You are on page 1of 52

Dynamic Electrochemistry

Year 3, Frontiers of Chemistry 1 Lecture Course


Dr Alison Parkin, alison.parkin@york.ac.uk, B022
Course overview
Year 2 Learning Objectives................................................................................................................................. 3
Core 5 Exam Question........................................................................................................................................ 4
Limitations to the thermodynamic-only assessment of redox reactions .......................................................... 7
Dynamic electrochemistry ............................................................................................................................... 11
Types of experiments....................................................................................................................................... 14
Film versus solution electrochemistry ............................................................................................................. 19
Film electrochemistry of electrocatalysis ........................................................................................................ 19
Film voltammetry of non-catalytic reversible redox processes ...................................................................... 29
Solution voltammetry of reversible redox processes ...................................................................................... 34
Solution voltammetry of irreversible redox reactions .................................................................................... 45
Warning from electrochemistry: Old American conventions can be confusing! ............................................ 47
Worked problems ............................................................................................................................................ 49

Figure 1 Overview of the experimental types discussed in this lecture course

1
Reading
Lecture Notes:

• “Equilibrium Electrochemistry” taught by Alison

Specialised textbooks:

• Electrode Dynamics (Oxford Chemistry Primers) by A. C. Fisher


• Electrode Potentials (Oxford Chemistry Primers) by Richard G. Compton

General chemistry texts with good section on electrochemistry, includes self-test question and answers:

• Shriver and Atkins’ Inorganic Chemistry, 5th Edition onwards


• Atkins Physical Chemistry (any edition) and dynamic electrochemistry section within

An excellent paper on solution voltammetry:

• A Practical Beginner’s Guide to Cyclic Voltammetry, Noémie Elgrishi, Kelley J. Rountree, Brian D.
McCarthy, Eric S. Rountree, Thomas T. Eisenhart, and Jillian L. Dempsey, Journal of Chemical
Education 2018 95 (2), 197-206, DOI: 10.1021/acs.jchemed.7b00361

Self-directed learning
• Workshop
• VLE “Dynamic Electrochemistry Quiz”
• Timetabled revision session
• Sample exam questions (two this year)
• Work through what was done in lectures as if for the first time (re-work through a blank handout
and check you can complete it)

General tips for staying calm and dealing with electrochemistry


• Look at the units
• Remember that everyone in your year is in the same boat- you all had the same lectures, at the
same time, you will all sit the same exam, and your questions will all be marked by the same person
• Chat any problems over with chemistry friends
• Email Alison and arrange to have a quick chat if you are stuck or confused

2
Revision: Year 2 Learning Objectives
• Know how to assign oxidation states, recognise redox reactions and identify the
oxidation and reduction processes

• Learn battery terminology: anode, where oxidation occurs; cathode, where


reduction occurs; cell voltage equation 𝐄𝐜𝐞𝐥𝐥 = 𝐄𝐜𝐚𝐭𝐡𝐨𝐝𝐞 − 𝐄𝐚𝐧𝐨𝐝𝐞 ; write out cell
diagrams
• Understand the meaning of the term “standard electrode potential”
• Apply the equations ∆𝐆 = −𝐧𝐅𝐄 and ∆𝐆 = −𝐑𝐓𝐥𝐧𝐊 to enable:
o the rationalisation of why some redox reactions are spontaneous
o the prediction of what redox reactions might occur
o the quantification of cell voltages of different battery designs
• Construct Frost diagrams from Latimer diagrams and use them to calculate
electrode potentials and predict whether a species is stable with respect to
disproportionation or comproportionation.
• Learn the Nernst equation and apply it to calculate electrode potentials and cell
potentials under non-standard conditions
• Derive an equation to describe how the electrode potential of an n-electron, m-
proton process changes with pH
• Construct and interpret Pourbaix diagrams
• Understand how activation energies are reflected in overpotential values

3
Core 5 Exam Question

4
5
6
Limitations to the thermodynamic-only
assessment of redox reactions
Learning Objectives

• Appreciate the need to measure rates of reaction

Figure 2 Zinc copper battery.

Reaction of the cell shown in Figure 2: Cu2+ + Zn  Zn2+ + Cu.


