You are on page 1of 7

Experimental Validation of Marcus

Theory for Outer-Sphere


Heterogeneous Electron-Transfer
Reactions: The Oxidation of
Substituted 1,4-Phenylenediamines
Antony D. Clegg, Neil V. Rees, Oleksiy V. Klymenko,
Barry A. Coles, and Richard G. Compton*
[a]
Introduction
In this Communication, we report a quantitative investigation
of Marcus theory
[16]
using high-precision measurements of fast
outer-sphere heterogeneous electron-transfer rate constants.
Theoretical developments since Marcus' initial work have con-
sidered the outer-sphere electron-transfer process to involve
the formation of a precursor complex between the reactant
molecule and the electrode surface,
[7]
leading to the commonly
found expression for the standard electrochemical rate con-
stant [Eq. (1)]
k
0
= k
el
K
p
n
n
exp

DG

RT

(1)
where K
p
is the equilibrium constant for the precursor complex
formation, DG

is the free energy of activation for the electron


transfer, n
n
is the frequency of crossing the free energy barrier,
and k
el
is the probability of electron tunnelling in the transition
state.
[68]
The pre-exponential terms are sensitive to the system,
and in particular the nuclear frequency and electronic trans-
mission factors have been considered in detail.
[6, 7, 9, 10]
It has
been shown that if the reaction free energy is zero, and the
[a] A. D. Clegg, N. V. Rees, O. V. Klymenko, Dr. B. A. Coles, Prof. R. G. Compton
Physical and Theoretical Chemistry Laboratory
Oxford University, South Parks Road
Oxford OX1 3QZ, (UK)
Fax: (+44) 1865-275410
E-mail : richard.compton@chemistry.ox.ac.uk
1234 2004 Wiley-VCH Verlag GmbH&Co. KGaA, Weinheim DOI: 10.1002/cphc.200400128 ChemPhysChem 2004, 5, 1234 1240
weak overlap limit is assumed, that is the electronic coupling
is small, then Equation (2) holds,
[8]
n
n
= t
1
L

DG

4pRT
1
=
2
(2)
where t
L
is defined as t
L
=t
D
e

/e
s
, t
D
is the experimental
Debye relaxation time and e

and e
s
are the high-frequency
and static dielectric permittivities, respectively.
[7]
In order to ac-
count for any nonadiabaticity in the electron transfer, which is
often observed in outer-sphere reactions,
[7]
the electronic
transmission probability, k
el
, can be stated as in Equation (3),
k
el
= k
0
el
exp[B(r
/
s)[ (3)
where r is the moleculeelectrode separation, s is the distance
of closest approach of the molecule and electrode, and B is a
constant.
[7]
If the free energy of activation is solely due to the
outer-sphere reorganisation energy then Equation (4) applies:
l
o
=
N
A
e
2
8pe
0

1
r

1
2d

1
e
op

1
e
s

(4)
where e is the electronic charge, r is the radius of the mole-
cule, and d is the distance from the reactant to the metal sur-
face
[11]
which is usually set to infinity following Hale,
[12]
then
Equation (1) can be expressed as Equation (5),
k
0
= Q

