Lecture Electrode Processes

You might also like

Download as pptx, pdf, or txt
Download as pptx, pdf, or txt
You are on page 1of 108

Electrode Processes

Dr. Neil V. Rees

School of Chemical Engineering


n.rees@bham.ac.uk

Introduction to Electrochemistry
21-25 October 2013
Introduction
• The transfer of electrons is a key process in chemistry,
biological chemistry, materials science and chemical physics.
• Processes such as metal corrosion, fuel cells, oxidation and
reduction reactions, cell metabolism and photosynthesis, all
depend on the transfer of electrons, either heterogeneously
or homogeneously, in a solution environment.
• In general, the electron transfer does not occur in isolation
and is often coupled to chemical reactions (e.g. proton
transfers, bond cleavage, and dimerisation).
• Voltammetric methods can be used to probe these kinetic
processes, also allowing for the careful control of the
experimental conditions.
Faradaic Electrochemistry
• In electrochemical experiments, a current is generated by the
transfer of electrons between the electrode surface and any
electroactive species in the solution in contact with it.

• The current is measured and recorded as a function of the potential


applied to the electrode, and is dependent on the mass transport of
the species to the electrode, the rate of electron transfer and any
subsequent homogeneous kinetics that the species undergoes.

• We need to address the following questions:


– What is the effect of placing an electrode into solution?
– What are the processes that occur at the electrode?
– How is material transported to it?
– What does our expt. look like and how does it work?
Electrodes in Solution
• Previously we saw that a
Pt wire
potential difference arises:

M Fe2+
Fe3+
• What does the interface look
X-
H2O
like?
• What consequences are
S
there?
Structure of the Metal-Solution Interface

• Various models developed:


– Helmholtz
– Gouy-Chapman
– Stern

The classical models of the metal-solution interface:


(a) Helmholtz, (b) Gouy-Chapman, and (c) Stern.
Helmholtz Model
• The simplest model of the metal-solution interface
• Charge on the electrode produces a layer of oppositely-charged ions
from the solution, hence the term “double layer”. The locus of the
centres of these solvated ions at the electrode surface forms the
Outer Helmholtz Plane (OHP).
• The charge densities on the two layers are equal but opposite in sign,
which exactly mimics a parallel-plate capacitor (with plate separation
equal to the OHP), with a linear drop in charge between the layers.
• However, this view of the double layer did not agree with
experimental findings
– in particular it predicted parabolic electrocapillary curves whereas
experiment invariably showed asymmetric curves, implying that the
capacitance of the double layer is not independent of potential.
The Gouy-Chapman Model (1910s)

• The model assumes the metal to be a perfect planar


conductor, and the ions behave as point charges within
the dielectric medium of the solution. The excess-
charge density on the OHP is less than that on the
metal electrode, not the same.
• There is a surfeit of oppositely-charged ions in a region
close to the electrode (the “diffuse layer”), which
extends a distance  10 nm for a concentration  10-2
M, and is akin to the ionic atmosphere of an ion.
• Further out into bulk solution, there is an equal
probability of finding a positive or negative ion.
• The diffuse charge region can be approximated to a parallel-
plate capacitor with the plates separated by a Debye length,
1/ . Where
• The model allows for the potential dependence of the double
layer capacitance, but still fails to account for the shape of real
electrocapillary curves.
• The departure from real solutions can be anticipated due to
three fundamental assumptions of the model.
– it assumes point charges
– it neglects inter-ion interactions which become important at higher
concentrations
– it assumes a constant dielectric constant throughout the diffuse layer.
The Stern Model (1930s)
• Replaced the point charges in the Gouy-Chapman theory
with ions of realistic dimensions.
• Divided the solution into three regions:
– a layer of oppositely-charged ions at the OHP which have a
lower excess-charge (q2) than the electrode itself (qM),
– a diffuse layer containing a surfeit of oppositely-charged ions
(qGC),
– the bulk solution.
• There are therefore two potential drops which
correspond to the linear drop in potential from the
electrode to the OHP (as in the Helmholtz model) and an
exponential decrease in potential from the OHP to the
bulk solution (as in the Gouy-Chapman theory).
• The total capacitance of the interface is now represented by
the Helmholtz and Gouy-Chapman capacitances in series, that
is:
+

where CDL, CH and CGC are the total double-layer,


Helmholtz and Gouy-Chapman capacitances respectively.
• The Stern model agrees well with experiment, provided that
the ions are not adsorbed to the electrode, e.g. electrolytes
such as NaF (aq) which have very strongly bound solvation
spheres.
• Grahame added to the Stern model by allowing for specific
adsorption of ions onto the electrode (e.g. Cl-, Br-, I-, or PO43-),
the locus of whose centres forms the Inner Helmholtz Plane
(IHP).
Electrical Double Layer
- - +
+ - - +
- + +
- + + -
+ + + -
-
- +
- + - +
- -
- + +
- +
OHP Diffuse layer Bulk solution
φM

