You are on page 1of 48

MODULE IV ELECTROCHEMICAL AND ELECTROANALYTICAL

TECHNIQUES

CHAPTER 1: CHRONOAMPEROMETRY AND POLAROGRAPHY

LEARNING OBJECTIVES

After reading this chapter, you will be able to

(i) obtain an overview of electrochemical experiments

(ii) understand the salient features of the potential step experiments

and

(iii) comprehend the importance of polarography

There exist a plethora of electrochemical techniques which have been extensively


employed with varying objectives. It should be pointed out that all electrochemical
techniques lead finally either to numerical or analytical function such as
F (i, E, t, c) = 0
where i is the current, E denotes the potential and t is the ‘sampling time’, c being the
bulk concentration of the reactant.
As illustrative examples, we point that experiments based on current – potential response
( i vs E) are known as ‘ polarographic’ or ‘voltammetric’ techniques. If the working
electrode is a dropping Hg electrode, the experiment is customarily referred to as
‘polarography’. In all the other cases, where a stationary electrode is employed, the
technique ( i vs E) is known as voltammetry. In these techniques, either a constant or time
dependent potential is applied and the resulting polarogram or voltammogram is
recorded, most often by a work station provided with the necessary software for data
analysis too. Note that, in these techniques, current vs time is not analyzed.
Fig 1: Typical polarogram depicting the diffusion limited current and half wave
potential

Different potential profiles and the corresponding experimental techniques

Potential Profile Experiment Response


Potential step Chronoamperometry i vs t
Triangular
Potential sweep Cyclic voltammetry i vs E at different
scan rates

Potential variation
without any Polarography i vs E
time - dependent
characteristics
using dropping Hg
electrodes

Stair Case Stair Case i vs E


Square Wave Square Wave i vs E

Differential Pulse Voltammetry (DPV)

On the other hand, if our interest is in the dynamical response of a system, we can keep
the potential constant and obtain current vs time ( i vs t) as an output. These techniques
are known as chronoamperometric techniques since current is measured in ampere,
‘chrono’ indicating the time. Analogously, the potential vs time response obtained at
constant current is known as the chronopotentiometric technique. The functional form
where the bulk concentration of the reactant does not play an explicit role is the
fundamental feature of all electrochemical techniques. However, in experiments, wherein
the bulk concentration is deduced (or predicted) from a suitable electrochemical output
may be considered as electroanalytical techniques. In particular, the following types of
techniques can be envisaged.
Response Name of the quantitative analysis
i vs c Ampereometric titrations
E vs c Potentiometric titrations
q vs c Coulometric titrations
Cottrell equation
In chronoamprometric experiments, a potential step is applied as shown in Fig 1. By
adding an excess supporting electrolyte, the migration of the reactant is suppressed. Thus
the time-dependent solution of Fick’s first law of diffusion leads to the
chronoamprometric current as shown below. For the reduction process ox + ne- → Red

Cox Dox  2Cox


= (1)
t x 2
Cox ( t = 0, x ) = Cox (2)

Cox ( x = , t ) = Cox (3)

Cox ( x = 0, t ) = 0 (4)

where Cox denotes the bulk concentration. The current is defined as

 C 
i(t) = nFAD  ox  (5)
 x  x =0
The solution of the differential equation (1 ) subject to the initial and boundary conditions
yields
 x 
Cox ( x, t ) = Cox erf   (6)
 2 Dt 
where error function is defined as
2
x
erf ( x ) =    e− y dy
2
(7)
0

Cerf ( x )
Since = , (8)
x
the chronoamprometric current becomes

D
i = nFACox (9)
t

The above equation is known as the Cottrell equation and indicates that i varies as
t-1/2. The current response predicted by eqn (9) is known as the chronoamprometric
response, depicted in Fig. Despite the simplicity, Cottrell equation can be employed for
estimating (i) the diffusion coefficient of any unknown substance (ii) the true area of the
electrode and (iii) the bulk concentrations of analytes.

Diffusion layer thickness


It is customary to choose the potential such that the electron transfer is a rapid process,
the diffusion process being the rate – determining step. In view of this, the current in
equation (9) can be considered as the diffusion – limited current. Thus
nFAD1ox2 Cox
ilim =
t
However, from the definition of the diffusion flux under the Nernst diffusion layer
approximation, the limiting current is given by
nFADCox
i lim

Comparing the above eqns, we note that the diffusion layer thickness  is
 Dt (10)
The above t1/2 dependence of  has a microscopic origin since it can be derived in a
straight-forward manner from one dimensional random walk models. Further, the

quantity D is also known as the mass transfer rate constant (km) since it is a direct

measure of the amount of mass transferred.
Fig 1- Concentration-distance profile for semi-infinite linear diffusion to a planar

electrode surface. t 4  t3  t 2  t1

Note that the unit of mass transfer rate constant and heterogeneous electron transfer rate
constant is identical( cm sec -1).

