Sei sulla pagina 1di 51

J.A.V.

Butler

Max Volmer

Walter Nernst

Julius Tafel

Physical and Interfacial


Electrochemistry 2013
Lecture 6.
Phenomenological electrode kinetics:
The Butler-Volmer Equation.
1 F

F
i i0 exp
exp

RT
RT


iox ired

Kinetics of interfacial ET.


Estimation of equilibrium
redox potentials provides a
quantitative measure for the
tendency for a specific
redox reaction to occur.
Kinetic information is not
derived.
ET reactions at
electrode/solution interfaces
are activated processes.
ET rate effected by:

Applied electrode potential


Temperature

Activation energy barrier


height can be effected by
applied potential. This is in
contrast to ordinary
chemical reactions.

We seek an answer to the


following questions:

How can we quantitatively


model the rate of an ET
process which occurs at the
interface between a metallic
electrode and an aqueous
solution containing a redox
active couple?
How can kinetic information
about ET processes be
derived?

We shall also investigate the


influence of material
transport, and double layer
structure on interfacial ET
processes.

Basic concepts of electrode kinetics.


For an interfacial ET
process:
current flow is
proportional to reaction
flux (rate).
Reaction rate is proportional
to reactant concentration at
interface.
As in chemical kinetics:
the constant of
proportionality between
reaction rate fS (molcm2s-1) and reactant
concentration c (molcm-3)
is termed the rate
constant k (cms-1).

All chemical and


electrochemical reactions
are activated processes.
An activation energy barrier
exists which must be
overcome in order that the
chemical reaction may
proceed.
Energy must be supplied
to surmount the
activation energy barrier.
This energy may be
supplied thermally or also
(for ET processes at
electrodes) via the
application of a potential
to the metallic electrode.

The electrode potential drives


ET processes at interfaces.
Application of a potential to an electrode
generates a large electric field at the
electrode/solution interface which reduces
the height of the activation energy barrier
and thereby increases the rate of the ET
reaction.
Hence the applied potential acts as a driving
force for the ET reaction.
We intuitively expect that the current
should increase with increasing driving
force. This can be understood using a simple
pictorial approach.

Interfacial electron transfer at electrode/solution interfaces:


oxidation and reduction processes.
Electron sink electrode
(Anode).

Electron source electrode


(Cathode).

P
ne-

Oxidation or de-electronation.
P = reductant (electron donor)
Q = Product

ne-

Reduction or electronation.
A = oxidant (electron acceptor)
B = Product

The greater the applied voltage,


the larger the resulting current
flow, and the greater the rate
of the chemical reaction.
The rate at which charge is
moved across the M/S interface
= the rate at which chemistry
is accomplished at the M/S
interface.
# electrons transferred
Electrode area (cm2)

Current (A)

dq

In electrolysis we use an applied voltage


i
dt
to perform chemistry at a M/S interface.
The applied voltage drives the chemical
reaction which does not occur spontaneously. Charge
(C)
The current flowing across the M/S
interface is a measure of the rate of
Time (s)
the chemical transformation at the interface.

nFA

dN
nFAf
dt

Amount of
Faraday Material (mol)
Constant (Cmol-1)

Reaction flux (rate)


mol cm-2s-1

Interfacial electron transfer processes are of two types :


oxidation or de-electronation , and reduction or
electronation. In oxidation, a species present in the
electrolyte (termed a reductant or electron donor) may
donate an electron to the electrode and in so doing become
chemically transformed. The electrode acts as an electron
sink and is termed an anode (fig. 1(a)). In reduction, an
oxidant present in the electrolyte accepts an electron from
the adjacent electrode, and in so doing becomes chemically
transformed. In this case the electrode acts as an electron
source and is termed a cathode (fig.1(b)) . Hence interfacial
electron transfer has a chemical implication. Substances may
be transformed and chemistry done at interfaces via
application of an external electrical potential.
We present some typical examples of oxidation and reduction
reactions in the table across. We note from table 1 that
redox reactions may be of many different types, and can
vary from simple single step processes involving the transfer
of a single electron, to complex multistep processes involving
the transfer of many electrons. A further point should be
noted here. Besides the elementary act of electron transfer
itself, one must in many cases consider further complicating
processes such as the occurence of chemical reactions
preceding or following the electron transfer step (CE or EC
processes), adsorption processes involving reactants,
intermediates or products, and electrochemical nucleation
and phase formation reactions. Furthermore the process of
reactant transport to the electrode and product transport
from the electrode will be important especially when the rate
of electron transport is high.

Fig. 1(a)

Fig. 1(b)

Electron sink electrode


(Anode).

Electron source electrode


(Cathode).

P
ne-

ne-

Reduction or electronation.
A = oxidant (electron acceptor)
B = Product

Oxidation or de-electronation.
P = reductant (electron donor)
Q = Product

Table 1. Representative oxidation and reduction reactions.


Cathodic reduction process
Fe3 aq e Fe2 aq a

2H2 O 2e H2 g 2OH

O2 g 4H aq 4e 2H2 O c
Cu 2 aq 2e Cus d
2CH2 CHCN 2H2 O 2e

CH2 CH2 CN 2 2OH

Anodic oxidation process


Ce 3 aq Ce4 aq e a

Pt H2 O PtO 2H 2e
2Cl aq Cl2 g 2e g
Fe s Fe 2 aq 2e

CH3OH H2 O CO2 6H 6e

(a) Simple outer sphere single step single electron transfer reaction. This reaction type serves as a prototype
in fundamental kinetic studies.
(b) The cathodic evolution of hydrogen. A much studied multistep, multielectron transfer reaction which
involves the formation of intermediate species adsorbed on the electrode surface.
(c) The electrochemical reduction of dioxygen. Much studied in fuel cell technology. Again a complex
reaction involving multiple electron transfer and the involvement of adsorbed intermediates.
(d) Metal electrodeposition.

An example of an electrochemical phase formation reaction involving

nucleation and growth processes.


(e) Conversion of acrylonitrile (AN) to adiponitrile (ADN). An example of electrochemistry applied to
organic synthesis.
(f) Metal oxidation. An example of a surface electrochemical redox transformation. Again this involves the
generation of transient intermediates.
(g) Chlorine gas evolution. Important in the chlor-alkali industry. This is the anodic analogue to hydrogen
evolution. Again adsorbed intermediates are involved in the reaction mechanism.
(h) Metal dissolution reaction. Fundamental basis of low temperature corrosion.
(i) Oxidation of methanol. Much studied in fuel cell research. Typical example of a complex multistep ,
multielectron transfer process.

