Voltammetry Overview

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Voltammetry j Overviewq

Robert J Forster, Darren Walsh, Kellie Adamson, and Elaine Spain, Dublin City University, Dublin, Ireland
© 2019 Elsevier Ltd. All rights reserved.

Introduction

Cyclic voltammetry, CV, provides a powerful and direct insight into the energetics of redox reactions, the dynamics and reversibility
of the electron transfer as well as the rates of coupled chemical reactions. It involves scanning the potential applied to a working
electrode according to triangular waveform and monitoring the resulting current flow. The potential range is typically selected in
order to switch the oxidation state of an electroactive species that can be in solution or adsorbed at the electrode surface. The current
response and the potential at which it is observed depends on the identity of the species in solution, their concentration and their
electrochemical properties. Thus, quantitative information about the energetics and dynamics of the redox reactions can be
extracted and, while it is not the most sensitive of electrochemical techniques, the identity can be inferred and the concentration
determined.
Cyclic voltammetry has found applications in areas as diverse as elucidating complex reaction mechanisms and quantifying key
environmental, industrial and disease biomarkers. Not only can it be used to detect electroactive species, it provides deep insights
into electrochemical reversibility, the dynamics and energetics of electron transfer across the electrode/solution or electrode/film
interface, adsorption processes, following or coupled chemical reactions and the stability of reaction products.
From an instrumentation perspective, CV is easy to implement and even the ubiquitous smartphone has been used as a “poten-
tiostat” in which the audio output is used to generate and apply an appropriate waveform to an electrochemical cell. This approach
enables point-of-need analysis and the analytical data can be moved instantaneously around the globe. Their small size and robust
engineering allows smartphone-based cyclic voltammetry coupled with modified working electrodes to be applied in environ-
mental analysis, food quality monitoring and point of care diagnostics, for example, detection of glucose concentration in diabetics,
as well as biomarkers for infectious and noninfectious disease. This capability raises the interesting possibility of using mobile
phones and cyclic voltammetry for the early detection of emerging health epidemics, for example, influenza, Ebola or Zika virus,
by distributing ultralow cost sensors to entire populations and having them track their health status without the need for clinicians
to be involved. Equally, practical applications of smartphone-based systems integrated with CV for environmental monitoring, for
example, of metals and redox active organic pollutants in waterways, have been demonstrated. Another key area where the simple
instrumentation required to implement CV has a significant impact is in food safety. Diverse targets from antioxidant concentra-
tions to the identification of species specific DNA to confirm the origin of meat, virgin olive oil, wines and other foodstuffs have
been successfully implemented.
The other key development in the implementation of cyclic voltammetry has been the shrinking of the electrode size typically
from several millimeters to micrometer and even nanometer length scales. With the advent of instruments capable of measuring the
low currents generated at these small electrodes, it has been possible to routinely measure rapid electron transfer reactions and,
while still not very widespread, to even determine the electrochemical properties of species that exist for just a few nanoseconds,
for example, photoexcited molecules. Moreover, it is possible to perform CV in unusual media ranging from cartilage to cement
because of their insensitivity to ohmic or resistive effects.
In this article, the potential waveform employed to run cyclic voltammetry experiments is first described. The current–voltage
waveforms for both reversible and irreversible redox reactions are then presented with an emphasis on the various parameters
that can be extracted from cyclic voltammograms. Finally, contemporary applications of the technique are then outlined, such
as monitoring heterogeneous electron transfer dynamics across the electrode/solution interface and investigation of redox reactions
with following chemical reactions.

