You are on page 1of 9

Journal of Electroanalytical Chemistry 505 (2001) 100108

www.elsevier.nl/locate/jelechem
Adsorptive square wave voltammetry of metal complexes.
Effect of ligand concentration
Part I. Theory
Fernando Garay *
INFIQC, Departamento de F sico Qu mica, Facultad de Ciencias Qu micas, Uni6ersidad Nacional de Cordoba, Pabellon Argentina, Ala 1, 2
piso, Ciudad Uni6ersitaria, 5000 Cordoba, Argentina
Received 8 December 2000; received in revised form 19 February 2001; accepted 1 March 2001
Abstract
The electrochemical behaviour of non-labile metallic complexes under square wave voltammetry (SWV) conditions is analysed
theoretically, considering the inuence of ligand adsorptiondesorption processes as well as the ligand concentration on the
quasi-reversible redox reaction mechanism. The dependence of current and peak potentials on the transfer reaction rate and on
complex stoichiometry is considered. Voltammetric responses for processes in which the ligand is desorbed or remains adsorbed
after the electroreduction process are compared. Diagnosis criteria for the selection of optimum SWV parameters in view of
analytical applications are presented for each case. 2001 Elsevier Science B.V. All rights reserved.
Keywords: Adsorptive square wave voltammetry; Mechanistic studies; Mathematical models
1. Introduction
The adsorption of electroactive analytes on the elec-
trode surface is used commonly as a preconcentration
step in electroanalytic procedures. This step improves
the sensitivity of the analytical technique, allowing
trace level determinations [16]. Square wave voltam-
metry (SWV) is one of the electrochemical techniques
more widely applied in quantitative analysis, due to its
high sensitivity, which is mainly the consequence of the
rejection of most of the capacitive currents [7]. The
alternating application of oxidation and reduction
pulses, characteristic of SWV, allows the concentration
gradients of reagents and products close to the surface
to be rebuilt in each wave. This fact allows the determi-
nation of both species almost simultaneously. The dif-
ferential peak current (Dc
p
) and the peak potential (E
p
)
are the most useful parameters of SWV from the ana-
lytical point of view. Nevertheless, Dc
p
values also
have been employed in some kinetic analyses, since a
plot of Dc
p
versus the charge transfer rate constant (k
s
)
depicts a maximum for quasi-reversible redox reactions
with adsorbed species; this quasi-reversible maximum
can be employed to obtain k
s
[810].
One of the modeling advantages is the optimisation
of the experimental conditions for quantitative determi-
nations of metallic complexes at trace levels. Many
studies in the literature take into account adsorbed
species involved in charge transfers of different rates
and examine also the effect of irreversible chemical
reactions coupled to the electrochemical step [922].
Models that simulate the voltammetric response for the
adsorptive accumulation of metallic cations have also
been developed. In all cases, the strong and labile
complexation forms of the metallic cations by organic
and inorganic ligands have been considered, respec-
tively, under conditions where the ligand is in great
excess with regard to the metal concentration [1216].
Simulation of mechanisms involving adsorbed species
requires the relationship between bulk concentrations
of soluble species and their surface excesses at a given
temperature. One of the simplest models considers a
linear relationship between bulk and surface concentra-
tions. This type of isotherm is applicable at low surface
coverages, which are found frequently in trace analysis
* Tel.: +54-351-4334169; fax: +54-351-4334188.
E-mail address: fgaray@squim.fcq.unc.edu.ar (F. Garay).
0022-0728/01/$ - see front matter 2001 Elsevier Science B.V. All rights reserved.
PII: S0022- 0728( 01) 00459- 4
F. Garay / Journal of Electroanalytical Chemistry 505 (2001) 100108 101
[12]. Nevertheless, a great excess of organic ligand is
usually added to the analytical solution in the adsorp-
tive quantication of metallic cations. This excess does
not necessarily increase the sensitivity of the response,
but it invalidates the applicability of the linear
isotherm.
In the present work, a theoretical model for a reac-
tion mechanism considering diffusion effects and the
adsorption of reagents is proposed. The model assumes
inert complexes with stability constants high enough to
consider that no free metal will be detected after the
formation of the complex. The concentration and ad-
sorption effects of the ligand on the voltammometric
response are analysed. This concentration was always
kept above the assumed trace amounts of metal. The
ligand excess range with regard to the total metallic ion
concentration was studied from 10
5
to 10
4
. In this
way, the evaluation of experimental cases under linear
adsorption conditions can be accomplished.
2. Mathematical model
The formation in the bulk solution of only one
chemically stable complex species is assumed:
M
+
(sol)
+uL