In short, one may classify electrode processes into two


general classes : simple single step charge transfer
processes involving inorganic complex ions , and complex
multistep multi-electron transfer reactions which consist of
a number of elementary electron transfer and chemical
reactions which can occur either in a consecutive or a parallel
manner. Parallel electrode reactions are not common and are
usually encountered in the anodic oxidation of certain organic
compounds such as methanol in acidic or alkaline media.
Consecutive processes are much more common. Typical
examples of the latter type of processes are the cathodic
oxygen reduction reaction, metal dissolution processes and
various electro-organic transformations. Now analysis of the
kinetics of simple single step reactions (i.e. outer sphere
reactions involving inorganic complex ions which do not involve
bond breaking) at a phenomenological level is relatively
simple and is well understood. In contrast, the kinetics of
multistep processes is generally complicated particularly for
the case of electro-organic reactions. However some
consecutive electrode reactions are particularly suitable for
detailed attention since the number of steps involved is not
too large. A particular example of the latter is the cathodic
hydrogen evolution reaction.
An important point to note from the table presented and
from the previous discussion , is that many redox
processes which underline important technological
applications are complex. . Hence their study requires
considerable ingenuity and the use of a broad arsenal of
electrochemical and in situ spectroscopic techniques
which are surface specific. Hence the modern approach
to the study of electrode kinetics involves the use of
electrochemical and non-electrochemical techniques.

Indeed one may state that the examination of the


kinetics and mechanisms of electrochemical
reactions is one specific aspect of the broader area
of surface science.
The fundamental act in electrode kinetics, the
interfacial electron transfer reaction, can be
described in two ways. Firstly, one may adopt a
macroscopic phenomenological approach, which
results in the formal description of electron transfer
kinetics in terms of rate equations and current /
potential relationships. It is based largely on the
activated complex theory of chemical reactions.
Alternatively, one may consider a microscopic
molecular based approach. This pathway leads to
the subject of quantum electrode kinetics . Here
an attempt is made to examine the molecular
basis of interfacial electron transfer, and one
seeks to describe the effect of the molecular
structure of the reactant molecules and the
electronic band structure of the electrode, on
the kinetics of the electrode process. Quantum
electrode kinetics was discovered early on in the
development of electrochemistry, but was
neglected for many years and consequently has
not undergone the same degree of development
as its more traditional macroscopic counterpart.
At the present time there is little broad
consensus in this topic.

In this section we consider , in a very qualitative way, the


kinetic description of a simple single step, single electron
transfer process. A typical example might be the Fe2+(aq)
/ Fe3+(aq) redox reaction at a Pt electrode in an aqueous
electrolyte solution . The latter type of reaction is
termed an outer sphere electron transfer process since
no bonds are broken or made during the course of the
reaction. Consequently, we neglect complicating factors
such as diffusional transport of reactants and products ,
and adsorption effects.
How are the electron transfer kinetics examined
experimentally ? It is obvious that one must do the
measurement using an electrochemical cell containing at
least two electrodes. Indeed it is conventional to use a
three electrode arrangement as outlined across. The
electrode of interest is termed the working or indicator
electrode. The redox chemistry occurs at the interface
between the working electrode and the electrolytic
solution. A second electrode , termed the counter or
auxiliary electrode, is required to complete the circuit .
Current is passed in a circuit containing the working and
the counter electrodes. Finally , one wishes to determine
the potential difference across the working electrode
/ solution interface. The latter cannot be experimentally
determined.

What one can do however, is to measure changes


in the potential of the working electrode with
respect to a third electrode, the reference
electrode, placed in the solution near the
working electrode. The latter will have a very
high impedance to current flow and so the
potential of the reference electrode can be
considered to be constant, irrespective of the
current passed through the working and counter
electrodes. Thus, the measured change in
potential between the working and reference
electrodes will be equal to the change in
potential at the working electrode / solution
interface. Hence in a typical experiment a
potential waveform is applied to the working
electrode , chemistry is done, and the resultant
current response is monitored using a suitable
output device .
Current is passed between
working and counter
electrodes. The potential
is measured between
working and reference
electrodes/.

Reference
electrode

Working
electrode

Counter (auxillary)
Electrode.

The potential applied to the electrode is


controlled using an electronic device called a
potentiostat. It is now common practice to
perform experiments under microcomputer
control and to use the processing capabilities
of modern microcomputers to collect, store
and process the resultant data. How are the
electron transfer kinetics examined
experimentally ? It is obvious that one must
do the measurement using an electrochemical
cell containing at least two electrodes.
Indeed it is conventional to use a three
electrode arrangement .The electrode of
interest is termed the working or indicator
electrode. The redox chemistry occurs at the
interface between the working electrode and
the electrolytic solution. A second electrode ,
termed the counter or auxiliary electrode, is
required to complete the circuit . Current is
passed in a circuit containing the working and
the counter electrodes. Finally , one wishes to
determine the potential difference across
the working electrode / solution interface.
The latter cannot be experimentally
determined.

What one can do however, is to measure changes in the


potential of the working electrode with respect to a
third electrode, the reference electrode, placed in the
solution near the working electrode. The latter will
have a very high impedance to current flow and so the
potential of the reference electrode can be considered
to be constant, irrespective of the current passed
through the working and counter electrodes. Thus, the
measured change in potential between the working and
reference electrodes will be equal to the change in
potential at the working electrode / solution interface.
Hence in a typical experiment a potential waveform is
applied to the working electrode , chemistry is done, and
the resultant current response is monitored using a
suitable output device . The potential applied to the
electrode is controlled using an electronic device called
a potentiostat. It is now common practice to perform
experiments under microcomputer control and to use
the processing capabilities of modern microcomputers
to collect, store and process the resultant data.

A aq
Baq
e

We consider the latter reaction which occurs at the


interface between a metal and a solution. Let us assume that
both A and B are present in the bulk of the solution with
concentrations a and b. We now consider the effect that
charging the electrode (via application of an external
potential) has on the rate of electron transfer. Let us firstly
approach the situation at a qualitative level before we begin
detailed kinetic analysis.

Energy of electrons
in metal increases
upon application of a
potential more negative
than the thermodynamic
equilibrium value.

A net reduction (cathodic)


current flows from metal to
LUMO levels of redox active
species in solution.

n e-

LUMO
LUMO

EF

HOMO

Electron
energy

HOMO

Redox couple
in solution

Metallic electrode

Pictorial explanation of
current flow due to reduction.