φS
x 12
Electrode Processes
• Two main classes of processes that can occur at an electrode.
– First, where electrons are transferred across the metal-solution
interface, causing the oxidation or reduction of electroactive species
in solution.
• These are termed faradaic, or charge-transfer, processes, since
they are governed by Faraday’s law in that the amount of
chemical reaction is proportional to the amount of charge
passed.

– Second, an electrode-solution interface may display a range of


potentials where no faradaic processes occur, either because they
are kinetically or thermodynamically unfavourable, or trivially where
there is no electroactive species present.
• In this case, processes such as adsorption/desorption and
changes in the structure of the electrode-solution interface with
changing potential, will occur and these are therefore termed
nonfaradaic processes. Although no charge is passed across the
interface, external currents will flow when the potential changes.
• When an electrode process takes place, both faradaic
and nonfaradaic processes occur
• This complicates the analysis of the experimental
data since nonfaradaic effects generally affect the
appearance of the faradaic signal.
Nonfaradaic Processes
• The most important nonfaradaic process is related to the structure of the
electrode-solution interface.
• The double-layer can be loosely described in terms of a parallel plate
capacitor, and experimentally can be characterised by a double-layer
capacitance, CDL.
• Whenever a potential is applied to a capacitor, a charging current will flow,
as charge separates on the capacitor plates (an excess of electrons on one
and a deficit on the other). Exactly the same effect is observed
electrochemically, with a double layer charging current.
• Typical double layer capacitances lie in the range 10-40 F cm-2.
• However, unlike a capacitor, the value of CDL does generally depend on the
potential.
• The electrochemical cell is often approximated to an equivalent electrical
circuit, and the simplest of these is the RC circuit where the solution is
represented by a resistance, Rs, and the double layer by the capacitance CDL:
• For a linear voltage ramp (i.e. linear
potential sweep), the charging current
response, IDL is given by:

where v is the voltage scan rate, and


which is graphically illustrated by

It should be noted that the magnitude of the charging current is linearly proportional
to the voltage scan rate; this has important consequences for cyclic voltammetry.
Faradaic Processes
• Electrochemical cells in which faradaic processes occur are separated into
galvanic and electrolytic cells. The former relates to cells where a reaction
spontaneously occurs when the electrodes are connected, e.g. fuel cells, and
the latter cell requires an external voltage to be applied to initiate a reaction

• The electrode reaction that occurs is often complicated by several factors since
it is a heterogeneous process that takes place only at the electrode-solution
interface, and yet the total current passed may also depend on homogeneous
reactions.

• The rate of reaction can therefore be affected by:


– mass transfer
– electron transfer kinetics at the electrode
– chemical reactions that precede or follow the electron transfer, and
– surface effects such as adsorption or desorption.
Mass Transport

• Mass transport is the means by which electroactive


species reach the electrode surface from bulk
solution, and has three major components:
– diffusion,
– migration,
– convection
Diffusion
• Diffusion is the movement of species through a
chemical potential gradient.
• The rate of diffusion (flux) depends on the
concentration gradient, and has been described
mathematically by Fick’s Laws.
• Fick I =>
if D is a constant, called the diffusion coefficient.
• Fick II =>
Fick I
Consider a plane, x, of area A

If there are N(x) molecules of B


on one side of the plane and
N(x+Δx) on the other

And Δx is the distance travelled


by a molecule via random walk in
time Δt
Fick II
Describes change in concentration with time
We can write

and since
Aside: Diffusion rates
• Einstein found that
• How far can a small molecule (D  10-5 cm2s-1) diffuse
in (a) 1 s? and (b) 1 day ?
Migration
• Migration is the movement of molecules according to local electric field
gradients.

• In bulk solution, there would be minimal concentration gradients, and hence


migration would be the main form of mass transport. This arises during
electrolysis where an electric potential gradient is established in the interfacial
region between the electrode and the solution, inducing the migration of
charged species towards the electrode.