(sol)
Yk
K
st,o
ML
u(sol)
(1)
Here, the u value can be 1 or 2, depending on the
complex stoichiometry. For simplicity, the charge on
the oxidised complex is omitted, all M
n+
are symbol-
ised by M
+
and the sufx u in the stability (K
st,o
) and
adsorption (K
ad
) constants are also omitted; c* stands
for bulk concentrations. It is considered that c
L
* .c
M
+
*
and the distribution of these ionic species is dened
according to:
c
M
+
(ini)
* =c
ML
u
(eq)
* (2)
c
L (eq)
* =c
L (ini)
* uc
M
+
(ini)
* (3)
where the subscripts (ini) and (eq) indicate the species
concentration before the ligand is added and after the
complexing equilibrium is reached, respectively. The
ligandcomplex concentration ratio is dened as:
R
LM
=c
L
*(c
ML
u
* )
1
(4)
Considering the above conditions, two reaction
schemes were analysed. The rst one assumes that the
product of the adsorbed complex reduction is the lig-
and in solution, Eq. (5), while in the second scheme the
adsorbed ligand remains at equilibrium with its soluble
fraction, Eq. (6).
ML
u(sol)
Yk
K
ad,ML
ML
u(ad)
Yk
k
s
ne

M(Hg)
(sol)
+uL
(sol)
(5)
ML
u(sol)
Yk
K
ad,ML
ML
u(ad)
Yk
k
s
ne

M(Hg)
(sol)
+uL
(ad)
Yk
K
ad,L
uL
(sol)
(6)
where K
ad,ML
and K
ad,L
are the adsorption constants for
the oxidised metal complex and the free ligand, respec-
tively. As the adsorption constants are dened for the
forward reactions of Eqs. (5) and (6) their dimensions
are in centimetres. A complete list of symbols is sum-
marised in the nomenclature.
As stated before, adsorption isotherms for the metal-
lic complex and for the free ligand are assumed to be
linear. Also, no redox reactions involving the ligand are
supposed to occur within the working potential range.
Considering one-dimensional diffusion, Eqs. (5) and (6)
can be evaluated mathematically with the following
differential equations:
#c
ML
u
/#t =D(#
2
c
ML
u
/#x
2
) (7)
#c
M(Hg)
/#t =D(#
2
c
M(Hg)
/#x
2
) (8)
#c
L
/#t =D(#
2
c
L
/#x
2
) (9)
For simplicity, a common value of the diffusion coef-
cient, D=110
5
cm
2
s
1
was assumed for all diffus-
ing species. The following initial and boundary
conditions are considered.
t =0, x0:
c
ML
u
=c
ML
u
* ; Y
ML
u
ini
=K
ad,ML
c
ML
u
* (10)
c
M(Hg)
=0; c
L
=c
L
* (11)
Y
L
ini
=K
ad,L
c
L
* (12)
t .0, x:
c
ML
u
c
ML
u
* ; c
L
c
L
*; c
M(Hg)
0 (13)
x=0:
Y
ML
u
=K
ad,ML
c
ML
u
(14)
D(#c
ML
u
/#x)
x=0
=#Y
ML
u
/#t +I/nFA (15)
D(#c
M(Hg)
/#x)
x=0
=I/nFA (16)
D(#c
L
/#x)
x=0
=I/nFA (17)
Y
L
=K
ad,L
c
L
(18)
D(#c
L
/#x)
x=0
=#Y
L
/#t I/nFA (19)
Boundary conditions (12), (18) and (19) are applied
only to Eq. (6) whereas condition (17) is operative only
for Eq. (5). All the other conditions apply to both
reaction schemes. Provided that the rate of the electro-
chemical step is considerably faster than the complex
dissociation, the reduced complex species could be as-
sumed to be an intermediate species in equilibrium with
the dissociated products:
K
st, r
=Y
ML
u
(c
M(Hg)
c
L
u
)
1
(20)
F. Garay / Journal of Electroanalytical Chemistry 505 (2001) 100108 102
where K
st,r
describes the chemical equilibrium between
the adsorbed reduced complex and its dissociated spe-
cies, M (in the amalgam) and L (in solution, close to
the electrode surface). The rate of the charge transfer
reaction is given by the well-known equation:
I(t)/nFA=k
s
exp[ h(t)]{Y
ML
u
Y
ML
u
exp[(t)]}
(21)
where the symbols have their usual meaning and (t) is
a function of the reaction scheme considered, according
to the following expressions [11,13]:
(t)
Eq. (5)
=nF[E(t) E]/RT+ln(K
ad,ML
)
+ln(K
st,o
/K
st,r
) (22)
(t)
Eq. (6)
=nF[E(t) E]/RT+ln(K
ad,ML
/K
ad,L
)
+ln(K
st,o
/K
st, r
) (23)
where E(t) is the square wave (SW) potential function
and E is the standard potential for the simple redox
reaction involving soluble free metal species.
In order to simulate the reactions described in Eqs.
(5) and (6), the current is normalised according to
c(t) =I(t)/(nFAfY
ML
u
ini
) where f is the SW frequency. A
numerical integration method is employed in the resolu-
tion of differential equations (7), (8) and (9) under the
relevant boundary conditions, Eqs. (10)(19) [23]. The
following results were obtained.
Reaction scheme of Eq. (5), assuming u=1:
0=c
(m)
2
+c
(m)
{Z
a
[T
(m)

