Volume 2 Interfacial Kinetics and Mass Transport
Volume 2 Interfacial Kinetics and Mass Transport
Volume 2 Interfacial Kinetics and Mass Transport
1.1
Ernesto J. Calvo
Universidad de Buenos Aires, Buenos Aires,
Argentina
1.1.2
1.1.1
Scope
Introduction
The currentpotential relationship at electrodes is set by a number of complex physical and chemical phenomena, depending on the experimental
conditions.
Electrode reactions are heterogeneous
chemical processes that may involve one
or more electron-transfer steps across the
electrochemical double layer [1, 2]. Electrode reactions provide a switch for charge
to ow between phases of different type of
electrical conductivity: electrodes and electrolyte [3]. Therefore, their response can
be analyzed either on the basis of electrical
or chemical models. The distinctive feature of reactions at electrodes is the strong
dependence of both the surface concentrations and the kinetics on the electrode
potential [410].
The following reactions at electried
interfaces are examples of electrode reactions:
Feaq 2+ Feaq 3+ + e(metal)
Cu Cuaq
2+
(1)
+ 2e(metal) (2)
Fundamentals
Xaq Xorg
(3)
dN
dt
(5)
(6)
Hg
2.0
Sn
1.8
Bi
Au
1.6
Cu
Ni
1.4
1.2
1.0
0.0
0.00
0.05
0.10
0.15
log I
1.1.3
(7)
O + ne
(8)
O+
n
H2 = R + nH+
2
(9)
is given by
G = nF E = RT ln Keq
(10)
(11)
ln
Ee = E +
nF
aR
Fundamentals
O+e
(12)
kb
(13)
=
vf = kf CO
ic
F
(14)
and,
vb = kb CR =
ia
F
(15)
i
F
(16)
=
Gi
kB T
(17)
exp
ki =
h
RT
where Gi is the free energy of activation for the reaction, kB the Boltzmann
constant, h the Plank constant and the
transmission coefcient which for simplicity we shall consider here equal to one.
In the case of electrode reactions that occur at an electried interface and therefore
under the inuence of a strong electrical eld (c.a. 107 V cm1 ), the driving
force for the reaction of charged species
=
Gi has a term of electrostatic work
wel = F E as shown in the scheme of
Fig. 2. The negative sign corresponds to
the negatively charged electron crossing
the electrical double layer.
B
A
G f (E)
FE
aFE
(1 a)FE
A
G b (E)
Gf (Ee)
FE
O + e
(18)
and,
=
Gb (E)
It is convenient to dene
=
Gf (Ee )
kB T
kf ,o =
exp
h
RT
and
kb,o
=
Gb (Ee ) (1 )F E
(19)
where 0 0 is the fraction of the
applied potential that drives the reaction
and a measure of the symmetry of the
energy barrier.
Replacing in Eq. (17):
=
Gf (Ee )
kB T
kf =
exp
h
RT
F E
exp
(20)
RT
=
Gb (Ee )
kB T
exp
kb =
h
RT
(1 )F E
exp
(21)
RT
(22)
=
Gb (Ee )
kB T
=
exp
h
RT
And therefore,
kf (E) = kf ,o exp
F E
RT
(23)
and
(1 )F E
kb (E) = kb,o exp
RT
(24)
i = F kf ,o CO exp
kb,o CR
RT
(1 )F E
exp
(25)
RT
which has an exponential dependence on
the electrode potential, E, as observed in
experiments.
Fundamentals
ko = kf ,o exp F
RT
E
= kb,o exp (1 )F
(28)
RT
1
CR . ko is the
and thus, io = F ko CO
standard rate constant for unit surface
concentration of reactants at the standard
equilibrium potential.
Substituting in Eq. (25) and introducing
the overpotential for the reaction as
= (E Ee ) and substituting io into
i
+ log io
log
1 exp
RT
(30)
log ( j/jo)
0
1
2
150
100
50
50
100
150
E Eo
Fig. 3
The Tafel relationship is a linear freeenergy relation of the rate coefcient for a
net electrode reaction (neglecting the back
reaction) and the transfer coefcient:
=
1 Gf
(31)
=
F
E
Ee
(32)
RT 1
nF io
(33)
10
Fundamentals
Rudolf Marcus was Chemistry Nobel laureate in 1992 for his work on the theory
of homogeneous and heterogeneous electron transfer (ET) in such diverse processes
as the bioenergetic of photosynthesis and
respiration, homogeneous redox and heterogeneous electrode reactions, corrosion,
color photography, and so on. Marcus theory provides an insight on the physical
and chemical phenomena of the heavy
particle subsystem during electron transfer in redox reactions. Furthermore, it can
accommodate disparate rates over many
orders of magnitude for soluble redox
species, intramolecular electron transfer
in biological reactions of utmost importance, and a common framework for a
comparison of redox processes in solution
and at electrode surfaces.
In its simplest form the Marcus expression for the electron-transfer rate coefcient is given by [1520]:
G 2
1+
kET = Z exp
(34)
4RT
with the reorganization free energy
of the reactants to yield products and
G the standard free energy of the
reaction (thermodynamic driving force)
and Z a frequency factor taken by Marcus
as 104 cm1 for heterogeneous electrode
reactions [19]. The reorganization free
energy, comprises the solvation, o , and
(35)
op
s
The inner sphere component, i is
given by
i =
1
p
kj (qjr qj )2
2
(37)
=
Gi
F (E E ) 2
=
1+
4
(38)
since G = FE .
The Marcus free-energy relation predicts
a nonlinear dependence of the transfer
coefcient, on electrode potential:
1 F (E E )
1 Gi
= +
F E
2
2
w o wR
(39)
2
11
Fundamentals
(40)
Fe
S
S
S
S
S
AuAu Au Au Au Au Au AuAu
(a)
105
104
kf + k b
[s1]
12
103
102
101
100
1.0
(b)
0.5
0.0
E Eo
[V]
0.5
1.0
1.1.6
Ion-Transfer Reactions
log i
[a cm2]
Zn Hgx/Zn2+
i(calculated)
i (experimental)
6
i+ (from Zn55)
0.80
0.75
0.70 V
0.65
eh
13
14
Fundamentals
z+
(solvated)
+ ze(metal)
(41)
Ei
i+ = k+ cM cT exp
RT
= k+ cM cT exp
and
zF
E
RT
E
i
i = k co exp
RT
(42)
= k co exp
(1 )zF
E
RT
(43)
RT
(1 )zF
exp
(45)
RT
a typical ion-transfer electrode reaction
is the dissolution of zinc amalgam into
solvated Zn2+ cations in the aqueous
electrolyte:
Zn(Hg) Zn2+ (aq) + 2e(Hg) (46)
For this reaction a Tafel plot is shown in
Fig. 5 as predicted by Eq. (45). A kinetic
analysis of this data yields for the trans
fer coefcient: = RT /zF ln io /o =
0.72 for the charge-transfer valence z = 2.
The free-energy barrier for the ow of
ionic charge across the oxideelectrolyte
interface has an electrical contribution and
consequently the currentpotential curve
can be described by a ButlerVolmer type
equation [8].
The ion-transfer reactions at the oxideelectrolyte interface:
Mz+ (oxide) Mz+ (electrolyte) (47)
and,
O2 (oxide) + 2H+ (electrolyte)
H2 O(electrolyte)
(48)
1 F
RT
Cuaq 2+ + e Cuads + + e Cu
2F
1 exp
RT
io,1
(1 + 2 1 )F
1+
exp
io,2
RT
(51)
where = E E and E is the standard
electrode potential for the overall twoelectron process. In Fig. 7 we represent
the Tafel plot for the current potential
relationship described by Eq. (51).
We can consider two limiting cases:
(50)
Since the time scale for the electrontransfer event, ca. 1016 s is much smaller
than the time scale of the fastest chemical
k1
+ H+ (aq) + e H2
(49)
or,
k1
k2
15
Fundamentals
Standard free-energy diagram
for a two-step electron-transfer
electrode reaction with two potential
dependent transition states.
Fig. 6
=/ 1
E1
E*
=/ 2
E2
2
k1
k2
k2 = k1
k2
i0.2
Log i
16
k1
i0.1
i0
Eo2
E*E o
Eo1
aA + bB + . . . .ne = cC + dD + . . . .
(52)
which takes place in several consecutive
intermediate steps, the rate of one of
them controlling the overall kinetics.
The reactants and products in the ratedetermining step (rds) will be identied by
R and P , respectively.
We shall assume that the completion of
the overall process represented by Eq. (52)
requires the formation and decomposition
of identical activated complexes.
According to Parsons the rate-determining step can be represented [46] by
a
b
n
c
d
A + B + .... e = C + D + ....
(53)
The stoichiometric number indicates the
number of times the rate-determining
step occurs in the overall stoichiometric
reaction. For example, the Tafel mechanism for the reduction of H+ on metals
occurs in two steps with Reaction (54) occurring twice for each time Reaction (56)
takes place.
(54)
2 H+ + e Had
2Hads H2
(55)
(56)
1
H2
2
(57)
then, = 1 and n = 1.
It is worthwhile noticing that, while
and n are arbitrary quantities that depend
on the way we write the stoichiometric
equation, the electron number n/ is
characteristic of the electrode reaction
kinetics. In the Tafel hydrogen reduction
mechanism, n/ = 1 and indicates the
17
18
Fundamentals
=
of activation, Gi , for charged particles.
Gf = R
(58)
=
Gb
(59)
= P
n n
+ e + F M
c
d
n
P = C + D + F S
R =
nF E
(64)
=
where Gi represents the nonelectrical
contribution to the standard free energy of
activation and E = (M S ) + const. It
should be noted that the right-hand side of
Eqs. (63 and 64) should be divided by the
activity coefcient of the activated complex,
which could be potential dependent [46].
Introducing the overpotential, = E
Ee , and replacing
nF
i = io exp
RT
(1 )n
(65)
exp
(60)
(61)
=
Gb
(1 ) nF E
kB T
kf =
exp
+
h
RT
RT
(62)
=
Gf
kB T
nF E
(63)
exp
+
kb =
h
RT
RT
io =
1
1
n
F ko [A] [B] [C] [D]
(66)
RT
=
ln i T,P,ci
nF
and
ln i
=
T,P,ci
RT
(1 )nF
(67)
=
RT
E a
E c
(68)
if the rate-determining step is the same
over the whole range of potentials from
which Tafel slopes are obtained. Note
that from a Tafel plot we can only
determine the product n/, not the
quantities separately.
+ ci
Ji = Di ci + ci
RT
(70)
which accounts for the ux of species i
(mol cm2 s1 ) at a given distance from
the electrode surface, Di is the diffusion
coefcient of the diffusing electroactive
species (cm2 s1 ). The rst term on the
right hand side of Eq. (70) represents the
transport mechanism by diffusion in a
uid velocity
(cm s1 ) and the particles
(71)
(73)
2 c
D
r r
(74)
19
20
Fundamentals
or,
Xcylindrical =
1 c
D
r r
(75)
r (2Dt).
Two other sources of a term X that
can compensate for the curvature of the
concentration gradient at the electrode
,
surface are a convective term X = ci
AH
A +H
(76)
k1
(77)
D
(nF DAci )
=
Kd,i
i
(81)
C0/C*0
i/iL
ButlerVolmer
id,a
C0/C*0
C0/C*0
id, c
X
1
X
Plot of i/iL versus calculated with Eq. (83) (solid line) and
ButlerVolmer currentpotential curve calculated with Eq. (29) for
= 0.5 and n = 1. Inset: concentration proles of the oxidized species
normalized to the analytical concentration at limiting current and at
equilibrium.
Fig. 8
predicted currentpotential curve calculated with Eq. (29) (dashed line) is shown
for comparison.
We can reexamine Eq. (83), both in the
Tafel region, for || RT /nF :
i
nF
log i = log io + 1
exp
iL,c
RT
(84)
and for the linear polarization region, for
|| RT/nF:
nF
1
1
1
RT 1
+
+
io
iL,c
iL,a
(85)
which in the absence of concentration
polarization is coincident with Eq. (32).
Expressing Eq. (84) in terms of rate
coefcient rather than current densities
and rearranging terms, the rate coefcient obtained experimentally at constant
i=
potential, kobs :
kobs =
i
nF ci
(86)
(87)
21
22
Fundamentals
4rAB DAB
4rAB DAB
1+
k(hom)
(88)
dened by
pk =
log ij
log ck
(90)
ci
=k ,E
(91)
at pH < 8 (93)
and
H2 O + e Hads + HO
at pH > 9
(94)
1.1.10
(95)
kapp = kM =2 CO2
nF (M 2 )
(97)
exp
RT
By combining Eqs. (96 and 97) and replacing the concentration of O at the plane of
reaction, the apparent rate constant can be
expressed in terms of the applied electrode
potential.
nF (M S )
S
kapp = kM =2 CO
exp
RT
(n z)F (2 S )
exp
RT
(98)
The rst exponential term shows the
potential dependence of the apparent
23
24
Fundamentals
ln iapp
2
F
=
(n zo )
E
RT
E
(101)
PO
=
ln iapp
ln cO
= PO
E
+ (n zo )
F
RT
2
ln CO
E
(102)
the double-layer correction in the second
term of Eq. (102) due to interfacial potential distribution is zero at the pzc.
1.1.11
Mixed Potential
|Ia,i (E)|
|Ic,i (E)| = 0
(103)
The additive individual component currents are based on the simplifying assumption that the anodic and cathodic processes
are statistically independent and that the
electrode surface sites are indistinguishable for both reactions. Since the total
anodic and cathodic currents equal at the
25
26
Fundamentals
O R + n e
O R + n e
ja
Ee1
Ee2
EM
jc
R + n e O
R + n e O
M MZ+ + z e
M MZ+ + z e
jcorr
Eo2
Eo1
Ecorr
MZ+
+z
jcorr
Eo4
Eo3
H2O O2 + 4H+ + 4e
2H+ + 2e H2
O2 + 4e + 4H+ 2H2O
Fig. 10
27
28
Fundamentals
Electrocrystallization
that is,
B
i = A exp
n
(105)
5. J. OM. Bockris, K. N. Reddy, Modern Electrochemistry, Plenum Press, New York, 1970.
6. B. E. Conway, Theory and Principles of Electrode Processes, Ronald Press, New York, 1964.
7. P. Delahay, Double Layer and Electrode Kinetics, 2nd edn., Interscience Publishers, New
York, 1966.
8. K. J. Vetter, Electrochemical Kinetics (English
Translation), Academic Press, New York,
1967.
9. C. M. A. Brett, A. M. Oliveira Brett, Electrochemistry: Principles, Methods and Applications, Oxford University Press, Oxford, 1993.
10. E. J. Calvo in Comprehensive Chemical Kinetics (Eds.: C. H. Bamford, R. G. Compton),
Elsevier, Amsterdam, 1986, Vol. 26.
11. J. Tafel, Z. Phys. Chem. 1905, 50A, 641.
12. J. A. V. Butler, Trans. Faraday Soc. 1932, 28,
379.
13. J. A. V. Butler, Proc. R. Soc. London, Ser. A
1936, 157, 423.
14. T. Erdey-Gruz, M. Volmer, Z. Phys. Chem.
Abt. A 1931, 157, 165.
15. R. A. Marcus (Nobel Lecuture), Angew.
Chem., Int. Ed. Engl. 1993, 32, 1111.
16. R. A. Marcus, J. Chem. Phys. 1956, 24, 966.
17. R. A. Marcus, Discuss. Faraday Soc. 1960, 29,
21.
18. R. A. Marcus, J. Phys. Chem. 1963, 67, 853.
19. R. Marcus, J. Chem. Phys., 1965, 43679.
20. R. Marcus, J. Phys. Chern. 1968, 72, 891.
21. J. R. Miller, L. T. Calcaterra, G. L. Closs, J.
Am. Chem. Soc. 1994, 106, 3047.
22. A. L. Barker, J. V. Macpherson, C. J. Slevin
et al., J. Phys. Chem. B 1998, 102, 1586.
23. A. L. Barker, P. R. Unwin, S. Amemiya et al.,
J. Phys. Chem. B 1999, 103, 7260.
24. M. Tsionsky, A. J. Bard, M. V. Mirkin, J.
Phys. Chem. 1996, 100, 17 881.
25. M. Tsionsky, A. J. Bard, M. V. Mirkin, J. Am.
Chem. Soc. 1997, 119, 10 785.
26. C. J. Miller in Physical Electrochemistry, Principles, Methods and Applications (Ed.: I. Rubinstein), Marcel Dekker, New York, 1995,
Chap. 2.
27. H. O. Finklea, Electroanal. Chem. 1996, 19,
109.
28. C. E. D. Chidsey, J. Am. Chem. Soc. 1997,
119, 10 563.
29. S. Creager, S. J. Yu, D. Bamdad et al., J. Am.
Chem. Soc. 1999, 121, 1059.
30. R. Holmlin, R. F. Ismagilov, R. Haag et al.,
Angew. Chem., Int. Ed. 2001, 40, 2316.
29
30
Fundamentals
31. D. I. Gittins, D. Bethell, D. J. Schiffrin et al.,
Nature 2000, 408, 67.
32. X. D. Cui, A. Primak, X. Zarate, J. Tomfohr,
O. F. Sankey, A. L. Moore, T. A. Moore,
D. Gust, G. Harris, S. M. Lindsay, Science
2001, 294, 571.
33. F. F. Fan, J. Y. S. M. Dirk, D. W. Price et al.,
J. Am. Chem. Soc. 2001, 123, 2454.
34. C. E. D. Chidsey, Science 1991, 251, 919.
35. R. J. Forster, L. R. Faulkner, J. Am. Chem.
Soc. 1994, 116, 5444.
36. V. G. Levich in Physical Chemistry. An Advanced Treatise (Eds.: H. Eyring, D. Henderson, W. Jost), Academic Press, New York,
1970, pp. 9851074, Vol. IXB.
37. H. Gerischer, Adv. Electrochem. Electrochem.
Eng. 1961, 1, 139.
38. W. Lorenz, Z. Elektrochem. 1953, 57, 382.
39. H. Gerischer, Z. Elektrochem. 1953, 57, 604.
40. K. Heusler, Electrochim. Acta 1983, 28, 439.
41. E. J. M. O Sullivan, E. J. Calvo in Reactions at
Metal Oxide Electrodes Comprehensive Chemical Kinetics (Ed.: R. G. Compton), Elsevier,
New York, 1987.
42. J. Horiuti, M. Ikusima, Proc. Imp. Acad.
(Tokyo) 1939, 15, 39.
43. J. Horiuti, J. Res. Inst. Catal. Hokkaido Univ.
1948, 1, 8.
44. R. Parsons, Pure Appl. Chem. 1974, 37, 501.
45. R. Parsons, Pure Appl. Chem. 1979, 52, 233;
reprinted in Electrochim. Acta 1981, 26, 1869.
1.2
Introduction
red
ox + e (Me)
(1)
31
1 Fundamentals
ox + e
Red
Energy
32
Eact
G
qi
qf
Reaction coordinate
(2)
(3)
= 12 k(qi qf )2
(4)
Eact =
where
( + G)2
4kB T
(5)
()f ()
exp
( + e0 )2
d
4kB T
(6)
where we have used the fact that a negative overpotential favors a reduction.
Equation (6) has also been derived by
Gerischer [7, 8], but with a different
interpretation. He denes
Wox (, ) = (4kB T )(1/2)
exp
( + e0 )2
4kB T
(7)
33
1 Fundamentals
0
= 0.5 eV
ln (kred/klim)
34
= 0.75 eV
= 1 eV
12
0.0
0.4
0.8
1.2
[V]
Logarithm of the normalized rate constant kred /klim vs.
overpotential for the reduction of a redox couple at a metal electrode
according to Marcus theory (Eq. 9).
Fig. 2
1.6
1.2.3
Energy of Reorganization
As outlined above, the energy of reorganization plays a major role in the electrontransfer theory. Generally it contains two
contributions: one from the surrounding
solvent and the other arising from changes
in the bond lengths and vibration frequencies of the reacting complex itself. They are
also referred to as outer-sphere and innersphere reorganization, respectively.
The reorganization of the solvent can
be expressed through its polarization,
which also contains two contributions: one
from the electronic polarizability of the
solvent molecules, and the other from the
librational and vibrational motion. Only
the latter are slower than the electron
exchange as such, and contribute to the
solvent reorganization energy out . This
takes a form that is reminiscent of the
Born formula for the energy of solvation:
1
1
1
out =
20
s
(Dox Dred )2 dV (10)
Here, and s are the optical and the
static dielectric constants of the solution;
the former appears because the contribution from the electronic polarizability has
been subtracted. Dox and Dred are the dielectric displacements when the reactant is
in the reduced and in the oxidized form, respectively. The integral is to be performed
over the space lled by the solution. When
the reactant is close to the electrode surface, image terms arise, which contribute
to the displacement.
The energy of reorganization is a molecular concept, and an equation such as (10),
which is based on macroscopic electrostatics, can only be a rough approximation.
Several other expressions, on the basis
(11)
Quantum-mechanical Theory
A Model for Electron Exchange
Between a Metal and a Solvated Reactant
Before we set up the model Hamiltonian
for electrochemical electron transfer, we
have to specify the models for the various parts of the system. For the electrons
in the metal, we use the quasi-free electron model in which the electronic states
are labeled by their quasi-momentum k.
For outer-sphere electron transfer on metals, it is usually permissible to ignore the
spin index keeping it would introduce an
additional factor of two, which can be incorporated into the interaction constants.
On the reactant, we consider a single orbital, labeled a, with which the electrons
are exchanged.
An important aspect of electron transfer
is the accompanying reorganization of the
solvent. The latter can be modeled as a
phonon bath, or a collection of harmonic
oscillators. Other models in terms of the
1.2.4.1
35
36
1 Fundamentals
+ Vk ca+ ck ] +
na
1
2
h g q
h (p2 + q2 )
(12)
a na , where
a = a
h g q
(13)
|Vk
h
|2
(a k )
(14)
kr (k ) =
4kB T
h
exp
(a k )2
(15)
4kB T
1
h g2
2
(16)
kred = C
h
kB T
( + e0 )2
()f () exp
4kB T
(18)
which differs from the MarcusHush
expression in the preexponential factor,
which is here determined by the square
of the transition amplitude Vk . This is
characteristic for the nonadiabatic, weak
coupling case. Solvent dynamics, which
is so important for the adiabatic, strong
coupling case plays no role.
Quantum Modes
When quantum modes are reorganized
during the electron exchange, they have
to be treated separately from the classical
modes. We consider explicitly the case
in which one such mode is present;
the generalization to many modes is
straightforward, though it results in rather
cumbersome formulae.
The situation is simplest if the frequency
i of the quantum mode is so high that its
quantum of energy is much higher than
the thermal energy, hi kT , so that it
is not excited in the initial state. In this
case, we only have to consider transitions
from the ground state 0 to the nth excited
state. Let
1.2.5.2
37
38
1 Fundamentals
f
1
n!
i
hi
i
hi
n
(19)
kr (0 n) = M0n
kB T
h
exp
( + nh e0 )2
4kB T
=
|Vk |2 ( k )
(23)
(20)
2
[ (q )]2 + 2
(24)
1
a
arccot
(25)
ln(a2 + 2 )
2
(26)
1
a na 2n0
arccot
(27)
(29)
are fullled, where 2 = /2.
Within this formalism, it is in principle possible to calculate the potentialenergy surfaces for electron-transfer reactions [17]. The important system parameters are , and a ; they can be
estimated from quantum chemical calculations and computer simulations. At
present, such surfaces involve fairly rough
approximations; however, they are useful
visualizations for understanding the dynamics of electron-transfer reactions. It
is convenient to restrict the system to a
minimum set of coordinates using a generalized solvent coordinate that follows a
reaction path from one minimum of the
surface via a saddle point to the other
minimum. This coordinate can be normalized in such a way that its value at the
two minima corresponds to the charge on
the reactant in these congurations [18].
It, then, is identical to the reaction coordinate in the theory of Hush [2]. If quantum
modes are reorganized during the electron
transfer typically, these would be innersphere modes they have to be singled out
and treated separately.
Figure 3 shows an adiabatic potentialenergy surface for one single solvent
coordinate q. Two minima are separated by
a barrier. This barrier is the lower and the
atter the larger the electronic coupling
between the reactant and the metal. For
39
1 Fundamentals
15.0
12.5
10.0
Energy
[kB T ]
40
7.5
5.0
2.5
0.0
1
Solvent coordinate q
Adiabatic potential-energy curves for various values of the energy
broadening ; full line: = 0.01 eV; long dashes: = 0.05 eV; short dashes:
= 0.1 eV. The energy of reorganization was taken as 6 kB T.
Fig. 3
Solvent Dynamics
i
Eact
exp
2
kB T
(30)
where Eact is the energy of activation. However, as was pointed out by Kramers [20],
the transition-state theory overestimates
the rate since the population at the barrier
may be depleted, and also because a system that has passed the saddle point can
return, a phenomenon denoted as barrier
recrossing. These effects can be described
in terms of the solvent friction, which encompasses the inuence of all the solvent
modes that are not explicitly considered in
the one-dimensional model. As a result,
the preexponential factor is modied by an
( 2 /4 + b2 )1/2
/2
(31)
2b
(32)
1
,
l
where
l =
D (33)
s
Computer Simulations
We have already emphasized the important role of the solvent dynamics and structure in the kinetics of electron-transfer
reactions. During the last few years, a
number of classical molecular dynamic
(MD) simulations have been performed
to obtain free energy surfaces of the reaction [2127]. These simulations require
explicit interaction potentials between the
constituents of the system: the reactants,
the solvent, and the electrode. Again, a
generalized solvent coordinate is used to
41
1 Fundamentals
100
80
G
[kcal mol1]
42
60
40
20
Fe3+
0
4
Fe2+
(E E )
[eV]
Diabatic solvent reorganization free energy curves for Fe2+
and
[22]. The bold solid lines represent the molecular dynamics
result, the thin solid lines the best parabolic t of the region, near the
bottom of each well.
Fig. 4
Fe3+
30
G(E)
[kcal]
20
Classical
10
0
Quantum
10
250
300
350
400
450
500
E
[kcal]
Adiabatic free energy curves for the Fe2+ /Fe3+ couple from
molecular dynamics [25]. The solid line represents results for the
quantized high-frequency degrees of water, the dashed line for the
classical degrees of freedom only.
Fig. 5
43
1 Fundamentals
I(2)
F(E1, E2)
[kcal]
44
30
25
20
15
10
5
P
200
150
50
0
100
E1
[kcal]
I(1)
50
150
0
50
200
100
50
E2
[kcal]
Adiabatic free energy surface for the electron transfer between two different redox
couples and a metal electrode from molecular dynamic simulations [27].
Fig. 6
(34)
0.70
0.65
(a)
0.60
g = 1.98
Distance
0.72
0.70
0.68
0.66
(b)
0.64
g = 18.1
0.750
0.625
0.500
0.375
0.00
(c)
0.25
0.50
0.75
1.00
Time
45
46
1 Fundamentals
(35)
Conclusion
F (E, r )
[kcal mole1]
100
50
0
50
0
50
100
(a)
r
[Angstroms]
F (E, r )
[kcal mole1]
100
50
0
50
0
50
100
E
150
[kcal mole1]
200
250
300
(b)
r
[Angstroms]
Fig. 8
47
48
1 Fundamentals
1.3 Electronic Tunnel Factors in Molecular Electron Transfer and Molecular Conduction
1.3
1.3.1
Introduction
Abraham Nitzan
Tel Aviv University, Tel-Aviv, Israel
Abstract
49
50
1 Fundamentals
B
D
(a)
(b)
(c)
1.3 Electronic Tunnel Factors in Molecular Electron Transfer and Molecular Conduction
{n}
(a)
{l}
{n}
{r}
(b)
51
52
1 Fundamentals
2
|VDA |2 F
h
(1)
D
Pth (D (D ))
4kB
2
F(EAD ) =
(3)
f (k )F(k e)|VkA |2
df ()F( e)
k
( k )|VkA |2
(4)
1.3 Electronic Tunnel Factors in Molecular Electron Transfer and Molecular Conduction
(a)
E
N+1
=A
0=D
(b)
Fig. 3 Simple level structure models (a) for molecular electron transfer and (b) for electron
transmission. The molecular bridge is represented by a simple set of levels that represent local
orbitals of appropriately chosen bridge sites. This set of levels is coupled to the donor and
acceptor species (with their corresponding nuclear environments) in (a), and to electronic
continua representing metal leads (say) in (b). In the latter case, the physical meaning of states 0
and N + 1 depends on the particular physical problem: They can denote donor and acceptor
states coupled to the continua of environmental states (hence the notation 0 = D, N + 1 = A),
surface localized states in an metal-molecule-metal junction, or they can belong to the right and
left scattering continua.
where
1
f () =
1 + e/kB
(5)
ket =
2
h
e(e+) /4kB
4kB
2
|V ()|2 f ()
(7)
53
54
1 Fundamentals
Much of the early work on electron transfer have used expressions such as (1 and
7) with the electronic coupling term VDA
being used as a tting parameter. More
recent work has focused on ways to characterize the dependence of this term on the
electronic structure of the donor/acceptor
pair and on the environment. In particular,
studies of bridge-mediated electron transfer, where the donor and acceptor species
are rigidly separated by molecular bridges
of well-dened structure and geometry,
have been very valuable for characterizing the interrelationship between structure
and functionality of the separating environment in electron transfer processes. As
expected for a tunneling process, the rate
is found to decrease exponentially with the
donor-acceptor distance
ket = k0 e RDA
(8)
where
is the range parameter that characterizes the distance dependence of the
electron transfer rate. The smallest values for
are found in highly conjugated
organic bridges for which
is in the
range of 0.2 to 0.6 A 1 [2132]. In contrast, for free space, taking a characteristic
ionization
barrier UB = 5 eV, we nd
=
2
8mUB /h 2.4 A 1 (m is the electron
mass). Lying between these two regimes
are many motifs, both synthetic and natural, including cytochromes and docked
proteins [3341], DNA [4250], and saturated organic molecules [5157]. Each
displays its own characteristic range of
values, and hence its own timescales and
distance dependencies of electron transfer. A direct measurement of
along
a single molecular chain was recently
demonstrated [58].
In addition, to bridge assisted transfer
between donor and acceptor species, electron transfer has been studied in systems
1.3 Electronic Tunnel Factors in Molecular Electron Transfer and Molecular Conduction
acceptor is given by
TDA (E) = VDA + VD1 G1N (E)VN A
(9)
N
1
1
Vn,n+1
E EN
E En
(10)
n=1
2
h
V1D VN A 2 VB 2N
F
%E
V
B
B
(11)
where %EB = EB E. Similarly, for
a bridge-assisted transfer between a
molecule and an electrode, Eq. (7) applies
with |V ()|2 given by
2
|V ()| =
VB
%EB
2N
( k )
V1k VN A 2
VB
(12)
55
56
1 Fundamentals
where
(Ex ) =
2m2 e
(2 h)3
dy
f (E + e)] =
dz [f (E)
4me
(2 h)3
[f (E) f (E + e)]
dEr
0
(16)
(18)
but mainly from the voltage dependence of
T. The simplest model for a metal-vacuummetal barrier between identical electrodes
without an external eld is a rectangular
barrier of height above the Fermi energy
given by the metal work function. When a
uniform electric eld is imposed between
the two metals, a linear potential drop
from EF on one electrode to EF e on
the other is often assumed (see Fig. 4). In
addition, the image potential experienced
by the electron between the two metals will
considerably modify the potential barrier.
For a point charge e, located at position
x between two conducting parallel plates
that are a distance d apart, the image potential is
e2
VI =
4
nd
1
1
+
2x
[(nd)2 x 2 ] nd
n=1
(19)
1.3 Electronic Tunnel Factors in Molecular Electron Transfer and Molecular Conduction
EF
EF
EF
EF
Fig. 4
e2 ln 2
2d
(20)
This negative contribution to the electrons energy reduces the potential barrier
(Fig. 4), and has been invoked to explain
the lower than expected barrier observed
in STM experiments [90, 91]. Some points
should however, be kept in mind. First,
the classical result (19) fails close to
the metal surface where quantum mechanical and atomic size effects change
both the position of the reference image plane and the functional form of the
image potential [9296]. Second, consideration of the dynamic nature of the image
response should be part of a complete theory [97100]. The timescale of electronic
response of metals can be roughly estimated from the plasma frequency to be
1016 s. This should be compared to the
time during which a tunneling particle can
respond to interactions localized in the barrier. For transmitted particles, this is the
traversal time for tunneling [101, 102] that,
for an electron traversing a 10 A wide 1 eV
barrier is of the order of 1 fs. This comparison would justify the use of the static
image approximation in this context, but
this approximation becomes questionable
for deeper tunneling or narrower barriers.
l r l r ) (22)
dS(
57
58
1 Fundamentals
I
4e2
|Mlr |2 (El EF )
=
h
l,r
(Er EF )
(23)
1.3.4
e2
T (EF )
h
(25)
kE
0
dk(k)n(k) = e
e
f (Ek )
=
E
0
kE
dk
0
dE f (E )
hk
m
(26)
1.3 Electronic Tunnel Factors in Molecular Electron Transfer and Molecular Conduction
e2
h
(27)
e2
T (EF )
h
e2
Tij (E)
h
e2
e2
Tij (EF ) =
Tr(SS )EF
h
h
i,j
(29)
where Tij = |Sij |2 is the probability that
a carrier coming from the left (say) of
the scatterer in transversal mode i will be
transmitted to the right into transversal
mode j (Sij , an element of the S matrix, is
the corresponding amplitude). The sum
in (29) is over all transversal modes
whose energy is smaller than EF . The
analog of Eq. (29) for the microcanonical
chemical reaction rate was rst written by
Miller [109]. More generally, the current
for a voltage difference between the
electrodes is given by
g(E)
dE[f (E) f (E + e)]
I=
e
0
(30)
(31)
i,j
dkz T E
(28)
g(E) =
Ly Lz 2m
(2)2 h2
0
h2
2
2
(ky + kz )
2m
dEr T (E Er )
(32)
59
60
1 Fundamentals
Tij (E) = 4 2
i,j
|Tlr |2
=
l,r
g(E)
e
(36)
where
g(E)
4e2
|Tlr |2
h l,r
(E El )(E Er ) (37)
l,r
(E El )(E Er )
4
[f (El )(1 f (Er + e))
h
l,r
(38)
Molecular Conduction
l,r
g = g(EF )
(34)
(35)
(39)
1.3 Electronic Tunnel Factors in Molecular Electron Transfer and Molecular Conduction
VB =
(L)
(R)
#1 (EF )#N (EF )
(40)
e
h
EF
(L)
EF e
(R)
#N (E
+ e)
(41)
with
(L)
#1 (E) = 2
(R)
#N (E)
= 2
The Greens function in Eq. (40) is itself reduced to the bridges subspace
by projecting out the metals degrees of
freedom. This results in a renormalization of the bridge Hamiltonian: in the
bridge subspace
(E HB ;B (E))1
(43)
N
n=1
En |nn|;
(44)
PP
2
dE
#n (E
)
(E E
)
(47)
(E H + i)1
Vn,n |nn |
n=1 n =1
<n (E) =
i,j
N
N
V1,2
(E E1 ;1 (E))(E EN ;N (E))
N1
j =2
Vj,j +1
E Ej
(48)
61
62
1 Fundamentals
electron transfer (e.g. [113, 114] and references therein). For applications of variants
of this formalism to electron transport in
specic systems, see Refs. [86, 87, 115,
116]. In the following text, we discuss more
general forms of this formulation.
1.3.6
|N 1
|VB
(EB E)N
(50)
GDA (E)
=
VD1 VN A
(E ED ;D (E))(E EA ;A (E))
G1N (E)
(52)
1.3 Electronic Tunnel Factors in Molecular Electron Transfer and Molecular Conduction
16e2
|VD1 VN A |2
h # (L) (EF )# (R) (EF )
D
|G1N (EF )| ;
EF = ED = EA (53)
e2 kDA
8h
h F # (L) # (R)
D
(54)
1.3.7
H(E) = EZ H
63
64
1 Fundamentals
(56)
(57)
e2
T r G(M) (E)# (R) (E)
h
(59)
G(M) (E)# (L) (E)
1.3 Electronic Tunnel Factors in Molecular Electron Transfer and Molecular Conduction
(61)
which is related to the conduction by
g(E) = (4e2 /h)s(E) (c.f Eq. 37) can be
represented by (c.f. Eq. 59)
s(E) =
1
T r G(M) (E)# (R) (E)
4 2
(62)
G(M) (E)# (L) (E)
65
66
1 Fundamentals
1
l|V G(M) # (R) G(M) V |l (65)
2
1
= l|L GABC (E)R GABC (E)L |l
(66)
1.3 Electronic Tunnel Factors in Molecular Electron Transfer and Molecular Conduction
The second part of Eq. (66) is obtained by using the identity R |l =
0 to write R G V |l = R (1 + G V )|l =
R G (G1 + V )|l, which, together with
G1 = E H0 V + i, (E H0 )|l = 0
and |l = l |l, yields the desired result.
The results (63 and 66) are very useful
for computations of transmission probabilities in models in which the interaction
between the transmitted particle and the
molecular spacer is given as a position dependent pseudopotential. Applications to
electron transmission through water and
other molecular layers are discussed in
Refs. [151158].
1.3.9
Density functional methods provide a convenient framework for treating metallic interfaces [100]. Applications of this methodology to the problem of electron transport
through atomic and molecular bridges
have been advanced by several workers. In
particular, Langs approach [90, 159165]
is based on the density functional formalism [166, 167] in which the single electron
wave functions 0 (r) and the electron density n0 (r) for two bare metal (jellium)
electrodes is computed, then used in the
LippmanSchwinger equation
(r) = 0 (r) + dr
dr
G0 (r, r
)V
(r
, r
)(r
)
(67)
Greens function of the bare electrode system and V is the difference between
the potential of the full system containing an atomic or a molecular spacer and
that of the bare electrodes. In atomic units
(|e|, , m = 1), it is
V (r, r ) = Vps (r, r ) + Vxc (n(r)) Vxc
n(r
)
n0 (r) + dr
|r r
|
(68)
J (r) = 2
dE d2 K I m{+
+ }
L
(69)
The factor 2 accounts for the double occupancy of each orbital. This approach
was used recently [169] to calculate current through a molecular species, Benzene
1,4-dithiolate molecule (as used in the experiment described in Ref. [67]), between
two jellium surfaces. The result demonstrates the large sensitivity of the computed
current to the microscopic structure of the
moleculemetal contacts.
67
68
1 Fundamentals
(70)
To end this brief overview of densityfunctional-based computations of molecular conduction, we should note that this
approach suffers in principle from problems similar to those encountered in using
the HF approximation, that is, the inherent inaccuracy of the computed LUMO
energy and wave functions. The errors
are different, for example HF overestimates the HOMOLUMO gap (since the
HF LUMO energy is too high [143146,
174176]) while density functional theory
(DFT) underestimates it [167, 177]. Common to both approaches is the observation
that processes dominated by the HOMO
level will be described considerably better
by these approaches than by processes controlled by coupling to the LUMO [137, 178].
1.3.10
Potential Proles
1.3 Electronic Tunnel Factors in Molecular Electron Transfer and Molecular Conduction
Rectication
D
mR
A
B
69
70
1 Fundamentals
Carriercarrier Interactions
EF + e2/(2R)
EF
EF
EF e2/(2R)
Fig. 6
1.3 Electronic Tunnel Factors in Molecular Electron Transfer and Molecular Conduction
Hubbard model with [199201]or without [202] the mean eld approximation.
In particular, Malysheva and Onipko have
derived a tight binding analog of the
model for negative differential resistance
originally proposed by Davydov and Ermakov [203] (see also Refs. [204206]).
Numerical simulations [207] can assist in
gauging the performance of the mean eld
approximations used in these calculations.
Such models may be relevant to the understanding of recent experimental observation of negative differential resistance in
a metal-self-assembled monolayermetal
junction with the SAM containing a nitroamine redox center [208].
We conclude this discussion by emphasizing again that understanding correlated
carrier transport in molecular junctions
continues to be an important experimental and theoretical challenge. Recent work
by Gurvitz et al [209211], using exactly
solvable models of electron transport in
two and three barrier structures, has indicated that new phenomenology may arise
from the interplay of inelastic transitions
and intercarrier interactions in the barrier.
In fact, dephasing transitions in the barrier
may prove instrumental in explaining the
charge quantization that gives rise to the
single electron transport behavior of such
junctions (Sect. 6.3 in Ref. [212]).
1.3.13
This section discusses some subtle difculties that are glossed over in most of
the treatments of electron transmission using the formalisms described above. These
should be regarded as open theoretical issues that should be addressed in future developments. The source of these problems
is our simplied treatment of what is actually a complex many-body open system.
71
72
1 Fundamentals
have a choice: either imposing the reaction eld on the electronic Hamiltonian
in the position representation, thus modifying all Coulomb interaction terms, then
calculate the electronic wave functions under the new potential, or compute the
electronic wave functions with the original
Hamiltonian under the imposed dielectric
boundary conditions. The fact that the two
representations are not equivalent is associated with the approximate nature of
the approach, which replaces a detailed
treatment of the electronic structure of the
solvent by its electronic dielectric response
and with the fact that the Schrodinger
equations derived from them are nonlinear in the electronic wave functions).
Examination of the energies and timescales involved suggests that assuming
instantaneous metal response to the electron position is more suitable in most
situations than taking instantaneous response to the charge distribution dened
by a molecular orbital, but the corresponding timescales are not different enough to
make this a denite statement.
A similar issue appears in attempts to
account for the electronic polarizability
of a solvent in treating fast electronic
processes involving solute molecules or
excess electrons in this solvent. For
example, in treating electron transmission
in MIM junctions, the potential barrier that
enters into expressions such as Eq. (14)
depends on the electronic structure of the
insulating spacer. For vacuum tunneling,
a rectangular barrier, whose height above
the metal Fermi energy is the metal work
function, modied by image interactions
as discussed above and in Sect. 1.3.3,
seems appropriate. For a dielectric spacer,
the barrier should be further modied
by the fast (electronic) dielectric response
of this spacer in the same way that it
is modied by the electronic response of
1.3 Electronic Tunnel Factors in Molecular Electron Transfer and Molecular Conduction
73
74
1 Fundamentals
3. A. M. Kuznetsov, J. Ulstrup, A. M. Kuzne,
T. Sov, Electron Transfer in Chemistry and
Biology: An Introduction to the Theory, John
Wiley & Sons, New York, 1998.
4. D. R. Lamb, Electrical Conduction Mechanisms in Thin Insulating Films, Methuen,
London, 1967.
