165
RUHR-UNIVERSITÄT BOCHUM DEVELOPMENT OF THE INTERMODULATED DIFFERENTIAL IMMITTANCE SPECTROSCOPY FOR ELECTROCHEMICAL ANALYSIS DISSERTATION SUBMITTED FOR THE DEGREE OF DOCTOR OF NATURAL SCIENCES (DR. RER. NAT.) FAKULTÄT FÜR CHEMIE UND BIOCHEMIE Zentrum für Elektrochemie - CES BOCHUM, JUNE 2014 ALBERTO BATTISTEL

Development of the intermodulated differential immittance ... · ruhr-universitÄt bochum development of the intermodulated differential immittance spectroscopy for electrochemical

  • Upload
    lyquynh

  • View
    220

  • Download
    0

Embed Size (px)

Citation preview

RUHR-UNIVERSITÄT BOCHUM

DEVELOPMENT OF THE

INTERMODULATED DIFFERENTIAL IMMITTANCE

SPECTROSCOPY FOR

ELECTROCHEMICAL ANALYSIS

DISSERTATION

SUBMITTED FOR THE DEGREE OF

DOCTOR OF NATURAL SCIENCES (DR. RER. NAT.)

FAKULTÄT FÜR CHEMIE UND BIOCHEMIE

Zentrum für Elektrochemie - CES

BOCHUM, JUNE 2014

ALBERTO BATTISTEL

The work presented in this thesis was carried out during my doctoral studies from

November 2010 to June 2014 in the group of Dr. Fabio La Mantia; Center for

Electrochemical Sciences (CES) - Semiconductor Electrochemistry & Energy

Conversion, Ruhr-Universität Bochum.

Date of submission

Chair of examination board

First supervisor

Dr. La Mantia

Second supervisor

Prof. Dr. W. Schuhmann

Dedico questo lavoro di tesi a mio padre

i

ACKNOWLEDGEMENTS

This thesis represents some of the professional achievements I collected in the last

three and a half years, but it would have been not possible without the help and

support of many.

I am extremely grateful to my supervisor Dr. Fabio La Mantia. I must say he was far

more than a simple guide during these years. He helped and supported me inside

and outside the lab. From him I have learnt that dedication, hard work, and

especially studying always pay back. I enjoyed a lot our endless discussions about

science, history, sociology, and politic… I regret I did not have time and energy

enough to put into practice some of the fabulous ideas we discussed. Dr. Fabio La

Mantia is for me the SuperV, an example, and a friend.

I thank Prof. Wolfgang Schuhmann my second supervisor. We first met four years

ago during my master and he gave me the possibility to come back here for my PhD.

He is the person created and keep running the group of analytical chemistry where I

could meet, discuss, and collaborate with many persons in the last years.

I should also mention that I am very thankful for the great effort both my

supervisors put in the last weeks in order to allow me to submit this thesis in time.

I want to thank Bettina Stetzka and Monika Niggemeyer for their patience with me: I

know I am really a disaster about administration!

I would like to thank who helped me in correcting and improving this thesis: Giorgia

Zampardi, Dr. Rosalba Rincon, Andrea Contin, Jan Clausmeyer, Dr. Guoqing Du,

Andjela Petkovic, and Rebecca Straub.

Heartfelt thanks goes to my girlfriend Rebecca Straub who rescues my soul from

science and shows me how important are other things of the life. Thank you!

A special thanks goes to Dr. Mauro Pasta who showed me that you have to be crazy

and determined in the life. You have not to care about what the others think and that

what you do you do it for yourself.

I am thankful to a lot of persons who in these years were close to me. Persons I met

here in Bochum, persons with whom I worked, studied, discussed, joked, hanged

out… A simple list cannot show my thanksgiving: Giorgia Zampardi, Dr. Rosalba

Rincon, Andrea Contin, Jan Clausmeyer, Dr. Guoqing Du, Dr. Aleksandar Zeradjanin,

Mu Fan, Dr. Jelena Stojadinovic, Dr. Edgar Ventosa, Dr. Rafael Trocoli Jimenez,

Bernhard Neuhaus, Dr. Jorge Eduardo Yánez Heras, Dr. Freddy Oropeza, Dr. Mauro

Pasta, Alberto Ganassin, Dr. Magdalena Gebala, Dr. Nicolas Plumere, Alex

Alborghetti, and Dr. Stefan Klink. These persons helped directly or indirectly in my

human and professional development and only a small fraction of what I have shared

with them can be placed on paper.

ii

I want to thank also my flatmates: Dr. Stefan Klink, Christian Sorgenfrei, and Arne

Wege for the nice time shared together at home.

I am grateful to Dr. Nicolas Plumere, Stéphanie Huss and the little Luop for the nice

time spent together, for the crazy Fridays, the walk to the lake, and the long travels

together (which I mostly spent sleeping ). Furthermore, they fed and influenced my

vision and interpretation of the life and of the world.

I want to thank Prof. Salvatore Daniele and all the members of his group who taught

me electrochemistry back in the time in Venice.

Un ringraziamento va alla mia famiglia e in particolare a mio padre. Sono fortunate:

mi é stato insegnato cosa vuol dire usare la propria testa e mantenere uno spirito

critico, ma senza cattiveria, nella vita. Quello che sono e dove sono arrivato lo devo a

voi. Grazie di cuore!

Grazie anche a tutti gli amici in Italia, che mi accolgono sempre a braccia aperte

quando torno a casa anche se non mi faccio mai sentire.

iii

CONTENTS

Acknowledgements .................................................................... i

Contents ................................................................................... iii

List of symbols and abbreviations ............................................ vi

1 Introduction ......................................................................... 1

1.1 State of the art .................................................................................... 2

1.1.1 History of impedance spectroscopy .................................................. 2

1.1.2 Nonlinear analysis .......................................................................... 4

1.1.3 Electrochemical noise ................................................................... 10

1.1.4 Gas-evolving electrodes and bubble effects ..................................... 12

1.2 Motivation and aims ........................................................................... 18

1.3 Final remarks and outlines .................................................................. 20

2 Theory ................................................................................ 22

2.1 Basic concept of electrochemistry ........................................................ 23

2.2 Linear systems and electrochemical impedance spectroscopy .................. 26

2.2.1 Transfer functions ........................................................................ 26

2.2.2 Electrochemical impedance spectroscopy ........................................ 27

2.3 Development of the nonlinear system ................................................... 35

2.3.1 Intermodulation ........................................................................... 36

2.3.2 Definitions and development of new transfer functions ..................... 36

2.3.3 Simplified case ............................................................................ 39

2.3.4 Ideal nonlinear case: diode ........................................................... 40

2.3.5 Electrochemical case: redox couple ................................................ 41

2.3.6 Mathematical simulation of the redox couple ................................... 43

2.3.7 Resistance compensation .............................................................. 53

2.4 Development of the nonlinear time varying system ................................ 55

3 Experimental part ............................................................... 59

3.1 Potentiostat and impedance measurement ............................................ 60

3.1.1 Potentiostat ................................................................................ 60

iv

3.1.2 Impedance measurements ............................................................ 62

3.2 Development of the IDIS instrumental setup ......................................... 67

3.2.1 Lock-in setup .............................................................................. 68

3.2.2 Oscilloscope setup........................................................................ 70

3.3 Electrochemical cell ............................................................................ 74

3.3.1 Capacitor bridge .......................................................................... 74

3.3.2 Coaxial cell ................................................................................. 75

3.4 Materials and procedures .................................................................... 78

3.4.1 Chemicals and electrodes .............................................................. 78

3.4.2 Standard electrochemical characterization and procedures ................ 80

3.4.3 IDIS instrument and parameters.................................................... 81

3.4.4 Potentiostat transimpedance and lock-in amplifier transfer function ... 82

3.4.5 Fitting procedure ......................................................................... 83

3.4.6 Uncompensated resistance correction ............................................. 83

4 Results and discussion ....................................................... 85

4.1 Instrument calibration and artifacts in impedance .................................. 86

4.1.1 Potentiostat transimpedance ......................................................... 86

4.1.2 Lock-in transfer function ............................................................... 89

4.1.3 Cell geometry and capacitor bridge ................................................ 90

4.2 Ideal nonlinear system: diode .............................................................. 96

4.2.1 Diode characterization .................................................................. 96

4.2.2 IDIS of the diode: nonlinear capacitance ......................................... 97

4.2.3 Uncompensated resistance correction ........................................... 101

4.3 Real electrochemical system: redox couple ......................................... 105

4.3.1 Characterization of the redox couple ............................................ 105

4.3.2 IDIS of the redox couple ............................................................. 107

4.3.3 Fitting of the differentials ............................................................ 109

4.4 Time variant system: gas-evolving electrode ....................................... 114

4.4.1 Characterization of the oxygen evolution reaction .......................... 114

4.4.2 Bubble evolution as a time variant system .................................... 116

v

4.4.3 Normalized impedance ............................................................... 119

4.5 Final remarks .................................................................................. 127

5 Conclusion ........................................................................ 129

5.1 Main contributions ............................................................................ 130

5.2 Further development ........................................................................ 133

Appendix ............................................................................... 135

A. Mass transport operator .................................................................... 135

B. Intermodulated differential immittance spectroscopy in impedance format

137

C. Relaxation of the a priori separation of faradaic and capacitive current ... 138

Bibliography .......................................................................... 140

List of publications ................................................................ 150

Patent ...................................................................................................... 150

Published peer-reviewed articles ................................................................. 150

Accepted work .......................................................................................... 150

Works in preparation ................................................................................. 150

Talks at international conferences ................................................................ 151

Posters at international conferences ............................................................. 151

vi

LIST OF SYMBOLS AND ABBREVIATIONS

Abbreviations

AC Alternating current

CE Counter electrode

CPE Constant-phase element

DC Direct current

EFM Electrochemical frequency modulation

EIS Electrochemical impedance spectroscopy

FFT Fast Fourier transform

FRA Frequency response analyzer

IDIS Intermodulated differential immittance spectroscopy

MICTF Modulation of interface capacitance transfer function

NLEIS Nonlinear impedance spectroscopy

PDS Power density spectrum

PSD Phase-sensitive detector

RE Reference electrode

WE Working electrode

Lower case symbols

a Surface area of the electrode [cm-2]

c Concentration [M]

cOx Concentration of oxidized species [M]

cRed Concentration of reduced species [M]

dG Differential conductance [S V-1]

vii

dR Differential resistance [Ω V-1]

dX Differential reactance [Ω V-1]

dY Differential admittance [S V-1]

dZ Differential impedance [Ω V-1]

e Elementary charge [C]

f Generic function -

i Current [A]

i0 Exchange current [A]

ic Capacitive current [A]

iF Faradaic current [A]

il Leakage current [A]

j Imaginary unit -

kB Boltzman constant -

m Mass transport operator [M A-1]

mOx Mass transport operator for the oxidized species [M A-1]

mRed Mass transport operator for the reduced species [M A-1]

n Number of electrons involved in the reaction -

u Potential [V]

Upper case symbols

A Generic vector -

B Susceptance [Ω]

C Capacitance [F]

Cdl Double layer capacitance [F]

viii

CSC Semiconductor capacitance [F]

D Diffusion coefficient [cm s-2]

DOx Diffusion coefficient of the oxidized species [cm s-2]

DRed Diffusion coefficient of the reduced species [cm s-2]

F Faraday constant [C mol-1]

G Conductance [Ω]

H Hessian matrix -

HF Faradaic Hessian -

Im(•) Imaginary argument -

J Jacobian vector -

ND Doping level [cm-3]

R Gas constant; Resistance [J K-1 mol-1]

Rct Charge transfer resistance [Ω]

Re(•) Real argument -

T Absolute temperature [K]

V Different of potential [V]

X Vector of variables -

Y Admittance [S]

Yc Capacitive admittance [S]

YF Faradaic admittance [S]

Z Impedance [Ω]

ZA Normalized impedance [Ω]

Zm Mean impedance [Ω]

ZW Warburg impedance [Ω]

ix

Greek symbols

Symmetry factor -

Symmetry of mass transport -

Double layer capacitance variation [F V-1]

electric permittivity [F m-1]

F m-1

r relative permittivity -

overpotential [V]

a Activation overpotential [V]

ohm Concentration overpotential [V]

conc Ohmic overpotential [V]

Warburg coefficient [Ω rad0.5 s-0.5]

Double layer capacitance time constant [s]

Stimulus angular frequency; stimulus subscription [rad-1]; -

Probe angular frequency; probe subscription [rad-1]; -

Signs

•T Transpose vector -

F Fourier transform operator -

Sign of time derivative -

Sign of time oscillation -

Sign of anti-differential -

•* Complex conjugated -

1

1 INTRODUCTION

The first chapter of this thesis starts with an historical overview of the

electrochemical impedance. Later on, I proceed with the main features of the

nonlinear analysis in electrochemistry. In particular, I describe the characteristics of

the nonlinear impedance spectroscopy, of the electrochemical frequency modulation,

and of the modulation of interface capacitance transfer function technique. By

showing the features of these approaches, I introduce the main problems which are

not considered in these techniques. Most of these problems are addressed in the

development of the intermodulated differential immittance spectroscopy in the

following chapter.

In the last part of the section, I discuss about the electrochemical noise and how this

is related with the bubble generation during electrochemical evolution of gas.

Besides, I report the main achievements concerning gas-evolving electrodes

especially by the means of approaches based on the interpretation of the

phenomenon in the frequency domain.

Later on, I proceed with the motivation which moved this work and the aims I had in

mind in developing and applying the concept of intermodulation. In particular, I show

a couple of examples where the electrochemical impedance spectroscopy cannot

provide all the information of the investigated system, which is the starting point of

this thesis.

Besides, a short outline of the thesis is reported. In this outline, the most important

parts of the work are listed.

2

1.1 STATE OF THE ART

The first section contains a brief review of the state of the art. Initially, I shortly

report the history of the impedance spectroscopy. In the second subsection, I discuss

about the nonlinear analysis in electrochemistry. In particular, I speak of nonlinear

impedance spectroscopy, of electrochemical frequency modulation, and of the

modulation of interface capacitance transfer function technique. The latter represents

the starting point of this thesis. I describe which nonlinear parts of the

electrochemical systems are usually considered and in which way. This part is

connected with the development of the theoretical part for the intermodulation in the

second chapter (Section 2.3) and with the results concerning the diode and the redox

couple in the fourth chapter (Section 4.2 and 4.3).

In the third subsection, I proceed speaking about the electrochemical noise and the

problems concerning performing electrochemical impedance spectroscopy in these

conditions. The last part discusses the electrochemical noise generated by bubble

evolution during electrochemical production of gas as for example hydrogen and

oxygen. I describe which models are available and which effects the gas phase has in

regard to the electrochemical reaction. In the second chapter (Section 2.4), I report

which approach was used in this work to explore the gas-evolving electrode during

oxygen bubble formation.

1.1.1 HISTORY OF IMPEDANCE SPECTROSCOPY

In the late nineteenth century, during his studies on telegraphy and electrical

circuits, Heaviside introduced what later became the basis for operational calculus

[1]. The extraordinary achievement was accomplished through Laplace transforms.

This allowed converting differential equations into algebraic equations. Every electric

element of an electric circuit can be written as a simple equation in which the

variable is the Laplace frequency or the angular frequency. Table 1–1 shows these

relationships. As one can see, the differential equations, defining the potential or the

current in the second column are converted into simple equations in the third and

fourth column. These equations are called transfer functions and are the base for the

operational calculus.

The fact that the names of these transfer functions such as impedance, admittance,

and reactance, which were given by Heaviside, are still used nowadays, underlines

the importance of his approach. Bode much later introduced the concept of

immittance to characterize both impedance and admittance [2]. Furthermore, the

conversion of the Laplace frequency into the imaginary frequency, used in the Fourier

3

transforms, is named Heaviside transform. The Fourier transform remains the most

used transform in the electrochemical impedance spectroscopy (EIS).

This was the starting point for the impedance spectroscopy, which later on was also

applied to the study of physical systems by Nernst [1]. EIS has several advantages.

It is based on the linear theory system, a well-developed and solid framework, it can

deliver a large amount of information within a single experiment, and it contains an

implicit validation method, the Kramer and Kronig relations, which allows estimating

or correcting possible sources of errors [1]. The concept of transfer function and

linear system is discussed in the second chapter (Section 2.2).

Another advantage of the EIS is the use of electrical analogues, also known as

equivalent circuits. These are composed by simple elements like resistors, capacitors

and allow an easy interpretation of a system. However, these analogues are only

useful when associated to some physical-chemical properties of the system, that is

when they represent a suitable differential equation. This is indeed a recognized

problem and some equivalent circuits are used only because they fit the response of

an electrochemical system [1].

Probably the most famous equivalent circuit is the one proposed by Randles to

represent an electrochemical reaction at the electrode interface. This is represented

in Figure 1–1. This equivalent circuit considers the main features of a charge transfer

reaction. The resistance of the electrolyte Re represents the ohmic drop in the

electrolyte, the double layer capacitance Cdl describes the accumulation of charged

Table 1–1: equations that describe some elements used in the electrochemical

impedance spectroscopy according to their representation in the time, Laplace,

and Fourier domain.

Electric

component

Differential

equations

in the time domain

Algebraic

equations

in the Laplace

domain

Algebraic equations

in the Fourier

domain

Resistor

Capacitor

Inductor

Warburg

element –

4

species at the double layer, the charge transfer resistance Rct expresses the

electrokinetic limitation of the faradaic reaction, and the Warburg impedance ZW

considers the limitation imposed by the diffusion of the electroactive species at the

electrode.

In 1899 in the work titled ―Über das Verhalten sogenannter unpolarisirbare

Elektroden gegen Wechselstrom‖ (About the behaviour of non-polarizable electrode

in alternating current), Warburg solved the diffusional transport of electroactive

species at the electrode [3]. From the Fick laws he derived the diffusional impedance

which still bears his name. This is reported in the last row of Table 1–1. The time

representation of these differential equations is too large to fit into the table.

