Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/328674174

Square-wave voltammetry

Article  in  ChemTexts · November 2018


DOI: 10.1007/s40828-018-0073-0

CITATIONS READS

11 7,450

3 authors, including:

Valentin Mirceski Leon Stojanov


Ss. Cyril and Methodius University Ss. Cyril and Methodius University in Skopje
138 PUBLICATIONS   2,784 CITATIONS    7 PUBLICATIONS   18 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

MSc thesis research (Sasho Stojkovikj) View project

All content following this page was uploaded by Valentin Mirceski on 15 November 2018.

The user has requested enhancement of the downloaded file.


ChemTexts (2018) 4:17
https://doi.org/10.1007/s40828-018-0073-0

LECTURE TEXT

Square-wave voltammetry
Valentin Mirceski1,2 · Sławomira Skrzypek1 · Leon Stojanov2

Received: 30 August 2018 / Accepted: 15 October 2018


© Springer Nature Switzerland AG 2018

Abstract
This text is written as a very basic, first introduction to square-wave voltammetry, as one of the very specific, but most
versatile techniques in the family of pulse voltammetric techniques. To make the text self-consistent, a brief introduction to
voltammetry is initially given, to make the next elaboration of square-wave voltammetry more easily understandable. For
more comprehensive introduction to voltammetric techniques, the previous chem-text is recommended (Scholz in ChemTexts
1:17, 2015). The aim of the current text is to explain the basic features of square-wave voltammetry, to reveal the essential
idea behind the specific shape of the excitation signal used, as well as to demonstrate that square-wave voltammetry is highly
suited to study the mechanism of electrode reactions. The latter is illustrated by addressing the simplest experimental situa-
tion encountered in voltammetry to make the text didactically suited for broader student population studying chemistry and
related sciences.

Keywords  Square-wave voltammetry · Pulse voltammetry · Mechanism of electrode reactions · Chemistry

Voltammetry: a brief overview specific cases, electrodes are used with dimensions ranging
from the micrometre to nanometre scale [3]. As redox reac-
Voltmammetry is based on microelectrolysis tions of the studied electroactive species take place at the
electrode|electrolyte interface only, minute amounts of the
Voltammetry encompasses a large group of electrochemi- electroactive reactant are consumed and the overall chemi-
cal techniques where nonspontaneous, interfacial charge cal composition of the system remains practically constant.
transfer processes (electron or ion transfer across the The voltammetric measurement is organised in a spe-
electrode|electrolyte solution interface) are driven by exter- cific way to provide information on the phenomena taking
nally applied electrical potential difference in an electrolytic place at one of the electrodes only, which is termed working
cell [1, 2]. The aim of voltammetry frequently is charac- electrode. A scheme of a typical cell is depicted in Fig. 1,
terisation of redox species and the analytical determina- consisting of the working electrode (WE), with the smallest
tion. Additionally, the thermodynamics and kinetics of the electrode surface area, the second electrode, called reference
charge transfer can be studied and possibly coupled chemical electrode (RE), and the third, the counter electrode (CE). A
reactions or adsorption phenomena. Voltammetric experi- current is flowing only between the working and the coun-
ments are frequently conducted in a few millilitres of an ter electrodes. The working electrode is the bottleneck for
electrolyte solution, using small electrodes with a surface the current flow; thus it carries information solely from the
area less than a square centimetre; alternatively, in more charge transfer across the small working electrode|electrolyte
interface. Electrochemical reactions at the counter electrode
are usually unknown, and finally irrelevant for the outcome
* Valentin Mirceski of the measurement.
valentin@pmf.ukim.mk The reference electrode should be robust, with a constant
1
chemical composition, which determines its constant electri-
Department of Inorganic and Analytical Chemistry, Faculty
of Chemistry, University of Lodz, Tamka 12, 90‑403 Lodz,
cal potential, [4]. Its role is to serve as a reference point to
Poland control the relative potential of the working electrode with
2
Institute of Chemistry, Faculty of Natural Sciences
the aid of an instrument called potentiostat. The electrical
and Mathematics, P.O. Box 162, 1000 Skopje, potential difference between the solution and the working
Republic of Macedonia

13
Vol.:(0123456789)
17   Page 2 of 14 ChemTexts (2018) 4:17

plus the suffix “metry” with a meaning of measuring; hence,


POTENTIOSTAT volt + am + metry ≡ voltammetry.

The simple shape of the voltammogram reflects


a variety of interfacial phenomena

A graphical representation of the outcome of a voltammetric


experiment is called voltammogram (Fig. 2a). It is an I–E
curve, representing the variation of the current I versus the
controlled potential E of the working electrode, relative to
WE the constant potential of the reference electrode. In Fig. 2a,
the simplest form of a voltammetric experiment has been
assumed. The potential is varied linearly in time within a
certain potential interval around the formal potential ( E∅  )