5. R. J. Miller, G. McLendon, A. Nozik et al.,
Surface Electron Transfer Processes, VCH
Publishers, New York, 1995.
6. F. Schmickler, Interfacial Electrochemistry,
Oxford University Press, Oxford, 1996.
7. R. M. Metzger in Molecular Electronics Science and Technology (Ed.: A. Aviram), American Institute of Physics Conference Proceedings, New York, 1992, Vol. 262, p. 85.
8. C. A. Mirkin, M. A. Ratner, Annu. Rev.
Phys. Chem. 1992, 43, 719754.
9. K. Sienicki, (Ed.), Molecular Electronics and
Molecular Electronic Devices, CRC Press,
New York, 1994.
10. M. C. Petty, M. R. Bryce, D. Bloor, (Eds.),
An Introduction to Molecular Electronics,
Oxford University Press, Oxford, 1995.
11. C. Joachim, S. Roth, (Eds.), Atomic and
Molecular Wires, Kluwer, Dordrecht, The
Netherlands, 1997, Vol. 341.
12. J. Jortner, M. Ratner, (Eds.), Molecular Electronics, Blackwell Science, Oxford, 1997.
13. L. Kouwenhoven, Science 1997, 275, 1896,
1897.
14. A. Aviram, M. Ratner, (Eds.), Molecular
Electronics: Science and Technology, New
York Academy of Sciences, New York, 1998.
15. C. Dekker, Phys. Today 1999, 52, 2228.
16. M. A. Reed, Proc. IEEE 1999, 87, 652658.
17. A. G. Davies, Philos. Trans. R. Soc. London
Ser. A Phys. Sci. Eng. 2000, 358, 151172.
18. R. Landauer, IEEE Trans. Electron Devices
1996, 43, 16371639.
19. R. A. Marcus, J. Chem. Phys. 1965, 43, 679.
20. S. Gosavi, R. A. Marcus, J. Phys. Chem.
2000, 104, 20672072.
21. M. R. Wasielewski, M. P. Niemczyk, D. G.
Johnson et al., Tetrahedron 1989, 45, 4785.
22. S. B. Sachs, S. P. Dudek, L. R. Sita et al., J.
Am. Chem. Soc. 1997, 119, 10 563.
23. V. Grosshenny, A. Harriman, R. Ziessel,
Angew. Chem., Int. Ed. Engl. 1996, 34, 2705.
24. J. M. Tour, Chem. Rev. 1996, 96, 537.
25. A. Osuka, N. Tanade, S. Kawabata et al., J.
Org. Chem. 1995, 60, 7177.
26. A. C. Ribou, J. P. Launay, K. Takahashi
et al., Inorg. Chem 1994, 33, 1325.
1.3 Electronic Tunnel Factors in Molecular Electron Transfer and Molecular Conduction
51. K. P. Ghiggino, A. H. A. Clayton, J. M.
Lawson et al., New J. Chem. 1996, 20, 853.
52. H. Oevering, M. N. Paddon-Row, M. Heppener et al., J. Am. Chem. Soc. 1987, 109,
3258.
53. M. T. Carter, G. K. Rowe, J. N. Richardson
et al., J. Am. Chem. Soc. 1995, 117, 2896.
54. H. L. Guo, J. S. Facci, G. Mclendon, J. Phys.
Chem. 1995, 99, 8458.
55. B. Paulson, K. Pramod, P. Eaton et al., Phys.
Chem. 1993, 97, 13 042.
56. M. D. Johnson, J. R. Miller, N. S. Green
et al., J. Phys. Chem. 1989, 93, 1173.
57. C. A. Stein, N. A. Lewis, G. Seitz, J. Am.
Chem. Soc. 1982, 104, 2596.
58. V. J. Langlais, R. R. Schlittler, H. Tang et al.,
Phys. Rev. Lett. 1999, 83, 28092812.
59. E. E. Polymerpopoulos, D. Mobius,
H. Kuhn, Thin Solid Films 1980, 68, 173.
60. M. Fujihira, K. Nishiyama, H. Yamada, Thin
Solid Films 1985, 132, 77.
61. S. Roth, M. Burghard, C. M. Fischer, Resonant tunneling and molecular rectication in Langmuir-Blodgett lms in
Molecular Electronics (Eds.: J. Jortner,
M. Ratner), Blackwell Science, Oxford,
1997, pp. 255280.
62. E. Delamarche, B. Michel, H. A. Biebuyck
et al., Adv. Mater. 1996, 8, 718.
63. L. A. Bumm, J. J. Arnold, M. T. Cygan et al.,
Science 1996, 271, 17051707.
64. U. Durig, O. Zuger, B. Michel et al., Phys.
Rev. B 1993, 48, 1711.
65. A. Dhirani, P. H. Lin, P. Guyot-Sionnest, J.
Chem. Phys. 1997, 106, 52495253.
66. L. Ottaviano, S. Santucci, S. D. Nardo et al.,
J. Vac. Sci. Technol., A 1997, 15, 1014.
67. M. A. Reed, C. Zhou, C. J. Muller et al.,
Science 1997, 278, 252254.
68. D. Porath, O. Millo, J. Appl. Phys. 1997, 81,
22412244.
69. D. Porath, Y. Levi, M. Tarabiah et al., Phys.
Rev. B 1997, 56, 98299833.
70. C. Joachim, J. K. Gimzewski, R. R. Schlittler
et al., Phys. Rev. Lett. 1995, 74, 2102.
71. S. Datta, W. D. Tian, S. H. Hong et al.,
Phys. Rev. Lett. 1997, 79, 25302533.
72. W. D. Tian, S. Datta, S. H. Hong et al., J.
Chem. Phys. 1998, 109, 28742882.
73. J. R. Hahn, Y. A. Hong, H. Kang, Appl.
Phys. A -Mater. Sci. Process. 1998, 66,
S467S472.
74. W. Han, E. N. Durantini, T. A. Moore et al.,
J. Phys. Chem. 1997, 101, 10 71910 725.
75
76
1 Fundamentals
99. B. G. Rudberg, M. Johnson, Phys. Rev. B
1991, 34, 9358.
100. A. Liebsch, Electronic Excitations at Metal
Surfaces, Plenum Press, New York, 1997.
101. M. Buttiker, R. Landauer, Phys. Rev. Lett.
1982, 49, 17391742.
102. M. Buttiker, Phys. Rev. B 1983, 27,
61786188.
103. J. Tersoff, D. R. Hamman, Phys. Rev. B
1985, 5031, 805.
104. J. Bardeen, Phys. Rev. Lett. 1961, 6, 57.
105. R. Landauer, IBM J. Res. Dev. 1957, 1, 223.
106. R. Landauer, Philos. Mag. 1970, 21,
863867.
107. Y. Imry, Physics of mesoscopic systems
in Directions in Condensed Matter Physics
(Eds.: G. Grinstein, G. Mazenko), World
Scientic, Singapore, 1986, p. 101.
108. Y. Imry, Introduction to Mesoscopic Physics,
Oxford University Press, Oxford, 1997.
109. W. H. Miller, S. D. Schwartz, J. W. Tromp,
J. Chem. Phys. 1983, 79, 48894898.
110. M. Galperin, D. Segal, A. Nitzan, J. Chem.
Phys. 1999, 111, 15691579.
111. V. Mujica, M. Kemp, M. A. Ratner, J. Chem.
Phys. 1994, 101, 68496864.
112. D. M. Newns, Phys. Rev. 1969, 178, 1123.
113. J. N. Onuchic, D. N. Beratan, J. Chem. Phys.
1990, 92, 722.
114. P. C. P. D. Andrade, J. N. Onuchik, J. Chem.
Phys. 1998, 108, 42924298.
115. A. Onipko, Phys. Rev. B 1999, 59,
999510 006.
116. A. Barraud, P. Millie, I. Yakimenko, J. Chem.
Phys. 1996, 105, 69726978.
117. A. Nitzan, J. Phys. Chem. A 2001, 105, 2677.
118. D. Segal, A. Nitzan, M. A. Ratner et al., J.
Phys. Chem. 2000, 104, 2790.
119. P. Sautet, C. Joachim, Chem. Phys. 1989,
135, 99.
120. P. Sautet, C. Joachim, Chem. Phys. Lett.
1991, 185, 23.
121. P. Sautet, J. C. Dunphy, D. F. Ogletree et al.,
Surf. Sci. 1994, 315, 127.
122. C. Chavy, C. Joachim, A. Altibeli, Chem.
Phys. Lett. 1993, 214, 569.
123. P. Doumergue, L. Pizzagalli, C. Joachim
et al., Phys. Rev. B 1999, 59, 15 91015 916.
124. C. Joachim, J. K. Gimzewski, Europhys. Lett.
1995, 30, 409.
125. C. Joachim, J. F. Vinuesa, Europhys. Lett.
1996, 33, 635640.
126. M. Magoga, C. Joachim, Phys. Rev. B 1997,
56, 47224729.
1.3 Electronic Tunnel Factors in Molecular Electron Transfer and Molecular Conduction
151. D. Evans, I. Benjamin, T. Seidman et al.,
Abstr. Papers Am. Chem. Soc. 1996, 212,
194.
152. A. Haran, A. Kadyshevitch, H. Cohen et al.,
Chem. Phys. Lett. 1997, 268, 475480.
153. A. Mosyak, A. Nitzan, R. Kosloff, J. Chem.
Phys. 1996, 104, 15491559.
154. A. Mosyak, P. Graf, I. Benjamin et al., J.
Phys. Chem. A 1997, 101, 429433.
155. A. Nitzan, I. Benjamin, Acc. Chem. Res.
1999, 32, 854861.
156. I. Benjamin, D. Evans, A. Nitzan, J. Chem.
Phys. 1997, 106, 12911293, 66476654.
157. R. Naaman, A. Haran, A. Nitzan et al., J.
Phys. Chem. B 1998, 102, 36583668.
158. U. Peskin, A. Edlund, I. Bar-On et al., J.
Chem. Phys. 1999, 111, 75587566.
159. N. D. Lang, Phys. Rev. B 1988, 38, 10 395.
160. N. D. Lang, A. Yacoby, Y. Imry, Phys. Rev.
Lett. 1989, 63, 1499.
161. N. D. Lang, Phys. Rev. B 1992, 45, 13 599.
162. N. D. Lang, Phys. Rev. B 1995, 51, 2029(E).
163. N. D. Lang, Phys. Rev. B 1995, 52,
53355342.
164. N. D. Lang, P. Avouris, Phys. Rev. Lett. 1998,
81, 35153518.
165. N. D. Lang, P. Avouris, Phys. Rev. Lett. 2000,
84, 358361.
166. R. G. Parr, W. Yang, Density Functional Theory of Atoms and Molecules, Oxford University Press, Oxford, 1989.
167. O. Gunnarsson, R. O. Jones, Rev. Mod.
Phys. 1989, 61, 689.
168. N. Lang, Phys. Rev. Lett. 1986, 56, 1164.
169. M. Di Ventra, S. T. Pantelides, N. D. Lang,
Phys. Rev. Lett. 2000, 84, 979982.
170. M. Di Ventra, S. T. Pantelides, Phys. Rev. B
1999, 59, R5320R5323.
171. G. Galli, M. Parrinello in Computer Simulations in Material Science (Eds.: V. Pontikis,
M. Meyer), Kluwer, Dordrecht, The Netherlands, 1991.
172. K. Hirose, M. Tsukada, Phys. Rev. B 1995,
51, 52785290.
173. V. Mujica, A. E. Roitberg, M. A. Ratner, J.
Chem. Phys. 2000, 112, 68346839.
174. A. Szabo, N. S. Ostlund, Modern Quantum
Chemistry: Introduction to Advanced Electronic Structure Theory, McGraw-Hill, New
York, 1989.
175. H. J. Silverstone, M. L. Yin, J. Chem. Phys.
1968, 49, 2020.
176. S. Huzinaga, C. Arnan, Phys. Rev. A 1970,
1, 1285.
177. H. Burke, E. K. U. Gross in Density Functionals: Theory and Applications (Ed.: D. Joubert), Springer, Berlin, 1998.
178. D. Lamoen, P. Ballone, M. Parrinello,
Phys. Rev. B: Condens. Matter 1996, 54,
50975105.
179. A. Aviram, M. A. Ratner, Chem. Phys. Lett.
1974, 29, 277.
180. D. H. Waldeck, D. N. Beratan, Science 1993,
261, 576, 577.
181. R. A. Marcus, J. Chem. Soc., Faraday Trans.
1996, 92, 39053908.
182. M. Pomerantz, A. Aviram, R. A. McCorkle
et al., Science 1992, 255, 1115.
183. A. S. Martin, J. R. Sambles, Adv. Mater.
1993, 5, 580582.
184. A. S. Martin, J. R. Sambles, G. J. Ashwell,
Phys. Rev. Lett. 1993, 70, 218221.
185. C. M. Fischer, M. Burghard, S. Roth et al.,
Europhys. Lett. 1994, 28, 129134, 375.
186. K. Nagesha, J. Gamache, A. D. Bass et al.,
Rev. Sci. Instrum. 1997, 68, 38833889.
187. R. M. Morsolais, M. Deschenes, L. Sanche,
Rev. Sci. Instum. 1989, 60, 27242732.
188. G. Makov, A. Nitzan, L. E. Brus, J. Chem.
Phys. 1988, 88, 50765085.
189. D. K. Ferry, S. M. Goodnick, Transport
in Nanostructures, Cambridge University
Press, Cambridge, 1997, Vol. 6.
190. R. Wilkins, E. Ben-Jacob, R. C. Jaklevic,
Phys. Rev. Lett. 1989, 63, 801804.
191. M. A. Kastner, Rev. Mod. Phys. 1992, 64,
849.
192. S. H. M. Persson, L. Olofsson, L. Gunnarsson, Appl. Phys. Lett. 1999, 74, 25462548.
193. M. Dorogi, J. Gomez, R. Osifchin et al.,
Phys. Rev. B: Condens. Matter 1995, 52,
90719077.
194. R. P. Andres, S. Datta, M. Dorogi et al., J.
Vac. Sci. Technol., A 1996, 14, 11781183.
195. R. P. Andres, T. Bein, M. Dorogi et al., Science 1996, 272, 13231325.
196. A. N. Korotkov, Coulomb blockade and digital single electron devices in Molecular Electronics (Eds.: J. Jortner, M. Ratner), Blackwell Science, Oxford, 1997, pp. 157189.
197. A. M. Kuznetsov, J. Ulstrup, J. Electroanal.
Chem. 1993, 362, 147152.
198. F.-R. F. Fan, A. J. Bard, Science 1997, 277,
17911793.
199. L. I. Malysheva, A. I. Onipko, Phys. Rev. B
1992, 46, 39063915.
200. V. Mujica, M. Kemp, A. Roitberg et al., J.
Chem. Phys. 1996, 104, 72967305.
77
78
1 Fundamentals
201. Y. Kawahito, H. Kasai, H. Nakanishi et al.,
J. Appl. Phys. 1999, 85, 947952.
202. Y.-Q. Li, C. Gruber, Phys. Rev. Lett. 1998, 80,
10341037.
203. A. S. Davidov, V. M. Ermakov, Physica 1987,
28D, 168180.
204. H. L. Berkowitz, R. A. Lux, J. Vac. Sci.
Technol. 1987, B5, 967970.
205. F. W. Sheard, G. A. Toombs, Appl. Phys.
Lett. 1988, 52, 12281230.
206. N. C. Kluksdahl, A. M. Kriman, D. K. Ferry
et al., Phys. Rev. B 1989, 39, 77207735.
207. A. Nakano, R. K. Kalia, P. Vashishta, Phys.
Rev. B: Condens. Matter 1991, 44,
81218128.
208. J. Chen, M. A. Reed, A. M. Rawlett et al.,
Science 1999, 286, 15501552.
209. S. A. Gurvitz, H. J. Lipkin, Y. S. Prager,
Mod. Phys. Lett. B 1994, 8, 1377.
210. S. A. Gurvitz, H. J. Lipkin, Y. S. Prager,
Phys. Lett. A 1996, 212, 9196.
211. S. A. Gurvitz, Y. S. Prager, Phys. Rev. B
1996, 53, 15 93215 943.
81
2.1
Fundamental Concepts
82
2 Transport Phenomena
dU = T dS +
i
i dni
(1)
dU = T dS +
i dni
(2)
(3)
When the transport of species i is not
coupled to that of other species, each
individual term of this sum is positive.
In this case, dni is determined only by
( i i ) and species i moves towards
the region in which its electrochemical
potential is lower, that is, dni < 0 when
i < i and vice versa. This happens, for
instance, in the case of ionic species in
diluted solutions. When the transport of
different species is coupled, one or more
terms in the sum can be negative, but the
sum is always positive.
The transport of species i can be
described in terms of either its velocity vi or
its ux density Ji , which are related to each
other by the molar concentration, Ji =
ci vi . If the area of the surface separating
the two volume elements is dA and its
orientation is given by the unit vector
n(from
(4)
(6)
(7)
2 =
e
F
zi ci
=
(8)
83
84
2 Transport Phenomena
d
dt
ci (r , t)dV =
ci
dV
t
ri dV Ji ndA
=
Ji )dV
(ri
(10)
(11)
2.1.1.3
zi Ji
(12)
i = 0, i)
(c
(13)
I F
F2 2
zi Di ci
RT
(14)
(15)
(16)
F
zi
(18)
i
zi Di c
RT ti
ln ci
F
zi
(19)
This is due to the diffusion potential gradient that originates from the diffusion
process, and therefore an expression similar to Ohms law can only be recovered
by subtracting the diffusion potential gradient from the total potential gradient, as
shown by Eq. (17). A theoretical simplication sometimes used in the literature
consists in neglecting the diffusion potential by assuming all ionic diffusivities to be
equal [17, 18].
Diffusion-conduction Flux
Equation
The introduction of the modied Ohms
law, Eq. (17), in the NernstPlanck equation, Eq. (7), leads to the following form of
the ux equation
2.1.1.5
i + zi ci f
dif ) + ti I
Ji = Di (c
zi F
(20)
For the sake of simplicity, consider the case
of a binary solution obtained by complete
dissociation of a strong electrolyte A1 B2
into 1 ions Az1 of charge number z1
and 2 ions B z2 of charge number z2 ;
obviously, the relation z1 1 + z2 2 = 0 is
satised. The ionic concentrations are then
related to the stoichiometric electrolyte
concentration c12 by the relations c1 =
1 c12 and c2 = 2 c12 which allows Eq. (20)
to be rewritten as
ti I
12 + ti I
Ji = i J12 +
= i D12 c
zi F
zi F
(21)
where J12 is the electrolyte ux density, and
D12 =
(1 + 2 )D1 D2
2 D1 + 1 D2
(22)
85
86
2 Transport Phenomena
DBL
Electrode
Bulk solution
I
fb
f(x)
Potential drop in DBL
f fb fs
fs
Electrode potential
E fe fs
fe
(23)
zi ir
zi F Jis
=
(25)
I
n
which is called the integral transport
number of species i [23]; this magnitude
is zero for the species that are not
involved in the electrode reaction (i.e. the
electroinactive species).
It is important
to observe that i Tis = 1 but Tis is not
bounded between 0 and 1. For example, in
the electrode reaction
Tis
87
88
2 Transport Phenomena
(27)
(28)
(30)
s
s
THNO
= 0 but JHNO
= I /2F , as can
2
2
be easily deduced from Eq. (24). (For the
neutral species, the transference number
i Ti /zi constitutes a good alternative to
Ti [24].)
Eeq = E +
RT r cis
i ln
nF
c
(31)
the standard concentration. The concentrations cis of the active species in Eq. (31)
are those at the electrode surface, which
are evaluated by extrapolation to the electrode surface of the concentration prole
within the DBL [19]. In other words, the
electrode surface at position x = 0 should
be understood as the neutral solution just
outside the electrical double layer (EDL).
Similarly, % refers to the potential difference between the bulk solution and
that point in the solution (that is, between
the bulk solution|DBL and the EDL|DBL
boundaries). Hence, no details of the EDL
structure are considered when evaluating
the equilibrium interfacial potential drop
and the surface concentrations of the active species.
Since the interfacial concentrations of
the active species vary with the electric
current density, Eeq also does and its
variation is often described in terms of
the concentration overpotential [19, 26]
b
c Eeq Eeq
RT r cis
i ln b (32)
nF
ci
i
which represents the potential of the electrode in equilibrium with the interfacial
concentrations that are established when
a current density I is driven through the
system, Eeq , relative to the (constant) potential of a similar electrode equilibrated
b . The potential
with the bulk solution, Eeq
difference between the electrode and the
b + %.
bulk solution is then Eeq
c
Note, nally, that when the supporting
electrolyte is used, the ionic strength
of the DBL is not signicantly affected
by the concentration polarization of the
active species, and therefore, their activity
coefcients are practically independent of
the current density [26]. This justies the
use of concentrations instead of activities
in the above equations.
2.1.2
The nite transport rate of the electroactive species toward the electrode surface
implies the existence of a limiting current
density, that is the most important consequence of mass transport. This limiting
current density is dened by the condition
of maximum concentration gradient of the
electroactive species, which is achieved
when the concentration of this species
tends to zero at the electrode | depleted
DBL interface (or, more accurately, at the
EDL | DBL boundary).
In this section, the mass transport effects
on the steady state currentvoltage curves
are described for the (relatively common)
case of systems with only one electroactive
species in solution. This species is denoted
by subscript 1. The integration procedure
of the transport equations in multi-ionic
systems is presented rst and the general
currentvoltage curve is derived. Except
in Sect. 2.1.2.5, homogeneous reactions
are assumed to be absent. The interesting case of excess supporting electrolyte,
in which the active species is present in
very small concentration and transport
takes place mostly by diffusion, is discussed. Finally, the simple cases of binary
and ternary solutions are considered to illustrate the conclusions drawn from the
general solution.
Integration of the Transport
Equations
The fact that the ux density of the
electroinactive species, i = 1, is zero
simplies signicantly the solution of the
transport equations. These species are
in electrochemical equilibrium within the
DBL, and hence their concentrations ci
are related to the electric potential by the
2.1.2.1
89
90
2 Transport Phenomena
where %ci cib cis . Introducing the limiting diffusion current density
Boltzmann equation
d
dci
= zi ci f
dx
dx
(i = 1, Ji = 0) (33)
(i = 1, Ji = 0)
(34)
where % b s is the electric potential drop across the DBL (see Fig. 1).
Furthermore, the LEN assumption
zi ci = z1 c1 +
zi ci = 0
(35)
i =1
1 s
1 b zi f %
zi ci =
zi ci e
z1
z1
i =1
i =1
(36)
The currentvoltage relation can be obtained from the ux equation of the
electroactive species
I
d
dc1
J1 =
= D1
+ z1 c1 f
z1 F
dx
dx
(37)
by transforming the migration term with
the help of Eq. (35) to
z1 c1 f
dci
d
d
zi ci f
=
=
dx
dx
dx
i =1
i =1
(38)
Integration of Eq. (37) over the DBL
then gives
I
= %c1 +
%ci
z1 F D1
=
i =1
i =1
zi
1
z1
%ci (39)
ILd,1
z1 F D1 c1b
(40)
and using Eqs. (34 and 36), the currentvoltage relation takes the form
cb
zi
i
1
(1 ezi f % )
b
z
c
1
i =1 1
(41)
Note that the denition of ILd,1 has not
been deduced from the condition c1s = 0,
but it rather corresponds to a situation
in which the transport of species 1 takes
place mostly by diffusion, and hence its
name. Moreover, Eq. (40) corresponds to
the particular case in which there is only
one active species; see Eq. (78) below for
the general denition.
It is interesting to observe that species
1 is the only one that moves, while the
inactive species are standing still over
the system. In a sense, this is similar to
the conduction of electrons in a metal,
in which electrons move and the ions
are standing on the crystal lattice and,
therefore, Ohms law is expected to be
satised. Indeed, Eq. (37) shows that the
electric current density is proportional
to the electric eld, I eff (d/dx),
in which
I = ILd,1
eff (x) =
F2
zi2 ci (x)
D1
RT
(42)
zi =z1
zi =z1
(45)
1 ez1 f %
1 ez1 f %L
2C b
c1b
1 1
c1b
Cb
(46)
1/2
(47)
is the limiting current density (i.e. that
corresponding to c1s = 0 and % = %L ).
The value of IL varies between 2ILd,1
in the absence of supporting electrolyte
(C b = c1b ) and ILd,1 in the presence
of excess supporting electrolyte (C b
c1b ). Migrational effects in homovalent
solutions can therefore account for up to a
factor 2 in IL .
Binary Electrolyte Solutions
When the solution contains only two
ionic species, the concentration proles
2.1.2.3
91
2 Transport Phenomena
IL
c1b
c2b
(50)
which can be obtained by integration of
Eq. (37) from x to , after use of Eq. (38).
In the depleted DBL, the interfacial
(48)
1.0
0.5
Concentrated
DBL
where
(z1 z2 )F D1 c2b
z1
IL = 1
ILd,1 =
z2
(49)
is the limiting current density. This curve
has been plotted in Fig. 2. The gure
I/IL
92
Depleted DBL
0.0
0.5
1.0
1
z2 ff
2.0
1.0
1.5
0.5
ci(x)/cbi
ci(x)/cbi
Depleted DBL
0.0
0.4
0
0.0
0.5
1.0
0.0
x/d
0.5
1.0
0.8
1.5
z2 f [fb f(x)]
z2 f [fb f(x)]
1.0
0.0
4
x/d
z2 z1
f (c %)
I = IL 1 exp
z1 z2
(52)
where c % is the change in the
potential difference between the electrode
and the bulk solution due to the passage
of the electric current.
z1 c1b
I
x
1
z1 z3 ILd,1
(53)
which has been obtained by integration of
Eq. (37) and making use of Eq. (38). The
c3 (x) = c3b
93
2 Transport Phenomena
2z1 F D1 c3b
(1 ez1 f % )
%L =
(57)
Equations (56 and 57) have been represented in Fig. 5 as a function of the electroactive/supporting electrolyte concentration ratio. When the electroactive species
0.02
f [fb f(x)]
0.04
1
f
0
0.0
0.5
1.0
0.00
1.5
x/d
Electric potential and ionic concentration proles
(marked with the respective subscripts 1, 2, 3) in a ternary
system with z1 = z2 = 1, z3 = 2, for a concentration ratio
electroactive to inactive cation c1b /c2b = 0.2, and two values of the
electric current density: I/ILd,1 = 0.5 (continuous line) and 1.0
(dashed line).
Fig. 4
(56)
1/2
2z1 F D1 c3b
c1b
1 1
IL =
c3b
3
2
RT
ln(1 + c1b /c2b )
2z1 F
and
0.06
(55)
ci(x)/cb1
94
3
2.0
1.5
1.0
z1 ffL
IL/ILd, 1
102
100
102
104
cb1/cb2
Dependence of the limiting current density
(continuous line) and the electric potential drop (dashed
line) in the DBL under limiting conditions on the
electroactive/supporting electrolyte concentration ratio.
Fig. 5
4c2
(58)
On the contrary, in the absence of
supporting electrolyte, c1b /c2b 1, IL
2ILd,1 and %L diverges; the ratio IL /ILd,1
tends to (1 z1 /z3 ) when c1b /c2b 1 in
systems with other charge numbers, as
can be deduced from Eq. (49).
The currentvoltage curves given by
Eq. (55) have been represented in Fig. 6.
b ,
The initial slope of these curves is eff
which obviously increases with decreasing
1000
50
5
I/ILd, 1
1.5
Currentvoltage curves of a
depleted DBL for different values of the
electroactive/supporting electrolyte
concentration ratio, c1b /c2b , shown on the
curves. % is the electric potential drop
in the DBL.
(59)
is the concentration overpotential. The potential drop % c is the change in electrode potential (relative to bulk solution)
due to the concentration polarization.
2.0
Fig. 6
cs
1
ln 1b
z1 f
c1
1
0.1
1.0
0.5
0.0
z1 ff
95
2 Transport Phenomena
I/ILd, 1
96
2.0
1000
50
1.5
5
1
1.0
0.1
0.5
0.0
z1 f(f hc)
z 2 c1
d
J1
= 1
dx
zi2 ci
i
c1
=
J1
2c3
(60)
is not affected by the passage of the electric current. It must be stressed, however,
that this statement refers to the concentration gradients in relative terms to the bulk
concentration of the respective species. In
absolute terms, the concentration gradients of the inactive species are similar to
that of the active species, and by no means
the concentration gradient of the active
species can be neglected. In fact, in the
homovalent case, it can be easily shown
that dc1 /dx 2dc3 /dx 2dc2 /dx when
migration is negligible. A straightforward
implication of these comments is that
the LEN assumption cannot be used in
conjunction with the excess supporting
electrolyte assumption.
Weak Binary Electrolyte
Consider an electrolyte A1 B2 dissociating into 1 ions Az1 of charge number z1
and 2 ions B z2 of charge number z2
2.1.2.5
z1
z2
A1 B2
1 A + 2 B
(62)
concentration. When the dissociation Reaction (62) is fast compared with the transport process, it can be assumed that the
dissociated and undissociated electrolyte
fractions are in equilibrium and the mass
action law
K=
1 2 12
c11 c22
12 1 1 2
= c12,T
c12,u
1
dc12,u
dx
(64)
12 (1 )D12,u + D12
12 (1 ) +
J12,T =
(63)
(66)
t2 I
z2 2 F
(67)
t2 I
z2 2 F
IL,u
b
z2 2 F D12,u c12,T
(71)
t2
and
IL
b
z2 2 F D12 c12,T
t2
b
z1 12 F D1 c12,T
(72)
97
2 Transport Phenomena
cb
RT
RT
12 D12,u
ln 12
s F z ( 1)D
z2 F
c12
2 12
2
b
s
c12,u
c12,u
(74)
s
b
c12
c12
% =
(cb12)2/K
102
100
102
104
2
IL, T /IL
98
Fig. 8
0.5
0
102
100
102
cb12/K
104
Fig. 9
0.01
1.0
IL, T /IL
0.1
1
10
0.5
100
1000
0.0
10
ff
s
that the value of s is independent of c12,T
,
which is inconsistent with Eq. (63). It is
then concluded that Eq. (63) is not valid in
the vicinity of the electrode. The study of
the limiting current without the assumption that the dissociation reaction is fast
compared with the electrodiffusion process has been carried out by Kharkats and
Sokirko [36, 37]. The validity of the dissociation equilibrium assumption and the
drastic changes in concentrations that occur when I IL have been considered by
Vorotyntsev [38].
2.1.3
99
100
2 Transport Phenomena
cb c1s
I
dc1
D1
= D1 1
J1 =
z1 F
dx
(75)
and the interfacial concentration can be
expressed as
I
s
b
(76)
c1 = c1 1
ILd,1
where ILd,1 is the limiting diffusion
current density dened by Eq (40). Equation (76) can be rewritten in terms of the
concentration overpotential, Eq. (32), as
I = ILd,1 (1 ez1 f c )
(77)
(81)
1 I /ILd,Fe3+
1
ln
f
1 I /ILd,Fe2+
1 + (DFe2+ /DFe3+ )I /ILd,Fe2+
1
ln
f
1 I /ILd,Fe2+
(82)
which has been represented in Fig. 10.
2.1.3.2
ILd,i
nF Di cib
ir
(78)
Diffusion-reaction Processes
The general strategy for solving transport
problems in which the electrode reaction,
Eq. (23), is coupled to a homogeneous
reaction in the DBL
i Cizi = 0
(83)
2.1.3.3
Fig. 10
I/ILd, Fe2+
1.0
0.5
0.0
0.0
2.0
4.0
6.0
fhc
means that the forward and backward reaction rates are much larger than the
mass transport by diffusion, and hence
their difference can be neglected in a
rst approximation. Actually, there is a
small difference between the forward and
backward reaction rates, the net reaction rate
r
1 dJi
ri
=
i
i dx
(84)
101
102
2 Transport Phenomena
the form
dJj
dJi
j,i
=
=0
dx
dx
(85)
j,i Ji =
j,i Di
dci
dx
I
=
j,i ir
nF
(86)
(89)
(90)
ci(x)
[mM]
50
1.0
I
[CdI4]2
104
[CdI4]2
102
Cd2+
0.5
ci(x)/cbi
100
Cd2+
0.0
I
ci(x)
[mM]
100
[CdI4]2
2 104
2 104
50
0
0.0
0.5
1
104
102
Cd2+
1.0
x/d
0.0
0.5
1.0
r
[mol L1 s1]
0
1.5
x/d
Fig. 11
103
104
2 Transport Phenomena
is then needed either to derive approximate expressions for the electric potential
distribution in some other way [50] or
to calculate the exact distribution as described below. In any case, if the electric
eld distribution E(x) or, equivalently, the
electric potential distribution
Edx
(93)
Ji
zi Di f E
(94)
N
zik ci ,
k = 0, 2, 3, . . . , N (95)
i=1
(92)
Sk
(x) = b +
ezi f [
method, that uses the concept of valency classes [32], these authors proposed
to rewrite the NernstPlanck equation
system for N species in terms of the
magnitudes
Gk
N
zik Ji /Di ,
i=1
k = 0, 1, 2, . . . , N 1
(96)
(97)
(98)
(x) = b +
where
3
1 G1
z1 z2 G0
(99)
(100)
S0b I
G0
(101)
cis
IL
(1 + zi 3)ILd,i
I 1+zi 3
I zi 3
1
1
1
IL
IL
cib
(104)
which reduces to cis = cib (1 I /IL ) in the
binary case. In deriving Eq. (104), note that
Ji /Di cib = I /ILd,i .
In the case of nonhomovalent multiionic systems, the solution procedure
is necessarily more complicated because
there is no simple relation such as S2 =
z2 S0 , a problem that was rst encountered
and solved in different ways by Schlogl [59]
and Brady and Turner [60]. If the system
contains N different ionic species, the N
NernstPlanck equations for the N ux
densities Ji are replaced by
dSk
+ f ESk+1 ,
dx
k = 0, 1, 2, . . . , N 1
Gk =
% =
(102)
zi f ( b )
dx =
(103)
the interfacial concentration is nally
given by
(105)
(106)
where
q1 = z1 + z2 + + zN
(107)
q2 = z1 z2 + z1 z3 + + z1 zN + z2 z3 +
+ z2 zN + z3 z4 + + zN 1 zN (108)
S0zi 3 dS0
G0 (S0b )zi 3
1+zi 3
S0b
G0
1+
=
1
G0 (1 + zi 3)
S0b
0
1+
q3 = z1 z2 z3 + + z1 z2 zN + z1 z3 z4 +
+ z1 z3 zN + + zN 2 zN 1 zN (109)
..
.
qk =
+2k
N
+1k N
i1 =1 i2 =i1 +1
N
ik =ik1 +1
(110)
105
106
2 Transport Phenomena
dS2
+ f E S3
dx
S 3 = q1 S 2 + q3 S 0
G2 =
(111)
(112)
(114)
= 0.1904
G0 =
D2
D1 F
F D1
(115)
z2
z1 I
I
G1 =
= 1.5712
D2
D1 F
F D1
(116)
z22
z12 I
I
= 6.7136
G2 =
D2
D1 F
F D1
(117)
where D1 = 0.719 105 cm2 s1 and
D2 = 0.604 105 cm2 s1 [39]. Equation (113) can be rewritten in terms of the
4cb ILd,1
(118)
+ 11.456 1 + 2b
I
3c1
where ILd,1 = F D1 c1b / is the limiting
diffusion current density of ferrous ion.
Equation (118) can be integrated by
standard numerical methods, such as
fourth-order RungeKutta method [61],
starting from position x = , in which
the boundary condition eb = G1 /S2b is
applied as
x
0.2619
I
(119)
=1 =
e
1.0
I/ILd, 1
10
0.5
0.1
0.5
0.0
0.00
0.05
0.10
0.15
ff
The ferrous ion concentration is obtained from Eq. (93) and using the relation J1 /D1 c1b = I /ILd,1 . The chloride
ion follows Boltzmanns equation c3 =
b
c3b ef ( ) , and the ferric ion concentration is obtained from the LEN assumption
c2 = (c3 2c1 )/3.
Figures 12 and 13 evidence that the
potential drop in the DBL is small and
is distributed quite uniformly. It is then
expected that the GCF assumption, E =
%/, provides reasonable estimates.
Indeed, by setting c1s = 0 in Eq. (94),
the limiting current density can be
estimated as
IL
2f %
=
ILd,1
1 e2f %
(120)
dS3
+ f E S4
dx
S 4 = q1 S 3 q2 S 2 q4 S 0
G3 =
(121)
(122)
0.09
ci(x)/c1b
0.06
f
0.03
Fe3+
f[f(x) fb]
Cl
Fe2+
0
0.0
0.00
0.5
1.0
1.5
x/d
Electric potential and concentration proles of Fe2+ ion
(i = 1), Fe3+ ion (i = 2), and Cl ion (i = 3) for a bulk
concentration ratio c2b /c1b = 1.0 and the limiting current density
I = IL = 0.9261ILd,1 .
Fig. 13
107
108
2 Transport Phenomena
+ (G2 q1 G1 )
E dx 2
dx
E2
f
dE
(G3 q1 G2 + q2 G1
dx
+ q4 S0 f E)(f E)2 = 0
(123)
Needless to say, transient transport conditions involve a higher level of mathematical complexity to the description of
ion transport in electrochemical systems.
The only case that can be worked out with
ease is that of binary solutions. Ternary
and multi-ionic solutions necessarily involve systems of partial differential equations coupled through the migration term
that require advanced numerical methods.
Given this situation, it is by no means
surprising that most theoretical studies of
transient problems neglect migration and
use the diffusion equation (or Ficks second law) to describe the transport of ionic
species. (The analysis of the role of supporting electrolytes in Sects. 2.1.2.2 and
2.1.2.4 was restricted to steady state conditions, but it is expected that a large excess
of supporting electrolyte makes migration
negligible under transient conditions too.)
This section describes diffusion and migration under transient conditions in some
simple illustrative cases. For the sake
of simplicity, an innite DBL thickness
is considered, which is justied in the absence of stirring. Current
and voltage steps in semi-innite planar geometry (and in the absence of
homogeneous reactions) are described.
Both Laplace transforms and Boltzmanns
change of variables are used. The inuence of migration is discussed analytically
(whenever possible) by using the latter
technique, which is introduced in Appendix B. Other solution techniques of
the diffusion equation [6264], and the
electrodiffusion equations are available in
the literature [13, 6572]. Further information on electrochemical techniques can
be found in Ref. [73].
Ficks Second Law for a Strong
Binary Electrolyte
The diffusionconduction equation,
Eq. (21), expresses the ux density of
an ionic species from a strong binary
electrolyte in terms of the electrolyte
ux J12 and the current density I .
Under transient conditions, both Ji
and J12 are position-dependent, but
I is not. Thus, taking into account
Eq. (21) and the relation ci = i c12 , the
continuity equation, Eq. (11), leads to
Ficks second law
2.1.5.1
c12
2 c12
= D12
t
x 2
(125)
c1
2 c1
= D1 2
t
x
(126)
Eeq (t) = E +
ln 1
z1 F
c
c1b
1/2
+ Aex(s/D1 )
s
+ Bex(s/D1 )
1/2
When the system is perturbed by applying a known function, I (t), Eq. (127) is
used to evaluate the chronopotentiometric response of the system. Analogously,
when the system is perturbed by applying a
known function Eeq (t), Eq. (128) describes
the chronoamperometric response of the
system.
(130)
c1b
1 + (ez1 f %E 1)
s
1/2
ex(s/D1 )
(131)
(128)
(129)
1/2
(132)
109
110
2 Transport Phenomena
z1 F c1b
D1
t
1/2
(134)
c1b
I0
1/2
s 3/2 ex(s/D1 )
+
1/2
s
z1 F D1
(135)
whose inverse transform provides the
concentration distribution
c1 (x, s) =
2I0 t 1/2
c1 (x, t) = c1b +
1/2
z1 F D1
[ 1/2 e erfc( )]
2
(136)
Eeq (t > 0) = E +
RT
z1 F
cb
2I0
ln 1 +
c
z1 F c
t
D1
1/2
(137)
Sk
Gk
= D
t
x
(139)
G0 =
(140)
(141)
(142)
I (t) r
i
nF D
(143)
Gs2 (t) =
I (t) 2 r
I (t)
(z1 + z2 )
zi i =
nF D
FD
i
(144)
which are used as boundary conditions
in chronopotentiometric techniques or to
evaluate the system response in chronoamperometric techniques. In the second
equality of Eq. (144), the common ion
is considered to be inactive, 3r = 0. The
parameter G1 = I (t)/FD is only a function of t.
Combining Eq. (139) for k = 0 and
Eq. (140), the partial differential equation
for G0 is
G0
2 S0
S0
= D
=D 2
t
x
x
(145)
(148)
(149)
(S b )2 dS2
d2 S2
dS2
+ 2
+ 22
(150)
2
d
d
S2 d
111
2 Transport Phenomena
Current density in a voltage
step of large width relative to that in
absence of migration, given by Cottrell
Eq. (152), as a function of the bulk
concentration ratio product/reactant.
The charge numbers z1 , z2 , z3 are
shown close to the curves.
Fig. 14
4, 1, 1
2.0
4, 1, 2
I/ICottrell
112
1.5
Reduction
2, 1, 1
3, 2, 1
1.0
2, 3, 1
1, 2, 1
1, 4, 1
0.5
3
Oxidation
Log10(cb2/cb1)
I =
=
F (S2b )2
2q3 S0b
D
t
1/2
ICottrell (151)
where
ICottrell = (z2 z1 )F c1b
D
t
1/2
(152)
d =0
q3 S0b
(153)
e
3
=
t
tx 2
Id
=
=
(155)
x t
x
x
where Id is the displacement current,
that is, the time derivative of the electric displacement D = E. Alternatively,
= F
zi
Ji
I
=
x
x
(156)
where I is the conduction electric current density. Equation (156) represents the
conservation of electric charge and its combination with Eq. (155) leads to
I
Id
IT
=
+
=0
x
x
x
(157)
2
tx
i
(158)
The current density IT is equal to that
due to the electrons owing through the
external circuit connected to the electrodes.
In the absence of concentration gradients, no space charge density can exist in
the solution under steady state conditions.
Moreover, if by any means some electrical
charge density is generated at any position
within the system, the time required for
this charge to disappear (in fact, to migrate
to the system boundaries) is of the order
of nanoseconds. This time is the so-called
electrical relaxation time of the system, e ,
and can be obtained from Eqs. (8, 13, and
156) as
IT = I + Id = F
zi Ji
e
I
2
=
= 2
t
x
x
e
e
=
(159)
113
114
2 Transport Phenomena
x
dif
=
(160)
x
x
x
x
=
t
x
equation, Eq. (8) tells that the eld associated to this charge density over a region
of thickness LD is of the order of e LD /.