Use 1 L volumes of solution for both sides of the cell, start with 0.2 mol Zn2+ and 0.4 mol Cu2+.
Production of one molecule of Zn2+ = consumption of one molecule of Cu2+, so state

7
Equation 1
0.2 + 𝑋
[𝑍𝑛2+ ]𝑒𝑞 =
1
Equation 2
0.4 − 𝑋
[𝐶𝑢2+ ]𝑒𝑞 =
1
Thus, we can re-state the equilibrium constant as:

Using the equilibrium constant calculated previously, K = 2.943 x 1037, this equation can be solved to yield
X:

Inserting this into

9
Equation 1
0.2 + 𝑋
[𝑍𝑛2+ ]𝑒𝑞 =
1
and Equation 2 shows that at the equilibrium point the concentrations in each half-cell will have changed
so that:
[𝑍𝑛2+ ]𝑒𝑞 = and [𝐶𝑢2+ ]𝑒𝑞 =

10
Charge
For every molecule of Zn2+ that is produced and every molecule of Cu2+ consumed, 2 electrons flow. In the
cell in Figure 2, the total number of moles electrons which will flow = 2X =

Equation 3 𝑇𝑜𝑡𝑎𝑙 𝑐ℎ𝑎𝑟𝑔𝑒 (𝐶) = 𝑎𝑚𝑜𝑢𝑛𝑡 𝑒𝑙𝑒𝑐𝑡𝑟𝑜𝑛𝑠 (𝑚𝑜𝑙) × 𝐹 (𝐶 (𝑚𝑜𝑙 𝑒𝑙𝑒𝑐𝑡𝑟𝑜𝑛𝑠)−1 )

Using Equation 3, the total charge which will therefore flow = 76995 C.

However, from the above analysis we gain absolutely no information about the rate at which the
equilibrium conditions will be achieved.

Dynamic electrochemistry
Learning Objectives

• Know the role of the different electrodes in a 3-electrode dynamic


electrochemistry experimental set up and understand the requirement for
electrolyte
• Understand what current measures and differentiate between faradaic and non-
faradaic processes
• Be able to apply a reference electrode correction factor to quote measured
values as versus SHE

Dynamic electrochemistry: monitoring the rate and energetics of an electrochemical reaction.

All basic dynamic experiments

SI units of current =

11
The 3-electrode cell
The most common way to make dynamic electrochemistry measurements is to use a 3 electrode cell,
Figure 3. The potential of the working electrode is measured relative to that of the reference electrode,
while current flows between the working and the counter electrode.

Figure 3 Schematic of a three electrode cell setup.

Working electrode

IUPAC terminology for working electrode current


• Negative current = electrons moving out of working electrode surface
• Positive current = electrons moving into working electrode

Faradaic versus non-faradaic current


Current arising from a redox process occurring at a working electrode is described as “faradaic current”
Positive faradaic current indicates
Negative faradaic current indicates

Current due to non-redox reactions is described as “non-faradaic” current

12
Reference and counter electrodes

Reference electrodes
Standard electrode potentials are defined relative to a standard hydrogen electrode but working with pH 0
solutions, 1 bar H2 and Pt electrocatalysts is not experimentally practical. Instead we make/use different
reference electrodes, which have highly stable redox potentials that are precisely known versus the
standard hydrogen electrode (SHE) potential.
The reference electrode used in teaching labs is a silver/silver chloride reference electrode, as shown in
Figure 4.

Figure 4 Ag/AgCl reference electrode.

𝑜
AgCl(s) + e− ⇌ Ag(s) + Cl− 𝐸𝐴𝑔𝐶𝑙/𝐴𝑔 = +0.235 V vs SHE
The solution inside the reference electrode contains 3 M Cl-, not 1 M, but the Nernst equation can be used
to correct for this deviation from standard conditions:

3𝑀 𝐶𝑙 − 𝑅𝑇
𝐸𝐴𝑔𝐶𝑙/𝐴𝑔 = 0.235 − ln[𝐶𝑙 − ] = 0.207 𝑉 𝑣𝑠 𝑆𝐻𝐸
𝐹
A value for the standard zinc couple can therefore be measured via recording the cell voltage for a setup
like that shown in Figure 5. If we express the reference electrode potential in V vs SHE then the potential
for the Zn/Zn2+ reaction is also defined in V vs SHE without having to build a dangerous H2 half-cell.

Figure 5 Using a Ag/AgCl reference electrode to measure the electrode potential for the standard zinc couple.

13
3𝑀 𝐶𝑙 − 𝑜
𝐸𝑐𝑒𝑙𝑙 = 𝐸𝐴𝑔𝐶𝑙/𝐴𝑔 − 𝐸𝑍𝑛 2+ /𝑍𝑛

𝑜 3𝑀 𝐶𝑙 −
𝐸𝑍𝑛 2+ /𝑍𝑛 = 𝐸𝐴𝑔𝐶𝑙/𝐴𝑔 − 𝐸𝑐𝑒𝑙𝑙 = 0.207 − 𝐸𝑐𝑒𝑙𝑙

[This is no different to measuring a temperature in ᵒC and then correcting to versus K

E.g. T (K) = T (ᵒC) + 273

Boiling temperature of water = 100 ᵒC

Therefore in K, boiling temperature of water = 100 + 273 = 373 K

Counter electrodes
The reason the “third” counter electrode (also called “auxiliary” electrode on some equipment) is added is
because if current flows through the reference electrode this will perturb the ratio of oxidised:reduced
species and therefore change the redox potential.