Energy of electrons
in metal decreases upon
application of a potential
more positive than the
thermodynamic equilibrium
value.
Electron
energy

A net anodic (oxidation)


current flows from the
HOMO level of the redox
species in solution to the
metallic electrode.

LUMO

LUMO
n e-

EF

HOMO

HOMO

+
Redox couple
in solution
Metallic
electrode

Pictorial explanation of current


flow due to oxidation.

Consider the situation depicted in the previous two


slides, one for a net reduction process and the other or
a net oxidation process.
In this pictures (which in effect distill some very
complex quantum mechanical considerations)
we
indicate, in a very schematic way, the filled and empty
electronic states in the metallic electrode and the
highest occupied and lowest unoccupied energy levels of
the donor species A and acceptor species B in the
solution. The demarcation line between filled and empty
electronic states in the metal is designated the Fermi
energy EF. Now if species A and B are both present in
solution and if no external potential is applied to the
working electrode, then after a certain time a steady
open circuit potential termed the equilibrium potential
Ee
may be measured. Indeed the value of this
potential will depend on the logarithm of ratio of the
concentrations of A and B via the Nernst equation as
discussed previously in lecture 4. Under such conditions
we may set EF = Ee.
In contrast, when the working electrode becomes
positively charged via application of an external
potential more positive than Ee then the energy of the
electrons in the metal will be lowered and EF shifts
downwards in energy. If the applied potential is
sufficiently positive then a stage will be reached such
that EF is lower in energy than the HOMO level of the
donor species A and one can obtain a net flow of
electrons from the donor to the metal. An anodic
oxidation current flows.

Conversely, if a potential more negative than


the equilibrium value is applied to the electrode
then the energy of the electrons in the filled
levels of the metal will be raised. A stage will
be reached when EF is now higher in energy
than the LUMO level of the acceptor species
B and electrons will be transferred from the
metal to B in solution. A cathodic reduction
current will flow. We shall consider this
picture in greater detail in the next chapter
when we discuss the microscopic quantum
mechanical approach to interfacial electron
transfer.

A survey of electrochemical
reaction types.
Electrochemical reactions
are usually complex multistep
processes involving the
transfer of more than one
electron.
In this course we focus on
simple single step ET
processes involving the
transfer of a single
electron.
The kinetics of simple ET
processes can be understood
using the activated complex
theory of chemical kinetics
(see SF Kinetics notes).

Factors effecting the current/potential


response at electrode/solution interfaces.

The current observed at an


electrode/solution interface
reflects two quantities:

Charging of electrical double


layer : non Faradaic charging
current iC
Interfacial ET across
electrode/solution interface :
Faradaic current iF.

Our aim is to be able to


understand the shape that the
general current/potential curve
adopts and to be able to
interpret the voltammogram in
terms of ET and MT effects.

We initially focus attention on


the Faradaic component.
The Faradaic current iF in turn
can have components:

at low potentials : arising from


rate determining interfacial ET,
at high potentials : arising from
material transport (MT) due to
diffusion mechanisms.
These components can be
quantified in terms of
characteristic rate constants :
k0 (units: cms-1) for ET and kD
(units: cms-1) for MT.

MT

ET

Simple equivalent circuit representation of


electrode/solution interface region.
DL charging
current

iC

CDL
RS

i
Faradaic
current

iF

Electrode

Resistance of
solution

i iC iF
RCT

Solution
Double layer charging
current always presen
in addition to
Faradaic current in
electrochemical
measurements.

Temperature effects in chemical kinetics.

Chemical reactions are activated processes : they require an


energy input in order to occur. The same is true for
electrochemical reactions which occur at electrode/solution
interfaces.

Many chemical reactions are activated via thermal means. In


contrast and in addition electron transfer reactions may also be
activated via application of an electrode potential which alters the
Fermi energy of the electrons in the electrode material.
Activation

The relationship between rate constant k and temperature T is


Energy
given by the empirical Arrhenius equation (refer back to CH1101 Pre-exponential
factor
for details). The form of this equation is independent of the
physical model used to explain it. The exact interpretation of the
the pre-exponential factor A and the Activation energy Term EA

The activation energy EA is determined from experiment, by


A
measuring the rate constant k at a number of different
temperatures. The Arrhenius equation is then used to construct
an Arrhenius plot of ln k versus 1/T. The activation energy
is determined from the slope of this plot.

E
k A exp
RT

The activation energy measures the sensitivity


Of the rate constant to changes in temperature.

d ln k
d ln k
RT 2
E A R

d
T
dT
1
/

ln k

Slope

1
T

EA
R

Vant Hoff expression:


Energy

U 0
d ln K c


2
dT

P RT

TS

Standard change in internal


energy:

U 0 E E

R
P

E
R

U0
P

k
Kc
k
d k
d ln k d ln k U 0
dT ln k dT dT RT 2
P

Reaction coordinate

d ln k
E

dT
RT 2
d ln k
E

dT
RT 2

This leads to formal


definition of Activation
Energy.

Arrhenius equation: more elaborate situations.


In some circumstances the Arrhenius Plot is curved which implies that
the Activation energy is a function of temperature.
Hence the rate constant may be expected to vary with temperature
according to the following expression.

E
k aT m exp
RT
We can relate the latter expression to the Arrhenius parameters A and EA
as follows.
ln k ln a m ln T

E
RT

E
d ln k
2 m
E mRT
E A RT 2
RT
2
dT
T RT
E E A mRT

Hence

E
E
k aT m e m exp A A exp A
RT
RT
A aT m e m

Svante August
Arrhenius

A more
comprehensive
Picture of
chemical
reactivity is
given by
Transition State
Theory.

Refer to earlier lectures


Presented in this
Module by Prof. Bridge.