The migrational ux density associated to
this eld is of the order of F Dce LD /RT .
The condition for the existence of an equilibrium space charge density is Dc/LD
F Dce LD /RT , which yields the expression of the Debye length
LD
RT
Fe
1/2
RT
F 2c
1/2
(161)
Space charge distributions are then expected to exist over regions of thickness
LD , which is of the order of 107 cm
for a 100 mM aqueous solution. A practical consequence of this comment is that
the behavior of electrochemical systems
comprising microgeometries is affected by
space charge layers [8082].
Note, nally, that ionic motions associated to space charge redistribution involve
distances of the order of LD , which suggests that e can be interpreted as the time
e L2D /D required for the ions to diffuse
over LD [79, 8385].
Deviations from Local
Electroneutrality
Planck suggested that the electric potential calculated from the LEN assumption,
(0) , could be used to check the validity
of this assumption [27, 31]. When this approximate electric potential is introduced
in the Poisson equation, the space charge
density obtained
2.1.6.2
E(0)
2 (0)
=
x
x 2
(162)
(164)
dS0
f E S1
dx
(165)
E2
2RT
(166)
dS1
dS1
f E S2 =
+ z1 z2 f ES0
dx
dx
(167)
and introducing the limiting current density IL S0b I /(G0 ) dened in Eq. (101),
the equation for E becomes
L2D d2 E
x
I
1
IL
z
+ 1
=
z
1
2
E dx 2
IL f E
I
IL
+ (f ELD )2
(168)
2I
where LD (RT /F 2 S0b )1/2 is the Debye
length.
Analytical studies of Eq. (168) have
used different changes of variables to
write the electric eld in terms of either
Painleve transcendents [68] or Jacobian elliptic functions [88]. Alternatively, asymptotic expansions have also been used [68,
87, 89, 90]. The approximate solution
methods have neglected different terms
of Eq. (168). Thus, while Urtenov and
Nikonenko [55, 91, 92] have considered
that the space charge density is quasiuniform, dS1 /dx 0, Bass [93] assumed
that E2 RT S0b and I /IL /LD to express the electric eld as the sum of the
electroneutral electric eld, which can be
derived from Eq. (99), and a modied Gouy
distribution corrected for the presence of
current. In any case, both analytical and
numerical [94, 95] solutions have shown
that in the range of underlimiting currents,
the deviations from LEN are conned to
the EDL and that the latter is not signicantly perturbed by the electric current.
In conclusion, r in Eq. (163) is a
physically meaningful magnitude that is
very similar to the actual e when LEN is
a good approximation. Contrarily, when
deviations from LEN occur, the actual
electric potential surely differs from that
obtained by using the LEN, and then the
actual e is likely to show no resemblance
to r .
Although the equations presented in
this chapter cannot account for this fact,
115
116
2 Transport Phenomena
=
Fc
F c dx 2
F c L2
LD 2
=
1
(169)
L
2.1.6.3
ci = ci
(1)
+ ci
(2)
+ 2 ci
(170)
d2 (0)
dx 2
(172)
Since (0) is the electric potential calculated from the transport equations at zero
order in (that is, making use of the LEN
assumption), the residual space charge
density dened in Eq. (162) is identied as
(1)
e . The use of higher-order terms in the
Expansions (170 and 171) is only needed
(1)
when e is of the order of F c [13, 27, 31,
67]. A different validity test for the LEN as(2)
(1)
sumption, namely e e , has been
recently proposed by Feldberg (with a different notation) [99].
Note that it is possible to use either the
full Poisson equation or the terms of its
asymptotic expansion, such as the LEN as(0)
sumption or Eq. (172), but e can never
be interpreted as the full e in the Poisson
equation. This would lead to the wrong
conclusion that the LEN assumption implies a constant electric eld.
Appendix A
F2
= z1 z2
(2 D1 + 1 D2 )
c12
RT
(174)
where the last equality is known as the
NernstEintein relation [3]. Equation (17)
then implies
;12
I
I dx
dohm = dx =
;12 c12
(175)
z1 1 F
(D12,u dc12,u + D12 dc12 )
t2 I
(176)
and using Eq. (63) to deduce that d lnc12,u=
12 d ln c12 , Eq. (175) is transformed to
dx =
z1 1 F 1
(D12,u dc12,u
t2 ;12 c12
RT 12 t1 dc12
+ D12 dc12 ) =
F
z1 1 c12
D12,u dc12,u
RT 12 t1
d ln c12
=
z2 D2 c12
F
z1 1
12 D12,u
c12,u
d
(177)
z2 (12 1)D2
c12
dohm =
c1
c1
=
2
x t
x t x t t t
2
1
c1
=
(179)
4D1 t 2 t
c1
t
c1
c1
+
t t x
t
c1
c1
=
+
(180)
2t t
t
=
x
(181)
117
118
2 Transport Phenomena
position (m)
charge number, dened with sign
x
zi
Greek symbols
dielectric permittivity (C V1 m1 ),
Eq. (8)
Latin symbols
C
class concentration (mol m3 ),
Eq. (43)
ci
molar concentration (mol m3 )
D
diffusion coefcient (m2 s1 )
Eeq
equilibrium
electrode potential (V), Eq. (31)
,E
E
electric eld (V m1 )
e
f E dimensionless
electric eld, Eq. (118)
F
Faraday constant (C mol1 )
f
F /RT (V1 )
Gk
i zik Ji /Di (mol m4 ), Eq. (96)
I
electric current
density (A m2 ), Eqs. (12, 158)
Ji
ionic ux density (mol m2 s1 ),
Eq. (5)
K
equilibrium
constant
3
[(mol m ) i ], Eq. (63)
LD
Debye length (m), Eq. (161)
N
number of ionic species, Eq. (95)
n
= i zi ir stoichiometric
number of the electron, Eq. (23)
ni
number of moles of species
i (mol)
qk
auxiliary variable, Eq. (110)
R
gas constant (J mol1 K1 )
r
chemical reaction
rate (mol m3 s1 ), Eqs. (11, 84)
Sk
i zik ci (mol m3 ), Eq. (95)
s
Laplace variable (s1 ), Eq. (130)
T
temperature (K)
Ti
integral transport number, Eq. (25)
t
time (s)
ti
migrational transport number,
Eq. (15)
ui
ionic mobility (m2 V1 s1 ),
Eq. (4)
vi
ionic velocity (m s1 ), Eq. (4)
Ld, i
mig
ohm
T
(i)
b
r
s
limiting diffusion
(current density), Eq. (78)
migration, Eq. (60)
ohmic, Eq. (18)
total, Eq. (158)
perturbation order, Eq. (170)
bulk solution (x = )
standard state, Eq. (31)
electrode reaction, Eq. (23)
electrode surface (x = 0)
17.
18.
19.
20.
21.
References
1. R. Haase, Thermodynamics of Irreversible
Processes, Addison-Wesley, New York, 1969.
2. K. Kontturi, Acta Polytech. Scand. 1983, 152,
140.
3. R. A. Robinson, R. H. Stokes, Electrolyte Solutions, Butterworths, London, 1955.
4. N. Ibl in Comprehensive Treatise of Electrochemistry (Ed.: E. Yeager, J. OM. Bockris,
B. E. Conway et al.), Plenum Press, New
York, 1983, Vol. 6, Chap. 1, pp. 163.
5. J. S. Newman, Electrochemical Systems,
Prentice-Hall, Englewood Cliffs, N.J., 1991.
6. A. V. Sokirko, J. Electroanal. Chem. 1994,
364, 5162.
7. E. J. Calvo in Comprehensive Chemical Kinetics (Eds.: C. H. Bamford, R. G. Compton),
Elsevier, Amsterdam, 1986, Vol. 26,
Chap. 1, pp. 178.
8. R. Schlogl, Stofftransport durch Membranen, Steinkopff-Verlag, Darmstadt, 1964,
Chap. 1, pp. 615, and Sect. 6.5, pp. 7993.
9. E. A. Guggenheim, Thermodynamics, North
Holland, Amsterdam, 1967, Chap. 8, pp.
298302.
10. D. Kondepundi, I. Prigogine, Modern Thermodynamics, John Wiley & Sons, Chichester, 1998.
11. R. P. Buck, J. Membr. Sci. 1984, 17, 162.
12. J. Pellicer, S. Mafe, V. M. Aguilella, Ber.
Bunsen-Ges. Phys. Chem. 1986, 90, 867872.
13. S. Mafe, J. Pellicer, V. M. Aguilella, J. Phys.
Chem. 1986, 90, 60456059.
14. D. G. Miller, Chem. Rev. 1960, 60, 1537.
15. D. G. Miller, J. Phys. Chem. 1967, 71,
35883592.
16. N. Ibl in Comprehensive Treatise of Electrochemistry (Eds.: E. Yeager, J. OM. Bockris,
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
119
120
2 Transport Phenomena
38. M. A. Vorotyntsev, Sov. Electrochem. 1988,
24, 11501154.
39. Y. Marcus, Ion Properties, Marcel Dekker,
New York, 1997.
40. J. G. Stark, H. G. Wallace, Chemistry Data
Book, John Murray, London, 1982.
41. V. K. Indusekhar, P. Meares in Physicochemical Hydrodynamics (Eds.: D. B. Spalding), Building & Sons, Guilford, 1977,
pp. 13011313, Vol. II.
42. K. Aoki, J. Electroanal. Chem. 2000, 488,
2531.
43. K. B. Oldham, T. J. Cardwell, J. H. Santos
et al., J. Electroanal. Chem. 1997, 430, 2537,
3946.
44. M. F. Bento, L. Thouin, C. Amatore et al., J.
Electroanal. Chem. 1998, 443, 137148.
45. M. F. Bento, L. Thouin, C. Amatore, J. Electroanal. Chem. 1998, 446, 91105.
46. A. M. Bond, S. W. Feldberg, J. Phys. Chem.
B 1998, 102, 99669974.
47. C. Amatore, L. Thouin, M. F. Bento, J. Electroanal. Chem. 1999, 463, 4552.
48. K. B. Oldham, J. Phys. Chem. B 2000, 88,
47034706.
49. A. D. MacGillivray, D. Hare, J. Theor. Biol.
1969, 25, 113126.
50. A. V. Sokirko, J. A. Manzanares, J. Pellicer,
J. Colloid Interface Sci. 1994, 168, 3239.
51. H. A. Kramers, Physica 1940, 7, 284330.
52. W. E. Morf, Anal. Chem. 1977, 49, 810813.
53. U. Behn, Ann. Phys. Chem. N.F. 1897, 62,
5467.
54. M. Kh. Urtenov, Methods of Solution of the
Nernst-Planck-Poisson Equation System [in
russian], Kuban State University, Krasnodar,
Russia, 1998.
55. V. V. Nikonenko, M. K. Urtenov, Russ. J.
Electrochem. 1996, 32, 187194.
56. A. Guirao, S. Mafe, J. A. Manzanares et al.,
J. Phys. Chem. 1995, 99, 33873393.
57. P. Ramrez, S. Mafe, A. Tanioka et al., Polymer 1997, 38, 49314934.
58. J. A. Manzanares, G. Vergara, S. Mafe et al.,
J. Phys. Chem. B 1998, 102, 13011307.
59. R. Schlogl, Z. Phys. Chem. N. F. 1954, 1,
305339.
60. J. F. Brady, J. C. R. Turner, J. Chem. Soc.,
Faraday Trans. 1 1978, 74, 28392849.
61. M. Abramowitz, I. A. Stegun, Handbook of
Mathematical Functions, Dover Publications, New York, 1965.
62. J. Crank, The Mathematics of Diffusion, 2nd
ed., Oxford University Press, Oxford, 1977.
121
122
2 Transport Phenomena
2.2
Introduction
Digital Simulations
The advent of the personal computer has
led to the rapid development of digital
simulations in the eld of electrochemistry. The motivation for these advances
has been driven by many factors including: the desire to gain quantitative insights
from electroanalytical measurements, extract kinetic and mechanistic information
regarding the pathways of electrolysis reactions, and assist in the design of new
quantitative electrochemical methods.
Quantication of the related phenomena proceeds by the denition of a model
that describes the physical processes (e.g.
mass transport, chemical reactivity, etc.)
in terms of a set of mathematical expressions. Traditionally, workers have attempted to solve these relationships directly via standard mathematical methods;
however, in many cases the complexity of the problem does not permit this
analytical-solution approach. Digital simulation breaks down (discretizes) the problem into a series of steps that can be
solved sequentially by the composition of
a suitable computer program. This discretization process gives rise to a variety
of different digital strategies with which
electrochemical or related problems can
be solved.
In this overview, three digital simulation
approaches (the nite difference method
(FDM), the nite element method (FEM),
2.2.1.1
2.2 The Digital Simulation of Voltammetry under Stagnant and Hydrodynamic Conditions
O
Reactant
transport
Physical model
R
Product
transport
O+ e
R
Electrode
2c
2c
c0
= D 20 + D 20
y
t
x
Mathematical description
y
Numerical discretisation
Grid points
x
Electrode
c0
Computational prediction
Analytical solution
Computational predictions/
model validation
Distance
Fig. 1
Simulation procedure.
the form
c
= D 2 c Vx c k
t
(1)
c
cBULK
(2)
(3)
123
124
2 Transport Phenomena
(4)
c
2c
2c
=D 2 +D 2
t
x
y
(5)
2.2 The Digital Simulation of Voltammetry under Stagnant and Hydrodynamic Conditions
Finite
difference
Original
formulation
Electrode
Series of grid
points
Integration
Finite element
Weak formulation
Electrode
Mesh of interconnecting
triangular elements
Integration
Boundary
element
Inverse
formulation
Electrode
Mesh of elements around
boundary of domain
Fig. 2
125
126
2 Transport Phenomena
c
2c
=D 2
t
x
2.2.2
cti1
cti
(x)3
f (x)
3!
(7)
where f , f , and f are the rst, second,
and third derivatives of f respectively.
If x is small, then the terms from
(x)3 /3! may reasonably be neglected and
t
expressions for the concentration at (ci1
t
and ci+1 ) predicted
t
ci+1
cit
+ x
c
x
(x)2
+
2!
2c
x 2
(8)
cti+1
Bulk
solution
x
X=d
Fig. 3
(x)2
2!
f (x)
Electrode
X=0
(6)
2.2 The Digital Simulation of Voltammetry under Stagnant and Hydrodynamic Conditions
t
ci1
cit
x
c
x
(x)2
+
2!
2c
x 2
(9)
(11)
cit = cBULK
t>0
cit
cit
i=1
(14)
=0
i=0
(15)
= cBULK
i = NI
(16)
i = NI
Electrode
Bulk
solution
X=d
X=0
(a)
i = 0, N I
cb
(b)
Fig. 4
127
128
2 Transport Phenomena
and is known as the von Neumann stability criterion [17], and it has been converted
into a complete proof by John [18]. The
stability criterion can prove problematic
when a mesh contains small increments
of x. For a xed diffusion coefcient,
this requires a small value of t, thus
considerably increasing the real simulation time.
Extension of the EFD method to tackle
convection and chemical reactions provides no conceptual problems. The convective term from Eq. (1) can be cast into
the EFD form
t
cit ci1
c
= Vx
(18)
t
x
and summed with Eq. (10) to yield the EFD
expression for the convectivediffusion
equation. Similarly, kinetic complications
arising from homogeneous chemical reactions can be introduced (e.g. for a
rst-order chemical reaction)
c
= kcit
t
(19)
2.2 The Digital Simulation of Voltammetry under Stagnant and Hydrodynamic Conditions
2 c(t+t/2)
1 t
t
=
2cit + ci+1
c
2
x
2(x)2 i1
t+t
t+t
+ ci1
2cit+t + ci+1
(20)
The concentration at (t + t/2) is evaluated via linear interpolation using
(t+t/2)
ci
(cit + cit+t )
2
(21)
2( 1) t
t
t
= ci1
+
ci ci+1
t+t
ci1
(22)
2( 1) t
= c0t +
(23)
c1 c1t
2( + 1) t+t
t+t
t+t
+ ci+1
ci
ci1
2( 1) t
t
t
= ci1
+
(24)
ci ci+1
2( + 1) t+t
t+t
t+t
cN
cN I + cN
I 1
I +1
2( 1) t
t
t
cN I cN
= cN
I 1 +
I +1
(25)
Application of appropriate boundary conditions yields a set of simultaneous equations that can be solved using a variety of
methods including Gaussian elimination,
of which the Thomas Algorithm [22] is a
simplied version. Although formulation
and implementation of the CN scheme is
considerably more involved than the EFD,
it does have advantages. In principle, the
stability of the method is not restricted
by the value of , therefore permitting
the use of larger time increments within
simulations. It has been noted that the CN
method does possess some limitations that
can be overcome by modication and the
reader is directed to the text by Britz [3] for
further details. The CN technique has been
applied to simulate one-dimensional diffusion to planar and spherical electrodes and
diffusion in redox polymers [23].
c0t+t
129
2 Transport Phenomena
One reported strategy [28] breaks each individual time increment into two half steps
(in the case of a two-dimensional problem). In the rst half-time increment,
one of the dimensions is solved explicitly whilst the other is retained in implicit
form, for example,
t+t/2
t
ci,j
ci,j
t+t/2
t+t/2
t+t/2
+ ci,j 1
= x ci,j +1 2ci,j
t
t
t
2ci,j
+ ci1,j
(26)
+ y ci+1,j
where
x =
Dt
(x)2
y =
Dt
(y)2
(27)
ci, j+1
ci1, j
130
ci, j
ci+1, j
ci, j1
2.2 The Digital Simulation of Voltammetry under Stagnant and Hydrodynamic Conditions
c
c
2c
= D 2 Vx
t
y
x
(31)
Vx t
(x)
(33)
131
132
2 Transport Phenomena
cit+t ci1
D1i
2.2.2.7
(34)
where
D2i
= D exp 2
i
3
= D exp 2
i
D1i
4
i > 1 (35)
i 2 (36)
Dt
(x)2
(37)
(38)
where y and x are given by Eq. (27).
Using this approach, the system equations
were cast into matrix form to yield
[A]{p} = {q}
(39)
(40)
This approach permits the efcient iterative solution of the matrix equations using
a standard NAG routine (NAG FORTRAN
Library (D03EBF)). The approach has been
compared to the ADI and HS methods with
the authors concluding that the SIP provides a highly efcient competitor to these
strategies in both diffusion and convectivediffusion problems [75].
2.2 The Digital Simulation of Voltammetry under Stagnant and Hydrodynamic Conditions
2.2.3
Formulation
The key steps involved in the formulation
of a one-dimensional diffusional problem
under steady state conditions now follow:
2.2.3.2
2c
=0
x 2
(41)
(42)
Na =
(43)
(44)
ca
cb
N e = [ Na
(45)
Nb ]
Node
numbers
1
Electrode
2
1
3
2
4
3
ca
Fig. 6
5
4
(46)
6
5
b
cb
Element
numbers
133
134
2 Transport Phenomena
e=1
w(x)R(x) dx = 0
(48)
c e
b
N e N e
dx C D
(N )
= 0
a
x x
x
a
(54)
Noting the second term of Eq. (54) relates
to the boundary of the element and may
be applied later; evaluation of Eq. (54)
now requires the remaining integral. From
Eq. (45), we can calculate
b
(50)
Ne
2C
dx = 0
x 2
(51)
1
l
1
N
=
x
l
(49)
The weighting function effectively distributes any residual error over the space
interval dened by the bounds a and b.
As noted, a number of possible routes to
apply Eq. (49) exist, including: orthogonal
collocation, the subdomain, and Galerkin
methods [89 pg. 78, 91]. In the FEM, the
Galerkin method exploits weighting functions that take the same form as the
interpolation functions Eqs. (43, 44). The
residual takes the same form as the original differential
(53)
2C
=0
R(x)
x 2
b
e C
N
=0
x
a
(52)
N e C
dx D
x x
D
1 1
C=0
1
l 1
(55)
(56)
1
[K] =
1
e
1
1
(57)
(58)
2.2 The Digital Simulation of Voltammetry under Stagnant and Hydrodynamic Conditions
Node
numbers
1
1
2
3
4
11
c1
k1
1 + 1 1
c2
k2
c3
k3
c4
k4
1+1
1+1
1+1
c5
k5
c6
k6
5
6
Contributions
from each
element
Fig. 7
Unknown
concentration
vector
Known
concentration
vector
Applications
Early applications of the MWR/FEM in
electrochemistry were reported by Whiting and coworkers [98] and Speiser and
coworkers [99105], who exploited the
method for electroanalytical purposes. In
1984, Penczek and coworkers [91, 106]
demonstrated the application of the FEM
to nite and innite one-dimensional
diffusion problems relating to the voltammetry at mercury amalgam lms.
Early two-dimensional simulations focused on the evaluation of the current
distribution at microdisc electrodes [107,
108] and simulations of a variety of
electrode geometries [109111] including
the inuence of recessed microelectrode
congurations [112]. Work has been also
extended to cases involving coupled homogeneous kinetics, adsorption [113], and
time-dependent redox polymer electrochemistry [114].
More recently, workers have reported
the use of automatic mesh generation
routines [115] and strategies that permit
2.2.3.3
135
136
2 Transport Phenomena
w C
C b
D
=0
dx D w(x)
x a
a x x
(59)
A second integration gives the inverse
formulation
2.2.4
C b
2w
C
dx
+
D
w(x)
x 2
x a
w b
D C
= 0 (60)
x a
(61)
2.2 The Digital Simulation of Voltammetry under Stagnant and Hydrodynamic Conditions
(x ) such that
2w
= (x )
x 2
(62)
Applications
Despite the potential of the BEM to reduce
the dimensionality of the numerical solution and provide a direct measure of the
interfacial ux, it has been poorly exploited
by workers in the electrochemical eld in
comparison with the FDM and the FEM.
By comparison, heat and mass transfer
have been widely treated using the BEM
in the engineering literature [148151]. In
1984, the BEM was employed to calculate
the primary current distribution during an
electropolymerization reaction [152], the
potential of the BEM for applications to irregular geometries was also noted. Hume
and coworkers [153] used the approach to
analyze mass transport effects of electrodeposition through polymeric masks.
Of related interest, Ramachandran
and coworkers have reported a range
of papers on the inuence of mass
transport [154156], including diffusionreaction problems [157]. Further work was
reported on a variety of current distribution
problems [158161] in the early 1990s,
with a comparison of the FEM and the
BEM efciency reported by Matlosz and
coworkers [162]. A two-dimensional study
of coplanar auxiliary electrodes was reported by Mehdizadeh and coworkers [163]
and was used to assess the inuence of the
electrode conguration on uniform growth
over the cathode electrode. Electroplating
and corrosion protection in industrial cell
congurations have also been addressed
by Druesne and coworkers [164, 165].
In 1994, Barbero and coworkers [166]
extended BEM applications in the electrochemical eld to investigate convective electrodiffusion problems in charged
membranes. Qiu, Wrobel, and Power have
also outlined the application of the BEM to
transport-related problems including detailed procedures for assessing current
distribution effects controlled by combined
2.2.4.3
(63)
where A and B are arbitrary constants. Selecting two solutions of Eq. (63)
for example,
w1 = x
(64)
w2 = B
(65)
+
=0
(67)
x
x
where Ca , Cb (Ca /x) and (Cb /x) correspond to the concentrations and uxes
at the boundary points a and b respectively. Application of boundary conditions
(dening the ux or concentration at each
boundary) permits the solution for the unknown values. Of particular interest to the
electrochemist is the ability to gain a direct
measure of the ux from the simulation
unlike the FEM and the FDM that rely on
a ne mesh normal to the surface.
Extension of the approach to convective
and indeed migratory transport has been
achieved in electrochemical applications
and readers are referred to the article
by Qiu, Wrobel, and Power [147] for
further details.
137
138
2 Transport Phenomena
diffusion, convection, migration [147], examining binary and three electrolyte ion
systems [167], and in the prediction of electrode topography [168].
The BEM has been applied in scanning
electrochemical microscopy applications
by Fisher and Denuault [169] to examine
the inuence of probe and substrate
surface topography. In addition, timedependent phenomena have been assessed
in oil droplets [170, 171] and a range of
microelectrode geometries using the dual
reciprocity method (DRM) [172] closely
related to the BEM [173].
2.2.5
References
1. S. W. Feldberg in Electroanalytical Chemistry
(Ed.: A. J. Bard), Marcel Dekker, New York,
1969, pp. 199296.
2. A. J. Bard, L. R. Faulkner, Electrochemical
Methods, John Wiley & Sons, New York,
1980.
3. D. Britz, Digital Simulation in Electrochemistry, 2nd ed., Springer-Verlag, Berlin, 1988.
4. D. K. Gosser, Cyclic Voltammetry. Simulation and Analysis of Reaction Mechanisms,
Wiley-VCH, Weinheim, Germany, 1993.
5. http://www.bath.ac.uk/chsacf.
6. P. A. Ramachandran, Boundary Element
Methods in Transport Phenomena, 1st ed.,
Elsevier, Southampton, UK, 1994.
7. L. F. Richardson, Philos. Trans. 1911, 210,
307357.
8. R. Courant, K. Friedrichs, H. Lewy, Math.
Ann. 1928, 100, 32.
9. H. W. Emmons, Q. Appl. Math. 1944, 2,
173195.
10. J. E. B. Randles, Trans. Faraday Soc. 1948,
44, 327338.
11. T. Joslin, D. Pletcher, J. Electroanal. Chem.
1974, 49, 171.
12. S. W. Feldberg, J. Electroanal. Chem. 1981,
127, 1.
13. L. K. Bieniasz, J. Electroanal. Chem. 1993,
360, 119138.
14. J. B. Flanagan, K. Takahashi, F. Anson, J.
Electroanal. Chem. 1977, 81, 261.
15. J. W. Dillard, J. A. Turner, R. A. Osteryoung, Anal. Chem. 1977, 49, 1246.
16. R. Seeber, S. Stefani, Anal. Chem. 1981, 53,
10111016.
17. J. von Neumann, R. D. Richtmyer, J. Appl.
Phys. 1950, 53, 10211099.
18. F. John, Common Pure Appl. Math. 1952, 5,
155211.
19. B. Speiser in Electroanalytical Chemistry
(Eds.: A. J. Bard, I. Rubinstein), Marcel
Dekker, New York, 1996, pp. 1108.
20. J. Crank, P. Nicolson, Proc. Cambridge Philos. Soc. 1947, 43, 5067.
21. J. Crank, The Mathematics of Diffusion,
Clarendon Press, Oxford, 1956.
22. L. Lapidus, G. F. Pinder, Numerical Solution
of Partial Differential Equations in Science and
Engineering, John Wiley & Sons, New York,
1982.
23. M. Storzbach, J. Heinze, J. Electroanal.
Chem. 1993, 346, 127.
2.2 The Digital Simulation of Voltammetry under Stagnant and Hydrodynamic Conditions
24. D. W. Peaceman, H. H. Rachford, J. Soc.
Ind. Appl. Math. 1955, 3, 28.
25. J. Heinze, M. Storzbach, Ber. Bunsen-Ges.
Phys. Chem. 1986, 90, 10431048.
26. J. Heinze, M. Storzbach, J. Mortensen, J.
Electroanal. Chem. 1984, 165, 6170.
27. J. Heinze, M. Storzbach, J. Mortensen, J.
Electroanal. Chem. 1988, 240, 2743.
28. M. K. Jain, Numerical Solution of Differential
Equations, Wiley Eastern Limited, New
Delhi, 1979.
29. J. Heinze, J. Electroanal. Chem. 1981, 124,
73.
30. D. J. Gavaghan, J. Electroanal. Chem. 1998,
456, 135.
31. A. J. Bard, F. R. F. Fan, F. F. Fu-Ren et al.,
Anal. Chem. 1989, 61, 132.
32. C. Demaille, P. R. Unwin, A. J. Bard, J.
Phys. Chem. 1996, 100, 14 13714 143.
33. J. L. Amphlett, G. Denuault, J. Phys. Chem.
B 1998, 102, 99469951.
34. A. J. Bard, M. V. Mirkin, P. R. Unwin et al.,
J. Phys. Chem. 1992, 96, 18611868.
35. P. R. Unwin, A. J. Bard, J. Phys. Chem. 1991,
95, 78147824.
36. J. V. Macpherson, P. R. Unwin, J. Phys.
Chem. 1996, 100, 19 47519 483.
37. J. V. Macpherson, P. R. Unwin, J. Phys.
Chem. 1994, 98, 17041713.
38. J. Booth, R. G. Compton, J. A. Cooper et al.,
J. Phys. Chem. 1995, 99, 10 94210 947.
39. A. R. Gourlay, J. Inst. Math. Appl. 1970, 6,
375.
40. D. Shoup, A. Szabo, J. Electroanal. Chem.
1982, 140, 237245.
41. D. Shoup, A. Szabo, J. Electroanal. Chem.
1984, 160, 117.
42. D. Shoup, A. Szabo, J. Electroanal. Chem.
1986, 199, 437441.
43. P. Pastore, F. Magno, J. Lavagnini et al., J.
Electroanal. Chem. 1991, 301, 113.
44. H. R. Corti, D. L. Goldfarb, A. S. Ortiz et al.,
Electroanalysis 1995, 7, 569573.
45. S. Moldoveanu, J. L. Anderson, J. Electroanal. Chem. 1984, 175, 67.
46. J. L. Anderson, S. Moldoveanu, J. Electroanal. Chem. 1984, 179, 107119.
47. A. C. Fisher, R. G. Compton, J. Phys. Chem.
1991, 95, 75387542.
48. A. C. Fisher, R. G. Compton, J. Appl. Electrochem. 1992, 22, 3842.
49. R. G. Compton, B. A. Coles, A. C. Fisher, J.
Phys. Chem. 1994, 98, 24412445.
139
140
2 Transport Phenomena
75. J. A. Alden, R. G. Compton, J. Electroanal.
Chem. 1996, 402, 110.
76. B. A. Finlayson, Br. Chem. Eng. 1969, 14,
5357.
77. R. Courant, Bull. Am. Math. Soc. 1943, 49,
123.
78. J. H. Argyris, Aircraft Eng. 1954, 26, 347.
79. M. J. Turner, R. W. Clough, H. C. Martin
et al., J. Aerosol Sci. 1956, 23, 805.
80. R. W. Clough, Proceedings of 2nd ASCE
Conference on Electronic Computation,
Pittsburg, Pa., 1960.
81. C. V. Girijavallabhan, L. C. Reese, J. Soil
Mech. Proc. ASCE 1968, 94, 473496.
82. S. S. Rao, Proc. Int. Symp. Disc. Methods Eng.
Milan, 1974, pp. 512525.
83. S. G. Ravikumaur, K. N. Seetharamu, P. A.
Aswathanarayana, Proc. Int. Conf. On Finite
Elements in Computational Mechanics 1985,
861, Vol. 2.
84. P. P. Silvester, R. L. Ferrari, Finite Elements
for Electrical Engineers, 3rd ed., Cambridge
University Press, Cambridge, 1996.
85. O. C. Zienkiewicz, R. L. Taylor, The Finite
Element Method, 5th ed., ButterworthHeinemann, Oxford, 2000.
86. S. S. Rao, The Finite Element Method in
Engineering, Pergamon Press, Oxford, 1982.
87. G. Dhatt, G. Touzot, The Finite Element
Method Displayed, John Wiley & Sons, New
York, 1984.
88. C. Cuvelier, A. Segal, A. A. van Steenhoven,
Finite Element Methods and Navier-Stokes
Equations, D. Reidel, Dordrecht, The
Netherlands, 1986.
89. J. Villadsen, M. L. Michelsen, Solution of
Differential Equations by Polynomial Approximation, Prentice Hall, Englewood Cliffs, NJ,
1978.
90. G. Fairweather, Finite Element Galerkin
Methods for Differential Equations, John
Wiley & Sons, New York, 1978.
91. M. Penczek, Z. Stojek, J. Osteryoung, J.
Electroanal. Chem. 1984, 170, 99108.
92. E. Isaacson, H. B. Keller, Analysis of Numerical Methods, John Wiley & Sons, New York,
1966.
93. L. W. Johnson, R. D. Riess, Numerical Analysis, 2nd ed., Addison-Wesley, Reading,
Mass., 1982.
94. W. H. Press, S. A. Teukolsky, W. T. Vetterling et al., Numerical Recipes in Fortran 77:
The Art of Scientic Computing, 2nd ed.,
95.
96.
97.
98.
99.
100.
101.
102.
103.
104.
105.
106.
107.
108.
109.
110.
111.
112.
113.
114.
115.
116.
117.
118.
2.2 The Digital Simulation of Voltammetry under Stagnant and Hydrodynamic Conditions
119. B. A. Coles, R. G. Compton, C. M. A. Brett
et al., J. Electroanal. Chem. 1995, 381,
99104.
120. Q. Fulian, N. P. C. Stevens, A. C. Fisher, J.
Phys. Chem. B 1998, 102, 37793783.
121. K. A. Gooch, N. A. Williams, A. C. Fisher,
Electrochem. Commun. 2000, 2, 5155.
122. Q. Fulian, A. C. Fisher, D. J. Riley, Electroanalysis 2000, 12, 503508.
123. F. L. Qiu, K. A. Gooch, A. C. Fisher et al.,
Anal. Chem. 2000, 72, 34803485.
124. N. P. C. Stevens, K. A. Gooch, A. C. Fisher,
J. Phys. Chem. B 2000, 104, 12411248.
125. J. Josserand, J. Morandini, H. J. Lee et al., J.
Electroanal. Chem. 1999, 468, 4252.
126. R. Ferrigno, H. H. Girault, J. Electroanal.
Chem. 2000, 492, 16.
127. H. J. Lee, C. Beriet, R. Ferrigno et al., J.
Electroanal. Chem. 2001, 502, 138145.
128. Y. Lee, S. Amemiya, A. J. Bard, Anal. Chem.
2001, 73, 22612267.
129. A. Beckmann, B. A. Coles, R. G. Compton
et al., J. Phys. Chem. B 2000, 104,
764769.
130. B. A. Coles, R. G. Compton, M. Suarez
et al., Langmuir 1998, 14, 218225.
131. S. Coen, D. K. Cope, D. E. Tallman, J. Electroanal. Chem. 1986, 215, 2948.
132. M. Fleischmann, J. Daschbach, S. Pons, J.
Electroanal. Chem. 1989, 263, 189203.
133. J. Daschbach, S. Pons, M. Fleischmann, J.
Electroanal. Chem. 1989, 263, 205224.
134. M. Rudolph, J. Electroanal. Chem. 1990, 292,
17.
135. U. Kalapathy, D. E. Tallman, D. K. Cope, J.
Electroanal. Chem. 1990, 285, 7177.
136. D. K. Cope, C. H. Scott, D. E. Tallman, J.
Electroanal. Chem. 1990, 285, 4969.
137. D. K. Cope, D. E. Tallman, J. Electroanal.
Chem. 1990, 285, 7992.
138. D. R. Baker, M. W. Verbrugge, J. Newman,
J. Electroanal. Chem. 1991, 314, 2344.
139. M. V. Mirkin, A. J. Bard, J. Electroanal.
Chem. 1992, 323, 127, 2951.
140. L. K. Bieniasz, Comput. Chem. 1992, 16,
311317.
141. L. K. Bieniasz, J. Electroanal. Chem. 1993,
347, 1530.
142. B. Pillay, J. Newman, J. Electrochem. Soc.
1993, 140, 414420.
143. P. K. Banerjee, R. Buttereld, Boundary Element in Engineering Science, McGraw-Hill,
New York, 1981.
141
142
2 Transport Phenomena
169. F. Qiu, A. C. Fisher, G. Denuault, J. Phys.
Chem. B 1999, 103, 4387, 4393.
170. F. L. Qiu, J. C. Ball, F. Marken et al., Electroanalysis 2000, 12, 10121016.
171. J. C. Ball, F. Marken, F. L. Qiu et al., Electroanalysis 2000, 12, 10171025.
172. F. L. Qiu, A. C. Fisher, Electrochem. Commun. 2000, 2, 738742.
173. D. Nardini, C. A. Brebbia in Boundary
Element Methods in Engineering (Ed.:
C. A. Brebbia), Springer-Verlag, Berlin,
1982, pp. 312326.
174. B. Marner, W. Schmickler, J. Electroanal.
Chem. 1986, 214, 589596.
175. B. Aurian-Blajeni, M. Kramer, M. Tomkiewicz, J. Phys. Chem. 1987, 91, 600605.
176. T. Pajkossy, L. Nyikos, Electrochim. Acta
1989, 34, 171179.
177. F. Sagues, J. M. Costa, J. Chem. Educ. 1989,
66, 502506.
178. S. Zalis, N. Fanelli, L. Posps il, J. Electroanal. Chem. 1991, 314, 111.
147
3.1
Introduction
148
3.1.2
NMR amplitude
(a)
(b)
800
600
400
200
d
[ppm]
149
150
K T1 T =
e
n
2
h
4kB
=S
(2)
Hhf
= B Hhf D(EF )
B
(3)
151
152
Cmatching
Ctuning
Counter
electrode
NMR coil
Reference
Lcounterelectrode
Ccounter
Lref.
Working
electrode
Lwk.
Cwk.
Cref.
Potentiostat
Schematic diagram (left) and picture (right) of an EC-NMR probe and its circuitry,
showing the interface between NMR and electrochemistry.
Fig. 2
NMR of Carbon-supported
Electrocatalysts and a Potential-scangenerated Sintering Effect
At the beginning of the 1980s, Slichter
and coworkers discovered several unique
features of the 195 Pt NMR of oxidesupported small platinum particles [28].
They found that the overall 195 Pt NMR
lineshape was extremely broad, extending
downeld some 4 kG from the position
of bulk platinum (1.138 G kHz1 ), and
contained a feature on the low-eld side
(1.089 G kHz1 ), which arose from the
oxidized Pt surface atoms. Later on, van
3.1.3.1
F
E
A
D
B
C
153
NMR Amplitude
154
1.08
1.10
1.12
1.14 1.08
1.10
1.12
1.14
Field/frequency
[G kHz1]
Fig. 4 Typical point-by-point 195 Pt NMR spectra showing electrochemical
cleaning, sintering by potential cycling, and a layer-model analysis of the
2.5-nm sample: (a) as-received catalyst; (b) electrochemically cleaned in
0.5 M H2 SO4 by holding electrode potential at 0.45 V versus reversible
hydrogen electrode (RHE); (c) cleaned by extensive potential cycling; and
(d) layer-model deconvolution of spectrum (b). The solid line in (b) is the
result of the simulation.
155
Pt/clean
Pt/Ru
Pt/CO
Pt/O
Pt/H
Pt/CN
Pt/S
NMR amplitude
156
1.08
1.10
1.12
1.14
Field/frequency
[G kHz1]
Superimposition of point-by-point, 8.47 T 195 Pt NMR spectra of
a 2.5 nm, carbon-supported Pt electrocatalyst without and with
different chemisorbed ligand: Ru, CO, O, H, CN , and S. The spectra
were normalized by equalizing the amplitude at 1.131 G kHz1
(indicated by the arrow). The invariance of signals beyond
1.131 G kHz1 provides experimental conrmation of the
FriedelHeine invariance theorem. The surface peaks range over
11 000 ppm. (Reproduced with permission from Ref. [10], Copyright
by 2000 American Chemical Society.)
Fig. 5
NMR amplitude
Pt/clean
Simulation
Pt/CO
Pt/CN
Pt/Ru
Pt/S
Pt/H
Pt/O
1.08 1.10 1.12 1.14 1.08 1.10 1.12 1.14 1.08 1.10 1.12 1.14 1.08 1.10 1.12 1.14
Field/frequency
[G kHz1]
The NMR layer-model simulations of the
spectra shown in Fig. 5 and the small dots
are the variances between the experimental and
tted values. (a) clean-surface Pt; (b) Simulation
of (a); (c) Pt with adsorbed Ru; (d) Pt with
Fig. 6
195 Pt
157
d ( 103)
[ppm]
158
10
Subsurface
20
Pt bulk shift
30
1.5
2.0
2.5
3.0
3.5
AllredRochow electronegativity
Correlation between surface/subsurface frequency shifts
(with respect to the Pt NMR reference H2 PtCl6 ) and the
AllredRochow electronegativity. The dashed horizontal line
indicates the Knight shift of bulk platinum atoms. The solid
straight lines are linear ts to the surface and subsurface shifts
as a function of the electronegativity. Both have R2 values of
ca. 0.92. (Reproduced with permission from Ref. [10], Copyright
by American Chemical Society 2000.)
Fig. 7
0.9
0.8
0.7
0.6
H
Li
0.5
K
0.5
1.0
1.5
2.0
2.5
3.0
3.5
4.0
Electronegativity
Solid curve showing the relationship between the
healing length and electronegativity as determined from
Eq. (7). For comparison, we also show the three experimental
points based on previous observations. See the text for
details. (Reproduced with permission from Ref. [10],
Copyright by American Chemical Society 2000.)
Fig. 8
159
Tab. 1
Sample
Clean-surface EF -LDOS
(Ry1 atom1 )
0
14.8
17.7a
18.6
0
24
160
330
351a
383
203
500
PtCO7
COPt/oxides (dry)
COPt/carbon (wet)
COPt/carbon (wet)
PdCO7
COPd/oxides (dry)
a For
an 8.8-nm sample.
500
400
Shift
[ppm]
160
300
200
10
15
20
25
30
Fig. 9
Tab. 2
Size
(nm)
5 -D(EF )
(Ry
molecule)1
2 -D(EF )
(Ry
molecule)1
IR
frequency
(cm)1
8.8
3.9
3.2
2.5
2.0
0.6
0.6
0.6
0.7
0.6
6.4
6.6
6.5
6.8
7.3
2044
2043
2044
2038
2028
161
1/T1
(s1)
162
Pt/13CO
Carbon support
0
0
200
Temperature
[K]
800
600
400
200
200
Fig. 10
and backdonation, which are the integrals of the LDOS from the bottom of
the conduction band to the Fermi level.
Nevertheless, as can be seen in Table 2,
the variation in the total EF -LDOS between different samples are mainly due to
changes in the 2 -D(EF ). This suggests
that the changes in chemical properties
of adsorbed CO are mainly determined
by variations in backbonding. In particular, if the Fermi level cuts the tail of the
2* band as it rises, then the increase in
the 2 -D(EF ) would indicate an enhancement in backdonation, and a decrease in
the corresponding vibrational CO stretch
frequency.