In this subsection, I shortly described the advantages of the electrochemical

impedance spectroscopy and I named who gave some of the most important

contributions. In the next subsection, I introduce the nonlinear impedance

spectroscopy and the main aspects of this technique.

1.1.2 NONLINEAR ANALYSIS

In this subsection, I give a short overview of the nonlinear analysis in

electrochemistry, in particular, how this appears as a natural extension of the EIS.

Following, I draw a mild distinction to generalize the aspects of several different

approaches in order to guide the reader through their strong points and their

weaknesses.

In the equivalent circuit shown in Figure 1–1, only Re is a linear element, whereas

the others show a potential dependence. In fact, Rct derives from the equations which

controls the electrokinetics which is dependent on the potential; Cdl is also function of

Figure 1–1: schematic of the Randles circuit, where Re stands for the resistance of

the electrolyte, Rct for the resistance of the charge transfer, ZW for the Warburg

element, and Cdl for the capacitance of the double layer.

5

potential as shown by Grahame [4]; and ZW which represents the diffusion limitation

is dependent on the current which is function of potential.

During electrochemical impedance spectroscopy, the system is perturbed by a

sinusoidal potential perturbation and the current responds with a periodic transient.

In the case of a linear system the current is a sinusoidal wave as well. However, in

the case of a nonlinear system, the current is not a pure sinusoidal oscillation, but

contains other waves. These are named harmonics and represent the nonlinearity of

the system. The harmonics oscillate at an integer multiple of the frequency of the

fundamental wave. Figure 1–2 shows the current transients of a diode during EIS in

conditions where the current is a strong nonlinear function of the potential. When the

potential perturbation is small enough (5 mV) the diode can be well-approximated by

a linear system which shows only a pure sinusoidal wave (Figure 1–2–a). Whereas,

when the perturbation is large (50 mV), the system departs from linearity and

several harmonics appear (Figure 1–2–b). One can notice that the impedance already

changes passing from 5 to 20 mV of amplitude perturbation (Figure 1–2–c). In fact,

EIS is based on the principle of linearity which neglects the interference of the

harmonics. This point is discussed in more details in the second chapter (Section

2.2).

Of course, considering an electrochemical system as linear is an approximation which

is valid only for small perturbations. Furthermore, the linearization procedure

discards some pieces of information regarding the system which come from the

harmonics.

Senda and Delahay [5], Neeb [6], and Kooyman et al.[7] were among the first to

employ nonlinear analysis in electrochemistry. The main reason for it was to recover

more information compared to what was normally accessible in a linear framework.

Although there is a large number of studies concerning nonlinear analysis conducted

over the last five decades, a more or less clear separation can be made on the nature

of the input signals. Wherever the input occurs under the form of potential or

current, this can be composed by a single perturbation falling under the name of

nonlinear impedance spectroscopy (NLEIS) and sporadically harmonic impedance

spectroscopy [8–37]. Whereas, when composed by two signals, it is mainly known in

literature as electrochemical frequency modulation (EFM) [38–53] and modulation of

interface capacitance transfer function (MICTF) technique [54–57].

6

In the NLEIS, the authors look at the higher harmonics given by a sinusoidal

perturbation to find additional valuable information or to correct the errors arising

from the oversimplified linearization of the system. One can understand this point

Figure 1–2: comparison of linear and nonlinear system. a) current transient of a

nonlinear system well-approximated to a linear system by the small perturbation

amplitude (5 mV); b) current transient of a nonlinear system with a large

perturbation amplitude (50 mV); c) impedance of the system as function of the

amplitude of the potential perturbation.

7

looking at Figure 1–2. The presence of the harmonics in Figure 1–2–b influences the

impedance of Figure 1–2–c. These harmonics can be used either to correct the

impedance spectrum and recover the impedance as if there were no distortions or to

grasp more information. In fact, the linearization is an approximation mainly

employed to reduce the complexity of the system. Once this approximation is relaxed

the system can be approached more accurately. In electrochemistry these extra

information regard some parameters of the system EIS cannot recover directly. For

instance, the EIS is often used to obtain the free corrosion current of a corrosion

reaction, but it does not give directly the value of the Tafel slopes. However, with the

nonlinear analysis these are also available.

Contrary to the NLEIS, in the EFM and the MICTF technique the system is perturbed

by two sinusoidal potential signals and the main source of information are the

intermodulation sidebands, which are shown in Figure 1–3–a. The perturbations

create an amplitude modulation of the current which is used to recover information

from the system. The current transient of an amplitude modulation is exemplified in

Figure 1–3–b.

The EFM is only applied in the field of corrosion research, where it is a powerful and

convenient tool to recover the Tafel slopes of the system. In fact, these can be

simultaneously obtained with a single experiment at the OCP. Additionally, a

validation procedure was suggested based on the relation between the

intermodulation sidebands and the second and third harmonic of the current [39].

Their ratio was called causality factors. In spite of its large use, the lack of an

adequate theoretical framework and of complex analysis (only the modulus of the

current is considered) hindered its further development. In fact, resistance of

electrolyte, double layer capacitance and mass transport are systematically

neglected.

The complex analysis is fundamental in the impedance spectroscopy. Indeed, it

permits to consider the delay the current transient shows compared to the potential

perturbation. This delay is correlated with some capacitive behavior of the system as

for instance given by the charging of the double layer. Also the mass transport

produces some capacitive effects. By discarding the complex analysis, these

phenomena are difficult to account for.

8

The MICTF technique, introduced by Keddam and Takenouti [54–57], deserves a

classification as a standalone methodology, although it shares the perturbation

system composed by two signals with the EFM. In the EFM, the focus was on the low

frequency faradaic component of the current without any consideration about the

non-faradaic one. Whereas, given a new transfer function, the MICTF focuses only on

nonlinearities of the capacitive components, which appear at high frequency. In

particular, the aim was to recover the time constant of the interfacial capacitance.

Another characteristic feature was the original instrumental setup composed by a

lock-in amplifier. This was used to demodulate the amplitude modulated current on

the high frequency carrier. Apparently, this technique developed completely

independently of the EFM. This explains the large discrepancy in respect to the EFM.

The MICTF was the inspiring idea for this thesis, in which a new class of generalized

transfer function, capable to interpret both faradaic and non-faradaic terms, are

introduced [58].

Figure 1–3: representation of intermodulation. a) intermodulation sidebands in the

case of a diode measured with two perturbations of 10 and 1000 Hz; b) simulated amplitude modulation of a 10 Hz signal at 1 Hz.

9

The main fields of application of the NLEIS were corrosion [10–12,15,18,23–25,29]

and fuel cells [26,30,59,33,60,31]. Given a very solid theoretical background, several

systems were investigated. A diode has shown to be a good benchmark for the NLEIS

and was employed several times to simulate the exponential dependency of the

current on the potential [12,13,32]. The primary focus was on tafelian systems

[9,12,13,17–21,23,36], but several works dealt also with reversible systems

following the Butler and Volmer equation [7–9,16,27,28,37], and adsorptions or

multistep reactions [14,20,21]. Despite the large amount of literature, many authors

neglected the influence of the resistance of electrolytes, double layer, and mass

transport.

The double layer was mainly either considered linear (or with a negligible

nonlinearity) or neglected in the treatment using the low frequency approximation.

These limitations were overcome by some authors, which addressed the dependency

of the double layer capacitance on the potential [5,9,35,61–66]. Dickinson and

Compton considered, that the non-faradaic current is also dependent on the

frequency i.e. on time [67]. This point was the reason for the development of the

MICTF technique.

Although the resistance of the electrolyte is a linear term, i.e. it does not depend on

the potential, its presence affects the observation of the nonlinearities. The nonlinear

components of the current, the harmonics or the intermodulation sidebands, are

dumped and phase-shifted by other linear elements. Any nonlinear term can be

described by a virtual potential source which produces a current [64]. This current,

passing through a passive resistance, generates an ohmic drop opposite to the virtual

source. Diard et al. considered this problem as a source of error [9].

Mass transport is hardly considered in the nonlinear analysis. Although its effect can

be negligible, it does affect the system, especially when electrokinetics plays a minor

role. Xu and Riley computed the diffusion in their treatment numerically [37]. Some

other authors did solve it analytically [7,9,16] and Darowicki used their results to

calculate the diffusion coefficients of one of the redox species which were involved in

the reaction [8].

The lack of an adequate mathematical framework in the nonlinear analysis often

leads to problems and some authors tried to shape the nonlinear terms as impedance

[16,28]. This is fundamentally wrong, because second terms cannot have the same

dimension of the linear terms (i.e. the resistance is measured in Ohm, but its

variation on the potential is measured in Ohm V-1). A second possibility is to consider

only the current instead of a transfer function. This was done, for example, by Xu

and Ripley [35,37] and is common used with the EFM. However, doing so prevents

from performing a complex analysis, which results in a large loss of information. In

10

this case, the plots always show the amplitude of the current as a function of the

frequency. The meaning of the transfer function lies also in normalizing the values:

for example, in the case where the potential perturbation is not constant in all the

frequency range, the current amplitude is not only the function of the pure answer of

the system, but also of some instrumental conditions. Therefore, a normalizing

procedure is required.

A mathematical and theoretical framework was introduced in the 1970s by

Rangarajan. He showed an elegant matrix formalism on which he built a theoretical

framework to embrace all kinds of phenomena in electrochemistry under the view of

impedance or admittance [68–71]. He showed ―how to arrive at the results for even

the most complex models without tears‖ [68]. Later on, he expanded his framework

to consider nonlinear systems [62,63]. Instead of Taylor series, he employed Volterra

series which give a memory effect to the series expansion. Despite his elegant and

sophisticated approach, this proved to be too complex.

In this subsection, I showed a brief historical overview of the nonlinear analysis in

electrochemistry and its link with EIS. Some general limitations and approximations

were discussed according to the distinction between the single input approach,

nonlinear impedance spectroscopy (NLEIS), and the double input approaches,

electrochemical frequency modulation (EFM) and modulation of interface capacitance

transfer function (MICTF) technique. In particular, the specular character of the last

two techniques was underlined. In the end, a small parenthesis was opened on the

necessity and advantage of having a proper mathematical and theoretical framework.

In the next section, the concept of EIS and nonlinear analysis is expanded to the field

of time variant systems. Also, a short overview of the gas-evolving electrode is

presented and discussed in the framework of impedance spectroscopy.

1.1.3 ELECTROCHEMICAL NOISE

Almost a century ago, Morgan observed the formation of bubbles while adding formic

acid to concentrated sulphuric acid [72]. He recorded the pressure fluctuations and

carefully described his findings. The reaction initially proceeded with an oscillatory

character and then reached a steady-state condition when the reactant was

consumed. This phenomenon was later called gas oscillator [73–76]. Much later,

Barker was one of the first to describe flicker noise in electrochemistry [77,78]. This

was produced by the hydrogen evolution reaction on a mercury electrode. He could

not properly explain the source of this noise which appeared as fluctuation of the

recorded current, but the term noise remained in the literature meaning a non-

stationary behavior of a reaction giving rise to periodical or stochastic variation in the

11

recorded signal (current or potential). He also suggested the use of electrodes of

small size and believed that the study of noise could give important mechanistic

information.

For certain, the effect of a gas evolution on the current, which begins to fluctuate

because of the birth, growth and departure of many fine bubbles on the surface of

the electrode, is well known. This is not only a source of disturbance, which

deteriorates the quality of the recorded data, but also limits the possibilities to

perform certain experiments. Appearing of noise in a cyclic voltammetry affects the

quality of the result, but the collected data can still be interpreted in a meaningful

way. The case of EIS is different. This technique requires a suitable steady-state

condition to operate, which is often missing in the case of bubble formation on the

electrolyte-electrode interface. Besides being an instrumental requirement (the

instrument would return unreliable values), a steady-state condition is also dictated

by the theory of linear time-invariant systems, which is the base of EIS [1]. In fact,

several authors reported the problem connected with performing EIS in presence of

gas evolution [79–83]. Figure 1–4–a is an example of this problem. It shows the

Nyquist plot of an impedance measurement performed on a cavity microelectrode

filled with ruthenium oxide during the generation of oxygen bubbles. At high

frequencies the variation of the current due to the gas-evolution was slow compared

with the measurement perturbation, the system was well approximated by a steady

state, it was steady as ‖seen‖ by a quick perturbation. There was a condition stable

enough in the operating time scale. When the frequency decreased approaching the

frequency range of the noise the steadiness was lost and the data appear scattered.

In this case, the impedance is a time-variant quantity and the EIS has no meaning in

this condition [1].

Steady state means that all the properties of the system are not changing in a

suitable lapse of time, i.e. for potentiostatic control the current is flat. When instead

bubbles are produced at the electrode the current shows periodic, pseudo-periodic,

or stochastic variations as reported in Figure 1–4–b, which is an example of a current

transient measured at the gas-evolving electrode. A strategy to recover some control

over these phenomena is to partially relax the requirement of steady state to

embrace the concept of periodic steady state [84–87]. In this framework, the

properties of the system are allowed to vary, but their variation has to be periodic, so

that the behavior of the system can still be predicted.

12

1.1.4 GAS-EVOLVING ELECTRODES AND BUBBLE EFFECTS

In this subsection, the specifics of the electrochemical noise produced by a gas-

evolving electrode are shown and the mechanism of bubble formation from a

supersaturated solution is briefly stated. I explain the effect of a bubble on the

electrode surface through its local influence on the current lines and on the

overpotentials. Several models and approaches are shown to attack the problem.

In general, bubbles form in a supersaturated solution according to four mechanisms:

classical homogeneous nucleation, classical heterogeneous nucleation, pseudo-

classical nucleation, and non-classical nucleation [88–91]. The concept of classical

nucleation was connected with the fact that those nucleations needed an activation

energy, conceived from a classical point of view [88]. The first two mechanisms did

not require the existence of a gas cavity, which represents a discontinuity on the

Figure 1–4: effect of the electrochemical noise produced by gas bubble evolution.

a) EIS during electrochemical noise; b) current transient of the electrochemical noise.

13

surface in contact with the solution which provides a small concavity with a small gas

pocket. However, the curvature radius of these gas pockets is a key point of

understanding nucleation in gas evolution reactions. If this radius is smaller than the

critical curvature radius for bubble nucleation, this is the pseudo-classical nucleation

mechanism, the gas phase preferentially forms at the cavity because less activation

energy is required there.

Without a gas cavity, the bubble would form only at very high levels of

supersaturation [91]. In the case where the curvature radius of the gas cavity is

bigger than the nucleation radius, this is referred to as the non-classical nucleation

mechanism: no activation energy is required to form a bubble leading to a

spontaneous evolution at that spot. These last two mechanisms of bubble nucleation

are of primary importance, because they link the morphological properties of an

electrode to its efficiency in evolving a gas phase.

Of course, a gas evolving electrode produces a great supersaturation in its proximity

and this leads to the formation, growth and departure of bubbles. These bubbles

have several effects on the electrochemical reaction, on the electrode surface, and on

the mass transport of the products. This point is sometimes mentioned as

macrokinetics [92].

The total overpotential at the electrode surface is given by the sum of the ohmic

overpotential, the activation overpotential, and the concentration overpotential. The

first one is given by the resistance of the electrolyte, the second is connected with

the activation energy needed to drive the reaction and the third is connected with the

energy required to drive the mass transport of the reactant and product at the

surface. All three overpotentials are affected by the presence, birth, growth and

departure of bubbles at the electrode surface.

The ohmic overpotential is given by the ohmic drop produced by the passage of

current through the electrolyte and it is strongly dependent on the section of solution

directly in contact with the electrode. Several studies addressed how the ohmic drop

is influenced by the presence of bubbles on the electrode surface. Most of the works

about influence of bubbles on the gas-evolving electrodes concern with this aspect

[93–99] that is usually referred to as bubble curtain or screen effect. The bubbles

form and stick to the surface of the electrode creating an insulating layer, a curtain,

which decreases the cross-sectional area of the electrode in contact with the

electrolyte. The presence of bubbles in the curtain in industrial application is so large

that there is more gas in this layer than dissolved in the remaining solution between

the two electrodes [94]. Moreover, the bubble is strongly bond to the surface and it

is hardly removed by convection means[79].

14

The screening effect is particularly important at high current density where the ohmic

overpotential plays the dominant role [100,101]. This is the case of industrial

generation of chlorine, where the current density is in the order of 5 kA m-2 [102]. To

control the characteristics of the bubble curtain is of primary importance in this

sector. It is noteworthy to consider the example of the dimensionally stable anode

(DSA®) for chlorine evolution. They are considered the biggest industrial

breakthrough of the last fifty years [102,103]. Because of the decreased bubble

screening effect they brought down the associated overpotential considerably and

allowed for a more compact cell design, reducing the distance between the electrodes

and consequently further lowering the ohmic drop.

Gabrielli and Huet and Bouazaze et al. studied the effect of an insulating body

standing on the electrode surface [98,100,104]. They mimicked a bubble by a glass

sphere. They found that the variation of the ohmic overpotential was proportional to

the square of the size of the obstructing object.

As for the ohmic overpotential, the activation overpotential is also influenced by the

screening effect. The presence of a bubble on the electrode surface decreases the

actual area available for the electrochemical reaction, increasing the local current

density. The effective electrodic area shrinks because of the bubble curtain. As in the

case of the ohmic overpotential there is direct proportionality between the magnitude

of the activation overpotential and the square of the size of bubbles adhering to the

electrode [98,100,104].

Although the law that links the supersaturation concentration with the concentration

overpotential was known [95,105,106] and the level of supersaturation was known

as well [93,105,106], this overpotential was difficult to evaluate. In fact, the bubbles

change the distribution of the concentration overpotential, as explained later.