y
4p
1
=
2

1
r
1
=
2
exp

Br
y
r

(5)
where Q=K
p
k
0
el
t
1
L
exp[B(ds)] , y=(N
A
e
2
/32pe
0
RT)[(1/e
op
)(1/
e
s
)], and r =r +d.
There have been few attempts to interpret experimentally
determined rates of electron transfer using expressions such as
(1) and (5), and most comparisons with Marcus theory remain
in a qualitative sense only. One reason for this may lie in the
necessity to obtain high-precision measurements of k
0
, togeth-
er with the need to be able to measure very fast as well as
slow values of k
0
, to perform a satisfactory test of the theory.
In most cases, the method of choice for the kineticist would
be cyclic voltammetry including the recently developed field
of fast-scan voltammetry.
[1315]
However, the experimental diffi-
culties of obtaining precise and accurate data via these meth-
ods are documented.
[16, 17]
The exacting requirements for preci-
sion and reliability of the experimental methodology lend
themselves to steady-state methods, since there is negligible
distortion from iR or capacitative effects. We have therefore se-
lected the high-speed channel electrode (HSChE) for this study
because of its proven ability to unambiguously measure kinetic
information for the fastest electrode processes,
[1821]
although
we note that small amplitude a.c. voltammetry offers some
merits in this context since capacitance contributions are re-
duced (but not eliminated) compared to fast-scan cyclic vol-
tammetry.
[2224]
In a recent publication, we developed a methodology for
measuring k
0
for fast electron transfers with a view to testing
Marcus Theory expressions such as (5), using the HSChE to
make the necessary high-precision measurements of electron
transfer rates.
[25]
We also considered what measures of the mo-
lecular radius r might be used. Of the several methods availa-
ble, such as crystallographic or computed values from the
mean spherical or ellipsoidal approximations,
[8, 26, 27]
it was sug-
gested that the hydrodynamic radius would provide the most
physically meaningful value, despite having been rarely used
as such in the literature.
[8]
It is recognised that this method as-
sumes that the electron transfer is orientation-independent, al-
though in practice some weighted average of molecular orien-
tations and their respective radii will probably be relevant.
Nevertheless, in the absence of specific orientational require-
ments or adsorption, the hydrodynamic radius should be relat-
ed to the true effective radius by some factor, and therefore
should show the correct trend for our interpretation. The hy-
drodynamic radius is also conveniently measurable from exper-
imental voltammetry, since it can be simply calculated from
the StokesEinstein equation [Eq. (6)]:
[28]
r =
kT
PphD
(6)
in which h is the viscosity, D is the diffusion coefficient, and P
is either 4 or 6 depending whether the stick or slip limit is
assumed for Equation (6).
[28]
To further assist the fitting of theo-
retical results to the experimental data, Equation (5) can be lin-
earised, and also rendered dimensionless for convenience by
making the substitutions y=r/y, b=yB, K=k
0
/k and q=Q/k',
where k =1 cms
1
; it then becomes Equation (7).
ln(K