Supporting electrolyte
The solvent system must have sufficiently high ionic conductivity so that the solution resistance does not
convolute the experimental results. In aqueous solution, NaCl is commonly used, whereas in organic
solvents a common salt is tetrabutylammonium hexafluorophosphate (Figure 6).

Figure 6 Tetrabutylammonium hexafluorophosphate

Types of experiments
Learning Objectives

• Plot the potential-time trace relevant for chronoamperometry, linear sweep or


cyclic voltammetry, including knowing what “anodic sweep” or “cathodic
sweep” means and making calculations that
o allow for different reference electrodes
o account for scan rates in different units, e.g. mV s -1, V s-1 etc
• Understand how double-layer charging gives rise to non-faradaic “baseline”
current response

14
Chronoamperometry
Chronoamperometry means to measure current (amperometry) as a function of time (chrono), while a
constant voltage is applied.

Experimental settings:
Applied voltage, Eapp (V vs Ref)
Time of hold, thold

Control:

Measure:

e.g. a chronoamperometry experiment at -1 V vs Ag/AgCl for 500 s means that the voltage waveform
applied to the working electrode is:

If use a saturated calomel reference electrode instead of a Ag/AgCl electrode:

3𝑀 𝐶𝑙 −
𝐸𝐴𝑔𝐶𝑙/𝐴𝑔 = 0.207 𝑉 𝑣𝑠 𝑆𝐻𝐸 and 𝐸𝑆𝐶𝐸 = 0.244 𝑉 𝑣𝑠. 𝑆𝐻𝐸

15
Voltammetry
In voltammetry, the voltage is modulated as a function of time, while the current is monitored. A
voltammogram is a plot of current versus voltage.

Linear sweep voltammetry


A linear voltage sweep is applied with either a positive (anodic) or negative (cathodic) gradient.

Experimental settings:
Initial voltage, Einitial (V vs Ref)
End voltage, Eend (V vs Ref)
Scan rate, υ (V s-1)

Control:

Measure:

e.g. a linear sweep voltammogram from -0.5 V to +0.5 V vs Ag/AgCl at 100 mVs-1:

16
Cyclic voltammetry
A forward and back voltage sweep is applied with both a positive and negative gradient.
Repeated voltage cycles are called “scans” or “sweeps” or “CVs”

Experimental settings:
Initial voltage, Einitial (V vs Ref)
Turning point voltage, Eturn (V vs Ref)
End voltage, Eend (V vs Ref)
Scan rate, υ (V s-1)

Control:

Measure:

e.g. a cyclic voltammogram between -0.5 V and +0.5 V vs Ag/AgCl at 0.1 Vs-1:

17
Non-faradaic current
Even in the absence of any redox active species, a baseline current response is still recorded in
electrochemistry. This is most clearly seen in cyclic voltammetry (Figure 7). This can be thought of as
equivalent to the solvent and cuvette absorbance which has to be accounted for in UV/vis spectroscopy.

Figure 7 (Left) The applied voltage-time trace and (right) resultant current-voltage response for a working
electrode in the absence of any redox active species.

This non-redox baseline current is referred to as “non-faradaic” or “capacitive” current and it arises
because of double layer rearrangements at the working electrode surface-solution interface (Figure 8).

Arrow in Figure 7: Steadily decreasing the voltage of the working electrode “charges” the surface of
the electrode with cations (+ve charge) because of electrostatic attraction (see Figure 8), and anions (-ve
charge) move away, resulting in a negative current (electrons moving away from the electrode). The rate of
movement of ions is directly proportional to the rate of change of potential, so a constant negative current
is measured.

Arrow in Figure 7: Reversing the direction of the voltage sweep reverses the direction of ion
movement resulting in an equal current of opposite sign as anions move towards the electrode surface and
cations move away.

Figure 8 Schematic of double layer in a liquid at contact with a negatively-charged solid.


18
Since scan rate directly controls the rate of ion movement, there is a linear relationship between capacitive
current and scan rate (Figure 9).

Figure 9 Cyclic voltammograms of graphite working electrode at scan rates of 10–100 mV s-1 the range of 0–
3.5 V.

Film versus solution electrochemistry


Learning Objectives

• Understand the terms “film electrochemistry” and “solution electrochemistry”

Experiments can be conducted with analyte adsorbed to the working electrode, this is “film”
electrochemistry, or experiments can be conducted with the analyte dissolved in the electrochemical cell
solution, this is “solution” electrochemistry.