Potential Sweep and Current Responses

Fig. 1 shows the triangular potential waveform employed in cyclic voltammetry. Typically, the potential is ramped linearly from an
initial potential, Ei, to the switching potential, Emax. The direction of the potential sweep is then reversed and scanning continues
until Emin is reached. The potential sweep may be terminated at the end of the first cycle or it may continue for an arbitrary number
of cycles. The primary experimental parameters are the initial potential, the switching potentials and the potential sweep or scan
rate. Typical scan rates for cyclic voltammetry, employing electrodes of conventional sizes (e.g., disk electrodes of radii 1–2 mm)
are in the range 1–1000 mV s 1. However, due to the low currents passed, the accessible range of experimental sweep rates may
be extended to the MVs 1 range by using microelectrodes.

q
Change History: November 2018. Robert J Forster updated the text and references.
This is an update of R.J. Forster, D.A. Walsh, Voltammetry j Overview, Editor(s): Paul Worsfold, Alan Townshend, Colin Poole, Encyclopedia of Analytical
Science (Second Edition), Elsevier, 2005, Pages 181–188.

Encyclopedia of Analytical Science, 3rd edition, Volume 10 https://doi.org/10.1016/B978-0-12-409547-2.14562-7 209


210 Voltammetry j Overview

Potential (V)

Emax

Ei

Emin


Time (s)

Fig. 1 Triangular potential waveform employed in cyclic voltammetry. Emax and Emin are the switching potentials. tl is the switching time. Ei is the
initial potential.

Fig. 2 Cyclic voltammogram for a reversible, one-electron redox reaction. The electrode area, A, is 0.1 cm2, the concentration, C, of the molecule in
solution is 5 mM, the diffusion coefficient, D, is 4  10 6 cm2 s 1, the rate of heterogeneous electron transfer, ko, is 0.1 cm s 1, the formal poten-
0
tial, Eo , is 0.300 V and the scan rate is 100 mV s 1. The initial potential is þ0.700 V.

Fig. 2 shows the CV response of an electrochemically reversible electroactive species in solution. In order for a species to be
electro-chemically reversible, the time constant for electron transfer should be shorter than that of the CV experiment and the
species must be chemically stable in both oxidation states. In this example, the analyte exists in the state and so the initial poten-
tial is set at a value that is well positive of the formal potential so that no current flows when the cell is first turned on, in this
case, þ 0.7 V. As the potential is scanned in a negative potential direction, the energy of the electrons in the electrode is increased
and, at a certain potential, approximately 0.5 V, electron transfer to the molecules in solution that are adjacent to the working
electrode surface is thermodynamically downhill and a reduction current is observed. This process causes a cathodic current, asso-
ciated with reduction of the oxidized form of the analyte, to flow which increases rapidly as the surface concentration of oxidized
species becomes zero. The magnitude of this current is controlled by the rate of heterogeneous electron transfer across the inter-
face. However, as the potential is made more negative and the rate of heterogeneous electron transfer increases, diffusional mass
transport begins to dominate the response and a diffusion-controlled current peak is observed, at approximately 0.32 V in this
example. At more negative potentials, the current magnitude is now controlled by the rate at which oxidized material diffuses to
the electrode surface and becomes reduced. At potentials negative of 0.32 V, the current decays as t 1/2 according to the Cottrell
equation.
At the switching potential, 0 V, the potential sweep is reversed and the potential remains sufficiently negative to cause a reduction
reaction at the electrode surface. This small cathodic current continues until the working electrode potential becomes sufficiently
positive to cause oxidation of the reduced species adjacent to the electrode surface (c. 0.25 V). This causes a diffusion-controlled
anodic current to flow, which increases until the thin layer up against the electrode is depleted of reduced material and the anodic
current peaks at approximately 0.38 V. As the potential returns to the initial potential (0.7 V), a small anodic current still flows as the
Voltammetry j Overview 211

reduced species diffuses to the electrode surface and is oxidized. In the cyclic voltammogram, important quantitative parameters of
interest are the cathodic and anodic peak potentials (Ep,c and Ep,a), the magnitudes of the cathodic and anodic peak currents (ip,c and
ip,a), the difference between the cathodic and anodic peak potentials (DEp ¼ jEp,a  Ep,cj), the potential at half height of the peak
(Ep/2 ¼ E at i ¼ ip/2) and the redox formal potential (E0 ¼ [Ep,a þ Ep,c]/2).