(m)
+1] +2]
a(m)
}
+Z
a
T
(m)
[]
b(m)
f
1
] +]
a(m)
[Z
a
+]
a(m)
] (24)
Reaction scheme of Eq. (5), assuming u=2:
0=c
(m)
3
+c
(m)
2
{Z
a
+3]
a(m)
}
+c
(m)
{(Z
a
/2)
2
[T
(m)

(m)
+1]
+]
a(m)
(2Z
a
+3]
a(m)
)}
+(Z
a
/2)
2
[T
(m)
(]
b(m)
f
1
) +]
a(m)
]
+(]
a(m)
)
2
[Z
a
+]
a(m)
] (25)
Reaction scheme of Eq. (6), assuming u=1:
0=c
(m)
2
+c
(m)
{Z
b
[T
(m)

(m)
+1] +]
a(m)
+]
c(m)
}
+Z
b
T
(m)
[]
b(m)
f
1
] +]
a(m)
[Z
b
+]
c(m)
] (26)
Reaction scheme of Eq. (6), assuming u=2:
0=c
(m)
3
+c
(m)
2
{Z
b
+2]
c(m)
+]
a(m)
}
+c
(m)
{(Z
b
/2)
2
[T
(m)

(m)
+1] +]
c(m)
(Z
b
+]
c(m)
)
+]
a(m)
(Z
b
+2]
c(m)
)}
+(Z
b
/2)
2
[T
(m)
(]
b(m)
f
1
) +]
a(m)
]
+]
c(m)
]
a(m)
[Z
b
+]
c(m)
] (27)
where ]
a(m)=

m1
j =1
c
( j )
S
(i )
; ]
b(m)
=
m1
j =1
c
( j )
Q
(i )
;
]
c(m)
P
(1)
=
m1
j =1
c
( j )
P
(i )
;
(m)
=exp[h
(m)
]k
s
1
+Q
(1)
;
S
(i )
=(i )
1/2
(i 1)
1/2
; i =mj +1; Q
(i )
={uS
(i )
+
Y
1(i )
}a
1
1
; P
(i )
={uS
(i )
+Y
2(i )
}a
2
1
; Y
y(i )
=
{exp[a
y
2
(li )] erfc[a
y
(li )
1/2
] exp[a
y
2
(l(i 1))] erfc[a
y
(l-
(i 1))
1/2
]}a
y
1
. In this latter equation, y subscripts can
be either 1 or 2 for the complex or ligand parameters,
respectively. Accordingly, a
1
=D
ML
1/2
K
ad,ML
1
; a
2
=D
L
1/2
-
K
ad,L
1
; Z
a
=R
LM
K
ad,ML
D
L
1/2
[ fu]
1
; Z
b
=R
LM
-
K
ad,ML
[K
ad,L
fP
(1)
]
1
; T
(m)
=D
M(Hg)
1/2
{ur
s
exp[
(m)
]}
1
.
The constant r
s
=1 cm determines the surface and bulk
standard concentration relationship [22]. The parameter
u=2(l/p)
1/2
depends upon the period l=(qf )
1
in-
volved in each numerical integration step, where q
stands for the number of subintervals considered in
each wave and a value of q=40 was employed.
3. Results and discussion
Fig. 1 shows the theoretical voltammograms for both
reaction schemes, Eqs. (5) and (6). Direct (c
f
) and
reverse (c
b
) normalised currents as a function of poten-
tial at 100 Hz are presented for different values of k
s
(curves af). The rst scheme, Eq. (5), is analysed for
u=1 and 2, whereas Eq. (6) is analysed for u=2. The
normalised currents present asymmetric bell-shaped
proles, with marked differences, depending on k
s
. The
sweeps start at the positive potential limit and the
reducing current, c
f
, is considered to be negative.
Curves calculated with Eqs. (24), (25) and (27) using
R
LM
=10 are arranged in Fig. 1A, B and C, respec-
tively. Almost identical cE proles were obtained for
the lowest k
s
value irrespective of the reaction scheme
and the complex stoichiometry, as is shown in curves a,
Fig. 1A, B and C. This is a clear indication that neither
the chemical state of the reduced products nor the
stoichiometry of the complex has any inuence on the
current potential proles, since k
s
is the only parame-
ter controlling the oxidation step. As k
s
is increased, the
quasi-reversible (curves be), and the reversible voltam-
metric responses (curves f) are sensitive to the chemical
properties of the complex, depending on both the u
value and the nal state of the ligand. The voltammet-
ric proles presented in Fig. 1C are independent of R
LM
values for R
LM
0.1. This indicates that ((c
L
/(x)
x=0