POLAROGRAPHIC TECHNIQUES
The next experiment in the hierarchy of potential perturbation techniques is the classical
direct current polarography. In polarography, the working electrode is a dropping
mercury electrode (DME), whose schematic setup is shown in Fig. Under the pressure of
a mercury reservoir of height h, mercury flows with the predetermined flow rate ‘m’ from
a capillary arrangement. The drop forming at the capillary tip grows and then detaches
after a time td. After its detachment, a new drop starts to form and detach, thus the cycle
is repeated. These growing mercury drops, till the time they are not detached, act as the
cathodes for reduction of various organic and inorganic compounds.
The chronoamperometric current for the potential step experiments has been deduced
earlier as (eqn 9)

nFACox Dox
i=
πDt
where Cox is the bulk concentration of the reactant. Since the Hg drops are spherical,

4 3 mt
VHg = rHg = d
3 Hg
23
 3mt d 
Hence r =
2
 (11)
 4 
where V Hg denotes the volume of the mercury
Since the area of the sphere is given by 4  r 2 , the area of the electrode becomes

A = ( 4 ) 32 3 m 2 3 t d2 3−2 3
13
(12)
Substituting the above equation for ‘A’ in the Cottrell equation (9), the diffusion current
becomes
i d = 706nCox Dox1 2 m 2 3 t1d 6 (13)

where id is in A (amperes) if Dox is in cm 2s −1 , Cox is in mol cm-3, m is in mg/s, and td is

in seconds. While the above equation is the maximum diffusion current, the average
diffusion current is lower and an equation for the mean diffusion current is as follows:
6
id = i d = 607nCox Dox1 2 m 2 3 t1d 6
7
Equation (13) is the Ilkovic equation, describing the diffusion current at the dropping
mercury electrode.While the Ilkovic equation yields the diffusion current, it is essential to
construct a polarogram which enables the quantative and qualitative identifications of
diverse species.
Heyrowsky – Ilkovic equation
For the reduction process, ox + ne- → Red,
the Nernst equation is
RT Cox( 0,t )
E = Eo + ln
nF Cred( 0,t )

However, as per conservation of mass,


( ) ( )
Cox 0, t d + C red 0, t d = C bulk

i = 706n Cox − Cox ( 0, t d )  D1ox2 m 2 3 t1d 6 = i d − 706nCox ( 0, t d ) D1ox2 m 2 3 t1d 6

Hence
id − i
Cox ( 0, t d ) =
706nD1ox2 m2 3 t1d 6

i = 706n Cred ( 0, t d ) − Cred


b
 D1red2 m 2 3 t1d 6

i
Cred ( 0, t d ) = 12
706nD m 2 3 t1d 6
red

12
RT  Dred  RT  id − i 
E=E + o
ln   + ln  
nF  Dox  nF  i 

Assuming that D red = Dox , we rewrite the above equation as


RT i d − i
E = Eo + ln
nF i
We note that when the current i equals id/2, the potential E becomes the half-wave
potential E1/2. The experimental current – potential response is analyzed by constructing a
polarogram. The above equation indicates that the potential at which the current is half of
the maximum current is characteristic of the species being analysed. Since the potential
corresponds to the ‘half’ of the diffusion current in the polarographic ‘wave’, this
potential is denoted as the ‘half wave potential’. Since E1/2 is approximately ( D red = Dox )

equal to E o for most compounds, this polarographic technique is the simplest method of
estimating standard reduction potentials of any organic/inorganic compounds or polymers
or nanomaterials. This significant advantage is the reason for the popularity of the
polarographic technique, even after several decades of its discovery.
Fig 2: Suppression of polarographic maxima using the surfactant Gelatin

The half – wave potential is related to the formal potential Eº as


1/2
RT  DRed 
E1/2 =E +
o
ln  
nF  Dox 
where DRed and Dox are the diffusion coefficients of the reduced and oxidized forms.

Table: Half wave potentials of various functional groups

Functional group E1/2 (V)


-N = N- -0.4
- C = C- -2.3
−C  C − -2.3
C=O -2.2
S-S -0.3
NO2 -0.9
C-X -1.5

For irreversible reduction, the potential can be deduced as

RT  Dox  RT   id − i   t  
1/2

E = E1/2 + ln  + ln 1.35k f    
nF  D Red  nF   i   Dox  
 
where  is the transfer coefficient and k f is the rate constant of the reaction.
Polarographic study for complex formation
The half – wave potential becomes more negative for the reduction of metal complexes,
since additional energy is required for their decomposition. If a complex MXq undergoes
reduction on Hg,
Reduction
Mqx ⎯⎯⎯⎯⎯
→M + qX

where X is the ligand and q is the stoichiometric number. The difference between the half
– wave potential for the complexed and uncomplexed metal ion can be deduced as
1/2
RT RT RT  Dfree 
( E1/2 )complex − ( E1/2 )free = ln K f − q ln[X] + ln  
nF nF nF  Dcomplex 
where K f is the formation constant. The stoichiometric number can thus be computed
from the slope of ( E1/2 )complex versus ln [X]. while the slope yields the formula of the

complex.

The polarographic wave sometimes exhibits a maximum before reaching the current
plateau. This is known as the polarographic maximum and is attributed to a
hydrodynamic flow of the solution surrounding the Hg drop. The polarographic maxima
can be suppressed by adding a small amount of surfactants such as Triton X – 100,
gelatin etc as shown in Fig 2.

WORKED OUT EXAMPLES


1. In a potential step experiment using a glassy carbon electrode of 1mm diameter,
the steady state current was 70 µA after a time duration of two seconds, while
employing 0.1M Ag+ as the reactant. Estimate the diffusion coefficient of Ag+.