(Fig. 2b) [3]. This sort of voltammetry is known as voltam-


metry with a linear potential scan or linear sweep voltamme-
RE try (LSV). The main parameter of the technique is the rate of
CE Red Ox the potential variation with time (sweep rate, or scan rate, v).
The sweep rate is defined as a ratio of the potential interval
electrolyte solution (sol) ΔE (ΔE = Ef − Es, where Ef and Es is the final and starting
potential, respectively) and the time interval Δt (i.e., the time
required to cross the potential path ΔE), thus v = ΔE/Δt.
Fig. 1  Scheme of a voltammetric cell where the interfacial charge Obviously, v determines the overall time of the experiment,
transfer is a redox reaction at the working electrode|electrolyte inter- which could range from minutes to milliseconds.
face, and the current flows between the small working electrode (WE)
In Fig. 2, it is further assumed that the experiment is
and the large counter electrode (CE). The potential of the working
electrode is controlled versus the constant potential of the robust ref- conducted at a static working electrode (e.g., chemically
erence electrode (RE) with the aid of the potentiostat  inert platinum electrode with a surface area in the range of
square centimetres, i.e., macroscopic electrode). The elec-
trode is immersed in an unstirred solution. Most frequently,
electrode is the driving force for the electrochemical reaction water, or some other polar organic liquid is the solvent. The
and the resulting flow. solvent contains dissolved electroactive compound [e.g.,
Most potentiostats for 3-electrode measurements are hexacyanoferrate(II) ions] typically in millimolar concentra-
working this way that the potential of the electro- tion and a supporting electrolyte (e.g., K
­ NO3), commonly at
lyte solution is equal to the potential of the metallic a concentration at least one order of magnitude higher. The
phase of the reference electrode. The potential differ- theoretically predicted voltammogram has a simple, asym-
ence between the metallic phase of the working elec- metric peak shape (Fig. 2a).
trode and the electrolyte solution is thus equal to that Although the voltammetric curves may look very sim-
between the working and the reference electrode. The ple, a variety of physicochemical processes (e.g., diffu-
voltage between the working and the counter electrode sion, migration, convection) and chemical reactions (redox
is adjusted by the potentiostat so as to have the desired reactions and accompanying chemical reactions) can be
potential difference between the working electrode involved. This can make a correct interpretation of the
and electrolyte solution (= reference electrode). For data quite challenging. The simplest possible voltammetric
voltammetric experiments, i.e., where the potential is experiment envisaged in Fig. 2 is at least affected by intrinsic
a certain function of time, the potentiostat has to be thermodynamic features of the interfacial redox reaction (1)
interfaced with a potential wave form generator. (e.g., Gibbs free energy) and the mass transfer phenomenon
known as diffusion [3]:
The foregoing discussion implicitly follows that in vol-
tammetry one plays with the physical parameters such as Redsol ⇄ Oxsol + ne− . (1)
electrical potential and current. The former is an independ- For the sake of simplicity, the charges of the species
ent variable, controlled by the instrumentation, whereas the are omitted and the abbreviation (sol) indicates dissolved
latter is the dependent one (i.e., an outcome of the experi- species, whereas n is the stoichiometric number of elec-
ment). Hence, the term “voltammetry” is coined from abbre- trons. The redox reaction (1) takes place at the working
viations referring to the units, i.e., volt (V) and ampere (A), electrode|electrolyte interface only. It is heterogeneous

13
ChemTexts (2018) 4:17 Page 3 of 14  17

Ep
Ef

potential
current

Ip E

Es
potential time
(a) (b)

Fig. 2  a Typical voltammogram obtained at a macroscopic work- rate. Such technique is called linear sweep voltammetry. The maxi-
ing electrode by varying the electrode potential in the simplest lin- mum of the asymmetric peak-shaped voltammogram, cf. a, is defined
ear fashion from certain starting (Es) to the final (Ef) potential. b The by the peak-current (Ip) and the peak-potential (Ep)
time variation of the potential at a certain rate v, known as the scan

interfacial chemical process, as the two participants ­Oxsol concentrations at the interface, which differ from the cor-
and ­Redsol reside in the electrolyte solution. The third par- responding bulk concentrations. The concentrations at the
ticipant, electrons, resides in a different phase, i.e., the interface are termed as surface concentrations, cs,Red and
working electrode. Electrons are exchanged between the cs,Ox , respectively. As R­ edsol is consumed and O ­ xsol is pro-
electrode and the redox species across the interface [3]. duced, cs,Red < cb,Red , cs,Ox > cb,Ox , respectively. Hence, con-
Thus, the redox reaction (1) is referred to electrode reac- centrations of redox species vary with time and distance. In
tion, being different to common homogeneous redox reac- mathematical terms, they are a function of the two variables,
tions between two redox couples in a single phase (e.g., distance (x), measured from the electrode surface, and time
­Red1sol + Ox2sol ⇄ Ox1sol + Red2sol) [3]. (t), measured from the beginning of the experiment, [written
In this context, we can recall again the physical meaning as c(x,t)]. Hence, a surface concentration can be also noted
of the electrical potential of the working electrode, which as c(0, t).
can be alternatively understood as a measure of the activity Concentration gradients established in the vicinity of the
of electrons in the electrode (1). Therefore, one can intui- electrode surface prompt the natural mass transfer known as
tively assume that variation of the potential enables shifting diffusion [3]. The diffusion of the species frequently affects
of the redox equilibrium (1) in a desirable direction, i.e., to the voltammetric experiment and the shape of the voltam-
the right-hand side (oxidation), or to the opposite side, i.e., metric curve to a large extend. Due to diffusion, the maxi-
the reduction. mum of the voltammetric curve, known as peak-current (Ip),
At the beginning of the experiment, it is assumed that is a linear function of the bulk concentration of the initial
only ­Redsol species are homogeneously distributed in the electroactive reactant:
electrolyte solution at a bulk concentration designated as
cb,Red.1 Accordingly, Ox is absent at the beginning of the
Ip = k × cb,Red , (2)
experiment ( cb,Ox = 0 ). In the course of the experiment, where k is the constant for a given experimental conditions.
the potential of the working electrode is progressively The linearity between the peak-current and the bulk concen-
increased commencing from a lower toward higher values tration provides a basis for analytical application of voltam-
that is E∅  . Consequently, the electrode reaction (1) is pro-

metry [1].
gressively shifted toward the right-hand side, causing the Beyond the maximum, the decaying tail of the curve is
current to flow. Both Red and Ox species establish certain also a consequence of diffusion, causing the overall voltam-
mogram to gain an asymmetric peak-like shape. Besides
the peak-current, the curve is additionally attributed with
1
  For simplicity reasons concentration instead of activity. another important parameter known as peak-potential (Ep).

13
17   Page 4 of 14 ChemTexts (2018) 4:17

Ip,a and Ep,a


Ef > E
potential

current
E

potential

Es < E
time Ip,c and Ep,c

(a) (b)

Fig. 3  a Typical potential variation and b the shape of the I–E curve in the popular technique known as cyclic voltammetry 