We begin by noting that chemical reactions are


thermally activated processes. Consequently in order
for chemical to occur, the reactant species must
initially come together in a molecular encounter and
then gain enough thermal energy to subsequently pass
over the activation energy energy barrier . We
recall that the activation energy is simply the energy
required to bring the reactants to some critical
configuration from which they can rearrange to form
products. It is clear that during the course of a
chemical reaction bonds will be stretched and broken
in the reactants and new bonds will be formed. Hence
we see that the potential energy of the system will
vary during the course of the reaction. Hence a more
complete and rigorous description involves the
examination of potential energy changes during the
course of a chemical reaction. This approach leads to
the development of multidimensional potential energy
surfaces where the potential energy of the system is
plotted as a function of various bond distances and
bond angles.
Consider the following simple reaction between an
atom A and a diatomic molecule BC in the gas phase :

A BC ABC* AB C

In the latter expression the reactants are transformed to


products via formation of a high energy activated complex
[ABC]* . If a potential energy surface is to be calculated for
this reaction then the potential energy V should be plotted
as a function of the two bond distances RAB and RBC and the

bond angle = ABC


.
Of course, this procedure would necessitate use of a four
dimensional diagram.
However to make matters less
complicated, one usually fixes one of the parameters, say ,
at a particular value, and then one examines the
corresponding three dimensional surface defined by the axes
V(R), RAB and RBC . The latter is a particular three dimensional
cut from the four dimensional energy hypersurface. A
particular three dimensional potential energy surface
corresponding to a fixed value is presented in the next
slide. We note that the course of the reaction is represented
by a trajectory on the potential energy surface from P to Q,
where P denotes the classical ground state of the diatomic
molecule BC and Q represents the ground state of the
diatomic molecule AB. Note that the variation of V(R) with
bond distance R is described by the Morse potential energy
function :

V R Do 1 exp a R Re

where Do denotes the bond dissociation energy , Re is


the equilibrium bond separation distance, and a is the
anharmonicity constant.

Activated complex
Transition state

Energy

Now the system will tend to describe a trajectory of


relatively low energy along the potential energy surface.
Typically, this pathway involves two valleys meeting at a
saddle point or col (labelled ) located in the interior of the
potential energy surface. This pathway is outlined in more
detail in the figure. Hence we see that in order for the
system to pass from the reactant state to the product state,
it will tend to travel along the bottom of the first valley
(reactant region), over the col, and down into the second
valley (product region). This minimum energy trajectory,
termed the reaction coordinate , q, is illustrated as a dashed
line in the figure presented across. This represents the
most probable pathway along the three dimensional energy
surface for the P / Q transformation. Finally, a cross section
through this minimum energy path, known as a reaction
profile is illustrated in the fig. outlined below. We pay
particular attention to the point at the top of the V(q)
profile. This point corresponds to the saddle point on the
potential energy surface, and corresponds not only to a
position of maximum energy with respect to the reaction coordinate, but also defines a position of minimum energy with
respect to trajectories at right angles with respect to the
reaction coordinate.
Hence the activated complex or transition state theory
focuses attention on the chemical species at the saddle point,
i.e., at the point where the reactants are just about to
transform into products. At this point one has the activated
complex which is that special configuration of atoms of a
system in transit between reactants and products.

Ea

Reaction coordinate

Henry Eyring
1901-1981
Developed (in 1935) the
Transition State Theory
(TST)
or
Activated Complex Theory
(ACT)
of Chemical Kinetics.

Reaction co-ordinate

0
Our presentation to date has been couched in
Where G denotes the electrochemical Gibbs energy of
qualitative terms. We can of course get more
activation. Note that the term
has got units of s-1
quantitative but this has already been covered and is expressed in cm.
Hence k ' has got units of cm s-1 as it should.
adequately earlier on in the module by Prof.
Bridge. Now we have stated that chemical
activation occurs via collisions between molecules
and that reaction involves formation of a discrete
activated complex of transient existence. The
latter species represents a configuration in which
the reactant molecules have been brought to a
degree of closeness and distortion, such that a
small perturbation due to a molecular vibration in
an appropriate direction will transform the
complex to products. The overall rate of reaction
is then equal to the rate of passage of the
activated complex through the transition state.
The Activated Complex Theory has been
discussed earlier on but we will deal with a special
case of a unimolecular reaction at an electrode
surface. Since the reaction is heterogeneous we
must include a characteristic reaction layer
thickness which we label (the length of which is
of the order of a molecular diameter).
Hence for an interfacial electron transfer
reaction we have the following.

kB T
G 0
k
exp

h
RT

Phenomenological electrode kinetics


In this section we consider the rate of electrochemical
reactions. This is the domain of kinetics . Thermodynamics is
very useful, it is a well developed subject, but it can only tell
us whether a given reaction will occur in a spontaneous
manner . In short thermodynamics sets the ground rules : it
determines whether a given reaction is energetically feasible
. If we wish to be able to determine how quickly an electrode
reaction occurs we need to apply chemical reaction kinetics .
The rate or velocity of an electrochemical reaction is
expressed in electrical terms . Reaction rate is expressed as
a current i . The quantity of charge passed across an
interface can be directly related to the amount of material
that has undergone chemical reaction as outlined via the
following expression

q nFAN k

f k Ea0

(3)

(1)

In the latter expression q denotes the charge passed , n is


the number of electrons transferred , A is the geometric
surface area of the electrode, F is the Faraday constant
and Nk represents the amount of species k chemically
transformed per unit area of the electrode surface .
hence the rate of reaction is given by :

This is a very important relationship . It expresses


the connection between the rate at which charge is
moved across the interface and the rate at which
chemistry is done at the interface . The quantity f
denotes the reaction flux or reaction velocity of the
heterogeneous electrochemical process , and is
expressed in units of mol cm-2 s-1 . We recall that the
velocity of a chemical reaction is proportional to the
concentration of the reactants , and that the constant
of proportionality is termed the rate constant . So it
is also with heterogeneous electrochemical processes
at interfaces . We consider a simple electrochemical
transformation
which occurs irreversibly
at an
ne
electrode/solution interface A
.The
flux
B
or reaction velocity will be given by the following
expression :

dN
dq
nFA k nFAf
dt
dt

(2)

In the latter we note that a0 denotes the


interfacial concentration of the reactant species A
(units : mol cm-3 ) and kE represents the
heterogeneous electrochemical rate constant (units
: cm s-1) .
Note that the corresponding units for a first order
chemical reaction occurring in the bulk of an
aqueous solution are s-1 . The difference in units
used in the electrochemical situation reflects the
heterogeneous nature of the process . Charge is
transferred between two phases of differing
composition .

Now we come to the crux of the matter .


The heterogeneous electrochemical rate
constant possesses a characteristic that is
not exhibited by ordinary chemical rate
constants . Its magnitude is strongly
influenced by the magnitude of the potential
applied to the electrode . Put another way :
the rate of the electrode process is strongly
influenced by the energy of the electrons
near the Fermi level in the electrode
material . By varying the energy of the
electrons in the metallic electrode, one can
significantly alter the rate of chemical
transformation at the interface .
This is a very significant result . We see
that the electrode potential acts as a driving
force for interfacial electron transfer . It
is the objective of phenomenological
electrode kinetics to derive a quantitative
relationship between the heterogeneous
rate constant and the driving force for
electron transfer .
Electrochemical reactions are thermally
activated processes . In order for any chemical
reaction to occur the reactant species must
initially come together in a molecular encounter
and then gain enough thermal energy to
subsequently pass over the activation energy
barrier . It is clear that during the course of a
chemical reaction bonds will be stretched and
broken in the reactants and new bonds will be
formed when the products are generated .