Reversible (Nernstian) Behavior

For a reversible system in which an oxidized species is reduced with the transfer of n electrons, the voltammetric peak current (in
amperes) at 298 K is given by the Randles-Sevçik equation:

ip ¼ 2:69  105 n3=2 A D1=2 C y1=2 (1)

where A is the electrode area (cm2), D is the diffusion coefficient (cm2 s 1), C is the bulk concentration of the oxidized species
(mol cm 3) and y is the potential sweep rate (V s 1).
A common diagnostic of a reversible system is the difference between Ep and Ep/2:

RT 56:5
Ep  Ep=2 ¼ 2:2 ¼ (2)
nF n
where R is the gas constant (8.314 J K 1 mol 1), T is the temperature (K) and F is the Faraday constant (9.6485  104 C mol 1).
The peak potential difference, DEp, depends on the value of the switching potential relative to the peak potential. When the
switching potential exceeds the peak potential by an infinite amount, DEp is equal to 57/n mV. DEp increases slightly as the
difference between the peak and switching potentials decrease. For example, when the switching potential is 71.5 mV past Ep, DEp is
60.5 and when the potential is switched 121.5 mV past the peak potential, DEp equals 59.2/n mV.
Overall, for a reversible (Nernstian) process, ip is proportional to y1/2, Ep is independent of y, r Ep  Ep/2 r is 56.5/n mV, DEp is
approximately 57 mV and | ip,c/ip,a | is 1.

Irreversible Behavior

Irreversible behavior arises either because the rate of electron transfer is very low or because the electrogenerated species decompose/
undergoes a following chemical reaction. For a completely irreversible system involving the reduction of an oxidized species with
the transfer of one electron, the peak current in amperes at 298 K is:

ip;c ¼ 2:99  105 a1=2 A C D1=2 y1=2 (3)

where a is the transfer coefficient and is related to the shape of the energy barrier to electron transfer. As the redox reaction is
completely irreversible, the cyclic voltammogram for this system only contains a cathodic wave.
The difference between the peak potential and the potential at half the peak height is given by:

RT 47:7 
Ep  Ep=2 ¼ 1:847 ¼ mV at 25 C (4)
aF a
For a completely irreversible wave, Ep will shift in a negative (for a reduction reaction) potential direction with increasing
sweep rate.

Cyclic Voltammetry of Surface Confined Species

The properties of electrodes can be dramatically altered by binding electroactive specific species, for example, molecular receptors or
catalysts, to their surfaces, to give rise to a “modified electrode.” These electrodes find widespread application as sensors,
electrocatalysts and in energy applications. Fig. 3 shows the typical cyclic voltammogram for a redox reaction where a monolayer
of redox active material has been deposited on the surface and there is no electroactive material in solution. Both the anodic
and cathodic waves are entirely symmetrical across the potential axis and the peak current for the cathodic and anodic reactions
are given by:
n2 F 2
ip ¼ yAG (5)
4RT
where G is the surface coverage of electroactive species in mol cm 2.
This voltammetric behavior is also observed for thin films (multilayers) of electroactive material at electrode surfaces provided
that the experimental timescale is sufficiently long so as to exhaustively electrolyse all of the deposited material (i.e., slow scan rates).
Under these conditions, finite diffusion, rather than semiinfinite linear diffusion as is the case for reagents in solution, predominates
and the voltammetric response will be similar to that shown in Fig. 3.
212 Voltammetry j Overview

Fig. 3 Cyclic voltammogram for a surface confined, one-electron redox reaction. The electrode area, A, is 0.1 cm2, the surface coverage is
0
1  10 10 mol cm 2, the rate of heterogeneous electron transfer, ko, 1  103 s 1, the formal potential, Eo , is 0.300 V and the scan rate is
1
100 mV s . The initial potential is þ0.700 V.