It is the position of the maximum at the potential axis. The (CV), with a typical voltammogram shown in Fig. 3b. Obvi-
latter is related to the formal potential ( E∅  ), which is ther- ously, the first half of the cyclic voltammetry is identical as

modynamic parameter of a critical importance for the elec- in Fig. 2, while in the second half, the reaction is drifted in
trode reaction (1) [3]. the opposite direction (reduction of O ­ xsol to R
­ edsol) with the
At the end of this section, it has to be mentioned that apart aid of the reverse potential sweep (cf. Fig. 3a). The corre-
from the foregoing discussed, the outcome of the experiment sponding cyclic voltammogram (Fig. 3b) is attributed with
is frequently additionally affected by the kinetics of the elec- forward and reverse components, reflecting the oxidative and
trode reaction, a variety of surface phenomena, chemical reductive shifts of the electrode equilibrium, respectively. By
reactions coupled to the electrode reactions, unavoidable convention, the oxidation (termed as anodic) and reduction
electrical phenomena related to the formation of electric (termed as cathodic) currents are assigned with a positive
double layer at the electrode|electrolyte interface, etc. and negative sign, respectively.
The main advantage of cyclic voltammetry is its inherent
The overwhelming beauty of a cyclic voltammogram ability to drive the electrode system (1) in both anodic and
cathodic directions. Any additional physicochemical pro-
At the end of the experiment described in Fig. 2, the elec- cesses involving redox species (e.g., adsorption, homogene-
trode reaction (1) is strongly drifted toward the right-hand ous chemical reactions, disproportionation, etc.), could be
side, resulting in formation of ­Oxsol species in the vicinity inspected with cyclic voltammetry. Thus, CV is the main
of the electrode. The initial reactant, R ­ edsol, is exhausted tool in hand to study electrode mechanisms, making the
from the electrode region, while the diffusion is trying to technique by far the most popular among electrochemists
resupply it from the bulk. Thus, the interfacial region where and chemistry-related community.
the diffusion mass transfer exchanges the material with the Cyclic voltammogram (cf. Fig. 3b) bares overwhelming
bulk is termed as a diffusion layer. The decaying tail of the aesthetic features, reflecting the chemical drift of a given
voltammogram (cf. Fig. 2a) reflects the mass transfer in the equilibrium in a perfectly precise and controllable way. The
expanding diffusion layer. In proportion to the time of the latter could be hardly achieved by any other technique in
experiment, the region becomes richer with O ­ xsol and poorer chemistry. Again, the main instrumental parameter is the
with ­Redsol. scan rate. The key features of the voltammogram are the
Considering the foregoing facts, one can intuitively pre- peak-current and peak-potential of both forward (anodic;
dict that the electrode reaction (1) can be drifted in opposite, Ip,a and Ep,a) and backward (cathodic; Ip,c and Ep,c) compo-
left-hand side (reduction), if the experiment continues with nents (cf. Fig. 3b). By careful inspection of the interrelation
oppositely oriented potential sweep (Fig. 3a). Such a sort of of these parameters as a function of the scan rate, one can
voltammetric experiment is known as cyclic voltammetry deduce information regarding thermodynamics and kinetics

13
ChemTexts (2018) 4:17 Page 5 of 14  17

double layer diffuse layer

solvent molecule
-
-
-
+ - anion

+
-
-
-
+ cation

+
-
-
- +
electrode

- +
- +
- - If
-
current
+
- -
- + Ic
-
+
- me

Fig. 4  A primitive sketch of the electric double layer at the electrode|electrolyte interface and the assumed variation of the charging (Ic) and fara-
daic (If) current in time (inset)

of the electrode reaction, the mechanism of interfacial elec- A microscopic double layer is formed due to an excess
tron transfer, adsorption of redox species, the role of coupled of charge at the electrode surface and the corresponding
chemical reactions, etc. response (arrangement) of the counter, free moving charged
particles at the side of the electrolyte solution (i.e., elec-
A current flows even in the absence of an electrode trolyte ions). In addition, polar solvent molecules are also
reaction accordingly orientated at the interface (Fig. 4). The formed
microscopic double layer acts as a capacitor with a signifi-
Unfortunately, the electrode processes involving charge cant capacitance due to extremely small, microscopic dis-
transfer [e.g., reaction (1)], termed as faradaic processes, tance between the two oppositely charged layers. Thus, for-
are not the sole source of current in voltammetry. There mation of the double layer (i.e., charging of the interfacial
are interfacial phenomena which do not involve an inter- capacitor) is associated with the current flow. Such current is
facial charge transfer, yet giving rise to the current. Such termed as charging current (Ic). The charging current inter-
phenomena are consequently termed as non-faradaic pro- feres with the faradaic one, because both are the additive
cesses. They are primarily related to the formation of an constituents of the overall current. The charging current can
electric double layer at the electrode|electrolyte interface, be even significantly higher than the faradaic one, when the
which is inevitably present at any given electrode potential. concentration of the electroactive reactant is very low (e.g.,

13
17   Page 6 of 14 ChemTexts (2018) 4:17

Ef

E
potenal

∆E
Es
(a)
(b)
me
time

(∆Ip, Ep)
sampling period (Ifor)

Inet
∆E Ifor

Esw

current
τ
current

∆E
tp Esw

(c) (d) Irev (e)


sampling period (Irev)

time potential
time

Fig. 5  a Potential wave-form in cyclic staircase voltammetry. The square-wave voltammetry; c single potential cycle in square-wave
inset shows the duration of the potential step (τ) and the scan incre- voltammetry; d variation of the current in time in the course of the
ment of the staircase potential (ΔE); b the potential modulation in square-wave voltammetry; e typical square-wave voltammogram

micromolar bulk concentration). Thus, the charging current diminish the contribution of the charging current. This can
is the limiting factor in terms of the sensitivity of voltam- be achieved by replacing the continuous potential ramp (cf.
metry in general [1]. Fig. 3a) with a staircase potential-time function in cyclic
Fortunately, for a given small potential step or potential voltammetry (Fig. 5a). Measuring the current at the end of
increment (a few mV), the formation of the double layer each potential step diminishes the charging current con-
typically is completed in a few milliseconds (or less) and tribution significantly, thus improving the sensitivity and
the charging current (Ic) drops to negligible small values quality of the voltammetric data profoundly. This sort of
after such time interval, as depicted in the inset of Fig. 4. technique should be precisely termed as current sampled
On the other hand, faradaic current (If) is still significant cyclic staircase voltammetry (CSCSV). The main features
after certain time, as envisaged in the inset. Though both of the technique are the scan increment ΔE (cf. Fig. 5a) and
current components decay with time, the ratio If/Ic increases the duration of each potential step τ. Obviously, the scan
for a given potential, which is particularly important in the rate is defined as the ratio of the scan increment and the step
context of the following section. duration, v = ΔE∕𝜏 .
Voltammetric experiment can be conducted even in more
advanced fashion, if the staircase potential function is com-
Square‑wave voltammetry bined with small potential pulses, as depicted in Fig. 5b. The
latter potential shape is typical for square-wave voltammetry,
Stepwise potential variation is beneficial which is considered to be one of the most advanced forms
in voltammetry of voltammetric techniques [5]. To understand the overall
potential modulation, let us consider the potential variation
From the foregoing discussion, one can plausibly assume during a single potential step (Fig. 5c). Indeed, at each
that the sensitivity of voltammetry could be significantly potential step, two small, oppositely oriented potential
improved by performing the experiment in the way to pulses are imposed. The duration of the two pulses is