All that has been said so far applies to any


chemical reaction . However for electron transfer
processes at interfaces , the externally applied
potential affects the interfacial potential
difference or electric field at the interface , and
this in turn has a marked effect on the activation
energy barrier . It is the effect of the externally
applied potential on the activation energy barrier
which differentiates an electrochemical reaction
from an ordinary chemical reaction . The rate of
the latter can only be affected by temperature .
Electrochemical reactions are not only temperature
dependent, they are also potential dependent . The
heterogeneous electrochemical rate constant
depends on applied potential according in the
following manner.

F E E 0
0
k E k exp
k exp
RT

(4)

In the latter expression k0 denotes the standard


electrochemical rate constant which is a measure of
the kinetic facility of an electrode process . If k0 is
large than the ET kinetics will be fast and vice versa .
The parameter is called the symmetry factor and
determines how much of the input electrical energy
fed into the system will affect the height of the
activation energy barrier for the electrode process .
Note that will exhibit values between 0 and 1 and in
many cases will be close to 0.5 . Note that the positive
sign in the exponential refers to an oxidation process
and the negative sign to a reduction process .

It is also possible to express the information contained in


eqn.4 in the following way since kinetic expressions in
electrochemistry may be expressed either in terms of rate
constants or in terms of currents . Both representations are
equivalent .

F
i i 0 exp
(5)
RT

In the latter expression i0 is termed the exchange current


density (note that current density is simply the current
divided by the area of the electrode and so has units of A
cm-2). This quantity is related to the standard heterogeneous
rate constant k0 via :

i 0 FAk 0 a 0 b 0 1

(6)

Note also that denotes the overpotential . The


latter quantity is defined as the extra potential ,
over and above that dictated by thermodynamics,
which one must apply in order to make an electrode
reaction occur at a net finite rate . Hence
symbolically the overpotential is defined as :

e E i E e

(7)

In the latter expression we recall that E(i)


represents the potential measured with respect
to a chosen reference electrode when a net
current i is flowing and Ee is the potential
recorded at equilibrium .
We can see that the overpotential can be thought
of as an example of the concept that one cannot
get something for nothing .
It is not possible for an electrode reaction to
take place at a significant rate under conditions
at which thermodynamics says it may occur . The
deviation of actuality from the thermodynamic
expectation is the overpotential . Overpotential
has to be inputted into the system in the form of
electrical energy in order to make the electrode
reaction proceed at a finite rate . The larger the
rate , the greater is the overpotential
contribution necessary .
Hence we note that:
E(i) = - ref and Ee = e - ref ,
where ref denotes the Galvani potential of the
chosen reference electrode .

Let us now assume that the interfacial ET


reaction is reversible . Hence we can write :

k'E
A

B
k'-E

where we assume that the forward process is a deelectronation oxidation process and the reverse
reaction is an electronation reduction process .
Both processes involve the transfer of a single
electron . The net flux is given by :

f k E a0 k E b0
and the heterogeneous
constants are given by :

electrochemical

k E k 0 exp

k E k 0 exp 1

rate

In the latter expressions we introduce the


normalised potential as :

F E E 0

0
RT

Alternatively we can state

F
1 F
exp
i i 0 exp

RT
RT


This is the Butler-Volmer equation which is the
fundamental equation of electrochemical ET kinetics .
We recall that electrochemical equilibrium implied, on a
microscopic level, that the electronation and deelectronation fluxes were in balance . No net currents
passed across the electrode/electrolyte interface .
Now if a net current is passed, this balanced situation
is perturbed , a net current will flow in one specific
direction . One has a departure from equilibrium . The
greater the departure from equilibrium , the larger will
be the observed net reaction rate and the larger will
be the overpotential . Positive overpotentials
correspond to oxidation processes , negative
overpotentials to reduction processes . Zero
overpotential corresponds to equilibrium . This can be
summarised as follows.

Net oxidation : > e

Net reduction : < e


Equilibrium : = e

0
0

Transport and kinetics at


electrode/solution interface
We consider two
fundamental processes when
considering dynamic events
at electrode/solution
interfaces:

Reactant /product transport


to/from electrode surface
Electron transfer (ET)
kinetics at electrode
surface.

We first consider the


kinetics of interfacial ET
from a classical, macroscopic
and phenomenological (non
quantum) viewpoint.
This approach is based on
classical Transition State
Theory, and results in the
Butler-Volmer Equation.

DMT

ETK

In electrochemistry the rate


constant k varies with
applied potential E because
the Gibbs energy of activation
G* varies with applied potential.

Energy

Amount of
Barrier lowering

Transition state
Activated complex

Reaction
Flux
mol cm-2 s-1

=0

G0 *
Reactant
state

G *

Product
state
finite

dN
dq
nFA
nFAf
i
dt
dt

f k ' ET c0
Interfacial
reactant
Heterogeneous
concentration
ET rate constant mol cm-3
cm s-1

G * G0 * F
Total added
Electrical energy

Reaction coordinate Application of a finite overpotential


lowers the activation energy barrier
by a fixed fraction .

Symmetry
factor

overpotential

E EN

Applied
potential
Thermodynamic
Nernst potential

We use the result of TST to obtain a value for


the ET rate constant.

f k ' ET c0
Apply Eyring eqn. from TST

k ' ET

G *
Z exp

RT

k T
Z B
h
Transmission
coefficient

overpotential

G * G0 * F
Electrochemical Gibbs energy
of activation

Characteristic ET
distance (molecular
diameter).

Symmetry
factor

k 'ET

G0 *
F
exp
Z exp

RT
RT
F
0
k ET exp
RT

The important result is that the rate constant for


heterogeneous ET at the interface depends in a
marked manner with applied electrode potential.
As the potential is increased the larger will be the
rate constant for ET.

i nFAf

Butler-Volmer Equation.