• However, the dependence on small variations in localised electric fields makes


migration an undesirable mode of mass transport in a quantitative experiment.

• The conditions can be difficult to recreate consistently

• Accurate modelling of mass transport (and hence the currents passed) by


analytical or numerical methods can be challenging.
– Nernst-Planck-Poisson equations
• In order to eliminate migration, an excess of supporting
electrolyte, which is both chemically and electrochemically
inert, is usually added to the reaction solution, typically at
a concentration 50-100 times higher than the species of
interest.
• This creates an abundance of ions near the interface
region which
– maintains electroneutrality during electrolysis, ensuring that
electric fields do not build up outside the electrical double layer,
– compresses the double layer to allow the quantum mechanical
tunnelling of electrons between the electrode and species in
solution.
Convection
• Two forms of convection:
– Natural (or thermal)
– Forced.

• Natural convection occurs as a result of density gradients in solution


caused by localised thermal variations.
– generally irreproducible and is undesirable in quantitative experiments.
– can usually be eliminated by ensuring that the time duration of the
experimental measurement does not exceed 10-20 seconds, thus providing a
lower limit to the voltage scan rate used in current-potential experiments.

• Forced convection, however, is the deliberate stirring or agitation of


the solution by mechanical means.
– often designed to be part of the experiment so as to dominate the mass
transport in the system and be of a well-defined and easily interpreted
character.
– Forced convection has found widespread use in hydrodynamic electrodes,
such as the rotating disk.
The Electrochemical Cell

• The simplest arrangement for the


measurement of current-voltage
responses would be to use a two-
electrode system, consisting of a working
electrode (WE) and a reference
electrode (RE)
• The reaction of interest occurs at the WE
and hence it is typically made of an inert
conducting material, whilst the RE
should produce a stable potential so that
when the voltage is applied between the
electrodes, the potential drop between
them remains precisely defined.
If the potentials for the WE, RE and solution are given by M, REF,
and S respectively, then the applied voltage, E, can be expressed
as
E   M   S   iR S   S   REF 

where Rs is the solution resistance and i the current flowing between the electrodes.
• Of the three terms in the equation the last term is fixed due to the stable
potential of the RE.
• The aim of the voltammetric experiment is to measure the current response
as a function of the first term .
• Provided that the current flowing is very small then the iRs term can be
ignored and reduces to

and changes in the applied potential E are directly reflected in the driving force .
• However, where the currents are larger
– iRs cannot be ignored, as changes in E do not
accurately reflect changes in ,
– the voltammetry becomes meaningless.
• In this case a three electrode arrangement
is required, with separate reference and
counter electrodes (RE and CE) as shown in
• In this arrangement variations in E are now
directly reflected in .
• However, the iRs term is still not completely
eliminated and an uncompensated
resistance Ru persists.
Uncompensated resistance
• The magnitude of the iRu term, usually called the
ohmic or iR drop, can be reduced in three clear ways:
– decreasing the current passed,
– increasing the conductivity of the solution, or
– decreasing the path length between WE and RE.
Voltammetric Techniques
• Voltammetric techniques fall broadly into two classifications.
• A transient technique typically involves the application of a
short perturbation to the system, followed by measurement
of the current response.
– The perturbation applied to the system, must have a timescale
shorter than the rate of interest.
– The shape of the current-voltage response will depend on the
nature of the perturbation, but will be determined by a
combination of the kinetics under investigation as well as the mass-
transport to the electrode.
• Steady-state methods are those where mass transport is
sufficiently rapid that the electrode kinetics, and not mass
transport, is the rate-limiting factor at all points of the
voltage scan.
– The current-voltage response has a characteristic sigmoidal shape.
Voltammetric Techniques
• There are relative merits of both approaches.
• Transient techniques are commonly used for
fast kinetic measurement,
– but the application of a rapid perturbation to the
electrode-solution interface requires a relaxation
of the double layer to adjust to it,
– nonfaradaic processes such as capacitative
charging currents often complicate the
voltammetry.
• Steady-state methods are without this
complication, since the voltage scan is
relatively slow compared to mass transport,
thereby allowing adjustment of the double
layer to the electrode potential more rapidly
than the electrode kinetics.
Voltammetry
• Measurement of current (I in Amps) as a function of
potential (E in Volts)
• Linear sweep or cyclic
• Requires 3 electrodes:
• Working (WE)
• Reference (RE)
• Counter/Auxilliary (CE)
The Rate of Heterogeneous Reactions
e.g. Electro-oxidation of ferrous ions to ferric ions
Fe2+(aq) → Fe3+(aq) + e-(m)