13
ChemTexts (2018) 4:17 Page 7 of 14  17

identical, designated as tp = 𝜏∕2 , where τ, as previously components, which is another important feature. Only in
defined, is the duration of the potential step. The height of a very specific cases, of completely irreversible electrode pro-
single pulse is termed as square-wave amplitude (Esw). Thus, cesses, the net SW peak can be smaller than the forward one
the main parameters of the SW potential modulation are the (see Fig. 8, in the following “Quasireversible electrode reac-
scan increment ΔE of the staircase ramp, the SW amplitude tion”). Finally, the contribution of the charging current to
(Esw) and the duration of the potential pulse tp. Often, the the net current may be even less significant than in particu-
later parameter is expressed in terms of the square-wave lar forward and reverse components, for most of the cases.
frequency f = 2t1  . We can see latter, it is the critical time The reason is that the charging current measured at the end
of the pulse is generally small. When small amplitudes are
p

parameter of the SW voltammetric experiment. By analogy used (nEsw ≤ 50 mV; n—is the stoichiometric number of
to cyclic staircase voltammetry, the SW potential wave-form electrons), the small charging current at two neighbouring
can be attributed with scan rate defined as the product of the pulses is identical in the sign and vary close in the mag-
frequency and the scan increment, v = f ΔE. nitude. Thus, the subtraction procedure in calculating Inet
The two neighbouring potential pulses imposed at a sin- diminishes further the contribution of the charging current,
gle potential step complete a potential cycle in SWV (cf. making SWV one of the most sensitive voltammetric tech-
Fig. 5c). The first and the second pulse are termed as for- niques [1, 5–7].
ward and reverse pulse, respectively. The idea behind impos- Typical SW voltammogram encompasses three I–E
ing two oppositely oriented potential pulses is to drive the curves: two experimentally measured (forward and reverse),
electrode reaction (1) in both anodic and cathodic direc- and the third calculated net current (Fig.  5e). Thus, to
tions for the purpose of gaining mechanistic information; each potential of the staircase ramp, three current values
such procedure is repeated at each potential step, resulting are associated. Each voltammetric curve is discontinuous,
in the overall potential modulation shown in Fig. 5b. One and the point number depends on the scan increment. One
can inspect the mechanistic aspects of the electrode reactions should be aware that the forward and reverse currents are
at the potential value corresponding to each potential step, measured at the real potential of the corresponding pulses,
starting from the Es to the final Ef potential, thus avoiding Efor = Estep + Esw, and Erev = Estep − Esw, for the forward and
the necessity to reverse the overall potential modulation, as reverse pulses, respectively; they are, however, plotted ver-
in CV. sus the step potential Estep. Hence, the significant effect
The current variation in the course of the SW voltam- of the Esw to the features of the response could be even
metric experiment for the reaction (1) is depicted in Fig. 5d. intuitively predicted. The overall shape of the forward and
During each potential cycle, one measures corresponding reverse components resembles the cyclic voltammogram,
anodic and cathodic currents that reflect the shift of the and intrinsically they provide analogous mechanistic infor-
electrode redox equilibrium in anodic and cathodic direc- mation as CV does. For all these reasons, one claims that
tion, respectively. For the sake of improved sensitivity the SWV unifies the advantages of cyclic voltammetry, as a tool
current is sampled at the end of each potential pulse only for mechanistic studies, with the advanced sensitivity, typi-
(cf. Fig. 5c). The series of current values measured at the cal for pulse voltammetric techniques [8].
end of each forward and reverse pulse constitute the forward
(Ifor) and reverse components (Irev) of the SW voltammetric Reversible electrode reaction
response, respectively (Fig. 5e). Obviously, at each potential
step two current values are measured. Moreover, another One of the simplest electrode mechanisms in electrochemis-
voltammetric curve can be constructed by subtracting the try is the so-called reversible electrode reaction of a solution
corresponding values of the forward and reverse current val- resident redox species, as described with Eq. (1) in “The
ues, termed as net current, Inet = Ifor − Irev. simple shape of the voltammogram reflects a variety of
Several justifications could be provided for such sort of interfacial phenomena”. Besides all previous suppositions,
simple mathematical transformation of the experimental let us further simplify the system by assuming that the stoi-
data. For instance, in most of the cases, the net SW voltam- chiometric number of electrons is n = 1, thus a one-electron
mogram is a well-defined peak, whereas the experimental reversible electrode reaction is considered.
forward and reverse components have a wave-like shape, The term “reversible” in the context of general chemistry
or in some cases could have a more complicate shape. The most frequently means a two-directional chemical reaction
peak-like shape of net SW component is advantageous in system (A ⇄ B). The reaction system A ⇄ B can reach the
analytical context as it provides readily measurable param- state of chemical equilibrium, which is dynamic state, where
eters ‘peak height’ and ‘peak position’, thus improving the rate of the forward (A → B) and reverse (B → A) chemi-
the resolution of the technique. The intensity of the net cal reactions are equal. In the thermodynamics, “reversible”
response is higher compared to the forward and backward refers to any process which goes from an initial to a final