0 for all t. In this way, further increments of c


L
* will not
affect the current prole, therefore Eqs. (4), (20) and
(21) allow the same expression found by other authors
when c
L
* c
M
* to be obtained [11,20]:
I(t)/nFA=k
s
exp( h(t))
[Y
MLp
r
s
c
M(Hg),(x=0)
exp((t))] (28)
F. Garay / Journal of Electroanalytical Chemistry 505 (2001) 100108 103
Curves bf in Fig. 1A and B show wider and lower
peaks than those of Fig. 1C. Accordingly, the R
LM
value of 10 is not high enough to discard the ligand
diffusion effect, which is enhanced for u=2, Fig. 1B.
The scheme of Eq. (6) for u=1 gave a voltammometric
response that was indistinguishable from that of Fig.
1C.
Fig. 2 shows a theoretical analysis of the dependences
of Dc
p
(Fig. 2A) and E
p
(Fig. 2B) with log(k
s
) at
different R
LM
values. In both gures, full lines corre-
spond to the scheme of Eq. (5), in which a ligand
desorption step during the complex reduction is consid-
ered, whereas dashed lines correspond to the limiting
case given by Eq. (28). The symbols indicate the differ-
ent values of R
LM
ranging from 10 to 10
4
. Open sym-
bols stand for 1:1 complexes, Eq. (24), whereas the full
ones refer to 1:2 stoichiometry, Eq. (25). The resolution
of Eq. (28) depicts the typical Dc
p
maximum for
quasi-reversible reactions involving adsorbed reactants,
which is clearly seen in Fig. 2A. Nevertheless, if the
ligand desorption takes place during the reduction pro-
cess, this quasi-reversible maximum gradually decreases
with R
LM
, until it virtually disappears for R
LM
10
1
and 10
2
for u=1 and u=2, respectively. Both com-
plex stoichiometries analysed present very dissimilar
current potential dependences when R
LM
100; these
differences can be employed to determine the u value
for a given reaction. The differences at the voltammet-
ric proles for both stoichiometries decrease gradually
as R
LM
is increased, becoming non-detectable for
R
LM
10
4
. For this concentration ratio, Eqs. (24) and
(25) tend to Eq. (28). The same behaviour is observed
for reaction schemes where the ligand remained ad-
sorbed, Eq. (6), indicating that Eqs. (26) and (27) also
tend to Eq. (28), but when R
LM
1 (not shown). In this
case, the required R
LM
value is four orders of magni-
tude lower than the reaction scheme of Eq. (5). This
fact points to a strong levelling off effect due to the
ligand adsorption equilibrium, which guarantees a con-
stant ligand concentration close to the electrode and
validates the employment of the quasi-reversible maxi-
mum for the evaluation of k
s
[11,12,20].
The dependences of E
p
on log(k
s
) show three well-dif-
ferentiated zones (Fig. 2B). When log(k
s
) 2, all
systems are controlled basically by the kinetic contribu-
tion, with a slope of 2.3RT/nF V dec
1
, regardless
of the effects of R
LM
and u values. A completely
different behaviour is observed for reversible redox
reactions for which log(k
s
) .2. In this case, E
p
is
strongly dependent on R
LM
and on the complex stoi-
chiometry, although it is not affected by k
s
. Finally,
there is an intermediate zone corresponding to the
quasi-reversible redox reactions, in which the different
variables analysed, k
s
, R
LM
and u, affect the E
p
values
jointly. As for Fig. 2A, all curves approach that of Eq.
(28) as R
LM
is raised (dashed line).
The effect of R
LM
on the theoretical SW voltammet-
ric proles calculated according to Eqs. (24), (25) and
(27) is examined in Fig. 3A, B and C, respectively.
These responses were obtained for k
s
=0.3 s
1
, f =100
Hz, changing R
LM
over four orders of magnitude. If the
reduction process does not affect the ligand gradient at
the electrode surface, all the reaction schemes consid-
ered will give the same voltammetric response. The
Fig. 1. Theoretical cE proles from Eq. (5) for u=1 (A) and u=2 (B) and from Eq. (6) for u=2 (C), f =100 Hz, E
sw
=50, dE=5 mV,
K
ad,ML
=0.1 cm, R
LM
=10, n=1, h=0.5 and k
s
=0.01 (a); 0.1 (b); 0.3 (c); 0.6 (d); 1 (e) and 10
3
s
1
(f).
F. Garay / Journal of Electroanalytical Chemistry 505 (2001) 100108 104
Fig. 2. Dependence of (A) Dc
p
and (B) E
p
as functions of log(k
s
),
obtained from Eq. (5) for u=1 (open symbols) and u=2 (full
symbols). R
LM
=10
4
( ); 10
3
(); 10
2
() and 10
1
(). Other
parameters as in Fig. 1.
high enough to ensure that the ligand gradient will not
be affected by the electrochemical redox reaction, Fig.
3C. However, completely different currentpotential
proles for this scheme are observed, when R
LM
is
below the lower limit. As R
LM
is diminished, the back-
ward current (in which the complex re-oxidation takes