For an immobilized film of electroactive species, under finite diffusion conditions the peak current is proportional to y, rather
than the y1/2 dependence observed for freely diffusional species. The area under the wave represents the charge passed during elec-
trolysis of the film and is given by:
Q ¼ nFAG (6)
A useful insight into the extent of lateral interactions between adsorbates may be obtained from the full width at half
maximum (FWHM), that is, the total width of the anodic or cathodic wave at half the peak current. Under Langmuir isotherm
conditions (i.e., when there are no interactions between adjacent adsorbates), the FWHM for an ideal Nernstian reaction at 298 K
is given by:
RT 90:6
FWHM ¼ 3:53 ¼ mV (7)
nF n
Deviations from this value can indicate the presence of attractive (< 90.6/n mV) or repulsive (> 90.6/n mV) interactions between
adjacent adsorbates.
The formal potential, E0 , contains extremely useful information about the ease of oxidation of redox species within a surface
confined film. A shift of E0 of an electroactive species, upon immobilization at an electrode surface, towards more positive poten-
tials indicates that it becomes thermodynamically more difficult to oxidize the film. This suggests a reduced electron density on the
redox center when immobilized.

Microelectrodes

One of the key barriers to the more widespread adoption of voltammetric techniques was the limited range of media in which
analysis could traditionally be performed, for example, aqueous solutions containing a relatively high concentration of support-
ing electrolyte. This restriction arose because resistance between the working or sensing electrode and the reference electrode
limited the precision with which the applied potential could be controlled. However, microelectrodes, also commonly known
as ultramicroelectrodes, whose critical dimension is in the micrometer range open up new dimensions of space, time and
medium for electroanalysis. The currents observed at microelectrodes typically lie in the picoamp to nanoamp range, which
are several orders of magnitude smaller than those observed at conventional millimeter dimensioned macroelectrodes. These
small electrolysis currents often completely eliminate ohmic effects. Thus, electrochemical processes can be investigated in
high resistance solvents, in solids, in supercritical fluids, and even in gases. Moreover, they facilitate measurements at short time-
scales and at low temperatures.

Reduced Ohmic Effects


When Faradaic and charging currents flow through a solution, they generate a potential that acts to diminish the applied potential
by an amount iRu, where i is the total current and Ru is the uncompensated cell resistance. This is an undesirable process that leads to
distorted voltammetric responses.
Voltammetry j Overview 213

The uncompensated solution resistance for a disk-shaped microelectrode is inversely proportional to the electrode radius,
1
Ru ¼ (8)
4krO
where k is the conductivity of the solution and rO is the radius of the microdisk. Eq. (8) shows that Ru increases as the electrode radius
decreases. However, the current observed depends on the electrode area (directly proportional to rO2) and at microelectrodes the
currents are typically six orders of magnitude smaller than those observed at macroelectrodes. These small currents often completely
eliminate ohmic drop effects even when working in organic solvents. For example, the steady-state current observed at a 5 mm radius
microdisk is approximately 2 nA for a 1.0 mM solution of ferrocene in acetonitrile. Taking a reasonable value of 0.01 U 1 cm 1 as
the specific conductivity, then Eq. (8) indicates that the resistance will be of the order of 50,000 U. This analysis suggests that the iRu
drop in this organic solvent is a negligible 0.09 mV. In contrast, for a conventional macroelectrode the iRu drop would be of the
order of 5–10 mV. Under these circumstances, distorted current responses and shifted peak potentials are observed in cyclic
voltammetry.