y
_
)
1
y
= by ln

q
2

p
_

(7)
Herein we present experimental results for comparison with
Equations (2) and (7) to investigate the validity of the applica-
tion of these theoretical results to heterogeneous electron
transfer over a range of molecular radii from 3 to 9 in alkyl
cyanide solvents. This range of molecular sizes is significantly
wider than previously attempted
[25]
and hence a more rigorous
test of the validity of the above equations. First, the solvent
effect on the rate of electron transfer is studied for the oxida-
tion of N,N,N,N-tetramethyl-1,4-phenylenediamine dihydro-
chloride (TMPD) in the solvents acetonitrile (MeCN), propioni-
trile (EtCN), butyronitrile (PrCN), and valeronitrile (BuCN) with
reference to Equation (2). Then, the variation of k
0
for a series
of related compounds are investigated by considering the first
oxidations of the ten tetraalkyl-1,4-phenylenediamines (TRPDs)
shown in Table 1, in MeCN.
Experimental Section
Reagents: The chemical reagents used for these experiments that
were available from commercial sources were TMPD and N,N-di-
ethyl-1,4-phenylenediamine (DEPD), and N,N-diphenyl-1,4-phenyl-
enediamine (DPPD; Aldrich, 98%), tetrabutylammonium perchlo-
ChemPhysChem 2004, 5, 1234 1240 www.chemphyschem.org 2004 Wiley-VCH Verlag GmbH&Co. KGaA, Weinheim 1235
rate (TBAP; Fluka, Puriss. >99%), and the solvents MeCN (Fisher
scientific, >99.99%), EtCN (Fluka, purum>99%), PrCN (Aldrich,
98%), and BuCN (Fluka, puriss >99%). These were the highest
grades available, and were used without further purification. TBPD,
THxPD, THpPD, TOPD, DEDB, and DEDHx (see Table 1 for abbrevia-
tions) were synthesised by Prof. D. J. Walton.
[2931]
Solvents were
stored over molecular sieves (Linde 5 , Aldrich) for several hours
prior to use and thoroughly degassed with argon (Pureshield
Argon, BOC Gases Ltd, UK) before and after solution preparation.
All solutions contained 0.10m TBAP as supporting electrolyte and
experiments were conducted at a temperature of 2942 K.
Instrumentation: The high-speed channel electrode (HSChE) and
pressurised apparatus have been described previously
[18, 19, 21]
(see
Figure 1). High flow rates across a microband electrode are ach-
ieved by pressurising a chamber containing the solution and elec-
trode assembly up to 1.5 atm. The solution passes through the
flow-cell (width, d=0.200 cm and height, 2 h=126 mm) and out
through one of three capillaries of varying internal bore size, to
the exit at atmospheric pressure. This achieves volume flow rates
of between 0.103.2 cm
3
s
1
(corresponding to linear flow velocities
close to the electrode of 0.7 to
20 ms
1
), and the Reynolds
number, R
e
, given by Equa-
tion (8),
[32]
R
e
=
3V
f
2hdn
(8)
can attain maximum values of
9000 under well-defined laminar
conditions,
[18, 19]
since the channel
flow-cell has been designed to
ensure these Reynolds numbers
are present for less than 2 mm
before the electrode, whilst a
lead-in length of approximately
4 mm is needed for the development of turbulent flow.
[19]
Voltam-
mograms are measured by means of an in-built potentiostat at a
scan rate of 400 mVs
1
with a platinum microband electrode of
length (x
e
) 40.5 mm.
The potentiostat used for microdisk voltammetry was a mAutolab
Type II (Eco Chemie BV, Utrecht, Netherlands) controlled by a Dell
Optiplex GX110 Pentium III computer using General Purpose Elec-
trochemical System v4.8 software (Eco Chemie BV, Utrecht, Nether-
lands).
Microelectrodes: Microband and microdisk electrodes were fabri-
cated by fusing platinum (99.95%, Johnson Matthey plc, London,
UK) into soda glass according to literature methods,
[19, 25]
and their
working surfaces ground and polished to a mirror finish. The mi-
croband electrode width (w), as measured by microscope, was
0.096 cm. For both types of electrode, the dimensions were con-
firmed by electrochemical calibration,
[33]
and the radius (r
d
) of the
microdisk found to be 14.1mm. The electrodes were cleaned with
ultrapure water and polished using 0.25 mm alumina slurry on soft
lapping pads, then finally rinsed in ultrapure water and dried care-
fully before use. The counter electrode was a smooth, bright, plati-
num mesh, and a silver wire (99.95%, Johnson Matthey plc,
London, UK) was used as a quasi-reference electrode. Microdisk
voltammetry was performed within a Faraday cage to reduce elec-
trical noise.
Analysis of hydrodynamic voltammetry: The analysis of steady-
state voltammograms recorded at the HSChE has been reported.
[20]
The current response for each voltammogram was normalised by
division by the respective limiting current, I
lim
, and the middle 60%
of the wave was plotted against the potential. These data were
then input into a program and compared against a calculated vol-
tammogram of (I/I
lim
) versus E for a selected range of a, k
0
and E
0
f
values. The quasi-reversible electron transfer to be simulated can
be treated as the Reaction (9):
Ae
k
f
k
b

B (9)
where
k
f
= k
0
exp

(1a)F
RT
(EE
0
f
)

= k
0
exp[(1a)q[ (10)
k
b
= k
0
exp

aF
RT
(EE
0
f
)

= k
0
exp(aq) (11)
and its simulation was achieved by using the following analytical
results [Equations (12) to (14)]:
[18, 20, 34, 35]
Table 1. The tetraalkyl-1,4-phenylenediamines under investigation.
R
1
R
2
R
3
R
4
Abbreviation
H H H H PPD
ethyl ethyl H H DEPD
methyl methyl methyl methyl TMPD
n-butyl n-butyl n-butyl n-butyl TBPD
n-hexyl n-hexyl n-hexyl n-hexyl THxPD
n-heptyl n-heptyl n-heptyl n-heptyl THpPD
n-octyl n-octyl n-octyl n-octyl TOPD
ethyl ethyl n-butyl n-butyl DEDB
ethyl ethyl n-hexyl n-hexyl DEDHx
H phenyl H phenyl DPPD
Figure 1. Schematic diagram of the channel cell and microband electrode
showing the geometrical parameters.
1236 2004 Wiley-VCH Verlag GmbH&Co. KGaA, Weinheim www.chemphyschem.org ChemPhysChem 2004, 5, 1234 1240
i
i
rev
= 12u 2u
2
ln(1 u
1
) (12)
where
u =
0:6783D
2
=
3
B
(3V
f
=4dx
e
h
2
)
1
=
3
k
0
exp[(1a)q[ (D
A
=D
B
)
2
=
3
exp(aq)
(13)
and
i
rev
=
0:925nFw[A[
bulk
(x
e
D
A
)
2
=
3
(h
2
d)