The concentration overpotential effect was shown first by Dukovic and Tobias [95].

They casted a model to describe how the current line distribution was influenced by

the presence of a bubble on the electrode. They found that the primary current lines

were concentrated in the bubble proximity, because of the screening effect. They

accounted for adsorption of dissolved gas from the solution by the gas phase and

found the supersaturation profile given by the bubble (see Figure 1–5). The

concentration of dissolved gas was low in proximity of the bubble and increased with

the distance from the bubble, in accordance with the prediction of Vogt [93]. The

bubble acts as a sink keeping the concentration of the dissolved gas lower.

Therefore, the reaction encountered lower concentration overpotential in proximity of

the gas phase and there the reaction rate was larger. They called this phenomenon

―enhancement effect‖ and concluded that this effect was large at lower current

density, while the ohmic overpotential had the dominant weight at large current

15

densities. Figure 1–5 is a schematic of the result predicted by the model of Dukovic

and Tobias. The bubble is surrounded by a boundary layer in which the concentration

of dissolved gas in solution is in equilibrium with the gas phase. The point in which

the tertiary current density is the highest is where the gas phase touches the

electrode.

Other authors dealt with the concentration overpotential as well [93,100,106]. The

main problem was related with the difficulty of measuring local quantities. In fact, a

way to measure the supersaturation level next to a bubble was not available.

Therefore, only averaged quantities were used. Vogt showed that the concentration

overpotential depends not only on the current density, but also on the bulk

concentration of dissolved gas and especially on the mass transport [106].

Some authors focused on the mass transport produced by the bubbles [107–112].

These can stir the solution by several means: growing, detaching, and, in the case of

vertical reactors, flowing parallel to the electrode. It is usually referred to as

microconvection, when the mass transport is locally induced by the growing and

detaching bubbles, whereas the term macroconvection is often used for the collective

stirring produced by columns of bubbles [108]. The two phenomena are closely

related, because the latter can influence the growth of the gas phase and, therefore,

indirectly the microconvection. Müller et al. considered the quantity of products

transported away from the electrode in the gaseous state [109]. They noted that the

Figure 1–5: schematic representation of the effect of a gas bubble on the concentration of dissolved gas and on the concentration overpotential.

16

rate of transport of products from the electrode carried by the bubbles was higher

than that performed by conventional mass transport of dissolved species in solution.

Mass transport affects the concentration overpotential, but the detachment of the

bubble from the surface has the additional feature of freeing the surface of the

electrode. The ohmic and activation overpotential drops instantaneously, but the

enhancement effect is lost, the volume previously occupied by the bubble is replaced

by supersaturation solution and the concentration overpotential rises [100].

All the overpotential showed to fluctuate with time because of the stochastic behavior

of the gas evolution. Although the equations, which controlled the variations of the

overpotentials, were known [96,97,99], the experimental separation of all the

contributions was challenging [96,97].

Huet at al. studied the gas evolution in the frequency domain converting the output

(current or potential) by the means of the Fourier transform into its power density

spectrum (PDS) [99,101,113]. They found a particular shape of the spectrum given

by the bubble noise. Starting from low frequencies, the shape was flat up to a cutoff

frequency and then rolled down with a certain slope; very similar to the shape of a

low-pass filter. They were able to link the cutoff frequency and the roll-off slope to

some mechanistic parameters of the gas evolution as average bubble size and mean

departure time [113].

Gabrielli et al. showed also that EIS could be performed on a gas-evolving electrode

[114]. They employed white noise as excitation, consisting of several frequencies

injected at once. The acquisition time for this technique is longer than for a single-

frequency impedance spectroscopy (although, since all the frequencies are swept at

once, the total time for the experiment is shorter). A FFT algorithm recovers all the

signals [115,116]. In the experiment of Gabrielli et al., since the recorded time was

longer and dictated by the lowest frequency, it happened that at least one complete

period of the pseudo-periodical bubble detachment was recorded allowing for a

proper demodulation of the signals.

Gabrielli et al. were able to follow the concentration overpotential given during

hydrogen evolution [100]. They found that at high current density the bubbles had

an ohmic contribution. In fact, in these conditions the bubble effects were associated

with screening the electrode surface. At low current density, there were two

contributions. The bubbles contributed to the ohmic and concentration overpotential

in opposite direction. They confirmed the prediction of Dukovic and Tobias who found

that at low current densities the bubble enhanced the reaction rate because of the

decreasing the supersaturation level of dissolved gas in proximity of the electrode

surface [95].

17

In this subsection, I gave an overview of the phenomenon of bubble formation at the

electrode-electrolyte interface. I described the mechanism of nucleation of the gas

phase and how this is related with the morphology of the electrode. Furthermore, I

reported the effects of the bubble dynamics on the overpotentials and the mass

transport. In particular, I showed that the activation and ohmic overpotential always

increase because of the bubble effect. Whereas, the concentration overpotential

decreases because of the enhancement effect. One problem was that these

overpotentials change in time and space and only averaged quantities were usually

available.

Besides, the bubbles affect the mass transport. In fact, through micro- and

macroconvection they influence the transfer of products. It also happens that the

bubble itself carries more efficiently the products away from the electrode interface,

hence decreasing the concentration overpotential.

Therefore, the gas phase at the interface has a double role in the electrochemical

reaction. It can either hinder or enhance the reaction rate. Several approaches were

proposed by the authors. Some valuable observations came from those approaches

which looked at the phenomenon in the frequency domain (PDS). In fact, a periodic

or pseudo-periodic event is more meaningful when decomposed in its characteristic

oscillations. However, the fact that the phenomenon was non-steady played several

problems for the main frequency domain approach: the EIS.

18

1.2 MOTIVATION AND AIMS

The main aim of this doctoral thesis was to introduce, characterize, and apply a new

electrochemical technique, the Intermodulated Differential Immittance Spectroscopy

(IDIS), designed to implement the electrochemical impedance spectroscopy (EIS).

The IDIS is based on the phenomenon of the intermodulation which appears when

two periodic stimuli with different frequencies interact in a nonlinear system creating

an amplitude modulation. As described in the state of the art (Subsection 1.1.2)

similar approach was used by other two techniques: the electrochemical frequency

modulation (EFM) and the modulation of interface capacitance transfer function

(MICTF) technique. The former aimed only at the faradaic process, whereas the latter

investigated only the double layer response. The IDIS merges both features and

represents a generalization of the two.

I developed the IDIS because this technique gives a deeper insight than a simple

electrochemical impedance spectroscopy. In fact, the symmetry of energetic barrier

and of mass transport, the time constant and the variation of the double layer

capacitance are all available. Usually several experiments are necessary to recover

these information (if possible) whereas with the IDIS, these can be obtained in a

single spectrum. This is not only an achievement concerning the number of

experiments one has to perform: for example, to recover the symmetry factor of an

electrochemical reaction either the potential or the concentration of the redox species

in solution must be changed. With the IDIS this is not necessary. Therefore, the

assumption that the symmetry factor is neither dependent on potential nor

concentrations is required.

In particular, the time constant of the double layer capacitance and the symmetry of

mass transport are usually not achievable through EIS. Furthermore, these

parameters were usually overlooked in most nonlinear analysis. However, because of

the nature of the intermodulation, the IDIS can easily recover them.

In order to develop the IDIS, a mathematical framework for the intermodulation is

necessary. This is required by the complexity of the nonlinear analysis. However, a

virtue is made out of necessity. The aim is to provide a general and elegant model

which can be used later for future development. This framework has to extend that of

the impedance spectroscopy where the use of transfer functions allows a handy

employment of several differential equations. Furthermore, it has to be suitable to

account for all the nonlinearities of the faradaic and capacitive current and to

consider the effect of the resistance of the electrolyte.

The electrochemical noise generated by the dynamics of gas bubbles evolution during

electrochemical generation of oxygen is a challenging system. In this condition a

19

reasonable steady state is never reached and the electrochemical impedance

spectroscopy can be only partially applied. However, when the noise is periodic the

system response is similar to that of an intermodulation. Starting from the

mathematical framework of the IDIS, the aim is to create a model to parameterize

the system in order to understand the effect of the dynamicity of the bubble

evolution on the electrochemical reaction. In particular, it is important to understand

the balance between the enhancement effect and the screening effect given by the

bubbles on the electrode surface. Besides, the effect of mass transport given by

microconvection is also important and it can be identified with a proper complex

analysis.

In the next section, I summarize what presented so far and I give a brief outline of

the thesis.

20

1.3 FINAL REMARKS AND OUTLINES

In the first chapter with the state of the art, I introduced the main aspects of the

nonlinear analysis (Subsection 1.1.2). I drew attention on how other authors dealt

with the description of the faradaic and capacitive current. In particular, I showed

that although the electrokinetics was subject of several investigations the mass

transport and the double layer were often disregarded. Besides, also the resistance of

the electrolyte was mainly neglected.

In the Subsection 1.1.4, I described the phenomenon of bubble evolution at the

electrode-electrolyte interface. In particular, I showed the effects of the gas phase on

the electrochemical evolution of gas and how this was observed by other authors.

In the second section, the motivation and the aims of this work were reported. In

particular, in which way the intermodulated differential immittance spectroscopy

(IDIS) represents a generalization of the electrochemical frequency modulation (EFM)

and the modulation of interface capacitance transfer function (MICTF) technique. The

aim of the IDIS is to recover information concerning the electrokinetics, the mass

transport, and the double layer. Besides, the intermodulation can be used to

parameterize the gas-evolving reaction when there is periodic formation of gas

bubbles.

In the next chapter, the basis of electrochemistry necessary for this work is reported.

Furthermore, the mathematical framework in which the intermodulation is developed

is described. One can find the model used to consider the faradaic and the capacitive

current in the Subsection 2.3.5 and 2.3.6. The parametric model to study the gas-

evolution reaction is reported in the Section 2.4 which discusses in general the

nonlinear time varying systems. In particular, the model shows how is possible to

normalize the impedance by the surface area really available for the electrochemical

reaction.

In the third chapter, I give an overview on the instruments, materials and procedures

employed in this thesis. In particular, the instrumental setups developed for the IDIS

and the expedients employed to mitigate the artifacts arising during the impedance

measurements are shown in Section 3.3 and 3.4, respectively.

The results are described in the fourth chapter. The application of the model

developed in the second chapter (Subsection 2.3.5 and 2.3.6) is reported and

discussed in the Section 4.3 where a best fit is used to recover the parameters of the

electrokinetics, mass transport, and double layer. Following the model reported in the

second chapter (Section 2.4), the results of the parameterization of the gas-evolving

electrode are described in the Section 4.4. In particular, the impedance normalized

by the surface area of the electrode free from gas phase is shown.

21

In the last chapter, I summarize the main contributions of this thesis and suggest

some future developments. In particular, I propose some modification of the IDIS

where the intermodulation can be also used with non-electric perturbations.

22

2 THEORY

This chapter contains all the definitions and the models developed in this work. In the

first section, I briefly introduce some concepts of electrochemistry necessary during

the elaboration. In particular, I describe the faradaic and capacitive current and the

basic equations which control them.

Later on, I explain the linear system theory and how this is connected with the

electrochemical impedance spectroscopy. Starting from the definition of transfer

function, some examples of applications of impedance spectroscopy are presented.

These are the starting points for the further development of the nonlinear treatment

and intermodulation.

In the third section, I introduce the concept of nonlinear system and show the origin

of the intermodulation. Following the examples of the second section, I describe the

model developed for intermodulated differential immittance spectroscopy (IDIS)

using three different cases. In the last example, the IDIS spectra are mathematically

simulated. These simulations are used as comparison and pattern recognition later in

the fourth chapter (Subsection 4.2.3 and 4.2.4).

In the last section of this chapter, I discuss about the nonlinear time varying system.

In particular, I introduce the case of gas-evolving electrodes where the periodic

formation, growth, and departure of a gas bubble obstruct the electrode surface.

Recalling the concept of periodic steady state and linear parametric varying system, I

show the model developed in this work to recover the effect of the bubble dynamics

on the electrochemical reaction.

23

2.1 BASIC CONCEPT OF ELECTROCHEMISTRY

Electrochemistry concerns with the study of the heterogeneous reduction and

oxidation (redox) reaction and accumulation of charged species at an electrode-

electrolyte interface. The electrode can behave as an inert spectator of the reaction,

in which case it solely provides the polarization of the interface, or being actively part

of the electrochemical reaction as in the case, for example, of metal dissolution. The

electrolyte is an ionic conducting phase, usually liquid, composed by a solution with

some dissolved ionic solutes.

Through the use of a potentiostat, whose working principle is reported in the third

chapter (Subsection 3.1.1), it is possible to finely control the potential difference

between the electrode and the electrolyte, without considering the treatment of an

additional electrode.

The electrode, referred as working electrode, is usually composed by a metallic phase

which cannot support an electric field inside. For this reason the electrons, needed to

balance the difference of potential, accumulate in a nearly infinitesimal region at the

metal boundary.

Compared with the electrode, the electrolyte has a finite and low amount of charged

species. These accumulate at the electrolyte interface forming the so-called double

layer, which has a finite thickness. The inner layer, represented by the centre of

mass of the desolvated ions adsorbed to the metal surface, is called Inner Helmholtz

plane (IHP). The outer Helmholtz plane (OHP), instead, represents the centre of

mass of the solvated ions. This is the layer of the closest proximity to the electrode a

nonspecifically adsorbed species can reach. Beyond the OHP, the charged species

necessary to balance the charge at the electrode are distributed because of thermal

agitation in a three-dimensional region called diffuse layer. This extends from the

OHP to the bulk of the solution. The total charge given by the double layer is equal

and opposite in sign to the charge accumulated on the metal side.

When the electrode potential is changed a flow of current arises at the interface

because of migration of the charged species in the double layer. This current which

has a transient nature is called capacitive or non-faradaic current and is given by the

accumulation and depletion of ions in the double layer. The capacitive current ic is

given by:

dt

dCV

dt

dVC

dt

dQi dl

dldl

c (2.1)

24

Where Qdl is the charge accumulated in the double layer, Cdl is the differential

capacitance of the double layer, which is potential dependent, V is the difference of

potential between the working electrode and the solution and t is the time. If the

dependence of Cdl on the time is neglected the capacitive current becomes dependent

only on the variation of the potential, which is the usual assumption.

When a redox reaction occurs a faradaic current flows. This is related to the electron

transfer between the electrode and one or more species in solution.

The law which binds potential and current is historically known as Butler and Volmer

equation:

aa0

RT

nFexp

RT

nF1expii (2.2)

Where i0 is the exchange current density, n the number of electrons involved in the

reaction, the transfer coefficient or symmetry factor, ηa the activation

overpotential, and F, R, and T have the usual meaning. This equation assumes that

the concentration of the electroactive species does not change at the electrode

interface. To allocate this change the Butler and Volmer equation has to be

substituted with the current-overpotential equation:

RT

nFexp

c

c

RT

nF1exp

c

cii

b,Ox

0,Ox

b,dRe

0,dRe

0 (2.3)

Where c represents the concentration of the electroactive species, the subscriptions

Red and Ox stand for reduced and oxidized species, and 0 and b for zero distance

from the electrode and bulk of the solution, respectively. The overpotential η in this

case contains an activation and a mass transport term. When the surface

concentrations do not vary from those of the solution bulk Equation (2.3) falls into

Equation (2.2).

Since the faradaic reaction produces or consumes electroactive species, a gradient of

concentration is built up in front of the electrode. The region in which this gradient

exists is called Nernst diffusion layer and it extends from the electrode toward the

solution bulk for several micrometres or even hundreds of micrometres until the

natural convection destroys the gradient maintaining the concentrations as those of

the bulk. This concentration gradient at the electrode surface drives the mass

transfer and produces the concentration overpotential.

The total overpotential η is given by:

25

eqconcaohm uV (2.4)

Where ηohm is the ohmic drop given by the passage of current through the solution;

ηa is the activation overpotential, which drives the electron transfer; ηconc is the

concentration overpotential, which represents the activation energy required to drive

the mass transport of reactant and product forward and backward the electrode

surface at the rate needed to support the current density; V is the voltage applied

between the electrode and the electrolyte; and ueq is the equilibrium potential of the

reaction.

The usual assumption in electrochemistry is that the capacitive current and the

faradaic current are completely independent, which leads to the fact that these

currents can be taken as isolated and analyzed separately. This is referred as a priori

separation of the capacitive and faradaic current. The assumption is valid when a

supporting electrolyte is present and when the electroactive species are in low

amount compared with the supporting electrolyte. The total current i is then:

cF iii (2.5)

The concepts developed in this first section are important for the following parts. In

particular, the equations provided here are employed in the electrochemical

impedance spectroscopy and in the intermodulated differential immittance

spectroscopy.

26

2.2 LINEAR SYSTEMS AND ELECTROCHEMICAL IMPEDANCE SPECTROSCOPY

A linear system is a mathematical model based on the use of a linear operator. An

operator is linear when it satisfies the properties of superposition and scaling. Given

two inputs x1(t) and x2(t) and two outputs y1(t) and y2(t) such that:

txHty 11 (2.6)

txHty 22 (2.7)

The operator H is defined as linear if it satisfies the following relation:

tbxtaxHtbytay 2121 (2.8)

Where a and b are two scalar numbers. The system is then defined by the operator

H.

The two properties of linearity allow the system to be decomposed in the response of

simple signals. Almost any system can be linearized in a small domain to be suitable

for a linear treatment. This is for example the case of the electrochemical systems

when EIS is employed.

2.2.1 TRANSFER FUNCTIONS

The transfer function is the mathematical representation of the relation between

(co)sinusoidal input and output signals in a linear system and it describes the

amplitude and phase relation between the two. It is usually defined through the

Laplace or the Fourier transform of the two signals. As common in electrochemistry,

the Fourier transform and, therefore, the frequency domain is employed also in this

thesis. The advantage of the transfer function is that differential equations are

converted into algebraic equation.