Capacitive Effects
Altering the potential that is applied to an electrode causes the charge on the metal side of the interface to change and some
reorganization of the ions and solvent dipoles in the double layer on the solution side of the interface will occur. This process causes
electrons to flow into or out of the surface, giving rise to a charging or capacitive response. The double layer capacitance for a dis-
k-shaped microelectrode is proportional to the area of the electrode surface and is given by:
Cdl ¼ p rO 2 CO (9)
where CO is the specific double layer capacitance of the electrode. Thus, shrinking the size of the electrode causes the interfacial
capacitance to decrease with decreasing rO2. These low capacitive currents are particularly important for analytical applications of
cyclic voltammetry where the ability to discriminate a Faradaic signal above a charging current often dictates the limit of detection
that can be achieved.

Mass Transport
A distinct advantage of microelectrodes over larger electrodes is that a range of diffusional mass transport regimes can be observed at
timescales that can be easily accessed experimentally. As discussed previously, oxidation or reduction of an electroactive species in
solution at a relatively larger electrode (mm2) using relatively fast scan rates (100 mV s 1) causes a diffusional current to flow that
peaks at Ep, after which depletion of electroactive materials occurs and the current decays. However, the timescale of the experiment
significantly affects the diffusional mass transport regime and significantly influences the response observed. By examining the time
dependence of the diffusional mass transport process it is possible to predict the cyclic voltammetry response and extract
quantitative electroanalytical information.
For example, consider a spherical electrode of radius rO placed in a solution that contains only supporting electrolyte and
a redox-active species at concentration C. The concentration gradient at the electrode surface after applying a potential step to
the electrode is obtained by solving Fick’s second law in spherical coordinates.
 2 
vCðr; t Þ v Cðr; t Þ 2 vCðr; t Þ
¼D þ (10)
vt vr 2 r vr
The boundary conditions for the potential step experiment are,
lim Cðr; t Þ ¼ CN (11)
r/N

Cðr; 0Þ ¼ CN (12)

Cðr0 ; tÞ ¼ 0 (13)
where r is the distance from the center of the sphere, r0 is the radius of the electrode, D is the diffusion coefficient for the redox active
species, and C is the concentration as a function of distance r and time t.
Eq. (10) can be solved using Laplace transform techniques to give the time evolution of the current, i(t), subject to the boundary
conditions described resulting in Eq. (14):

nFADCN nFAD1=2 CN
iðt Þ ¼ þ (14)
rO p1=2 t 1=2
Eq. (14) reveals that the current contains both time-dependent and independent terms. The differences in the electrochemical
responses observed at macroscopic and microscopic electrodes arise because of the relative importance of these terms at
214 Voltammetry j Overview

(A) (B) Spherical


Planar
Diffusion Field
Diffusion Field

Electrode Insulator Electrode Insulator

Fig. 4 Diffusion fields observed at microelectrodes. (A) Linear diffusion observed at short times. (B) Radial diffusion observed at long times.

conventional electrochemical timescales. It is possible to distinguish two limiting regimes depending on whether the experimental
timescale is short or long.
(i) Short times. At sufficiently short times, the thickness of the diffusion layer that is depleted of reactant is significantly
smaller than the electrode radius and the spherical electrode appears to be planar to a molecule at the edge of this diffusion
layer. Under these conditions, the electrode behaves like a macroelectrode and mass transport is dominated by linear diffusion
to the electrode surface as illustrated in Fig. 4A. At short times the first part of Eq. (14) becomes insignificant compared to the
second part due to the t 1/2 dependence of the current. Therefore, the current decays over time in accordance with the Cottrell
equation:

nFAD1=2 CN
iðt Þ ¼ (15)
p1=2 t 1=2
(ii) Long times. At long times (i.e., slow scan rates), the second term in Eq. (14) becomes negligible and the current attains a time-
independent steady-state value given by Eq. (16):

nFADCN
iss ¼ (16)
rO
At these long times, the spherical character of the electrode becomes important and the mass transport process is dominated by
radial (spherical) diffusion as illustrated in Fig. 4B.
Fig. 5A shows the sigmoidal-shaped responses that characterize steady-state mass transfer in slow scan-rate cyclic voltammetry.
In contrast, as illustrated in Fig. 5B, at short experimental timescales (high scan-rates), peaked responses similar to those observed at
conventional macroelectrodes are seen.
The preceding analysis considered a spherical electrode because its surface is uniformly accessible, and a simple closed-form
solution to the diffusion equation exists. The microdisk is the most widely used geometry, but derivation of rigorous expressions
describing their experimental response is complicated because the surface is not uniformly accessible. For disks, electrolysis at the
outer circumference of the disk diminishes the flux of the electroactive material to the center of the electrode. However, microdisk