For the moment we neglect the fact that mass


transport may be rate limiting and focus attention
on the act of electron transfer at the electrode/
solution interface.
We examine the kinetics of a simple ET process
in which bonds are not broken or made, involving the
transfer of a single electron in a single step.
Symmetry
factor

A(aq ) B (aq )

e
B(aq )
A(aq )

Normalised
potential

Net rate
Oxidation
and
i iox ired
Reduction
i i0 exp
exp 1
processes are
BV equation
microscopically
reversible.
Exchange
Reduction
Net current i at interface
Oxidation
Exchange current
current
component
reflects a balance between
component
provides a measure
iox and ired .
of kinetic facility
Thermodynamic
overpotential Nernst potential
Symmetry factor determines
of ET process.
how much of the input electrical
energy fed into the system will
0 1
F F E E N Exchange
i

FAk
a
b
0

current
affect the activation energy
RT
RT
barrier for the redox process.
Standard rate
Note 0< < 1 and typically = 0.5.
Applied
constant
potential

Delineating the regions of the simple


Butler-Volmer Equation.

i i0 exp exp 1
reduction

oxidation
15

i iox i0 exp

Tafel Region

10
Linear Ohmic
region

= i/i0

The Butler-Volmer
equation describes
the shape of the
current density /
overpotential
characteristic of an
interfacial ET
reaction . Typical i
versus curves are
presented across for
typical values of the
exchange current
density . Note that
when = 0 the net
current i density is
zero and the reaction
is at equilibrium .
The current density
is seen to rise
rapidly when the
overpotential
deviates from its
equilibrium value of
zero . The rate of
current increase with
overpotential
depends on the
magnitude of the
exchange current
density and hence on
the value of the
heterogeneous
standard rate
constant .

i ired i0 exp 1

-5
-10
-15
-6

Tafel Region

-4

-2

F/RT

6
1/ 2

i 2i0 sinh
2

The situation in
terms of
potential energy
curves of the
type used in
transition state
theory are
presented
across.

ired iox
i ired

Net oxidation

iox ired

Net Reduction

i iox

15
10

= i/i0

5
0

iox ired

-5

i 0

-10

Equilibrium
-15
-6

i iox ired

-4

-2

F/RT

The shape of the current/potential curve


depends on the numerical value of the
symmetry factor.
100

0.9

80
60

0.7

= i/i0

40

0.5

20
0
-20

When differs from


0.5 the i vs curve
becomes asymmetrical.

-40 0.3
-60

0. 1

-80
-100
-6

-4

-2

F/RT

Approximations to the BV equation (I).


High overpotential Tafel Limit.
F
1 F
The BV equation reduces to the

exp
i i0 exp

Tafel equation when the overpotential


RT
RT


is large, typically > 120 mV.
iox ired
At high overpotentials we assume
that the ET reaction occurs in
If >> 0 then i iox : net de-electronation
the forward direction and the
or oxidation.
reaction occurring in the reverse
F
F
direction can be neglected.
i iox i0 exp
ln i ln i0

This results in the derivation of a


RT
RT
logarithmic relationship between
If << 0 then i ired : net electronation or
current and overpotential.
reduction.
A plot of ln i vs is linear. This is
called a Tafel plot.
1 F
i ired i0 exp
Evaluation of the slope of the linear
RT

Tafel region enables the symmetry


Factor to be evaluated, whereas the exchange
Current i0 is obtained from the intercept at = 0. ln i ln i 1 F
0

RT

Tafel Plot Analysis


Linear Tafel
Region
100

Oxidation

log (i/i0)=log

10

log i
Reduction

Linear Tafel
Region

0.1

0.01

0.001
-8

d
RT
2.303
bA
F
d log i
d
RT
2.303
bC
1 F
d log i

-6

-4

-2

= F/RT

Exchange
Current evaluated
At = 0

Tafel approximation
not valid at low
overpotential

Approximations to the BV equation (II).


The low overpotential linear limit.
F
1 F
i i0 exp
exp
We note that the logarithmic

RT

RT
Tafel behaviour breaks down as
iox ired
0.
Taylor expansion at small x.
Tafel behaviour is characteristic
of totally irreversible (hard driven)
x2
expx 1 x
ET kinetics and will only be valid
2!
if the driving force for the
x2
exp x 1 x
electrode process is very large
2!
which will be the case at high
F
F
overpotentials.
exp
1

RT
RT
In the limit of low overpotentials
1 F 1 F F
1 F
exp
1
( < 10 mV) the exponential terms

RT
RT
RT RT

in the BV equation may be simplified


via use of a Taylor expansion to produce
F F
F
i
i
1
1

0
a linear relationship between current
RT
RT
RT

and overpotential (Ohms Law).


F

Linear
approximation

i i0

RT

Linear i versus
relation

Charge Transfer
Resistance

= i/i0

We introduce the concept of a


1
charge transfer resistance RCT as
being a measure of the resistance
0
to ET across the metal/solution
interface.
-1
Since RCT and i0 are inversely
proportional to each other, then a
-2
large value of RCT implies a small
-2
-1
value of i0 and vice versa.
Large RCT implies sluggish ET kinetics,
iF
and small RCT implies facile ET kinetics.
i 0
Useful idea : physical act of ET modelled
RT
in terms of a resistor of magnitude RCT.
Hence ET process described using an
electrical equivalent circuit element.

= F/RT

RT
d
RCT

di
i 0 i0 F

iRCT

Ohms Law
RCT

Interfacial ET

Transport effects in electrode


kinetics
Influence of reactant transport (logistics) becomes important
when the applied overpotential becomes very large.
The current/potential curve bends over and a current plateau
region is observed.
This observation is explained in terms of rate control via
diffusion of the reactant species to the electrode surface.
We consider a two step sequence : diffusion to the site of ET
at the electrode/solution interface followed by the act of ET
itself.
When the overpotential is very large, the driving force for
interfacial ET is very large, and so ET becomes facile and hence
no longer controls the rate.
Matter transport via diffusion (i.e. getting the reactant species
to the region of reaction) becomes rate limiting.

Transport and kinetics at electrode/solution


interfaces

Low overpotential
situation:
Charge Transfer
Control

ETK

DMT

Large overpotential
situation :
Mass Transport
Control

DMT

ETK

Bulk Phase
Reactant

Bulk Phase
Reactant

Reactant at
electrode

Fast diffusive mass


transport, slow rate
determining interfacial
ET kinetics

Product at
electrode

Reactant at
electrode

Slow rate determining


Diffusive mass transport,
Fast interfacial ET kinetics

Product at
electrode

Transport and kinetics


at electrodes. Current density

1.0

i
J

nFA nF

Net flux

ET & MT
D

kD

= i / iD

Diffusion layer approximation used.