If the system is perturbed by a potential sinusoidal signal as it is common in EIS, the

admittance should be defined as the complex ratio of the Fourier transform of the

potential, u(t), to the Fourier transform of the current, i(t), at the frequency ω:

U

I

tu

tiY F

F (2.9)

27

Where the italic F stands for the Fourier transform at the angular frequency ω. It is

often used, and preferred in this thesis, to represent the transformed variables with

capital letters followed by the angular frequency. For a function x(t), the Fourier

transform is defined as:

dtetxX tj (2.10)

These variables pass from real values to complex values in the frequency domain.

Therefore, the admittance Y is also in general a complex number. Using this

definition of transfer function, the linear operator H defined by Equations (2.6) and

(2.8) can be converted into its frequency response.

Although the admittance should be employed in the case the potential is used as

input, the impedance is often preferred. This is simply the reciprocal of the

admittance. They are collected together into the concept of immittance. The

possibility to pass from a transfer function to another is a peculiarity of this

mathematical framework.

The Impedance can be seen as the generalization of the Ohm‘s law to alternating

current regimes as function of the angular frequency.

The real part of the admittance is called conductance and the imaginary part is called

susceptance. Similar separation is valid for the impedance, where the resistance is

the real part and the reactance is the imaginary one.

In the EIS the systems can be seen as a combination of electric elements to form a

circuit called electric analog or equivalent circuit. This is particularly useful when

associated with the a priori separation of the capacitive and faradaic current. In fact

two circuits can be used to represent the capacitive current and the faradaic current

and they can be combined to construct the electric analog of the total electrochemical

system.

2.2.2 ELECTROCHEMICAL IMPEDANCE SPECTROSCOPY

In this subsection, the concept of transfer function is applied into the electrochemical

systems, leading to the basis of electrochemical impedance spectroscopy.

Let assume a potential (co)sinusoidal perturbation:

tj*tj

0eUeUutu (2.11)

28

Where the amplitude U is small enough to accomplish the assumption of small signal

approximation. The system is assumed to be linear. Therefore the current responds

with a (co)sinusoidal wave:

tj*tj

0eIeIiti (2.12)

Which possesses the same frequency but it is scaled in modulus and shifted in phase.

Following, some cases are demonstrated in order to show how to develop the

impedance.

2.2.2.1 DIODE

The first case I show is that of a diode. According to the Equation (2.5) the total

current is simply the sum of the capacitive and faradaic current, which leads to the

fact that the admittance is also simply the sum of the capacitive and faradaic

admittances as if there are two independent branches in the circuit.

A diode in reverse bias fits nicely this situation and the Nyquist plot of the impedance

is a perfect semicircle, where the faradaic current is substituted by the leakage

current il of the device, given by the Shockley equation:

Tk

euexp1ii

B

sl (2.13)

Where is is the saturation current in reverse bias given by thermo-emission of

electrons and movement of the holes in the valence band, u is the polarization

potential, e is the electron charge, and kB is the Boltzmann constant. For large and

positive potential il ≈ is.

The capacitive current is given by Equation (2.1), where time dependence of the

capacitance is neglected and Cdl is substituted by the following semiconductor

capacitance:

2

1

Bfb

D0r

SCe

Tkuu

2

NeεεuC

(2.14)

29

where εr is the relative dielectric constant, ε0 the permittivity of vacuum, ND the

concentration of dopants, and ufb the flat band voltage. Equation (2.14) is known as

the Mott-Schottky equation.

The current can be linearized around the point i0 of interest:

uuCuiu

ii 0SCl0

(2.15)

The partial derivative of the leakage current has the unit of Siemens and corresponds

to a leakage conductance Gl.

Passing to the admittance as shown by Equation (2.9), one yields:

SCl CjGY (2.16)

Which gives as impedance:

2SCl

SC

2

l

CR1

CRjRZ

(2.17)

Where Rl corresponds to the resistance at the leakage current.

The case of a real electrochemical system is shown in the next subsection.

2.2.2.2 ELECTROKINETICS OF A REDOX COUPLE IN SOLUTION

In this subsection the simplest electrochemical case is presented. Suppose that the

redox reaction of a fast redox couple in solution is given as:

RedneOx (2.18)

Where Ox and Red are the oxidized and reduced species, respectively. In order to

model the impedance several assumptions are made:

1. Pure diffusion mass transport. The redox couple is present in such a low

amount with respect to the supporting electrolyte that the transport number

of the redox species can be neglected.

2. A priori separation of faradaic and charging currents. The faradaic current and

the charging current are fully separable and independent.

30

3. The electrochemical reaction is first order with respect to reactants and

products.

4. The small signal approximation holds. In respect to the perturbation, the first

and second order terms are linear and quadratic, respectively.

5. The semi-infinite diffusion model is valid.

6. A negligible resistance of the electrolyte is present.

It should be noted that the first assumption is a necessary condition of the second.

As shown by Equation (2.3) the faradaic current is a function of the overpotential (u)

and the concentration of redox species (cred,0 and cox,0) at the electrode-electrolyte

interface. The capacitive current is defined in general as a function of only the

potential drop (u) and its variation in time ( ). As consequence of the a priori

separation of faradaic and capacitive current the total current i is:

u,uic,c,uii c0,Ox0,dReF (2.19)

In order to linearize Equation (2.19) it is convenient to employ immediately the

Fourier transform with the means of a matrix notation as shown by Rangarajan [68–

71]. For the faradaic current this leads to:

XJIT

FF (2.20)

Where JF is the faradaic Jacobian and X is the vector of physical variables x (u, cRed,0

and cOx,0), and the superscription T stands for transpose. JF and X are defined by:

Ox,0

F

oRed,

F

F

F

c

i

c

iu

i

J (2.21)

And

Ox,0

oRed,

C

C

U

X (2.22)

31

The usefulness of such a notation is that additional operators can be employed as in

the case of the mass transport. The variations of the concentrations at the interface

are dictated by the flux J, which is eventually related to the faradaic current.

Everything is translated by the operator mi as:

ωIωmωC Fii,0 (2.23)

Where the subscription i runs on Red and Ox. A plain derivation of m is reported in

Appendix A. With equation (2.23) it is possible to substitute Ci in Equation (2.22).

Equation (2.20) contains now only IF and U and after rearrangement leads to:

FOx

FdRe

T

FF

Ym

Ym

1

JY (2.24)

Where YF is the faradaic admittance.

After computing the matrix multiplication and rearranging:

Oxd

F

mm

Y

Ox,0

FRe

oRed,

F

F

c

i

c

i1

u

i

(2.25)

Which leads to the faradaic impedance ZF:

u

i

mc

im

c

i1

ZF

Ox

Ox,0

FdRe

oRed,

F

F

(2.26)

The charge transfer resistance is defined as the reciprocal of the derivative of the

faradaic current on the potential:

1

F

ctu

iR

(2.27)

32

And it represents the kinetic resistance at the electron transfer. At the OCP Rct is

related to the exchange current i0:

0

ctnFi

RTR (2.28)

And from i0 the standard heterogeneous rate constant k0 can be derived:

b,dRe

1

b,Ox00 ccaFki (2.29)

Where a is the surface area of the electrode and the subscription b refers to the bulk

of the solution.

Collecting all the terms dependent on the frequency in Equation (2.26), one has:

WctF ZRZ (2.30)

Where ZW is the Warburg impedance defined for a semi-infinite diffusion profile as:

j

1ZW (2.31)

Where Warburg coefficient ζ is given by:

OxbOx,RedbRed,

2Dc

1

Dc

1

nFa

RT (2.32)

Where a is the surface area of the electrode and Di are the diffusion coefficients. The

Warburg impedance represents the resistance given by the mass transport of

reactants and products driven by diffusion.

The capacitive current is given by Equation (2.1), which leads to the capacitive

admittance Yc:

dlc CjωY (2.33)

The total impedance is the reciprocal of the total admittance, which is the sum of the

faradaic and capacitive one. In order to visualize the effect of the terms of the

33

impedance, it is useful to simulate it. Figure 2–1 shows the simulated impedance of a

redox couple in solution with and without diffusion or kinetics limitation, using the

following normalized parameters: Cdl = 20 μF cm-2, Rct = 1.5 Ω cm2, and ζ = 15 Ω s–

0.5 cm2.

As extreme case, when the charge transfer kinetics is very facile, which means Rct

tends to zero, the faradaic impedance is represented only by the Warburg

impedance. In the Nyquist plot, this is characterized by a 45° straight line. In this

case only the effect of the mass transfer is visible (dotted line in Figure 2–1).

Another extreme case is considered when the kinetics is very sluggish or the mass

transport does not impose any limitation. The impedance resembles that of the diode

in reverse bias, which is a perfect semicircle (dashed line in Figure 2–1). When both

kinetics and diffusion are taken into account the impedance is a combination of the

two extreme cases. At high frequency the diffusion does not play a large role and the

impedance starts as a semicircle. As the frequency decreases, the Warburg

impedance grows and a 45° straight line rises. At very low frequency only the mass

transport is significant and the total impedance is simply given by the Warburg

impedance.

The EIS can recover the charge transfer resistance, the double layer capacitance, and

the Warburg impedance, but it cannot distinguish different values of and the

diffusion coefficients. Therefore, not all information on the electrochemical kinetics

and mass transport can be revealed by a single experiment. Instead, when the

Figure 2–1: Nyquist plot of the impedance of a redox couple in solution from 100

kHz to 10 Hz with different kinetic and diffusion limitation.

34

electrochemical system is treated as nonlinear, the amount of information available is

larger as shown in the next section.

35

2.3 DEVELOPMENT OF THE NONLINEAR SYSTEM

So far only the linear system theory was shown, which is connected with the concept

of transfer function and impedance spectroscopy, but a careful observation of

Equations (2.2) and (2.14) unveils that neither the faradaic current of a redox couple

nor the capacitive current in the case of a diode are linear functions of the potential.

In this case when the system is perturbed by a potential signal as that of Equation

(2.11) the current is not equivalent to that of Equation (2.12), namely it is not a pure

sinusoidal wave. Figure 2–2 shows the frequency domain of a linear and nonlinear

system perturbed by a sinusoidal potential as input (first row).

When the system is linear the frequency domain of the current (output) shows only

one peak as predicted by Equation (2.12) (second column of Figure 2–2). Instead,

when the system is nonlinear the current shows additional peaks, corresponding to

the higher order harmonics (third column of Figure 2–2) which are intrinsically

connected with the nonlinearity of the system. Their intensity and order depends on

the nonlinearity, as dictated by a Taylor series polynomial.

The case of the intermodulation, represented by the second row of Figure 2–2, is

reported in the next subsection.

Figure 2–2: schematic of the Fourier transform for a single (first column) and

double input measurement (second column) on a linear (second row) and

nonlinear system (third row). represents the fundamental harmonic, the

higher order harmonics and the intermodulation sidebands. (Adapted from [58])

36

2.3.1 INTERMODULATION

During the intermodulation the system is perturbed by two sinusoidal potential waves

of different frequencies, as schematized by the first column and second row of Figure

2–2. The lower frequency signal is called stimulus (frequency fω and angular

frequency ω) and the higher frequency one probe (frequency fΩ and angular

frequency Ω). Under such conditions, the output of a linear system is composed by

two sine-waves with the same frequencies fω and fΩ, but different modulus and phase

(second column of Figure 2–2), as for a single sinusoidal perturbation.

However, for nonlinear systems the situation is considerably different: in addition to

the higher order harmonics of the stimulus and probe, symmetric peaks around the

probe frequency are observed (third column of Figure 2–2). These are called

sidebands. Sidebands are generated by the second order polynomial in the Taylor

series of the nonlinear system and are located at fΩ – fω and fΩ + fω. The output of

the probe and the sidebands in the time domain (t-domain) represent the envelope

of the amplitude modulation of the stimulus on the probe signal. This is the so-called

phenomenon of the intermodulation, and it contains information coming from the

stimulus, the probe and the characteristics of the system. An important feature of the

sidebands is that the intensity of these signals is in general higher than the intensity

of second harmonics [38].

Employing the small signal approximation, it is possible to develop the system in a

similar fashion of the linear system theory. This approximation leads to the fact that

the first order signals in the third column of Figure 2–2 are linear with the input and

that the second order signals, second harmonics and sidebands, are proportional to a

combination of the square of the inputs. In this way the immittance transfer

functions can still be successfully employed together with a new set of transfer

functions defined in the next subsection.

2.3.2 DEFINITIONS AND DEVELOPMENT OF NEW TRANSFER FUNCTIONS

In this subsection, I introduce the new class of transfer function developed in this

work to handle the intermodulation.

The potential perturbation u(t) during an intermodulation experiment is composed by

two sine-waves of angular frequency ω and Ω superimposed to a DC potential u0:

tj*tjωtj*ωtj

0 eUeUeωUeωUutu (2.34)

37

The two waves are defined stimulus, that at angular frequency ω, and probe, that at

angular frequency Ω. For sake of simplicity, they are referred to as uω and uΩ. As in

the case of the impedance spectroscopy the amplitude of the perturbation has to be

small enough to guarantee a limited number of higher order terms in the current.

This is fundamental prerequisite to achieve the small signal approximation.

The current i(t) generated during the intermodulation by the potential of Equation

(2.34) is:

tωΩj*tωΩjtj*tj

tωΩj*tωΩjt-j*tj

0

eωΩIeωΩIeIeI

eωΩIeωΩIeωIeωIiti

(2.35)

Where i0 is the steady state current, the terms in ω and Ω are the linear answers of

the system, and the terms in Ω ± ω represent the intermodulation sidebands (for

clarity sake the second harmonics are discarded). As for the potential, also the

current can be divided into stimulus and probe signals. While iω contains only one

wave (at angular frequency ω), iΩ contains three signals, the linear answer at

frequency Ω and the intermodulation sidebands.

In order to demodulate the intermodulation, the same process of a phase-sensitive

detector is adopted (see Subsection 3.1.2.3 and 3.2.1 for details). This works

multiplying twice the current iΩ(t) by the probe perturbation uΩ, once in phase and

once out of phase. The result is normalized by the absolute value of the probe. By

neglecting the terms which contains ± 2Ω, the conductance GΩ and the susceptance

BΩ at the angular frequency Ω are derived as:

,G~

G2

1G

2

1

U4U

eUeUti

*

tj*tj

(2.36)

And:

,B~

B2

1B

2

1

U4jU

eUeUti

*

tj*tj

(2.37)

These coincide with Equations (3.15) and (3.16) in the third chapter. The tilde ‗~‘

represents a variable oscillating. Contrary to the case of a standard impedance

spectroscopy, GΩ and BΩ are time-variant quantities which are composed by a steady

state value G(Ω) and B(Ω), correlated with the admittance Y(Ω), plus an oscillatory

term (ω,Ω) and (ω,Ω), which oscillates at the frequency ω. During the

38

intermodulation, the stimulus modulates the conductance and the susceptance of the

system and hence its admittance.

The two waves (Ω) and (Ω) are:

tj

*

*tj

tjtj

*

*

eU

Ie

U

I

eU

Ie

U

I

G~

(2.38)

tj

*

*tj

tjtj

*

*

eU

Ie

U

I

eU

Ie

U

I

jB~

(2.39)

Two new transfer functions can be defined from the time-variant conductance and

susceptance. These are the differential conductance, dG, and the differential

susceptance, dB, which are defined as for the case of the admittance from the

complex ratio of the Fourier transform of the variables to the transform of the

stimulus:

ωU

ωG~

Ωω,dG

F

(2.40)

ωU

ωB~

Ωω,dB

F

(2.41)

Where capital letters are used to indicate the Fourier transforms of the corresponding

quantities. dG and dB represent the variation of the conductance and of the

susceptance given by the stimulus and they are in general complex values. They can

be combined into a differential admittance dY as:

Ωω,jdBΩω,dGΩω,dY (2.42)

Which represents the total variation of the admittance measured at Ω produced by

the stimulus.

From Equations (2.40) and (2.41) dG and dB become:

39

ω2U

ΩU

ωΩI

ΩU

ωΩI

dG*

*

(2.43)

ω2U

ΩU

ωΩI

ΩU

ωΩI

jdB*

*

(2.44)

Plugging them into Equation (2.42) dY becomes:

ωUΩU

ωΩIΩω,dY

(2.45)

Which is given by the right (upper) sideband alone. Similarly, the anti-differential

admittance d can be defined for the left (bottom) sideband as:

ωUΩU

ωΩIΩω,Yd

*

* (2.46)

By Equations (2.45) and (2.46) dG and dB can be rewritten as:

2

ωΩ,dYωΩ,YdωΩ,dG

(2.47)

2

ωΩ,dYωΩ,YdjωΩ,dB

(2.48)

In the Appendix B, from dY also the differential impedance dZ is derived.

In the next subsection, there are some examples where these transfer functions are

employed.

2.3.3 SIMPLIFIED CASE

In order to understand which information lay in the intermodulation, in this

subsection a simple case is introduced. Let assume a current i = f(u), which can be

expressed in a Taylor expansion up to the second order as:

40

2

2

2

0 Δudu

fd

2

1Δu

du

dfufi (2.49)

The derivatives are real values and both current, i, and potential, u, are given by

Equation (2.34) and (2.35). Expanding the term Δu2 and passing to the frequency

domain the following identities are found:

ΩUωUdu

fdωΩI *

2

2* (2.50)

ΩUωUdu

fdωΩI

2

2

(2.51)

And from the definition of dY and d given by Equations (2.45) and (2.46):

2

2

*

*

du

fd

ΩUωU

ωΩIωΩ,Yd

(2.52)

2

2

du

fd

ΩUωU

ωΩIωΩ,dY

(2.53)

Equations (2.52) and (2.53) show the connection between the intermodulation and

the second derivatives of the investigated system in a similar fashion as for the EIS

and the first derivatives.