Fig. 5 Effect of scan rate on the cyclic voltammetry at microelectrodes. The electrode radius is 5 mm, the concentration, C, of the molecule in solu-
tion is 5 mM, the diffusion coefficient, D, is 4  10 6 cm2 s 1, the rate of heterogeneous electron transfer, ko, is 0.1 cm s 1, the formal potential,
0
Eo , is 0.300 V and the initial potential is þ0.700 V. (A) Scan rate is 0.1 V s 1; (B) scan rate is 10 V s 1.
Voltammetry j Overview 215

and microring geometries share the advantage of spherical microelectrodes in that quasi-spherical diffusion fields are established in
relatively short periods of time. The steady-state current is given by:
iss ¼ gnFDCrO (17)
where g is 4 and 2p for disk and hemispherical shaped electrodes, respectively.

Electroanalysis

Voltammetry provides both qualitative and quantitative analytical information about the concentration of redox active species
within samples as diverse as foods, polymers, beverages, drugs, fine chemicals, cosmetics, plating baths, animal feed, raw materials,
water and wastewater. Versatility and simplicity distinguish the technique and the following list highlights some of the main
advantages.

• Sensitivity
• Speed
• Speciation
• Multicomponent capability
• Versatility
• Response immune to colored or turbid samples
• Reliability
A wide range of elements can be analyzed using cyclic voltammetry and this breadth of application is dramatically expanded when
one considers the diverse range of inorganic and organic species that is electroactive. Cyclic voltammetry provides routine analysis at
low levels using instrumentation and consumables that are widely available at low-cost. Also, a wide variety of methods have been
developed that can allow complex mixtures to be analyzed often with a minimum of clean-up. For example, by correlating the
formal potentials of unknown components within mixtures with reference compounds it may be possible to identify unknowns
within complex mixtures. However, this approach is limited since the experimental response for a single analyte takes up a signif-
icant fraction of the available analytical window. While techniques such as differential pulse voltammetry offer significantly greater
sensitivity, voltammetry is capable of measuring concentrations at ppb level levels. Modifying the electrode surface, for example,
with antibodies or nucleic acids capture strands, makes it selective towards a particular target analyte frequently eliminating the
need for a separation step, for example, by HPLC.

Heterogeneous Electron Transfer Dynamics

Beyond an insight into the ease with which compounds can be oxidized or reduced by measuring the formal potential, cyclic vol-
tammetry can be used to determine the rate of electron transfer across the electrode/solution or electrode/film interface. Optimizing
the rate of heterogeneous electron transfer is important for technological applications ranging from the analysis of metals in poly-
mers, foods and cosmetics to the development of biosensors and molecular electronic devices.
Electrolysis (oxidation or reduction) of a freely diffusing electroactive species at an electrode surface during cyclic voltammetry
consists of a number of elementary steps. Initially, diffusion of the redox species to the electrode surface under the influence of
a concentration gradient must occur. Thermal activation of the redox center and electronic coupling of the activated redox center
with the electrode surface must then occur. Finally, the instantaneous elementary electron transfer event occurs. In investigating
heterogeneous electron transfer dynamics, one must ensure that the rate of mass transfer to the electrode surface exceeds the rate
of electron transfer, for example, by stirring the solution, rotating the electrode or by using microelectrodes.
Research into this area is dominated by microelectrodes. At short times, the diffusion layer thickness is much smaller than
the microelectrode radius and the dominant mass transport mechanism is planar diffusion. Under these conditions, the
classical theories, for example, that of Nicholson and Shain, can be used to extract kinetic parameters from the scan rate
dependence of the separation between the anodic and cathodic peak potentials. Using this approach, the standard heteroge-
neous electron transfer rate constant, k , may be determined from the published working curves relating DEp to a kinetic
parameter, J. The variation of DEp with y is determined and, from this, J is calculated. k is then determined by the
following equation:
  1 2   a 2
= =
nF DR
ko ¼ J DO py (18)
RT DO
However, this approach can be limited due to factors such as convolution of Faradaic and charging currents at high scan rates. In
addition, uncompensated Ohmic effects at high scan rates may result in inaccurate measurements of heterogeneous electron transfer
kinetics. One particular approach that can overcome these limitations is to simulate the entire cyclic voltammogram. In addition to
this, positive feedback circuitry can be used to compensate for Ohmic drop. This approach has been used successfully for a wide
216 Voltammetry j Overview