In situ subtractively normalized interfacial Fourier transform infrared reectance
spectroelectrochemistry (SNIFTIRS) studies conrm this prediction [37]. They also
2050
(cm1)
2040
2030
2020
6.0
6.5
7.0
7.5
8.0
D2p*(Ef)
[Ry1 molecule1]
spectroscopy and solid-state NMR, in providing insights into the details of surface
chemistry of electro-chemisorbed CO on
platinum. While SNIFTIRS a major research technique of the interfacial electrochemical community provides important qualitative information regarding the
bond strengths of adsorbed CO, it offers
somewhat less insight into the electronic
relocations that occur within the bond, the
electron distribution as a function of sample composition and morphological detail.
For instance, how the bond changes as a
function of particle size is difcult to learn
using infrared alone. However, a 13 C NMR
analysis of the 5 - and 2 -EF -LDOS at
the carbon atom offers a quantitative description of metalCO bonding in terms
of changes in the EF -LDOS. The joint
SNIFTIRS/NMR approach provides new
information on the electronic structure
of the metalsolution interface, which
is relevant to the characterization of industrial fuel cell catalysts, as well as
other interfacial electrochemical systems
of practical importance, by combining
bond strength (IR) and quantitative EF LDOS results (NMR).
Surface Diffusion
The rationale behind using NMR to study
surface diffusion is that the correlation
time for diffusion, , can be related to the
temperature dependence of the NMR relaxation rates observed for diffusing atoms
or molecules [12]. Under certain circumstances, such a relation can be expressed
analytically, permitting a quantitative analysis of the data in dynamic and thermodynamic terms.
When using the temperature dependence of the nuclear spinspin or nuclear
spinlattice relaxation rate to study molecular motion, as is the case with the
surface diffusion we are dealing with here,
there exist so-called strong and weak
collision limits. Different mathematical relationships are needed to describe these
limits. Consider the nuclear spinspin
relaxation rate (1/T2 ) as measured by a conventional Hahn-echo pulse sequence, and
suppose that is the amplitude of the
local eld uctuation responsible for relaxation. Also assume that is the correlation
time for the motion, say a jump, which
causes the local eld to uctuate. The
strong collision limit is dened such that
3.1.3.5
163
164
1/. At t = 0, that is, just after the
rst /2 pulse, which ips all nuclear spins
into the xy plane, all spins are in phase, and
they retain their phase memory until t = ,
when the collision changes the direction of
the local eld. Since is sufciently long
to dephase the spins, they lose all phase
memory before the next collision. Therefore, 1/T2 = 1/ . In contrast, in the weak
collision limit, 1/. Since is relatively short, insufcient dephasing builds
up between two consecutive jumps. When
the spin jumps randomly, many jumps are
needed in order to accumulate sufcient
dephasing, and one has 1/T2 = ()2 .
In summary then, we have for these two
limiting cases:
1
1
1
= for
T2
(8)
1
1
= ()2 for
T2
(9)
and
(10)
RL
+aT +
1600
Pt/13CN
E = 10.3 kcal/mol
1400
a = 0.8 s1 K1
w = 966 Hz
N = 1222
1200
1/T2
(s1)
1000
2500
Pt/13CO
E = 7.8 kcal/mol
2000
a = 5.4 s1 K1
w = 1053 Hz
N = 977
1500
1000
500
100
200
300
Temperature
[K]
165
2.0
Pt-13CNCH3
1.5
1/T1
(s1)
166
1.0
0.5
0.0
0
50
100
150
200
Temperature
[K]
250
300
500
400
[ppm]
13CO, 13CN
Shift
450
13
CO
13CN
350
200
150
0.8
0.4
0.0
0.4
Potential
[V]
167
168
Future Perspectives
Acknowledgments
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
169
170
3.2
Introduction
This section outlines the theoretical concepts of EPR. The theory is similar to
the related technique of NMR described
in Volume 2 Chapter 3.1. In the following, a basic knowledge of quantum
mechanics will be assumed. Throughout
the following, the symbol B refers to the
magnetic ux density (or magnetic induction eld); the symbol H being reserved
for the magnetic eld strength. In SI units,
H = 4 107 B.
EPR Theory
The origin of EPR spectroscopy lies in the
fact that electrons have both electrostatic
charge and spin angular momentum. The
former is observed by electron-deection
from negatively charged surfaces, in for
example, a cathode ray oscilloscope; the
latter is easily veriable in the classic SternGerlach experiment [29, 30] in
which a beam of S-state silver atoms is
observed to split into two separate beams
3.2.2.1
(1)
(3)
171
172
Fig. 1
1h
2
Ms = + 1
2
Ms = 1
2
E=
ge e Ms Bz
h
Ms = + 1
2
2phn = gegeBZ
B=0
E
Ms = 1
2
BZ
B
dA
dB
Bpp
(5)
ge e Bz
2 h
(6)
ge e (1 )Bz
2 h
(7)
(8)
173
174
2T2
1 + 4 2 T22 ( )2
(10)
f () = T2 2 exp[2 2 T22 ( )2 ]
(11)
where the symbol T2 is dened later. The
Gaussian lineshape is observed when the
line is the superposition of a large number of unresolved individual components,
such a line being referred to as inhomogeneously broadened, and is commonly
encountered in solid-state EPR spectra.
More important is the Lorenztian lineshape, which is often observed for radicals
in solution, when complications due to unresolved hyperne splittings do not occur.
3.2.2.3.1 Spin-lattice Relaxation The transitions between eigenstates induced by
the electromagnetic radiation necessarily
involves a perturbation of the Boltzmann
Distribution Law,
(E2 E1 )
n2
= exp
(12)
n1
kT
(13)
absorption lines occur. The former situation is often encountered in the EPR
spectra of transition metal ions, whilst the
latter is usually found for organic radicals
in solution.
Another form of spinlattice relaxation
occurs when there is a quadrupolar nucleus (vide infra) present in the paramagnetic species. Associated with this nucleus
is an electric quadrupolar eld, which
will interact with gradients in the electric
component of the microwave radiation.
However, as this electric quadrupolar interaction is generally weak, it does not
contribute much towards the relaxation.
Spinspin Relaxation The second major nonradiative phenomenon is
termed spinspin relaxation and is characterized by a relaxation time T2 . The
processes that contribute to T1 also inuence T2 , which controls the linewidth
(vide supra). However, the random forces
that modulate the electron spin energy
levels at low frequencies (below 1010 Hz)
contribute to T2 , whereas they have no
inuence upon T1 . Again, the source of
the uctuating magnetic eld lies in the
anisotropic magnetic interaction within
the molecule, mainly due to the anisotropy
in the g-tensor and to the dipolar hyperne coupling with magnetic nuclei. This
has assumed that the radical (or parent
species) concentration is not too high;
otherwise, spin exchange occurs as a result of radicalradical collisions, thereby
causing a broadening of the EPR lines.
The spinspin process involves the interchange of the spins of the colliding radicals. At very high radical concentrations,
the spin sees an average environment as
it moves from radical to radical, and so the
hyperne splitting is lost and a single line
results. This is the reason solid state EPR
spectra are generally of a single line, as
3.2.2.3.2
175
176
Fig. 3
5G
N = gN N I
(14)
Tab. 1
Electron
1H
2D
6 Li
7 Li
12 C
13 C
14 N
15 N
19 F
23 Na
31 P
Natural
abundance
%
Spin
angular
momentum
99.98
0.02
7.43
92.5
98.9
1.1
99.63
0.37
100
100
100
1/2
1/2
1
1
3/2
0
1/2
1
1/2
1/2
3/2
1/2
g-value
2.0023
5.585
0.857
0.8219
2.1707
0
1.405
0.404
0.567
5.257
1.477
2.263
The anisotropic term, (E)anisotropic , represents the dipolar interaction that depends
upon the relative positions of the magnetic moments of the unpaired electron
and the nucleus (e and N ). In single crystals, this gives helpful information
on the crystal geometry; in polycrystalline
substances, it causes line broadening.
In liquid samples, however, since the
molecules are constantly tumbling, the
magnetic dipoledipole interactions average to zero, except for a small amount
that is dependent upon the liquid viscosity
and hence gives rise to line broadening.
The hyperne structure due to radicals
in solution, therefore, results from an
isotropic term, (E)isotropic , which is a
consequence of a nonclassical interaction,
known as the Fermi-Contact interaction.
This interaction arises because of the presence of a small, but nevertheless nite
spin density at the position of the nucleus.
In a strong magnetic eld applied in the
z-direction, this term, referred to as the
coupling constant, a, is proportional to the
square of the electronic wave function at
the nucleus,
8
ge gN e N |(0)|2
(16)
a=
3
Since only s-orbitals have a nite probability density at the nucleus (other orbitals all have nodes in their wave functions there), the contact interaction can
only occur when the electron occupies a
molecular orbital in which there is some
s-orbital character.
Thus, for each nuclear spin state, a separate transition in the EPR spectrum will
I = 1/2
I=1
MI
MI
+1
0
1
+1/2
1/2
MS
+1/2
MS
+1/2
2phn
1/2
1/2
+1
0
1
1/2
+1/2
a
B
(a)
MI +1/2 1/2
(b)
MI +1
177
178
MI = 0
(17)
MS
M1
M2
M 1
+1/2
+1/2
1/2
1/2
+1/2
1/2
+1/2
1/2
+1
0
1
M 1
I=1
+2
+1
0
1
2
MS
+1/2
+1/2
2phn
1/2
1/2
1/2
1/2
+1/2
+1/2
1/2
+1/2
1/2
+1/2
1
0
+1
+2
+1
0
+1
+2
M1 +1 0
M1 +2 +1 0 1 2
Splitting of an EPR signal by the hyperne interaction between the unpaired electron
and two equivalent nuclei.
Fig. 5
I = 1/2
1
1
1
1
1
1
2
3
4
5
6
n=0
1
1
10
15
1
3
1
1
3
1
5
15
4
1
10
16
19
16
10
2
1
10
20
I = 1:
5
1
Fig. 6 EPR line intensities for the coupling between two equivalent nuclei when I = 1/2
(Pascals Triangle) and I = 1.
179
180
Fig. 7
Canonical structures of
the CH fragment.
Fig. 8
(19)
R
R
q
C
a(H) = A + B cos2
(20)
Waveguide
Klystron
Sample
tube
Phase-sensitive
detection
Modulation coils
Signal out
Electromagnet
Fig. 10
Magnetic field
modulation (100 kHz)
181
182
Sample tube
Magnetic vector
of radiation
Electric vector
3.2.3.1.1 Cell Design This in situ stationary solution cell developed by Bard
and coworkers [4345], is illustrated in
Fig. 12. It is similar to the in situ at cell
pioneered by Adams and coworkers [17].
It consists of a platinum mesh working
electrode (1 mm grid size) placed in a
at cell, with a U-shaped tungsten rod
counter electrode along the cell edges, and
a silver wire quasi-reference electrode located in the at cell above the working
electrode. Unlike the Adams cell design,
ohmic drop problems were removed by
the location of all three electrodes within
the EPR cavity for electrolyte solutions in
DMF. However, a careful choice of supporting electrolyte was necessary; ClO2
radicals may be produced at the counter
electrode when reductions occurred at the
working electrode in the presence of perchlorate background electrolytes. The use
Pt wire
Reference
Ag wire
Teflon sleeve
Counter electrode
Working
electrode
Glass wall
Schematic diagram of
the BardGoldberg cell.
Fig. 12
183
184
Fig. 13
0.4 mm od
0.3 mm id
5 mm
Working
electrode
Loop-gap
resonator
0.25 mm
0.8 mm od
0.6 mm id
A schematic illustration
of the active region of
Allendoerfers electrochemical
loop-gap resonator geometry.
Fig. 14
1 mm
185
186
Fig. 15
Quartz
Wire
Inactive volume
1 mm
Fig. 16
Reference electrode
Luggin capillary
Quartz tube
Auxiliary electrode
Working electrode
Teflon flow
baffles
Flow
Flow
187
188
iii
iv
ii
R
iii
Z
i
ii
ii
ii
(a)
TE102
X40
(b)
Loop gap
4G
(c)
radicals for EPR spectroscopy, this twoelectrode arrangement was adequate, with
the platinum electrode acting as a quasireference electrode in aqueous chloride.
However, for the purposes of in situ electrochemical EPR, this arrangement was
rened to accommodate a three-electrode
Fig. 18
189
190
x = 0
(h y)2
1
h2
1/3
2 c(x, y)
c(x, y)
x
=0
y 2
x
(22)
(23)
(25)
Fig. 19
Cavity length
y=0
xu
c(x, y)dy dx
(27)
0
This equation can be solved either analytically [55, 58] or numerically [6568] using
the Backward Implicit approach [6971]
for the case of macroelectrodes at steady
y = 2h
Flow
xc xe
xd
MK =
SVf
(29)
i
where K is the normalized rate constant
for the radical decay,
1/3
4h4 2 d 2
(30)
K=k
9Vf2 D
For a stable radical, M0 is constant.
Figure 20 shows how the ratio MK /M0
191
1.0
MK
M0
192
0.5
0.0
0.0
2.5
5.0
7.5
10.0
K
The variation of the EPR detection efciency, MK , with the
normalized rate constant, K, for different sized electrodes (+, 2 mm;
, 3 mm; O, 4 mm) located centrally in the cavity.
Fig. 20
M 0ref
MKref
1.0
0.5
0.0
0.0
0.5
1.0
xu/I
ref
The variation of Mref
k /M0 for a 5 mm electrode with position of the
electrode in the cavity for different rate constants (K = 0.0, ; K = 1.0, O;
K = 5.0, ; K = 7.5, ).
Fig. 21
(31)
S( ) = S( )
1 bn (K + p)n Jn
2n 5
p
n=0 ?
+
3
3
L1
bn K n J n
2n 5
n=0 ?
+
3
3
(32)
1/3
2/3
9Vf D
= t 4 2 2
(33)
4h d
L1 represents the inverse Laplace Transformation (variable p), and ?(x) is the
Gamma function [73]. The coefcients bn
are given by
1
1
?
n
+
3(n2)/3
3
3
bn =
n!
n=0
2
(n + 1)
(34)
sin
3
193
x xe
x
U
d
(35)
where U(x) is the unit step function.
Theoretical growth transients for various
values of normalized rate constant are
shown in Fig. 22 for an electrode located at
the cavity center. For fast mass transport
(or sluggish kinetics), the transient is
governed by both the mass transport of
material throughout the cavity and by
the kinetic decay or the growth of the
radical. As the normalized rate constant,
K, increases (or as the mass transport
decreases), the time taken for the transient
to reach a steady state value decreases,
and the transient changes shape until they
attain a simple exponential form, which
0.5
0.0
0.5
1.0
1.5
t
Calculated transients for a 4 mm electrode located at the cavity center
for different values of K (K = 0.0, O; K = 0.1, ; K = 1.0, ; K = 10.0, ;
K = 20.0, ).
Fig. 22
(36)
1.0
S(t)/S(t )
194
for the reduction of nitromethane in aqueous alkaline solution [78]. One last point to
note is that since the cell is constructed of a
transparent substance, silica, illumination
of the electrode is possible, permitting photoelectrochemical EPR [55, 6567, 7981].
3.2.3.4
Cell
3.2.3.4.1 Cell Design Albery and coworkers [914] used tubular electrodes for ex
situ electrochemical EPR experiments.
The tubular electrode is equivalent to the
channel electrode in all respects, except
that the cross section is circular rather
than rectangular [82, 137]. Like the laterdeveloped channel ow cell, this setup
(shown in Fig. 23) permits the interrogation of electrode reaction mechanisms of
relatively long-lived radical species, [914]
since the convective-diffusion equations
are mathematically well dened, which at
steady state are given by Eq. (37)
2
2 c 1 c
0 1
+
r 2
r r
(37)
where 0 is the velocity of ow at the center
of the tube, r0 is the radius of the tube, r is
the radial distance from the center of the
r
r0
c
=D
x
xe
195
d
50
S/%Mn2+
196
25
a
c
b
0
0.00
0.25
0.50
0.75
1.00
x/l
Variation of the sensitivity of the cavity with distance through
the cavity. Curve a is obtained with no electrode inside the cavity;
Curve b is obtained with a tube electrode consisting of a complete
annulus at the cavity center. Curves c and d are obtained with an
electrode at the center of the cavity consisting of a half annulus and an
annulus with a missing sector, respectively.
Fig. 24
albeit reduced by a factor of two, but nevertheless adequate for EPR measurements.
Curve d in Fig. 24 shows the sensitivity
when an annular electrode with a sector
subtending 10 has been removed. The
sin2 function is now distorted, but the sensitivity at the electrode is enhanced by a
factor of three when compared to Curve a.
The difference between a complete annular electrode and a part-annular one is
that in the former, the magnetic component of the standing microwave eld
induces eddy currents in the complete circle and therefore is unable to penetrate to
the electrolyte solution. These eddy currents are minimized using an incomplete
annulus. In the case of the semiannular
electrode, there is a greater loss of sensitivity, but the standing microwave eld is
preserved.
Application For a semiannular
tubular electrode, the limiting current is
given by the appropriate form of the Levich
3.2.3.4.2
equation [10],
2/3
1/3
(38)
i
2/3
nF D 1/3 Vf
(39)
3.2.3.5.1 Cell Design Waller and Compton [85] effectively mimicked the Allendoerfer coaxial design (vide supra), whilst
simultaneously maintaining the mathematically well-dened laminar ow of the
channel ow cell. This improved cell for
electrochemical EPR [85] allowed an improvement in the channel cell regarding
lifetimes of radicals amenable to study,
whilst retaining the hydrodynamic ow
that is essential for the investigation of
electrode reaction mechanisms.
197
198
35
A
B
115
40
D
E
40
35
15
55
F
G
11
2/3
(41)
2/3
Vf
log10 s/i
[a.u.]
0.0
0.5
1.0
Fig. 26
3.0
2.0
199
4
nF 0 h cbulk
3
(42)
3.2.3.6
Cell
3.2.3.6.1 Cell Design Bond and coworkers have designed various cells suitable
for in situ electrochemical EPR [9094].
These cells generally have a small volume and permit the recording of variable
temperature EPR. In this section, only
the most recently designed cell [94] will
be discussed.
Figure 27 illustrates the all-glass cell
used for electrochemical EPR over a wide
temperature range. The reference electrode (either an Ag|AgCl or platinum)
is xed in the center of a Pt workingelectrode coil, the latter extending to the
bottom of the quartz tube. These electrodes have diameters of less than 70 m,
so as to minimize the amount of metal
present in the EPR cavity; the counter electrode is situated 15 mm above the working
electrode outside the cavity, also permitting interference-free measurements from
species diffusing from the counter electrode during electrolysis. With this design,
Cu rods
Soldered connections
Nylon support
PVC sleeves
149
125
150
200
O ring
Average filling level
(approx. 0.25 ml)
Counter electrode
2
4
given in mm.
201
202
Fig. 28
Solution from
reservior
Reference
electrode
Altex T-piece
To
counter electrode
Glass
tubing
od = 0.4 cm
id = 0.2 cm
Jet
17.5 cm
Mercury working
electrode
Copper
wire
Solution
intlet
RE
3.2.3.7.2 Application Albery [99] and Matsuda [100] have derived an approximate
equation for the mass transport-limited
current for an n-electron transfer reaction
at a wall-jet electrode,
AE
Solution
outlet
3/4
(43)
Fig. 29
WE
of 1,4-benzoquinone in acetonitrile in
terms of Eq. (43), by plotting ilim against
3/4
Vf revealed a straight line, and gave a
diffusion coefcient in good agreement
with literature values, indicating that the
hydrodynamics of this cell are well dened
and understood.
It has been seen that the other in
situ hydrodynamic cells discussed above,
the channel electrode cell and the tube
electrode cell, both exhibit EPR signal (S)
characteristics dependent upon the ratio
2/3
i/Vf . However, in the case of the walljet electrode under steady state conditions,
in which the entire cell downstream of the
working electrode is lled with radicals,
the rate of loss of radicals from the cell
is proportional to Vf , while the rate of
their formation is proportional to i, thereby
anticipating
i
S
(44)
Vf
This relationship was veried experimentally for the benzoquinone/acetonitrile system. The EPR signals were observed with
excellent signal-to-noise ratio.
Analogous experiments in aqueous solutions gave similar responses. These
present a more stringent test in terms
of EPR compatibility since they are notoriously lossy; in order to reduce dielectric
loss to acceptable levels, the internal diameter of the glass tube was reduced
to 0.15 cm. Although satisfactory hydrodynamic behavior was observed, when the
cell dimensions were reduced for EPR
compatibility and using a cell of the exact
type to that described by Scholz [93], this
was no longer the case signicant deviations from wall-jet behavior were found
whatever the separation between the working electrode and the jet. Reproducibility
problems were additionally encountered,
in part due to the trapping of air bubbles
within the cell, and due to a greater sensitivity of the cell to the exact location of the
jet, pointing to loss of the wall-jet behavior and to the existence of edge effects
caused by the walls of the cell.
In the Scholz design, the separation between the mercury working electrode and
the jet is minimized so as to give rise to
a thin-layer cell as shown in Fig. 30.
Near-exhaustive electrolysis of the electroactive material is thus likely to occur,
leading to large currents and a high sensitivity towards the detection of radicals that
are stable on the experimental timescale;
the shape of the mercury electrode is
such that it may screen the workingelectrode|electrolyte interface from the
100 kHz magnetic eld modulation used
in conventional EPR spectrometers, so that
only radicals downstream of the shielded
area will be EPR active; the wall-jet approach may be more suited in not only
observing mechanistic pathways, but also
Flow
The ow pattern
between the jet and the mercury
working electrode under
thin-layer conditions.
Fig. 30
Hg
203
204
in homogeneous follow-on reactions. Furthermore, EPR alone cannot give structural information concerning diamagnetic reaction participants. Neudeck and
coworkers [102108] combined UV-VIS
and EPR spectroscopies to give insights
into whether primary or non-primary radicals are formed during electrolysis.
3.2.3.8
Working
electrode
Second
counter electrode
Adjustable
cell holder
To photodiode
array detector
Reference
electrode
Gas inlet
ESR-cavity
From light source
Modulation
coils
Fig. 31
First
counter electrode
(45)
205
206
3.2.3.9
(b)
(h)
(c)
(a)
5 cm
(g)
(e)
(h)
(i)
(d)
(j)
(f)
Fig. 32
inherent high resistivity of a at cell combined with low dielectric solvents (such
as 1,2-dichloroethane, DCE) can be partially overcome using this arrangement,
and reasonable linear sweep voltammetry
was observed using a four-electrode potentiostat. All experiments were undertaken
in a TE102 cavity.
3.2.3.9.2 Typical Applications EPR spectra were obtained for both the interfacial
reduction of 7,7,8,8,-tetracyanoquinodimethane (TCNQ) and oxidation of tetrathiafulvalene (TTF), when dissolved in DCE
by the aqueous phase ferri/ferrocyanide
redox couple, following the application
of a potential difference directly to the
liquid|liquid interface. Previous work [111,
112] suggested that a charge-transfer process occurs at the liquid|liquid interface,
due to the heterogeneous reduction of
TCNQ by the aqueous couple,
C
D
C
D
207
208
Open
circuit
f1 f2 f1 f2 f1
Open
circuit
t
[s]
3.2.4
3.2.4.1
C
PADH
D
NPQH + 2e + 2H (48)
+
C
PADH + NPQHD
2PAD
(49)
Fig. 34
The Reduction of C60 This compound undergoes a series of reductions [116, 117]. Electrochemical EPR has
been undertaken in toluene|acetonitrile
mixtures [118], but it is notoriously difcult to unambiguously interpret these
spectra [119, 120].
3.2.4.1.3
10G
stable paramagnetic species. This technique was rst introduced by Janzen and
Blackburn [121, 122], using N-tert-butyl-phenylnitrone (PBN) to spin-trap alkyl
and alkoxy radicals. In a similar manner,
Wadhawan and coworkers [123] were able
to propose the existence of alkyl peroxides
during sono-emulsion Kolbe electrosyntheses. It must be stressed that case should
be taken to avoid electroactive spin traps;
Bard and coworkers [124] showed that
spin trapping is possible during electrochemical experiments, and that PBN is
electroinactive in the potential range 1.5
to 2.5 V versus SCE in acetonitrile. The
interested reader is referred to Kemps
review on spin trapping [125] for further
information.
Electrode Reaction Kinetics and
Mechanism
This section shows the versatility of the
channel ow EPR cell for mechanistic
discrimination.
3.2.4.2
3.2.4.2.1
Electrode
C
A e D
Solution
B C
Electrode
C
C e D
products (iii)
(i)
(ii)
209
210
CN
b b
5G
(a)
CN
5G
(b)
Fig. 35
Steps (i) and (ii) dene a pure EC mechanism, while Steps (i), (ii), and (iii) correspond to an ECE mechanism. In effect, the
mechanisms differ in terms of the electroactivity of the product, C, of reaction
(ii) so that if C is more readily oxidized
or reduced than A, the process is of the
Neff
1.6
1.4
1.2
0.0
3
log10KECE
Working curve showing the relationship of Neff and K for an
ECE process.
Fig. 36
of simple rst-order kinetics (Eq. 36) permitted the inference of a rate constant,
2/3
k. Plotting this parameter against Vf ,
gave a straight line graph that allowed
extrapolation to zero ow rate to give
a value of 0.36 0.02 s1 for the rstorder decay of the radical anion. The
steady state EPR signal was found to
be directly proportional to the generating current, until after the onset of the
second voltammetric wave, when a dramatic decrease of EPR signal occurred.
These features are both consistent with
EC and ECE mechanisms. Conclusive evidence of the mechanistic discrimination
between EC and ECE reactions came from
analysis of the reduction current|ow rate
data, shown in Fig. 37. It can be seen
that at fast ow rates, the data tend towards one-electron behavior, whilst at slow
ow rates, they tend towards two-electron
behavior, indicative of an ECE reaction.
Further analysis of the EPR signal|ow
rates data permitted the determination of
EPR detection efciencies, from which a
rate constant of 0.28 s1 was obtained for
radical decay.
211
1500
1000
Iobs
[A]
212
500
0
0.0
0.3
0.2
Fig. 37
3.2.4.2.2
Electrode
A e B
Solution
Solution
C
F + e D
S + H+ SH+
(i)
C
SH+ + S
D
F + L
B C
(ii)
C
B + CD
A + D
(iii)
1.8
Neff
1.4
1.4
1.2
0.0
3
log10 KDISP1
Working curve showing the relationship of Neff and K for a
DISP1 process.
Fig. 38
Fig. 39
100
200
1
kapp
2 Vlim
f 0
(52)
213
214
C
A e D
(i)
B+P
A+Q
k
Q Products
(ii)
(iii)
Solution of the appropriate convectivediffusion equations at steady state permits the behavior of this mechanism to
be described in terms of a normalized
rate constant:
1/3
[P ]bulk h2 xe2
K = kK
(53)
[A]bulk 402
where [P ]bulk and [A]bulk are the bulk concentrations of P and A. Under conditions
in which the Leveque approximation is
valid [63, 71], the current is a unique function of K . The observed behavior is best
expressed in terms of the effective number of electrons transferred, Neff . The EPR
signal behavior may be deduced from numerical simulation of the concentration of
the ion-radical species, B, as indicated in
Sect. 3.2.3.
Comproportionation Reactions
Comproportionation reactions are of
the form Y + Y2 2Y . Spackman
and coworkers [75] studied the effect
of a homogeneous comproportionation
reaction on an electrode reaction that
proceeds stepwise via two singleelectron transfers,
3.2.4.2.4
Electrode
C
A + e D
(i)
Solution
C
B + e D
(ii)
Electrode
A+C
(iii)
2B
i
2/3
Vf
(55)
1.4
1.2
S2/S1
1.0
0.8
0.6
0.4
0.2
50
60
70
80
90
1.4
1.2
S2/S1
1.0
0.8
0.6
0.4
0.2
50
60
70
80
90
215
216
(56)
It has been shown that this process occurs at the three-phase boundary
of THPD|aqueous electrolyte|solid electrode [135, 136]. In situ EPR transients for
the oxidation and rereduction of THPD
were seen to be consistent with the anion
insertion process occurring at the triple
phase junction [135].
3.2.5
Conclusion
217
218
219
220
3.3
Introduction
221
222
Substrate
electrode
f
Ef + eVbias
ysub
Tip
electrode
ytip
Ef
electronic states of the tip and the sample surface. The tunneling current across
a vacuum or insulation gap between the
tip and the sample surface is given by [26]
4e +
I =
[f (EF eVbias + )
h
f (EF )]s (EF eVbias + )
T (EF + )|M|2 d
(2)
T (EF + ) d
(3)
(a)
Fig. 2
(b)
(c)
223
224
f0
2k
(5)
STM and AFM applications in electrochemistry as imaging tools are numerous. Novel nonimaging applications are
also abundant in the literature [57, 58].
Because of space limitations, here we focus on tunneling spectroscopy and force
spectroscopy applications using selective
examples that are directly related to electrochemistry. Other important nonimaging
applications that are relevant to electrochemistry include fabricating nanostructures [5968] and probing fast kinetics [6975].
Tunneling Spectroscopy
The electronic states of a sample can
be determined with STM by performing
Scanning Tunneling Spectroscopy (STS).
The concept of tunneling spectroscopy
was developed as early as in the sixties
using metal-oxide-metal junctions [26, 76].
The original idea of building the STM by
Binnig and Rohrer was actually to perform
tunneling spectroscopy locally on a small
area. According to Eq. (3), the tunneling
current is a convolution of the tip and
3.3.4.1
(6)
eVbias
Dox ( e + eVbias ) d
(7)
where is the overpotential and Dox is the
density of the oxidized states. When the
molecule couples weakly to the substrate
and the tip, Dox takes the Gaussian form
Dox () =
( )2
exp
kB T
4kB T
(8)
where is the reorganization energy. The
tunneling spectrum, dI /dVbias versus Vbias
gives approximately the density of states
(Dox ). In electrochemical STM, the overpotential () can be varied independent of
the tip-substrate bias voltage (Vbias ). So by
225
226
measuring the tunneling current as a function of the overpotential with a xed Vbias ,
one has a new way to perform tunneling
spectroscopy. It follows from Eqs. (7 and
8) that I versus gives also the density of
states for a small Vbias .
Kuznetsov and Ulstrup [22, 36, 41] have
considered another limit, a sequential
two-step process, in which the electron
transfers from the tip (substrate) to
the molecule and reduces the molecule.
The reduced states of the molecule are
located below the Fermi level of the
substrate, but a thermal uctuation can
shift the states up and then allow a second
electron transfer from the molecule to the
substrate (tip), returning the molecule to
the oxidized state. Neglecting backward
owing current, the current caused by this
two-step process is
I
Rsm Rmt
Rsm + Rmt
(9)
( E0 + eVbias )2
exp
d
4kB T
(10)
and
Rmt
[1 f ( + eVbias )]
( E0 + + eVbias )2
d
exp
4kB T
(11)
0.03
FeBr
0.025
FeCl
0.02
Mn
dI/dV
0.015
2H
0.01
0.005
0
0.005
1000
Zn
Bare electrode
800
600
400
200
Bias
[mV]
Derivatives (dI/dVbias ) of the tunneling current versus bias voltage
curves of Fe, Mn, Zn, and H embedded in tetraphenol porphyrin (TPP).
(From Ref. [85] with permission.)
Fig. 3
227
228
3.3.4.1.3 Molecular Electronics As silicon-based microelectronics is heading towards the increasingly difcult road of
nanometer scale, building electronic devices with individual molecules (molecular
electronics) becomes a promising alternative [103, 104]. Indeed, many molecules
possess wonderful electronic properties
229
S
8
2 Br
Au nanoparticle
6 nm
Counter
electrode
Redox
gate
3 nm
Au substrate
Current
[nA]
30
dI/dV
[nA V1]
230
20
5
0
2
Substrate bias
[V]
10
0
2.0
1.5
1.0
0.5
0.0
Substrate bias
[V]
(a) Schematic representation of a nanoscopic device. Electrons
can be injected into the redox gate by applying a suitable potential
between the substrate and the counterelectrode. (b) A currentvoltage
curve (inset) and its derivative (dI/dVbias ) in mesitylene. The dotted line
corresponds to the control experiment, in which gold nanoparticles were
deposited on a hexanedithiol monolayer. (From Ref. [114] with
permission.)
Fig. 4
3.3.4.2.1
Fdl =
2R
[( 2 + S2 )e2KD D
0 KD T
+ 2T S eKD D ]
(12)
(13)
231
232
(14)
V
[mV]
F/R
[mN m1]
0.6
1.0
0.8
F/R
[mN m1]
0.8
700
500
400
300
200
100
0
+100
0.4
0.6
V
[mV]
0.4
0.2
700
400
200
100
0
+100
0.0
0.2
0.4
0.2
8 10 12 14 16 18 20
d
[nm]
0.0
0.2
0.4
0
10
20
30
40
50
60
d
[nm]
Fig. 5 Force between silica sphere and gold electrode in an aqueous solution of 1 mM KCl at
25 C and pH = 5.5 as a function of the applied potential at gold electrode. The curves
correspond to, from top to bottom, electrode potentials of 700, 500, 400, 300, 200,
100, 0, and 100 mV (vs SCE). Electrostatic repulsion decreases as the electrode potential
increases from 700 to 100 mV. Inset: force data for silica sphere and gold substrate in
10 mM KCl solution. (From Ref. [137] with permission.)
2D
cos
d
eD/d
(15)
233
234
(b)
10
Z position
[nm]
Fig. 6 Two independent measurements of force spectra of the MAC mode
AFM at OMCTS-graphite interface. The oscillation amplitude of the magnetic
cantilever driven by an external magnetic eld oscillates in both approaching
(solid line) and retracting (dotted line) curves in the region of a few nanometers
away from the surface due to ordered layers of OMCTS molecules at the
interface. The arrows on this and subsequent plots correspond to
repulsive-force maxima. (From Ref. [142] with permission.)
235
236
Et 2
z
4L2 (1 )
(17)
where E, t, L, and are the thickness, length, Youngs modulus, and Poisson ratio of the cantilever, respectively.
This equation has been widely used to
determine changes in surface stress although one end of the cantilever is often
clamped in experiments. Dahmen and
coworkers [153] recently performed a nite
element analysis of single crystal plates
(anisotropic) with one end clamped. They
concluded that Eq. (10) derived for the unconstrained bending could still be used in
most cases with appropriate parameters.
It is clear from Eq. (10) that the bending
of a cantilever because of a surface stress
is (L/t)2 , which requires long and
thin cantilevers for high sensitivity. When
trying to improve the sensitivity, one has to
consider also the noise caused by environmental vibration in the measurement. In
order to minimize the environmental vibration, one requires a high-resonance fre
quency, f = 0.163(t/L2 ) E/, in which
is the density of the cantilever material.
In AFM, microfabricated silicon or silicon nitride cantilevers coated with various
suitable if one wants to study the surface stress of a well-dened single crystal.
This task can be performed with an STM
setup [151, 167169]. The sample in STM
can be either a single crystal plate or thermally evaporated metal lms on glass plate
with one end clamped. One side of the single crystal plate must be covered with, for
example, nail polish. An STM tip is placed
over the other end to detect the surface
stress-induced bending in the plate. The
bending causes a change in the tunneling gap between the tip and sample and
is reected in a change of the tunneling
current. The STM feedback maintains a
constant current by adjusting the vertical
position of the STM via a z-piezoelectric
transducer. The amount of the adjustment
gives the deection, z, from which the
surface stress is extracted using Eq. (10).
A nice feature is that the same STM tip
allows one also to image the surface.
3.2
Au(100) 1 1
4.4
2.8
Au(111) 1 1
4.2
2.4
Au(111) rec
4.0
2.0
(111) (1 1)
3.8
(111) (rec)
1.6
400
200
200
400
600
800
Fig. 7
1000
3.6
(100) (hex)
Au(100) hex
4.6
(100) (1 1)
237
238
Conclusion
One important example is scanning tunneling spectroscopy that allows extraction of electronic states of the probed
molecules. The electronic information can
be used to identify structurally similar
molecules, to obtain chemical reactivity
of the molecules, and is directly relevant to
the current effort of developing molecular
electronics. However, in order to reliably
extract the information, improvements in
both the experimental method for measuring tunneling spectroscopy and theoretical
description of the STM imaging mechanism in the electrochemical environment
are needed. Another important example is
the local force spectroscopy with AFM,
which reveals not only various surface
forces but also mechanical properties of
individual molecules. The microfabricated
cantilevers used as force sensors in AFM
have found broad applications in chemical and biological sensors, in detecting
infrared light, magnetic eld, and other
physical quantities and properties. Other
nonimaging applications of STM and AFM
in conjunction with the electrochemical
techniques include nanofabrication and
nanomanipulation, which will continue to
play important roles in the emerging eld
of nanoscience and nanotechnology.
Acknowledgments
239
240
241
242
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
3.4
Historical Overview
243
244
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
245
Reection of the s-() and p-() polarized light on a at mirror surface. Note the
direction of the electric eld upon reection.
Fig. 1
Angle
0
Phase shift
246
45
90
180
Phase shift for the reected s-( ) and p-( ) polarized light upon
reection on a metallic surface. (Data from Ref. [3] with permission.)
Fig. 2
90
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
R R0
R0
(1)
1/2
= (1 + RS ) + 2RS
z
cos Sr + 4
1 (2)
where RS is the s-polarized light reectivity, Sr the phase angle of the reected
radiation, is the wavelength of the radiation, z is the position above the reecting
surface, and = n cos , where n is the
refraction index and the incidence angle.
On the surface plane (z = 0), the electric
eld is given by [6]
2
ES1
02
ES1
1/2
= (1 + RS ) + 2RS cos Sr (3)
1/2 2
= 1 RS
(4)
1/2
= cos2 1 (1 + Rp ) 2Rp
z
cos pr + 4
1
1/2
+ sin2 1 (1 + Rp ) + 2Rp
z
cos pr + 4
(5)
1
247
2.0
E z20
n3 = 3.0
k3 = 30.0
n1 = 1.0
E z2
248
1.0
45
90
Angle
Change of the mean square electric eld for the air/metallic surface with
the incidence angle. The simulation parameters are shown in the gure.
Fig. 3
+ n1 sin 1
02
EP2
(6)
z
C = cos 22
z
S = sin 22
(8)
(9)
and
= cos 1 (1 rP )C
+ i2
2
(1 rP )
S
2
(7)
and
2
EP2
n1 (1 rP )
S
n22
where
= [(1 + rS )C + i(1 rS )S]2
2
n1 (1 + rP )
S
n22
rS =
(10)
rP =
(11)
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
Electrode
Absorbate
Electrolyte
Window
q2
Air
q1
(a)
Electrode
Absorbate
Electrolyte
Prismatic window
Air
q
Air
(b)
Fig. 4
and r12 and r23 are the Fresnel coefcients for the window/electrolyte and
electrolyte/metal interfaces, respectively.
is the energy attenuation in the solution
phase and is given by
d
(12)
2
= 2
249
cos cos 1
Equation (13) is useful in understanding the different components affecting
3.4.2.1.2
2.0
E 20
n3 = 3.0
k3 = 30.0
n1 = 1.0
E 2
250
1.0
5.0
10.0
Z
[m]
Change of the intensity of the reected mean square electric eld
with the distance from the reecting surface for the s- and
p-polarized radiation.
Fig. 5
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
251
252
s-pol. light
1100 cm1
R/R0 = 0.05%
p-pol. light
950 cm1
1196
1500
1400
1300
1200
1100
1000
900
800
Wave number
[cm1]
s- and p-polarized light spectra for the adsorbed sulfate species on
Au(111) in 0.69 mol L1 HF, 0.5 mol L1 KF, and 102 mol L1 K2 SO4
solution. Ref. Pd/H2 . The s-polarized spectrum is extracted. (From Ref. [32]
with permission.)
Fig. 6
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
C3
sv
C2
sv
sv
sh
C2
C2
Fig. 7
Sketch of the D3h symmetry elements for the free nitrate ion.
Proposed NO
vibrations for the different
modes of the adsorbed NO3
ions under a C2v symmetry.
Fig. 8
N
O
O
A1
A1
O
N
O
O
B1
253
254
C2v
C2 (z)
(xz)
(yz)
A1
A2
B1
B2
+1
+1
+1
+1
+1
+1
1
1
+1
1
+1
1
+1
1
1
+1
Tz
Rz
Tx , Ry
Ty , Rx
xx , yy , zz
xy
xz
yz
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
1.2 V
1.1 V
1V
1024
R/R0 = 0.025%
1436
1600
1400
1200
1000
Wave number
[cm1]
In situ FTIR p-polarized light spectra of adsorbed NO3 ions on
Au(100) in 1 mol L1 and 0.05 mol L1 HNO3 . Ref. Pd/H2 . (From Ref. [33]
with permission.)
Fig. 9
255
256
Screw for
fixation of the electrode
Connection
PTFE support
Gas entrance
Position for voltammetry
Reference electrode
Auxiliary electrode
Position for the in situ
FTIR
Window
support
PTFE ring
Window
Fig. 10
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
1.02
1.00
0.98
R/R
0.96
0.94
0.92
0.90
0.88
0
(a)
20
40
60
80
Angle
[degree]
Fig. 11
257
1.0
0.9
0.8
R /R 0
258
0.7
0.6
0.5
0.4
0
40
60
80
Angle
[degree]
(b)
Fig. 11
20
(Continued)
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
259
p-pol. light
0
50
400
600
600
R/R0 0.2%
E
[mV]
400
E
[mV]
260
R/R0 0.5%
1000
1000
1200
1100
1000
1200
1100
1000
Wave number
[cm1]
Fig. 12
HPO4 2 + H3 O+
H2 PO4 + H2 O
(15)
This is an example of how pH changes
caused by H+ desorption during the
potential step used for the measurement
cause equilibrium changes in solution,
and any possible band due to adsorbed
phosphate is completely buried by the
solution spectral signals. To minimize
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
261
1150
1050
Wave number
[cm1]
1250
1050
R/R0 = 0.1%
1100
(i)
(ii)
(b)
950 1300
1100
1100
(i)
952
(ii)
1000
(c)
1300
(ii)
960
(i)
1100
1000
R/R0 = 0.05%
1150
1100
Wave number
[cm1]
1200
1247
1240
Rh(111)
900 1400
R/R0 = 0.01%
1050
Wave number
[cm1]
1200
1192
1185
Au(111)
Fig. 13
In situ FTIR spectra of adsorbed sulfate species on (a) Pt(111). (Data extracted from Ref. [45] with permission.) (b) Au(111). (Data extracted
from Ref. [32] with permission.) (c) Rh(111) in (i) 0.69 mol L1 HF, 0.5 mol L1 KF, and 102 mol L1 K2 SO4 and (ii) 7.3 mol L1 HF and 102 mol L1
K2 SO4 . (Data extracted from Ref. [46] with permission.)
(a)
1350
1272
1254
Pt(111)
262
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
1193
4 104 a.u.