The Cottrell equation is


nFAD1ox2 Cox
i=
t
1 96500C mol−1  3.14 ( 0.05cm )  D1ox210−4 mol cm −3
2
−6
70 10 A =
3.14  2 sec

1.754 10−4
D 12
Ag +
= cm sec−1 2
0.07575
DAg+ = 5.36 10−6 cm 2 sec −1

2. The diffusion coefficient of potassium ferrocyanide is 1.52 x 10-5 cm2/sec at 298K


deduced from chronoamperometry. Estimate the time required to traverse the
diffusion layer, whose thickness is 0.01 cm.
 Dt

0.01 cm 3.14 1.52 10−5 cm 2 sec −1 t


( 0.01)
2
sec
t =
3.14 1.52 10−5
2.09 seconds

3. Show that the concentration of the reactant at any time t for chronoamperometry
given by
 x 
Cox ( x, t ) = Cox erf  
 2 Dox t 
satisfies all the initial and boundary conditions of Fick’s law of diffusion
pertaining to potential step experiments.
At x = 0, erf (0) = 0  Cox ( x, t ) = 0

At x =  , erf (  ) = 1  Cox ( x, t ) = Cox

1
At t = 0, erf   = erf (  ) = 1  Cox ( x, t ) = Cox
0

4. The mass transfer coefficient of Cd2+ is deduced as 10-2 cm/sec when the area of
the electrode was 0.1 cm2 and the bulk concentration was 20 x 10-6 moles cm-3.
Estimate the chronoamperometric current, assuming planar diffusion.

nFADCox
i lim = nFACox k m

= 2  96500 C mol−1  0.1cm 2  0.01 cm sec−1  20 10−6 mol cm −3
= 3.86 10−3A

5. Integrate the Cottrell equation and interpret the same

nFAD1ox2 Cox
i(t) =
t
nFAD1ox2 Cox −1 2
  i ( t ) dt = q ( t ) =  t dt

2nFAD1ox2 Cox t −1 2
=

This equation represents the total charge consumed in a reduction process involving n
– electron transfer. The above equation is also known as the Anson equation.

EXERCISES
1. The Cottrell equation pertains to the planar diffusion. If the diffusion is towards a
spherical electrode of radius r, what will be the transient current in
chronoamperometry?
2. If the dimension of the working electrode is <<1 µm, will the Cottrell equation be
valid?
FREQUENTLY ASKED QUESTIONS
1. What are the differences between diffusion layer thickness and double layer
thickness?
2. What are the recent developments in polarography?
3. What are the limitations of polarographic techniques?

SUMMARY
The salient features of potential step experiments and classical direct current
polarography have been highlighted.

CHAPTER 2 - CYCLIC VOLTAMMETRY

Introduction to Cyclic Voltammetric techniques


The Cyclic voltammetry is a technique wherein the time – dependent potential is applied
to a system and the resulting current vs potential response is recorded. Hence it is also
known as stationary electrode polarography.
A schematic representation of the experimental set up is given below.
Figure 1: A typical electrochemical cell for cyclic voltammetry.
WE-Working Electrode; RE-Reference Electrode; CE-Counter Electrode

The characteristic feature of the cyclic voltammetry is the triangular potential pulse (E vs
t) shown in the Fig 1 as the function generator. For the reaction Ox + ne- = Red, the
potential pulse applied in cyclic voltammetry is given by the equation
E(t) = Ei - t 0<t (1)
indicating that the potential is made more negative (‘cathodic’).  denotes the time at
which the potential is reversed (‘switching time’). E i denotes the initial potential,  being
the scan rate. The scan rates may range from 10-3 volt per second to 106 volts per second.
While carrying out cyclic voltammograms, it is customary to specify the initial potential,
scan rate and the potential range. Consequently, it follows inter alia, that one follows the
given reaction dynamically since
t = (Ei - E)/
Thus cyclic voltammetry is also sometimes referred to as a potentiodynamic technique.
Instead of equation (1), one can also write the potential as
E(t) = Ei + t
In this case, it implies that the potential is made more positive (anodic).
Hence it is imperative to specify the scan direction while representing cyclic voltammetry
data.

Fig 2: The triangular potential sweep employed in cyclic voltammetry


E = Ei + t; 0 < t < 
E = Ei + t - 2
As in most electrochemical experiments, it is customary to solve the one-dimensional
Fick’s second law of diffusion viz
COx ( x, t )   2COx ( x, t ) 
= DOx   (2)
t  x 2 
subject to appropriate initial and boundary conditions. We consider the evaluation of the
current for planar electrodes. We note that such partial differential equations arise also in
the heat transfer processes with the replacement of ‘concentrations’ by ‘temperature’.
However, a simplifying feature in the cyclic voltammetric analysis (in contrast to the heat
transfer exercise) is that one seldom requires the reactant concentration as a function of
time and distance throughout the ‘reaction layer’. On the other hand, the experimental
observable in cyclic voltammetry is the current defined in terms of the flux at the
electrode surface. Thus, the complete concentration profile C(x, t) is un-necessary. With
this proviso, we proceed to solve the equation (2). In view of the structure of the partial
differential equation (2), one requires one condition in ‘time’ (eqn 3) and two conditions
in ‘distance’ (eqns 4 and 5) for obtaining the complete solution.

Reversible Electron Transfer


It is convenient to consider the reversible electron transfer to begin with, in order to
comprehend the complexity for other reaction schemes. The conditions required for
solving equation (2) are as follows:
COx ( x, 0 ) = COx
*
(Initial condition at t =0) (3)

lim COx ( x, t ) = COx


*
(Boundary condition at x → ) (4)
x →

From the Nernst equation, we write


RT Cred
E = E0 − ln
nF Cox
RT C
ln Ox = E − E 0
nF CRe d

where E = Ei - t. Hence,


 nF
( Ei − t − Erev ) (surface condition at x = 0)
COx (t )
= exp  (5)
CRe d (t )  RT 
The expression for current arising from the Fick’s law is
I  C 
= DOx  Ox 
nFA  x  x =0
The seminal analysis provided by Nicholson and Shain, in 1966 cleverly defines a
dimensionless current function such that it becomes possible to consult the tabular
compilation for each value of the ‘potential’ as shown below. Such compilations are
available for almost all reaction schemes that can be envisaged.
The dimensionless current function  (t) for reversible electron transfer processes is
defined by Nicholsm and Shain as
I = nFACOx
*
( DOx )1 2  ( t )