13
17   Page 8 of 14 ChemTexts (2018) 4:17

state at an infinitesimally slow rate (thus requiring an infinite while cOx (x, 0) = 0.2 When the experiments start (t > 0), the
time), being considered to be permanently in the state of electrode reaction (1) takes place at the initial potential Es
equilibrium. In voltammetry, reversible electrode reaction is (cf. Fig. 5b), giving rise to surface concentrations cRed (0, t)
the one in which the redox species at the interface reach the and cOx (0, t) interconnected through the Nernst Eq. (3). As
equilibrium concentrations at any potential. In other words, Red is consumed ( cRed (0, t) < cb, Red while Ox is formed
at the interface, the Nernst equation holds for any poten- ( cOx (0, t) > cb, Ox ), Red diffuses from the bulk toward the
tial of the working electrode. The latter equation, written in electrode surface, while Ox diffuses in the opposite direc-
terms of surface concentrations, is as follows: tion. The rate of diffusion depends on the concentration
c (0, t) gradients established in the vicinity of the electrode and the
� RT
E(t) = E� + ln Ox , (3) diffusion rate constant, i.e., the diffusion coefficient D. The
F cRed (0, t)
latter is commonly expressed in units of ­cm2 s−1, and the
where E∅ is the formal potential related to the standard

typical value is in order of ­10−6 cm2 s−1, for most of spe-
potential of the redox couple and the activity coefficients of cies in an aqueous electrolyte. Fortunately, the assumption
redox species. Alternatively, Nernst equation can be writ- that both redox species have the same diffusion coefficient
ten as is frequently close to the reality.
After complex mathematical modelling and computer
cOx (0, t) = cRed (0, t) exp (𝜑), (4) simulations, the response of the one-electron reversible elec-
where trode reaction under conditions of SWV can be expressed as

𝜑=
F ( �)
E − E� , (5)
I = AΨ(𝜑), (6)
RT where A could be termed as an amperometric constant
where 𝜑 is the dimensionless potential, which can be under- defined as
stood as a driving force of the interfacial electron transfer √
process. As mentioned in “The simple shape of the voltam- A = FScb, Red Df , (7)
mogram reflects a variety of interfacial phenomena”, the
where S is the electrode surface area, D is the common dif-
notation c(0,t) means the surface concentration. From the
fusion coefficient, and f is the frequency. Ψ(𝜑) is the dimen-
Nernst equation, we learn that the variation of the electrode
sionless function, which could be revealed through math-
potential E(t), as an independent, instrumentally controllable
c (0,t) ematical modelling and simulations only. It represents the
variable, causes variation of the concentration ratio c Ox (0,t) .
Red effect of the potential wave-form and the diffusion of redox
It means a shift of the redox equilibrium (1) at the electrode species in the course of the experiment. Thus, it can be intui-
surface manifested as a flow of electric current. tively assumed that Ψ(𝜑) depends on the amplitude (Esw)
The Ner nst equation (3) holds only at the and scan increment (ΔE), as typical features of the poten-
electrode|electrolyte interface and the overall system in the tial modulation. Most importantly, Ψ(𝜑) does not depend of
voltammetric cell is not at equilibrium. As explained in the frequency and concentration of the redox species for a
“The simple shape of the voltammogram reflects a variety reversible electrode reaction. It means that the real current
of interfacial phenomena”, the electrode reaction (1) makes I, including the net peak-current ΔIp (cf. Fig. 5e), is a linear
the surface concentrations different than in the bulk, thus function of the square-root of the frequency, which is the
prompting spontaneous mass transfer by diffusion in the first diagnostic criterion of a reversible electrode reaction
vicinity of the electrode. Consequently, the reversible elec- of a dissolved redox couple. In the reality, one varies the
trode reaction in voltammetry is affected by both potential frequency typically from the minimal instrumental value
of the working electrode and the diffusion of redox species. [e.g., f = 5 Hz (tp = 100 ms) to 100 Hz (tp = 5 ms)]. Beyond
In the interplay between the electrode potential and the mass the latter interval, deviations from linearity frequently occur
transfer, the latter is the limiting step. for most of the experimental systems, the reason of which is
To establish a theoretical background, besides Nernst discussed in next section.
Eq. (3), Fick’s law of diffusion has to be considered [3]. Understanding the voltammetric behaviour of particu-
Unfortunately, diffusion is described by a complex math- lar electrode mechanism commonly requires the study of
ematics, which is frequently beyond the common educa- the dimensionless function. The evolution of Ψ(𝜑) for dif-
tion in chemistry. Nevertheless, we are able to describe the ferent SW amplitudes is depicted in Fig. 6. It consists of
experiment qualitatively in simple mathematical terms as
follows. At the beginning of the experiment, t = 0, only Red
is present in the system, being equally distributed in the elec- 2
  Strictly, there is always a certain concentration of Ox existing in a
trolyte solution (x ≥ 0) at a concentration cRed (x, 0) = cb, Red , reversible equilibrium; however, in analytical context, its concentra-
tion is negligible to Red.

13
ChemTexts (2018) 4:17 Page 9 of 14  17

1.1 Ψ
1.1
Ψ

0.7
0.7

0.3
0.3 (b)
(a)
E vs. E E vs. E

-0.4 -0.2 -0.1 0 0.2 0.4


-0.4 -0.2 -0.1 0 0.2 0.4

-0.5
-0.5

1.1 1.1
Ψ
Ψ

0.7 0.7

0.3 0.3
(c) (d)
E vs. E E vs. E

-0.4 -0.2 -0.1 0 0.2 0.4 -0.4 -0.2 -0.1 0 0.2 0.4

-0.5 -0.5

Fig. 6  Evolution of the dimensionless function Ψ(𝜑) for a one-electron reversible electrode reaction calculated for the SW amplitude of 25 (a);
50 (b); 75 (c) and 100 mV (d). The scan increment is ΔE = 5 mV

well-developed forward and reverse current components, and of the net peak is 120 mV. For Esw < 50 mV, the forward
the net peak with the peak-potential identical to the formal (anodic) component is shifted toward more positive poten-
potential of the electrode reaction (1). Taking into account tials than the reverse (cathodic) one, while for Esw > 50 mV,
Eq. (7) and recalling that Ψ(𝜑) is independent of the fre- the interposition is reversed. At the same time, by enhanc-
quency, the real net peak has the peak-potential identical to ing the Esw the intensity of the overall response increases,
the formal potential of the electrode reaction, independent of together with the half-peak width. However, similar effect on
both SW amplitude and frequency, which is the second diag- the morphological evolution of the voltammetric response
nostic criterion of the reversible electrode reaction in SWV. is observed for many other mechanisms [5], implying that
More detailed analysis of the morphological evolution the influence of the amplitude cannot be used as a selective
of the response under the influence of the amplitude is rel- diagnostic criterion for characterising reversible electrode
atively complex (cf. Fig. 6). For the typical amplitude of reaction.
Esw = 50 mV (Fig. 6b), the forward and reverse components Scan increment affects the scan rate of the overall wave-
have the same peak potentials, equal to the formal potential, form, thus the time of the experiment increases by decreasing
while their peak-currents are different. The half-peak width the scan increment. Commonly, the scan increment is varied