From Equations (2.47) and (2.48), dG and dB are found. In this simple case, the left

and right sidebands are equal and only the real part of dG is non-zero. Furthermore,

if one calculated the causality factor as suggested by Bosch et al. [39], this would be

equal to two. In general, this is not true as it is proved in the fourth chapter

(Subsection 4.3.2).

In the next subsections, some real cases are examined when the intermodulation is

applied.

2.3.4 IDEAL NONLINEAR CASE: DIODE

In the Subsection 2.2.2.1, the example of the diode is reported in reverse bias as

ideal case during an impedance experiment. The stimulus, affecting the potential

41

across the system, modulates the value of the capacitance of the diode, which is the

origin of the amplitude modulation of the probe current at the stimulus frequency.

The leakage current can be considered linear with the potential and only the

capacitive current contributes to the secondary response of the system. Its second

derivative calculated using Equation (2.14) leads to:

23

Bfb0

D0r

e

Tkuu

2

Neεε

2

1jdG

(2.52)

23

Bfb0

D0r

e

Tkuu

2

Neεε

2

1ΩdB

(2.53)

It is worth to note that dG should be negative and imaginary, while dB, although

negative, should be real. Both terms are related to the variation of the capacitance

with the potential, however, the value of dG is proportional to ω, while the value of

dB is proportional to Ω.

An IDIS can be used in combination with an EIS to recover the flat band potential of

the device as shown in the fourth chapter (Section 4.2).

2.3.5 ELECTROCHEMICAL CASE: REDOX COUPLE

The case of a redox couple is shown here. This follows Subsection 2.2.2.2.

If the system is nonlinear, not only the current, but all physical variables show the

intermodulation sidebands in the frequency domain. These appear at angular

frequencies Ω + ω and Ω – ω (see Figure 2–2). In addition to the faradaic Jacobian,

JF, and the vector of physical variables, X, defined by Equations (2.21) and (2.22),

the faradaic Hessian, HF, is given by:

00cu

i

00cu

i

cu

i

cu

i

u

i

H

Ox,0

F

2

Red,0

F

2

Ox,0

F

2

Red,0

F

2

2

F

2

F (2.56)

42

Which represents the matrix collecting all the second derivatives. As for the previous

cases (Subsection 2.2.2.2), the currents as well as the sidebands are calculated for

the faradaic (IF) and capacitive (Ic) current.

ωXHΩXωΩXJωΩI F

TT

FF (2.57)

ωXHΩXωΩXJωΩI F

*T*T

F

*

F (2.58)

And

ωUΩΩωUu

CωUΩUωΩ

u

CjωΩI dldl

c

(2.59)

ωUΩΩωUu

CωUΩUωΩ

u

CjωΩI dldl*

c

(2.60)

It is noteworthy that Equations (2.57) and (2.58) contain both first and second

derivatives of the current contrary to Equations (2.59) and (2.60), because the

potential is the only quantity that does not oscillate at Ω ω. It is clear from this set

of equations the usefulness of a matrix formalism to develop the secondary effects

during an intermodulation.

As for X(ω) in Equation (2.24), also the vectors X(Ω+ω) and X*(Ω–ω) depend on the

values of IF(Ω+ω) and IF*(Ω–ω) through the mass transport operator m.

The faradaic sidebands are:

ωΩmc

iωΩm

c

i1

ωXHΩXωΩI

Ox

Ox,0

FRed

Red,0

F

F

T

F

(2.61)

ωΩmc

iωΩm

c

i1

ωXHΩXωΩI

*

Ox

Ox,0

F*

Red

Red,0

F

F

*T*

F

(2.62)

Which lead to the faradaic differential and anti-differential admittance dYF and d F:

43

ωΩmc

iωΩm

c

i1

ωAHΩA

ωUΩU

ωΩIωΩ,dY

Ox

Ox,0

FRed

Red,0

F

F

T

FF

(2.63)

ωΩmc

iωΩm

c

i1

ωAHΩA

ωUΩU

ωΩIωΩ,Yd

*

Ox

Ox,0

F*

Red

Red,0

F

F

*T

*

*

FF

(2.64)

Where the vector A is defined for brevity as:

ωYωm

ωYωm

1

ωU

ωCωU

ωC

1

ωA

FOx

FRed

oOx,

oRed, (2.65)

Capacitive differential and anti-differential admittance dYc and d c are given by:

Ωωu

CωΩ

u

Cj

ωUΩU

ωΩIωΩ,dY dldlc

c

(2.66)

Ωωu

CωΩ

u

Cj

ωUΩU

ωΩIωΩ,Yd dldl

*

*

c

c

(2.67)

The total differential and anti-differential admittance are then the sum of the faradaic

and capacitive terms, which are used through Equations (2.47) and (2.48) to recover

dG and dB.

Given the complexity of Equations (2.64) - (2.67), a straightforward interpretation of

dG and dB is impossible and the simulated results are reported in the following

subsections.

2.3.6 MATHEMATICAL SIMULATION OF THE REDOX COUPLE

The mathematical simulation of the intermodulation on a redox couple is shown in

this and next subsections. In Table 2–1 the parameters describing the system and

44

their relation to the kinetic equation, mass transport and interfacial capacitance are

reported.

Three cases are possible: the system is either under kinetic or mass transport

limitation or both limitations contribute. The following subsections show the different

cases.

The mathematical simulations are performed fixing the parameters of the impedance

which lead to the Nyquist plot in Figure 2–1. This means to freeze all the first

derivatives of the current. As reported in the Subsection 2.2.2.2, EIS cannot give all

the information on the electrochemical kinetics and mass transport in a single

experiment. It is interesting to understand if the differential conductance and

differential susceptance are able to separate the effect of and δ from the other

effects.

The simulations are performed with a probe frequency of 0.5, 5, or 100 kHz and the

stimulus frequency spans down to 0.1 Hz. All the results are reported as real and

imaginary parts of the differential conductance and susceptance as a function of the

stimulus frequency.

Table 2–1: Parameters used during the simulation of the impedance, differential

conductance, susceptance, and admittance of the redox couple in solution.

Parameter Definition Physical meaning

Rct Kinetic limitation

Symmetry of barrier

ζ Mass transport limitation

δ Symmetry of mass transport

Cdl Capacitance of the interface

Δdl Variation of Cdl with potential

ηdl Time constant of Cdl variation

45

Next subsection starts from the analysis of the charging current, which is an

extension of what already observed for the diode (Subsection 2.3.4).

2.3.6.1 CAPACITIVE CURRENT

The capacitive current has two nonlinear terms, which cannot be excluded a priori

(even in the simplified case of excess of supporting electrolyte and no adsorption

phenomena): the first is relative to the variation of the capacitance of the interface

with the potential, Δdl (e.g. for the diode), as given by the Stern model; the second is

the time constant of the charging of the interface, ηdl.

If only the capacitive terms are considered, dG and dB can be written as:

dl jωωΩ,dG (2.68)

dldl ηjω1ΩΔωΩ,dB (2.69)

dB is related to the interface capacitance transfer function proposed by Keddam and

Takenouti [54–57].

The simulation is performed keeping the probe frequency fixed at 5 kHz and

spanning the stimulus down to 0.1 Hz.

Figure 2–3–a shows the effect of Δdl on the differential immittance spectra of a redox

couple. Δdl affects both the imaginary part of the differential conductance, dG, and

the real part of the differential susceptance, dB. As seen before (Subsection 2.3.4),

Δdl affects dG and dB proportionally to ω and Ω, respectively. The immittance Bode

plot of dG shows the frequency dependence of a capacitor and the plot of dB that of a

resistor.

Figure 2–3–b reports the effect of ηdl on the differential immittance spectra. It

combines the effect of Δdl with a linear variation of the imaginary part of dB. This is

one of major contributions on Im(dB).

ηdl and Δdl are closely related. In fact, in presence of ηdl, Δdl cannot be zero. Indeed,

being ηdl the time necessary for the capacitance of the interface to change its value,

it has no physical meaning if the value of the capacitance of the interface does not

change.

The time constant of the double layer was predicted [67] and experimentally

observed by others [54–57].

In the differential immittance Bode plot, dG still shows the frequency dependence of

a capacitor, while dB resembles the admittance spectra of a resistor in parallel to a

46

capacitor. By changing the probe frequency Ω, the absolute value of dG does not

change, whereas the absolute value of dB changes proportionally, as shown by

Equations (2.68) and (2.69).

The next subsection focuses on the effect of the parameters that rule the faradaic

current.

2.3.6.2 FARADAIC CURRENT: ABSENCE OF MASS TRANSPORT

In this subsection, only with faradaic current is considered, ignoring the nonlinearities

of the capacitive one. In the beginning, the mass transport is neglected. In this case

the concentrations of the redox species at the interface are not a function of the

current density (the mass transport functional mi is zero) and, therefore, they are not

Figure 2–3: effect of the nonlinearity of the double layer capacitance on the

differentials. a) effect of dl = 20 F V-1 cm-2; b) effect of dl = 5s. Inlets: same

graphs in linear scale.

47

affected by the variation of potential. The impedance is represented by the dashed

line in Figure 2–1.

From equations (2.63) and (2.64), one can obtain the differential conductance dG

and the differential susceptance dB as:

ctR

α21

RT

nFωΩ,dG

(2.70)

0ωΩ,dB (2.71)

Only the differential conductance is non-zero. In addition, it is a real number

dependent on Rct and which represents the symmetry of the energetic barrier

between reactants and products. In particular, with higher than 0.5 the increase in

the reduction current is faster than in the oxidation current, whereas for lower than

0.5 the opposite is true. Figure 2–4 shows the effect of on dG when the other

parameters are kept constant.

When is 0.5 the second derivative is zero and Re(dG) is zero as well. Instead, when

is higher than 0.5, the real part of the differential conductance is negative. The

more departures from 0.5, the more dG becomes large. When is lower than 0.5,

the real value of dG behaves in a specular way with respect to the previous case.

Under this condition dG resembles the frequency dependence of a resistor.

Figure 2–4: effect of on the real part of the differential conductance.

48

2.3.6.3 FARADAIC CURRENT: ABSENCE OF KINETICS LIMITATION

Contrary to the previous subsection, here the charge transfer limitation is neglected.

The effect on the impedance is shown in Figure 2–1 by the dotted line. In this case,

the charge transfer resistance tends to zero and the differential admittance and the

differential conjugated admittance become:

ωΩjζ

δ21

RT

nFωΩ,dY

F

(2.72)

ωΩjζ

δ21

RT

nFωΩ,Yd

F

(2.73)

And the differential conductance and susceptance become:

ωΩωΩjωΩωΩ2

2

δ21

RT

nFωΩ,dG

(2.74)

ωΩωΩjωΩωΩ2

2

2δ1

RT

nFωΩ,dB

(2.75)

The differentials are only function of δ, ζ, and of the stimulus and probe frequencies.

One can notice that the real parts of dG and dB coincide, whereas the imaginary ones

are conjugated.

If there is symmetry in the mass transport (δ = 0.5) all the differentials are equally

zero. On the contrary, when the symmetry is broken, all the differentials show similar

frequency dependence. A coefficient higher than 0.5 produces a negative value for

Re(dG), Re(dB), and Im(dB), whereas the value of Im(dB) is positive. Figure 2–5

shows the effect of the mass transport on the differentials when δ is 0.55. The real

part of dG and dB coincide and result almost flat in the whole frequency range,

tending towards zero only when ω approaches Ω. As one can see from the inlet, the

imaginary part of dG and dB are perfectly linear in the whole range of investigated

frequencies and tend to zero as the frequency of the stimulus approaches zero. The

value of δ could be easily extrapolated from the slope of the imaginary parts at low

frequencies or from the intercept of the real parts at ω approaching zero frequency.

For ω << Ω, dG and dB show the characteristic behavior of a capacitor in series with

a resistor.

49

2.3.6.4 FARADAIC CURRENT: MASS TRANSPORT AND KINETICS LIMITATION

Here I show the combined effect of both mass transport and electrokinetic limitations

on the differential immittance spectra. As for the previous cases, Figure 2–1 shows

the Nyquist plot of the impedance when the system is limited by both mass transport

and electrokinetics.

As known from the previous part, the total mass transport and charge transfer

resistances are given by ζ and Rct, respectively, while δ and represent their

symmetry. By changing and δ, and keeping constant ζ and Rct, the total mass

transport and electrokinetic limitations of the systems remain constant, but the

symmetry between the partial reduction reaction and the partial oxidation reaction is

changed.

Figure 2–6–a shows the effect of the asymmetry of the charge transfer ( = 0.55 and

δ = 0.5). The real part of dG has a sigmoidal shape with larger values at higher

frequencies of the stimulus, while the imaginary part of dG is negative and it is of

bell-shape. The maximum of the bell is in correspondence of the frequency where

Re(dG) shows its flex. The value of dB is very small compared to the value of dG.

This is due to the frequency of the probe signal. Later, it is shown that at lower

frequencies, when diffusion becomes predominant over charge transfer, dB becomes

a significant fraction of the differential admittance.

Figure 2–5: effect of the diffusion on the differentials in absence of kinetics; the

real parts of the conductance and of the susceptance coincide. Inlet: same graph

in linear scale.

50

Figure 2–6–b reports the complementary case: only the mass transport is

asymmetric ( = 0.5 and δ = 0.6). Both real and imaginary part of dG and dB show

similar features as before. However, the sigmoidal shape of Re(dG) is inverted and

tends to zero for high frequencies. Instead, Im(dG) is positive. The presence of a

hook at the higher frequencies is a feature common to the real and imaginary parts

of all differentials and it is a special feature given by the asymmetry of mass

transport.

If both mass transport and charge transfer asymmetries are present, the final effect

is just the sum of the two. This is summarized in Figure 2–6–c and –d, where =

0.6, and = 0.55 and 0.45, respectively. All features observed in the two previous

cases appear in Figure 2–6–c. One can imagine the complete case as composed by

two parallel circuital branches each one having only one asymmetry. It is interesting

to observe Figure 2–6–d. In this case, the asymmetries of mass transport and of

electrokinetics are in opposite direction (δ > 0.5 and < 0.5) and the real part of dG

crosses the x-axis.

It is important to observe what happens when and are equal. This is shown in

Figure 2–7. All differentials are flat for a large range of frequencies and the imaginary

Figure 2–6: effect of mass transport and kinetics limitation on the differentials; a)

asymmetry of kinetics; b) asymmetry of diffusion; c) both kinetics and diffusion asymmetric; d) opposite symmetry for diffusion and kinetics.

51

parts of dG and dB are practically zero. The frequency dependence is lost apart for

the hook at the very end. It is noteworthy to mention that δ can be changed by

controlling the concentrations in solution. Therefore it is possible, once found the

value of DOx and DRed, to redesign the experiment to obtain δ = .

So far the probe frequency was fixed at 5 kHz. In order to understand its effect, this

is changed to 500 Hz and 100 kHz. At lower frequency, the mass transport has the

largest impact on the impedance. However, at higher frequency, the diffusion is

negligible. This is well-visible in Figure 2–1.

Figure 2–8–a shows the case with Ω = 500 Hz, = 0.55 and δ = 0.6. As anticipated,

the differences between Figure 2–8–a and Figure 2–6–c lays in the Re(dG) and in

Re(dB). The sigmoidal curve of Re(dG) is shifted positively and shrinks, as well as the

bell-shape of Im(dG). The real part of the differential susceptance is more negative

and flat in a large range of frequencies, whereas its imaginary part remains near

zero. The hook, which was rather small in the case of Ω = 5 kHz, now becomes more

pronounced in all differentials.

Figure 2–7: effect of = = 0.55 on the differentials. Im(dG) and Im(dB) overlaps.

52

At 100 kHz the mass transport limitation becomes negligible and the electrokinetics

dominates. In this case, the hook at high frequencies disappears, confirming that its

presence and intensity is related to the mass transport limitation (see Figure 2–8–b).

dB becomes very small and eventually zero for very high values of Ω. In this case, an

approximated solution exists which leads to:

Figure 2–8: effect on the differentials of the probe frequency = 0.55 and δ =

0.6. a) probe frequency of 500 Hz; b) probe frequency of 100 kHz.

53

ct

ct

ct

F R

α21

ζR

ζδ21Rα21

2RT

nFωΩ,dY

(2.76)

ct

ct

ct

F R

α21

ζR

ζδ21Rα21

2RT

nFωΩ,Yd

(2.77)

From these the differential conductance and differential susceptance are derived as:

ζR

ζαδ1Rα21

R

1

RT

nFωΩ,dG

ct

ct

ct

(2.78)

0ωΩ,dB (2.79)

As observed dB becomes zero and an easy solution for and δ exists.

What was simulated in this subsection is reported for a real experiment in the fourth

chapter (Section 4.3). In the next subsection, the effect of the resistance of the

electrolyte on the intermodulation, which was disregarded so far, is shown.

2.3.7 RESISTANCE COMPENSATION

So far the resistance of the electrolyte was considered negligible, but this is seldom

the case. In common electrochemical experiments the resistance of the electrolyte

affects the nonlinearities. The sidebands in the frequency domain of the current are

reduced in intensity and phase-shifted.

The quantities measured are, as in a normal electrochemical impedance

spectroscopy, the current i(t) and the total potential v(t). However, interfacial

admittance Yin, interfacial differential admittance dYin and interfacial anti-differential

admittance d in are defined with respect to u(t), the interfacial potential. A

correlation exists between interfacial potential and total potential v:

54

iRuv el (2.80)

Where Rel is the resistance of the electrolyte. Taking into account that an

intermodulation experiment is performed and passing from the time domain to the

frequency domain, one obtains:

ωIRωVωU el (2.81)

ωΩIRωΩU el (2.82)

ωΩIRωΩU *

el

* (2.83)

Because of the resistance of the electrolyte, the interfacial potential shows

intermodulation. U( ) represents the virtual source of the intermodulation.