0.5

MT

D
dc
f D k ET c0
c c0 k D c c0

dx 0

0.0
-6

-4

-2

= F(E-E0)/RT

ET

c
c0
k
1 ET
D

k ET k D c
f
k D k ET

f k ET c0
k ET k exp
0

Normalised
potential

1
1
1

f k ET c k D c

Mass transport corrected Tafel Equation


k ET k D c
kD c
fD
i
f

nFA
k ET k D 1 k D 1 k D
k ET
k ET
F E E
k ET k 0 exp
RT

fD

iD
kDc
nFA

fD
k
1 D
f
k ET

MT corrected
Tafel plot

i i
ln D
i

F E E0
iD
fD
f D f kD

1
1
0 exp
i
f
f
k
RT

f D f
iD i
kD F E E
ln
ln
ln 0
f
i
RT

k
ln( D0 )
k

ln k

D
0
k

F
RT

E E0

This is one form of the mass transport corrected Tafel equation. We see that a plot of ln (iD-i/i)
vs is linear . The slope of this plot yields - F/RT whereas the intercept directly
yields ln (kD/k0). Since kD may be readily evaluated, then the standard rate constant k0 may
be determined. This form of Tafel plot has been much used in the literature.

Other useful data analysis strategies.


The Tafel analysis may be applied with good accuracy to obtain
i0 and when the exchange current density for the electrode
process is low (typically when i0 < 10-3 A cm-2). In contrast , the
low overpotential linear approximation is useful when the exchange
current density is large (i0 > 10-3 A cm-2).

i0 FAk 0 a

A ne B

ln i0 ln FAk 0 1 ln a ln b

ln i0

1
ln a b

ln i0

lnb a

Hence we see that a plot of ln i0 versus either ln a or ln b should


be linear with slope producing a value for the symmetry factor .
We note that the Butler-Volmer equation may be used to determine
i0 and regardless of the magnitude of the overpotential using the
following procedure.

i
F
F
F F

ln

ln
i

i i0 exp
exp
exp
0

F
RT
RT RT
RT
1 exp RT

F
F
i0 exp
1 exp

RT
RT
Allen-Hickling Equation
Hence we see that a plot of ln i 1 exp F RT versus
should be linear, the slope yielding and the intercept at = 0 yielding i0 .

The molecular interpretation of electron


transfer.
We have presented a brief analysis of the
fundamentals of electrode kinetics in terms of
current/overpotential relationships . The
analysis was based on macroscopic or
phenomenological considerations . However a
proper understanding of interfacial electron
transfer requires the adoption of a
microscopic perspective . This requires
quantum mechanics .
The sharp rise in current density with increasing
overpotential as presented in the current vs
overpotential curve can be understood as follows .
Here we present a molecular interpretation of
interfacial electron transfer . It is a quantum
mechanical process and a quantitative development
requires some sophisticated mathematics .
Consequently we shall adopt the lazy mans
approach and present a qualitative pictorial
presentation of the essentials . Consider the
situation depicted in the next two slides which we
have shown previously.

In these pictures we indicate , in a very schematic manner,


the filled and empty electronic states in the metallic
electrode and the highest occupied and lowest unoccupied
energy levels of the donor species A and acceptor species B
in the solution . The demarcation line between filled and
empty electronic states in the metal is designated the Fermi
energy EF . Now if the redox species A and B are both
present in the solution and if no external potential is applied
to the metal, then as previously noted, an equilibrium
potential Dfe or Ee will be set up reflecting the balanced
Faradaic activity across the interface . The value of the
latter potential will depend on the logarithm of the ratio of
the activities of A and B via the Nernst equation as
previously discussed . Under such conditions we may set EF =
Ee . In contrast, when the electrode becomes positively
charged via application of an external potential more positive
than Ee , then the energy of the electrons in the metal is
lowered and EF shifts downward in energy . If the applied
potential is sufficiently positive then a stage will be reached
such that EF is lower in energy than the HOMO level of the
donor species A and one obtains a net flow of electrons from
donor species to metal . An anodic oxidation current flows .
Conversely, if a potential more negative than the equilibrium
value is applied to the electrode, then the energy of the
electrons in the filled levels of the metal will be raised . A
stage will be reached when EF is now higher in energy than
the LUMO level of the acceptor species B and electrons will
be transferred from the metal to species B in solution . A
cathodic reduction current will flow .

Energy of electrons
in metal decreases upon
application of a potential
more positive than the
thermodynamic equilibrium
value.
Electron
energy

A net anodic (oxidation)


current flows from the
HOMO level of the redox
species in solution to the
metallic electrode.

LUMO

LUMO
n e-

EF

HOMO

HOMO

+
Redox couple
in solution
Metallic
electrode

Pictorial explanation of current


flow due to oxidation.

Energy of electrons
in metal increases
upon application of a
potential more negative
than the thermodynamic
equilibrium value.

A net reduction (cathodic)


current flows from metal to
LUMO levels of redox active
species in solution.

n e-

LUMO
LUMO

EF

HOMO

Electron
energy

HOMO

Redox couple
in solution

Metallic electrode

Pictorial explanation of
current flow due to reduction.

The discussion has focused on the reactant species , the


electron donor or electron acceptor . Bonds become
stretched or activated , the potential energy of the system
changes and, at a critical configuration (the transition state)
the electron is somehow magically transferred . The
following question arises : how does the electron reach the
acceptor species in solution , or conversely, how does the
electron pass from the donor to the metal ?
The answer proposed from classical physics is that the
electron would be emitted from the metal, passing directly
over the potential energy barrier at the metal/solution
interface . However it can be shown that an electron emission
of this type would only generate a very tiny current , far less
than that observed experimentally . The problem is not unlike
the situation of a particle decay : radioactive substances do
decay , a particles leave the nucleus although classically they
are not meant to do so . What does happen is that the
electron (and indeed the a particle) does not pass over the
potential energy barrier, but actually tunnels through the
barrier from their levels in the metal to the vacant levels of
the acceptor ion in solution .
Hence we state :
the fundamental act of interfacial electron
transfer is a quantum event, governed by the
rules of quantum mechanics .

It is possible (Bockris and Khan 1983) to


quantitatively evaluate the tunnelling probability
PT
by solving
from first principles or
approximately (via the WKB approximation ) the
pertinent Schrodinger equation provided that a
suitable barrier geometry is proposed . Typical
barrier shapes used in calculations of this type
are rectangular or parabolic or those of the
Eckart type . For instance if we assume that
the electron tunnels through a rectangular
potential energy barrier and described by a
potential energy function of the form :

U( x) 0 x 0
U( x) U 0

0xa

U( x) 0 x a
In the latter a represents the width of the
rectangular barrier and U0 denotes the barrier
height ,
Total energy of particle

Hence we see that the overpotential for the production of a


definite reaction rate at a metal/solution interface is in fact
the shift of the Fermi Level in the metal from the value
which it had when the electrode reaction was taking place at
equilibrium .