958
1187
1103
900
1200
n
[cm1]
263
E
vs SHE
[mV]
1500
1234
1300
956
1100
R/R
1250
900
800
1170
700
957
Wave numbers
[cm1]
264
1225
1200
1175
0.001
1153
1400
(a)
1300
1200
1100
Wave numbers
[cm1]
1000
1150
600 800 1000 1200 1400 1600
900
(b)
E vs SHE
[mV]
(a) In situ FTIR spectra of adsorbed sulfate species on Au(111) in 0.5 mol L1 D2 SO4 in D2 O
and (b) band center energy as a function of the applied potential. (From Ref. [48] with permission.)
Fig. 15
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
More detailed results from the calculations are necessary to account for these
differences. The controversy on the nature
of the adsorbed sulfate species (sulfate or
bisulfate) can be attributed to the limitations of the in situ FTIR spectroscopy in
this case. Associated with the difculty in
identifying the nature of the adsorbed anion (sulfate or bisulfate) is the difculty
in determining the symmetry of the adsorbed species, which would give the clue
in determining the coordination to the surface. The problem of the adsorbed sulfate
is that only the symmetric modes will be
active, according to the surface selection
rule (see Sect. 3.4.2.1.2). It is assumed that
the Td symmetry is broken upon adsorption, since the oxygen atoms are no longer
equivalent. The nal symmetry will depend on the coordination to the surface.
Coordinating one or three oxygen to the
surface will lead to a C3v symmetry, while
coordinating two oxygen will lead to a C2v
symmetry. The C2v symmetry has no degenerated modes, while the C3v has two
degenerated modes, one for the stretching
modes. Therefore, for the free species in
solution, the C2v symmetry must present
one more stretching mode than the C3v .
Because of the fact that the surface selection rule limits the active modes to those
having the dynamic dipole moment perpendicular to the surface, two different
coordinations will present the same number of bands. Considering only the number
of bands, it is not possible to identify the
symmetry of the adsorbed sulfate species;
and consequently, it is not possible to assign unequivocally either the coordination
or the nature of the adsorbed species.
Other strategies have been used to clear
these points.
Shingaya and coworkers have simulated
the electrochemical environment in ultra
high vacuum (UHV) by adding water and
265
266
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
Pt(110)
1106
1234
Pt(100)
1108
1208
1100
Pt(111)
R/R0 = 0.1%
1254
1300
1200
1100
1000
Wave number
[cm1]
In situ FTIR spectra for the adsorbed sulfate species
on platinum single-crystal electrodes in 0.69 mol L1 HF,
0.5 mol L1 KF, and 102 mol L1 K2 SO4 . Ref. Pd/H2 . (Data
extracted from Ref. [45] (Pt(111)), Ref. [57] (Pt(100)), and
Ref. [58] (Pt(110)) with permission.)
Fig. 16
267
268
R/R0 = 0.033%
1.2 V
956
1091
1.0 V
1181
0.8 V
1165
1091
0.7 V
1300
1200
1100
1000
900
Wave number
[cm1]
In situ FTIR spectra for the adsorbed sulfate species on Au(100)
single-crystal electrode in 7.3 mol L1 HF and 102 mol L1 K2 SO4 .
Ref. Pd/H2 . (Data reproduced from Ref. [34] with permission.)
Fig. 17
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
269
270
2075
Pt
2061
PtRu
2010
Ru
R /R0 = 0.005
2250
2100
1950
Wave number
[cm1]
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
271
272
0.60 V
0.55 V
0.50 V
0.55 V
0.45 V
CO(L)
0.50 V
0.40 V
0.45 V
0.30 V
CO2
0.20 V
0.40 V
0.10 V
0.35 V
R/R = 5 103
2400
2300
2200
2100
2000
1900
Wavenumber
[cm1]
In situ FTIR spectra for saturated CO adsorbates on a
PtRu(50 : 50) alloy electrode in 0.1 M HClO4 from 100 mV to
600 mV at 50 mV interval. CO was adsorbed at 300 mV. For each
spectrum, 100 interferograms were collected at 8 cm1 resolution,
acquisition time ca. 44 s. Bands on the left side, corresponding to
CO2 produced during COads oxidation, were calculated against a
reference spectrum taken at 100 mV (CO2 was absent). Bands on
the right side, corresponding to adsorbed CO, were calculated
against a reference spectrum taken at 800 mV after complete
oxidation of CO. (From Ref. [62] with permission.)
Fig. 19
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
1.0
CO2
0.8
0.6
0.4
0.2
PtRu
Ru
Pt
0.0
0.2
0.4
0.6
Potential (RHE)
[V]
Integrated band intensity for CO2 as a function of potential.
The intensities were calculated from series of spectra obtained at each
catalyst as shown in Fig. 20 for the PtRu alloy. (From Ref. [62] with
permission.)
Fig. 20
273
Fig. 21
2085
Pt
2070
2055
Wave number
[cm1]
274
PtRu
2040
2025
Ru
2010
1995
0.0
0.2
0.4
0.6
0.8
E vs RHE
[V]
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
q = 0.12
q = 0.41
q = 0.43
q = 0.45
q = 0.69
q = 0.69
q = 0.75
q=1
q=1
q=1
R/R0 = 5 103
R/R0 = 5 103
Pt
2400
2200
2000
R/R0 = 5 103
PtRu (50:50)
2400
2200
2000
Ru
2400
2200
2000
Wave number
[cm1]
Comparison of in situ FTIR spectra for submonolayer CO adsorbates on Pt,
PtRu(50 : 50), and Ru electrodes in 0.1 M HClO4 . CO was adsorbed at 300 mV. Sample
spectra were acquired at 300 mV and computed against a reference spectrum taken at
800 mV; 100 interferograms were collected at 8 cm1 resolution, acquisition time ca. 44 s.
The CO coverages co indicated on each spectrum were calculated from the ratio of the
band intensity for CO2 at each coverage to the CO2 band obtained at saturation coverage.
(From Ref. [62] with permission.)
Fig. 22
3.4.4.1.2
275
2080
Pt
PtRu
2040
Wave number
[cm1]
276
Ru
2000
1960
0.0
0.2
0.4
0.6
0.8
1.0
qCO
Dependence of vCO on CO for submonolayer CO adsorbates
on Pt, PtRu(50 : 50), and Ru electrodes in 0.1 M HClO4 at a constant
potential of 300 mV: CO adsorption at 300 mV. Other details as in
Fig. 23. (From Ref. [62] with permission.)
Fig. 23
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
600
600
550
550
500
500
550
550
450
450
400
400
500
500
350
350
300
300
450
250
400
450
250
400
200
200
150
150
R/R0 = 0.01
1816
2058
qRu = 0.2
2400
2200
(a)
Wave number
[cm1]
2000
qRu = 0.75
1800
1982
2047
2008
R/R0 = 0.01
1982
2400
2200
(b)
Wave number
[cm1]
2000
1800
In situ FTIR spectra for adsorbed CO and for the CO2 produced during the CO
stripping in 0.1 M HClO4 on Pt(111) modied with Ru. Ru coverage: (a) 0.2 mL and
(b) 0.75 mL. After the formation of the CO adlayer at 0.1 V, CO was eliminated from the
solution by N2 bubbling. The potential was then changed from 0.1 V onwards in 50 mV
steps. The sample potentials are indicated in the corresponding spectra. The CO2 region
(band at 2341 cm1 ) was calculated with a reference spectrum taken at 0.1 V (a potential
in which CO2 is not formed). For the CO region, the reference spectrum was one taken at
0.8 V, that is, a potential in which COads was completely oxidized. (From Ref. [63] with
permission.)
Fig. 24
277
Pt(111)
Wave number
[cm1]
2060
2040
Pt(111)/Ru
2020
Ru
2000
0.0
0.2
0.4
0.6
Potential vs RHE
[V]
Comparison of the CO stretch wave number for the COL on the
Pt(111), Ru-modied Pt (111), and pure polycrystalline Ru as a function
of potential. Data were derived from IR spectra as those in the gure.
Experimental conditions as in Fig. 25. (From Ref. [63] with permission.)
Fig. 25
1.0
0.8
0.6
(a)
Ru
PtRu alloy
0.4
0.2
0.0
Pt(111)/Ru
0
0.6
Pt(111)/Ru
0.4
PtRu alloy
0.2
Ru
Pt(111)
0.0
Time
[min]
0.8
Pt(111)
278
0
(b)
Time
[min]
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
279
(a)
Ru(0001)
550 mV
500 mV
(b)
16
Band intensity
[a.u.]
5
1000 mV
800 mV
650 mV
12
COL
8
4
CO2
COH
450 mV
0
2040
400 mV
(c)
250 mV
2030
200 mV
2020
150 mV
2010
nCO
[cm1]
280
0 mV
COL
(d)
1800
100 mV
1790
200 mV
COH
R/R0 =
0.005 COL
2100
2000
1780
1900
1800
1770
1760
Wave number
[cm1]
COH
0.2 0.0 0.2 0.4 0.6 0.8 1.0
E vs Ag/AgCl
[mV]
(a) In situ FTIR spectra for CO adsorbed at Ru(0001) at different potentials (vs
Ag/AgCl/Cl ) as indicated. CO was adsorbed at 100 mV during 5 min, then the solution was
replaced with pure base electrolyte (0.1 M HClO4 ). Spectra were normalized versus a
spectrum taken at 1100 mV, after waiting for 2 min. in order to completely oxidize CO.
(b) Integrated band intensities for COL , COH from spectra is shown in (a). CO2 band intensity
from the same experiments. (c), (d) Potential dependence of the band center frequency for
COL and COH , respectively, from spectra shown in (a). (From Ref. [65] with permission.)
Fig. 27
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
281
282
0.15 V
0.20 V
0.30 V
0.30 V
0.35 V
0.40 V
0.45 V
0.50 V
0.40 V
0.50 V
0.55 V
0.50 V
0.55 V
2056
2046
0.55 V
1721
2059
1826
1719
R/R0 = 0.02
R/R0 = 0.03
R/R0 = 0.01
2341
2500
2341
Pt(111)
2000
2500
Pt(111)/Ruads
2000
2500
2341
PtRu alloy
2000
Wavenumber
[cm1]
In situ FTIR spectra for Pt(111), Pt(111)/Ru 39%, and PtRu alloy
(85 : 15) in 0.5 M CH3 OH + 0.1 M HClO4 . Potentials as indicated on each
spectrum; reference spectrum taken at 0.05 V. (From Ref. [70] with permission.)
Fig. 28
C2 H5 OH + 0.1M HClO4 , ethanol adsorption is minimized. Therefore, this procedure allows the monitoring of the complete
process of ethanol adsorption and oxidation as the potential is increased in the
positive direction. Because of the wellknown problem of surface poisoning by
strongly adsorbed organic residues, data
obtained under these conditions may not
be observable in the second and following
potential scans or when the electrode
comes in contact with ethanol without potential control.
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
0.15
0.15
0.35
0.25
0.40
0.30
0.50
0.50
0.60
0.60
E
[V]
E
[V]
0.55
1866
0.65
0.70
3000
2500
2000
Wave number
[cm1]
1500
2341
R/R0 = 1.5%
2625
Pt(111)
3000
1407
1282
1712
1370
R/R0 = 1.5%
0.80
2055
1282
2343
2632
1412
2061
1713
2983
2906
0.70
Pt(100)
2500
2000
1500
Wave number
[cm1]
In situ FTIR spectra (256 scans, 8 cm1 resolution) at Pt(111) (left side) and Pt(100)
(right side) in 0.1 M C2 H5 OH + 0.1M HClO4 . Reference spectrum collected at 50 mV, and
sample spectra collected at the indicated potential in 50 mV steps (some spectra were
omitted for clarity). (From Ref. [80] with permission.)
Fig. 29
283
Wave number
[cm1 ]
Functional group or
chemical species
Mode, comments,
references
2983, 2906
2632 (broad)
2341
20552060
1713
1402
1370/1281
1100
CH3 , CH2
COOH
CO2
Adsorbed CO
COOH or CHO
Adsorbed CH3 COO
COOH
ClO4
CH str., 81
OH str., 81
CO asym. str., 81
Linearly bonded, 60
C=O str., carbonyl, 81
CO sym. str., 82
Coupl. CO str. OH def., 81
ClO str. (F), 35
Pt(111)
Pt(100)
20
45
CO2
15
Band intensity
[a.u.]
Band intensity
[a.u.]
284
10
COL
COL
CO2
30
COB
15
5
0
0
0.2
(a)
0.4
0.6
0.8
1.0
Potential vs RHE
[V]
1.2
0.2
(b)
0.4
Band intensity for CO and CO2 from spectra collected during the
oxidation of 0.1 M ethanol in 0.1 M HClO4 at (a) Pt(111) and (b) Pt(100).
Experimental conditions as in Fig. 29. (From Ref. [80] with permission.)
Fig. 30
0.6
0.8
Potential vs RHE
[V]
1.0
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
1.0
Current density
[mA cm2]
0.8
0.6
0.4
0.2
0.0
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
Potential vs RHE
[V]
Cyclic voltammogram for a Pt(111) electrode in 1 M C2 H5 OH + 0.1 M
HClO4 solution. Sweep rate 0.05 V/s.
Fig. 31
285
(a)
8
9
COB
2
0
CH3COOH
Band intensity
[a.u.]
Band intensity
[a.u.]
286
COL
1.5
5
4
CH3CHO 3
3
1.0
2
CO2
0.5
0
0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2
(b)
Potential vs RHE
[V]
0.0
0.2
0.4
0.6
0.8
1.0
1.2
Potential vs RHE
[V]
(c)
Potential dependence of the band intensities for CO, CO2 , acetaldehyde, and acetic
acid, from spectra obtained at Pt(111) during the oxidation of ethanol in 1 M C2 H5 OH + 0.1 M
HClO4 . After taking a reference spectrum at 0.05 V, potential steps were applied towards higher
positive potentials and then back to the initial value. All spectra are referred to the rst one taken
at 0.05 V. A 60 prismatic CaF2 window was used for CO and CO2 and a at ZnSe window for
acetaldehyde and acetic acid. (From Ref. [80] with permission.)
Fig. 32
Final remarks
The examples presented in this chapter show the power of in situ FTIR
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
FAPESP and CNPq are gratefully acknowledged for their nancial support.
References
1. L. H. Little, Infrared Spectra of Adsorbed
Species, Academic Press, London, 1966.
2. S. A. Francis, A. H. Ellison, J. Opt. Soc. Am.
1959, 49, 131138.
3. R. G. Greenler, J. Chem. Phys. 1966, 50,
310315.
4. A. H. Reed, E. Yeager, Electrochim. Acta
1970, 15, 13451354.
5. H. Neff, P. Lange, D. K. Roe et al., J. Electroanal. Chem. 1983, 150, 513519.
6. W. N. Hansen, J. Opt. Soc. Am. 1968, 58, 380.
7. K. Ataka, M. Osawa, Langmuir 1998, 14,
951959.
8. M. Osawa, K. Ataka, K. Yoshii et al., J. Electron. Spectrosc. 1993, 64, 371379.
9. Z. Ping, G. E. Nauer, H. Neugebauer et al.,
Electrochim. Acta 1997, 42, 16931700.
10. A. Bewick, K. Kunimatsu, B. S. Pons, Electrochim. Acta 1980, 25, 465480.
287
288
60. Beden, C. Lamy, A. Bewick et al., J. Electroanal. Chem. 1981, 121, 343.
61. D. K. Lambert, Electrochim. Acta 1996, 41,
623630.
62. W. F. Lin, T. Iwasita, W. Vielstich, J. Phys.
Chem. B 1999, 103, 32503257.
63. W. F. Lin, M. S. Zei, M. Eiswirth et al., J.
Phys. ChemB 1999, 103, 69686977.
64. S. Cram, K. A. Friedrich, K. P. Geyzers et al.,
Fresenius J. Anal. Chem. 1997, 358, 189192.
65. W. F. Lin, P. A. Christensen, A. Hamnett
et al., J. Phys. Chem. 2000, 104, 66426652.
66. N. Ikemiya, T. Senna, M. Ito, Surf. Sci. 2000,
464, L681L685.
67. T. E. Madey, C. Benndorf, Surf. Sci. 1985,
164, 602624.
68. H. Pfnur, D. Menzel, F. M. Hoffmann et al.,
Surf. Sci. 1980, 93, 431452.
69. B. N. J. Persson, R. Rydberg, Phys. Rev. B
1981, 24, 69546970.
70. T. Iwasita, H. Hoster, A. John-Anacker et al.,
Langmuir 2000, 16, 522529.
71. H. Hoster, T. Iwasita, H. Baumgartner et al.,
PCCP 2001, 3, 337346.
72. B. E. Nieuwenhuys, W. H. M. Sachtler, Surf.
Sci. 1973, 34, 225.
73. B. E. Nieuwenhuys, R. Bouwman, W. H. M.
Sachtler, Thin Solid Films 1973, 21, 5.
74. E. Ticianelli, J. G. Beery, M. T. Paffett et al.,
J. Electroanal. Chem. 1989, 258, 6177.
75. M. Watanabe, S. Motoo, J. Electroanal. Chem.
1975, 60, 267273.
76. T. Iwasita, B. Rasch, E. Cattaneo et al., Electrochim. Acta 1989, 34, 10731079.
77. S.-C. Chang, L.-W. H. Leung, M. J. Weaver,
J. Phys. Chem. 1990, 94, 60136021.
78. F. Cases, M. Lopez-Atalaya, J. L. Vazquez
et al., J. Electroanal. Chem. 1999, 278,
433440.
79. T. Iwasita, E. Pastor, Electrochim. Acta 1994,
39, 531537.
80. X. H. Xia, H.-D. Liess, T. Iwasita, J. Electroanal. Chem. 1997, 437, 233240.
81. G. Socrates, Infrared Characteristic Group
Frequencies, Wiley, New York, 1966.
82. A. Rodes, E. Pastor, T. Iwasita, J. Electroanal.
Chem. 1994, 376, 109118.
83. J. Shin, W. J. Tornquist, C. Korzeniewski
et al., Surf. Sci. 1996, 364, 122130.
84. X. H. Xia, T. Iwasita, F. Y. Ge et al., Electrochim. Acta 1996, 41, 711718.
85. T. Iwasita, X. H. Xia, E. Herrero et al., Langmuir 1996, 12, 42604265.
3.4 In situ FTIR as a tool for mechanistic studies. Fundamentals and applications
86. T. Iwasita, X. H. Xia, J. Electroanal. Chem.
1996, 411, 95102.
87. T. Iwasita, X. H. Xia, H.-D. Liess et al., J.
Phys. Chem. B 1997, 101, 75427547.
Appendix
(A1)
(A2)
1
S =
Re(E H )
2
(A3)
1
(A4)
divS =
E E
2
where is the effective conductivity of the
absorbing medium at the eld frequency
and the quantity E E is the time average
of the E2 , therefore
1
divS =
(A5)
E2
2
Hence the rate of energy dissipation
is proportional to the square of the
mean electric eld of the incident radiation and it is the result of the
spectroscopic measurement. The intensity of the E2 is directly related to the
amount of dissipated energy by the absorbing medium.
289
295
4.1
Introduction
296
Methods
4.1.2.1
X-ray Diffraction
sin2 (Qa3 )
(1)
n
fj exp(iQrj )
(2)
j =1
Calculated intensity
103
102
10
2p
3p
4p
Fig. 1
297
298
(0 1 l)
(1 1 l)
(0 1 l)
(015)
(114)
(0 0 l)
(013)
(1 0 l)
(0 0 l)
(012)
(003)
(0 k 0)
(104)
(004)
(0 k 0)
(1 0 l)
(112)
(011)
(103)
(110)
(002)
60
(a)
(101)
(101)
(h 0 0)
(b)
(h 0 0)
The real-space surface structures and corresponding reciprocal space lattices for:
(a) (111); (b) (001); and (c) (110) surface terminations of an fcc crystal. The unit cells are shown
in the real-space structures and the corresponding Bragg reections indicated in the reciprocal
space lattices.
Fig. 2
(Continued)
(1 1 l)
(0 1 l)
(0 0 l)
(1 0 l)
(112)
(002)
(011 )
(0 k 0)
(101)
(110)
(c)
(h 0 0)
299
300
2
f
1 exp[i(h + k + l)]
(5)
Finally, the (110) reciprocal surface unit
cell is rectangular as shown in Fig. 2(c) and
the reciprocal lattice notation is such that
h is along [1 1 0], k along [0 1 0], and l is
along the [1 1 0] surface normal. The units
for h, k, and l are a = c = 2/aNN and
b = 4/ 2aNN , where aNN is the nearestneighbor distance in the crystal. For this
surface, the CTR scattering reduces to
the form
I (h, k, l)
2
f
1 exp[2i(h/2 + k/2 + l/2)] (6)
Equations (46) are by no means rigorous
but they essentially form the basis for
analyzing CTR data from the three lowindex surfaces of fcc transition metals.
The equations are easily modied to
include surface relaxation effects, surface
roughness, or adsorbed adlayer species
and are used in the analysis of the results
presented in Sects. 4.1.3 and 4.1.4 of
this report.
4.1.2.1.3 Experimental Details The general experimental procedures used in X-ray
diffraction measurements of electrochemical systems have been described in detail
in previous articles [68]. After surface
preparation of the electrode to be studied,
the crystal is transferred (preferably whilst
being protected from the atmosphere by a
drop of electrolyte on the surface) into an
X-ray electrochemical cell. The design of
the electrochemical cell that we have used
in our experiments is shown schematically
in Fig. 3. In this design, the crystal is held
in place by tightening down the central column of the cell, which is screw-threaded
into the main body (all parts are made of
Kel-F material, which can be chemically
cleaned and is relatively easy to machine).
The crystal surface then forms the highest
part of the cell and so is easily accessible to the incident X-ray beam. Electrical
contact to the working electrode and to a
counter and reference electrode are made
through feedthrough connections on the
side of the cell. In our experiments, a palladium/hydrogen reference electrode was
used, which was in direct contact with
the electrolyte in the main body of the
Electrolyte out
Electrolyte in
Counterelectrode
Working electrode
Fig. 3
cell. Care has to be taken in this situation as, depending on the nature of the
electrochemical reactions that are being
studied, there can be shifts in the reference potential. Two further connections
form electrolyte feedthroughs to the cell
and the liquid is contained within the cell
by a thin (10 m) polypropylene window
sealed with an o ring. During the X-ray
experiment, the electrochemical cell can be
deated to trap a thin layer of electrolyte
(10 m) between the electrode surface
and the X-ray window to shorten the solution pathlength of the X-ray beam. Note
that this cell design allows the sample to be
oriented in any direction as the electrolyte
is contained, that is, there is no need for
any specialized diffractometer equipment
in order to maintain a horizontal surface
geometry. In order to minimize contamination from the atmosphere, the cell is
surrounded by an outer shell containing a
nitrogen atmosphere.
Most of the X-ray measurements described in this report were carried out
at room temperature on beamline 7-2
at the Stanford Synchrotron Radiation
Laboratory (SSRL) utilizing a focused
monochromatic X-ray beam. Additionally,
301
302
(8)
where the diffusion limiting current is proportional to angular velocity 1/2 [15, 16].
The success of the RDE method has
stimulated the development of several
other rotating congurations. The RRDE
is perhaps the most useful extension of
the idea of the RDE. The RRDE was rst
developed by Frumkin and Nekrasov to
detect unstable intermediates in electrode
reactions. The RRDE, Fig. 4, consists of
a central disk electrode surrounded by
303
304
Fig. 4
interference from kinetic and doublelayer effects. The best example is UPD
of copper in solution containing Br , in
which the amount of Cu UPD and/or
bromide anion deposited/adsorbed on the
disk electrode can be estimated using
the two ring-electrode properties ring
shielding and ring collection [16]. For both
modes, the UPD species is reduced to
metal at the ring, and electroactive species
in solution must undergo convectivediffusion controlled reduction at the ring
electrode. The ring-shielding property of
the RRDE is illustrated in Sect. 4.1.4.2.3 in
which it is used to assess the mass ux of
both the Cu2+ and the Br from and to the
Pt(111) disk electrode. If either Br or Cu
is adsorbed at the Pt(111) disk, the surface
coverages Br,Cu can be assessed from the
ring currents in either potentiodynamic or
potentiostatic experiments:
Br,Cu
1
=
Q
(1/v)
(ir ir ) dE
AnN
(9)
Br,Cu
1
=
Q
(ir(t) ir ) d
AnN
(10)
305
306
Surface Structures
Surface Reconstruction
4.1.3.1.1 Gassolid
Interface Under
UHV conditions, the clean low-index faces
due to the difculty in preparing clean surfaces and transferring them to the electrochemical environment (see Sect. 4.1.2.3).
Ex situ techniques, whereby the electrode
is emersed from solution and transferred
back into UHV for surface analysis, cannot give denitive results due to the loss of
potential control and/or solution species.
The advances in the application of in situ
methods, for example, scanning tunneling
microscopy (STM) [17, 21] and X-ray scattering, have alleviated this emersion gap
problem, and provided denitive structure
determination of metal surfaces in electrolyte and under potential control.
Studies of the reconstruction of (111)
metal surfaces date back to the early
1970s, when it was reported that UHVprepared Pt(111)-(1 1) surface remained
intact after contact with solution [33, 34].
More recently systematic, in situ X-ray scattering [35] and STM [21, 28] studies have
demonstrated that well-ordered P(111)(1 1) structures, prepared either in UHV
or by the ame-annealing procedure, are
stable in aqueous solutions. As for Pt(111),
the (111) surfaces of Ag and Cu exhibit
a (1 1) termination in the absence of
any strongly adsorbing species, such as
anion adsorption or UPD metals. In contrast, the Au(111) single-crystal surface
exhibits a potential-dependent reconstruction that is unique among the fcc metal
surfaces. As in UHV, the surface forms
the (23 3) structure at negative electrode potentials whereas at positive potentials the unreconstructed (1 1) surface
is observed. The driving force for the
307
308
d = 0.206
(1, 1, 0)c
(1, 0)h
1.206
(0, 0, 0)
(1, 1, 0)c
(0, 1)h
(a)
(b)
1.6
1.6
1.6
g
b
(c)
(d)
Fig. 5
the real driving force for the potentialinduced reconstruction of Au(111) is. At
present, we consider the issue to be unresolved; for further details the readers
are refered to Kolbs review on reconstruction phenomena at metalelectrolyte
interfaces [17].
The structural behavior of the (001) surfaces in aqueous electrolytes has been
more controversial. For example, the rst
ex situ experiments indicated that the
Pt(001) 5 20 structure transforms
into the (1 1) phase upon contact with
aqueous solution or even when in contact with water vapor [38]. In contrast,
Zei and coworkers found that the UHVprepared reconstructed Pt(001) surface
remained stable in contact with electrolyte, although no convincing evidence of
potential-induced reconstruction was ever
detected [39]. X-ray scattering measurements have unambiguously demonstrated
that the Pt(001)-(1 1) surface is stable in
the potential range between the adsorption of hydrogen and hydroxyl species and
the reconstructed surface was never observed [40]. In contrast with Pt(001), the
Au(001) surface in electrolyte exhibits a
potential-induced reconstruction that is
very similar to the temperature-induced
reconstruction in UHV [7]. To illustrate
the general phenomenon of surface reconstruction in electrolyte, representative
results for this system are now presented
to demonstrate the power of the X-ray
diffraction technique in elucidating the
potential-induced structural transformation of the Au(001) surface.
The potential stability of the gold surface reconstruction in the electrochemical
environment has been studied by CV,
ex situ emersion LEED experiments [17],
STM, and X-ray diffraction [7]. In both
acid and alkaline solutions, there is good
agreement between the experimental techniques that the 5 20 reconstruction
(see Fig. 5) is formed at cathodic potentials and that it can be reversibly lifted
and formed upon cycling the applied
potential anodically. Figure 6 shows representative in-plane X-ray diffraction results
in the form of rocking scans through
309
0.1 M KOH
0.16
0.12
(a)
0.12
0.1 M HClO4
0.11
Intensity
[a.u.]
310
0.10
0.09
(b)
0.30
(1, 1, 0.4)
0.20
0.10
3
(c)
f
[degrees]
Rocking scans through: (a) (1.206, 1.206, 0.4) in 0.1 M KOH (at
E = 0.35 V); (b) (1.206, 1,206, 0.4) in 0.1 M HClO4 (E = 0.05 V);
and (c) (1, 1, 0.4), a CTR position for a Au(001) electrode. These
positions are illustrated in Fig. 5. The closed symbols in (a) show the
measured prole in both solutions after the lifting of the reconstruction
at positive potentials.
Fig. 6
0.17
0.16
(a)
0.15
0.14
Intensity
[a.u.]
0.1 M KOH
0.12
0.10
(b)
(0, 0, 1.3)
0.40
0.20
0.0
(c)
0.4
0.8
1.2
E
[V]
XRV measurements at: (a) the rotated peak position in 0.1 M HClO4 (see
Fig. 6b); (b) (1.206, 1.206, 0.4) in 0.1 M KOH; and (c) at (0, 0, 1.3), a position on
the specular CTR, in the same solution as (b).
Fig. 7
311
3.0
2.5
Intensity
[a.u.]
312
2.0
1.5
1.0
1.48
0.4
1.52
(h, 0, 0.1)
(a)
(b)
0.0
0.4
f
[degrees]
The measured X-ray intensity at (0, 1.5, 0.1): (a) along the [0 1 0] direction; and
(b) along the [1 0 0] direction for a Pt(110) electrode in 0.1 M NaOH (E = 0.1 V). The
solid lines are ts of a Lorentzian lineshape to the data.
Fig. 8
distributed single-height steps on the surface. The Lorentzian lineshape does not
perfectly produce the data in the tails of the
Bragg reection (Fig. 8a) and this could be
due to a residual step distribution. It is
important to emphasize that the (1 2)
surface reconstruction was stable over a
wide potential range, for example, even at
1.2 V there is still a strong (110)-(1 2)
diffraction pattern. Upon sweeping the
potential negatively, however, the (1 2)
reconstruction is nally lifted as the surface atoms move to accommodate the
oxide reduction. It should also be noted
that the Pt(110)-(1 2) structure was so
stable in aqueous electrolytes that adsorption of CO on this surface did not
induce the (1 2) (1 1) transition
that is observed in UHV upon adsorption of CO [44]. The high stability of
the Pt(110)-(1 2) structure implies that
the potential-induced mobility of Pt surface atoms at the solidliquid interface
is signicantly reduced compared to the
temperature-induced mobility in UHV.
In contrast with Pt(110), the reconstruction at the Au(110)liquid interface can
be reversibly lifted and reformed, implying that the mobility of gold surface
atoms is appreciable even in contact with
electrolyte. Depending on the nature of
the supporting electrolyte, however, either
the (1 2) or (1 3) reconstruction was
found to be stable at negative potentials. In
particular, Magnussen and coworkers [45]
demonstrated that in perchloric acid the
(1 2) phase was stable at the potential
of zero charge (denoted hereafter as the
pzc). Positive of the pzc, the reconstruction
is lifted and the Au(110)-(1 1) structure
is clearly observed in STM experiments.
The observed structural transformation
is highly reversible, the (1 2) structure
being fully restored when the applied potential is negative of the pzc. Because of
the fact that the phase transition was observed at the pzc, it was proposed that the
structural transformation of the Au(110)
surface was governed by the specic (contact) adsorption of anions. In contrast, in
an X-ray diffraction experiment, a (1 3)
reconstruction was observed in salt solutions (0.1 M NaCl) [46]. As for Au(111),
Ocko and coworkers postulated that the
formation of the surface reconstruction
was driven by the induced negative surface charge. It was also suggested that
the (1 3) reconstruction was stabilized
over the (1 2) structure due to the negative surface charge, by analogy to UHV
results showing that Ag(110) underwent
a (1 1) (1 2) transition after the
adsorption of K on the surface [47]. As
for the other Au surfaces, it is very difcult to resolve unambiguously the driving
force for the potential-induced structural
transformation, especially given that surface charge and anion adsorption are so
strongly correlated.
Surface Relaxation
The relaxation of metal surfaces under
UHV conditions has been well studied
by utilizing LEED I V analysis and a
signicant database now exists [23]. For
unreconstructed low-index single-crystal
surfaces, it is often found that the
outermost layer of atoms is contracted
toward the second layer, compared to the
bulk layer spacing. Adsorption on the clean
surface in UHV, however, usually reverses
this trend and results in the outermost
layer of metal atoms being expanded away
from the second layer. The experimental
data has led to interesting theoretical
work aimed at establishing a fundamental
understanding of surface relaxation [48].
The same concepts can be applied to the
M(hkl)electrolyte interface, and indeed
relaxation has also been observed at the
4.1.3.2
313
314
Change density
[c cm2]
200
100
0
10 A
QH
QOH
(a)
Side view
5.2
(0, 0, 1.55)
5.0
X
Intensity
[a.u.]
4.8
4.6
4.4
Top view
4.2
0.0
(b)
0.2
0.4
0.6
0.8
1.0
E
[V]
Fig. 9
315
Intensity
[a.u.]
101
102
(a)
103
101
Intensity
[a.u.]
316
102
103
0.5
(b)
1.0
1.5
2.0
2.5
l
[rlu]
The measured CTR data: (a) (0, 0, l); and (b) (1, 0, l) for the
Pt(110) electrode at an electrode potential of 0.1 V. The dashed lines are
calculated for a contracted (1 2) missing-row model that was found in
UHV studies [52] and the solid lines are ts to the data indicating a 20%
expansion of the surface [41].
Fig. 10
Adsorbate-modied Surfaces
317
3
E0
Current
[A]
2
1
0
1
2
3
0.4 0.2
0.0
0.2
0.4
0.6
0.8
(0, 0)
(0, 1)
(i) (1, 1) (X, X)
(ii)
(1, 0) (3X, 0)
<T, 1>
EAg/AgCl
[V]
(a)
<1, 1>
(b)
4.25
4.20
318
4.15
4.10
4.05
(c)
4.00
(a) The CV for a Au electrode in 0.1 M HClO4 + 0.1 M NaBr. (b) The
in-plane X-ray diffraction pattern observed at potentials positive of E0 .
Positions A, B, and C correspond to the rst-, second-, and third-order
diffraction peaks from the bromide adlattice, respectively (lled circles). Also
shown is a schematic illustration of the corresponding bromide structure.
(c) The bromidebromide nearest-neighbor spacing (obtained from the X-ray
measurements) as a function of the applied potential and solution
concentration; 0.1 M (circles), 0.033 M (squares), 0.01 M (triangles), and
0.001 M (diamonds). The lines are polynomial ts to the data. (This gure is
taken from Ref. [55].)
Fig. 11
319
0.10
P1
0.08
10
c(2 22)
0.04
5
0.02
P2
Intensity
[a.u.]
0.06
c(2 p)
Current
[A cm2]
320
P3
0.00
0
P3
P2
5
0.3
0.0
0.3
0.6
ESCE
[V]
The CV for a Au(001) electrode in 0.05 M NaBr solution and the
corresponding X-ray intensity at the position (0, 1, 0.1) in which scattering from a
bromide adlayer was observed. (Taken from Ref. [57].)
Fig. 12
K
(2, 2)
(c)
pa
22a
2a
(a)
(2, 2)
2a
(b)
(d)
id
= 0.99
(11)
321
Pt(111) + 8.105 M Br
Disk current
[A]
0.4
q Br
[ML]
10
(a)
0.3
0.2
0.1
0.0
0.2 0.0
(c)
Ring current
[A]
322
Br
0.2
0.4
E
[V]
+1.08 V Br + e
2
ir
1
0.2
(b)
0.0
0.2
0.4
E
[V]
0.05
0.65 V
0.55 V
0.04
0.35 V
Intensity
[a.u.]
0.03
0.25 V
0.02
0.01
0.2 V
Fig. 15
0.00
0.66
0.68
0.70
(h, 0, 0.1)
0.72
323
324
Disk current
[A]
Pt(001) + 104 M Br
(900 rpm; 50 mV/s; 0.1 M HClO4)
[C cm2]
300
Qd
200
Qr/N
20
(c)
100
(a)
Ring current
[A]
2
Br
+1.08 V
0.2
Br2 + e
0.0
0.2
0.4
E
[V]
i
r
0.2
(b)
0.0
0.2
0.4
E
[V]
Fig. 16
bromide adsorption isotherm (charge) versus disk potential is shown in Fig. 16(c)
yielding a maximum coverage of 0.42 ML
at the positive potential limit of 0.5 V.
In the X-ray scattering experiments, no
superlattice peaks were found at any potential for Br/Pt(001) in agreement with
the STM results of Bittner and coworkers [64]. Despite the lack of adsorbate
structures with long-range order in the
surface plane, information about the surface normal structure (the PtBr spacing)
and local bonding sites for specically
adsorbed anions can be obtained from
analysis of the specular and nonspecular
CTRs [63]. Figures 17(a, b) show the specular (0, 0, l) CTR and nonspecular (1, 0, l)
CTR measured at an electrode potential of
0.2 V, where no bromide (or based on the
adsorption isotherm for Br , a negligible
325
Intensity
[a.u.]
10
0.1
(a)
(b)
1.4
326
1.2
1.0
0.8
0.6
1
(c)
(0, 0, l)
1
(d)
(1, 0, l)
CTR data for the Br /Pt(001) system: (a) (0, 0, l); and
(b) (1, 0, l) measured at 0.2 V where no Br is adsorbed on the
surface. The solid lines are a t to the data including relaxation
of the Pt surface. (c, d) show the changes in intensity measured
at 0.2 V as referenced to the data at 0.2 V. The ts to the data
are described in the text.
Fig. 17
that copper UPD onto Pt(111) in the presence of halide anions also causes the
formation of an aligned, incommensurate
hexagonal CuBr (or CuCl) bilayer structure prior to formation of a full copper
monolayer.
The absence of any ordered structures
for the Brad adlayer on Pt(001) contrasts
with the ordered structures observed on
Au(001), Au(111), and Pt(111). The ordered Brads adlayer structures observed
on Pt(111) and Au(111) are incommensurate, which means that there is no
preference for any particular site. However, the energy minimum for adsorption
at the hollow sites on (001) surfaces is
considerably deeper than for the hollow sites on the (111) surfaces. This
explains the appearance of the com
327
328
(02)
1000
( 23 43 )
(01)
(11)
(11)
( 23 23 )
( 13 23 )
1 1
( 43 13 )
(3 3)
(00)
(10)
800
( 43 43 )
F2
[eu]
( 23 53 )
(20)
600
400
200
(23 13 )
( 43 23 )
(a)
300
500
F2
[eu]
F2
[eu]
200
600
400
300
200
100
100
0
0
(c)
400
L
[r/u]
(b)
500
L
[r/u]
0.5
(d)
1.0
1.5
2.0
2.5
3.0
L
[r/u]
Fig. 18
l) and (2/3, 5/3, l); (d) (4/3, 1.3, l). The solid lines
are ts to the data according to the structural
model shown in Fig. 19. (Taken from Ref. [67].)
329
330
Fig. 19
(a)
(b)
Pt(111)
0.8
0.05 M H2SO4
5 105 M Cu2+
w = 900 rpm
Cu
0.6
0.4
0.2
i
[A]
Ei
0.0
0.0
0.1
0.2
0.3
0.4
E
[V]
(a)
0.30
0.20
100
0.10
Intensity
[a.u.]
0.40
(0, 1, 2.5)
0.35
0.30
0.25
0.20
(0, 0, 3.9)
0.25
0.20
0.0
(b)
0.1
0.2
0.3
E
[V]
331
1.3
E = 0.15 V
E = 0.0 V
1.2
1.1
1.0
0.9
1.0
Fig. 21
Intensity
[a.u.]
332
0.1
model derived for the ( 3 3)R30 Cusulfate structure on Au(111) (Fig. 19), it
can be seen that the CTR results are in
good agreement with that study. The bond
lengths obtained for the Cu-sulfate bilayer
are in good agreement with the Au study
in which it was proposed that the sulfate
is chemically bonded to the hexagonal Cu
layer. At 0.0 V, the results give a Cu coverage of 0.81 ML per Pt surface atom with
0.21 ML of sulfate anions still adsorbed on
top of the Cu layer. Because of the increase
in Cu coverage, the sulfate anions are located at an increased height above the Cu
adlayer presumably as the hollow sites in
the hexagonal Cu adlayer have been lled
by Cu ad-atoms. It is interesting that the
Cu coverage derived by the CTR measurement are in good agreement with those
obtained from the RRDE experiments and
it appears, therefore, that it is not possible to complete a full Cu monolayer on
the Pt(111) surface. The lower limit of
potential (0.0 V) was chosen to avoid the
formation of Cu+ via one-electron reduction of Cu2+ without further reduction
to metallic copper. Fits to the nonspecular CTR data at both potentials (0.0 and
0.15 V), again to both raw data and ratio
data sets, gave occupation of both types of
threefold hollow sites, fcc and hexagonal
close packed (hcp), although the fcc sites
are favored (fcc/hcp = 0.51/0.09 at 0.15 V
and 0.67/0.16 at 0.0 V). In contrast with the
Au(111) system in which the Cu ad-atoms
occupied only the fcc sites, it appears that
on Pt(111) the energy difference between
these sites may be smaller. Occupation of
both types of threefold hollow site may
333
20
0.0
900 rpm
5 mVs1
qdisk
qring
qCu
[ML]
Disk current
[A]
15
0.5
10
1.0
0.0
0.1
0.2
0.3
0.4
E
[V]
0
0.1 M HClO4
5 105 M Cu2+
1 104 M Br
(a)
ir
Ring current
[A]
334
Cu2+ + 2e
Cu
2 A
0.0
(b)
0.275 V
0.2
0.4
E
[V]
Fig. 22
qBr
[ML]
0.50
900 rpm
20 mVs1
Pt(111)
0.25
Disk current
[A]
Pt(111)/Cu
0.0
0.0
0.2
0.4
E
[V]
0.1 M HClO4
1 104 M Cu2+
1 104 M Br
5 A
Ring current
[A]
I R = 16.2 A
1 A
Br
+1.06 V
0.0
Br2 + e
0.2
0.4
E
[V]
Top: Potentiodynamic curve for Cu UPD on a Pt(111) disk
electrode. Bottom: Ring current corresponding to the change in the
Br ux during copper deposition. Insert: Surface coverage of
bromide at the disk electrode with and without copper present in
solution (assessed from ring transients).
Fig. 23
335
2.0
0.12
Intensity
Intensity
0.14
0.10
0.08
1.5
1.0
0.5
0.06
2
0.0
(O k)
0
f
[]
(a)
0.4
0.0
0.4
f
[]
(b)
(h O)
0.77
6
4
2
0
(c)
Integrated intensity
336
0.76
CuBr
CuCl
0.75
0.74
0.73
0.2
0.3
E
[V]
0.4
0.2
(d)
0.3
0.4
E
[V]
Fig. 24
337
c(2 2)
Intensity
[a.u.]