1
=
3
V
1
=
3
f
1 (D
A
=D
B
)
2
=
3
exp(q)
(14)
V
f
is the volume flow rate, [A]
bulk
is the bulk concentration of the
electroactive species, D
i
is the diffusion coefficient of species i, and
the geometrical parameters are as given in Figure 1. The quantity
i
rev
is the current that would flow if the electrode kinetics were re-
versible. These approximate results have been verified against nu-
merical simulations using the backward implicit finite difference
method,
[18]
and found to be valid over the range of flow rates ach-
ievable in the HSChE.
For each simulated voltammogram, a mean scaled absolute devia-
tion (MSAD) given by Equation (15)
MSAD(a,k
0
,E
0
f
) =
X
N
k=1
[I
exp
(E
k
)I
th
(a,k
0
,E
k
E
0
f
)[
I
exp
(E
k
)
(15)
was calculated as the sum of the differences between each simu-
lated point (I
th
) and each experimental point (I
exp
). Here, E
k
, k=
1,,N are the potentials of the experimental data points under
analysis, and N is typically above 20. Each simulated voltammo-
gram was therefore awarded its own MSAD value and this enabled
a minimum to be found (by a Direct Search Method) correspond-
ing to the optimum values of a, k
0
and E
0
f
. Contour plots of MSAD
as a function of a and k
0
, and as a function of k
0
and E
0
f
were pro-
duced showing the existence of a single minimum in every case.
Results and Discussion
First, a solution containing 1.12 mm TMPD and 0.10m TBAP in
MeCN was introduced into the pressure chamber of the HSChE
apparatus fitted with the 40.5 mm Pt microband electrode. A
linear sweep voltammogram was then recorded at an arbitrary
flow rate (1.14 cm
3
s
1
) which yielded an effectively steady-
state response, enabling a limiting current (I
lim
) to be measured
for the first oxidation wave as shown in the inset of Figure 2.
This was repeated for a range of volume flow rates (V
f
) from
0.15 to 3.10 cm
3
s
1
. Figure 2 also shows a Levich plot of
measured limiting currents against V
1
=
3
f
, according to the Levich
equation
[36]
for channel flow cells [Eq. (16)]:
I
lim
= 0:925nFw[A[
bulk
(x
e
D)
2
=
3