Equation (2.81) can be used to calculate U(Ω) and U(ω) from v(t), i(t), and Rel.

Equations (2.82) and (2.83) can be combined with Equations (2.79) to extract the

experimental values of dY0 and d 0:

inelin0

YR1dYdY (2.84)

*

inelin0 YR1YdYd (2.85)

These represent the values of the differential admittance and anti-differential

admittance as if the experiment were conducted in presence of a negligible resistance

of the electrolyte. The subscription ―in‖ means that the quantity is measured using

U() and U(). The experimental values can be reconstructed knowing Rel and the

admittance at the frequencies – and + . Equations (2.84) and (2.85)

represent the fact that the passage of current through the resistance of the

electrolyte produces an ohmic drop which reduces the intermodulation. The algorithm

to recover the differential is reported in the third chapter (Subsection 3.4.6).

In this section, the intermodulation was employed to recover an IDIS spectrum. dG

and dB were used to obtain some information from the investigated system. In the

next section, the intermodulation is used to parameterize a nonlinear time varying

system.

55

2.4 DEVELOPMENT OF THE NONLINEAR TIME VARYING SYSTEM

In the previous section, the intermodulation was employed to recover the

differentials (dG and dB). These were correlated with some properties of the system

which could not be recovered with a simple impedance. In this section, the concept of

intermodulation is introduced in the study of a time variant system.

As seen in the Subsection 2.3.2 during the intermodulation, the conductance and the

susceptance of the system oscillate and a new set of transfer functions, dG, dB, and

dY, can be used to correlate the oscillation to the stimulus perturbation. The probe

sees the system as a periodic time varying one and this is the link between the

intermodulation and the study of nonlinear time variant systems.

In a periodic steady state all the properties of the system oscillate and the

impedance and admittance lose their normal meanings. It does not exist a unique

impedance for the system, but there is a different one at every time.

One could imagine the dynamics of the system dependent on a measurable

parameter called scheduling parameter. In the case of a periodic variation, the

scheduling parameter oscillates periodically. In this frame, the system is treated as a

periodic linear parametric varying system.

Let assume a system which possesses a natural oscillation ωn. As in the case of the

intermodulation differential immittance, the admittance should be measured with a

phase-sensitive detector leading to an oscillating conductance and susceptance at

frequency ωn which can be combined into an oscillating admittance. During

intermodulation, these oscillating quantities were transformed in respect of their

source, i.e. the stimulus potential for the previous examples. In the case of a time

varying system the stimulus is given by the scheduling parameter.

The task is to model the system upon the scheduling parameter. Some assumptions

are required:

1. The system is nonlinear. This is a natural requirement to have a time variant

system.

2. The system oscillates in a pure periodic fashion. This means that either the

oscillation is (co)sinusoidal or the harmonics are neglected in the treatment.

All the quantities oscillate accordingly.

3. The scheduling parameter can be recovered during the experiment.

4. The signals used to recover the impedance and the scheduling parameter do

not intermodulate. This means that the scheduling parameter is a rather linear

component of the electrochemical system.

The second assumption integrates the fourth assumption about small signal

approximation in the Subsection 2.2.2.2.

56

The model is built upon a system composed by a gas-evolving electrode working at a

potential high enough to evolve oxygen bubbles. These bubbles give rise to the

electrochemical noise in the current. As scheduling parameter, the surface area a(t)

of the electrode is taken. This is periodically occupied by a bubble.

In order to recover the surface area, a high frequency signal perturbation is used.

Simultaneously, the impedance is measured at the frequency . The current i(t) is

given by:

k

tkω-j

n

*tkω-j

n

tj*tj

k

tkω-ωj

n

*tkω-ωj

n

tj-*tj

k

t-jk

n

*tjk

n0

nn

nn

nn

ekω-Iekω-IeIeI

ekω-ωIekω-ωIeωIeωI

ekωIekωIiti

(2.86)

The terms in kωn belong to the electrochemical noise and they are collected into iωn.

The terms containing only ω and Ω are the linear responses of the system. The

electrochemical noise intermodulates with both ω and Ω. Figure 2–9 shows the

schematic of the frequency domain for Equation (2.86), according to the assumption

k = 1.

Following the names used for Equation (2.35), the current i(t) is divided into iωn, iω

which contains the intermodulation of the noise at the frequency ω, and iΩ which

contains the intermodulation of the noise at the frequency Ω. From iΩ the scheduling

Figure 2–9: Fourier transform of a nonlinear time variant system given by Equation (2.84) with k = 1. represents the electrochemical noise, the fundamental harmonic, and the intermodulation sidebands.

57

parameter, a(t) is recovered from the real part of the admittance as done in Equation

(2.36). Subsequently, the variation of the surface area a(t) in its Fourier

representation ( (ωn,Ω)) is defined as:

*

00

n

*

n

*

0

n

nYY

U

I

U

I

YRe

,G~

,A~

(2.87)

Where Re[Y0(Ω)] is taken from an experiment performed without bubble evolution.

This is required to normalize the admittance to the surface area. The scheduling

parameter represents the stimulus which perturbs the impedance. Therefore, from

Equation (2.40) and (2.41), one can write:

,A~

,dZZ~

nnn (2.88)

As shown in Appendix B, dZ can be derived from dY as:

2

n

nY

,dY,dZ

(2.89)

Where Y(ω) is the admittance at the probe frequency. In this case, dY is defined as:

,A~

U

I,dY

n

n

n (2.90)

Where it is clear that the perturbation ω represents the probe and ωn the stimulus.

Combining Equation (2.88) - (2.90) one yields:

I

I

,A~

Z

,A~Z~

n

nn

n (2.91)

Which is a new transfer function named ZA. This transfer function represents the

normalization of the oscillating impedance on the source of the oscillation which is

the changing of the surface area of the electrode.

In the fourth chapter (Section 4.4), I describe how this model is employed, focusing

on the oscillation of the admittance and of the scheduling parameter. From the

58

comparison of the mean impedance Zm and ZA the effect of the bubble on the oxygen

evolution reaction is explained.

59

3 EXPERIMENTAL PART

In this chapter, I give an overview on the instruments, materials and procedures

employed in this thesis. In the first section I describe the most common working

process of a potentiostat and of the electrochemical impedance spectroscopy. In

particular, I focus on the limitations of the instruments. In the second section, I

report the two instrumental configurations for the intermodulated differential

immittance spectroscopy (IDIS) developed in this work and I explain their

advantages and disadvantages. In section 3.3 I discuss the electrochemical cell

geometry according to the artifacts and the distortions observed in an

electrochemical impedance spectrum. In order to obtain reliable IDIS experiments,

and therefore reliable impedance experiments, particular care was posed in the

optimization of the electrochemical cell employed in this thesis. The optimizations

concerned both the artifacts coming from the stray capacitance and the improper

disposition of the electrodes in the cell.

Finally, in the last section the materials, the instruments, and the procedures

employed in this thesis are listed.

60

3.1 POTENTIOSTAT AND IMPEDANCE MEASUREMENT

The first section deals with the basic working principle and structure of the

potentiostat. I focus on the main problems connected with this instrument when

operating at high frequencies. In particular, I describe the transimpedance of the

current follower. In the last section of this chapter (Subsection 3.4.4), I show a

strategy to recover this transfer function and the results are shown in the fourth

chapter (Subsection 4.1.1).

In the second part, the most common instruments and techniques that are employed

to recover impedance spectroscopy, such as the fast Fourier transform (FFT)

analysis, the frequency response analyzer (FRA), and the phase-sensitive detector

(PSD) are described. The FFT analysis and the PSD are fundamental in order to

understand how to perform an IDIS. Often in the second chapter, I referred to the

phase-sensitive detector when I spoke of the demodulation of the signals. In this

section, this concept is described.

3.1.1 POTENTIOSTAT

The potentiostat was introduced in the 1950s. Its main scope is to achieve a fine

potentiostatic (or galvanostatic) control of the electrochemical experiment by the

means of three electrodes: a working electrode WE, a reference electrode RE, and a

counter electrode CE. It is mainly composed by a high impedance input to measure

the potential, a means to measure the current and a control loop to rule the potential

(potentiostatic mode) or the current (galvanostatic mode) in the cell. These parts are

schematized in Figure 3–1 for a potentiostatic control. The potentiostat is equipped

with two outputs. One of them is proportional to the measured potential, E-monitor,

and the other one is proportional to the current, I-monitor. Both are used to feed

other instruments or just to record the signals. The input is connected to a function

generator that provides the desired potential or current profile which is then applied

to the cell. In most cases, the potentiostat itself is equipped with a function

generator.

The potential is measured between the WE and the RE, while the current flows

between the WE and the CE. The WE is usually grounded and the potential of the

electrochemical solution is changed. The current needed to maintain the potential

difference flows through the CE which acts as a simple current sink in order to avoid

any ohmic drop in the RE branch.

61

A potentiostat can present problems when not operating in DC mode. The operational

amplifiers which compose the instrument depart from the ideality when the

instrument works in AC condition and increasing the frequency of operation the

results get worse and worse [117]. There are three kinds of problem.

One is connected with the control loop. At high frequency the response of the

potentiostat is slower and weaker. This means that the signal applied to the cell is

phase-shifted and smaller in amplitude than the one received from the input. This

problem can be circumnavigated by using larger inputs and monitoring the real signal

applied to the cell, how it is done usually.

The second problem, somehow more involved, is connected with the current reading

which is implemented as a current to voltage converter. The conversion unit is volt

on ampere, that is ohm as for the impedance. This conversion is also called

transimpedance which is a shorthand for transfer impedance.

The transimpedance Ht of a current to voltage converter is frequency dependent. The

converter gain is not constant and decreases with the frequency. It is overlapped

with a low-pass filter: the greater the gain, i.e. the lower the current range, the

lower the cutoff frequency of the filter. The cutoff is sometimes called the bandwidth

of the converter (although the same term is also used for the control loop). Typical

potentiostats show a bandwidth of 1 MHz at 1 mA full scale, but they can go as low

as 10-100 Hz at 1 nA full scale. In normal cases, the transimpedance can also be

dependent on the impedance of the investigated system.

In order to stabilize the converter or to reduce the noise, additional filters are placed

on the current reader. These influence the transimpedance decreasing the bandwidth

of the device. It is important to know the transimpedance of the instrument,

Figure 3–1: schematic of a potentiostat electric circuit for potentiodynamic experiments from reference [117].

62

especially when operating near or above its bandwidth. Section 3.4.4 explains the

strategy to study the potentiostat transimpedance and in fourth chapter (Subsection

4.1.1) some results are shown.

The third problem of the potentiostat is connected with the non-infinite impedance of

the RE input, which at high frequencies becomes small enough to allow some leakage

current to flow into the reference circuit. In Subsection 3.3.1, this issue is discussed

regarding the artifacts which can be originated during an electrochemical impedance

spectroscopy.

In the next subsection, the most common strategies to recover the impedance from

an EIS are shown.

3.1.2 IMPEDANCE MEASUREMENTS

The most common setups and instruments to perform the electrochemical impedance

spectroscopy are shown in this subsection. In potentiostatic mode, the impedance is

measured superimposing a sinusoidal perturbation to a DC potential.

All methods described in this subsection require the connection with the E- and I-

monitor outputs of the potentiostat and, if the potentiostat is not ideal, the correction

of its transimpedance. The methods can be divided into time domain, fast Fourier

transform (FFT) analysis, and into frequency domain techniques, frequency response

analyzer (FRA) and phase-sensitive detector (PSD). Their difference lies in the way

and in the domain in which they perform the integral transformation which then leads

to the impedance.

The impedance, Z, is defined as the complex ratio of the Fourier transform of the

potential, u(t), to the Fourier transform of the current, i(t), at the frequency ω:

I

U

ti

tuZ F

F (3.1)

Where capital letters are shorthand for the Fourier transform. It is clear from

Equation (3.1) that Z(ω) is a function of the frequency. Besides, it is in general a

complex number. Therefore, it is expressed in cartesian coordinates as:

ZImjZReZ (3.2)

Where Re and Im stand for the real and the imaginary argument and j is the

imaginary unit. In polar coordinates the impedance is:

63

jeZZ (3.3)

Where |Z| is the absolute value of the impedance and θ is the phase-shift.

For the sake of simplicity, the different methods to recover the impedance are

discussed only for a single input i(t), the I-monitor. Let assume that this input is

composed by a sinusoidal signal with angular frequency ω, amplitude I and phase-

shift θ. Let also disregard any conversion factor, transimpedance correction, and DC

component. Hence, i(t) is defined as:

tcosIti (3.4)

The impedance is simply calculated from the ratio of the integral transform between

the potential and the current.

3.1.2.1 FAST FOURIER TRANSFORM ANALYSIS

The fast Fourier transform (FFT) analysis first requires the digitalization of the signal

and then its elaboration; the latter can be carried out by a computational unit in the

same device or by a personal computer. The main advantage of this method is its

flexibility, because a large range of mathematical calculation can be performed on a

digital signal.

The analog to digital conversion (ADC) is the key point of this method and plays a big

role in the resolution of the technique. Nowadays, the ADCs are capable of extremely

high performances with a sampling rate as high as several megahertz and almost

endless recording time. The only limitation, under this point of view, appears to be

the capability and the memory of the calculation unit.

The ADC transforms the continuous signal into a time- and band-limited signal. This

places some restrictions. First, the digital signal is limited in time (this is indeed a

limitation always presents), and second, it is chopped into a sequence of data

separated by a time step given by the sampling rate of the acquisition. Ideally every

data point should be recorded for an infinitesimal small amount of time; in real

devices this is not the case and every single data is actually an average over a short

time of the recorded signal. This is usually referred as the analog bandwidth of the

ADC.

After the digitalization, the signal is only suitable for the discrete Fourier transform,

implemented as FFT, which is an algorithm to optimize calculation. This algorithm

64

was a breakthrough in the field of signal processing allowing the expansion of the

nuclear magnetic resonance and infrared spectroscopy.

The FFT analysis, as opposite to the other methods described later on (Subsections

3.1.2.2 and 3.1.2.3), shows the total frequency domain of the signal, noise,

harmonics and spurious signals included. A glance at the whole spectrum is a good

way to control the quality of one‘s instrumental setup and to understand the noise

sources. Moreover, this technique can resolve also synchronous noise closely

interacting with the investigated signal.

From the spectrum the complex value of the transform of the signal is taken at the

frequency of interest. This corresponds to the first coefficient c of the complex

Fourier series. For the signal i(t) given by equation (3.4):

T

0

tj dteiT

2c (3.5)

Where T is the recorded time.

The inverse of the recorded time represents the frequency steps in the frequency

domain. A low frequency signal or a dense frequency domain require a long

acquisition time. Moreover, according to the Nyquist theorem, the sampling rate has

to be at least twice as high as the highest investigated frequency in order to achieve

a perfect reconstruction of the signal in the discrete domain. This is connected with

the aliasing. Half of the sampling rate (called Nyquist frequency) represents the top

limit of the discrete Fourier transform in the frequency domain. After this frequency

the spectrum is simply repeated endless times.

If, for example, the investigated signal is at 1 kHz but the sampling rate is 200 Hz,

only one point is taken every five periods of the signal. Intuitively this leads to an

improper picture of the real signal. The signal cannot be reconstructed properly,

which leads to the appearance of some ghost waves (aliases) that actually do not

exist and are only a consequence of the low sampling rate. One can recall this

phenomenon in old western movies when the wheels of the wagons sometimes

seemed spinning in the opposite direction from their true rotation. The aliasing and

its implication are discussed in more detail in the Subsection 3.2.2 about

undersampling in the oscilloscope setup for the IDIS.

3.1.2.2 FREQUENCY RESPONSE ANALYZER

The frequency response analyzer (FRA) is probably the most used device to recover

the impedance in electrochemistry and it is usually integrated in most commercial

65

potentiostats. It employs the cross-correlation between the input signal i(t) and two

sinusoidal reference signals, Rin (Rin = cos(ωt)) and Rout (Rout = –sin(ωt)) at the

investigated frequency ω.

nT

0

indtRinT

2a (3.6)

nT

0

outdtRinT

2b (3.7)

The integral is performed on n periods T of the reference wave. The reference signals

are in quadrature and the results of the Equations (3.6) and (3.7) correspond to the

first two Fourier coefficients of a Fourier series expressed as:

cosIa (3.8)

sinIb (3.9)

The cross-correlation is extremely effective in rejecting the asynchronous noise and

the harmonics present in the signal. The accuracy in the recovery of the signal

increases with square root of the integration time nT.

3.1.2.3 PHASE-SENSITIVE DETECTOR

The phase-sensitive detector (PSD), also known as lock-in amplifier, is composed by

a frequency mixer and one or more filters. It has the particularity that it outputs two

signals. As for the FRA, two references in quadrature are used to recover the signal.

The frequency mixer multiplies the input signal with the references. The results, one

for each reference, pass subsequently to an integrator and to a low-pass filter.

Contrary to the FFT analysis and to the FRA, a lock-in amplifier does not provide a

value but two outputs, X and Y, which can be used to feed other instruments or

simply be recorded for elaboration. The outputs after the mixer and the integrator

are:

t

t-t

in

CC

dtRit

1X (3.6)

66

t

t-t

out

C C

dtRit

1Y (3.7)

Where tc represents the time constant of the integrator. Equations (3.6) and (3.7)

lead to the demodulation of i(t) as:

tscos2

ItX

X (3.8)

tssin2

ItY Y (3.9)

Which are two time dependent signals. sX(t) and sY(t) contain the higher harmonics

of the signal i(t) and those arising from the demodulation plus any other synchronous

and asynchronous noise. The value of the time constant is critical in the noise

rejection. A large tc allows for a clean demodulation, while a small one enables the

demodulation to follow fast variation in the signal.