338
(a)
0.4
(b)
0.0
0.4
f
[degrees]
(c)
Fig. 25
Cu
dPt Cu
Cu
1.0 0.06
1.75 0.05 A
0.13 0.05 A
Br [c(2 2)]
dCu Br
Br
0.9 0.1
1.79 0.08 A
0.3 0.2 A
339
Fig. 26
1.6
1.4
1.2
1.0
0.8
0.6
101
Intensity
[a.u.]
340
102
103
1
2
3
(0, 0, l)
2
3
(1, 0, l)
dCu-Br
dPt-Br
Pt(111)
jd
[A cm2]
200
3.
- Anions
(4 4)
100
- Cu
0
4.
O2
2.
1.
(a)
3.
Pt(001)
jd
[A cm2]
100
(2 2)
4.
0.2
0.4
(b)
0.6
0.8
1.0
E vs RHE
[V]
Fig. 27
by Cuupd (0.5 < %Cu < 1), anions are either entirely displaced from the surface
by Cuupd (state 3a) or Cuupd and anions
form two-layer structures in which anions are adsorbed on both Pt as well as
Cuupd sites (state 3b in Fig. 27). Note that
state 3a is representative of the ordered
(4 4) Clupd Clad (Brad ) bilayer structure.
In contrast with the Pt(111)Cuupd anion
system, on Pt(001), in the same potential
341
342
O2, ad
k2
H2O2, ad
k4
k3
H2O
k5
H2O2
1G
RT
F E
RT
(12)
343
344
of these results. Rather, we will show, using representative examples, the kind of
information that can improve our understanding of the role of the local symmetry
of platinum surface atoms in O2 reduction electrocatalysis. There are two general
observations concerning the structure sensitivity of the ORR: (i) the same activation
energy in both acid (at the reversible potential ca. 42 kJ mol1 [93]) and alkaline
solution (at 0.8 Vca. 40 kJ mol1 [94]) has
been found for all three Pt(hkl) surfaces;
(ii) the structural sensitivity is most pronounced in electrolytes in which there is
strong adsorption of anions. As a consequence, the structure sensitivity of the
ORR on Pt(hkl) surfaces is mainly determined by the preexponential (1 %ad )
coverage term.
ORR at the Pt(hkl)halide Interface For the purpose of demonstrating
the importance of the (1 %ad ) term in
the kinetics of the ORR on Pt(hkl), two
representative sets of polarization curves
are shown in Figs. 28 and 29. Figure 28
shows that in the presence of Cl , the
variation in activity increases in the order (001) < (110) < (111). Figure 28(a)
shows that the ORR on the Pt(111) surface in 0.1 M HClO4 + 103 M Cl is
strongly inhibited, with an activity in a
solution containing Cl being several orders of magnitude lower than in 0.1 M
HClO4 [88, 92]. Figure 28(a) also shows
that between 0.3 < E < 0.9 V, the ring
currents were a small fraction of the disk
currents, implying that on the surface
highly covered with Clad , oxygen reduction proceeds almost entirely through the
4e reduction pathway. The appearance
of peroxide oxidation currents on the ring
electrode begins at potentials negative of
0.25 V, and parallels with the adsorption
of hydrogen on Pt(111). In contrast with
4.1.5.2.1
IR
[A]
50
iD
[mA cm2]
0
2
4
6
Pt(111)
IR H2O2
iD
;
[mA cm2] [A] [%]
(a)
2
0
2
4
Pt(110)
IR H 2 O2
;
[A] [%]
(b)
0.05 M H2SO4 + 103 M Cl
0.05 M H2SO4
50
iD
[mA cm2]
0
2
4
Pt(001)
6
0.2
0.4
0.6
0.8
E
(c)
[VRHE]
Fig. 28
345
346
900 RPM
10 A
0.4
qBr
[Ml]
0.3
0.2
Pt(111)
0.1 M HClO4
0.1
(a)
0.0
IR
[A]
150
100
50
0
Pt(111)/Br
2500
1600
900
400
RPM
ID
[mA]
400
1.0
2500
2.0
0.2
(b)
0.0
0.2
0.4
0.6
E
[V]
Fig. 29
more negative than E = 0.1 V, the surface coverage with Br decreases substantially, for example, from Br 0.35 ML
to Br 0.25 ML, and the O2 -reduction
347
348
process, the ORR currents at the Cuupd modied surfaces can also be measured
under steady state conditions, see the
closed circles in Fig. 30. In these experiments, the ORR is more inhibited than
in the potentiodynamic experiments. As
shown in Fig. 30, the deactivation is more
pronounced in the presence of Cl ions,
presumably due to stronger interaction of
specically adsorbed chloride (Clad ) than
bisulfate (H2 SO4(ad) ) anions with the PtCuupd surface. Analysis of the RRDE data
(Fig. 30) also revealed that the mechanism
for the ORR on Cuupd -modied electrodes
is the same as on unmodied Pt(111),
that is, proceeds mostly as a 4e reaction pathway with negligible solution
phase peroxide formation, (ca. 3%). It
is surprising, however, that a relatively
small amount of Cuupd has such a devastating effect on the rate of ORR. This
anomalously large inhibition by a very
small amount of Cuupd is attributed to
enhanced anion adsorption on platinum
atoms adjacent to Cuupd atoms [98], see
Sect. 4.1.4.2 and Fig. 27. The predominant
role of anions in determining the rate of the
ORR on Cuupd -modied platinum singlecrystal surfaces is supported by the fact
that the activity of the ORR decreases in
the same sequence as the Cu-anion bond
strength increases: PtCuupd HSO4,ad
PtCuupd Br . The model that rationalizes these results is one in which the active
sites for the adsorption of molecular O2
are the small number of platinum islands created in state 2 in Fig. 27. It is
important to note that in the Cu UPD
potential region, Cuupd is always covered
with anions, even at the most negative
potentials at which Pt(001) and Pt(111)
are covered by nominally 1 ML of Cuupd .
It is obvious, therefore, that is impossible to determine the true catalytic activity
of an anion-free pseudomorphic (1 1)
349
Cu
IR
[A]
6
0
0
10
20
0.2
Ei
0.2
0.4
0.6
[VRHE]
(b)
[VRHE]
1.0
0.8
0.05 M H2SO4 + Cl
10
0
0
0.8
100 A cm2
Pt(111)
20
0.0
0.5
0.6
0.4
50 A cm2
Pt(111)
1.0
1.0
Fig. 30
The effect of Cuupd on the ORR at the Pt(111) interface in (a) 0.05 M H2 SO4 and (b) 0.05 M H2 SO4 + 103 M Cl , represented at the bottom of
(a) and (b). In both cases, the solution contains 104 M Cu2+ . The corresponding Cu coverages (open circles) obtained from RRDE experiments (disk
currents are shown by solid lines) are represented at the top.
(a)
iD
[mA cm2]
Cu
IR
[A]
iD
[mA cm2]
350
(b)
Iring
[A]
2.5
0.8
2.0
1.5
1.0
0.5
0
0.0
50
0.6
Edisk vs SCE
[V]
0.4
Idisk
[A]
Cu(111)
@ 1600 rpm
0.8
10
0.4
0.2
0.5 M H2SO4
0.1 M HClO4
Edisk vs SCE
[V]
0.0
0.0
0.8
0.6
0.4
2 0.8
0.4
0.0
0.2
0.5 M H2SO4
0.1 M HClO4
Edisk vs SCE
[V]
Edisk vs SCE
[V]
Cu(001)
@ 1600 rpm
Idisk
[A]
4
0.0
0
0
50
100
(d)
(c)
Fig. 31
Polarization curves (positive sweeps) of the ORR on (a, b) Cu(111) and (c, d) Cu(001) electrodes in 0.5 M H2 SO4 (dashed lines) and 0.1 M
HClO4 (solid lines). Inserts: The corresponding CVs obtained on the disk electrodes in 0.1 M HClO4 in the absence of O2 .
(a)
Idisk
[mA]
Iring
[A]
Idisk
[mA]
100
352
the reason for this remarkable effect remains unclear. It was reported [104] that
the unique catalytic activity of the Au(001)
surface only appears in the potential region in which OHad is adsorbed on the
surface, see Fig. 32. Figure 32 also shows
that at potentials negative of the potential
for Au(001)-OHad formation, the surface
provides only 2e reduction. On the basis of this evidence, it was proposed that
the change in the catalytic activity of the
Au(001) surface from 4e to 2e reduction was due to the existence of a AuOH
layer. On the other hand, it was also proposed that the higher activity of Au(001)
for the ORR is related to the dissociative
adsorption of HO2 anions on the fourfold symmetry sites that occur only on the
(001)-(1 1) surface [105]. Consequently,
the change in the oxygen-reduction pathway from 4e to 2e reduction is due
to potential-induced reconstruction of the
(1 1) surface to a structure similar to the
hex reconstruction found in UHV. The
results in Sect. 4.1.3.2.1 (Fig. 7) indicate
that the (1 1) hex reconstruction
is not the dominant mechanism for this
change in reaction pathway since the transition in reduction kinetics occurs in a
much narrower potential range than the
(1 1) hex transition. Furthermore,
the time constant for the two transitions
differs by approximately two orders of magnitude, with the structural transition being
the slower.
Summary of the ORR
Just as in heterogeneous catalysis, the
ultimate challenge in electrocatalysis is
to relate the microscopic details of adsorbed states of reaction intermediates
to the macroscopic measurement of kinetic rates. There are many strategies
that may be employed in this endeavor.
Here we develop the relation between
4.1.5.5
353
Current
100 A cm2
900
1600
2500 rpm
2 mA cm2
1.2
1.0
0.00
0.8
0.25
0.50
0.6
0.75
Reconstruction fraction
1.4
Intensity (normalized)
354
0.4
1.2
0.8
0.4
0.0
0.4
1.00
0.8
E
[V]
Top: Polarization curves for the ORR on Au(001) in 0.1 M
KOH at different rotation rates. The dashed line shows the CV.
Bottom: Corresponding changes in the X-ray scattering signal at (0,
0, 2.3) where the scattered intensity depends on the presence of the
surface reconstruction, that is, deconstruction leads to an increase
in the intensity. The fraction of the surface that is reconstructed is
shown on the right-hand axis.
Fig. 32
355
356
4.6
Conclusion
Acknowledgment
This work was supported by the Director, Ofce of Energy Research, Materials
Sciences Division (MSD) of the U.S.
Department of Energy (DOE) under contract no. DE-AC03-76SF00098. Research
was carried out in part at SSRL, which
is funded by the Division of Chemical
Sciences (DCS), U.S. DOE and at the
XMaS UK-CRG beamline at the ESRF,
Grenoble. CAL is grateful for the support of an EPSRC Advanced Research
Fellowship.
References
1. R. Feidenhansl, Surf. Sci. Rep. 1989, 10,
105.
2. P. H. Fuoss, S. Brennan, Annu. Rev. Mater.
Sci. 1990, 20, 365.
3. I. K. Robinson, D. J. Tweet, Rep. Prog. Phys.
1992, 55, 599.
4. I. K. Robinson, Phys. Rev. B 1986, 33, 3830.
5. B. E. Warren, X-ray Diffraction, Dover Publications, New York, 1990.
6. M. G. Samant, M. F. Toney, G. L. Borges
et al., J. Phys. Chem. 1988, 92, 220.
7. B. M. Ocko, J. Wang, A. Davenport et al.,
Phys. Rev. Lett. 1990, 65, 1466.
8. I. M. Tidswell, N. M. Markovic, C. A. Lucas
et al., Phys. Rev. B 1993, 47, 16 542.
9. C. A. Lucas, J. Phys. D: Appl. Phys. 1999, 32,
A198.
10. M. F. Toney, D. G. Wiesler, Acta Crystallogr., Sect. A 1993, 49, 624.
11. C. A. Lucas, N. M. Markovic, P. N. Ross,
Surf. Sci. 2000, 448, 65.
12. C. A. Lucas, N. M. Markovic, B. N. Grgur
et al., Surf. Sci. 2000, 448, 77.
13. E. D. Specht, F. J. Walker, Phys. Rev. B 1993,
47, 13 743.
14. H. Stragier, J. O. Cross, J. J. Rehr et al.,
Phys. Rev. Lett. 1992, 69, 3064.
15. W. J. Albery, M. L. Hitchman, Ring-Disc
Electrodes, Clarendon Press, Oxford, 1971.
16. A. J. Bard, L. R. Faulkner, Electrochemical
Methods, John Wiley & Sons, New York,
1980.
17. D. Kolb, Prog. Surf. Sci. 1996, 51, 109.
18. J. Clavilier, J. Electroanal. Chem. 1980, 107,
211.
19. M. Samant, M. Toney, G. Borges et al., Surf.
Sci. 1988, 193, L29.
20. Y. S. Chu, I. K. Robinson, A. A. Gewirth,
Phys. Rev. B 1997, 55, 7945.
21. O. M. Magnussen, J. Scherer, B. M. Ocko
et al., J. Phys. Chem. 2000, 104, 1222.
22. G. A. Somorjai, M. A. Van Hove, Prog. Surf.
Sci. 1989, 30, 201.
23. G. A. Somorjai, Introduction to Surface
Chemistry and Catalysis, John Wiley & Sons,
New York, 1993.
24. P. A. Thiel, P. J. Estrup in CRC Handbook
of Surface Imaging and Visualization (Ed.:
A. T. Hubbard), CRC Press, Boca Raton,
Fla., 1995.
357
358
359
360
4.2
Charge-transfer reactions at interfaces between two immiscible electrolyte solutions (ITIES) have been the subject of
a tremendous amount of research from
an experimental and theoretical point of
view in the last thirty years. The relevance
of this kind of molecular interfaces has
been addressed in various review articles
and books, expanding from fundamental
charge transfer at interfaces to drug delivery in biological systems, ion-selective
electrodes, sensors, articial photosynthesis, and so forth [15]. The scope of this
chapter is to review the theoretical models
developed in recent years for rationalizing
the kinetic behavior of electron and ion
transfer across ITIES. This chapter does
not intend to provide an exhaustive account of experimental developments and
specic models but an overview of the
more general theoretical descriptions in a
critical fashion. In doing so, we have made
an attempt to present the basic description
of each model employing a homogenized
terminology.
The discussions will be mostly centered
on ideal systems in which the reactants
are considered as particles with given
redox properties and solvation energies
present at the interfacial region between
two dielectric media. In the case of
electron transfer (ET), we shall concentrate
4.2 Charge-transfer Kinetics across Interfaces between Two Immiscible Electrolyte Solutions
Thermodynamic Aspects in
Connection to Heterogeneous ET
The thermodynamic driving force for an
ET process across liquid|liquid junctions is
determined by the difference in the redox
potential of both species as well as the
Galvani potential difference. Considering
the general reaction
+
+ Fc+
(DCE) H(w) + Fc(DCE)
(4)
where the subscript w and DCE correspond to the water and 1,2-dichloroethane
electrolyte phases, respectively. In order to
estimate the formal Gibbs energy for Reaction (4), the following thermodynamic
cycle can be introduced:
kobs
+
Fc+
(DCE) Fc(w)
4.2.2.1
Ow
1
+ R2
kobs
Rw
1
+ O2
(1)
the equality in the electrochemical potential determines that the ratio of the
activities of the redox species is linked to
the Galvani potential difference by
w
cO1 cR2
RT
w
w
=
ln
(2)
o
o et
w c
nF
cR1
O2
1
2 H2(w)
1
2 H2(w)
(5)
+
+ Fc+
(w) H(w) + Fc(w) (6)
Fc(w) Fc(DCE)
(7)
SHE
w
GFc
(8)
361
Tab. 1
[E ,DCE ]SHE /V
Redox couple
Decamethylferrocene+/0
Dimethylferrocene+/0
Diferrocenylethane+/0
Butylferrocene+/0
Ferrocene+/0
Trianysilamine+/0
Tri(bromophenyl)amine+/0
Tetraphenylborate0/
Tetrakis(4-chlorophenyl)borate0/
Tetracyanoquinodimethane0/
Chloranil0/
Benzoquinone0/
0.07
0.55
0.55
0.56
0.64
0.76
1.10
0.93
1.15
0.29
0.17
0.43
value of w
o et in the presence of Fc
should be located close to 0.16 V. The
cyclic voltammogram features a response
with a half-wave transfer potential that co
incides with w
o et . The signal observed
at more negative potentials corresponds
to the transfer of tetrapropylammonium
employed as internal reference for the estimation of the Galvani potential difference
(w
o TPA+ = 0.093 V [23]). It should be
noted that the half-wave potential associated with the ET across the ITIES is
not necessarily correlated to the formal
Li2 SO4
1.5 M
Fe(CN)6
0.1 M
104 M
(aq)
40
TPA+
20
BTPPACl
0.01 M
0.001 M
AgCl Ag
0.01 M
20
ET
40
60
80
0.2
BTPPATPBCl
Fc 4 103 M LiCl
106 i
[A]
362
0.0
0.2
w
of
[V]
(DCE)
(aq)
4.2 Charge-transfer Kinetics across Interfaces between Two Immiscible Electrolyte Solutions
106 i
[A]
Fig. 2
180
0.24
0.04
0.06
0.16
0.26
w
o f
[V]
(a)
Absorption
2.0
1.5
1.0
0.5
0.0
600
(b)
700
800
Wavelength
[nm]
900
1000
363
364
4.2.2.2.1
w
Gact = Gact,eq obs (w
o o eq )
(9)
where Gact,eq and w
correspond
to
o eq
the Gibbs activation energy and the Galvani potential difference at the equilibrium
state. The effect on the ET rate constants
can be phenomenologically described as
obs nFw
o
kobs = kobs exp
(10)
RT
and
(1 obs )nFw
o
kobs = kobs exp
RT
(11)
where the preexponential parameters can
be evaluated at the standard potential
obs nFw
o
k = kobs exp
RT
(1 obs )nFw
o
= kobs exp
RT
(12)
From Eqs. (10 to 12), the faradaic current
as a function of the applied potential can
be expressed in terms of the concentration
of each of the reactants at the respective
interfacial reaction planes (position along
the z-axis),
w
I = nFAkobs cO1
(zO1 )cR2 (zR2 )
w
obs nF (w
o o )
exp
RT
w
(zR1 )cO2 (zO2 )
nFAkobs cR1
w
(1 obs )nF (w
o o )
exp
RT
(13)
Equation (13) can be rewritten in terms of
w
the overpotential = w
o o eq , as
w (z )c (z )
cO1
O1 R2 R2
w c
cO1
R2
obs nF
exp
RT
I = I0
w (z )c (z )
cR1
R1 O2 O2
w c
cR1
O2
(1 obs )nF
exp
(14)
RT
4.2 Charge-transfer Kinetics across Interfaces between Two Immiscible Electrolyte Solutions
where
w
w
I0 = nFAkobs [cO1
cR2 ]1obs [cO1
cR2 ]obs
(15)
If we assume that the concentrations at
the reaction planes are equal to that in
the bulk, Eq. (14) is simplied to the
corresponding ButlerVolmer expression
for ET at ITIES,
obs nF
I = I0 exp
RT
(1 obs )nF
exp
(16)
RT
This simplistic derivation opens up some
important questions that we shall address
in the next sections. Firstly, the observed
ET rate constant needs to be considered in
terms of microscopic parameters such as
the distance separating the redox species,
the adiabatic properties of the systems,
and the thermodynamic driving force. If
the reactants feature ionic species, the concentration proles at the interfaces arising
from polarization effects can introduce
substantial deviations from Eq. (16). Another parameter to be dealt with from rst
basis is obs . Does this parameter only reect the fraction of the Galvani potential
difference acting on the Gact as indicated
in Eq. (9)?
The Rate Constant of Electron
Transfer We shall commence by assuming that the liquid|liquid interface can be
described in terms of a sharp boundary between water and a lower dielectric
medium. The single ET process of Eq. (1)
involves the approaching of the reactants
to the interface as schematically represented in Fig. 3. The observed rate of ET is
given by
4.2.2.2.2
w
cR2
jobs = kobs cO1
zR2
(17)
aR2
zR2
aR2
l
l
qmax
z=0
z=0
q
a O1
zO1
(a)
a O1
(b)
(a) Schematic representation of the coordinates for the calculations of the hypervolume
as described by Marcus in Ref. [10]. (b) The maximum value of the angle for a given zR2 and l
arrangement.
Fig. 3
365
366
(18)
l=aR2 +aO1
laO1
zo =aR2
l2
l=aR2 +aO1
k(l)dl
k(lmax )
= 2(aO1 + aR2 ) 3
(19)
(21)
Equation (21) is associated with the limiting case for a sharp phase boundary
between both liquids. Marcus also estimated the hypervolume for the case in
which both reactants are allowed to penetrate the interfacial region in such a way
that their centers can effectively reach the
liquid boundary. For such conguration,
the limits of the integrations in Eq. (19)
are different and approaches to
(aO1 + aR2 )3 1
=0
k(l)
sin d d dzo dl
k(lmax )
(l aR2 aO1 )2 l
=
(22)
2
(1 + )
(23)
where
= 1 exp(2)
(24)
4.2 Charge-transfer Kinetics across Interfaces between Two Immiscible Electrolyte Solutions
and
|H |2 3/2
=
2h(kB T )1/2
(25)
exp
=0
Gact (l) 2
l gR1 (z)gO2 (z)
kB T
sin d d dz dl
(27)
(28)
367
ks
[cm4s1]
1018
2
1019
1020
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
0.8
1.0
1.2
1.4
(a)
1015
3
1016
ks
[cm4s1]
368
1017
1
1018
1019
0.0
(b)
0.2
0.4
0.6
Fig. 4
The Solvent Reorganization Energy The Gibbs energy of activation already included in Eq. (18) can be described
in terms of the total reorganization energy
4.2.2.2.3
4.2 Charge-transfer Kinetics across Interfaces between Two Immiscible Electrolyte Solutions
1.0
z=3
0.8
z = 3.25
0.6
Integrand curves
for z = 3 to 4.75
in 0.25 increments
Ho = 0.01 eV
z = 3.5
0.4
0.2
0.0
0
(a)
3
R a1 a2
1.2
z=3
1.0
z = 3.25
0.8
z = 3.5
Integrand curves
for z = 3 to 4.75
in 0.25 increments
Ho = 0.01 eV
0.6
0.4
0.2
0.0
0
(b)
3
R a1 a2
369
370
the interface
Get + ws wp
1+
4
Gact =
(29)
w
Get = nF (w
o o et )
(30)
1
1
(Df Di )2 dv
s =
20
st
20
Vw Vo
Vw Vo
dielectric media and for the general geometrical arrangement depicted in Fig. 3(a),
provides an expression for s of the form
(ne)2 1
1
1
s =
s
40 oop
o 2aR2
(ne)2 1
1
1
+
op s
40 w
w 2aO1
(ne)2
1
+
40 4zR2
op
op
s
o w
os w
op op
op
s)
o (o + w ) os (os + w
(ne)2
1
+
40 4zO1
op
op
s s
w o
w
o
op op
op s s
(w + os )
w (w + o ) w
(ne)2 2
1
1
op
op s
s
40 l
o + w
o + w
(32)
where op and s are the optical and static
dielectric constants, n is effective number
of electrons transferred, and e is the
op
op
elementary charge. Taking w = o =
s = s = s , Eq. (32) reduces to
op and w
o
the well-known expression obtained for
homogeneous ET,
hm
s
static
(Df Di )2 dv
op
n2
1
1
1
=
+
40 2aO1
2aR2
l
1
1
s
(33)
op
optical
(31)
w|DCE
n2
40
0.189 0.275
+
aR2
aO1
0.029 0.014 0.494
+
zR2
zO1
l
(34)
4.2 Charge-transfer Kinetics across Interfaces between Two Immiscible Electrolyte Solutions
wp =
s
40
4do os
4dw w
s
2 zO1 zR2
os
+
(35)
w
s + s
s + s
w
l w
o
o
It should be noticed that this expression is
obtained for a sharp boundary between
the dielectric; therefore it may not be
directly applicable in the mixed-solvent region model. The equivalent expression for
the successor complex (ws ) can be easily
obtained by replacing the charges of the
reactants by the corresponding products.
For the sake of comparison, if we introduce the parameters employed earlier for
w|DCE
estimating s
, it can be obtained that
they are of the order of 10 and 80 meV,
respectively. For this calculation we have
also assumed that zO1 = zO2 = +1 and
zR1 = zR2 = 0. It can be clearly seen that
wp and ws are expected to be rather small
w|DCE
in comparison to s
, therefore the
activation energy (see Eq. 29) is mostly
controlled by the latter. It should also be
mentioned that Benjamin and Kharkats
have developed expressions for s for all
R = 2aO1 AVm
(36)
371
372
where
pzc
kobs
= 2aO1 Vm B exp
4RT
w
nFo et
exp
RT
(39)
1 nF zO1
+
(zR2 w
o et ) (40)
2
4
(43)
Kp = R exp
(44)
RT
Similar to the previous consideration, R
corresponds to the reaction zone where
the molecules feature the appropriate
conguration for the ET step. Considering
Eqs. (20, 29, 30, and 41 to 44), an
expression for kobs identical in form
4.2 Charge-transfer Kinetics across Interfaces between Two Immiscible Electrolyte Solutions
exp
4RT
exp
wp
RT
(wp ws + nFw
o et )
RT
(45)
here eff is the effective electron hopping
frequency. It should also be noticed that if
the potential independent components of
the work terms are negligible with respect
to the reorganization energy and the thermodynamic driving force, Eq. (45) reduces
to an expression similar to Eq. (39). Apparently, this might be physically plausible if
w|DCE
, wp , and
we compare the values of s
373
pzc
ln (kobs/kobs)
3.0
(a)
O1
102 zR2
f
[V]
374
2
1
0
0.20 0.15 0.10 0.05
(b)
0.00
wof
[V]
4.2 Charge-transfer Kinetics across Interfaces between Two Immiscible Electrolyte Solutions
separating the redox couple, therefore increasing the solvent reorganization energy
(see Eq. 32). In this case, a decrease in
kobs with increasing potentials at high values of driving force has been observed.
This phenomenon was taken as rst evidence of the Inverted Marcus Region at a
polarized interface. However, the limited
number of experimental points does not
allow quantifying parameters such as the
reorganization energy term.
Barker and coworkers employed essentially the same approach but reducing the
concentration of the redox couple in the
aqueous phase. Although the analysis of
the approaching curves is further complicated by the coupling of diffusion regimes
in either side of the interface, the most
important aspect is that the overall rate
of ET decreases. Consequently, rate constants of fast reactions, which previously
appeared diffusion-controlled, were accurately measured. This analysis provided a
dependence of the rate constant on the
driving force as illustrated in Fig. 7(b). The
authors rationalized the limited number of
experimental points employing Eqs. (18,
29, and 30), providing values for the reorganization energy and the preexponential
factor of 0.55 eV and 8.3 1020 cm4 s1 ,
respectively. In principle, the value for s is
consistent with Eq. (32), assuming a faceto-face approach and for l aO1 + aR2 .
The previous works based on SECM have
not addressed the specic contributions of
the Galvani potential difference and the
formal ET potential to the thermodynamic
driving force. Our recent work based on
photoinduced ET at externally polarizable interfaces quantitatively tackles this
problem employing the thermodynamic
relations discussed in Sect. 4.2.2.1 [48].
The advantage of this approach is that
the driving force can be changed potentiostatically as well as by changing the redox
375
logk f
[cm s1]
log (k12)
[cm s1 M1]
(a)
(b)
10.0
ket
[M1 s1 cm]
376
1.0
: Fc
: DMFc
: ButylFc
: DFCEt
: DCMFc
0.1
0.01
0.001
0.0
(c)
0.5
1.0
1.5
2.0
Geto'
[eV]
energy, which shift the inverted region beyond 1 eV. The number of experimental
points obtained from a family of ferrocene
derivatives allows a condent analysis in
terms of Eq. (29). This rather high reorganization energy can also be explained in
terms of Eq. (34), assuming aFc 0.2 nm,
apor 0.7 nm, and l 1 nm.
4.2 Charge-transfer Kinetics across Interfaces between Two Immiscible Electrolyte Solutions
Finally, it is rather interesting to notice that Figs. 7(b and c) provide preexponential factors of the order of 5
1020 cm4 s1 . Comparing with the simulations in Fig. 4, it can be concluded that
the experimental data is consistent with
a highly nonadiabatic limit ( 0), in
which the expressions derived by Marcus
and Smith and coworkers do converge.
This low electronic coupling between the
redox species may arise from the difference in the solvation properties, which
induces interfacial separation distances of
the order of 1 nm.
4.2.3
4.2.3.1
Ion-transfer Reactions
I = zFA k cw k c
(46)
377
Gibbs
energy
378
m0, w + zFfw
RT ln(co/cw)
m0, o + zFfo
Reaction coordinate
Standard Gibbs energy prole for global ion-transfer
reactions.
Fig. 8
w
o i =
o,w
Gtr,i
zi F
(47)
Go,w
tr,i
I = zFA k ca k cb
(50)
4.2.3.2
4.2 Charge-transfer Kinetics across Interfaces between Two Immiscible Electrolyte Solutions
0.5
log k
[cm s1]
a = 0.48
1.0
a = 0.53
1.5
150
100
50
50
(wof wof)
[mV]
Tafel plots for ion transfer of acetylcholine. From water to DCE () and from DCE to water ().
Reaction plane
Gibbs
energy
Fig. 9
m0,w + zFfa
a
RT ln(c b/c a)
m0,o + zFfb
b
Reaction coordinate
Standard Gibbs energy prole for an elementary
ion-transfer reaction.
Fig. 10
Ion
in water
Scheme 1
Preequilibrium
Ion at
interface
in position a
Transfer
Ion at
interface
in position b
Preequilibrium
Ion in
organic phase
379
380
cb = c ezFo /RT
b
(51)
(52)
k c ezFo /RT
(53)
Furthermore, if the positions a and b are
close, we can assume that the electrical
potential difference between these two
points is negligible. In this case, Eq. (53)
reduces to
w
k c e(1)zFo /RT (54)
The proportion of the Galvani potential
difference occurring in the aqueous phase
is termed . At equilibrium, at the standard
transfer potential, we have
w
zFwo
RT = k e(1)zFo /RT
ke
=k
(55)
k = k0 ezFb /RT
a
k = k0 e(1)zFb /RT
(57)
(58)
k0 c e(1)zFa /RT ezFo /RT
(60)
4.2 Charge-transfer Kinetics across Interfaces between Two Immiscible Electrolyte Solutions
(61)
b# = (1 )ba
(62)
k0 c ezF# /RT
(63)
#w = w
(64)
#
(65)
(1 )w
k0 c e(1)zFo /RT (66)
At the formal standard transfer potential,
we have
w
zFwo /RT
= k0 e(1)zFo /RT
k0 e
=k
(67)
Ji = ci u i grad i
(68)
(69)
u i (x)grad[(x) i (x)]
with i (x) = i /zi F .
(70)
381
382
(71)
Ji
b
[(b) i (b)]
RT
zi F
ya =
[(a) i (a)]
RT
yb =
(73)
(74)
we can write
zi F [(x) i (x)]
x
= yb + (ya yb )
RT
L
(75)
If it is assumed that the ionic mobility is
constant through the interface, we have
Ji
u i
i RT
0
ya
y
b
e e
=
(a ab i )
2
2RT b
(79)
The ion ux can now be expressed as
2u i RT y cia e2y cib
Ji =
L
e2y 1
y=
u i RT y
(ca ey cib ey ) (80)
L sinh y i
(81)
k =
L
sinh y
y
u i RT
ye
k =
(82)
L
sinh y
Since it may be noted that
yey
yey
= lim
=1
y0 sinh y
y0 sinh y
lim
(83)
(84)
ki =
L
4.2 Charge-transfer Kinetics across Interfaces between Two Immiscible Electrolyte Solutions
u i RT y w zi Faw /RT
c e
L sinh y i
a
(85)
(86)
u i RT y
L sinh y
a
#
ciw ezi F (w +1/2b i )/RT
a
#
Ji =
u i RT y
L sinh y
#
w
ciw ezi F (w +1/2o i )/RT
#
w
u i RT y
L sinh y
w
w
ciw ezi F (o 1/2o i )/RT
w ]/RT
(89)
It should be noted that in order to apply
this rened model in general it is necessary
to know the potential drop across the inner
layer. We can recover Eq. (56) if we make
the assumption that is independent
of w
o and dene k according to the
following equation valid at equilibrium at
the formal standard transfer potential:
u i RT y
w
w
ezi F (o i 1/2o i )/RT =
L sinh y
u i RT y
L sinh y
w 1/2w )/RT
o i
ezi F ((1)o i
=k
(90)
Note, however, that this approach leads to
a different dependence of k on w
o i .
An interesting point concerns the fact that
when is close to 0.5, the dependence of
k on w
o i vanishes.
383
384
u w
i
Ji
b
d(ezi F (x)/RT )
zi F (x)/RT
u i
(91)
dx = RT
d(ezi F (x)/RT )
(92)
RT u w
a
i
=
cia ezi Fw /RT ciw
(93)
w
A
b
ci ezi F /RT cib ezi F /RT
Ji = RT u i
ezi F (x)/RT dx
=
RT u i
(94)
A =
=
(95)
b
b
(96)
+
c
(98)
cib = ezi Fo
i
RT u i
In the case of small Aw and A , the traditional Boltzmann distribution is recovered.
This situation corresponds to the case in
which the potential drops between a and
w and between b and o are small.
These expressions can be inserted into
Eq. (50) to obtain the following result:
w zi Faw /RT
k ci e
b
k ci ezi Fo /RT
(99)
Ji =
Aw zi Faw /RT
1+ k
e
RT u w
i
A
b
k
ezi Fo /RT
RT u i
Note that a potential dependence may
easily be incorporated into an expression
4.2 Charge-transfer Kinetics across Interfaces between Two Immiscible Electrolyte Solutions
transport process:
(ciw ey ezi Fw /RT
u i y
b
ci ey ezi Fo /RT )
L sinh y
w
Ji =
A
a
1+
ey ezi Fw /RT
RT u w
i
A y z Fb /RT u i y
i
o
e
e
RT u i
L sinh y
(100)
As in Eq. (80), we have here specied a
linear potential prole across the interface.
potential distribution when realistic parameters of the ion and diffuse layers
are used [62]. From these considerations
it may thus be concluded that traditional
steady state current voltage curves generally do not give much information about
the inner layer processes.
In general, experimental studies on iontransfer reactions lead to values of apparent rate constants. In this connection it is
interesting to note that the nominator in
Eq. (99) is essentially similar to the expression for Ji in Eq. (66) obtained from the
simple ButlerVolmer approach. It thus
immediately follows that the correction
of an experimentally determined apparent
rate constant based on Eq. (66) (equivalent
to the Frumkin correction used for ET at
solid electrodes) is not in direct agreement
with the more general treatment leading to
Eq. (99). This point was rst recognized by
Senda [63], who termed the denominator
of Eq. (99) the Levich correction.
The Marcus Model for Ion Transfer
at Liquid|Liquid Interfaces
Recently, Marcus has attempted to treat
the ion transfer as a combined activationcontrolled and transport-controlled process [68]. In the Marcus theory for ion
transfer between two immiscible solutions, the ion transport is described using
a NernstPlanck type equation [57]. This
equation leads to an expression of k similar to that obtained by Kakiuchi [60],
4.2.3.6
k =
uRT
D
=
L
L
(101)
385
386
(102)
(104)
1
A
kassn
1
Ak
Keq
diff
1
B kB
Keq
diss
(105)
In this equation the rate constants and
equilibrium constants may be estimated
using the transition state theory. In this
treatment an activation energy can easily
be incorporated in the expressions for k A
B using the transition state theory.
and kdiss
A
kassn
= k A P (ha )ha
A
Keq
=
2
1
2g r P (ha )ha
B
=
Keq
2
1
2g r P (hb )hb
(106)
(107)
(108)
4.2 Charge-transfer Kinetics across Interfaces between Two Immiscible Electrolyte Solutions
1
for z 0
z/
2e
f1 (z) =
1
for z 0
2ez/
(110)
f2 (z) = 1 f1 (z)
(111)
387
388
4.2 Charge-transfer Kinetics across Interfaces between Two Immiscible Electrolyte Solutions
389
390
4.3
Introduction
This equation makes it clear that electrocatalysis consists of two terms: an electrical
one (the term depending on the potential difference across the interface) and
a chemical one (the Arrhenius term),
which depends on the nature of the electrode material.
Although by applying high overpotentials one can increase the rate of an
electrochemical reaction on a given electrode/catalyst even by more than 10 orders
of magnitude, the activity of the electrode,
the Arrhenius term, still remains the most
important fact in electrocatalysis. This is
obvious, since the energy efciency of any
cell is determined mainly by the overpotentials necessary at the anode and cathode.
The lower the overpotentials, the better
is the energy efciency. Furthermore, the
application of high overpotentials usually
introduces problems due to the interference of undesired reactions (e.g. the
hydrogen or the oxygen evolution reaction,
the electrodissolution of the electrode,
etc.).
The catalytic activity of an electrode
depends on its electronic and perhaps
crystallographic properties. When the reactants or intermediates are adsorbed on
the electrode surface, the reaction rate will
depend additionally on the heat of adsorption, on the strength of bond-forming or
bond-breaking, and on the properties of
the adsorption isotherm. All these factors
may be changed by the modication of the
electrode surface from adsorbed species
existing in the solution, not participating
in the reaction. These species can sometimes improve the catalytic activity of the
bare material. For example, heavy metal
atoms (e.g. Pb, Tl, and Bi) deposited in the
underpotential region have been shown
to increase the catalytic activity of noble metal electrodes (e.g. Pt, Au, Ag, Pd,
etc.) for several technologically interesting
391
392
E = Eads ENernst 0
(2)
M z+ + ze
Mads
(3)
Ep = 0.5
(6)
393
394
Electrocatalysis by Ad-atoms
(7)
395
396
O2 + 2H2 O + 4e 4OH
(alkaline solutions) (8)
O
M
O O
O O
2H2 O
(acid solutions)
(9)
OH
O2 + H2 O + 2e HO2
(H2 O + 2e )
3OH
(alkaline solutions)
(10)
Depending on the electrode material, the
reduction of oxygen occurs by one or the
other reaction pathway. The two reaction
pathways may also take place in parallel.
The most powerful technique for the
quantitative determination of the extent
of these reactions was proved to be the
rotating ring-disc electrode voltammetry,
which allows the detection of hydrogen
peroxide on the ring electrode.
On some electrodes, hydrogen peroxide cannot undergo further reduction and
is obtained as the nal reaction product. Electrodes on which the two-electron
reduction to hydrogen peroxide predominates are Au, Hg, and various carbon
materials. Electrodes on which the fourelectron reduction to water is predominant
include Ag, Pt, and the Pt family metals.
Platinum is considered to be the best electrocatalyst for oxygen reduction, since the
reduction on this electrode occurs with the
lowest overpotential. The reaction pathway that is more favorable depends on the
way molecular oxygen is adsorbed on the
metal surface. Three different models for
adsorption have been proposed, the Grifths model, the Paulings model, and the
bridge model [38, 39]:
j
[A cm2]
4
0
4
(b)
8
0.0
j
[mA cm2]
0.4
(c)
E vs NHE
[V]
0.1
0.2 5
5.10
5.103
cPb2+/M
0.8
4
2 3
1
1.2
(a)
0.2
0.0
0.2
E vs NHE
[V]
Fig. 1
397
0.2
iR
[A]
iR
[mA]
0.8
0.1
0.0
Pt
Pt/Pb
0.0
0.8
0.0
Au
Au/TI
400
iD
[mA]
iD
[mA]
398
0.4
1.6
0.8
2.4
1400 rpm
1.2
0.6
(a)
0.4
0.2
0.0
ED vs Hg/HgO
[V]
0.0
(b)
0.4
0.8
ED vs RHE
[V]
Fig. 2
(2e + 2H+ )
RNO2 RNO
H2 O
(2e + 2H+ )
RNHOH
(2e + 2H+ )
RNH2
H2 O
(11)
399
Au(100), Br
(i)
Pt(111), Br
50
0
i
[A]
j
[A cm2]
0
(3 3)Br
4
50
0.8
0.0
0.4
0.0
(ii)
c(2 22)
0.0
0.4
i
[mA]
Intensity
(ii)
j
[mA cm2]
400
(iii)
Pt(111), Br
0.8
Au(100), Br
0.2
Pt(111)
Au(100)
1.2
1.4
0.0
(a)
0.4
0.8
E vs NHE
[V]
0.0
(b)
0.4
0.8
E vs NHE
[V]
j
[mA cm2]
0.0
0.1
2 34 5
0.2
(a)
j
[mA cm2]
0.0
0.1
1
(b)
0.2
j
[mA cm2]
0.1
0.0
0.1
0.2
(c)
0.0
0.2
0.4
E vs NHE
[V]
X-ray diffraction [15] and STM [20] studies. Also, the second reduction wave
of nitrosobenze, which corresponds to
the reduction of phenylhydroxylamine to
401
402
+ 2e + 2H+
NO2
PhCH2NO2 + HNO2
(12)
(Pt/Mads): Ph-NO 2
Scheme 1
**
Ph-N
*
O
*
O
2e + 2H+
Ph-NO
H2O
2Hads
H2O
2e + 2H+
NO2
PhCH
+ 2e + 2H+
NO2
NOH
PhC
NO2
+ H2O
(13)
Reaction (12), producing phenylnitromethane, predominates on Pt/Tl(upd),
while on Pt/Pb(upd) the major reaction is the reduction to -nitrobenzaldoxime.
Also, in the case of the reduction of
benzofuroxan [59] the UPD adlayers of Pb,
Tl, and Bi strongly modify the catalytic
activity and selectivity of the platinum
electrode. Figure 5 shows the rotatingdisc voltammograms for benzofuroxan
reduction on Pt and Pt/M(upd), M = Pb,
Tl, Bi.
On bare platinum the reduction proceeds via catalytic hydrogenation to
o-nitroaniline (Scheme 2), which is reduced further to o-phenylenediamine. On
the other hand, on UPD-modied platinum surfaces, the reduction occurs at
much more positive potentials and follows
the electronation mechanism, that is, the
direct exchange of electrons between the
dinitroso tautomeric form and the modied electrode surface. The rst step is now
the reduction to o-benzoquinone dioxime
that appears as a stable intermediate over
a wide potential range (0.60 to 0.45 V)
before it is reduced further to the nal
products.