V
f
h
2
d

1
=
3
(16)
where n, F, V
f
, and [A]
bulk
have their usual meaning, and the
other geometrical terms are shown in Figure 1. The gradient of
this plot was used to obtain a value for the diffusion coeffi-
cient of TMPD in MeCN of (2.200.35) 10
5
cm
2
s
1
, which
compares with a literature value of 2.110
5
cm
2
s
1
.
[37]
These
data were then put into the computer program described
above and optimum values of k
0
, a and E
0
f
were found simulta-
neously for each voltammogram and the mean values taken.
Figure 3 shows the three-dimensional surface plots for k
0
,E
0
f
and k
0
,a as they vary independently with MSAD. The plots
Figure 2. Levich plot for 1.12 mm TMPD in MeCN/0.1m TBAP (gradient =10.03 -
mAcm
1
s
1/3
; R
2
=0.984). The inset shows a typical steady-state linear-sweep
voltammogram (V
f
=1.14 cm
3
s
1
).
Figure 3. Contour plots for TMPD in MeCN, with V
f
=1.87 cm
3
s
1
, showing a) k
0
vs. E
f
0
, and b) k
0
vs. a. Numbers shown on contours are the MSAD values.
ChemPhysChem 2004, 5, 1234 1240 www.chemphyschem.org 2004 Wiley-VCH Verlag GmbH&Co. KGaA, Weinheim 1237
show single minima, which demonstrates that there is only
one set of optimised values.
The same procedure was repeated for solutions of TMPD
and 0.10m TBAP in EtCN (1.16 mm), PrCN (1.16 mm), and BuCN
(1.32 mm). In all cases, excellent Levich plots were obtained
(with R
2
_ 0.990), and Table 2 lists the diffusion coefficients
and kinetic parameters derived from analysis.
A logarithmic plot of the determined vales of k
0
against the
reported longitudinal dielectric relaxation times (MeCN=
0.20 ps, EtCN=0.31 ps, PrCN=0.52 ps, BuCN=0.74 ps)
[38]
ac-
cording to Equation (17), yields a linear plot with gradient
0.92 as shown in Figure 4. This result compares with the the-
oretically expected inverse relationship with t
L
. However, there
are several studies that have investigated the electron transfer
rates in a variety of different solvents
[3944]
using an empirical
form of Equation (1) to account
for observed departures from
the inverse dependence on t
L
,
namely Equation (17).
lnk
0
= q lnt
L
lnA (17)
These reports have confirmed
a linear relationship between
lnk
0
and lnt
L
for various values
of the constant q ranging from
10.2, and also indicate the
presence of an outer-sphere
rather than an inner-sphere
pathway for TMPD.
[3941, 43]
In these papers, q=1 indicates an
adiabatic, purely outer-sphere electron transfer, and therefore
q<1 indicates a degree of either nonadiabaticity or inner-
sphere contribution to the reorganisation energy.
In particular, our results for TMPD are consistent with report-
ed findings which considered the relationship between the for-
ward rate constant k
f
[see Eq. (9)] and t
L
for 12 solvents, which
gave an average value of q=
0.53, although the value recorded
in MeCN reflects a higher value
of q.
[39]
Considering these results,
we infer that the departure of
the gradient from unity is due to
a failure of the solvent continu-
um model, perhaps attributable
to the solvation of the amino
groups by RCN molecules caus-
ing additional solvent inertial effects.
Next, the oxidation of different TRPDs was investigated by
the HSChE method using solutions of TBPD (1.16 mm), DEDB
(1.16 mm), DEDHx (1.84 mm) and DPPD (1.00 mm) in 0.10m
TBAP and MeCN. In all cases, excellent Levich plots were ob-
tained (with R
2
_0.980) which yielded values for the respective
diffusion coefficients that were all in good agreement (see
Table 3) with those predicted by the WilkeChang expression
[Eq. (18)] .
[45]
D ~
(7:4 10
8
)T