place) is also diminished until it becomes undetectable.
Provided that c
L
* is very low, the ligand desorption is
almost complete during the reduction pulse, followed
by its diffusion towards the bulk. Consequently, during
the following oxidation pulse, Y
L
will be very low,
hindering the oxidation of the amalgamated metal at
this potential.
Fig. 4 shows how Dc
p
(Fig. 4A) and E
p
(Fig. 4B),
arising from the scheme of Eq. (5) vary as functions of
log(R
LM
) for different values of k
s
. Open and full
symbols stand for complex stoichiometries 1:1, Eq. (24),
and 1:2, Eq. (25), respectively. Three distinct types of
behaviour are observed. First, for log(R
LM
) .3, both
parameters are controlled basically by k
s
, and conse-
quently, neither R
LM
nor the complex stoichiometry has
a noticeable effect on the voltammetric responses. Sec-
ond, there is a transitional region for 0log(R
LM
) 3,
where complex behaviour is observed. The strong
changes produced by k
s
and R
LM
on Dc
p
predict that
either the enhancement or the diminution of Dc
p
can
be commanded by changing the ligandcomplex ratio.
These tendencies depend directly on the value of k
s
with
regard to the quasi-reversible maximum. The enhance-
ment of the peak current when R
LM
is diminished
depends on the driving force of the c
L
gradient, which
promotes the occurrence of the redox reaction. There-
fore, experiments performed with low values of R
LM
are
useful to differentiate the complex stoichiometry,
whereas the higher values help to determine the k
s
value. Finally, when log(R
LM
) 0, the responses are
determined by diffusion and Dc
p
becomes independent
of k
s
. In this region, Dc
p
has a constant value, charac-
teristic of the complex stoichiometry.
In the case of E
p
(Fig. 4B), provided log(R
LM
) .3,
the kinetic contribution essentially controls the voltam-
metric responses and hence the E
p
values remain con-
stant for increasing log(R
LM
) values. For log(R
LM
) 0,
E
p
varies linearly with log(R
LM
) presenting slopes of
0.060 and 0.120 V dec
1
and pointing to a diffusional
control for both stoichiometries, Eqs. (24) and (25),
respectively. This dependence of E
p
on both R
LM
and k
s
is similar, since both parameters govern the reversibility
of the global reaction, the rst controlling the chemical
kinetics of the complex formation and the second com-
manding the electrochemical rate. For the reaction
scheme of Eq. (6), which considers the ligand adsorp-
tion step, E
p
values are constants provided log(R
LM
) .
2, since they are determined mainly by k
s
(not
shown).
latter can be achieved when R
LM
1, Eq. (28), or when
the ligand adsorption produces the levelling off effect
previously described, Eqs. (26) and (27). Differences
between each mechanism start to be evident at the
voltammograms when R
LM
is lower than a given limit-
ing value, which is strongly dependent on the ligand
state at the electrode surface (Fig. 3A (curve e), B
(curve c) and C (curve i), respectively). However, when
the ligandcomplex ratio is diminished by three orders
of magnitude in relation to the limiting value, the
current potential proles maintain their shapes, but a
shift in the E
p
value towards more negative potentials is
observed (Fig. 3A (curves g and h), B (curves eg) and
C (curves k and l)). The diffusion contribution to the
current is evident in Fig. 3A (curves eg) and B (curves
ce) in which the diffusional current tails at more
positive potentials than E
p
are increased as R
LM
is
decreased. This effect is more important for u=2, Fig.
3B. On the other hand, if the ligand remains adsorbed
after the complex reduction, an R
LM
value of just 0.1 is
F. Garay / Journal of Electroanalytical Chemistry 505 (2001) 100108 105
Fig. 3. Theoretical cE proles obtained from Fig. 5 for u=1 (A) and u=2 (B) and from Fig. 6 for u=1 (C). k
s
=0.3 s
1
and R
LM
=10
4
(a);
10
3
(b); 10
2
(c); 30 (d); 10 (e); 3 (f); 1 (g); 0.1 (h); 10
3
(i); 310
4
(j); 10
4
(k) and 10
5
(l). Other parameters as in Fig. 1.
Fig. 5 compares, according to the scheme of Eq. (6),
the dependence of Dc
p
with log(K
ad,L
) obtained for
different values of R
LM
(symbols), for u=1 (dashed
lines) and for u=2 (full lines). Considering k
s
=1 s
1
,
below the quasi-reversible maximum (k
s,max
), the curves
show two kinds of well-dened behaviour. At low
log(K
ad,L
), Dc
p
increases steadily with log(K
ad,L
) until it
reaches a constant value. High R
LM
and 1:1 stoi-
chiometries expand the range of K
ad,L
values for which
Dc
p
is constant, which is achieved for R
LM
K
ad,L