Fig. 6 Background-subtracted experimental (solid line) and simulated (open circles) cyclic voltammograms for a solution of ferrocene (10 mM) in
acetonitrile containing 0.6 M tetra-butyl ammonium perchlorate. The scan rate is 500 kV s 1 and the gold microelectrode radius is 5 mm. Simulations
allow for ohmic drop and RC cell time constant and k is 2 cm s 1.

range of organic and inorganic compounds and Fig. 6 shows an example of a comparison between a simulated and experimental
voltammogram. By simulation of the entire voltammogram, the standard heterogeneous electron transfer rate constant, k , has been
determined as 2 cm s 1 for the ferrocene/ferrocenium redox couple.
An attractive approach to solving the mass transfer limitations of these investigations is to immobilize the electroactive species at
the electrode surface within a monomolecular film. Clearly, if the electroactive species is immobilized at the electrode surface, diffu-
sion of the species to the electrode does not need to occur prior to electron transfer. In addition, immobilization at an electrode
surface can preconcentrate the species of interest, resulting in higher currents that are easier to detect. Electroactive adsorbed mono-
layers have been developed that exhibit close to ideal reversible electrochemical behavior under a wide variety of experimental
conditions of timescale, temperature, solvent and electrolyte. These studies have elucidated the effects of electron transfer distance,
tunneling medium, molecular structure, electric fields and ion pairing on heterogeneous electron transfer dynamics.

Following Chemical (EC) Reactions

Beyond the ability to probe heterogeneous electron transfer reactions which occur on the micro, or even nano timescale, high-speed
cyclic voltammetry has found application in the investigation of complex reaction mechanisms. An example of this is the following
chemical (EC) reaction. Oxidation or reduction (E reaction) of an electroactive species can produce a chemically unstable product
that undergoes a chemical reaction (C reaction) to form an electro-inactive product. Electron transfer reactions can also lead to
isomerizations and ligand loss in organometallic compounds, such as:
CoIII Br 2 en2 þ 6H2 O þ e/CoII ðH2 OÞ6 þ 2Br  þ 2en (19)
1
where en is ethylenediamine. Fig. 7A shows a slow scan rate (0.2 Vs ) cyclic voltammogram for an EC reaction:
A þ e/B (20)

B/C (21)
where the reduced product, B, undergoes a reaction to form C, which is electro-inactive. The ratio of anodic peak current to cathodic
peak current is less than 1. Fig. 7B illustrates the same system while using a sweep rate of 50 Vs 1. The smaller timescale of the
voltammogram in Fig. 7B illustrates the ability of higher sweep rate cyclic voltammetry to outrun the following chemical reaction.