m
s
_
hs
0:6
(18)
where T is the absolute temperature, f is the solvent affinity
factor (unity for aprotic solvents), m
s
is the molecular mass of
the solvent, h is the solvent viscosity, and s the molecular
volume.
[45]
The analysis confirmed that these compounds ex-
hibit a simple one-electron oxidation, and the kinetic parame-
ters derived are shown in Table 3 (meanstandard deviation).
Previously published results for the kinetics of oxidation of
PPD and DEPD have been included for completeness.
[20]
In the case of THxPD, THpPD, and TOPD, initial HSChE ex-
periments appeared to indicate some adsorption effects, as
well as having k
0
values apparently approaching the lower
limit of reliability for the HSChE analysis method.
[46]
It was
therefore decided that quantitative kinetic measurements were
Table 2. Diffusion coefficients and kinetic parameters obtained for the oxidation of TMPD in alkyl cyanides.
Solvent D10
5
[cm
2
s
1
] k
0
[cm
1
s
1
] a E
f
0
[V
1
(vs. Ag)]
MeCN 2.200.15 0.550.08 0.500.01 0.3740.002
EtCN 1.200.10 0.450.03 0.530.02 0.4360.025
PrCN 1.110.10 0.330.03 0.540.02 0.2430.003
BuCN 0.840.06 0.180.01 0.520.01 0.2040.002
Figure 4. Plot of lnk
0
vs. lnt
L
for TMPD in alkyl cyanides (gradient =0.92,
R
2
=0.944).
Table 3. Diffusion coefficients and kinetic parameters obtained for the oxidation of substituted 1,4-phenylenedi-
amines in MeCN. D
WC
is the WilkeChang estimate for the diffusion coefficient.
Compound D10
5
[cm
2
s
1
] D
WC
10
5
[cm
2
s
1
] k
0
[cm
1
s
1
] a E
f
0
[V
1
(vs. Ag)]
PPD
[20]
2.09 2.1 0.840.16 0.50 0.2870.002
DEPD
[20]
1.88 1.6 1.640.25 0.50 0.2080.003
TMPD 2.200.35 1.6 0.550.08 0.500.01 0.3740.002
TBPD 1.240.12 1.0 0.300.05 0.440.05 0.1630.007
THxPD 0.820.08 0.8 0.0080.001 0.400.02 0.2240.005
THpPD 0.720.07 0.7 0.0140.003 0.330.07 0.1810.007
TOPD 0.770.08 0.7 0.0070.001 0.310.02 0.2320.010
DEDB 1.050.12 1.1 0.360.03 0.500.01 0.2090.009
DEDHx 0.960.10 1.0 0.0370.006 0.400.06 0.2440.017
DPPD 1.700.10 1.3 1.100.15 0.530.03 0.4070.008
1238 2004 Wiley-VCH Verlag GmbH&Co. KGaA, Weinheim www.chemphyschem.org ChemPhysChem 2004, 5, 1234 1240
best obtained using steady-state voltammetry at a microdisk
electrode, due to the ease of mechanically cleaning the elec-
trode surface. Solutions of 1.82 mm THxPD, 0.92 mm THpPD
and 0.85 mm TOPD were investigated in this way, with careful
polishing of the electrode between voltage scans. In each case,
several steady-state voltammograms were recorded and ana-
lysed according to the method of Mirkin and Bard
[47]
to extract
values for k
0
and a. The results are also presented in Table 3.
These results can be interpreted by mean of Equation (7),
and a graph of this function is shown in Figure 5, yielding a
linear plot with a gradient of 86.7 and R
2
=0.970. From this
plot, values of Q and B can be extracted using the known
value for y of 37.3 . This gives B=2.31
1
and Q=1.79
10
8
cms
1
. These values compare reasonably with other values
for B of 1.37
1[25]
and within the range 12
1
.
[7]
These
values can then be put into Equation (5) to visualise the direct
relationship between k
0
and r, as shown in Figure 6, showing a
very good correlation between the values determined for k
0
and Equation (5). This justifies the use of the hydrodynamic
radius in this context, as well as providing a rigorous test of
Marcus theory over a range of k
0
that spans more than two
orders of magnitude.
In using Equations (5) and (7), we assume that the contribu-
tion to the activation energy due to inner-sphere reorganisa-
tion is not significant. The basis of this assumption is that l
i
has been calculated to be no more than 10% of the value of
l
o
for PPD and TMPD,
[27]
and would be expected to be less for
the larger TRPDs. In choosing the hydrodynamic radius as a
measure of molecular size, it is noted that Kapturkiewicz has
demonstrated the usefulness of the ellipsoidal model, using an
ellipsoidal radius calculated from the lengths of the molecu-
lar semi-axes in the crystallographic literature.
[27]
Due to the ab-
sence of published structural data for most of the compounds
in the present study (including those which are liquid at ambi-
ent temperatures), a thorough comparison of ellipsoidal and
hydrodynamic radii is not appropriate. However, for the com-
pounds PPD and TMPD the calculated ellipsoidal radii are
0.28 nm and 0.40 nm respectively,
[27]
which compare favourably
with those determined for the hydrodynamic radii (under the
slip limit) of 0.29 nm and 0.37 nm.
Conclusions
Hydrodynamic and stationary electrodes have been used to
measure the heterogeneous electron-transfer rates for the 1,4-
phenylenediamines under steady-state conditions, thereby
avoiding the problems of charging currents and iR losses which
are associated with alternative conventional methodology
based on transient measurements such as cyclic voltammetry.
The dependence of k
0
on the hydrodynamic radius (r) has been
investigated over the significantly wide range 3 _r _9 and
the experimental results found to be in excellent agreement
with the values determined for k
0
, which range from 710
3
to
1.64 cms
1
, and Equation (7). This justifies the use of the hydro-
dynamic radius in this context, as well as providing a rigorous
test of Marcus theory. The dependence of k
0
on the solvent lon-
gitudinal dielectric relaxation (t
L
) has been investigated for
TMPD and a minor departure from the theoretical t
L
1
relation-
ship in Equation (2) was found, and these have been ascribed
to the solvation effects of the amino groups.
Acknowledgements
We thank the Clarendon Fund for partial funding for O.V.K. , and
both the EPSRC for a studentship and Avecia Ltd for CASE sup-
port for N.V.R.
Keywords: electrochemistry electron transfer kinetics
Marcus theory oxidations phenylenediamines
[1] R. A. Marcus, J. Chem. Phys. 1956, 24, 966.
[2] R. A. Marcus, J. Chem. Phys. 1956, 24, 979.
Figure 5. Best fit plot of [ln(K