I
=  ( t )
nFAC ( DOx )1 2
*
Ox

where  = nF/RT. The unit of  is (time)-1 and hence the argument of the current
function viz (t ) is dimensionless. The l h s of the above equation is the experimentally
observed current for a given potential or equivalently (E - E0). For a given scan rate, all
the quantities on the l..h..s are also known. This implies that for a given potential (E - E0),
experimental values of  (t) are also known. The Table given below provides the
theoretical current functions for various potentials. The most salient features of the
analysis by Nicholson and Shain are the diagnostic criteria for mechanistic analysis of
electron transfer reactions. In the case of reversible electron transfer processes, the
following quantitative expressions arise:
Randles-Sevcic equation:

12
 F  12 12
I p = 0.4463n FACOx    DOx
32

 RT 

RT
Peak potential: E p − E 0 = −1.109
nF
RT
Half peak width: E p − E p 2 = −2.20
nF
I p ,a
Ratio of peak currents: =1
I p ,c
57.0
Peak separation: E p ,a − E p ,c = mV
n

Table 1: Values of current function 1/2(t) for linear sweep voltammetry in the
case of reversible electron transfer reaction ox + ne- Red

n(E – E0) / mV 1/2(t) n(E – E0) / mV 1/2(t)

120 0.009 -5 0.400

100 0.020 -10 0.418


80 0.042 -15 0.432

60 0.084 -20 0.441

50 0.117 -25 0.445

45 0.138 -28.50 0.4463

40 0.160 -30 0.446

35 0.185 -35 0.443

30 0.211 -40 0.438

25 0.240 -50 0.421

20 0.269 -60 0.399

15 0.298 -80 0.353

10 0.328 -100 0.312

5 0.355 -120 0.280

0 0.380 -150 0.245


Figure 3 (A): The variation of the dimensionless current function with potential for
reversible electron transfer at planar electrodes predicted by the tabular compilation of
Nicholson and Shain (B) Typical cyclic Voltammogram for reversible electron transfer.
Irreversible electron transfer reactions

Fig 4: Cyclic Voltammogram of irreversible electron transfer processes


It is well known that for irreversible electron transfer processes, the rate constants and
transfer coefficients should enter the description for cyclic voltammograms.
Consequently, the current is given by the equation

I  C 
= DOx  Ox  = khet COx (0, t )
nFA  x  x =0
where khet denotes the rate constant for the reduction process given by

khet = k e0 ( )
− nF E − E 0 / RT
het

k0het being the value of khet at E = E0


In this case, the dimensionless current function for irreversible electron transfer process is
defined as

  nF 
12

I = nFACOx DOx 
12 12
    (bt )
 RT 
 nF
where b = and has the dimensions of (time)-1.
RT
Table 2: Values of the dimensionless current function 1/2(bt) for linear sweep
voltammetry pertaining to irreversible electron transfer process Ox + ne- → Red

Potential/mV 1/2(bt) Potential/mV 1/2(bt)

160 0.003 15 0.457

140 0. 008 10 0.462

120 0.016 5 0.480

110 0.024 0 0.492

100 0.035 -5 0.496

90 0.050 -5.34 0.4958

80 0.073 -10 0.493

70 0.104 -15 0.485

60 0.145 -20 0.472

50 0.199 -25 0.457

40 0.264 -30 0.441

35 0.300 -35 0.423

30 0.337 -40 0.406

25 0.372 -50 0.374

20 0.406 -70 0.323


The diagnostic criteria in this scheme are as follows.
  nF 
I p = 0.227nFACOx
* 0
khet exp  −
 RT
( E p − Erev ) 

RT   D1/ 2    nF  
1/ 2

E p = Erev −  0.780 + ln  Ox
0 
+ ln   
 nF   khet   RT  
RT 47.7
E p 2 − E p = 1.857 = mV at 298.1 K
 nF  n
We note that there are three unknown parameters in the equation for current function viz
(i) standard heterogenous rate constant; (ii) transfer coefficient and (iii) standard
reduction potential.

Quasi-reversible electron transfer reaction

Fig 5: Cyclic Voltammogram of quasi – reversible electron transfer processes

The boundary condition in this eqn is as follows:


  COx ( x, t )  − n
F
 E ( t ) − Erev   n
F
 E (t ) − Erev  
 = khet e COx (0, t ) − CRe d (0, t )e
0
DOx  RT RT

 x  x =0  
The dimensionless current function is defined as
12
 nF 
I = nFAC D    ( E )
* 12 12
Ox Ox
 RT 
The tabular compilations for the quasi-reversible electron transfer process is more
involved but have been provided by Nicholson and Shain. We note that although there
57
exist two peaks as in the case of reversible processes, the peak separation is not mV .
n
The following criteria is useful for classification for evalution of reversibility, when
D =10-5 cm2/s and T = 298K.
Reversible: k0 > 0.3 v1/2 cm/s
Quasi – reversible: k0 > 2 x 10-5 v1/2 cm/s
Totally irreversible: k0 < 2 x 10-5 v1/2 cm/s
where v denotes the scan rate
The objectives regarding the construction of cyclic voltammogrm are varied and are
dictated by the target in mind. In particular, if the aim is to deduce the redox property of a
newly synthesized material, one looks for the reversible cyclic Voltammogram as shown
in Fig 3, since E = ( E pc + E pa ) / 2 . On the other hand, if the objective is to study any

electrochemical reaction, it is customary to construct cyclic voltammogram for


irreversible electron transfer processes as shown in Fig 4, in order to estimate the rate
constants and transfer coefficients. However, the cyclic voltammograms seldom are
solely reversible/ irreversible/ quasi – reversible. One comes across dimerization reaction,
complexation process, disproportionation etc. Scheme 1 depicts a plausible method of
classification of electron transfer schemes using the mass transfer rate constants vis a vis
electron transfer rate contants.
Table 1: - Time-dependent surface concentration for electron transfer reactions.
Dimensionless surface concentration
Reaction type
of the oxidized species, COx(0, t)/COx*
kw
Reversible
1 + kw