0.1 (for u=1) and 0.3 (for u=2), respectively. Con-


sequently, Eqs. (26) and (27) will present similar
voltammetric responses if R
LM
K
ad,L
exceeds these
limiting values.
Fig. 6 shows the theoretical dependence of Dc
p
versus log(K
ad,L
) for k
s
values below and above the
k
s,max
value, considering u=1 and R
LM
=10. The curve
for k
s
=1 s
1
, shown in Fig. 5, is included here for
comparison. The tendency of increasing Dc
p
with
log(K
ad,L
) can be reversed when k
s
.k
s,max
or remain
independent of K
ad,L
for highly irreversible reactions.
When k
s
k
s,max
, Dc
p
will be controlled completely by
k
s
, regardless of the ligand desorption or its surface
concentration. The condition of R
LM
K
ad,L
0.1
holds for all k
s
values analysed, since Dc
p
does not
change with K
ad,L
for log(K
ad,L
) .2. For slow elec-
tron transfer reactions (k
s
k
s,max
), the adsorption equi-
librium can be restored before the current sampling at
the end of the pulse; therefore, the diminution of R
LM
and K
ad,L
will decrease the value of Y
L
, diminishing in
this way the value of Dc
p
. On the other hand, for more
reversible redox reactions (k
s
.k
s,max
), Dc
p
is enhanced
by decreasing K
ad,L
. This effect depends directly on the
increase of the ligand concentration close to the elec-
trode surface. After the ligand desorption takes place,
Y
L
will not be high enough to drain the whole reduced
Fig. 4. Dependence of (A) Dc
p
and (B) E
p
as functions of log(k
s
),
obtained from Eq. (5) for u=1 (open symbols) and u=2 (full
symbols). k
s
=10
3
(); 1 (); 0.3 () and 0.1 s
1
(). Other
parameters as in Fig. 1.
F. Garay / Journal of Electroanalytical Chemistry 505 (2001) 100108 106
Fig. 5. Dependence of Dc
p
versus log(K
ad,L
) obtained from Eq. (6)
for u=1 (open symbols) and u=2 (full symbols), k
s
=1 s
1
, R
LM
=
10 (); 1 () and 0.1 (). Other parameters as in Fig. 1.
the net peak height. Likewise, the E
p
shifts slightly
towards more positive values, Fig. 7B. By nE
sw
100
mV, the skew in the c
f
and c
b
has dened a shoulder
at the Dc (curves d and e).
For irreversible redox reactions, the peak width re-
mains constant, with DE
p/2
=(63.5_0.5)/hn mV, for
every SW amplitude and the increment of E
sw
will just
enhance the Dc
p
value [11]. Nevertheless, for reversible
and quasi-reversible redox reactions, larger values of
E
sw
increase the peak width linearly (not shown).
The calculated dependences of the Dc
p
DE
p/2
1
ratio
versus E
sw
are shown in Fig. 8A and B, considering
reversible and quasi-reversible charge transfer reactions,
respectively. Both gures show the changes in
Dc
p
DE
p/2
1
when R
LM
is varied from 1 to 10
3
(symbols)
according to the scheme of Fig. 5 and for u=1. The
behaviour for both charge transfer kinetics indicates
that Dc
p
DE
p/2
1
is enlarged by the increment of R
LM
and E
sw
. Particularly for E
sw
40 mV, the analytical
signal increases more rapidly. For the case of a quasi-
reversible system (k
s
=1), Fig. 8B, the response quality
increases for E
sw
up to 100 mV, except for the case of
R
LM
=10
3
, which has a maximum at 80 mV. For the
reversible case (k
s
=10
4
), Fig. 8A, the maximum at
E
sw
=40 mV is only for R
LM
=10
3
, whereas for R
LM
=
10
2
the maximum is shifted to 50 mV. For R
LM
=10
and 1 there is no maximum in the E
sw
range exploited.
Fig. 6. Dependence of Dc
p
versus log(K
ad,L
) obtained from Eq. (6)
for u=1, k
s
=10
4
(); 10 (); 1 () and 10
2
s
1
(). Other
parameters as in Fig. 1.
Fig. 7. Theoretical cE proles from Fig. 5 for u=1, k
s
=1 s
1
,
E
sw
=5 (a); 10 (b); 20 (c); 40 (d) and 100 mV (e). Other parameters
as in Fig. 1.
species at the beginning of the oxidation pulse. Low
values of K
ad,L
will favour the ligand desorption during
the reduction pulse, increasing the c
L
near the electrode
surface and producing an enhancement of Dc
p
.
Fig. 7 shows how the morphology of forward and
backward currents (Fig. 7A) as well as of Dc (Fig. 7B)
varies with E
sw
, considering a 1:1 complex according to
Eq. (5). Reduction currents are observed for direct and
reverse pulses for E=5 mV, Fig. 7A. As the SW
amplitude increases, an oxidation current is observed at
the reverse pulse making a substantial contribution to
F. Garay / Journal of Electroanalytical Chemistry 505 (2001) 100108 107
Fig. 8. Dependence of Dc
p
DE
p/2
1
on E
sw
obtained from Eq. (6) with
u=1. For (A) k
s
=10
4
and (B) k
s
=1 s
1
, R
LM
=1 (); 10 (); 10
2
() and 10
3
(). Other parameters as in Fig. 1.
From the analytical point of view, the theoretical
curves provide a comprehensive way to maximise the
Dc
p
DE
p/2
1
ratio. Also, the elucidation of the mecha-
nism points out how the ligand concentration and the
other experimental parameters should be tuned to max-
imise the current response and to produce one well-
shaped peak at a favourable potential.
Acknowledgements
Financial support from the Consejo Nacional de
Investigaciones Cient cas y Tecnolo gicas (CONICET),
Consejo de Investigaciones de la Provincia de Co rdoba
(CONICOR) and Secretar a de Ciencia y Tecnolog a de
la Universidad Nacional de Co rdoba is gratefully ac-
knowledged. The author also wishes to thank Dr
Milivoj Lovric and Dr Velia Solis for very helpful
discussions.
Appendix A
A electrode surface area (cm
2
)
=D
ML
1/2
K
ad,ML
1
(s
1/2
) a
1
a
2
=D
L
1/2
K
ad,L
1
(s
1/2
)
c
L
*; c
ML
u
* bulk ligand and complex concentra-
tions (mol cm
3
)
c
M(Hg)
; c
ML
u
; concentrations of M, ML
u
and L