Current
I/Amps

Fe2+ BULK Fe3+


Fe 3+
Fe2+
Diffusion
Fe2+ Fe2+
Fe2+
Fe2+
SURFACE
Fe3+
I = nFAj
I = Current (Amps)
n = Number of electrons transferred
F = The Faraday Constant
A = Area of electrode (cm2)
j = Flux of reactant (mol cm-2 s-1)
j is the rate of the heterogeneous reaction, the
analog of d[X]/dt of homogeneous kinetics.
Heterogeneous Rate Laws

j  k n Fe 
2 n
0

n = order of reaction
kn = nth order rate constant
[ ]0 = surface concentration

Generally [ ]0 ≠ [ ]bulk
k depends exponentially on the difference in potential between
the electrode (m) and the solution (s).
A change of 1V in (m - s) can lead to a change in k of over 109
Cyclic Voltammetry

• Cyclic voltammetry (often abbreviated “CV”) is the most common


method used for investigating electrochemical kinetics.
• It involves the application of a triangular voltage ramp to the cell
• The characteristic voltammogram (I-V curve) enables much
information to be determined, even from very complicated
electrode reactions.
Cyclic Voltammetry
• In cyclic voltammetry a graph is plotted of the current measured versus
the potential applied.
e.g. Consider a cell containing species X which can be electro-
oxidised.
X - e- Y
• The potential that applied is varied over time. For an oxidation, the
potential is gradually swept up to more positive, more oxidising
voltages, in an attempt to drive the reaction.
X – e- → Y
• After this the potential is gradually reduced again down to less
positive, more reducing voltages, in an attempt to drive the reverse
reaction. (*Providing that Y is a stable species - see later)
Y + e- → X
The Variation of Potential with Time
Slope = “scan rate”

E2

E/V

E1

Time/s

• The scan rate (υ) is measured in Volts per second, (V/s)


• The scan rate can be greatly varied and in general
(10-3 < υ < 106 )V/s
• A change in scan rate is useful as it allows a probe into
reaction mechanisms (see later).
• The current I is proportional to the flux j, where j = k[X]0
• Hence the current measured depends on two factors:
– the electrochemical rate constant k
– the concentration of the electroactive species on the
electrode surface [X]0
• The interplay of these two factors lead to the general shape of
the cyclic voltammogram curve.
– To start with when the potential is low, k is low and the rate of reaction is
governed by the rate of electron transfer.
– At higher potentials k is much higher, but the electrolysis already carried
out has depleted the surface concentration of X, and so the rate of
reaction is governed by the rate of diffusion of X towards the electrode
surface.
Typical Cyclic Voltammetry Curve
-
1.0 X-e Y
Diffusion
0.8
Kinetics
Electrode
0.6
Kinetics
0.4

E2
I/mA

0.2
EP E/V
0.0
E1
t
-0.2
E1 E2
-0.4

-
-0.6 Y+e X

0.0 0.1 0.2 0.3 0.4


E/V

EP = Potential at which the peak current is observed


Some relations

• The separation of the anodic and cathodic peaks, ΔEp


is used to define the electrochemical reversibility of
the process (where n electrons are passed):
– ΔEp = 59/n mV … reversible
– 59/n < ΔEp <120/n mV … quasi-reversible
– ΔEp >120/n mV … irreversible
Concentration Variation During a Scan X-e
-
Y
1.0
B
0.8
C
0.6

0.4
A

I/mA
0.2

0.0

-0.2

-0.4 F
-
-0.6
E D Y+e X

0.0 0.1 0.2 0.3 0.4


E/V

A X
Y
B X
Y

0.0010 0.0010
Concentration/(mol dm )

Concentration/(mol dm )
0.0008 0.0008
-3

-3
0.0006 0.0006

0.0004 0.0004

0.0002 0.0002

0.0000 0.0000

0.000 0.001 0.002 0.003 0.004 0.005 0.000 0.001 0.002 0.003 0.004 0.005
Distance from Electrode/cm Distance from Electrode/cm
Concentration Variation During a Scan 1.0
X-e
-
Y