 −k  
Irreversible exp  0 
 Dkw 

 −k  
kw + exp  0 (1 + kw ) 
Quasi-reversible  Dkw 
( w)
1 + k
(C0* refers to the bulk concentration,  denotes diffusion layer thickness and k0, the
standard heterogeneous rate constant; kw = exp[nF(E-E0)/RT], where E0 is the standard
electrode potential).

Scheme 1: Flowchart depicting the criteria for classification of electron transfer


schemes

Since E=Ei-t, surface concentration becomes time-dependent in an implicit manner.


Lest one may think that the cyclic voltammetric analysis follows any of the three
mechanisms viz reversible, irreversible or quasi – reversible, we hasten to add that there
exist innumerable mechanisms encountered in practice while studying the redox
behaviour of unknown compounds. In such cases, there exist several options to unravel
the reaction mechanism. Among them, mention may be made of the following: (i) Digital
simulation of cyclic voltammogram using the in – built software of the electrochemical
workstations; (ii) solving the Fick’s law of diffusion for the desired geometry of the
electrode and reaction conditions; (iii) Non – linear curve fitting for the entire cyclic
voltammogram and interpretation of the fitting constants; (iv) resorting to qualitative
diagnostic criteria advocated by Nicholson and Shain for EC, EC’ ECE reaction etc; (v)
judicious choice of scan rates to make the system totally reversible or irreversible and (vi)
employing the Convolution Potential Sweep Voltammetry (CPSV).

Convolution Potential Sweep Voltammetry

The convolution potential sweep voltammetry is a more valuable technique than the
conventional voltammetry due to the following reasons (i) The entire set of data points
of the voltammogram can be made use of unlike the conventional methods where the
peak parameters alone are employed; (ii) The elucidation of the mechanism is more
accurate than the conventional method since a wide range of scan rates (10 mV/sec to 106
V/sec) are employed; (iii) potential dependent rate constants (khet) and transfer
coefficients () can be estimated and (iv) kinetic laws such as Butler – volmer equation
need not be assumed a priori. The convoluted current is often estimated from the normal
voltammetric current using the semi - integral equation.
t
1 i (u )
I (t ) =
  1
du
0
(t − u ) 2

where I is the convoluted current and t is the time. The integration of voltammetric
current with respect to time yields the total charge. However, the convoluted current
which is the semi integration of current has the properties of current as well as charge and
hence its unit is Ampere sec1/2
I 10-5 / A s1/2
i  10-5 / A

Potential vs. SCE / V

Fig 6: Convolution Potential Sweep Voltammetry of a typical organic compound at low


scan rates. (<10 mV/sec ). Note the plateau for the convolution current.

A sigmoid shape convoluted current reaches its limiting value as shown in Figure 6. The
plateau of the sigmoidal curve is called limiting convoluted current (Il) and is given by
I l = nFACb D

where Cb is the bulk concentration of the reactant, D being the corresponding diffusion
coefficient, A is the surface area of the electrode and ‘n’ denotes the number of electrons
involved. The surface concentrations of the oxidized and reduced species can be easily
calculated from the limiting convoluted current as
Il − I (t )
Cox (0, t ) =
nFADo1/2

I (t )
CRed (0, t ) =
nFAD1/2
R

By substituting the surface concentrations in the convolution current, the potential -

dependent rate constant is calculated as

I l − I (t )
ln(k ET ) = ln( D ) − ln
i (t )
We note that the CPSV has the ability to (i) provide directly the surface concentrations
Cox (0, t ) and CRed (0, t ) ; (ii) yield the potential-dependent rate constants; (iii) provide the
transfer coefficients and (iv) verify the validity of the Butler – Volmer equation by
constructing Tafel plots.

WORKED OUT EXAMPLES

1. For a symmetrical cyclic voltammogram of a reactant at a concentration of 0.002


M, the peak current is 0.2 mA at the scan rate of 100 mV/s. Estimate the number
of coulombs consumed in the process, assuming the diameter of the electrode to
be 2 mm and the number of electrons to be 1.

For symmetrical cyclic voltammogram with ΔE p = 0,

n 2 F 2 Av
ip =
4 RT
where
 = Surface coverage of the absorbed reactant species
A = area of the electrode =  r 2 = 0.0314 cm2
Since i p = 0.2 mA and v = 100 mV/s,  is evaluated as

0.2 10−3  4  8.314  298


=
965  965  104  3.14  10−3
= 6.77  10−8 mol cm-2

Therefore,
Charge consumed = nFA = 1 96500  0.0314  6.77 10−8
= 20.5 10−5 Coulombs

2. For the reversible oxidation of catechol, (concentration 0.5 μM ) the anodic peak
current is 3.5 μA at a scan rate of 100 mV/s. using 2 mm diameter of Au
microelectrode. Calculate the concentration of catechol for which the five field
increase of current occurs at a scan rate of 50 mV/s.