*** *
Ph-N-O
mHads
r.d.s
Ph-NHOH
2e + 2H+
Ph-NH2
H2O
Products
(a)
0
1.5
(b)
3.0
4
j
[mA cm2]
j
[mA cm2]
(c)
0
4.5
(d)
0.2
6.0
0.0
0.4
0.8
1.2
E vs NHE
[V]
7.5
0.0
0.4
0.8
E vs NHE
[V]
Current-potential curves for benzofuroxan (103 M) reduction on a
Pt rotating-disc electrode in 0.5 M HClO4 without (1) and with (2) 103 M
TlClO4 ; (3) 103 M Pb(ClO4 )2 ; (4) 103 M Bi(ClO4 )3 ( = 10 mV s1 ;
f = 25 Hz). Inset: Cyclic voltammograms for the UPD of Pb, Tl, and Bi on
Pt in 0.5 M HClO4 . (a) Without ions; (b) with 5 104 M Pb2+ ; (c) with
5 104 M Tl+ ; and (d) with 5 104 M Bi3+ . Scan rate = 50 mV s1 .
(Reproduced from Ref. [59] with permission.)
Fig. 5
Pt
NO2
O*
O + 2Pt
N
2Hads
N*
NH2
O*
**
N
NO2
O*
+ 4Pt
mHads
Products
NH2
NH2
Pt/M(upd)
O
N
2e + 2H+
O
N
NOH
NO
NO
NOH
403
N
N
NO2
NO2
O2N
NO2
N
H
160
0
(i)
3
4
10
20
ER = 1.6 V
w1/2
[s1/2]
0.2
jD
[mA cm2]
(ii)
0.1
0.0
1
2
10
ER = 0.5 V
20
30
2
3
0.1
0.2
0.0
0.4
E vs SHE
[V]
2
1
(a)
80
1
2
n=4
0
Au
Au Pb
n=2
iR
[A]
1
2
iR
[A]
jL
[mA cm2]
j
[mA cm2]
j
[mA cm2]
404
0.8
(b)
0.0
0.4
ED vs SHE
[V]
UPD of Pb on Au in 0.2 M
HClO4 cPb 2+ = 5 104 M; = 50 mV s1 .
(b) AuAu rotating ring-disc (N = 0.18)
measurements for 3-nitro-1,2,4-triazole
(103 M) in 0.2 M HClO4 + 5 104 M
Pb(ClO4 )2 ( = 20 mV s1 ). f/Hz: (1) 12.5;
(2) 25; and (3) 50. (Reproduced from Ref. [56]
with permission.)
OH
OH
N
NO2
N
2e + 2H
OH
kfast
kslow
N
N
NHOH
NO
N
2e + 2H+
H
Au/M(upd) (low coverage)
O
Ar
N
2e
r.d.s
Ar
HO
2H+
OH
2e + 2H+
fast
ArNHOH + H2O
Ar
N
(M = Pb, Tl)
Au M
Scheme 3 Reduction mechanism of 3-nitro-1,2,4-triazole on Au and Au/M(upd)
(M = Pb, Tl).
H2O
405
406
rst two fuels to CO2 requires the donation of oxygen, which comes either
from water or hydroxyl ions depending
on the pH of the solution. In aqueous
acid solutions the overall reaction may be
written as
CHx Oy + (2 y)H2 O
CO2 + [2(2 y) + x](H+ + e ) (14)
Since all hydrogen atoms must be abstracted from the fuel, the electrocatalyst
must support the dehydrogenation. Electrodes that can be used as efcient anodes
in C1 electrocatalysis are those with great
afnity toward dehydrogenation reactions,
namely, Pt and some other metals of the
Pt-Group. Unfortunately, the current densities obtained with these electrodes are
very low owing to poisoning effects. The
main effect of poisoning is to block sites
on the electrode surface.
Many efforts have been made to identify
the nature of the poisoning species. The
most successful techniques are the in situ
IR reectance spectroscopies. Intensive
work has been done both on polycrystalline
and single-crystal electrodes. The work
done has been reviewed by Bewick and
Pons [62], Beden and Lamy [63], and more
recently by Sun [64]. It is now recognized
that CO (either linearly or bridge-bonded
to the surface) is the poison and not formyl
species (COH), as believed previously [65,
66]. Carbon monoxide is formed by dehydration in the case of HCOOH and by
dehydrogenation in the case of HCHO and
CH3 OH. The CO formation reaction constitutes the one branch of the so-called
dual mechanism, originally suggested
by Capon and Parsons for the electrochemical oxidation of C1 molecules [65, 66]. The
other branch is the reaction through reactive intermediates.
CO2
[RI]
407
Pt
Pt/Pb
60
j
[mA cm2]
408
40
20
0
0.0
0.4
0.8
1.2
1.6
E vs NHE
[V]
Voltammetric curves for formic acid (0.265 M)
oxidation on a Pt electrode in 1 M HClO4 without (dashed
line) and with (full line) 103 M Pb2+ ions. Scan rate
= 50 mV s1 . (Reproduced from Ref. [71] with permission.)
Fig. 7
(a)
j
[mA cm2]
j
[mA cm2]
20
10
0.1
0.0
0.5
E
[VRHE]
0.1
0.1
0.5
0.9
j
[mA cm2]
j
[mA cm2]
20
10
0.2
0.0
0.2
0.1
(b)
0.5
0.5
E
[VRHE]
0.9
E vs RHE
[V]
409
410
the increase of the carbon chain. The formation of CO results from the dissociative
chemisorption and the C1 C2 bond breaking. In the case of D-glucose oxidation on
Pt, the poisoning species is believed to be
a gluconolactone type intermediate [113,
114], which is also considered as the reactive intermediate producing gluconic acid
upon hydrolysis.
O
R
O Pt
OH
C1 + 2Pt
(2H+ + 2e)
C1 Pt
R
O
R
C1 O + 2Pt
(15a)
OH
O OH
HO
OH
+ 2Pt
OH
OH
O O
HO
OH
OH
Pt
Pt
OH
O
HO
OH
O + 2Pt
Hydrolysis
OH
(15b)
However, more recent studies by in situ IR
spectroscopy [115, 116] have demonstrated
that the catalytic decomposition of glucose
Selective oxidation
of glycolaldehyde on Pt and
Pt/Sbads .
Scheme 4
CH2OHCHO
Pt/Sbads
CHOCHO (glyoxal)
CH2OHCOOH (glycolic acid)
411
412
jp
[mA cm2]
j
[mA cm2]
3
2
1
0
20
10
v
[mV s1]
1
Ep
[V]
0.93
0.81
0.79
0.89
1.4
Pt
0.91
Pt/TI
0.77
1.0
0.6
0.2
log (v )
[V s1]
0.2
0.6
1.0
E vs NHE
[V]
Cyclic voltammograms for pyrocatechol (2.5 103 M) on Pt in
0.5 M HClO4 without (dashed lines) and with (full lines) 103 M TlClO4 .
Scan rate /mV s1 : (1) 50; (2) 100; (3) 200; (4) 300; and (5) 400. The
insets show plots of jp versus 1/2 and Ep versus log without (1) and
with (2) TlClO4 . (Reproduced from Ref. [120] with permission.)
Fig. 9
inhibition effect was correlated to the superlattice structures of the UPD ad-atoms
on the different Ag single-crystal surfaces.
4.3.5
Conclusion
413
414
to develop new, more efcient electrocatalysts for these reactions. On the other
hand, studies on ad-atom-modied electrodes have allowed us to understand how
surface geometric and electronic structure
inuences catalytic activity and reactivity.
Signicant progress has been made toward this direction by using single-crystal
electrodes with well-dened structures of
substrate and ad-atom layers in combination with in situ spectroscopies and
scanning probes. These techniques have
provided information on reaction intermediates (reactive or poisons) and surface
properties under operating conditions that
can help us to directly correlate the surface structure and composition with the
catalytic activity. We already know enough
about oxygen reduction, but there is still
much to learn about the oxidation of organic fuels.
Except the enhancement of the electrocatalytic activity, ad-atoms may also
improve the selectivity of the surface of
the electrocatalysts regarding the nal reaction products. Little work has been done
in this direction. Probably in the future,
electrodes modied by ad-atoms will be
more attractive as potential catalysts in selective electrocatalysis. The improvement
of the selectivity would require thorough
mechanistic studies in order to get a
better understanding of the various factors that may inuence the selectivity of
these electrodes.
References
1. J. O.M. Bockris, A. K. N. Reddy, Modern
Electrochemistry, Plenum Press, New York,
1977, Vol. 2.
2. W. J. Lorenz, H. D. Herrman, N. Wuthrich
et al., J. Electrochem. Soc. 1974, 121,
11671177.
3. D. M. Kolb in Advances in Electrochemistry and Electrochemical Engineering (Eds.:
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
415
416
92. D. Sazou, N. Xonoglou, G. Kokkinidis, Collect. Czech. Chem. Commun. 1986, 51,
24442454.
93. C. T. Hable, M. S. Wrighton, Langmuir
1993, 9, 32843290.
94. G. Kokkinidis, D. Jannakoudakis, J. Electroanal. Chem. 1982, 133, 307315.
95. F. Kadirgan, B. Beden, C. Lamy, J. Electroanal. Chem. 1982, 136, 119138.
96. F. Kadirgan, B. Beden, C. Lamy, J. Electroanal. Chem. 1983, 143, 135152.
97. M. Shibata, N. Furuya, M. Watanabe, J.
Electroanal. Chem. 1989, 267, 163170.
98. G. Pierre, A. Ziode, M. L. Kordi, Electrochim
Acta 1987, 32, 601606.
99. P. Ocon, B. Beden, H. Huser et al., Electrochim. Acta 1987, 32, 387394.
100. P. Ocon, B. Beden, C. Lamy, Electrochim.
Acta 1987, 32, 10951101.
101. M. Sakamoto, K. Takamura, Bioelectrochem.
Bioenerg. 1982, 9, 571582.
102. N. Xonoglou, G. Kokkinidis, Bioelectrochem.
Bioenerg. 1984, 12, 485498.
103. G. Kokkinidis, N. Xonoglou, Bioelectrochem.
Bioenerg. 1985, 14, 375387.
104. N. Xonoglou, I. Moumtzis, G. Kokkinidis,
J. Electroanal. Chem. 1987, 237, 93104.
105. G. Kokkinidis, J.-M. Leger, C. Lamy, J. Electroanal. Chem. 1988, 242, 221242.
106. M. Shibata, S. Motoo, J. Electroanal. Chem.
1985, 187, 151159.
107. M. Shibata, S. Motoo, J. Electroanal. Chem.
1986, 201, 2332.
108. B. Beden, M. C. Morin, F. Hahn, C. Lamy,
J. Electroanal. Chem. 1987, 229, 353366.
109. H. Hitmi, E. M. Belgsir, J.-M. Leger et al.,
Electrochim. Acta 1994, 39, 407415.
110. S.-G. Sun, D.-F. Yang, Z.-W. Tian, J. Electroanal. Chem. 1990, 289, 177187.
111. E. Pastor, S. Wasmus, T. Iwasita et al., J.
Electroanal. Chem. 1993, 350, 97116.
112. F. Hahn, B. Beden, F. Kadirgan et al., J.
Electroanal. Chem. 1987, 216, 169180.
113. S. Ernst, J. Heitbaum, C. H. Hamann, Ber.
Bunsen-Ges. Phys. Chem. 1980, 84, 5055.
114. Yu. B. Vassiliev, O. A. Khazova, N. N. Nikolaeva, J. Electroanal. Chem. 1985, 196,
105125, 127144.
115. T. T. Bae, E. Yeager, J. Electroanal. Chem.
1991, 309, 131145.
116. F. Largeaud, K. B. Kokoh, B. Beden et al., J.
Electroanal. Chem. 1995, 397, 261269.
117. M. Shibata, N. Furuya, M. Watanabe, J.
Electroanal. Chem. 1993, 344, 389393.
417
418
4.4
Introduction
Surfactant-stabilized Microheterogeneous
Fluids
Structures of Micelles and
Microemulsions
4.4.2.1
+
N
Br
CH3
DDAB-didodecyldimethylammonium bromide
O
O
P
O
DHP-dihexadecylphosphate
CH3
CH3
Br
CH3
CTAB-cetyltrimethylammonium bromide
Na+ SO4
SDS-sodium dodecylsulfate
CH2
OOC
CH
OOC
CH2
O3POCH2CH2N+(CH3)3
PC-phosphatidylcholine (lecithin)
N
+
Br
Cetylpyridinium bromide
HO(CH2CH2O)23
Brij-35 polyoxyethylene(23)dodecyl ether
Fig. 1 Amphiphilic molecules (surfactants) used to make
micellar solutions and microemulsions.
419
420
Water
Water
Globular micelle
Oil
Oil
Water
Oil
Rod-shaped micelle
Oil
Inverse micelle
Fig. 2
via [10]
Vc = m (27.4 + 26.9n )A
(1)
(2)
v = 27.4 + 26.9n A
(3)
Oil
Oil
Water
Water
Oil
Water
o/w microemulsion
Oil
Oil
Oil
Bicontinuous microemulsion
Water
Oil
Oil
w/o microemulsion
Fig. 3
421
422
Solute
+
Solute
Fig. 4
Solute
423
424
4.4.3
O + ne = R
(4)
nF
RT
1/2
1/2 D 1/2
(5)
where n is the number of electrons
transferred per molecule, F is Faradays
constant, A is the electrode area, C* is the
solute concentration, R is the gas constant,
T is the temperature in Kelvin, and is
the scan rate. Equation (5) predicts a linear
plot of Ip versus for a reversible reduction
or oxidation. The slope of this plot can be
used to estimate D in a micellar solution
or microemulsion.
Linear sweep voltammetry at ultramicroelectrode disks of radius r < 10 m under
mass transport control, usually achieved
at scan rates <50 mV s1 , provides a limiting current IL that depends directly on
D [25]:
IL = 4nF D C r
(6)
A number of other electrochemical methods can also be employed to obtain D [4, 9].
D is inuenced by binding to micelles,
microemulsion droplets, or surfactant
monolayers at o/w interfaces, as well
(7)
(8)
reaction:
StokesEinstein equation:
D1 =
kT
6rh
M + mX = MXm
(9)
Km
(10)
[MXm ]
=
[M][X]m
(11)
D0
1 + CM K m CXm1
CM K m CXm1 D1
1 + CM K m CXm1
(12)
10
107D
[cm 2 s1]
0.5
1.5
2.5
3.5
Fc
[mM]
Inuence of the concentration of ferrocene (Fc) on D measured by cyclic voltammetry
at a glassy carbon electrode in 0.15 M CTAB/0.1 M tetraethylammonium bromide. Points are
experimental: line is best t to Eq. (12) with n = 3 by nonlinear regression, giving
D1 = 3.3 107 cm2 s1 for the rod-shaped micelles. (Reproduced with permission from
Ref. [4], Copyright 1994 by Marcel Dekker.)
Fig. 5
425
106D
[cm2 s1]
426
0.00
0.50
1.00
1.50
B12
[mM]
Inuence of concentration of cobalt(II) corrin complex vitamin B12r
on D measured by linear sweep voltammetry at a 10-m radius carbon
ultramicroelectrode in w/o microemulsion made with 0.2 M aerosol AT,
4 M water, and isooctane. Points are experimental: line is best t to
Eq. (12) with n = 3 by nonlinear regression giving
D1 = 0.63 106 cm2 s1 for the water droplets. Using this value with
Eq. (9) gave a hydrodynamic radius of 7.5 nm. (Adapted with permission
from Ref. [27], Copyright by American Chemical Society.)
Fig. 6
K0 =
[AW ]
[AM ]
and
Kr =
[BW ]
[BM ]
r
EM
= EW +
RT
K0 (1 + Kr )
ln
nF
Kr (1 + K0 )
D
RT
ln 0
2nF
Dr
r
EM
= EW +
RT
K0
ln
nF
Kr
(14)
ln
ln 0
nF
Kr
2nF
Dr
(15)
An example is ferrocene bound to
positively charged alkyltrimethylammonium micelles, where oxidation product ferricinium ion dissociates when
formed [29].
r
EM
= EW +
AM + ne
K0
Scheme 1
(13)
M + AW + ne
BM
Kr
M + BM
427
428
k = k0 (1 ) + k1
(16)
#
MXn
$
MXn1 + X
(17)
#
X $
adsorbed on electrodes from microemulsions made with 3, oil, water, and alcohol.
Comparisons of electrochemical k values of surfactants 1 and 2 in microemulsions and micellar solutions helped establish qualitative pictures of dynamics
at relevant electrode-uid interfaces. Rate
constants for oxidations of ferrocene (Fc)
2-Fc (1) and 5-Fc (2) were similar in homogeneous DMF and DMSO on Pt and glassy
carbon electrodes [32, 33]. However, in
aqueous CTAB micelles, electron transfer
rates were in the order Fc > 2-Fc > 5-Fc,
with tenfold differences in successive values. This was attributed to 2-Fc and
5-Fc achieving head down orientations on
the electrode prior to electron transfer.
Adsorbed CTA+ on the electrode seems to
help order 1 and 2 on the electrode prior
to electron transfer.
In a bicontinuous CTAC microemul
sion, ferrocene had a k twice as large
as 1 and 2 [33], but values for 2-Fc and
5-Fc were similar. These relatively small
differences in electron transfer rates in
the microemulsion cannot be explained by
head downtail up orientation of 1 and
2 at the time of electron transfer as proposed for micellar solutions. The results
suggest an increased disorder and mobility in the electrodeuid interface in the
CTAC microemulsion compared to micellar CTAB solutions.
Alcohols are often used as cosurfactants in microemulsions, and insight has
been obtained from the electrochemistry
of ferrocene alcohols (4). Oxidations of
FcOHC10 , FcOHC14 , and FcOHC18 [34]
were nearly reversible and controlled by
diffusion in microemulsions of DDAB,
CTAC, or SDS. In micellar solutions,
electrode reactions were more complex
and reected strong adsorption of the
ferrocene alcohols onto the electrode.
CH3
CH3
N +
OOCFc
CH3
(1)
2-Fc
CH3
CH3
N+
OOCFc
CH3
(2)
5-Fc
CH3
CH3
CH2
(CH2)n1CH3
Fc-Cn, n = 8,12,16
(3)
Fc
CH3
HO
N
Fc
(CH2)n1CH3
(4)
FcOH-Cn, n = 10,14,18
Electroactive
ferrocene-containing surfactants.
Fig. 7
429
430
H
N
N
H
431
432
10
11
12
kM [Q]M [A]M
[Q][A]
(19)
P + e
Q+A
P+B
kM
2
M
(20)
kobs =
Q
k1
ArX + Q
E (at electrode)
ArX + P
k2
ArX
Electrochemical
catalytic reduction of aryl
halides.
Scheme 3
Ar + H+ + e
Ar + P
ArH
433
434
4.4.4.4
4.4.4.4.1
Co(II)L, 2e
+ 2 Br
(21)
In bicontinuous DDAB microemulsions,
CoII L mediators reside in the water
phase and DBCH in the oil phase, and
the reaction (Eq. 21) probably occurs at
the o/w interface. Cyclic voltammetry
(Fig. 8) shows the reversible CoII /CoI
reductionoxidation peaks of the mediator
and an increase in reduction current when
DBCH is added. Direct reduction of DBCH
occurs at a much more negative potential.
The catalytic current was used to estimate the apparent rate constant k1 for
the DBCH reaction with CoI L [58]. A
linear plot of log k1 versus E of the
mediator (Fig. 9) in DDAB microemulsions and organic solvents was obtained.
A similar linear plot was found for the
catalytic reduction of benzyl bromide in
microemulsions [59]. These plots suggest
that the bimolecular reactions in these
microemulsions are controlled mainly by
I
[A]
25
a
0
0.60
0.80
1.00
1.20
1.40
1.60
1.80
E vs SCE
[V]
Cyclic voltammograms at 0.1 V s1 on glassy carbon electrodes in a DDAB
microemulsion: a) 0.4 mM Co(salen) alone; b) 0.4 mM Co(salen) +1.5 mM DBCH; and c)
1.5 mM DBCH alone without catalyst. (Adapted with permission from Ref. [58], Copyright
by American Chemical Society.)
Fig. 8
435
6
Co(salen)
B12
CoOEP
log k1
[M1 s1]
436
4
CoTPP
2
CoPcTS
1.2
1.0
0.80
0.60
0.40
E ' vs SCE
[V]
Inuence of catalyst formal potential (E ) on log k1 for reaction of CoI L with DBCH
in a bicontinuous microemulsion (%) of DDAB/water/dodecane (21/39/40) and in
dimethylformamide () for dissolved catalysts vitamin B12, Co(salen), cobalt
phthalocyaninetetrasulfonate (CoPCTS), cobalt tetraphenylporphyrin (CoTPP), and cobalt
octaethylporphyrin (CoOEP). Points from reactions in the microemulsion (%) represent
apparent k1 values for 0.4, 0.5, 1.0, and 2.0 mM catalyst in order of decreasing log k1 .
(Adapted with permission from Ref. [58], Copyright by American Chemical Society.)
Fig. 9
CoIIL + e
RX + CoIL
RCoIIIL + X
R CH2CH2Z + CoIIL
Co(I)L + RI
O
RCo(III)L + I
O
hn
RCo(III)L +
+ Co(II)L
R
13
O
Br
Co(III)L
+ Co(I)L
+ Br
14
O
O
Co(III)L
hn
+ Co(II)L
15
Scheme 6
O
Br
Co(II)L
1.5 V
Scheme 7
16
17
437
438
Summary
This chapter describes the electrochemistry of small reactants dissolved in micellar solutions and microemulsions. A
major inuence of these microheterogeneous uids on reversible reactants is
slowing down mass transport. These phenomena enable electrochemical probes to
be used to characterize aggregate mass
transport and size in the uids. Tuning the
compositions of micelles and microemulsions can control pathways and kinetics of
direct organic reactions, polymerizations,
and mediated electrochemical reactions.
Acknowledgments
The authors work on micelles and microemulsions described in this chapter was
initially supported by US PHS Grant No.
ES03154 from the National Institute of Environmental Health Sciences, NIH. More
recent synthetic aspects were supported by
Grants nos. CTS-9306961, CTS-9632391,
and CTS-9982854 from NSF. The contents
of this chapter are solely the responsibility of the author and do not necessarily
represent ofcial views of NIH or NSF.
References
1. P. L. Luisi, L. Magid, CRC Crit. Rev. Biochem.
1987, 20, 409474.
2. M. Bourrel, R. S. Schechter, Microemulsions
and Related Systems, Marcel Dekker, New
York, 1988.
439
443
5.1
Electrocatalysis
Enrique Herrero, Juan M. Feliu, Antonio Aldaz
Universidad de Alicante, Alicante, Spain
5.1.1
444
(1)
E=E +
+
RT
kR
ln
nF
kOx
(j jlim,Ox )
RT
ln
nF
(jlim,R j )
(2)
Whenever a process studied with a stationary technique follows Eq. (2), the process
is mass controlled and improvement in the
reaction rates will not increase the overall
reaction rate. If not, it may be interesting
to nd a better electrocatalyst.
Improvement of the reaction rates is
not the only attribute sought in an
electrocatalyst. Often, a higher selectivity
versus any competitive reaction that may
take place is also required. For that reason,
catalysts given higher reaction rates but
lower material yields are normally not
used. Thus, a good electrocatalyst is the
one that gives high current densities at
low overpotential and with high selectivity
for the desired product.
When studying the electrocatalytic behavior of a given material, several properties may inuence its overall performance:
the electronic properties, the surface structure, and the different types of sites present
on it, the particle size, surface composition, and so on. The variety of these
5.1 Electrocatalysis
445
A
1
Energy
446
2
Ox
3
P
G
Reaction coordinate
Schematic free energy prole scheme for an electron
electroreduction. O, oxidized species; R, reduced species; A, transition
state; and P and S, the precursor and successor states, respectively, that
) for a large distance
precede and follow the transition state. (
between the reactive species and the electrode. (-----) Energetics of the
process in which the reacting species come closer to the electrode surface
and these species have attractive interactions with the electrode surface.
( . . . . ) Same as a dashed line but with a stronger overlap.
Fig. 1
5.1 Electrocatalysis
k1 = A1 exp
4RT
(3)
k2 = A2 exp
(4)
2RT
Dividing Eqs. (3 and 4) leads to
k1
=
A1
k2
A2
1/2
(5)
When the reaction rates for the homogeneous and heterogeneous reaction rates
fulll Eq. (3), it is clear that the reaction
is taking place through an outer-sphere
mechanism, and the electrode material is
not acting as a true electrocatalyst.
Adsorption and Electrocatalysis
As stated previously, adsorption of some
species involved in the electrochemical reaction is the key parameter to consider
an electrode material as an electrocatalyst.
Therefore, it is important to relate the activity of such a material with its adsorption
properties. Let us consider rst the simplest reaction A + e B, in which A is
the only species adsorbed on the catalyst,
and try to relate the reaction rate to the
adsorption properties of the surface S. A
simple reaction mechanism is
5.1.2.2
S + A S A
1
2
S A + e S + B
where S stands for an
site. Assuming that the
(6)
adsorption
adsorption
k2 K1 CA
1 + K1 CA
(7)
k2 = k2 exp
RT
2 F (E E )
exp
(9)
RT
where is the symmetry factor for
the electron-transfer step and E is the
standard potential for the overall reaction.
For a given series of homologous reactions
(in this case, the reaction S A + e B
using different surfaces), the activation
energy can be written as a function of
the Gibbs energy change of the reaction
using the Brnsted relationship [11]:
EA = EA G
(10)
where is the transfer coefcient (generally close to 0.5) and EA is the intrinsic
activation energy for the surface in which
the Gibbs energy of the reaction of the
adsorbed species is zero. Substituting
Eq. (10) into Eq. (9), the following equation
can be obtained:
447
448
EA2
k2 = k2 exp
RT
2 F (E E )
exp
RT
E + 2 G2
= k2 exp A2
RT
2 F (E E )
exp
(11)
RT
The Gibbs energy of the second step (G2 )
of the overall process is related to the
overall Gibbs energy (GT ) and the Gibbs
energy of adsorption (G1 ) according to
(12)
k2 = k2 exp A2
RT
2 F (E E )
exp
(13)
RT
and grouping all the constant terms for the
series of homologous reactions (GT and
log (k#2)
5.1 Electrocatalysis
G 1
Fig. 2
449
500
400
400
300
300
200
200
100
100
0
100
0
100
200
200
300
300
400
400
500
(a)
j
[A cm2]
j
[A cm2]
450
0.0
0.5
1.0
E vs. RHE
[V]
500
1.5
(b)
0.0
0.5
1.0
1.5
E vs. RHE
[V]
5.1 Electrocatalysis
E1
HCOOH
CO + H2O
E2
CO2
E2 > E1
CO2
(17)
The simplicity of this reaction scheme
makes formic acid oxidation a model
reaction in electrocatalysis. For these
reactions, it is important to assess the
catalytic activity of the surface for both
paths (specially the determination of
the reaction path through the active
intermediate) in order to investigate the
catalytic effects of a given surface. The
problem is the reciprocal interference
of the paths that makes difcult the
evaluation of the reaction rates of both
paths separately. The reaction rate for the
active intermediate path can be obtained
only in the absence of surface poison. In
order to achieve that condition, several
strategies have been used. The simplest
strategy is to measure the current density.
After cleaning the surface of the poison
(that normally can be achieved at potentials
above 0.75 V), the electrode potential is
then stepped to a lower value at which
the oxidation of the molecule takes place.
As the electrode surface at time zero (just
after the step) is free from the poison, the
current extrapolated at t = 0 is the reaction
rate through the active intermediate path.
This method has been used for formic acid
and methanol [26, 27]. This strategy has
451
20
i
[mA cm2]
452
0.2
0.5
E RHE
[V]
(a) Pulsed votammogram; and (b) stationary voltammogram for a
),
Pt(100) electrode in 0.5 M H2 SO4 + 0.1 M HCOOH solution (
pulsed voltammogram (-----). Pulse voltammetry conditions: sampling
window time: 0.30 s; poison oxidation pulse length: 0.360 s; potential of
pulse: 0.95 V; time between the end of the poison oxidation pulse and the
opening of the sampling window: 0.015 s. Sweep rate: 10 mV s1 .
(Reproduced with permission from Ref. [28].)
Fig. 4
5.1 Electrocatalysis
453
454
Pt (100)
qCO = 0.16
Pt (110)
qCO = 0.16
qCO = 0.2
qCO = 0.85
qCO = 1.0
4 103 a.u.
qCO = 0.6
Rh (111)
Rh (100)
qCO 0.2
qCO = 0.75
qCO = 0.24
qCO = 0.75
1800
2000
Rh (110)
qCO = 0.25
2000
1800
qCO = 1.0
2000
1800
n
[cm1]
FTIR spectra for CO irreversibly adsorbed on platinum and rhodium low-index
surfaces in 0.1 M HClO4 at 0.25 V versus SCE at saturation and low CO coverages. The
reference spectrum was acquired at 0.450.5 V after CO oxidation. (Reproduced with
permission from Ref. [19].)
Fig. 5
O
C
HN
NH
Pt (100)
O
C
HN
NH 2
Pt (111)
Low coverage
NH 2
HN
C
O
Pt (111)
High coverage
5.1 Electrocatalysis
Pt(111)
Pt(110)
Pt(100)
0.74
0.57
4.8
24.5
6.5
17.0
4.2
5.2
6.1
455
0.2
0.0
0.2
0.4
0.6
0.8
0.6
0.8
i
[mA cm2]
2.0
1.5
1.0
0.5
(a)
0.0
i
[mA cm2]
1.00
0.75
0.50
0.25
(b)
0.00
0.50
0.40
i
[mA cm2]
456
0.30
a3
0.20
0.10
a1
0.00
0.4
(c)
a2
0.2
0.0
0.2
0.4
E (Ag/AgCl)
[V]
Fig. 7
5.1 Electrocatalysis
j
[mA cm2]
5.1.3.2
(100)
(27, 1, 1)
(19, 1, 1)
(11, 1, 1)
(711)
1.0
0.5
(511)
(311)
0
0.5
1.0
E RHE
[V]
Voltammetric proles for the oxidation of 0.01 M glucose in 0.1 M
HClO4 on different Pt(s):[n(100) (111)] stepped surfaces. Scan rate
50 mV s1 . (Reproduced with permission from Ref. [44].)
Fig. 8
457
500
400
300
200
100
0
100
Current density
[A cm2]
Current density
[A cm2]
.2
.6
.4
.6
.2
.8 .9
E (Pd/H ref)
[V]
.6
.8 .9
500
400
300
200
100
0
100
0
(d)
.4
E (Pd/H ref)
[V]
(b)
Current density
[A cm2]
.2
500
400
300
200
100
0
100
.8 .9
500
400
300
200
100
0
100
0
(c)
.4
E (Pd/H ref)
[V]
(a)
Current density
[A cm2]
458
.2
.4
.6
.8 .9
E (Pd/H ref)
[V]
Voltammograms for glucose oxidation on the two chiral Pt(643) electrodes in 0.05 M
H2 SO4 supporting electrolyte. (a) Pt(643)S electrode in 5 mM D-glucose; (b) Pt(643)S electrode
in 5 mM L-glucose; (c) Pt(643)R electrode in 5 mM D-glucose; and (d) Pt(643)R electrode in
5 mM L-glucose. Scan rate: 50 mV s1 . (Reproduced with permission from Ref. [47].)
Fig. 9
5.1 Electrocatalysis
j
[A cm2]
j
[mA cm2]
100
0.5
E (RHE)
[V]
10
0.5
E (RHE)
[V]
Quasi-stationary voltammetric proles of Pt(100) (-----) and
) electrodes in 0.25 M
PdPt(100) (Pd = 0.32) (
HCOOH + 0.5 M H2 SO4 solution. Inset: blank voltammograms of the
same electrodes in 0.5 M H2 SO4 solution. Scan rate 50 mV s1 .
(Reproduced with permission from Ref. [49].)
Fig. 10
459
460
5.1 Electrocatalysis
461
j
[A cm2]
30
20
10
25
75
50
t
[s]
(a)
100
j
[A cm2]
462
50
0
(b)
15
10
20
25
t
[s]
Fig. 11
5.1 Electrocatalysis
j
[mA cm2]
40
Comparison between
experimental () and theoretical
currents versus coverage curves for the
electrooxidation of formic acid at
BiPt(111) electrode in 0.25 M
HCOOH + 0.5 M H2 SO4 . (Reproduced
with permission from Ref. [65].)
Fig. 12
30
20
10
0
0.0
0.2
0.4
0.6
qBi
0.8
1.0
463
464
electrode, the distinction between the 14. A. T. Kuhn, H. Wroblowa, Trans. Faraday.
Soc. 1967, 63, 1458.
effects of both types of ad-atoms is clear.
15. A. Capon, R. Parsons, J. Electroanal. Chem.
For the electronically effective ad-atoms,
1973, 44, 239.
the catalytic activity of the surface increases 16. R. R. Adzic, D. N. Simic, A. R. Despicm
with the ad-atom coverage until it reaches a
et al., J. Electroanal. Chem. 1977, 80, 81.
maximum when 75% of the initial surface 17. A. J. Appleby, Catal. Rev. 1970, 4, 221.
sites have been covered, that is, bismuth- 18a. S. Schuldiner, J. Electrochem. Soc. 1963, 110,
332.
modied Pt(111) surfaces (Fig. 12) [65, 68]. 18b. S. Schuldiner, J. Electrochem. Soc. 1968, 115,
For the ad-atoms acting only as a third
362.
body, the current of the oxidation of formic 19. S.-C. Chang, M. J. Waver, Surf. Sci. 1990,
238, 142.
acid diminishes with the ad-atom coverage,
20. S.-C. Chang, J. D. Roth, Y. Ho et al., J. Elecas observed with the selenium-modied
tron Spectrosc. Relat. Phenom. 1990, 54/55,
Pt(111) surfaces [65].
1185.
References
1. A. J. Appleby in Comprehensive Treatise
of Electrochemistry (Eds.: B. Conway,
J. OM. Bockris, E. Yeager et al.), Plenum
Press, New York, 1983, pp. 173239, Vol. 7.
2. L. I. Kristalik in Advances in Electrochemistry and Electrochemical Engineering (Eds.:
P. Delahay, C. Tobias), John Wiley & Sons,
New York, 1970, pp. 283339, Vol. 7.
3. R. Parsons, Proc. Conf. 75th Anniv. Real
Soc. Esp. de Fis. Y Quim, Madrid, 1981,
pp. 352358.
4. J. OM. Bockris, A. Reddy, Modern Electrochemistry, Plenum Press, New York, 1970.
5. J. OM. Bockris, S. U. M. Khan, Surface Electrochemistry, Plenum Press, New York, 1973.
6. R. Adzic in Modern Aspects of Electrochemistry (Eds.: R. E. White, J. OM. Bockris,
B. E. Conway), Plenum Press, New York,
1990, pp. 163236, Vol. 21.
7. J. Lipkowski, P. N. Ross, (Eds.), Electrocatalysis, Wiley-VCH, New York, 1998.
8. A. Wieckowski, (Ed.), Interfacial Electrochemistry, Marcel Dekker, New York, 1999.
9. M. J. Weaver in Chemical Kinetics (Ed.:
R. G. Compton), Elsevier, Amsterdam, 1987,
pp. 160, Vol. 27.
10. R. A. Marcus, J. Electroanal. Chem. 2000, 483,
2.
11. J. N. Brnsted, K. J. Pedersen, Z. Phys. Chem.
1923, 108, 185.
12. R. I. Masel, Principles of Adsorption and Reaction on Surfaces, John Wiley & Sons, New
York, 1996.
13. S. Trasatti, J. Electroanal. Chem. 1972 39, 163.
5.1 Electrocatalysis
39. S. Motoo, N. Furuya, Ber. Bunsen-Ges. Phys.
Chem. 1987, 91, 457.
40. S. Motoo, N. Furuya, J. Electroanal. Chem.
1985, 184, 303.
41. R. Gomez, J. M. Orts, J. M. Feliu et al., J.
Electroanal. Chem. 1997, 432, 1.
42. K. Popovic, A. Tripkovic, N. Markovic et al.,
J. Electroanal. Chem. 1990, 295, 79.
43. K. Popovic, A. Tripkovic, N. Markovic et al.,
J. Electroanal. Chem. 1991, 313, 181.
44. M. J. Llorca, J. M. Feliu, A. Aldaz et al., J.
Electroanal. Chem. 1991, 316, 175.
45. A. Rodes, M. J. Llorca, J. M. Feliu et al., Anal.
Quim. 1996, 92, 118.
46. C. F. McFaden, P. S. Cremer, A. J. Gellman,
Langmuir 1996, 12, 2483.
47. G. A. Attard, A. Ahmadi, J. M. Feliu et al.,
Langmuir 1999, 15, 1420.
48. G. A. Attard, A. Ahmadi, J. M. Feliu et al., J.
Phys. Chem. B 1999, 103, 1381.
49. M. J. Llorca, J. M. Feliu, A. Aldaz et al., J.
Electroanal. Chem. 1994, 376, 151.
50. M. J. Llorca, E. Herrero, J. M. Feliu et al., J.
Electroanal. Chem. 1994, 373, 217.
51. M. Watanabe, S. Motoo, J. Electroanal. Chem.
1975, 60, 267.
52. K. Franaszczuk, J. Sobkowski, J. Electroanal.
Chem. 1992, 327, 235.
53. T. Iwasita, F. C. Nart, W. Vielstich, Ber.
Bunsen-Ges. Phys. Chem. 1990, 94, 1030.
465
466
5.2
5.2.1
+ 6H2 O + 6e
The anodic oxidation of methanol was considered as one of the most interesting
subjects in electrochemistry during the
last 15 years. Besides very characteristic
reaction pathways and electrocatalytic effects, the methanol molecule with its four
hydrogen atoms is the basis of a highenergy density liquid fuel. The reaction of
methanol with oxygen follows the chemical route
CH3 OH + 1.5O2 CO2 + 2H2 O
(1)
with a reaction enthalpy of 726.6 kJ mol1
and with a Gibbs Free Energy of
702.5 kJ mol1 under standard conditions.
The high-energy density of a combination
of liquid methanol and liquid oxygen was
already used since 1940 by Wernher von
Braun, for the propulsion of the rst rocket
in space. But in the 1960s, the electric energy onboard the space ship was not taken
from a methanol/oxygen fuel cell, but from
a hydrogen/oxygen system. At this time,
the electric data of a methanol/oxygen fuel
cell had not been sufcient for this application. Especially in acid solution, the rate
of methanol oxidation
CH3 OH + H2 O CO2 + 6H+ + 6e
(2)
in combination with oxygen reduction at
the cathode
1.5O2 + 6H+ + 6e 3H2 O
(3)
(4)
(6)
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
Basic Facts
G = H T S
Thermodynamics, Open-circuit
Potentials
From thermodynamics, one obtains the
electrode potential on the basis of a
reversible process at the surface of the
catalyst. Our top example is the hydrogen
electrode with its high exchange current at
platinum in acid solution:
(8)
(9)
G
nF
(10)
5.2.2.1
Had + H2 O H3 O+ + e
G
[%]
H
(11)
467
468
Fuel
Reaction
H0
[kJ mol1 ]
G0
[kJ mol1 ]
0 [V]
Hydrogen
H2 + 12 O2 H2 Ol
286.0
237.3
0.000
83.0
CO
CO + 12 O2 CO2
283.1
257.2
0.163
90.9
Formic Acid
HCOOH + 12 O2 CO2 + H2 Ol
270.3
285.5
0.251
105.6
726.6
702.5
0.015
96.7
Methanol
CH3 OH +
3
2 O2
CO2 + 2H2 Ol
Reaction Pathways
In the introduction (Sect. 5.2.1), we
learned already that in alkaline solution
CO shows two possible reaction pathways.
In acid solution this is not the case. But we
have now to discuss the different pathways
for formic acid and methanol.
As already proposed by Capon and
Parsons [9], the anodic reaction of formic
acid at platinum proceeds indeed via
a dual-path mechanism. One pathway
is producing CO2 in a direct way via
dehydrogenation
5.2.2.2
(13)
(14)
and
COad + OHad CO2 + H+ + e
(15)
The direct pathway possibly involves a
reactive intermediate like COOH [9]. This
step is still unknown, but the formation
of hydrogen atoms has been shown using
a palladium membrane as sensor [10, 11].
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
CHOHad COHad + H+ + e
or
CHOHad COad + 2H+ + 2e
(18)
or
H2 COad HCOad + H+ + e (19)
The presence of adsorbed CO was
shown by in situ Fourier Transform
Infrared Spectroscopy (FTIRS) [14] and
the presence of adsorbed COH by
FTIRS [15], the relative amounts depending on methanol concentration. COH is
favored by low methanol concentrations
as proved via Thermal Desorption Spectroscopy (TDS) [16a]. For the complete
oxidation to CO2 , again OHad is required
as in the above case of formic acid.
The existence of formic acid formation
on a second path has been proved by the
DEMS-technique (Sect. 5.2.3). A possible
step of formation could be
CHOHad + OHad HCOOH
+ H+ + e or
CHOHad + H2 O HCOOH
+ 2H+ + 2e
(20)
469
Methanol
O
C
H
Pt-elektrode
CO
P
STO
H+
e e
CO2
H+
H2 O
O C
H+
Current density
[A cm2]
30
III
IV
II
15
0
CO2
m /e = 44
III
II
IV
5.1012A
Mass signal
470
0.0
0.2
0.4
0.6
0.8
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
40
Pt(111)
20
Current density
[A cm2]
Pt(100)
150
100
50
0.2
0.6
1.0
471
j
[A cm2]
(a)
0.8
0.4
0.0
i ion 1012
[A cm2]
m /e = 44
CO2
(b)
m /e = 60
HCOOCH3
i ion 1014
[A cm2]
472
0
0.0
(c)
0.5
1.0
1.5
E vs. RHE
[V]
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
0.3
Current
[mA]
0.2
0.1
0.0
(a)
3.0
m /e = 44
MI 1011
[A]
2.5
2.0
1.5
1.0
0.0
0.1
0.2
0.3
0.4
0.5
0.6
(b)
5.2.3.2
5.2.3.2.1 Carbon Monoxide Carbon monoxide adsorbed at the interface electrode/electrolyte shows a strong surfacesensitive behavior [34]. We will show
this for Pt(111) as a model substrate.
Depending on the potential, CO can
be adsorbed in three different forms:
linearly bonded (atop), bridge bonded,
and threefold bonded. With the shift
in the strength of the CO bond, the
frequency (or alternatively, the wave number) are varying in a clear manner.