B
0.8
C
0.6

0.4
A

I/mA
0.2

0.0

-0.2

-0.4 F
-
-0.6
E D Y+e X

0.0 0.1 0.2 0.3 0.4


E/V

C X
Y
D X
Y
0.0010
0.0010
Concentration/(mol dm )

0.0008
-3

Concentration/(mol dm )
0.0008

-3
0.0006
0.0006

0.0004 0.0004

0.0002 0.0002

0.0000 0.0000

0.000 0.001 0.002 0.003 0.004 0.005 0.000 0.001 0.002 0.003 0.004 0.005

Distance from Electrode/cm Distance from Electrode/cm


Concentration Variation During a Scan 1.0
X-e
-
Y

B
0.8
C
0.6

0.4
A

I/mA
0.2

0.0

-0.2

-0.4 F
-
-0.6
E D Y+e X

0.0 0.1 0.2 0.3 0.4


E/V

E X
Y F
X
0.0010 Y

0.0010
Concentration/(mol dm )

0.0008
-3

Concentration/(mol dm )
0.0008

-3
0.0006
0.0006

0.0004
0.0004

0.0002
0.0002

0.0000 0.0000

0.000 0.001 0.002 0.003 0.004 0.005 0.000 0.001 0.002 0.003 0.004 0.005
Distance from Electrode/cm
Distance from Electrode/cm
Mechanistic Notation
The CE Mechanism Step
OH
Homogeneous Chemical Reaction C CH2 H2C O + H2O
OH
+2H++2e-
Electron Transfer Step E H2C O CH3OH

The EC Mechanism Step OH O


-2e- -2H+
Electron Transfer Step E
NH2 NH
O
O
+H2O
Homogeneous Chemical Reaction C +NH3

N O
H
Reduction of Diethylfumarate
COOEt

Diethylfumarate (DEF) -1.38V


EtOOC DEF + e -
DEF∙-
Similar forward and
-1.0 backward peaks due to
E/V
-1.55 stable anion
t

0.2
-
DEF
-
DEF + e
I/mA

0.0 First Scan


Second Scan
-
DEF
-
DEF + e

-0.2
-1.0 -1.2 -1.4 -1.6
E/V

In DMF and [(n-Bu)4N]BF4


Reduction of Diethylmaleate
COOEt COOEt -1.58V
Diethylmaleate (DEM) Diethylfumarate (DEF) DEM + e -
DEM∙-
COOEt EtOOC
DEM∙- DEF∙-
-1.38V
-0.9
DEF + e -
DEF∙-
E/V
-1.7 π* orbital occupied in
t
anions, so bond order
decreased. This allows
- -
0.6 DEM + e DEM

0.4
free rotation around
0.2
bond.
I/mA

-
DEF
-
DEF + e

On reverse scan
0.0 -
DEM
-
DEM + e
-0.2

DEF + e
-
DEF
-
mostly DEF∙-
-0.4
-1.0 -1.2 -1.4 -1.6 -1.8 oxidation observed,
E/V First Scan
Second Scan
and on second scan
get DEF reduction.
In DMF and [(n-Bu)4N]BF4
Reduction of Diethylmaleate
Decreasing Scan Rate
3
- -
DEM + e DEM 0.6 DEM + e
-
DEM
-

0.4

1
I/mA

0.2

I/mA
- -
DEF DEF
- -
DEF + e DEF + e
0
0.0 -
DEM
-
DEF + e DEF
- -
-1 - DEM + e
DEM -0.2
-
DEM + e DEF + e
-
DEF
-

-2 -0.4
-1.0 -1.2 -1.4 -1.6 -1.8 -1.0 -1.2 -1.4 -1.6 -1.8

E/V First Scan E/V First Scan


Second Scan Second Scan

DEF∙- oxidation peak grows and DEM∙- oxidation peak decreases, with
decreasing scan rate. Thus, by altering the scan rate, the timescale of the
isomerisation can be probed.
E/V
0.3
An ECE Process
-0.25
t Product from C step is more easily
reduced than starting material, so do
OH OH not observe second reduction peak.
+2e+2H
- +
E
0.8