c
L
near the electrode surface (mol cm
3
)
D diffusion coefcient (cm
2
s
1
)
dE SW step amplitude (V)
DE
p/2
half-peak width (V)
E
(t)
SW potential program (V)
E standard potential for a simple redox
reaction of free soluble species (V)
E
p
peak potential (for the net current)
(V)
E
sw
half peak-to-peak SW potential ampli-
tude (V)
f SW frequency (Hz)
Faraday constant (C) F
i =mj+1
current (A) I
(t)
K
ad,L
; K
ad,ML
ligand and complex adsorption con-
stants (cm)
k
s
standard reaction rate constant (s
1
)
the value of k
s
at the quasi-reversible k
s,max
maximum (s
1
)
K
st,o
stability constant of oxidised species
(cm
3
mol
1
)
stability constant of reduced species K
st,r
(cm
4
mol
1
) or (cm
7
mol
2
)
n number of electrons
P
(i )
={uS
(i )
+Y
2(i )
}a
2
1
(s)
=40; number of time increments in q
each SW period
Consequently, a SW amplitude of 100 mV is preferred
for reversible reactions in which R
LM
10
1
.
In the case of irreversible charge transfer reactions, a
suitable Dc
p
DE
p/2
1
value was obtained for E
sw
=100
mV, regardless of the R
LM
value (not shown).
4. Conclusions
A description of the SW voltammetric behaviour
considering the ligand concentration effect over non-
labile metallic complexes has been formulated. The
theoretical voltammograms obtained according to the
proposed schemes permit the morphology of the re-
sponse for a wide range of experimental parameters to
be examined. These schemes can be found for broad
categories of reactions, especially for the case of SW
stripping analysis of adsorbed metallic complexes.
The schemes in this study provide a means to charac-
terise the electrochemical properties of adsorbed metal-
lic complexes without employing a great ligand excess.
Low ligand concentrations can be adjusted to discern
the complex stoichiometry as well as to determine the
ligand chemical state. In contrast, high ligandcomplex
ratios emphasise features concerning the rate process.
Thus, the quasi-reversible maximum is increased and
the effect of k
s
is stronger.
F. Garay / Journal of Electroanalytical Chemistry 505 (2001) 100108 108
={uS
(i )
+Y
1(i )
}a
1
1
(s) Q
(i )
gas constant (J mol
1
K
1
) R
R
LM
ligandcomplex concentration ratio
=1 cm r
s
=(i )
1/2
(i1)
1/2
S
(i )
T temperature (K)
=D
M(Hg)
1/2
{ur
s
exp[
(m)
]}
1
T
(m)
complex stoichiometry u