Incidence of particle
on barrier

Transmission of particle
Through barrier

Barrier width a

III

II
Barrier
Height U0

Distance
0

Now if the energy E of the transferring electron is


less than U0 , detailed solution of the pertinent
Schrodinger equation results in the following
expression for the tunnelling probability which we
label PT :

sinh 2 ka

PT 1

4
E
E

U0
U 0

In particular when the barrier is


very high (U0 / E >> 1) and wide
(so that ka >> 1) we can write
that

provides a measure of the opacity of the barrier , and in


the classical limit PT will be very small due to the fact that
the opacity parameter will be very large .

Barrier Type

Barrier Heigth (wrt


Fermi Level /eV

Barrier Width/

Tunneling
Probability

Parabolic Barrier

1.2 x10-4

10

10

9.1 x10-12

15

15

5.2 x10-21

20

20

6.0 x10-32

1.0x10-5

10

10

8.8x10-15

15

15

1.5x10-26

20

20

1.7x10-40

and so the tunnelling probability


PT reduces to :
Rectangular
Barrier

16E
16E
E
exp 2 ka
exp 2 ka
1
U0 U0
U0

For a parabolic barrier one can show that


The tunneling probability takes the following form
Where L = barrier width and U0 is barrier maximum.

2L
1/2
PT exp
2m U 0 E

2m U 0a 2 2

2 m U 0 E

sinh ka exp ka 2

PT

The latter expressions lead to the remarkable prediction


that a quantum particle exhibits a certain finite (although
small) probability of leaking through a potential energy
barrier which is completely opaque from the viewpoint of
classical physics (since E < U0) . Indeed the parameter

We note that the tunneling probability is very sensitive to the barrier


parameter.

A further consideration must be noted . In order for


electron tunnelling to occur, the electron must move to
an acceptor energy level exactly equal in energy to
that of the electron in the metal . In other words the
electron transfer process is said to be radiationless .
These acceptor states are of a special kind : they are
said to be vibrationally excited . We can quote a
concrete example at this point . In the electronation
of a hydrated proton, which is one of the constituent
steps in the multistep hydrogen evolution reaction,
The latter reaction is often used as a prototype in the
fundamental investigation of electrode processes .
+
H 3O + e + M
MH + H2O
In the latter expression M denotes a vacant metal site
and MH an adsorbed hydrogen atom , one supposes
that the O-H bonds which receive electrons are in
excited vibrational states . The degree to which this
excitation energy exceeds the ground state energy of
the reaction is a contribution to the activation energy .
Hence molecular activation via bond stretching is a
very important contribution to the net activation
energy . One can also show that the solvation of the
ion also plays an important role . The ion in solution has
associated with it a solvation shell .
The solvation state of the reduced and oxidised form
of a redox couple is different . Hence the solvation
shell of the reactant species must be configured in a
certain way to allow the electron transfer reaction to
occur . This re-organization of the solvation shell also
requires energy and so will also be a contributor to the
energy of activation .
These ideas form the basis for the
Marcus theory of interfacial electron transfer.

The Marcus model assumes that the electron transfer rate


constant is given by :

G
k E A exp

k BT
where G* is the Gibbs energy of activation . The preexponential factor is given by

A e n n
(Sutin 1982 ; Hupp and Weaver 1984 ) where e is the
electronic transmission coefficient and represents the
probability with which electron transfer will occur once the
transition state has been formed ( and so is equivalent to the
tunnelling probability) . Typically 0 e 1 . Also n (units :
s-1 ) is the nuclear frequency factor and is defined as the
frequency at which the configuration of nuclear co-ordinates
appropriate for electron transfer is attained . In other
words it represents the rate at which the reactant species in
the vicinity of the transition state is transformed into
products .
This factor depends on the mechanism of activation and will
include contributions from solvent re-organization and bond
stretching . Its value is typically in the range 0.5 - 1.0 x 1013
s-1 . Finally n is termed the nuclear tunnelling factor and is
a quantum mechanical correction which becomes important at
low temperature . All of these factors have been discussed in
detail in a recent review (Weaver 1987) .

2e

e

k E A exp
A exp
exp

4 k B T

4k BT
2k BT

Detailed development of this Marcus model which is


based on the transition state theory of chemical
reactions results in the following expressions for the
heterogeneous electron transfer rate constants :

e 2

k E A exp

k
T
4

B
k E

e 2

A exp
4k B T

where A is a pre-exponential factor the nature of


which has been discussed by Hupp and Weaver
(1983) and denotes the re-organization energy
which consists of both an outer sphere and an
inner sphere component reflecting the two
contributions (solvent shell re-organization and
bond stretching) to the activation energy . The
latter quantity can be evaluated approximately and
is typically 0.5 to 1.5 eV

We note that if

e
the quadratic term in the energy of activation
presented in the Marcus-Hush expressions
outlined above may be expanded to first order in
e and the heterogeneous electron transfer
rate constant for the forward de-electronation
process admits the form given on the left hand
side across.

This expression is just the simple Butler-Volmer equation


with the symmetry factor having a numerical value of .
More generally according to the Marcus formulation the
symmetry factor or transfer coefficient is given by :

G 1 G 0

2
G 0 2
where we note that G* is the activation energy for electron
transfer and G0 represents the standard Gibbs energy
change on proceeding from the reactant to the product state
which for an electrochemical reaction is given by - e . The
Gibbs energy of activation is given by (Albery 1975) :

G 0
G 0
*
G 1

4

4
2

The expression for the symmetry factor presented in the


expression above contains some important information . In
general the latter quantity is a measure of the location of
the transition state along the reaction co-ordinate .

As outlined in figure 2.23 we can consider a number


of simple situations . If, for example the transition
state is symmetric then G0 = 0 and = .
Alternatively when G0 is negative the electron
transfer proceeds in an energetically downhill manner
and < and the transition state is reactant like .
Finally when G0 is positive , the electron transfer
reaction is uphill and > and the transition state is
product like .
The latter statements are a quantitative way of
expressing the so called Hammond Postulate.
Note also that the facility of electron transfer will
depend on the magnitude of the re-organization
energy . When

G 0 e
the electron transfer kinetics are slow and as noted
in eqn.2.52. On the other hand when l is small and one
has fast electron transfer then the following three
cases must be considered . When
then .
Alternatively, if

G 0

then 1 . Finally, if

G 0
then 0 .

Potrebbero piacerti anche