0.7

0.6

NO HNO 0.5

0.4
HO
OH I/A 0.3

0.2

-H2O
C 0.1

0.0

-0.1

HNO NH -0.2
Back peak disappears
HO OH
-0.3
at slow scan rates as C
-0.4

+2e +2H
- + -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3
process completed
E E-E /V
0

before back scan can


happen.
NH NH2
Monitoring Kinetics of Very Fast Reactions –
Reduction of Pyrilium Cation
Use tiny electrodes (<10μm) Ph
+ e-
Ph

and very fast scan rates O + O ∙ E


1.0 Ph Ph Ph Ph
Ph
Back peak decreases H
0.8
with slower scan rate
2 O ∙ O H
O C
0.6 Ph Ph Ph

0
0.4
E/V Dimerisation Rate
Constant
I/A

-0.8
0.2
t
= 2.5x109M-1s-1
0.0 Scan Rate / (V/s)
10,000 Diffusion
-0.2
100,000 Controlled
500,000
-0.4 1,000,000 (20ns Half Life)
0.0 -0.2 -0.4 -0.6 -0.8
E/V
In MeCN and Bu4NBF4
Surface-bound Species
OH O
-2e-, -2H+
Gold Surface
CH3 CH3

OH S S
O
+2e-, +2H+

• Current drops to zero after peak


• Maxima of forward and reverse peaks coincide – no diffusion
• Area (charge) under peak gives indication of the surface coverage
(mol cm-2)
I

E
Variation of Peak Current with Scan Rate
The peak current Ip varies with the scan rate.
Surface-bound Species Solution Species
I
Ip
I Ip

E
E

Ip  υ Ip  √ υ
For solution species the value of Ip is determined by the flux of the electroactive
species (X) that is reacting at the surface of the electrode. Near the peak potential
Ep, the current Ip is controlled by the rate of diffusion of X.
• Fick’s First Law shows that at potentials near
the peak potential, Ip is determined by the
concentration gradient of X near the electrode
surface.
j = -DX(∂[X]/∂r)

DX is the diffusion coefficient of X


(∂[X]/∂r) is the concentration gradient
of X

• The concentration gradient of X is determined


by the thickness of the diffusion layer (δ).

δ
[X]BULK

[X]

[X]0
r
• At faster scan rates, there is less time for electrolysis
to occur.
• This means that there is not as much depletion of X
near the surface of the electrode.
• This leads to a thinner diffusion layer near the
electrode and so a higher diffusion gradient.
• Consequently there is a higher flux and so a higher
peak current.
e.g. 1mM Ru(NH 3)6Cl3 and 0.1M KCl
100 on Graphite (BPPG) Electrode
Scan Rate (mV/s)
80 Greater peak 10
60 height with 25
40 increasing scan 50
75
20 rate 100
0 200
I/ A

300
-20
400
-40 500
-60 750
1000
-80

-100

-120

-140
-0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3
(E vs Saturated Calomel Electrode)/V
Randles-Sèvčík Equation
• Relates peak current, Ip, to square
root of scan rate
• Curved R-S plots imply
adsorption, varying n, etc

I / mA
• For a reversible couple:
• Ip = (2.69x105)n3/2ACbulkD1/2v1/2
where D = diffusion coefficient /
cm2s-1 1/2/V1/2s-1/2

A = surface area / cm2


Cbulk = bulk concentration, mol cm-3
n = no. of electrons transferred
V = voltage scan rate / Vs-1
• For an irreversible couple

• Ip = (2.99x105)nα1/2ACbulkD1/2v1/2
where α = transfer coefficient of the RDS
Electrode Kinetics
Consider the general electrode process
k red

Ox ( aq )  e  ( m ) Red ( aq)
ko x
Using the relationships already discussed, the currents passing at the
electrodes can be given by (reductive current at cathode, oxidative
current at anode)
I red   FAk red [Ox]0
I ox  FAk ox [Red]0
I  FA (kox [Red]0  k red [Ox]0 )
The rate constants are given by G G ╪

transition state theory, where ΔG ╪


red
ΔG╪ox

kred = Zexp(-ΔG╪red/RT)
Gox

kox = Zexp(-ΔG╪ox/RT) Ox(aq) + e-(m)


Gred Red(aq)

Reaction Coordinate
where ΔG╪red = G╪ - Gox and ΔG╪ox = G╪ - Gred

In order to calculate the molar Gibbs energy of each species we


need to work out their chemical energy (the familiar chemical
potential from homogeneous kinetics), and their electrical energy.