Film electrochemistry of electrocatalysis


Learning Objectives

• Calculating rate of product production and total product yield assuming 100%
faradaic efficiency
• How the Nernst equation can be applied to predict the ideal sigmoidal shape of
a faradaic-only voltammogram measured for a surface-adsorbed uni-directional
catalyst reacting with substrate in solution
• How overpotential can be determined in catalytic film-voltammetry

19
Electrocatalytic film chronoamperometry

Figure 10 Experimental set up to monitor CO2 reduction catalysed by a Co protoporphyrin catalyst


immobilised on the surface of the working electrode.

If a constant potential is applied which is sufficient to reduce the Co 2+ to Co+ then the electrode transfers
electrons to the Co catalyst, the Co catalyst then transfers its electrons to CO2 and H+, converting them into
CH4 and H2O. This can be thought of as a spontaneous potential scale as shown in Figure 11, or as the inset
diagram in Figure 10.

Figure 11 Voltage diagram to envisage how electrons flow from the electrode to a substrate in solution via a
surface bound catalyst.

20
What a chronoamperometry experiment tells us

Figure 12 Current-time trace acquired from chronoamperometry of reductive electrocatalysis catalysed by a


species at the surface of the working electrode.

The current output of a film-electrocatalysis chronoamperometry reaction conducted on the system shown
in Figure 10 is as shown in Figure 12.

• What is the reaction stoichiometry of electrons per CH4?

• What is the maximum rate of production of CH4 in the experiment shown in Figure 12?

21
• Why is this a maximum rate?

• What is the maximum yield of CH4 in the experiment shown in Figure 12?

• Assuming 5 mg of the following catalyst was applied to the electrode surface, use Equation 4 to
calculate the rate of catalysis

Molecular Weight 655.03 g mol-1

Equation 4 𝑖𝑐𝑎𝑡 = 𝑛𝐹𝑚𝑐𝑎𝑡 𝑘𝑐𝑎𝑡

22
Analysing real electrocatalytic chronoamperometry data
Galia Maayan, Naama Gluz, George Christou, A bioinspired soluble manganese cluster as a water oxidation
electrocatalyst with low overpotential (2018) Nature Catalysis, 1, 48–54

23
Electrocatalytic film voltammetry
As well as chronoamperometry, catalytic redox chemistry such as that shown in Figure 10 can be also be
interrogated as a function of potential using cyclic voltammetry. Because the voltage of the working
electrode will be changing, it is helpful to consider the potential energy diagram shown in Figure 13.

Figure 13 A potential level diagram to envisage a cyclic voltammetry experiment interrogating reduction
catalysis.

In voltammetry, it is important to consider both the faradaic and non-faradaic current contributions.

Theoretical faradaic-only current


To understand the film voltammetry signal which arises when catalysis is analysed we must first
understand how the faradaic current (current from redox chemistry) changes as a function of the applied
potential. Continuing to consider CO2 reduction at a Co catalyst as our example reaction, we can see how
the simplest way to do this is to consider the Nernst equation and the catalytic scheme shown in Figure 14.

Figure 14 Catalytic scheme for CO2 reduction to CH4 at Co.

If the conversion of Co2+ to Co+ is sufficiently fast, then it can be approximated that:
𝑖(𝐸) = −8𝐹𝑘2 [𝐶𝑜+ ](𝐸)

24
For voltammetry experiments conducted at slow scan rates, 𝑚𝐶𝑜+ (𝐸), the function describing how the
molar amount of Co+ varies as a function of potential, can be derived from the Nernst equation for the
reaction Co2+ + e- ⇌ Co+.
𝑅𝑇 [𝐶𝑜 + ]
0
𝐸=𝐸 − ln
𝐹 [𝐶𝑜2+ ]
Also note that the total amount of catalyst in the experiment remains constant, thus:

[𝐶𝑜] = [𝐶𝑜 + ] + [𝐶𝑜2+ ]


Combining, yields:

𝑅𝑇 [𝐶𝑜+ ]
𝐸 = 𝐸0 − ln
𝐹 [𝐶𝑜] − [𝐶𝑜+ ]
Re-arrange:

[𝐶𝑜+ ] 𝐹( 𝐸 0 −𝐸)
= 𝑒 𝑅𝑇
[𝐶𝑜] − [𝐶𝑜+ ]
Simplify by stating:

𝐹( 𝐸 0 −𝐸)
𝑋= 𝑒 𝑅𝑇
Thus

[𝐶𝑜+ ]
=𝑋
[𝐶𝑜] − [𝐶𝑜+ ]
𝑋[𝐶𝑜]
[𝐶𝑜+ ] =
1+𝑋
Therefore, as a function of potential the faradaic current in slow voltammetry experiments is expected to
have a sigmoidal shape, as shown in Figure 15.