Here, n = 2

( )
i p1 = 2.69 105 n3/2 Ac1 D1/2 v11/2

( )
i p2 = 2.69 105 n3/2 Ac2 D1/2v1/2
2

1
i p1 c v  2
= 1  1 
i p2 c2  v2 
or,
1/2
c1  0.05 
5= −3
 
0.5 10  0.1 
or,
c1
5=  0.707
0.5 10−3
 c1 = 3.5 10−3 M = 3.5 mM

3. The cyclic voltammogram for a reversible system has the anodic and cathodic
peak potential separation as 0.59V at 298 K. The cathodic potential is 1.31 V. All
the potentials are relative to Ag | AgCl | 0.1 M KCl reference electrode. Calculate
the equilibrium constant.

Here,
ΔE p = 0.59V
E cp - E ap = 0.59V; E cp = 1.31V
 E ap = 0.72V
E cp + E ap 1.31 + 0.72
E = = = 1.01V
2 2
ΔG o = -nFE o = -RTlnK
or
RT
Eo = lnK
nF
or
nFE o 1 96500 1.01
lnK = =
RT 8.314  298
K = 1.2110 17

CHAPTER 3: BASICS OF ELECTROCHEMICAL IMPEDANCE


SPECTROSCOPY

LEARNING OBJECTIVES

After reading this chapter, you will be able to

(i)comprehend the basic concept of impedance

(ii)construct equivalent circuits and derive the corresponding impedances

and

(iii) deduce the electrochemical system parameters from the impedance data

The impedance measurements are today extremely popular on account of its applications
in electrode kinetics, electrical double layer, batteries, fuel cells, supercapacitors,
corrosion, solid state ionics, bioelectrochemistry etc. It is one of the linear response
methods, which means that the system is perturbed by a sine wave current or potential of
such a small amplitude that the response contains only the relationship between the input
and output.
A common method of impedance measurements is to apply a small amplitude sinusoidal
potential perturbation superimposed on a direct potential recorded at various frequencies.
The impedance spectrum is measured at various values of the potential. This method is
designated as “Electrochemical Impedance Spectroscopy” (EIS), in order to make the
linear response theory valid,it is customary to employ a very small amplitude, whose
maximum value is preferably 5mV.
If a sinusoidal potential pulse such as E = E sin (t ) is applied to a system, the current

response is dictated by the dynamics of the processes under consideration. Any


electrochemical system consists of electrodes, electrolytes and reactants encompassing
various phenomena such as activation, diffusion, charge transfer etc. In view of the
electrolytes possessing conductance, there exists ohmic resistance while diffusion causes
the so called ‘diffusional’ or ‘mass’ transfer resistance. In addition, the interfacial region
(electrical double layer) stores charges and hence behaves as a capacitor. In view of the
above, it is customary to consider any electrochemical system as composed of diverse
capacitors and resistors connected either in parallel or in series. The electrode also has a
fundamental role in view of its geometry (planar, spherical etc) and nature (porous, non
porous, fractal etc).
The connectivity of these capacitors and resistances in response to alternating potential /
current is the subject of the Electrochemical Impedance Spectroscopy (EIS). Since
several electrochemical processes are essentially impediments to the natural flow of
charges, the term ‘impedance’ is used. Since the impedance data is a precise mapping of
various processes with different time constants, the impedance technique is analogous to
conventional spectroscopy.

Introduction to the concept of Impedance


The classical Ohm’s law is given by
E
R=
i
where R denotes the resistance, E and I being the applied potential and current
respectively.
On the other hand, if the potential or current is made ‘time – dependent’, the concept of
impedance arises. Impedance essentially is a factor which ‘impedes’ the flow of charge
carriers. It is customary to define a complex impedance Z(t) as
E (t )
Z (t ) =
i (t )
If E (t) and i (t) are written respectively as
E ( t ) = E cos (t ) (1)
and
I ( t ) = I cos (t ) , (2)

the above equation for impedance becomes

E cos(t )  cos(t ) 
Z= =Z   (3)
i cos(t −  )  cos(t −  ) 

where Z 0 is the magnitude of the impedance,  is known as the phase angle,  being

the radial frequency (2  f) while f is in Hz.


We can write Z as
E e jt
Z= = Z e j = Z ( cos  + j sin  ) (4)
i e( jt − j )
using de Moivre theorem. We note here that Z has a real and imaginary part. The
identification of real and imaginary parts of the complex impedance is a fascinating
exercise and depends upon the properties of the ‘circuit elements’. It is convenient to
write the impedance Z as
Z = Zreal - jZim, (5)

for impedance of the real and imaginary parts.The inverse of impedance is known as
admittance.
There are several methods of representing the impedance data. Among them, the most
important ones are
i. Nyquist plot ( Imaginary vs real part of the impedance)
ii. Bode’ magnitude plot (|Z| vs log  )
iii. Bode’ phase angle plot  vs log  ( or tan  vs log  )
Any equivalent circuit in electrochemical system is a combination of capacitances,
resistances and inductances. The inclusion of the inductance ensures that all frequency
ranges are taken into account.
IMPEDANCE OF A FEW COMMON CIRCUIT ELEMENTS
TABLE 1: Impedance of a few common Equivalent Circuit Elements
Circuit Element Impedance
Resistor R
Capacitor 1/ jC
1/ Q ( j ) 

Constant Phase element  

Inductor j L
Warburg Impedance (  ) (1 − j )

RESISTOR
In the case of a pure resistor, the imaginary part of the impedance is zero and Z real = R, as
shown in Figure 1. The graph of – Zim vs Zreal for various frequencies is customarily
designated as the Nyquist plot.

Fig 1: Nyquist plot for a pure resistor

CAPACITOR
The impedance pertaining to a capacitor is slightly complex and can be derived as
follows. The fundamental definition of capacitance is
C = Capacitance = Charge = dq = idt
Potential dE dE

dq and dE being infinitesimal changes in charge and potential respectively.