The low-pass filters, added after the integrator, are necessary to remove sX and sY.

They apply their transfer function to X and Y, which is an important point to handle

the intermodulation as explained in Subsection 3.2.1 about the lock-in setup for the

IDIS. The variation of the impedance in time can be observed due to the

transformation of the lock-in amplifier, which is an important point in the time-

variant systems.

67

3.2 DEVELOPMENT OF THE IDIS INSTRUMENTAL SETUP

In this section, I show the principles of instrumental apparatuses developed to

perform an IDIS. The instrument can operate in two ways, lock-in setup and

oscilloscope setup which are described in the two subsections.

Figure 3–2 shows the schematic of the instruments necessary for the IDIS in both

setups [58].

The lock-in and the oscilloscope setup do have some parts in common. These are:

the probe Ω and the stimulus ω generators, the potentiostat, a multichannel digital

oscilloscope, and a personal computer to perform the FFT analysis. The details of the

two setups are shown in the following subsections. The specifications for the devices

which compose the instrument and the instrumental parameters are reported in the

Subsection 3.4.3

Figure 3–2: schematic of the instrument setup. Generator ω outputs the stimulus

signal and generator Ω the probe signal which is used as reference by the lock-in

amplifier. The potentiostat outputs the E- and I-monitor to feed the first two

channels of the oscilloscope. Furthermore, the I-monitor also feeds the lock-in

amplifier. The lock-in demodulates the current and sends the in phase X and out

of phase Y components to the second two channels of the oscilloscope. Through FFT analysis the IDIS spectrum is recovered.

68

3.2.1 LOCK-IN SETUP

The key part of lock-in setup is the phase-sensitive device that, as explained in

subsection 3.1.2.3, outputs two signals X and Y. The I-monitor of the potentiostat

feeds the input of the lock-in where the signal is further on demodulated according to

the probe generator as shown in Figure 3–2. Note that a 4-channel oscilloscope is

necessary in this configuration. The lock-in amplifier performs its activity only on the

probe signals. In this case, the first two channels of the oscilloscope are only

necessary to recover the stimulus signals. The best configuration for this setup

requires the splitting of the I-monitor in two and the use of four additional low-pass

filters (not shown in Figure 3–2). These low-pass filters are placed before the 4-

channel oscilloscope in such a way that all the recorded signals are filtered likewise.

This configuration was used by Keddam and Takenouti to implement the MICTF

technique [54–57]. It is useful to filter the probe signal from the E- and I-monitor,

to lower the sample frequency of the oscilloscope without incurring in aliasing

problems or in order to employ a multichannel FRA as done by Keddam and

Takenouti. The second connection of the I-monitor goes to the lock-in for the

demodulation. Besides, the X and Y outputs feed, possibly through the filters, the

oscilloscope.

The total current during an intermodulation experiment is:

ωΩωΩ

Ωω0

θtωΩcosIθtωΩcosI

θΩtcosIθωtcosIii

(3.10)

Where I is the current amplitudes, θ the phase shift, and the subscript refers to the

angular frequency to which the properties are related. The last two terms of the right

hand side of Equation (3.10) are the intermodulation sidebands. During the phase-

sensitive demodulation, the in phase component X and the out of phase component Y

from the lock-in amplifier are determined.

They are obtained by first multiplying i ∙ Rin and i ∙ –Rout, respectively, and then

integrating the results in the time constant tC, as described in subsection 3.1.2.3:

ωΩωΩωΩωΩΩ θωtcosI2

1θωtcosI

2

1θcosI

2

1X (3.11)

2

πθωtcosI

2

1

2

πθωtcosI

2

1θsinI

2

1Y ωΩωΩ-Ω (3.12)

69

Calculating the conductance G and the susceptance B from X and Y, the following

relations are obtained:

ωΩωΩΩΩ θωtcos

U

Iθωtcos

U

Iθcos

U

IG

(3.13)

2

πθωtcos

U

I

2

πθωtcos

U

Iθsin

U

IB ωΩωΩΩ (3.14)

Using the definitions of differential conductance and differential susceptance given in

Equation (2.39) and (2.40) and applying them to Equation (3.13) and (3.14), one

obtains:

ωΩ

ω

ω

ωΩ

ω

ω jθexpUU

Ijθexp

UU

I

2

1dG (3.15)

ωΩ

ω

ω

ωΩ

ω

ω jθexpUU

Ijθexp

UU

I

2

jdB (3.16)

Where Uω is the amplitude of the stimulus frequency and j the imaginary unit.

Equations (3.15) and (3.16) are equivalent to:

ω2U

ΩU

ωΩI

ΩU

ωΩI

dG*

*

(3.17)

ω2U

ΩU

ωΩI

ΩU

ωΩI

jdB*

*

(3.18)

So far two problems were disregarded: the phase-shift between the probe reference

and the potential signal inside the electrochemical cell, given by the bandwidth of the

control loop of the potentiostat, and the transfer function of the low-pass filter of the

lock-in amplifier. The first issue is easily manageable if the lock-in amplifier can be

synchronized with the probe frequency coming from the E-monitor. This is indeed the

70

case for the instrument built up during this work. Otherwise, an addition control

procedure, either performed during the experiment or aside, is necessary to get the

phase-shift of the control loop and to correct it.

The second issue requires a suitable way to find the transfer function of the built-in

filters of the lock-in amplifier. The easiest way is to compare the results of an IDIS

performed with the lock-in setup with those coming from the oscilloscope setup for

different time constants. This procedure is reported in details in subsection 3.4.4 and

the result is shown in the fourth chapter (Subsection 4.1.1).

The main advantage of the lock-in setup, especially if equipped with four additional

low-pass filters, is the possibility to have large probe frequencies and low stimulus

frequency at the same time without overloading the oscilloscope. Moreover, the

sensitivity of the PSD is higher than the one of the oscilloscope setup. A comparison

is shown in the Section 4.2.

The main disadvantage is the necessity of a 4-channel oscilloscope and eventually

additional low-pass filters. Furthermore, the lock-in amplifier places some limitations

on the stimulus and probe frequency. A ratio of one to ten is necessary to achieve a

suitable demodulation without overly reducing the time constant of the integrator,

and introducing a high level of noise.

3.2.2 OSCILLOSCOPE SETUP

The oscilloscope setup simplifies the instrumental design. Only a dual channel digital

oscilloscope is necessary and the IDIS is performed through FFT analysis. Another

advantage is the flexibility gained for the stimulus and probe frequencies. There is no

constrain in their relative values and the stimulus could even approach and cross the

probe frequency. The only limitation is the memory capability of the computer, where

datasets larger than few millions show to be difficult to handle.

There is one important aspect of the FTT analysis that is fundamental for

intermodulation: the spectrum leakage. The spectrum leakage is due to the finite size

of the recorded signal and appears as a blurring effect, given by the side lobes which

surround the frequency peaks in the F-domain. Figure 3–3 shows the spectrum

leakage of an IDIS dataset recorded on a Schottky diode. The side lobes extend to a

large part of the spectrum and easily obscure the intermodulated sidebands.

71

A common way to tackle this problem is to employ a windowing. This is achieved with

a window function w(t), which has a bell-shape that approaches zero at the edges of

the recorded time interval. This function multiplies to the digitalized data. In this

work a Blackman-Harris window, which is given by the following equation, was

employed:

T

t6πcosa

T

t4πcosa

T

t2πcosaatw 3210 (3.19)

Where T represents the total recorded time and the parameters are a0 = 0.35875, a1

= 0.48829, a2 = 0.14128, and a3 = 0.01168. The integral area of the function has to

be unitary. The window maintains a bell-shape also in the frequency domain, in

which the time multiplication is transformed into a convolution. This convolution

causes a slight broadening of the peaks, but reduces strongly the side lobes

flattening the background as in Figure 3–3. The window function can be seen as a

kind of band-pass filter applied on the frequency spectrum. The broadening of the

peaks is a favorite characteristic of the windowing which allows an easier detection of

high frequency signals recorded for a long time. In this case the width of the peaks,

which is proportional to the square root of the number of the recorded periods,

becomes smaller and smaller. When the peak is too sharp, a sophisticated search is

require in order to precisely allocate its position in the frequency domain. In fact, a

Figure 3–3: effect of the windowing. FFT of the current of the diode with a

stimulus frequency of 10 Hz and a probe frequency of 1 kHz performed with and without Blackman-Harris window function.

72

shift in the frequency of the signal as small as few decimals of hertz can produce a

wrong spotting of the peak in the frequency domain.

The disadvantage of the oscilloscope setup is that it has a lower sensitivity than the

lock-in setup. Also, a practical limit is imposed on the investigated frequency. As

already suggested previously, the size of the recorded data should be below few

millions of points to be processable by a normal computer. For example, a probe

frequency of 1 kHz, which implies a sampling rate of 10 kHz, and ten periods of a

stimulus frequency of 0.1 Hz requires a dataset collection of one million of points per

single value in an IDIS measurement. The limit of this approach is shown when

stimulus and probe frequency depart from each other of more than four orders of

magnitude.

To reduce the datasets, the undersampling can be employed. This involves the

sampling of a band-limited signal below its Nyquist frequency, which induces the

aliasing of the signal as shortly explained in the Subsection 3.1.2.1. The

undersampling can be seen as a folding of the frequency domain at the Nyquist

frequency, which is reproduced in the entire domain. Figure 3–4 shows a schematic

of the folding process that leads to the aliasing. The first two rows represent the FFT

Figure 3–4: schematic of the undersampling. The red triangles represent the

bandlimited signals and the green triangles are the aliases. In the first two rows

the signals a and b are sampled properly, the Nyquist frequency is higher than the

signal frequency; in the third row the signal b is undersampled and its Fourier

representation is indistinguishable from the one of signal a in the first row.

73

of the signal a and b performed respecting the Nyquist theorem. The signals are

band-limited for the sake of simplicity, i.e. they contain only a limited range of

frequencies. When instead the Nyquist theorem is not respected, as in the third row,

some aliases appear. In this situation, the signal a in the first row and the signal b in

the third row are not distinguishable.

In the undersampling, the aliasing is employed to recover a signal in a region of the

frequency domain different from that recovered with a proper sampling. In this way a

high frequency signal can be shifted to a lower frequency reducing the sampling rate

appropriately. There are two requirements to achieve the undersampling. The first

one is, that the band size of the signal to recover is preserved and the second, that

the ADC has an analogue bandwidth large enough to record the data properly.

Given a signal which extends from the frequency f1 to the frequency f2, as shown in

Figure 3–4, the first requirement is fulfilled by:

1

22 12

n

ff

n

fs

(3.20)

Where fs is the sampling frequency and n is an integer number representing the

undersampling magnitude. It is clear that n = 1 leads to the Nyquist theorem.

Equation (3.20) gives a set of possible sampling frequencies to employ an

undersampling of magnitude n and leads to:

12

21ff

fn

(3.21)

Which gives the set of available undersampling magnitude n. As for the example of

Figure 3–4 n is three.

In subsection 3.4.3, it is specified in which experiments the undersampling was

employed and at which frequency the probe frequency was folded.

74

3.3 ELECTROCHEMICAL CELL

In this section, I briefly discuss the main issues accompanying the experiments

conducted with electrochemical impedance spectroscopy (EIS) and, in the two

subsections, the solutions developed in this thesis.

The first problem is connected with the artifacts coming from the instrumental setup

(wiring, stray capacitance and others) that always arise in an electrochemical

experiment conducted with three electrodes and a potentiostat [118]. These usually

appear at high frequencies and are easily balanced by the use of a capacitor bridge

[119].

The second subsection deals with the improper positioning of the electrodes in the

cell. Here, also the particular geometry employed for the electrochemical

experiments of this thesis is described.

3.3.1 CAPACITOR BRIDGE

Figure 3–5 shows the schematic of the electric circuit of the three electrode

configuration proposed by Fletcher [118]. Aside the impedance of the three

electrodes (ZWE, ZCE, and ZRE), also the capacitive couplings between all the elements

of the cell are drawn (CCE/WE, CCE/RE, and CRE/WE). During an experiment, current flows

into the cell from the CE or WE node, while the potential is measured between the RE

and the WE node.

In an ideal configuration, ZCE and ZRE and all the capacitive couplings are zero. It is

clear that the impedance of the working electrode is measured precisely at any

frequency without any interference. This is not the real life scenario and all the parts

of the circuit contribute to the total impedance as soon as the experiment departs

from DC.

Two metallic objects always behave like a capacitor and this is the first source of the

capacitive coupling in the electrochemical cell. This shows that a three-electrode

setup is always worse, regarding stray capacitance, than a two electrode setup, since

more electrodes also means more couplings. Besides, the coupling is also increased

by the stray capacitance of the wirings which lead to the potentiostat. As discussed in

subsection 3.1.1, the input impedance of the reference in the potentiostat is not

infinite and shows a stray capacitance of 1-100 pF.

Problems arise when some current leaks into the RE electrode branch producing an

ohmic drop given by ZRE. Clearly, the impedance of the reference has to be kept as

small as possible, but usual reference electrodes have a resistance of 1-10 kΩ. The

ohmic drop changes the potential sensed at the reference node, which is not any

75

longer the same as at the central node. The leakage current is the result of the

capacitive coupling between WE and RE, which at high frequency becomes a possible

pathway for the current.

In his work, Fletcher suggested to keep all the capacitive coupling as small as

possible [118], but a wise observation of Figure 3–5 shows, indeed, the importance

of finite capacitance between the RE and the CE branches [119]. This can be used to

draw some current from the CE node directly toward the RE node to balance the

leakage current and decrease the ohmic drop in the reference branch. This is

achieved by the capacitor bridge.

The bridge is nothing else than a capacitor designed to short-circuit the RE and the

CE node increasing their capacitive coupling as shown in Figure 3–6. Of course, the

bridge has to be dimensioned accurately to avoid even greater artifacts [119]. In the

fourth chapter (Subsection 4.1.2), the effect of the capacitor bridge is shown.

3.3.2 COAXIAL CELL

In this subsection, I describe the electrochemical cell developed for the experiments

conducted in this thesis. A schematic of the electrode configuration is shown in Figure

3–6. The working electrode is a disk electrode, which is placed at the top end of the

Figure 3–5: schematic of the circuit suggested by Fletcher [116] to represent a three electrode cell. (Adapted from [119])

76

axis of the counter electrode. This is made out of a cylindrical platinum mesh. The

reference is placed outside the CE [119].

The CE behaves like a faradaic cage confining current and the electric field inside

leaving the outer region free of any electric perturbation. This is hence an

equipotential volume and represents the best place for the RE which can be placed

anywhere outside the CE.

The CE being composed by a mesh ensures ionic contact between the inner and the

outer solution. This configuration represents an optimum for the use of the capacitor

bridge. Inside the CE, the current is dominated by the cylindrical geometry, which

makes the current line distribution more homogeneous. In the case where the size of

the WE is small compared with the inner diameter of the CE, the cylindrical geometry

approaches the hemispherical one.

The configuration shown in Figure 3–6, which is referred as coaxial, has several

advantages. It reduces the problem of experimental reproducibility, because the

degrees of freedom for the position of the electrodes are minimized. The only

constrains are WE being inside the CE, which has to be cylindrical, and RE placed

outside. Strict control of the position of the WE and RE is not necessary.

The distortions arising from the current line distributions are minimized.

In the case where the WE is small compared with the diameter of the CE, this

geometry guarantees the smallest resistance of the electrolyte [119].

Some comparisons for EIS measured with different cell configuration are shown in

the fourth chapter (Subsection 4.1.2). There, it is evident also from a practical point

Figure 3–6: schematic of the coaxial cell geometry with the capacitor bridge

[119]. (Adapted from [119])

77

of view, that there are problems connected with an incorrect geometry which can

lead to erroneous results.

78

3.4 MATERIALS AND PROCEDURES

In this section, I report all the details about the experiments performed in this work.

The experiments are divided in four groups: cell geometry experiments, diode

experiments, redox couple experiments, and gas-evolving electrode experiments.

In the first subsection, all employed chemicals, electrolytes and electrodes are

shown. In subsection 3.4.2, the standard instruments and the procedures for the

experiments are listed. Subsection 3.4.3 reports the parts which compose the IDIS

instrument and the parameters employed for the experiments. The procedures to

recover the transimpedance of the potentiostat and the transfer function of the lock-

in amplifier are described in subsection 3.4.4. In the subsection 3.4.5 the best fit

approach of the EIS and IDIS spectra is explained and in the last part I report the

algorithm used to correct the electrolyte resistance.

3.4.1 CHEMICALS AND ELECTRODES

In order to prepare the solutions for the experiments, K2HPO4 (Fisher Chemical),

KH2PO4 (Sigma-Aldrich), KOH (J.T.Backer), KCl (J.T.Backer), K3Fe(CN)6 (Sigma-

Aldrich) and K4Fe(CN)6 (Sigma-Aldrich) were used.

For the cell geometry experiments, the working electrode was a platinum disc sealed

in a glass capillary. The total diameter of the electrode embodiment was

approximately 6 mm, while the diameter of the platinum disc was 1 mm. Two

solutions were employed. One, containing 10 mM equimolar K3Fe(CN)6 and K4Fe(CN)6

in 1 M KCl as supporting electrolyte. The second one, made from a 10 fold dilution of

the former, was supposed to simulate electrolytes which possess low conductivity.

The capacitor bridge consisted of two wires connected either through a capacitor in

the middle (the standard 100 nF bridge, total length 15 cm), or with a capacitor box

(C DEKADE, HCK, Essen). In both cases the extremities of the bridge were connected

to the metallic contact of the counter and reference electrode just next to the

standard connections of the potentiostat as shown in Figure 3–6.