Conclusions and Future Directions

Cyclic voltammetry is arguably the world’s most widely used electrochemical technique and the range of readily available informa-
tion that can be elucidated from the technique has ensured its place in the arsenal of techniques employed by electrochemists. The
ability to extract information regarding electron transfer rates and complex reaction mechanisms is established and available to an
ever-increasing number of chemists.
As electrode areas become ever smaller, it may soon be possible to work with well-defined electrodes that are of similar dimen-
sions to the molecules that are electrolyzed at their surfaces. However, it will remain important that instrumental advances proceed
Voltammetry j Overview 217

Fig. 7 (A) First three sweeps of a voltammogram for an EC reaction at a scan rate of 0.2 V s 1. (B) Voltammogram for the same system as in
(A) at a scan rate of 50 V s 1. The electrode area, A, is 0.1 cm2, the concentration, C, of the molecule in solution is 5 mM, the diffusion coefficient,
0
D, is 4  10 6 cm2 s 1, the rate of heterogeneous electron transfer, ko, is 0.1 cm s 1, the formal potential, Eo , is 0.300 V and the rate of the
1
forward chemical reaction is 30 s . The initial potential is þ0.700 V.

in tandem with our ability to construct electrodes at the nano and subnano level so that the very small Faradaic currents generated,
for example, pA, at very short times, ns, can be accurately measured. Beyond the fundamental research level, cyclic voltammetry has
a promising future as a tool to perform entire voltammetric experiments within the living cell. Cell signaling pathways may be eluci-
dated by not only performing cyclic voltammetry at a nanometer and subnanometer level, but also at extremely short timescales to
monitor transient species that exist for mere nanoseconds. The future for cyclic voltammetry looks as promising as ever before and
will, undoubtedly, provide the tool that electrochemists will need to explore new dimensions in space and time.

Further Reading

Batchelor-McAuley, C.; Kätelhön, E.; Barnes, E. O.; Compton, R. G.; Laborda, E.; Molina, A. Recent Advances in Voltammetry. Chem. Open. 2015, 4 (3), 224–260.
Bard, A. J.; Faulkner, L. R. Electrochemical Methods: Principles and applications, 2nd ed; Wiley, 2001.
Zhou, X. S.; Mao, B. W.; Amatore, C.; Compton, R. G.; Marignier, J. L.; Mostafavi, M. Transient Electrochemistry: Beyond simply temporal resolution. Chem. Comm. 2016, 52 (2),
251–263.
Compton, R. G.; Banks, C. E. Understanding Voltammetry, World Scientific, 2007.
Elgrishi, N.; Rountree, K. J.; McCarthy, B. D.; Rountree, E. S.; Eisenhart, T. T.; Dempsey, J. L. A Practical Beginner’s Guide to Cyclic Voltammetry. J. Chem. Educ. 2018, 95 (2),
197–206.
Marken, F.; Neudeck, A.; Bond, A. M. Cyclic Voltammetry. In Electroanalytical Methods; 2010; pp 57–106.
Amatore, C.; Guille-Collignon, M.; Lemaître, F. Recent Investigations of Single Living Cells with Ultramicroelectrodes. Nanoelectrochem. 2015, 439–468.
Bond, A. M.; Elton, D.; Guo, S. X.; Kennedy, G. F.; Mashkina, E.; Simonov, A. N. An integrated instrumental and theoretical approach to quantitative electrode kinetic studies based
on large amplitude Fourier transformed ac voltammetry: A Mini Review. Electrochem. Commun. 2015, 57, 78–83.
Bond, A. M.; Duffy, N. W.; Guo, S. X.; Zhang, J.; Elton, D. Changing the Look of Voltammetry. Anal. Chem. 2005, 77 (9), 186 A–195 A.
Forster, R. J.; Vos, J. G.; Keyes, T. E. Interfacial Supramolecular Assemblies, Wiley, 2003.
Bond, A. M. Broadening Electrochemical Horizons: Principles and Illustration of Voltammetric and Related Techniques, Oxford, 2002.
Pletcher, D. Instrumental Methods in Electrochemistry, Albion/Horwood, 2002, 2002.

You might also like