y
_
) +1/y] vs. y for the 1,4-phenylenediamines in
MeCN with gradient =86.46 and intercept =17.74 (R
2
=0.970). Concentrations
are as given in the text and the compounds are as follows: a) PPD, b) DEPD, c)
TMPD, d) TBPD, e) THxPD, f) THpPD, g) TOPD, h) DEDB, i) DEDHx, and j) DPPD.
Figure 6. Plot of k
0
vs. r for 1,4-phenylenediamines in MeCN using the values
B=2.3 and Q=1.7910
8
cms
1
derived above from the linear plot in
Figure 5 for the optimum fit of the data. Concentrations are as given in the
text and the compounds are as follows: a) PPD, b) DEPD, c) TMPD, d) TBPD, e)
THxPD, f) THpPD, g) TOPD, h) DEDB, i) DEDHx, and j) DPPD.
ChemPhysChem 2004, 5, 1234 1240 www.chemphyschem.org 2004 Wiley-VCH Verlag GmbH&Co. KGaA, Weinheim 1239
[3] R. A. Marcus, J. Chem. Phys. 1957, 26, 867.
[4] R. A. Marcus, J. Phys. Chem. 1963, 67, 853.
[5] R. A. Marcus, J. Chem. Phys. 1965, 43, 679.
[6] R. A. Marcus, Int. J. Chem. Kinetics 1981, 13, 865.
[7] M. J. Weaver, in Comprehensive Chemical Kinetics, Vol. 27 (Ed. : R. G.
Compton), Elsevier, 1987, pp. 1.
[8] M. Opallo, J. Chem. Soc. Faraday Transactions 1 1986, 82, 339.
[9] N. Sutin, Prog. Inorg. Chem. 1983, 30, 441.
[10] R. R. Dogonadze, A. M. Kuznetsov, Prog. Surf. Sci. 1975, 6, 1.
[11] D. F. Calef, P. G. Wolynes, J. Phys. Chem. 1983, 87, 3387.
[12] J. M. Hale, in Reactions of Molecules at Electrodes (Ed. : N. S. Hush), Wiley,
London, 1971.
[13] C. Amatore, C. Lefrou, F. Pfluger, J. Electroanal. Chem. 1989, 270, 43.
[14] C. Amatore, E. Maisonhaute, G. Simonneau, J. Electroanal. Chem. 2000,
486, 141.
[15] C. Amatore, E. Maisonhaute, G. Simonneau, Electrochem. Commun.
2000, 2, 81.
[16] C. P. Andrieux, P. Hapiot, J. M. Savant, Chem. Rev. 1990, 90, 723.
[17] R. F. Forster, in Encyclopedia of Analytical Chemistry, Vol. 11 (Ed. : R. A.
Meyers), Wiley, Chichester, 2000, pp. 10142.
[18] N. V. Rees, J. A. Alden, R. A. W. Dryfe, B. A. Coles, R. G. Compton, J. Phys.
Chem. 1995, 99, 14813.
[19] N. V. Rees, R. A. W. Dryfe, J. A. Cooper, B. A. Coles, R. G. Compton, S. G.
Davies, T. D. McCarthy, J. Phys. Chem. 1995, 99, 7096.
[20] N. V. Rees, O. V. Klymenko, B. A. Coles, R. G. Compton, J. Electroanal.
Chem. 2002, 534, 151.
[21] B. A. Coles, R. A. W. Dryfe, N. V. Rees, R. G. Compton, S. G. Davies, T. D.
McCarthy, J. Electroanal. Chem. 1996, 411, 121.
[22] A. S. Baranski, K. Winkler, J. Electroanal. Chem. 1998, 453, 29.