473
Current
[A]
60
40
8
6
4
2
0
2
20
0.1
0.2
0.3
0.4
0
50
Pt(110)
40
Current
[A]
30
2
20
1
0
10
1
2
0.1
0.2
0.3
0.4
30
Current
[A]
474
Pt(111)
20
10
0
0.0
0.2
0.4
0.6
0.8
1.0
1.2
Potential
[V]
Fig. 6 First potential scan after methanol adsorption at 0.05 V
as in Fig. 5, but single-crystal surfaces Pt(100), Pt(110), and
Pt(111), 50 mV s1 [26, 27].
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
2.0
I / 103
[A]
1.5
1.0
0.5
0
0.5
0.5
1.0
1.5
E rhe
[V]
Ii /
[A]
1011
m /e = 44
3
0.5
1.0
1.5
E rhe
[V]
Fig. 7 Formic acid oxidation at porous Pt, CV, and MSCV as in
Fig. 4, but 12.5 mV s1 and 0.01 M HCOOH solution [30, 31].
475
10.0
I /104
[A]
7.5
5.0
2.5
E ad
0.0
0.5
1.0
m /e = 44
Ii
[A]
/1011
1.5
E rhe
[V]
2.5
E ad
m /e = 45
0.5
Ii
1.0
1.5
E rhe
[V]
m /e = 44
[A]
/1011
476
0.5
1.0
1.5
E rhe
[V]
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
.01
2073
1850
.005
(b)
Absorbance
[a.u.]
2066
1773
(a)
.005
.01
2400
2200
2000
1800
Wavenumbers
[cm1]
Fig. 9 Potential-difference infrared spectra in a CO-saturated 0.1 M
HClO4 solution at (a) 0.01 V and (b) 0.34 V versus RHE, reference was
made by stepping to 0.75 V RHE [35].
( 19 19) structure.
Not only does the integrated band
intensity change with the potential, but
the frequency (or wave number) also shifts.
Figure 11 shows the potential dependence
of these two band parameters for atop
CO on polycrystalline ruthenium. In this
case, we follow the stripping of a saturated
adlayer formed at 0.3 V using a pure 0.1 M
477
0.20
478
0.15
Diffuse
(2 2)
0.10
( 19 19)
0.05
0.00
0.3
0.2
0.1
0.0
0.1
0.2
0.3
Potential vs SCE
[V]
Fig. 10 Plot of integrated infrared band intensities for linearly, bridge- and
threefold-bonded CO adsorbed on Pt(111), together with spectra for CO2 ,
obtained along with stepping to the 0.75 V versus RHE reference; all data were
taken with CO-saturated 0.1 M HClO4 [35].
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
CO
0.8
0.4
CO2
0.0
Wavenumber
[cm1]
2025
2010
1995
0.0
0.2
0.4
0.6
0.8
E vs RHE
[V]
6
Pt(111)
CO2
Band intensity
[a.u.]
4
CO ( 3)
2
= CO ( 3)
0
0.0
0.2
0.4
0.6
0.8
1.0
Potential
[V]
Fig. 12 Integrated IR band intensities of COL (atop), COB (bridge-bonded), and
CO2 during the rst anodic scan at Pt(111) in 0.1 M HCOOH/0.1 M HClO4 as
function of potential [38].
479
1.6
1.4
Current
[mA]
480
1.2
1.0
0.8
0.6
0.30
0.35
0.40
0.45
0.50
Potential vs RHE
[V]
Fig. 13 CV for a Pt(111) electrode in 0.1 M HCOOH/0.1 M
HClO4 [44, 45]. The potential is reversed at 0.5 V, sweep rate
50 mV s1 .
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
1640
Bandcenter frequency
[cm1]
1630
1620
1610
1600
dHOH
1590
0.0
0.2
0.4
0.6
0.8
1.0
1.2
Potential
[V]
Fig. 14 Potential dependence of the wave number for the in-plane
deformation (scissor mode) of water adsorbed at Pt(111) [44, 45].
481
0.25 V
0.30 V
0.35 V
H2O
0.40 V
0.50 V
1710
1826
1230
0.55 V
2059
482
0.60 V
HCOOCH3
COL
COB
R /R 0 = 0.01
CO2
2400
2000
1600
1200
Wavenumber
[cm1]
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
3.2
0.16
Pt(111)
0.12
2.4
COL
0.08
1.6
COB
CO2
Band intensity
[a.u.]
0.04
0.8
0.00
3.2
Pt(111)/Ru(45%)
2.4
0.08
CO2
1.6
0.04
0.8
COL
0.00
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.0
0.8
Potential vs RHE
[V]
Integrated IR band intensities of COL (atop), COB
(bridge-bonded), and CO2 during methanol oxidation at Pt(111) and at
Ru(45%)/Pt(111) in 0.5 M CH3 OH/0.1 M HClO4 as function of
potential [53].
Fig. 16
483
484
t
[min]
5
10
20
30
40
R /R
2100
1900
1300
0.1%
1100
Wavenumber
[cm1]
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
ruthenium (in the form of islands, clusters, or atoms) for the formation of a
Ru-OH species, being used nally for a
successive oxidation of methanol adsorbates (Fig. 1). It has to be noted that at
room temperature ruthenium does not
adsorb methanol [64], the rst step of adsorption takes place in the presence of Pt
sites only.
In the following, we discuss the PtRu
system as a model catalyst. Binary PtRu
electrocatalysts are presently studied in
many different forms, PtRu alloys [6568],
Ru electrodeposits on Pt [69, 70], PtRu
100
50
50
0
0
50
50
Current density
[A cm2]
(a) 0% Ru
(d) 63% Ru
100
50
50
50
50
(b) 11% Ru
(e) 13% Ru (H2red.)
50
50
50
50
(c) 39% Ru
0.0
0.2
0.4
0.6
0.8
0.2
0.4
0.6
0.8
1.0
Potential vs RHE
[V]
Fig. 18 CVs at 50 mV s1 for Pt(111), different Ru-modied Pt(111) surfaces and a
PtRu alloy in 0.1 M HClO4 , 25 C; (a) pt(111), (bd) using spontaneous Ru
adsorption, (e) Ru adsorption and simultaneous reduction by hydrogen [75].
485
30
0.50
15
0.25
0
15
0.00
Pt
Pt
50
25
0
25
PtRu(50 : 50)
Current density
[mA cm2]
Current density
[A cm2]
486
1.2
0.8
0.4
PtRu(50 : 50)
0.0
30
0.6
0
0.3
30
Ru
60
0.0
0.2
0.4
0.6
0.8
Potential vs RHE
[V]
Ru
0.0
1.0
Potential vs RHE
[V]
Fig. 19 CVs for oxidative stripping of CO (left side) and oxidation from
CO-saturated solution (right side), 0.1 M HClO4 ; for stripping saturation of
surface with CO at 0.3 V; dashed: second scan coinciding the CV of a clean
surface in the supporting electrolyte, 50 mV s1 [36].
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
120+/12 A cm2
100
Current density
[A cm2]
40
31+/2.4 A cm2
10
20
40
Ru%
60
80
100
487
488
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
Fig. 21 Comparison of CO bands in
FTIR spectra for methanol on different
samples of Fig. 20: Pt(111), Pt(111) with
spontaneously adsorbed Ru(39%), with
Ru reduced by hydrogen and PtRu alloy,
15% of Ru each [75].
Pt(111)
1834
Pt(111)/Ru(ads)
2057
Pt(111)/Ru(H2)-red.
2060
1965
PtRu alloy
R /R0 = 2 103
2048
2250
2000
1750
Wavenumber
[cm1]
489
5.2.3.3.3
12
j
[mA cm2]
490
0
0.0
0.5
1.0
1.5
2.0
E (RHE)
[V]
Fig. 22 Cyclic Voltammograms for formic acid oxidation at smooth
polycrystalline platinum, without addition (dotted lines) and with lead ions
in the electrolyte (full lines); 1 M HCOOH/1 M H2 SO4 , 105 M Pb2+ ,
50 mV s1 [94].
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
30
1.0
I/Imax
Ip
[mA cm2]
0.8
20
0.6
0.4
10
0.2
0.0
0.0
Sb
(a)
0.2
0.4
0.6
0.8
1.0
0.8
1.0
Adatom
(b)
1.0
40
30
I/Imax
I
[mA cm2]
0.8
20
0.4
10
0.2
0.0
0
0.5
(c)
0.6
1.0
Bi
0.0
(d)
0.2
0.4
0.6
Adatom
distributed ad-atoms may block neighboring places, making some isolated sites
available for the direct oxidation. In addition, the rate of reaction per free Pt site
is enhanced. In Fig. 23, we have in (a, b)
a comparison between the experimental
current/Sb plot and a simulation under the conditions of randomly distributed
ad-atoms and a reaction rate proportional
491
492
From the discussions in the above chapters, we learned already that surface
structures are of special importance for
methanol oxidation as a result of the required bifunctional catalyst properties. A
rst step to use STM images to gain a
better insight into this matter was done
by Cramm and coworkers [83]. Electrodeposits of Ru islands on Pt(111) in 0.1 M
HClO4 were characterized by STM images. Islands of monoatomic height and
between 2 to 5 nm of diameter are shown
for two different Ru coverages. More recent
results will be presented in the following paragraph.
Ex Situ STM Images Taken during
Methanol Oxidation via UHV Transfer
In order to follow the morphology of a
Ru/Pt(111) surface during methanol oxidation, the following modication of an
Omicron standard preparation and analysis chamber was made [102]. The manipulator was extended by a platinum shield
to avoid foreign metal deposition. The external load-lock chamber was modied to
5.2.4.1
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
50 nm
50 nm
(b)
(a)
20 nm
20 nm
(c)
(d)
Morphology of a Ru/Pt(111) electrode taken via STM before and after 50 min
methanol oxidation at 500 mV in 0.5 M CH3 OH/0.1 M HClO4 . Potential step from 50 to
500 mV RHE [103].
Fig. 24
493
1000
Current density
[A cm2]
F
100
E
D
C
10
Fig. 25
B
A
0
12
18
Time
[min]
40
30
20
Pt(111)
10
0
300
Current density
[A cm2]
494
Pt : Ru
50 : 50
200
100
0
400
Pt : Ru
75 : 25
200
0
400
Pt : Ru
85 : 15
200
0
10
Time
[min]
15
20
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
0.020
Current
[A]
0.015
0.010
0.005
0.000
200
400
600
800
1000
1200
1400
Time
[s]
Fig. 27 Current/time curve for a porous PtRu electrodeposited (81 : 19) on
a gold substrate. The steps at ca. 4 min and 14 min are caused by the
production of CO2 bubbles [104].
495
Potential vs RHE
[V]
0.25
0.00
250
(b) Current response
Current density
[A cm2]
496
200
150
100
10
20
30
40
50
60
70
80
90 100 110
Time
[min]
Fig. 28 Current/time curves for a smooth PtRu(85 : 15) alloy as in
Fig. 26, but with load interruptions by turning back the potential to its
initial value [104].
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
(21)
dN (t) = kN (t)
(23)
Path one
(1) CH3 OH
CO + 4H+ + 4e
(24)
+
(27)
(28)
(29)
Path three
(7) CH3 OH
HCHO + 2H+ + 2e
(30)
(8) HCHO + H2 O
CO2 + 4H+ + 4e
(31)
Path four
(9) CH3 OH
COH + 3H+ + 3e
(32)
(33)
497
methanol concentration for ve different porous catalysts and with ve different methanol concentrations between
0.1 and 3.0 M. Interestingly, for an ETEK PtRu(50 : 50) catalyst, again a strong
inuence of methanol concentration was
found. For PtRu(84 : 16) and PtRu(75 : 25),
a higher activity was observed, being
almost constant between 0.5 and 3 M
(Fig. 29). The last data are in good agreement with the plots in Fig. 20. Obviously,
in the high activity PtRu region and for
25 C, the rate-determining reaction is the
oxidation of adsorbed CO (Steps 2 and 3
in our scheme for the mechanism) while
outside this region (e.g. for the catalysts
PtRu(50 : 50)) the available Pt sites are not
enough for a sufcient rate of adsorption
of methanol. For Pt(90 : 10) and of course
for pure platinum, the number of Pt sites
is high, but the bifunctional mechanism is
hindered or even not possible at all.
For a methanol concentration smaller
than ca. 0.3 M, we obviously have in
all cases the effect of a to low rate of
methanol adsorption. In the region of
0.012
Pt75Ru25/C
0.010
Pt84Ru16/C
inorm.
[A cm2]
498
0.008
Pt50Ru50/C-(E-TEK)
0.006
Pt92Ru10/C
0.004
Pt/C-(E-TEK)
0.002
0.5 V vs. RHE
0.000
0.0
0.5
1.0
1.5
2.0
2.5
CH3OH conc.
[mol L1]
Fig. 29 Current density, taken 30 min after applying the load, as a
function of methanol concentration for different porous catalysts at
25 C [114].
3.0
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
N-shaped currentpotential
curves i() ( potential of the working
electrode) and load lines (U )/RE ; U
applied voltage, RE ohmic resistance:
(a) single stationary case; (b) three
stationary states (bistability); (c) plot of
current i against the external voltage U.
Fig. 30
(a)
(U j)/RE
Potential j
(U j)/RE
i (j)
i (j)
5.2.5
(b)
Potential j
i (j)
Voltage U
(c)
499
500
iC
W
CD
RE
ZF
iF
ZF
The current through the interface electrode/electrolyte has two pathways, iF and
iC . For the respective differential equation follows:
i=
(U )
= iC + iF
RE
= ACD
d
+ iF ()
dt
(34)
(35)
d
diF
+
and
dt
d
(U )
=
, or
RE
RE
= ACD
+ ZF 1 , or
RE
dt
d()
= (ACD )1 (ZF 1 + RE 1 ), or
dt
= (t = 0) exp{(ACD )1
(ZF 1 + RE 1 )t}
(37)
(38)
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
(39)
(40)
501
502
dCO
= k2 (1 CO H2 O )
dt
k4 (CO H2 O )
(42)
dH2 O
= k3 CO (1 CO H2 O )
dt
k3 H2 O k4 (CO H2 O )
(43)
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
0.8
0.4
0.0
10
12
14
70
72
74
76
78
80
Potential
[V]
0.8
0.4
0.0
0.8
0.4
148
150
152
154
156
Time
[min]
Potential
[V]
1.0
0.5
0.0
0
10
15
20
Time
[sec]
Fig. 32 Potential oscillations during formic acid oxidation at a smooth polycrystalline
platinum electrode: (a) 5 M HCOOH/0.5 M H2 SO4 , applied current density of
1.5 mA cm2 ; (b) 1 M HCOOH and 0.5 M H2 SO4 , 0.25 mA cm2 [127].
1
dE
=
[I 2Fh{k1 (1 CO H2 O )
dt
Cd
+ k4 (CO H2 O )}]
(44)
503
504
Im Z 8
[]
6
4
2
5 Hz
8
Re Z
[]
10 kHz
4
4
6
0.1 Hz
4
6
8
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
Reduction of Pt oxide, formed
in 2.5 M H2 SO4 , by dipping the
platinized Pt electrode in 0.1 M
fuel/2.5 M H2 SO4 solution [127, 135].
Fig. 34
1.2
1.0
1.4
0.8
0.6
0.4
CH3OH
0.2
HCOOH
CH2O
0.0
[min]
505
0.20
Current density
[A cm2]
c
0.16
0.12
0.08
50
100
150
200
250
300
Time
[s]
Fig. 35 Current Oscillations during methanol oxidation at a Pt catalyzed gas
diffusion electrode (0.4 mgPt cm2 ), supplied with nitrogen to the gas side;
2.0 M CH3 OH/0.5 M H2 SO4 ; controlled potentials versus RHE (a) 1.0; (b) 1.2;
and (c) 1.5 Volt [136].
0.15
0.4
0.12
I
[mA cm2]
506
0.3
0.00
0.2
3000
3500
4000
0.1
0.0
0
1000
2000
3000
4000
t
[s]
Fig. 36 Current/time transient of a rotating platinum electrode (1.000 rpm) in
CO-saturated 0.1 M HClO4 at 0.95 V versus RHE. Initial instants after potential
application, the insert shows a time window [140].
5000
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
the surface (e.g. via convection) is of importance. At a rotating electrode, one has
above 1.0 V versus RHE limiting current,
being constant also during the negative
scan and passing the sharp maximum of
the anodic scan (Fig. 4 in Ref. [140]).
5.2.6
Current density
[A cm2]
0.1
CH3OH
b, c
0.0
O2 + CH3OH
d
0.1
O2
0.2
0.0
0.2
0.4
0.6
0.8
1.0
1.2
Potential vs RHE
[V]
Fig. 37 Stationary current/potential curves of a Pt catalyzed gas diffusion
electrode as in Fig. 33. Points taken after 3 to 5 min, room temperature;
(a) 0.5 M H2 SO4 , oxygen supplied to the gas chamber only; (b) 2.0 M
CH3 OH/0.5 M H2 SO4 , nitrogen supplied to gas chamber and solution;
(c) methanol solution and nitrogen supplied to the gas chamber as in (b),
but oxygen supplied to the solution; (d) methanol in solution and oxygen
supplied to the gas chamber [136].
507
508
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
14. A. Bewick, K. Kunimatsu, S. Pons, J. W.
Russel, J. Electroanal. Chem. 1984, 160, 47.
15. T. Iwasita in Advances in Electrochemical Science and Engineering (Eds.: H. Gerischer,
C. W. Tobias), Wiley-VCH, Weinheim, Germany, 1995, pp. 123216, Vol. 4.
16a. S. Wilhelm, T. Iwasita, W. Vielstich, J. Electroanal. Chem. 1987, 238, 383391.
16b. K. I. Ota, Y. Nakagawa, M. Takahashi, J.
Electroanal. Chem. 1964, 179, 179.
17. M. Watanabe, S. Motoo, J. Electroanal.
Chem. 1975, 60, 267273.
18. B. Bittins, E. Cattaneo, P. Konigshoven
et al. in Electroanalytical Chemistry (Ed.: A. J.
Bard), Marcel Dekker, New York, 1991,
pp. 182219, Vol. 17.
19. H. Baltruschat in Interfacial Electrochemistry
(Ed.: A. Wieckowski), Marcel Dekker, New
York, 1999, pp. 577597.
20. P. A. Christensen, A. Hamnett, Techniques
and Mechanisms in Electrochemistry, Blackie
Academic, Chapman & Hall, London, 1994,
pp. 217224, 278283.
21. O. Wolter, C. Giordano, J. Heitbaum et al.,
Proc. Symp. Electrocatalysis, The Electrochemical Society, Pennington, N.J., 1982,
pp. 235253.
22. P. Stonehart, Electrochim. Acta 1973, 18, 63.
23. E. Santos, E. Leiva, W. Vielstich et al., J.
Electroanal. Chem. 1987, 227, 199211.
24. L. H. Leung, A. Wieckowski, M. J. Weaver,
J. Phys. Chem. 1988, 92, 69856990.
25. T. Iwasita, W. Vielstich, J. Electroanal.
Chem. 1986, 201, 403408.
26. X. H. Xia, T. Iwasita, F. Ge et al., Electrochim. Acta 1996, 41, 711718.
27. T. Iwasita in Handbook of Fuel Cells (Eds.:
W. Vielstich, A. Lamm, H. Gasteiger), John
Wiley & Sons, Chichester, UK; in press.
28. J. Clavilier, C. Lamy, J. M. Leger, J. Electroanal. Chem. 1981, 125, 249.
29. S.-Ch. Chang, L. H. Leung, M. Weaver, J.
Phys. Chem. 1990, 94, 60136021.
30. J. Willsau, Dissertation Uni Bonn, 1985,
pp. 8084.
31. J. Willsau, J. Heitbaum, Electrochim. Acta
1986, 31, 943948.
32. J. Clavilier, R. Parsons, R. Durand et al., J.
Electroanal. Chem. 1981, 124, 321.
33. R. R. Adzic, A. V. Tripkovic, W. E. OGrady,
Nature 1982, 296, 137.
34. R. R. Adzic in Handbook of Fuel Cells (Eds.:
W. Vielstich, A. Lamm, H. Gasteiger), John
Wiley & Sons, Chichester, UK; in press.
509
510
5.2 CO, Formic Acid, and Methanol Oxidation in Acid Electrolytes Mechanisms and Electrocatalysis
107. D. Chu, S. Gilman, J. Electrochem. Soc. 1996,
143, 1685.
108. A. Aramata, M. Masuda, J. Electrochem. Soc.
1991, 138, 19491957.
109. M. Watanabe, M. Uchida, S. Motoo, ECS
Meeting, Boston, Mass., Ext. Abstr. 339,
1986.
110. I. A. Rodrigues, J. P. I. Souza, F. C. Nart,
Langmuir 1997, 13, 6829.
111. R. Liu, H. Iddir, Q. Fan et al., J. Phys. Chem.
B 2000, 104, 35183531.
112. H. Kim, I. Rabelo de Moraes, G. TremiliosiFilho et al., Surf. Sci. 2001, 474, L203L212
113. T. J. Schmidt, H. Gasteiger, R. J. Behm,
Electrochem. Commun. 1999, 1, 14.
114. V. A. Paganin, W. Lizcano-Valbuena, E. R.
Gonzales, Electrochim. Acta 2000, 47,
37153722.
115. K. F. Bonhoeffer, G. Vollheim, Z. Naturforsch. 1953, 8b, 406.
116. U. F. Franck, R. FitzHugh, Z. Elektrochem.
1961, 65, 156.
117. M. T. M. Koper, Electrochim. Acta 1992, 37,
17711778.
118. T. Yamazaki, T. Kodera, Electrochim. Acta
1991, 36, 639646.
119. M. T. M. Koper, J. H. Sluyters, J. Electroanal. Chem. 1991, 303, 73.
120. M. T. M. Koper, J. H. Sluyters, J. Electroanal. Chem. 1994, 371, 149.
121. K. Krischer in Modern Aspects of Electrochemistry (Eds.: J. OM. Bockris, B. E. Conway,
R. White), Kluwer Academic Publishers,
New York, 1999, pp. 1142, Vol. 32.
122. K. Krischer, N. Mazouz, P. Grauel, Angew.
Chem. 2001, 113, 842863.
123. K. Krischer, N. Mazouz, P. Grauel, Angew.
Chem., Int. Ed. Engl. 2001, 40, 850861.
124. K. Krischer, M. Lubke, W. Wolf et al., Electrochim. Acta 1995, 40, 69.
125. M. Wolf, K. Krischer, M. Lubke et al., J.
Electroanal. Chem. 1995, 385, 85.
511
512
5.3
n1
m1
n+1
m+1
(1)
Thus, consecutive addition (and detachment) of atoms leads eventually to the
formation of supercritical nuclei that grow
irreversibly and are not further considered
in the steady state equations. Becker and
513
514
occurs at very early stages, the above expression soon reduces to N = N0 , and
nucleation is instantaneous. Conversely,
for very small A, the number density of
nuclei increases initially linearly with time,
N = N0 At, and nucleation is said to be
progressive.
In multiple nucleation conditions, the
current density may be expressed as a
convolution of the growth current of individual nuclei and their birth rate on the
surface. In the very initial stages, nuclei
are small in both size and number, and
they grow independently of each other.
The current transient is then given
by
3/2 (Mt/)1/2 N %( At)
i(t) = zF (2Dc)
0
At 2
At
e d.
%( At) = 1 (e / At) 0
For, At 20, %( At) 1 and the equation above describes a current density
proportional to N0 t 1/2 , corresponding to
instantaneous nucleation; for
At 0.2
nucleation is progressive, %( At)
(2/3)At and the current density is proportional to (2/3)N0 At 3/2 [10, 11].
An overall description of the nucleation
and growth process valid for all times
is in general complicated by interactions
among growing centers. When growth
is diffusion controlled, then overlap of
the spherical diffusional elds supporting
the growth of individual nuclei must be
considered, eventually collapsing into a
single planar eld to the plane of the electrode. The many-body problem that ensues
has been approached following several
strategies. General models not restricted
to the instantaneous case, and allowing
determination of A and N0 from current transients, consider two-dimensional
projections of the hemispherical diffusional elds around individual nuclei onto
the plane of the electrode [12] and estimate their overlap using the Avrami
zF Dc
(Dt)1/2
(At)1/2
eAt
2
1
e d
(At)1/2 0
1 eAt
1
At
cM 1/2
3/2
1 exp (2) D
1 eAt
(2)
N0 t 1
At
515
516
(3)
d(r 2 )
=0
dnL
(4)
2 L
=0
r
(5)
reactions such as
Mz+
(solution) + ze(electrode)
M(new phase)
(8)
for metal ion reduction to the corresponding metallic phase, or
M(solid) + aXz
(solution)
MXa(new phase)
+ aze
(electrode)
(9)
for anodic deposit formation.
According to ErdeyGruz and Volmer [15], the supersaturation c/c , where
c is the concentration of ions in solution
in equilibrium with a surface of radius of
curvature r, and c is the corresponding
concentration of ions in equilibrium with
a planar surface, is directly determined by
the overpotential, = E Erev , where E
is the electrode potential and Erev is the
equilibrium potential, in the absence of net
current ow for a single electrode reaction.
For the metal electrode|solution interface,
RT
c
=
(10)
ln
zF
c
where F is Faradays constant. This
relation expresses that through external
control of the electrical potential, the
conditions governing the electrochemical
nucleation process can be precisely and
reversibly controlled.
The Classical Nucleation Model
From Eq. (3), the Gibbs energy of formation of a nucleus may be formally described
as [3]
5.3.3.1
G = Gvolume + Gsurface
(11)
517
518
where
Gvolume
4
r 3
GV
=
3
(12)
where
GV is the Gibbs energy of
formation of the bulk phase per unit
volume, and
Gsurface = 4r 2
(13)
G
= 4(r )2
GV
r r
+ 8(r ) = 0
(14)
GV
(15)
Ghomo =
3
(
GV )2
=
(16)
4(r )2
3
which has been designated here as
Ghomo (in homogeneous phase) to emphasize that Eq. (16) refers to the work
of formation of a nucleus from the
bulk phase.
A comparison of Eqs. (15 and 7) reveals
that both the radius of the critical nucleus
and the work needed to form it are
functions of the ratio c/c . When c = c ,
that is, when the solution is saturated
in relation to an interface of innite
radius of curvature, the work required to
form a nucleus tends to innite, and the
phase transition does not occur. As the
supersaturation ratio c/c increases, the
radius of the critical nucleus, and hence its
reversible work of formation, diminishes.
Supersaturation is one of the principal
factors determining the work required to
form a critical nucleus, but there are
other factors worth considering too. In
the majority of cases and especially under
electrochemical conditions, formation of
a new phase occurs on a surface. The
(17)
Fig. 1
(19)
G =
GV V + 1,2 S1,2
+ 2,3 S2,3 + 1,3 S1,3
(20)
s1,3
(18)
s2, 3
519
520
Ghetero =
Ghomo ()
16(1,3 )3
=
()
3(
GV )2
(21)
GV =
zF
M
(23)
Ghetero =
16 3 M 2 ()
3(zF )2
(24)
2 M
zF ||
(25)
1018 G*
[J]
2.0
60 mV
1.0
90 mV
0.0
1.0
0
10
15
20
25
108
[r cm1]
Fig. 2 Free energy of formation of hemispherical mercury clusters from
Hg2 2+ aqueous solution at different overpotentials, as a function of
nucleus size. = 300 dyn cm1 , obtained from electrocapillary curves of
= 0.8 V.
Hg in KNO3 solution at EHg(II)/Hg
n1
n1
n
n+1
n
(26)
n+1
The change in concentration of clusters of n molecules may be written as dCn(t)/dt = n1 Cn1 (t) (n +
n )Cn (t) + n+1 Cn+1 (t), which has the
form of Kolmogorov differential equation
for Markov processes in discrete number
space and continuous time [21]. n and
n are respectively the net probabilities
of incorporation or loss of molecules by
a cluster per unit time, and these may
be dened formally as the aggregation or
detachment frequencies times the surface
area of the cluster of n molecules. Given
the small size of the clusters, n and n
are not simple functions of n and in general they are unknown. However, if n
and n are not functions of time, then
an equilibrium distribution Cn of cluster sizes exists, such that dCn /dt = 0 for
Cn (t) = Cn , and the following differential
521
522
Cn (t)
Cn (t)
=
n Cn
(27)
t
n
n
Cn
This expression holds if the time during
which the macroscopic phase transition
occurs is large compared to the characteristic times of the accretion and decay
processes, n1 and n1 .
Equation (27) is the FokkerPlanck
equation and describes diffusion in the
presence of an external force, n plays a
role similar to that of the diffusion coefcient in the Fick equation; for the special
case of Cn = C = constant, substituting
n for D, we obtain Ficks law,
2 Cn (t)
Cn (t)
=D
t
n2
(28)
Cn = C1 exp
Gn
kT
(29)
where
Gn is the standard Gibbs energy
of formation of a n-mer from n monomers
and C1 is the concentration of monomers
in the initial phase. Referring to Eq. (16)
describing the Gibbs energy of a critical
cluster (cf. Fig. 2), it is clear that Cn is
minimal at n = n (n is the number
of molecules constituting the critical
nucleus) or r = r and that
G =
Cn (t)
Cn (t)
=
n
t
n
n
+
n Cn (t) (
Gn )
kT
n
(30)
(31)
Js = n
n n
which is clearly a diffusion equation
that may be solved approximately with
the limiting layer model. Thus replacing
(Cn /n)n by Cn /
n, where
n is the
width of the critical region, (2kT / )1/2
and taking account that diffusion occurs at
both sides of the critical region, then [22]
Js =
2n Cn
(2kT / )1/2
(32)
0.655
kT
1018 G q
[J]
0.650
0.645
0.640
0.635
0.630
0.625
30
35
40
45
50
55
60
n
[atoms]
Fig. 3 Critical region around n within which the energies of hemispherical
mercury clusters at = 90 mV differ from
G less than kT.
523
524
n
Ra2
n
a
[(n + 1)Qad + En QD ]
exp
kT
(35)
where R is the incidence rate of atoms
to the surface, n is the capture cross
section of critical nuclei, a is the distance
between adsorption sites, is an attempt
frequency, Qad is the bond energy with the
surface, and QD is the activation energy for
surface diffusion. Since n can only adopt
integer values, the same size of critical
nucleus should operate over a range of
temperatures, and discontinuities in the
temperature-dependence of the nucleation
rate are to be expected upon changes
of the critical size along temperature
intervals. Regarding the clustering process
as an innite chain of intermediates
with relative concentrations determined
by their respective rates of growth and
decay [31] avoids altogether having to
consider the notion of critical nucleus but,
unfortunately, does not lead to analytical
expressions for the nucleation rate without
a number of additional assumptions,
although it is possible to obtain explicit
expressions for certain limiting cases [32].
Stoyanov [6] reexamined Waltons atomistic theory of heterogeneous nucleation
and derived the nucleation rate avoiding
J =R
Js =
1+
m
i=2
2 3 i
+2 +3 +i
(36)
QD
Ci
exp
N0
kT
and
(Ei Ei1 + QD )
i = i exp
kT
(37)
where i is the number of ways of forming
a cluster of size i + 1 incorporating one
ad-atom to a cluster of size i and i is the
number of ways an atom can detach from
a cluster of size i. Given that for each birth,
i i + 1, there is a corresponding death,
i + 1 i, i and i are related by
i = i+1
(38)
525
526
obtained as
i
m
Js = +1 C1
(j N0
1+
i=2 j =2
exp(Ej /kT )
(39)
+n 1 = k+n 1 c
Un 1 (1 e )ze0 (E0 + )
exp
kT
(41)
and
n =
k
Un + e ze0 (E0 + )
exp
kT
(42)
where Un 1 and Un are the energies
of transfer of electrodepositing ions from
solution to clusters and vice versa, c is
the concentration of electrodepositing ions
in solution, e is the electronic transfer
coefcient, e0 is the charge of the electron,
and E0 is the equilibrium potential of the
N0 N1
0
exp
(43)
N1 =
N0
kT
so that
kn 1
n
n
=
exp
+n 1
k+n 1 c
kT
ze0 (E0 + )
exp
(44)
kT
and the nucleation rate is expressed as
Un (1 e )ze0 E0
Js = ZcN0 exp
kT
%(n )
exp
kT
(n + 1 e )ze0
exp
(45)
kT
where Z contains the ratio of products of
frequencies of attachment and detachment
from clusters up to the critical and %(n) =
n1/2 ;i represents the difference in
energy of n atoms when they are part of an
innitely large crystal on the one hand, and
when they form an independent cluster on
the electrode surface on the other, that is,
%(n) represents the surface energy of the
cluster of n atoms.
5.3.4
Metal electrodeposition occurs at the interface between an electronically conducting substrate and an ionically conducting
(46)
zF dV
M dt
(47)
zF
dr
2r02
M
dt
(48)
and accordingly,
2zF Dcr0 =
(50)
527
528
(51)
(52)
zF (2Dc)3/2 M 1/2
1/2
t
dN
(t u)1/2
du
dt
0
(53)
where
eAt
%=1
(At)1/2
(At)1/2
e d
(55)
3/2
= (2)
cM
1/2
(t u) (57)
However, nuclei do not grow independently of each other and their interactions
results in overlap of the diffusion elds
around them. Overlap may be accounted
for using Avramis theorem, [13]
d = 1 exp(ex )
(58)
(t u) exp(Au) du
(59)
ex = 2
(56)
(a)
rd2
2MDc
1/2
= At>
(b)
(60)
(c)
Fig. 4
529
530
cM 1/2
3/2
1 exp (2) D
1 eAt
(66)
N0 t 1
At
(1 eAt )
At
(61)
The length of the diffusion layer is obtained from the balance between Eqs. (54
and 63) [11]
zF Dc
zF (2Dc)3/2 M 1/2 N0 t 1/2
%=
ex
1/2
(64)
as
>
= (Dt)1/2
(65)
%
According to this expression [11], the
rate of expansion of the diffusion layer
depends on the nucleation rate A, but
not on the density of active sites. For
instantaneous nucleation, = (Dt)1/2 ,
whereas for progressive nucleation, =
(3/4)(Dt)1/2 . From Eqs. (54, 60 and 64),
the current density is described as [11, 12]
i(t) =
zF Dc
(Dt)1/2
(At)1/2
eAt
2
e d
1/2
(At)
0
1 eAt
1
At
Two-dimensional
Electrocrystallization
Two-dimensional phase formation occurs
preferentially when a strong substratemetal interaction exists, a process that
typically involves the formation of growth
centers a few atoms thick, that expand and
coalesce to form a monolayer that serves
as a precursor deposit to subsequent twodimensional metal layers.
As discussed before, for a determinate
supersaturation level
, ion transfer
from solution to electrode across the electrochemical double layer occurs accompanied by the corresponding free energy
change,
GV (n) = nze||. A fraction
(n) of this energy is consumed in the
formation of the new nucleus|solution
and nucleus|surface interfaces, giving
for the total change of Gibbs free energy associated with the formation of a
cluster comprising n atoms,
GV (n) =
nze|| + (n). This expression has a
minimum
G for the equilibrium form
of the crystal with the lower surface energy i among the i crystalline faces, as
determined by the GibbsCurie expression [51], = ;i i Ai = min (for constant
volume V ).
In the case of epitaxial grow over
an ideal substrate, the surface energies
of the two-dimensional clusters reduce
to border energies i (J cm1 ), as the
emergence of the new phase is balanced
by disappearance of substrate area. Under
these conditions, the 2D crystal satises
the GibbsCurie condition for the excess
free energy, in analogy to the threedimensional case; = ;i i Li = min (at
5.3.4.2
b2
(ze||)2
(67)
(69)
2nF k 2 hM
t
(70)
I2D,prog =
2zF MhN0 k 2 t
N0 M 2 k 2 t 2
exp
(71)
2
zF MhN0 k 2 t 2
AN0 M 2 k 2 t 3
exp
3 2
(72)
531
A(r, t)
n +1
(2DcM/)1/2 1/2
(73)
= A0 1
r
where A0 is the noninhibited nucleation
rate at a sufciently large distance from a
growing nucleus. The local variation of the
nucleation rate due to the growth of nuclei
is shown in Fig. 5.
rd/r0
rd /r0
1.0
n* = 0
0.8
A(r, t)/A0
532
2
0.6
5
0.4
10
0.2
0.0
20
10
10
20
r/r0(t)
Local nucleation rate around a growing nucleus of radius r0 ,
for different critical nucleus sizes. Also shown is the exclusion zone of
radius rd , dening regions inside (A = 0) and outside (A = A0 ) the
exclusion zone.
Fig. 5
cM
1/2
N0
(Au 1 eAu )
A
eAu du
(74)
(2pNs)1/2p(r)
1.0
0.8
0.6
0.4
0.2
0.0
(2pNs)1/2r
Normalized probability density of nearest neighbor distances
for 2.2 106 drops/cm2 of mercury, electrodeposited onto vitreous
carbon at 220 mV from 0.01 mol dm3 Hg2+ aqueous solution
(
) [58]. Also shown are the probability densities corresponding to
uniformly distributed drops (thin line) and excluded nucleation for
distances less than rd = [(8 cM/)1/2 D(t u)]1/2 from each drop
(heavy line), obtained from digital simulations [59].
Fig. 6
533
534
2NS r dr
(75)
(76)
The probability density in Eq. (76) has maximum value for rm = 1/(2NS )1/2 , and we
may express it in nondimensional form in
terms of this characteristic distance, as
(2NS )1/2 p(r) = (2NS )1/2 r
exp( 12 [(2NS )1/2 r 2 ])
(77)
This normalized distribution does not depend on the number density of nuclei on
the electrode surface, and only denotes
the ordering pattern of nuclei on the electrode surface. If nuclei were, for example,
ordered on a square array, then the distance between nearest neighbors would
1/2
be 1/NS and the probability of nding the nearest neighbor would change
from 0 to 1 at a nondimensional distance
1/2
NS /(2NS )1/2 = (2)1/2 = 2.507. In
a hexagonal array, all nearest neighbors are
separated by a nondimensional distance
(4/ 3)1/2 = 2.694. For a random uniform distribution, the probability density
is maximum at the characteristic distance,
that is, (2NS )1/2 /(2NS )1/2 = 1. In
nondimensional coordinates, the maximum of the distribution shifts to higher
2.5
250
240
230
i
[mA cm2]
2.0
220
210
1.5
200
190
1.0
0.5
0.0
10
15
20
25
30
t
[s]
Fig. 7 Current transients recorded during Hg electrodeposition onto
vitreous carbon from 0.01 mol dm3 Hg2 2+ solutions in KNO3
1 mol dm3 at the indicated overpotentials (in mV) [58].
535
2.5
g
f
i
[mA cm2]
2.0
e
d
c
1.5
b
a
1.0
0.5
0.0
10
15
20
25
30
t
[s]
Current transients according to Eq. (66), for mercury
electrodeposition from 10 mM Hg2 2+ solutions, with D = 105 cm2 s1 ,
at nucleation rates A and number densities of active sites N0 ,
respectively, of: (a) 0.02 s1 and 0.4 106 cm2 ; (b) 0.04 s1 and
0.6 106 cm2 ; (c) 0.06 s1 and 0.8 106 cm2 ; (d) 0.08 s1 and
1.0 106 cm2 ; (e) 0.10 s1 and 1.2 106 cm2 ; (f) 0.12 s1 and
1.4 106 cm2 ; and (g) 0.14 s1 and 1.6 106 cm2 . The broken line
indicates the Cottrell diffusive limiting current to the electrode surface.
Fig. 8
102
101
100
A
[s1]
536
101
102
103
0.15
0.20
0.25
0.30
h
[V]
Nucleation rates of Hg onto vitreous carbon from 0.01 mol dm3
Hg2 2+ () and 0.01 mol dm3 Hg2+ (
) solutions, as a function of the
overpotential [58].
Fig. 9
N0
[cm2]
107
106
105
0.15
0.20
0.25
0.30
h
[V]
Number densities of nucleation sites on the surface of vitreous
carbon as a function of the overpotential, obtained from analysis of current
transients obtained during mercury deposition from 0.01 mol dm3 Hg2 2+
() and 0.01 mol dm3 Hg2+ (
) aqueous solutions [58].
Fig. 10
537
538
References
1. E. Budevski, G. Staikov, W. J. Lorenz, Electrochemical Phase Formation and Growth,
John Wiley & Sons, New York, 1996.
2. W. Thomson, Proc. R. Soc. Edinb. 1870, 7, 63.
3. J. W. Gibbs, Collected Works (Thermodynamics), Yale University Press, New Haven, 1948,
Vol. I.
4. R. Becker, W. Doring, Ann. Phys. 1935, 24,
719.
5. J. B. Zeldovich, Acta Phys. Chem. USSR 1943,
18, 1.
6. S. Stoyanov, Thin Solid Films 1973, 18, 91.
7. A. Milchev, S. Stoyanov, R. Kaischev, Thin
Solid Films 1974, 22, 255.
8. G. J. Hills, D. J. Schiffrin, J. Thompson, Electrochim. Acta 1974, 19, 657.
9. M. Sluyters-Rehbach, J. H. O. J. Wijenberg,
E. Bosco et al., J. Electroanal. Chem. 1987,
236, 1.
10. B. R. Scharifker, J. Mostany, M. PalomarPardave et al., J. Electrochem. Soc. 1999, 146,
1005.
11. L. Heerman, A. Tarallo, J. Electroanal. Chem.
1999, 470, 70.
12. B. R. Scharifker, J. Mostany, J. Electroanal.
Chem. 1984, 177, 13.
13. M. Avrami, J. Chem. Phys. 1939, 7, 1103.
14. W. J. Basirun, D. Pletcher, A. SarabyReintjes, J. Appl. Electrochem. 1996, 26, 873.
15. T. Erdey-Gruz, M. Volmer, Z. Phys. Chem.
1931, 157, 165.
16. T. Young, Philos. Trans. R. Soc. (Londres)
1805, 1, 65.
17. R. Aveyard, D. A. Haydon, An Introduction to
the Principles of Surface Chemistry, Cambridge
University Press, Cambridge, 1973, p. 74.
18. M. Volmer, A. Weber, Z. Phys. Chem. 1926,
119, 277.
19. L. Farkas, Z. Phys. Chem. 1927, 125, 236.
20. H. A. Kramers, Physica 1940, 7, 284.
21. E. Parzen, Stochastic Processes, Holden-Day,
San Francisco, 1962.
22. J. Frenkel, Kinetic Theory of Liquids, Clarendon Press, Oxford, 1946.
23. E. A. Guggenheim, Trans. Faraday Soc. 1940,
36, 408.
24. R. C. Tolman, J. Chem. Phys. 1949, 17, 333.
25. K. S. C. Freeman, I. R. McDonald, Mol. Phys.
1973, 26, 529.
26. J. W. Cahn, J. E. Hilliard, J. Chem. Phys.
1959, 31, 688.
539