Hence, i = C dE
dt

For the potential perturbation given by E = E e jt , the above equation yields the

following expression for the impedance of the capacitor.


dE
Since i = C ,
dt
we can write explicitly the current as
i ( t ) = C ( ) E e jt j

where C ( ) indicates that the capacitance is a function of the frequency  . Thus

E (t ) E e jt 1
Z= = = (6)
i(t ) C ( ) E e j jC
jt

Fig 2: Nyquist plot of a capacitor and resistor connected in series


Thus the impedance of a capacitor is given by 1/ jC . We note that the real part of the
impedance in the case of a capacitor is zero while the imaginary part is given by −1/ C .
Equivalent circuits in general are composed of several capacitors and several resistors
connected in series or parallel. If a capacitor and a resistor is in series, the total
impedance is given by
1
Ztotal = R + (7)
jC

Fig 2 depicts the corresponding Nyquist plot.

REPRESENTATIVE EQUIVALENT CIRCUITS IN ELECTROCHEMICAL


SYSTEMS

(A) Capacitor and resistor in series

Since the circuit elements are in series, Kirchhoff’s law leads to


ZTotal = ZR + ZC

where ZR and ZC denote respectively the impedance of a resistor and a capacitor.

1
Since ZR = R and ZC =
jωC

1 j
ZTotal = R + =R− (8)
jωC ωC

Since ZTotal = Zreal - jZim,

Zreal = R (9)
and
1
Zim = (10)
ωC

A popular method of representing the impedance data is to construct an Argand diagram


which is well known in the field of complex numbers. This plot involves the graph of Zim
or -Zim vs Zreal. For the above circuit, it follows from equations (9 ) and ( 10) that

2 2
 Rs   Rs 
 Z re −  + Zim =  
2

 2   2 

which is the equation for a semi – circle with the radius being Rs/2. By providing various
values of Zre , the corresponding Zim can be obtained, as shown in Fig ( ).
Another method of representing the impedance data is to construct Bode’ plots. There are
two types of Bode’ plots viz (i) Bode magnitude plot and (ii) Bode phase angle plot. The
Bode’ magnitude plot involves the variation of log |Z| vs frequency while the Bode’
phase angle plot involves constructing tan  vs  plots. In this case

Zim
tan  = − =  Rs C (11)
Z re
− = tan −1 ( Rs C ) (12)
Since the frequencies are in general varied from 10-3 Hz to 106 Hz, it is more convenient
to employ logarithmic scale for  . Fig 3 depicts the Bode’ phase angle and Bode’
magnitude plots.

Fig 3: (a) Nyquist; (b) Bode’ phase angle (c) Bode’ phase angle and plots for a capacitor
and resistor in series.

In circuits many equivalent, the capacitance C is often associated with the double layer
capacitance (Cdl). While the resistance R is often the solution resistance.

(B) Capacitor and resistor in parallel

Since the circuit elements are in parallel, Kirchhoff’s law leads to


1 1 1
= +
Z Total Z R ZC
ZTotal = R || C (13)

Z R .Z C
ZTotal =
Z R + ZC

1
R
jC
ZTotal =
1
R+
jC

R R (1 − j RC )
= = (14)
1 + j RC 1 +  2 R 2C 2

R  2 R 2C
= −j (15)
1 +  2 R 2C 2 1 +  2 R 2C 2
Since ZTotal = Zre - jZim,

R ωR 2C
Zre = ; Z Im = (15)
1 + ω2 R 2C 2 1 + ω2 R 2C 2

− Z im -ωR 2C
tan  = = = − RC (16)
Z re R
where  is the phase angle

Fig 4: Nyquist and Bode’ plots for a capacitor and resistor in parallel
From Fig 4, we note that if the mass transfer is not significant , the charge transfer
resistance (Rct ) can be directly obtained.

(C)

Note that this circuit involves an additional resistor in series with the circuit of B. Hence
1
Rct 
jC1
ZTotal = RS +
1
Rct +
jC1
Rct 1 + j Rct C1
= RS + 
1 + j Rct C1 1 + j Rct C1

− 2 Rct2 C12
tan  =
RS (1 +  2 Rct2 C12 ) + Rct

Rct (1 − j Rct C1 )
= RS +
1 +  2 Rct2 C12

since ZTotal = Z re − jZ Im
Rct  2 Rct2 C1
Z re = RS + , Z =
1 +  2 Rct2 C12 1 +  2 Rct2 C12
Im
(17)

The Nyquist plot is as follows


Fig 5: Nyquist and Bode plots for the circuit (C)

(D) A well known equivalent circuit in various contexts in electrochemistry such as (i)
supercapacitors (ii) underpotential deposition of materials and (iv) two – dimensional
phase transitions is as follows:

Capacitors sometimes function as constant phase elements and their impedances are
defined as
1
ZCPE =
( j )
n
C
(E) Randles Circuit

ZTotal = Rsol + Cdl || ( Rcr + ZW )

  (1 − j )  1
 Rct + 
  jCdl
= Rs + 
 (1 − j ) 1
Rct + +
 jCdl

   j 
 Rct + − 
   
= Rs +
 ( jCdl ) . Cdl
1 + jCdl .Rct + +
 

ZTotal =
(Z )(R
CPE1 ct + Z CPE2 ) (18)
Rct + ZCPE2 + Z CPE1

1    
  Rct + − j
jQ1    
ZTotal = Rs + (19)
1  
+ Rct + − j
jQ1  

Warburg impedance
The diffusion of species yields another impedance designated as Warburg impedance.
The equation for the ‘infinite’ Warburg impedance is:

Zw,infinite =σ ( ω ) (1- j)
-1
2
(20)