(m)
=exp[h
(m)
]k
s
1
+Q
(1)
(s)
={exp[a
y
2
(li )] erfc[a
y
(li )
1/2
] Y
y(i )
exp[a
y
2
(l(i1))] erfc[a
y
(l(i1))
1/2
]}
a
y
1
(s
1/2
)
Z
a
=R
LM
K
ad,ML
D
L
1/2
[ fu]
1
Z
b
=R
LM
K
ad,ML
[K
ad,L
fP
(1)
]
1
=
m1
j =1
c
( j )
S
(i )
]
a(m)
=
m1
j =1
c
( j )
Q
(i )
(s) ]
b(m)
=P
(1)
1

m1
j =1
c
( j )
P
(i )
]
c(m)
charge transfer coefcient h
time increments in each SW period (s) l
dimensionless SW potential program (t)
initial ligand and complex surface con- Y
L
ini
; Y
ML
u
ini
centrations (mol cm
2
)
ligand and complex surface concentra- YL; Y
ML
u
tions (mol cm
2
)
u =2(l/p)
1/2
(s
1/2
)
=I
(t)
/(nFAfY
ML
u
ini
) dimensionless c(t)
current function
forward and backward normalised cur- c
f
, c
b
rents
=c
f
c
b
; normalised net current Dc
normalised net peak current Dc
p
References
[1] G. Paneli, A. Voulgaropoulos, Electroanalysis 5 (1993) 355.
[2] H. Sawamoto, Bunseki Kagaku 48 (1999) 137.
[3] K. Bruland, E. Rue, J. Donat, S. Skrabal, J. Moffett, Anal. Chim.
Acta 405 (2000) 99.
[4] A. Bond, Anal. Chim. Acta 400 (1999) 333.
[5] O. Abollino, M. Aceto, C. Sarzanini, E. Mentasti, Electroanalysis
11 (1999) 870.
[6] M. Shi, F. Anson, Anal. Chem. 70 (1998) 1489.
[7] J. ODea, J. Osteryoung, Square wave voltammetry, in: A.J. Bard
(Ed.), Electroanalytical Chemistry, vol. 14, Marcel Dekker, New
York, 1986.
[8] P. Molina, M. Zon, H. Fernandez, Electroanalysis 12 (2000) 791.
[9] S& . Komorsky-Lovric, M. Lovric, J. Electroanal. Chem. 384 (1995)
115.
[10] M. Lovric, S& . Komorsky-Lovric, A. Bond, J. Electroanal. Chem.
319 (1991) 1.
[11] M. Lovric, S& . Komorsky-Lovric, J. Electroanal. Chem. 248 (1988)
239.
[12] S& . Komorsky-Lovric, M. Lovric, Fresenius Z. Anal. Chem. 335
(1989) 289.
[13] F. Garay, V. Solis, J. Electroanal. Chem. 476 (1999) 165.
[14] F. Garay, V. Solis, M. Lovric, J. Electroanal. Chem. 478 (1999) 17.
[15] M. Lovric, M. Branica, J. Electroanal. Chem. 226 (1987) 239.
[16] I. Pizeta, M. Lovric, M. Zelic, M. Branica, J. Electroanal. Chem.
318 (1991) 25.
[17] J. ODea, J. Osteryoung, Anal. Chem. 65 (1993) 3090.
[18] J. ODea, J. Osteryoung, Anal. Chem. 69 (1997) 650.
[19] M. Lovric, J. Electroanal. Chem. 465 (1999) 30.
[20] M. Lovric, Electrokhimiya 32 (1996) 1068.
[21] R. Carlin, P. Trulove, R. Mantz, J. ODea, R. Osteryoung, J. Phys.
Chem. 92 (1996) 3969.
[22] M. Lovric, I. Pizeta, S& . Komorsky-Lovric, Electroanalysis 4 (1992)
327.
[23] R.S. Nicholson, M. Olmstead, in: J. Matson, H. Mark, H. Mac-
donald (Eds.), Electrochemistry: Calculations, Simulations and In-
strumentation, vol. 2, New York, 1972.
.

You might also like