25
Figure 15 Faradaic-only current that is calculated for the Co catalyst at 298 K, if Eo = -1.1 V vs Ag/AgCl, [Co] =
3x10-14 and k2 = 1000 s-1.

Simplistically, this can be understood in terms of three separate voltage windows:


1. E> -0.9 V
The potential of the
working electrode is so
positive that none of the
Co2+ is reduced so catalysis
cannot proceed.

2. E < -1.3 V
The potential of the
working electrode is so
negative that all of the
Co2+ is converted to Co+ so
rapidly that the rate of
catalysis is independent of
potential

26
3. -1.3 V > E < -0.9 V
The current changes
dramatically with potential
in this region as the
equilibrium ratio of [Co2+] /
[Co+] is highly dependent
on potential across this
region

Summing faradaic and non-faradaic current


Real data is a sum of both the faradaic and non-faradaic current, i.e. Figure 16.

Figure 16 The total current measured is a sum of faradaic and non-faradaic current contributions.

27
Analysing real electrocatalytic cyclic voltammetry data
Looking at published data, Figure 17 shows a voltammogram from Shen, J., et al., 2015, Nature
Communications, 6, 8177.

E/ V vs RHE
Figure 17 1 mV s−1 cyclic voltammogram of the electrochemical reduction of CO2 on Co protoporphyrin
immobilized on a PG electrode in 0.1 M HClO4. Blue lines are negative-going (forward) scans; magenta lines
are positive-going (return) scans

• What is the overpotential of the catalyst?

• How would we expect the cyclic voltammogram of a catalyst with a higher turnover rate to
compare if the same experimental conditions (e.g. substrate supply, catalytic loading etc) were
used to compare?

28
• How long has the cyclic voltammogram shown in Figure 17 taken to run?

• Why has no electrolyte salt been added to the experiment?

Film voltammetry of non-catalytic reversible


redox processes
Learning Objectives

• Shape of the faradaic-only current response and the total current response,
including non-faradaic contributions
• Why faradaic-only peak current varies linearly with scan rate
• How the redox potential of a couple can be calculated from the voltammetric
response

Copper electron transfer proteins can be adsorbed onto graphite or gold electrodes (Figure 18).

Figure 18 Film voltammetry involves interrogating a redox active substance at an electrode surface, in this
example a copper protein is schemicatically depicted as adorbed at the electrode.

29
Theoretical faradaic current response
Again, the faradaic current can be understood based on the Nernst equation. e.g., continuing to consider a
Cu protein adsorbed on an electrode, as shown in Figure 18, the redox reaction is:
Cu2+ + e- ⇌ Cu1+ Standard electrode potential, Eo

Figure 19 Theoretical shape of faradaic current arising from a reversible one-electron redox reaction in a film
on the working electrode surface. Model data generated using equation and scan rate of 0.1 Vs-1, n = 1, A =
0.03 cm2, Γ = 1.00 pmol cm-2, T= 298 K, Eo = 0.2 V vs Ag/AgCl.

For interest:

Taken from Heering, H. A., et al., 1997, Journal of the American Chemical Society, 119 (48), 11628.

30
Interpreting Figure 19
1. Potential of the maximum positive current (oxidative peak potential), E ox, equals the potential of
the negative current peak (reductive peak potential), Ered.

i.e. in Figure 19, Eox = Ered = +0.2 V vs Ag/AgCl

2. These “peak potentials” equal the potential of the redox couple being investigated

i.e. if the process measured in Figure 19 is a Cu2+/1+ redox reaction, E(Cu2+/1+) = +0.2 V vs Ag/AgCl
under the conditions of pH and temperature used in the experiment

3. The peak area (peak integral) is equal to the charge passed, and therefore provides a measure of
number of moles of redox active material at the electrode surface
i.e. consider just the oxidative peak in Figure 19, can re-state the x-axis in terms of time rather than
potential using the scan rate of 0.1 V s-1 and basic maths:

Applied potential Potential difference between applied Time elapsed


(V vs Ag/AgCl) potential and starting voltage, -0.2 V vs through oxidative
Ag/AgCl sweep
(V) (s)

-0.2

0.0

0.2

0.4

0.6

0.8

So can re-plot Figure 19 as Figure 20.

Figure 20 Re-drawing Figure 19 with time as x-axis.


31
Units of peak area (∫ 𝑖 𝑑𝑡):

Note: this means that for film cyclic voltammetry of a reversible redox reaction, since the same amount of
redox active species is on the surface of the electrode in all experiments the current should scale linearly
with scan rate, as the time equivalent to the voltage window drops (Figure 21).

