem., 324
(1992) 1.
[59] R. Meunier-Prest and E. Laviron, J. Electroanal. Chem., 328
(1992) 33.
[60] E. Laviron, J. Electroanal. Chem., 365 (11994) 1.
[61] E. Laviron, L. Roullier and J.L. Decker, unpublished results.
[62] E. Laviron, R. Meunier-Prest and E. Mathieu, J. Electroanal.
Chem., 371 (1994) 251.
[63] E. Laviron, in preparation.

Potrebbero piacerti anche