13
17   Page 10 of 14 ChemTexts (2018) 4:17

0.8 0.8 0.8


Ψ Ψ Ψ
0.6 0.6 0.6

0.4 0.4 0.4

0.2 (a) 0.2 (b) 0.2


(c)
E vs. E E vs. E E vs. E
0 0 0
-0.3 -0.1 0.1 0.3 -0.3 -0.1 0.1 -0.3
0.3 -0.1 0.1 0.3
-0.2 -0.2 -0.2

-0.4 -0.4 -0.4

Fig. 7  Evolution of the dimensionless function Ψ(𝜑) for a one-electron reversible electrode reaction calculated for the amplitude of Esw = 50 mV
and the scan increment of 0.5 (a); 5 (b), and 10 mV (c)

over the interval from 1 to 10 mV. The net peak-potential Here, 𝜑 is the dimensionless electrode potential, i.e., the
is independent on ΔE, while the net peak-current slightly driving force of the interfacial electron transfer, being iden-
decreases by increasing ΔE. Figure 7 depicts the morpho- tical as in the context of the reversible electrode reaction
logical variation Ψ(𝜑) for very low (ΔE = 0.5 mV), moder- (Eqs. 3–5). In the case of the reversible electrode process,
ate (ΔE = 5 mV), and a large value (ΔE = 10 mV) of the scan the driving force 𝜑 dictates instantaneous redox equilibrium
increment, revealing a marked influence to both forward and at the electrode surface. For the quasireversible reaction, the
reverse components. For very low scan increment, when the same driving force determines the rate to reach the equilib-
overall time of electrolysis is long, the thickness of the diffu- rium, which is not an instant process, but requires certain
sion layer significantly increases, making the rate of diffusion time. Of course, if the rate is low and the time is short, the
very slow at the end of the experiment. As a consequence redox equilibrium is not established, which is typical situa-
both forward and reverse current components (Fig. 7a) gain tion for the quasireversible electrode reaction.
a peak-like shape, contrary to the wave-like shape, associated In the particular case when the electrode potential is
to the diffusion tail for the large scan increment and the fast E = E�  , the driving force is 𝜑 = 0 ; thus exp (0) = 1 , and

scan rate (Fig. 7c). both rate constants are equal, ka = kc = k�  . The latter param-

eter, k∅  , known as the formal rate constant, is an intrinsic


Quasireversible electrode reaction kinetic parameter of electrode reactions [9]; k∅ has dimen-

sions of length per time, due to the heterogeneous nature of


In the experimental reality reversible electrode reactions are the electron transfer occurring across the interface.
seldom encountered. Often, the electrode reaction is affected Several explanations can be offered to understand the
by the kinetics of the interfacial electron transfer; such elec- meaning of the formal rate constant. For instance, at a given
trode reaction is termed as quasireversible. Let us keep con- driving force 𝜑 , an electrode reaction attributed with small
sidering a one-electron electrode reaction (1), assuming that k∅ requires longer time to reach the equilibrium compared

the oxidation (anodic; Red → Ox + e) and reduction (cathodic; to a reaction attributed with larger k∅  . Alternatively, one can

Ox + e → Red) are attributed with anodic and cathodic elec- say that an electrode reaction with small k∅ requires larger

trochemical rate constant ka and kc , respectively. The elec- driving force 𝜑 than a reaction with larger k∅ to gain an iden-

trochemical rate constants depend on the electrode potential. tical rate. A useful concept is to consider an electrode reac-
Hence, they are different than the rate constants of a common tion in equilibrium at the formal potential E = E�  . From the

chemical reaction [3, 9]. The exact mathematical formulations Nernst Eq. (3) the surface concentrations of redox species
are as follows: are equal, cOx (0, t) = cRed (0, t) = c(0, t) . However, the equi-
librium is dynamic process, and both anodic and cathodic
reactions must proceed at equal rate, which fortunately is
�[ ]𝛽
ka = k� exp (𝜑) , (8)
true, considering Eqs. (8) and (9), i.e.,

va = vc = v = k� c(0, t). (10)
�[ ]𝛼
kc = k� exp (−𝜑) , (9)

13
ChemTexts (2018) 4:17 Page 11 of 14  17

Yet, there is a significant discrepancy between two reactions which is of critical importance for understanding voltam-
attributed with different formal rate constants in the state of metric behaviour [5].
equilibrium. From Eq. (10), an electrode reaction with small First of all, let us note that 𝜅 unifies the effect of the
k∅ has a small exchange rate between Ox and Red under equi- electron exchange rate ( k∅  ), the rate of mass transfer (D),
′ ′

librium conditions, compared to another electrode reaction and the critical time of the voltammetric experiment (f).
attributed with larger k∅  , where the turnover between Ox and Figure 8 shows sigmoid variation of the net peak-current

Red is faster. with the logarithm of the dimensionless kinetic parameter


In Eqs. (8) and (9), exponents α and β are dimension- κ. At first glance, Fig. 8 should be understood as a com-
less numbers related as α + β = 1, known as cathodic and parison of a series of electrode reactions attributed with
anodic electron transfer coefficients, respectively [4, 9]. For different formal rate constants k∅  , while other experimen-

the sake of simplicity, it is sufficient to define only a single tal conditions, including D and f, are identical. Roughly,
electron transfer coefficient, e.g., the anodic one β and the in electrochemical terms, three kinetic regions can be
cathodic coefficient is[ α = 1 − β. ]𝛽 Thus, the rate constants can
](1−𝛽) identified, i.e., irreversible [log(κ) < − 1.5], quasirevers-
be written as ka = k� exp (𝜑) and kc = k� exp (−𝜑)  . ible [− 1.5 ≤ log(κ) ≤ 1] and reversible [log(κ) > 1]. In the
� �[

Commonly, for most of the electrode processes α = β = 0.5. reversible kinetic region, log(κ) > 1, the electrode kinetics
If 𝛽 ≠ 0.5 , then an asymmetry appears in the degree of vari- is fast enough to enable the redox equilibrium to be estab-
ation of the rate constants ka and kc for the same driving force lished in the course of the potential pulse. Accordingly, the
(in absolute value) 𝜑 of the electrode reaction. For these diffusion is the rate limiting step and the electrode reac-
reasons, β is frequently termed as a symmetry factor in the tion has an electrochemically reversible behaviour. The
context of electrode kinetics [9]. dimensionless function Ψ(k� , 𝜑, f , D) (Eq. 12) is becoming