Figure 21 Linear relationship between peak current and scan rate in film voltammetry of non-catalytic
reversible redox reactions.

Total current response = sum of faradaic and non-faradaic


contributions
Figure 22 shows the type of total current-voltage response that would be predicted in ideal experiments
measuring non-catalytic reversible film voltammetry.

Figure 22 Showing total current and separate faradaic and non-faradaic contributions.
32
Analysing real film-voltammetry data of a reversible redox process

Figure 23 (left) Cyclic voltammograms of azurin at a 1-pentanethiol (n = 4) SAM coated Au electrode in 70%
[ch][dhp] solution (pH 4.6), T = 20 °C. Scan rates: 10 (grey line), 20 (orange line), 40 (green line), 60 (red line)
and 80 (blue line) V s−1. (right) “Trumpet” plot showing how peak potentials (Eox and Ered) and the average of
these potentials (Eavg) change as a function of scan rate.

In the absence of catalytic “amplification”, i.e. a reaction with substrate continually regenerating the
enzyme starting state, the faradaic current is much smaller, and therefore the non-faradaic contributions
to the total current are much more significant (Figure 24).

Figure 24 Cartoon to illustrate why catalysis “amplifies” the current response of a redox active molecule on
an electrode surface.

33
• Analyse Figure 23 to show if the current scales linearly with scan rate as it should:

Scan rate (V s-1) 𝒊𝒐𝒙


𝒑

10
20
40
60
80

• Why do faster scan rates show a greater the separation between E ox and Ered?

Solution voltammetry of reversible redox


processes
Learning Objectives

• Understand what processes contribute to the “duck shaped” response that


occurs in the voltammetric interrogation of the reversible redox chemistry of a
solution species interacting with a stationary electrode

34
• Be able to deduce the number of redox interconversions a molecule can undergo
from its solution voltammetry response
• Work out the midpoint potential and the number of electrons
• Know the relationship between peak-oxidation current and peak-reduction
current for a fully reversible reaction, and the Randles–Sevcik equation
• Interpret differences in the midpoint potential of different compounds in light of
undergraduate concepts of structure, bonding and solubility

Diffusion at a stationary macro electrode

Figure 25 Electrochemical set up for measuring the voltammetry of the reversible redox chemistry of
ferricyanide ([Fe(CN)6]3-) dissolved in solution.

When a “normal” disk shaped working electrode (i.e. radius > 1mm) is used to measure voltammetry of a
reversible redox reaction of a compound dissolved in the electrochemical cell solution (e.g. Figure 25), the
faradaic signal (current arising from redox reactions) is dominated by linear diffusion effects (Figure 26).
This means that diffusion is only significant in the direction normal to the electrode surface, and that the
influence of the electrode edge can be neglected, so reactions and transport are uniform across the
electrode surface.

Figure 26 Cartoon depicting the predominantly planar diffusion observed at a macro-electrode surface
35
Voltammetric response
Figure 27 shows a typical solution voltammogram measured for the reversible redox reaction [Fe(CN)6]3- +
e- ⇌ [Fe(CN)6]4-. It arises because of competition between the rate of electrolysis at the electrode surface
and the rate of transport of the reacting chemical species to that surface by diffusion.

Figure 27 (Top) Potential-time trace and (Bottom) corresponding current-time trace for cyclic voltammograms
of 1 mM potassium ferricyanide solution measured using a glassy carbon working electrode and scan rate 0.5
V s-1. The supporting electrolyte was 0.1 M aqueous KCl and the potentials are reported versus a 3 M Ag/AgCl
reference.

36
• At tim

e zero, a negative potential of -0.2 V vs Ag/AgCl is applied to the


working electrode.

• A small amount of redox current is


observed as the potential is swept from -0.2
up to approx. +0.1 V, as the Fe(CN)63– at the
electrode surface is reduced to ferrocya

nide, Fe(CN)64–

• As the voltage is swept from +0.1 V to


approx. +0.3 V, positive current is recorded,
indicating the oxidation of Fe(CN)64– to
Fe(CN)63–. As the potential becomes more
positive over this range the current
increases, indicating the oxidation reaction
is accelerated. This can be simply
understood by considering that the
thermodynamic driving force for the
reaction is increasing with increasing
potential, i.e. 𝑘𝑜𝑥 ∝ 𝐸.

37
• At approx. +0.3 V a current maximum is
reached, despite the continuing increase in
potential. This is because the rate of
diffusion of Fe(CN)64– from the bulk solution
to the surface of the electrode, kdiffusion,
cannot keep up with the rate of oxidation at
the electrode surface, kox.

• As the potential is raised from +0.3 V to +0.7 V the current falls.