[23] K. Winkler, A. S. Baranski, W. R. Fawcett, J. Chem. Soc. Faraday Trans.
1996, 92, 3899.
[24] M. Rosvall, Electrochem. Commun. 2000, 2, 791.
[25] A. D. Clegg, N. V. Rees, O. V. Klymenko, B. A. Coles, R. G. Compton, J.
Am. Chem. Soc. 2004, 126, 6185.
[26] G. Grampp, W. Jaenicke, Ber. Bunsenges. Phys. Chem. 1984, 88, 325.
[27] A. Kapturkiewicz, W. Jaenicke, J. Chem. Soc. Faraday Trans. 1 1987, 83,
2727.
[28] J. O'M. Bockris, A. K. N. Reddy, Modern Electrochemistry 1: Ionics, Vol. 1,
2nd ed. , Plenum Press, New York, 1998.
[29] J. D. Wadhawan, R. G. Evans, C. E. Banks, S. J. Wilkins, R. R. France, N. J.
Oldham, A. J. Fairbanks, B. Wood, D. J. Walton, U. Schroeder, R. G.
Compton, J. Phys. Chem. B 2002, 106, 9619.
[30] F. Marken, R. D. Webster, S. D. Bull, S. G. Davies, J. Electroanal. Chem.
1997, 437, 209.
[31] J.-H. Fuhrop, H. Bartsch, Liebigs Ann. Chem. 1983, 802.
[32] C. M. A. Brett, A. M. C. F. Oliveira Brett, in Comprehensive Chemical Kinet-
ics, Vol. 26 (Eds. : C. H. Bamford, R. G. Compton), Elsevier, Amsterdam,
1986, p. 355.
[33] B. A. Brookes, N. S. Lawrence, R. G. Compton, J. Phys. Chem. B 2000, 104,
11258.
[34] W. J. Blaedel, L. N. Klatt, Anal. Chem. 1966, 38, 879.
[35] L. N. Klatt, W. J. Blaedel, Anal. Chem. 1967, 39, 1065.
[36] V. G. Levich, Physicochemical Hydrodynamics, Prentice Hall, Englewood
Cliffs, New Jersey, 1962.
[37] B. A. Brookes, T. J. Davies, A. C. Fisher, R. G. Evans, S. J. Wilkins, K. Yunus,
J. D. Wadhawan, R. G. Compton, J. Phys. Chem. B 2003, 107, 1616.
[38] Krishnaji, A. Mansingh, J. Chem. Phys. 1964, 41, 827.
[39] H. Fernandez, M. A. Zon, J. Electroanal. Chem. 1992, 332, 237.
[40] H. Fernandez, M. A. Zon, J. Electroanal. Chem. 1990, 283, 251.
[41] M. B. Moressi, M. A. Zon, H. Fernandez, Electrochim. Acta 2000, 45, 1669.
[42] W. R. Fawcett, M. Opallo, J. Phys. Chem. 1992, 96, 2920.
[43] W. R. Fawcett, C. A. Foss Jr. , J. Electroanal. Chem. 1989, 270, 103.
[44] W. R. Fawcett, Electrochim. Acta 1997, 42, 833.
[45] C. R. Wilke, P. Chang, Am. Inst. Chem. Eng. J. 1955, 1, 264.
[46] N. V. Rees, O. V. Klymenko, B. A. Coles, R. G. Compton, J. Phys. Chem. B
2003, 107, 13649.
[47] M. V. Mirkin, A. J. Bard, Anal. Chem. 1992, 64, 2293.
Received: March 25, 2004
Revised: May 18, 2004
1240 2004 Wiley-VCH Verlag GmbH&Co. KGaA, Weinheim www.chemphyschem.org ChemPhysChem 2004, 5, 1234 1240

You might also like