σ is the Warburg coefficient defined as:


RT  1 1 
=  +  (21)
n 2 F 2 A 2  C * D0 C * DR 

where,
ω = 2 πf
D0 = diffusion coefficient of the oxidant
DR = diffusion coefficient of the reductant
A = surface area of the electrode
n = number of electrons involved
The word ‘infinite’ indicates that the diffusion layer thickness is infinite. σ is known as
the Warburg coefficient. On a Nyquist plot, the Warburg impedance appears as a
diagonal line with a slope of 45º. On a Bode’ Plot, the Warburg impedance exhibits a
phase shift of 45º.
If the diffusion layer thickness is finite,
  j  12 
Zw,finite =σ ( ω ) (1- j) tanh     
-1
2
(22)
 
D

where δ denotes the Nernst diffusion layer thickness
D0 + DR
D denotes the average value of the diffusion coefficients
2
   j
 Rct + −
  
ZTotal = Rs + (23)
(
1 +  .  .Cdl + j Cdl Rct +  . Cdl )
  

 Rct +
 
(
 1 +   Cdl +

) (
 Rct Cdl +   Cdl )
Z re = Rs + (24)
( ) ( )
2 2
1 +   Cdl +  Rct Cdl +   Cdl

  

 Rct +

(
  Rct Cdl +   Cdl +
 
)
1 +   Cdl ( )
Z Im = (25)
( ) ( )
2
1 +   Cdl +  Rct Cdl +   Cdl
(E)

1
( Rct + Z )
1 jCd
ZTotal = Rs + +
jCL 1
+ Rct + Z
jCd
1 Rct + Z
= Rs + +
jCL 1 + jCd ( Rct + Z )

(F)

1 1 1
R1  R2  R3 
jQ1 jQ2 jQ3
Z total = + +
1 1 1
R1 + R2 + R3 +
jQ1 jQ2 jQ3

R1 R2 R3
= + +
1 + j R1Q1 1 + j R2Q2 1 + j R3Q3

Time Constants
In simple RC circuits, the double layer capacitance and charge transfer resistance values
enable the estimation of the corresponding time constant  using
 = Rct Cdl
We note that the time constant is also influenced by the geometry of the electrode; in the
case of microelectrodes, the time constant increases with increasing radius. For practical
supercapacitors, low time constants are preferred so as to ensure fast charge – discharge
characteristics. There are two distinct approaches to the construction of equivalent
circuits for any electrochemical system viz (i) fitting of the impedance data using the in-
built software of any electrochemical work station or (ii) formulating the real and
imaginary parts of the impedance with subsequent non – linear regression analysis using
any programming languages. The impedance equations provide an insight into the role
played by each circuit parameter in order to decipher its influence. However, there are
instances when a given data can be fitted by many circuits.In such cases, physical
significance of each element serves as the criterion for choosing the most relebvant
circuit.

WORKED OUT EXAMPLES


1. For the Nyquist plot given below for the reation Cd2+ + 2e- = Cd, estimate (i) the
ohmic resistance, (ii) charge transfer resistance, (iii) double layer capacitance and (iv)
exchange current density at 298 K

(i) Ohmic resistance = 2.5 Ω cm2


(ii) Charge transfer resistance = 20 Ω cm2
(iii) The double layer capacitance at the maximum of the Nyquist plot is
given by
1 1
Cdl = = = 0.1Fcm−2
 Rct 5 2

RT
(iv) Rct =
nFi0

8.314VCK −1  298K
20cm = 2

2  96500Ci0
8.314  298V
 i0 = = 6.42 10−3 C cm-2
2cm  2  96500
2

2. A novel binary alloy synthesized at 500ºC gave the following Nyquist plot. Assuming
that a parallel RC circuit fits the data, estimate the mixed conductance of the alloy.

Rct = 0.08 103 


1
Conductance = = 12.5 10−3 −1
0.08 103

3. Draw schematically the Nyquist plot for the circuit.


4. (A) Identify the maximum frequency in the Nyquist plot shown below and provide its
interpretation (B) Identify the low frequency limit in the plot (C). What is the value of
the phase angle in the low frequency region?

1
(a) max =
Rct Cdl
(b) The straight line depicts ω→0 limit
(c) Phase angle ф = 45º
EXERCISES
1. Draw the equivalent circuits and the corresponding Nyquist diagrams for
(a) Ideally polarizable and (b) ideally non – polarizable electrode/electrolyte
interfaces.
2. Depict the Nyquist plot for an inhibitor which inhibits the corrosion of metals to
nearly 100%
3. If there are n resistors connected in series, write the total impedance.
4. What is the physical significance of the parameter Cdl Rohmic? Cdl and Rohmic
denote respectively the double later capacitance and ohmic resistance.
5. Draw the Nyquist and Bode’ plots for a capacitor.
6. Can the amplitude be made negative in the impedance experiments?
7. Which of the following is/are true?
A Semi circle will appear in the Nyquist plot for
(a) Parallel RC circuit
(b) Series RC circuit
(c) Series RL circuit
(d) Parallel RL circuit
8. Derive the unit for  in equation (21).

FREQUENTLY ASKED QUESTIONS


1. What specific information is deduced from EIS that is not available in any other
electrochemical experiments?
2. Why the applied frequency is varied over a large range such as 10-3 Hz to 106 Hz?
3. Why does imaginary parts arise in experimentally observable parameters?
4. What is the limit under which A.C response will lead to D.C results?
5. What is the role of the amplitude in the excitation potential?
SUMMARY
The fundamental principles underlying EIS are highlighted. A few typical equivalent
circuits pertaining to electrochemical systems are constructed.

You might also like