Molar Gibbs energy = constant + zF

Chemical Energy, (depends on P and T) Electrical Energy


z = charge,  = potential
e.g. consider example Fe3+(aq) + e-(m) Fe2+(aq)

Reactants (Fe3+(aq) + e-(m)) Products (Fe2+(aq))

Gox = constant + 3F s – F m Gred = constant’ + 2F s


Gox = constant + 2F s – F(m – s)
The transition state is assumed to have a potential difference in
between that of the reactants and products
G╪ = constant + 2F s – (1-α)F(m – s)
The value of α, the transfer coefficient, gives a measure of how
sensitive the transition state is to the drop in potential between
the metal and the solution. It is possible for α to take values
between 0 and 1. When α is near 0 then the transition state is
reactant like in character, and when α is near to 1 then the
transition state is product like in character. Often α ≈ 0.5,
indicating that the transition state is intermediate in character.
Hence
ΔG╪red = constant + 2F s – (1-α)F(m – s) – [constant + 2F s – F(m – s)]
ΔG╪red = constant + αF(m – s)

ΔG╪ox = constant + 2F s – (1-α)F(m – s) – [constant + 2F s]


ΔG╪ox = constant – (1-α)F(m – s)
Since
kred = Zexp(-ΔG╪red/RT)
ln (kred) = ln Z - (ΔG╪red/RT)
ln (kred) = ln Z – (constant/RT) - (αF/RT)(m – s)

Define
k0red as kred when m = s
ln(k0red) = ln Z – (constant/RT)

Hence
ln (kred) = ln(k0red) - (αF/RT)(m – s)
kred = k0redexp[- αF(m – s)/RT]
Similarly
kox = Zexp(-ΔG╪ox/RT)
ln (kox) = ln Z - (ΔG╪ox/RT)
ln (kox) = ln Z – (constant/RT) + (1-α)(F/RT)(m – s)

Let
β = (1 – α) [i.e. α + β = 1] **

Define
k0ox as kox when m = s
ln(k0ox) = ln Z – (constant/RT)

Hence
ln (kox) = ln(k0ox) + [βF/RT](m – s)
kox = k0oxexp[βF(m – s)/RT]
I Reversible

Irreversible

Overpotential η

There are two types of kinetics for the Butler Volmer equation.
• In the reversible case hardly any overpotential is needed to drive
the reaction in either direction. The overall current I will have a
significant contribution from both Iox and Ired. Only at very large
overpotentials will one of these become negligible.
• In the irreversible case a very high overpotential is needed to drive
a current. By the time the overpotential is large enough to drive I red
for example, then Iox becomes negligible. This allows the Butler-
Volmer Equation to be simplified.
When η is large and positive, exp[- αFη/RT] becomes very small
and insignificant. Hence the equation can be simplified to
I  [Red]0 exp[βFη/RT]
ln(I) = ln(I0) + βFη/RT
When η is very negative, exp[βFη/RT] becomes very small and
insignificant. Hence the equation can be simplified to
I  – [Ox]0 exp[- αFη/RT]
ln(-I) = ln(I0) - αFη/RT
This allows the values of α and β to be found graphically from
the gradient of a Tafel plot
Tafel Plot Region ln|I| Tafel Plot Region

ln|I0|

Overpotential η
CV as Characterisation:
Electrochemical Surface Area

• Classical example of
determining active area of Pt
• Scan CV of Pt electrode
(catalyst) in acid
• Integrate CV under H
adsorption peaks to find Qads
• Qads (polycrx Pt) related to
area by
Interpretation of Cyclic Voltammograms

• For inference of kinetic parameters from cyclic


voltammograms, it is usual for certain key features to be
used.
– peak currents,
– peak potentials, and
– anodic-cathodic peak separations.
• It is also possible to model the entire waveshape with some
commercial software packages (e.g. DIGISIM, DIGIELCH,
COMSOL, etc), or by using mathematical optimisation
methods in concert with simulation programs.
• The removal of ohmic drop effects is commonly achieved by use of on-line compensation.
There are few other methods, save perhaps the semi-integral or convolution
techniques.55,56 Such convolution methods, however, do have the added advantage of
identifying the double layer capacitance Cd, and therefore enables easier removal from the
faradaic response. Removal of capacitative charging currents from data is often
challenging, and most techniques are not completely without some degree of subjectivity.
It has been reasoned by Savéant et al, that there is an interaction between faradaic and
non-faradaic currents which renders any method which seeks to separate them by treating
as independent processes as inaccurate.14 These could include such methods as
subtraction of a “blank” voltammogram (i.e. an identical scan except electroactive material
has been removed from the solution), and simple subtraction of a baseline where it has
become well-established before the faradaic region of the scan.
• It should be recalled here that the faradaic current varies as the square root of scan rate,
whereas the charging current depends linearly on it (see equations (1.4), (1.18) and
(1.19)). Therefore as scan rates increase, there is a tendency of the faradaic signal to
become swamped by the charging current, especially when operating with low
concentrations.