In modelling of a quasireversible electrode reaction under identical with the function Ψ(𝜑) for the reversible reaction
voltammetric conditions, the most important equation is the (Eq. 6). The inequality log(κ) > 1 can be rewritten in the
Butler–Volmer equation, which relates the current to k∅  , following form:

the driving force 𝜑 , and the surface concentrations of redox ′


(13)

k∅ > 10 Df ,
species:
I (which has a very useful physical meaning. The left-hand side
k∅ represents the interfacial
�√ �electron transfer kinetics,
�[ ]𝛽 [ ′)
= k� exp (𝜑) cRed (0, t) − cox (0, t) exp (−𝜑) . (11)
]
FS while the right-hand side Df represents the mass trans-
It is derived taking into account Eqs. (8) and (9), and
fer kinetics (i.e., diffusion). Equation (13) is an important
considering that the net-rate of a heterogeneous electrode
criterion showing in what relation should be the kinetics of
reaction at a given driving force ( v = va − vc ) is equiva-
the electron transfer versus diffusion rate to exhibit a revers-
lent to the ratio FS
I
 . The latter is known as flux (in units of
−1 −2 ible electrochemical behaviour.
mol s  cm ), i.e., the amount of a reacted material per unit
Accordingly, the criterion√for a quasireversible behav-
of time and surface area. In addition, the mass transfer by
iour can be written as 0.032 Df ⩽ k∅ ⩽ 10 Df  . Within


diffusion has to be again taken into account, equivalent as
this region, the current is determined by the electrode
for the reversible electrode reaction. In the case of a quasi-
kinetics and the interfacial electron transfer is slower than
reversible electrode reaction, however, the rate of the elec-
the mass transfer. Considering a single voltammogram, the
tron transfer is the limiting step in the interplay between
latter statement is valid for the first part of the net SW
the electron and mass transfer for the major portion of the
voltammogram, including the net peak-current. At poten-
voltammetric experiment.
tials beyond the net peak-current, the driving force 𝜑 of
The voltammetric response under conditions of SWV can
the forward reaction (Red → Ox + e −) is becoming very
be expressed as
large, making the kinetics of electron transfer again faster

I = AΨ(k� , 𝜑, f , D), (12) than the diffusion, which is the reason for expansion of the
where A is the amperometric constant defined as for the diffusion layer and exhaustion of the initial
√ electroactive
reversible electrode reaction (Eq. 7). Yet, the dimensionless material Red. Whereas, for k < 0.032 Df the electrode
∅′

function is rather complex, depending on k∅ , 𝜑 , the fre-


′ kinetics is sluggish, causing the amount of formed Ox spe-
quency f (i.e., the duration of potential pulses), and the rate cies at the electrode surface(to be minute and the rate con-
](1−𝛽) )
stant of the reverse reaction kc = k� exp (−𝜑) to be
�[
constant of the diffusion D (i.e., the diffusion coefficient).
Fortunately, in the exact mathematical formula for the negligible for any 𝜑 , due to very small value of k∅  . As a

dimensionless function Ψ, the formal rate constant k∅  , fre-



consequence, the reverse (cathodic; Ox + e → Red) reac-
quency f, and the diffusion coefficient D are unified in� a tion does not proceed at a measurable rate at any potential;

single dimensional electrode kinetic parameter 𝜅 = √k  ,
Df

13
17   Page 12 of 14 ChemTexts (2018) 4:17

0.8

0.7 0.4
Ψ
reversible
0.6
E vs. E
0
-0.3 0.2 0.8
0.5 quasireversible Ψ
∆Ψ p

0.4
0.4 -0.4
0.8

E vs. E
Ψ 0
0.3 irreversible -0.3 -0.1 0.1 0.3
0.4

-0.4
0.2 E vs. E
0
-0.3 -0.1 0.1 0.3

0.1
-0.4

0
-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2

log(κ )

Fig. 8  Variation of the net peak-current of the dimensionless func- ΔE = 5  mV. The inset shows the dimensionless SW voltammetric
tion Ψ of a quasireversible electrode reaction with the electrode response typical for the irreversible (left), quasireversible (middle)
kinetic parameter κ, for amplitude Esw = 50  mV and scan increment and reversible (right) kinetic region

hence the reverse component of the response is missing the frequency of the potential modulation. Hence, the degree
(see the inset at the left-hand side of Fig. 8). For this rea- of reversibility of a single electrode reaction can be varied
son, such electrode reaction is termed as electrochemically by the frequency (i.e., duration of potential pulses). For illus-
irreversible. tration, let us assume an electrode reaction with
Besides remarkable variation of the net peak-current and k∅  = 0.08  cm s−1 and D = 5 × 10−6  cm2  s−1. At f = 5  Hz,

the morphology of the overall response with κ (cf. Fig. 8), log(κ) = 1.204; thus the reaction behaves as a reversible pro-
the position of the response at the potential axis is strongly cess (cf. Fig. 8). In other words, the duration of a single
sensitive to the electrode kinetics. Generally, the response potential pulse is long enough to achieve the redox equilib-
is shifted toward more positive potentials versus the formal rium at the end of the pulse [i.e., when the current is sam-
potential E∅ by decreasing the electrode kinetics, for the pled (cf. Fig. 5c)], whereas, for f = 1000 Hz, log(κ) = 0.054

case of the oxidative electrode mechanisms (Eq. 1), in which and the reaction behaves as a typical quasireversible process,
the experiments start in the presence of Red species only. in which the kinetics of the electron transfer is the rate deter-
Accordingly, the shift of the response will be in the opposite mining step. Obviously, the variation of the frequency from
direction (toward more negative values relative to E∅  ) for 5 to 1000 Hz transposes the system from the reversible into


reductive electrode mechanism (Ox + e  ⇄ Red). Hence, the the quasireversible kinetic region.
variation of the net peak-potential with the electrode kinetics The frequency exhibits even more complex effect to the
is an important criterion for characterisation of a quasire- real current (Eq.  12), as it affects simultaneously both
versible electrode reaction. � �Ψ(k , 𝜑, f , D) √a n d� t h e a m p e r o m e t r i c c o n s t a n t
��
��

Electrode kinetic parameter 𝜅 = √k unifies two typi- A = FScb, Red Df  . For these reasons, the real net peak-
Df

cal constants of a given experimental system ( k∅ and D) with current is complex, non-linear function on the square-root

13
ChemTexts (2018) 4:17 Page 13 of 14  17

0.45 0.45 0.45 Ψ


Ψ Ψ

0.25 0.25 0.25

(a) (b) (c)

0.05 0.05 0.05


E vs. E E vs. E E vs. E

-0.4 -0.2 0 0.2 0.4 -0.4 -0.2 0 0.2 0.4 -0.4 -0.2 0 0.2 0.4

-0.15 -0.15 -0.15

Fig. 9  Evolution of the dimensionless function Ψ of a quasireversible and 60 (c) mV. The frequency and the scan increment are f = 10  Hz
electrode reaction characterised with k∅  = 2 × 10−3  cm  s−1, β = 0.5, and ΔE = 5 mV

−6 2 −1
and D = 5 × 10   cm   s with the amplitude of Esw = 20 (a); 40 (b)

of the frequency, contrary to the reversible electrode reac- and the rate of the diffusion. Thus, it is expected to√exhibit
tion. So, the variation of the morphology of the response, an influence on the interrelation between k∅ and Df in

and the net peak-current and potential with the frequency is terms of the electrode kinetic regions. The influence of ΔE
the clear criterion for differentiation of the quasireversible on the electrode kinetic is yet insufficiently elaborated in
and reversible electrode reactions. the scientific literature, which remains to be a challenge for
Unfortunately, the voltammetric behaviour of the quasi- further studies.
reversible electrode reaction is even more complex and the
electrochemical reversibility is not solely determined by the
electrode kinetic parameter κ. The above given criterion for Conclusions
�t h e √q u a s i r e v e r√
s i b �l e kinetic region
0.032 Df ⩽ k ⩽ 10 Df is actually valid for the condi-
∅′ Square-wave voltammetry is certainly one of the most
advanced and versatile members in the family of pulse
tions of Fig. 8, i.e., the amplitude Esw = 50 mV and the scan voltammetric techniques. It possesses high analytical sen-
increment ΔE = 5 mV. Changing these intrinsic parameters sitivity and speed of measurements. The excitation signal,
of the SW potential wave-form affects the quasireversible i.e., potential modulation in the course of the voltammetric
kinetic region and the actual interrelation between k∅  , D and

measurement, enables the electrode reaction to be driven in


f. Let us recall that the amplitude determines the exact value both oxidative and reductive directions repeatable at each
of the pulse potential (cf. Fig. 5c). Thus, Esw affects the driv- step of the staircase potential, thus providing an insight into
ing force of the electrode reaction 𝜑 and finally, in a combi- the mechanistic aspects of the studied electrode reaction.
nation with k∅ and the surface concentrations c(0, t) (Eq. 11),

Hence, square-wave voltammetry combines the advanta-


determines the actual rate of the electrode reaction [10]. geous of pulse techniques, with respect to the high analyti-
Figure 9 clearly illustrates how the increasing amplitude cal sensitivity, and cyclic voltammetry, with respect to the
transposes a single electrode reaction from the irreversible electrode mechanism insight. With respect to the analyti-
to the quasireversible kinetic region, for a given experimen- cal performances it is quite comparable to differential pulse
tal conditions (i.e., constant frequency and scan voltammetry. However, SWV is yet superior in regard to
increment). electrode kinetics measurements and mechanistic aspects of
Finally, a similar effect is expected to be observed under electrode processes.
the influence of the scan increment for a single electrode The technique is generally fast. Typical time intervals for
reaction at a given set of values for Esw and f. As mentioned driving the electrode reaction in both cathodic and anodic
in “Stepwise potential variation is beneficial in voltamme- directions are in millisecond range. Hence, it is more suited
try”, the overall scan rate of the SW potential wave-form is to analyse reversible or quasireversible electrode processes
defined as v = f ΔE  . The scan rate determines the overall and electrode processes coupled with fast chemical reac-
time of the experiment, the expansion of the diffusion layer, tions, than to study sluggish electrode processes. When the

13
17   Page 14 of 14 ChemTexts (2018) 4:17

electrode reaction is very slow, square-wave voltammetry 2. Scholz F (2015) ChemTexts 1:17
provides even smaller response than other simpler form of 3. Bard AJ, Faulkner LR (2001) Electrochemical methods: funda-
mentals and applications, 2nd edn. Wiley, New York
voltammetry, e.g., linear sweep voltammetry, which is the 4. Inzelt G, Lewenstam A, Scholz F (eds) (2013) Handbook of refer-
main disadvantage of the technique. Therefore, cyclic vol- ence electrodes. Springer-Verlag, Berlin Heidelberg
tammetry is yet unavoidable in the preliminary mechanistic 5. Mirceski V, Komorsky Lovrić Š, Lovrić M, Scholz F (eds) (2007)
study of particular electrode mechanism. The strong synergy Square-wave voltammetry: theory and application. Springer-Ver-
lag, Berlin Heidelberg
between cyclic and square-wave voltammetry is achieved 6. Mirceski V, Gulaboski R, Lovric M, Bogeski I, Kappl R, Hoth M
in coupling both techniques in a form known as cyclic (2013) Electroanalysis 25:2411
square-wave voltammetry. Finally, it is worth mentioning 7. Mirceski V, Gulaboski R (2014) Maced J Chem Chem Eng 33:1
that square-wave voltammetry is particularly powerful tech- 8. Molina A, Gonzalez J, Scholz F (eds) (2016) Pulse voltammetry in
physical electrochemistry and electroanalysis: theory and applica-
nique when electrode processes are coupled with adsorption tions. Springer International Publishing, New York
phenomena, which could be the subject of elaboration of 9. Bard AJ, Inzelt G, Scholz F (eds) (2012) Electrochemical diction-
forthcoming chem-texts. ary, 2nd edn. Springer-Verlag, Berlin Heidelberg
10. Mirceski V, Laborda E, Guziejewski D, Compton RG (2013) Anal
Chem 85:5586

References
1. Wang J (2006) Analytical electrochemistry, 3rd edn. Wiley-VCH,
Hoboken

13

View publication stats

You might also like