This is because the magnitude of the current is now controlled
by the rate at which fresh Fe(CN)64– can diffuse up to the
electrode surface, kdiffusion. As electrolysis proceeds, the zone
around the electrode in which Fe(CN)64– is depleted becomes
thicker and so new Fe(CN)64– has further to diffuse, causing the
current to drop further and further.

38
• Upon reversing the scan direction, negative
current is recorded, corresponding to the
reduction of Fe(CN)63– at the surface of the
electrode to Fe(CN)64–. The same peak-like
response is seen since initially a high
concentration of Fe(CN)63– is present in the
diffusion layer and the kinetics for the
conversion of Fe(CN)63– to Fe(CN)64–
become more favourable as the potential
becomes more negative. Gradually, all of
Fe(CN)63– present in the diffusion layer is
converted to Fe(CN)64– and the current
drops to zero.

Interpreting solution-voltammetry of a reversible redox process


As shown in Figure 28, the symbols Epred and Epox are used to denote

Figure 28 Defining ip and Ep in solution cyclic voltammetry


39
Peak current in solution voltammetry

Comparing oxidative and reductive currents


ox
For a fully reversible process, |ip | = |ired
p |

Randles–Sevcik equation
When electrochemistry is under diffusion control, the peak current should be directly proportional to the
concentration of the analyte, and scale with the square-root of scan rate, as quantified by the Randles–
Sevcik equation, Equation 5.

Equation 5
𝑛𝐹𝜐𝐷
𝑖𝑝 = 0.4463𝑛𝐹𝐴𝐶 √
𝑅𝑇
Where A = electrode area in cm2, D = diffusion coefficient in cm2 s-1, C =
concentration in mol cm-3, υ = scan rate in V s-1 and other symbols have the usual
meaning.

This is because faster scan rates cause the diffusion layer to be established over a shorter distance.
Because the concentration varies from bulk to zero over a narrower length, the concentration gradient is
higher, increasing the local rate of diffusion, therefore, a faster voltage sweep rate results in a higher
current

Figure 29. Increase in peak current with scan rate.


40
Calculating redox potentials and number of electrons from the peak current
potentials
The average of the two Ep values defines the electrode potential for the chemical process, as shown in
Equation 6.

Equation 6
𝐸𝑝𝑜𝑥 + 𝐸𝑝𝑟𝑒𝑑
𝐸=
2
The number of electrons transferred in the reaction, n, is calculated using Equation 7.

Equation 7
𝑅𝑇
𝐸𝑝𝑜𝑥 − 𝐸𝑝𝑟𝑒𝑑 = 2.22
𝑛𝐹

Analysing real solution-voltammetry of reversible redox processes

Concentration effects
The cyclic voltammograms shown in Figure 31 are taken from Bogdan, M., et al., 2014, Journal of
Cheminformatics, 6 (30).

i / µA

Figure 30 (Left) Structure of ruthenium complex 2a, and (Right) cyclic voltammograms measured at a scan
rate of 0.5 V s−1 and initial concentrations of 0.2, 0.4, 0.6, 0.8 mM of 2a. Electroactive area: A = 0.064 cm 2;
potential values vs. a Ag/Ag+ reference electrode. Simulated cyclic voltammograms are indicated by dashed
lines, and measured data is shown by solid lines.

41
• Does the data in Figure 30 fit the Randles–Sevcik equation?

Concentration of 2a (mM) 𝒊𝒐𝒙


𝒑

0.2
0.4
0.6
0.8

Reversibility
E.g. examining CVs shown in Figure 31, taken from Yan, Q., et al., 2010, Journal of the American Chemical
Society, 132 (27), 9268.

Figure 31 (Left) Scheme from designing oxidation-cleaved drug delivery micelles, and (Right) cyclic
voltammograms measured before and after positive and negative electro-stimuli of micelles.

• After positive electro-stimuli, can confirm have free ferrocene in solution undergoing reversible
redox chemistry because:

42
Synthetic applications of reversible redox processes
Amatore, C.; Cammoun, C.; Jutand, A.; Electrochemical Recycling of Benzoquinone in the
Pd/Benzoquinone‐Catalyzed Heck‐Type Reactions from Arenes. (2007) Adv. Synth. Catal., 349, 292-296.

• Do the peak separations match a reversible 2-electron redox reaction?

43
Solution voltammetry of molecules capable of multiple reversible
redox reactions
As shown in Figure 32, for molecules capable of two reversible redox processes, two sets of peaks are
observed.

Figure 32 Cyclic voltammetry of 1.0 mM (E)-dimethylferrocenylethylene in dichloromethane, a redox molecule


capable of two reversible redox reactions. A freshly polished platinum working electrode (2 mm diameter)
was used and 0.1 M tetrabutylammonium hexafluorophosphate was added to the solution as the electrolyte.
The electrochemical cell was purged with a flow of N2 for 5 min before the measurement. A scan rate of 100
mV s-1 was used.