• Kinetic Information & Notation
• There are two basic kinetic processes occurring at or near to the
electrode. Heterogeneous kinetics in its widest sense encompasses the
kinetics of processes undergone at the electrode-solution interface,
therefore including adsorbed species, electrodeposition, corrosion, and
surface-modified electrodes, as well as simple electron transfer. In the
context of this work however, we shall restrict its use to describing the
act of electron transfer between electrode and electroactive species at
the electrode-solution interface.
• Homogeneous kinetics describes the processes occurring in solution.
Although homogeneous processes will include electron transfer between
species (e.g. disproportionation) as well as chemical reactions (e.g. bond
scission, dimerisation), it is the latter reaction steps within the diffusion
layer that are most relevant to this work.
• Heterogeneous Electron Transfer
• There are several approaches to rationalising the rate of electron transfer at
an electrode, ranging from the phenomenological, through classical
dynamics to the quantum-theoretical.
• In the case of a reversible reduction of species A to its stable radical anion B:

• (1.30)
• where kf and kb are the rate constants for the forward (reductive) and
backward (oxidative) processes respectively.
• Butler and Volmer found that these rates could be expressed in terms of a
standard electrochemical rate constant, k0.
• (1.31)
• (1.32)
• where α is the transfer coefficient, which provides a measure of the symmetry of the energy barrier, that is, if α =
0.5 the transition state lies midway between A and B.
• These can be used to determine the amount of current flowing, since the net current must be a sum of the
reductive (cathodic) and oxidative (anodic) components:

I  Ia  Ic
 nFAk b B x 0  FAk f Ax 0
 
 F E  E 0f
 nFAk 0 B x 0 exp
  A 
  1   F E  E 0f


x  0 exp 
  RT   RT 
   

Define an overpotential given by , and if the solutions are well stirred such that surface
and bulk concentrations are approximately equal.
Since the equilibrium position is given by Nernst:

RT  Abulk 
E  E 0f  ln  
nF  B bulk 
• then
RT  Abulk 
 E eq  E   E eq  E 0f  ln  
nF  B bulk 
RT  Abulk 
E  E 0f    ln  
nF  B bulk 

• Combining these
 F  Abulk   1   F  Abulk 
  1

I  nFAk 0 B x 0 exp     Ax 0 exp    1 
  RT  B bulk  RT  B bulk 
 A  B x 0   F   Ax 0    1   F 
 nFAk 0  1bulk  
  
 exp     exp 
 B bulk  B bulk   RT   Abulk   RT 
 B    F   Ax 0    1   F 
 I 0  x 0  exp     exp 
 B bulk   RT   Abulk   RT 
• where
is called the exchange current.
• If the mass transfer is sufficiently rapid (eg. vigorously stirred
solution, small electrodes) and the volume is large enough so
that bulk concentrations remain constant, we can
approximate bulk and surface concentrations, simplifying the
result to the Butler-Volmer equation:
  F    1   F 
I  I 0 exp   exp 
  RT   RT 
• Consider the extremes of overpotential:
– When  = large & positive,

 F 
I  I 0 exp 
 RT 
F
ln I   ln I 0
RT

– When  = large & negative,


  1   F 
I  I 0 exp 
 RT 
 1   F
ln I   ln I 0
RT

– Hence Tafel plots of ln I or ln j vs .


• An electrochemically reversible reaction is one where the
electrode kinetics are much faster than the mass transport, i.e.
• (1.34)
• where the mass transport coefficient, kMT, is related to the flux of
species at the electrode by15
• (1.35)
• By contrast, an electrochemically irreversible reaction is where
there is a significant energy barrier to reaction and so mass
transport is faster than the electrode kinetics. An overpotential, ,
is required to surmount the free energy barrier to electron
transfer.
• Quasi-reversible reactions display intermediate behaviour.
• Example 4: ECE reaction
• This is where the chemical
step yields a product which is
itself electrochemically active
in the experimental potential
range, e.g. the reduction of
1,2-bromonitrobenzene:70
AC Impedance / EIS

You might also like