The data is adapted from Ventura, K., et al., 2017, J. Chem. Educ. and indicates that the diferrocene moiety
can undergo two separate redox reactions:
Fe3+―Fe3+ + 1e- ⇌ Fe2+―Fe3+ Ered, 1
Fe2+―Fe3+ + 1e- ⇌ Fe2+―Fe2+ Ered, 2

• What species will be prevalent over what potential window?

44
Solution voltammetry of irreversible redox
reactions
Learning Objectives

• Know what the solution voltammetry of an irreversible redox process looks like

Figure 33. Scheme depicting how irreversible chemical reactions can follow electrochemical reactions.

In many cases, an electrochemical (E) reaction can be followed by an irreversible chemical reaction (C), as
depicted in the scheme in Figure 33, leading to a distinctively different CV shape, as shown in Figure 34.

Figure 34. (a) A solution CV of a reversible redox reaction, and (b) an irreversible reduction reaction.

No oxidative wave is seen in the scan to high potential in (b) in Figure 33 because

45
Analysing real solution-voltammetry data of an irreversible redox process
Wang, F.; Rafiee, M.; Stahl, S. S.; Electrochemical Functional-Group-Tolerant Shono-type Oxidation of Cyclic
Carbamates Enabled by Aminoxyl Mediators (2018) Angew Chem Int Ed Engl., 57(22), 6686-6690.

• What is the oxidation process?

• What does the CV show?

46
Warning from electrochemistry: Old
American conventions can be confusing!
Historically, American scientists had a (very annoying/confusing) habit of plotting voltammograms
backwards, on the y-axis, the potential gets more negative from left to right and oxidative current is defined
as negative, and reductive current as positive, see

Figure 35. Mostly this is being quashed, but it is a warning to be careful when reading primary literature (I
always rotate CVs to show you them in the IUPAC convention, so there will be no tutorial or exam question
surprises!).

47
Figure 35 Unique Americans conventions

48
Worked problems
Applying equilibrium electrochemistry knowledge when analysing catalytic cyclic
voltammetry

Figure 36 Enzyme electro-catalysis data for films of hydrogenase, a H2-oxidation enzyme.

Calculate E(2H+/H2) at pH 7.0, 100% H2

Does this data indicate there is an overpotential for H2 oxidation?

49
Analysing solution cyclic voltammetry

Figure 37 Solution voltammetry for a ferrocene-containing CO-releasing molecule.

Describe what features of the voltammogram indicate that this is solution, rather than film, voltammetry.

How can the data be analysed to indicate the reversibility of the redox process being analysed.

50
Question 4

(a) Figure 2 shows linear sweep voltammograms measured for the reduction of CO2 to CO
catalysed by four different working electrodes, each of which had a surface area
of 1.0 cm2. The experiment was conducted in CO2 saturated aqueous solution, and the
potential was swept at a rate of 50 mV s−1 from oxidising to reducing potentials.

Figure 2

(i) Using the graph paper provided, draw two quantitative voltage-time plots to (5)
show a linear sweep voltammogram and a cyclic voltammogram, each swept at a
rate of 50 mV s−1 and over the same potential range as that shown in Figure 2.

Note: attach your graph paper to your answer booklet

51
(ii) Briefly describe what is meant by the Faradaic efficiency of an electrocatalytic (1)
process.

(iii) Using the data in Figure 2, identify which of the four electrodes shows the fastest (5)
rate of CO production, and calculate the maximum rate at which it produces CO
in μmol s−1. Show your working and state any assumptions that you make.

52
(b) An experiment to measure CO production versus time was carried out using a WSe2
working electrode with a surface area of 5.0 cm2 at a constant voltage of −0.8 V vs RHE.
The volume of gas above the solution in the electrochemical cell was 20 cm3, from
which 0.010 cm3 samples were withdrawn for analysis by gas chromatography (GC).

The CO peak areas determined by GC at different times are given in Table 1, and a
calibration experiment showed that a CO peak area of 1.00 corresponded to 0.010 μmol
of CO.
Table 1
Time / min CO peak area
0 0
15 283
30 567
45 850
60 1133

(i) Using the data in Table 1, and assuming that the electrochemical cell was (6)
operating at 90% Faradaic efficiency, calculate the constant current that flowed
during this experiment. Show your working.

(ii) Based on the data in Figure 2 and Table 1, estimate the CO peak area that would (3)
be measured after 15 min if the experiment to measure CO production with the
WSe2 working electrode was carried out at a constant voltage of −0.4 V vs RHE
instead of −0.8 V vs RHE. Explain your approach.

53

You might also like