For the diode experiments, a Schottky diode 80SQ040 (International Rectifier), as

ideal nonlinear system, and a dummy cell composed by a 9.4 MΩ resistor in parallel

with a 2 nF capacitor, as ideal linear system were used. The diode was connected to

the potentiostat with the cathode attached to the working electrode and the anode to

the reference and counter electrodes. In order to simulate the effect of an

uncompensated resistance (Subsection 4.2.3), a 28.7 kΩ was added in series to the

diode.

79

In the redox couple experiments, the electrolyte was a 0.5 M phosphate buffer

solution at pH 7 containing 10 mM of both iron complexes. This solution was kept

protected from light and stored in a fridge under argon atmosphere.

The WE was a 125 μm diameter platinum disk electrode embedded in glass. This was

polished with merge paper down to grade 2000 and subsequently with lapping films

of 1 and 3 m grade (3 M).

The solution for the gas-evolving experiments was 0.1 M KOH. The gas-evolving

electrode was a gold cavity microelectrode. The cavity was circa 100 μm in diameter

and 10 μm in depth prepared according to the procedure reported in reference [120].

This was filled with RuO2, a catalyst for oxygen evolution. Figure 3–7 shows a picture

of the electrode filled with RuO2. This electrode was placed facing upward in the

electrochemical cell to enhance bubbles departure from the surface.

The RE was a homemade Ag / AgCl / 3M KCl (potential 0.210 V vs. NHE) used with a

double junction for the redox couple experiments. The CE was a cylindrical mesh

(Labor Platina) made of a platinum iridium alloy with an inner diameter of 10 mm

and a height of 12 mm. This guaranteed a coaxial geometry as described in the

Subsection 3.3.2.

A 100 nF capacitor bridge was employed for all the experiments in 3-electrode

configuration to avoid high frequency artifacts in the impedance spectra, as explained

in the previous section (Subsection 3.3.1). The effect of different capacitor size is

shown in the fourth chapter (Subsection 4.1.2).

Figure 3–7: details of the cavity microelectrode filled with the catalysts.

80

3.4.2 STANDARD ELECTROCHEMICAL CHARACTERIZATION AND PROCEDURES

Three different potentiostats were used in this work: a Zahner Zennium (Zahner), a

VSP-300 (BioLogic), and a Solartron ModuLab potentiostat/galvanostat (AMETEK).

The first was employed for the experiments concerning the Schottky diode, the

second for the cell-geometry and redox couple experiments, and the third for the

experiments with the gas-evolving electrode.

The potentiostatic EIS performed on the commercial instruments were made between

1 MHz (100 kHz for the experiments with the diode) and 100 mHz with 10

frequencies per decade and 10 mV of potential perturbation.

For the experiments with the Schottky diode, the cyclic voltammetry were performed

at 10 mV s-1 between 0 V and 2 V. The diode was connected according to the IUPAC

official setup, with the cathode attached to the working electrode and the anode to

the reference and counter electrodes. The Mott-Schottky analysis to obtain the flat-

band potential of the diode was performed between 0 V and 2 V with an EIS every

100 mV.

For the redox couple experiments, the following procedure was employed. The

Fe(II)/Fe(III) cyanide solution was first placed in the sealed electrochemical cell and

sparged with argon while waiting to reach a stable temperature. As control, a series

of cyclic voltammetries were performed with a freshly polished working electrode and

a freshly annealed counter electrode at 50 mV s-1 between -250 and +250 mV vs.

the OCP (circa 250 mV vs. RE) until the voltammogram was stable. In the case

where the voltammograms were not satisfactory, the WE was removed and polished

one more time. When major problems appeared, also the solution was changed. After

the cyclic voltammetry also some EIS were performed as control.

With careful regard to the argon atmosphere, the electrochemical cell was moved to

the IDIS instrument where an additional EIS and the intermodulation were applied.

These were performed at the OCP. As final validation of the series of experiments, a

last EIS was performed in the end to ensure that the argon atmosphere was still in

good conditions and that poisoning was not occurring at the last stage.

For the experiments with the gas-evolving electrode the working electrode was

placed facing upward. The linear sweep voltammetry was performed between 0.95

VRHE and 1.75 VRHE at 5 mV s-1. Before passing to the intermodulation an EIS at 1.45

VRHE was performed to recover the resistance of the electrolyte without bubble

evolution. The intermodulation was made between 1.6 VRHE and 1.75 VRHE.

In the next subsection, I report the parameters of the IDIS and of the EIS used in

the experiments with the diode, the redox couple, and the intermodulation of the

gas-evolving electrode.

81

3.4.3 IDIS INSTRUMENT AND PARAMETERS

In this subsection I present the IDIS instrument composition along with the relevant

parameters employed during the experiments.

The instrument was composed by a potentiostat PG_310USB (HEKA Elektronik), a 2-

channels lock-in amplifier HF2LI (Zurich Instruments), a 4-channels oscilloscope

PicoScope 4424 (Pico Technology), and a personal computer to perform the FFT

analysis. As generator for the stimulus and the probe, the two internal generators of

the lock-in amplifier were used. The FFT analysis and the instrument control were

performed through a homemade collection of routines run with Matlab (MathWorks).

The probe generator was used as internal reference by the lock-in to demodulate the

high frequency signal and the control loop phase-shift was controlled and

synchronized before proceeding with the data acquisition routine.

To improve the resolution and dampen the high frequency noise, the additional 4th

order Bessel filter (10 KHz cutoff frequency) of the potentiostat was employed

together with a lower control loop bandwidth (10-30 kHz).

The routine automatically controlled the lock-in amplifier time constant to match a

stimulus to cutoff frequency ratio as close as possible to ω/Δω = 0.2 for every

stimulus frequency.

The sampling rate of the oscilloscope was between five and twenty times as high as

the probe frequency. The other parameters, such as stimulus, probe frequency,

amplitude, and the acquisition time, are reported in Table 3–1. The amplitudes

shown in Table 3–1 were higher than those effectively applied in the cell, especially

at high frequency because of the speed of the control loop (Subsection 3.1.1). As

example, an amplitude of 20 mV led to circa 13 mV at 10 kHz of real perturbation at

the ends of the WE and RE.

To ensure the maximum accuracy, the acquisition was delayed for every stimulus

frequency of at least one period or one second.

For the gas-evolving electrode experiments, the undersampling was employed to

reduce the dataset size. The probe and sideband signals were folded either to circa

120 Hz or to 1000 Hz, keeping the total size of the dataset in the range of one million

points.

In the FFT analysis, a procedure to find the right frequency position of the peaks was

employed. The stimulus and the probe frequency were first searched in the frequency

domain of the potential, looking for local maxima. The results were then used for the

FFT analysis of the current.

The potentiostat transimpedance and the lock-in transfer function were applied to the

result after the FFT analysis.

82

3.4.4 POTENTIOSTAT TRANSIMPEDANCE AND LOCK-IN AMPLIFIER TRANSFER

FUNCTION

In this subsection, the strategy to recover the potentiostat transimpedance and the

lock-in amplifier transfer function are described. The potentiostat transimpedance HTi

was recovered in a two-electrode setup with a dummy cell composed by a resistor

and a capacitor in series. The capacitor was 0.5 nF and the resistor was scaled

according to the gain of the current range, from 1 M for the 1 A current range to 1

k for 1 mA current range. Such a dummy cell showed to be more suitable to

recover the transimpedance and the obtained results were more reliable than those

coming from a dummy cell constructed by a simple resistor. The result of such a

procedure is reported in the fourth chapter (Subsection 4.1.1).

The transfer function of the lock-in amplifier is connected with the time constant of

the integrator and the bandwidth of the subsequent low-pass filter [58].

Naming Δω the cutoff frequency in rad s-1 of the filter of the lock-in amplifier, the

relevant parameter which controls the transfer function of the device with respect to

the stimulus frequency is given by ω/Δω. The relation between the time constant tC

and the cutoff frequency is:

C

t

AΔω (3.22)

Where A is a constant that depends on the kind and order of the low-pass filter. The

transfer function of the lock-in, HL(ω/Δω), was obtained from the ratio of the

Table 3–1: IDIS settings for the experiments

Stimulus Probe Acquisition time

Diode 40 mV /

0.1 – 100 Hz 20 mV / 1 kHz

10 stimulus

periods or at least

5 s

Redox couple 15 mV /

1 Hz – 15 KHz 25 mV / 20 kHz

10 stimulus

periods or at least

0.6 s

Gas-evolving

electrode

5 or 10 mV /

50 mHz – 15 kHz 20 mV / 10 kHz

15 stimulus

periods or at least

20 s

83

differential admittance measured from the lock-in setup to the one measured by the

oscilloscope setup for the Schottky diode for different values of Δω and ω/Δω. To

improve the data quality, this function was fitted with a fourth order inverse

polynomial, which was thereafter used for the proper correction.

3.4.5 FITTING PROCEDURE

The impedance performed in the cell geometry experiments were fitted with the EIS

Spectrum Analyser [121], whereas the fits for the intermodulation were performed

with a homemade routine based on Matlab. This employed the minimization of the Χ2

function, defined as:

N

i ii

iifitiifit

df YY

YYYY

n 12

exp,

2

exp,

2

exp,

2

exp,2

ImRe

ImImReRe1 (3.23)

Yfit and Yexp represent the fitting function and the underlying experimental function.

ndf are the degrees of freedom of the system given by the difference between the

number of measured points and the parameters employed in the fitting.

In the best fit of the impedance spectra, the procedure was performed on the

impedance Z, whereas in the IDIS spectra the differential conductance dG and the

differential susceptance dB were fitted simultaneously. For the fitting, the model

shown in the second chapter was employed (Subsection 2.3.5 Equations (2.55) -

(2.64)).

The electrolyte resistance and the impedance necessary to correct dG and dB were

recovered from the EIS fitting and employed directly through the Equations (2.81)

and (2.82) in the fitting function for the differentials.

3.4.6 UNCOMPENSATED RESISTANCE CORRECTION

In this subsection, I report the algorithm used to correct the IDIS spectrum in the

experiments with uncompensated resistance. The differential admittance dYexp and

the anti-differential admittance d exp were calculated through Equation (2.47) and

(2.48). Later on, the following equations were employed:

)(YR1)(YR1)(YR1

,dY,dY

expelexpelexpel

exp

0

(3.28)

84

)(YR1)(YR1)(YR1

,Yd,Yd

*

expel

*

expelexpel

exp

0

(3.29)

Where the subscription ―exp‖ means that the quantity was recovered neglecting the

resistance of the electrolyte and dY0 and d 0 were the quantities corrected for the

resistance.

85

4 RESULTS AND DISCUSSION

In this chapter, I report and discuss the experimental results of this thesis. In the

first section, the results concerning the potentiostat transimpedance and the lock-in

transfer function are reported. Besides, I show the effect of the capacitor bridge and

of a proper electrodes‘ disposition in the electrochemical cell to achieve reliable

impedance measurements.

The intermodulated differential immittance spectroscopy (IDIS) was applied as proof

of concept to an ideal nonlinear system, a Schottky diode (Section 4.2). This section

follows the equations described in the Subsection 2.3.4. Through intermodulation the

flat-band potential and the doping level of device were recovered. The diode was also

used as benchmark to determine the resolution of the instrumental setups and the

uncompensated resistance correction (Subsection 2.3.7).

Subsequently, I show the results concerning the redox couple experiments and how

the EIS and IDIS were used to recover the electrokinetic and mass transport

parameters (Section 4.3). In this section the model developed in the Subsection

2.3.5 is employed.

In the last part, I present the gas-evolving electrode and how the concepts of

intermodulation and transfer functions developed in the second chapter (Section 2.4)

are applied.

86

4.1 INSTRUMENT CALIBRATION AND ARTIFACTS IN IMPEDANCE

In this section, I show the preliminary results. These results deal with the

characterization of the instruments and of the electrochemical cell. Initially, I report

the transimpedance of the current follower of the potentiostat and discuss the effect

this has on the impedance measurements.

In the second subsection, I describe the transfer function of the lock-in amplifier

which was derived as reported in the third chapter (Subsection 3.4.4). I also discuss

the implication of dealing with a transfer function according to distortions, noise

rejection, and signal detection.

Finally, in the third part, I present the most common artifacts in impedance

spectroscopy which I introduced in Section 3.3. These arise from capacitive couplings

and improper electrodes‘ disposition. In particular, I show how to use the capacitor

bridge in order to mitigate the harmful combination of stray capacitance between

working and reference electrode and non-negligible resistance of the latter. Besides,

I report the advantages of using coaxial cell geometry where the counter electrode

plays the most important role.

4.1.1 POTENTIOSTAT TRANSIMPEDANCE

As reported in Subsection 3.1.1, the potentiostat does not behave ideally when it

operates at high frequencies. In this subsection, I describe the transimpedance of the

current follower of the potentiostat employed during this doctoral work. This was

recorded through different dummy cells as explained in detail in the third chapter

(Subsection 3.4.4).

Figure 4–1 shows the transimpedance HTi of the potentiostat for four current ranges

from 1 mA to 1 μA full scale without additional filters. The modulus of HTi was

normalized by the gain of the current range in order to plot all the curves in the same

graph (Figure 4–1–a). In the case of ideal behavior, Abs(HTi) should be unitary in all

the frequencies and the phase should be zero. Instead, the modulus decreased at

high frequencies. The cutoff frequency, which is the frequency where there is an

attenuation of -3 dB (≈ –30 %), was lower and lower for every current range. This

was in agreement with the fact that the bandwidth of a current follower was inversely

proportional to its gain. In fact, the cutoff frequency was circa 10 kHz and 100 kHz

for 1 μA and 10 μA full scale, respectively. For the higher current ranges the cutoff

frequency was above the investigated range.

The proportionality constant between gain and cutoff represented the bandwidth of

the operational amplifier that composed the current follower. This was a quality

87

identifier of the electronic device. In the case of the potentiostat used for this study

the bandwidth of the device was circa 10 MHz at 1 mA full scale which was one order

of magnitude higher than usual potentiostats.

Figure 4–1–b shows the phase of HTi. The transimpedance had a larger influence on

the phase of the signal than on the modulus. Although the deviation on the amplitude

of a signal at 10 kHz recorded with 10 μA current range was minimal, the phase was

already delayed of 11°. This was of fundamental importance when a proper Fourier

analysis was necessary. Besides, the phase bent at high frequencies, which implied

that the current follower was more complex than a simple operation amplifier in the

inverting amplifier configuration.

If one neglected the influence of the non-ideality of the potentiostat, it would record

higher impedance than the real one at high frequency. In fact, under these conditions

the gain of the current follower is smaller. Moreover, one would encounter problems

with the phase-shift of the results.

Figure 4–1: Bode plot of the normalized transimpedance HTi of the potentiostat used in this thesis.

88

As discussed in Subsection 3.1.1, the transimpedance is usually dependent on the

impedance of the investigated system. Figure 4–2 shows HTi for the 10 A current

range with two different dummy cells. These were composed by a 0.5 nF capacitor in

series with a 100 k and 1 k resistor, respectively. The absolute values started

diverging at 10 kHz and they intercepted the same values at 100 kHz. In fact, the

cutoff frequency seemed to be only minimally affected by the impedance of the

dummy cell. The phases diverged at 10 kHz and at 100 kHz the difference between

them was more than 20°. As expected, the influence of the different impedance was

stronger and stronger at higher frequencies.

This demonstrated the importance of adapting the calibration of the potentiostat with

a dummy cell that resembled closely the impedance of the investigated system,

especially at high frequencies.

Figure 4–2: comparison of the transimpedance of the 10 μA full scale current range measured with two different dummy cells.

89

4.1.2 LOCK-IN TRANSFER FUNCTION

The transfer function HL of the lock-in amplifier was recovered comparing the results

of the IDIS for the diode between the oscilloscope setup and the lock-in setup as

explained in the third chapter (Subsection 3.4.4) [58].

Figure 4–3 reports the Bode representation of the transfer function HL of the lock-in

amplifier. A fourth order inverse polynomial was used to fit the experimental data.

Later on, this polynomial was used for the correction of the intermodulation signals.

ω/Δω = 1 represented the normalized cutoff frequency. At this frequency, the

attenuation of the signal was -3 dB and the phase-shift was circa -90°. When ω/Δω

was greater than one, HL strongly attenuated and delayed the intermodulation signal,

however larger frequencies were measurable. On the other hand, for ω/Δω < 1, HL

tended to unity and the phase-shift tended to zero degrees. Although in this case the

transfer function had little influences, stimulus frequencies close to the probe

Figure 4–3: Bode plots of the lock-in amplifier transfer function HLI, with fourth

order inverse polynomial fit. a) absolute value; b) phase-shift. (Replotted from [58])

90

frequency were not accessible. For later experiments, a good compromise was found

with ω/Δω = 0.2. Therefore, the homemade software routine automatically set Δω as

close to 5∙ω as possible for every stimulus frequency. In this way it was possible to

measure dG and dB with higher precision than by using the oscilloscope setup.

4.1.3 CELL GEOMETRY AND CAPACITOR BRIDGE

In the two previous subsections, I showed the characterization of the potentiostat

and the lock-in amplifier, the instruments used during this work. In this subsection, I

discuss the problems arising during impedance spectroscopy. As explained in the

third chapter (Section 3.3), these had two sources. The first was the non-negligible

impedance of all the parts that composed an electrochemical system with three

electrode configuration. This could be regarded as an instrumental problem and it

was easily circumvented using the capacitor bridge [119]. The second source of

artifacts was the improper electrodes‘ disposition in the cell [119].

In order to understand the influence of the impedance of all the cell components, the

artifacts were mathematically simulated using the model given in Figure 3–4

(Subsection 3.3.1), provided by Fletcher and Sadkowski and Diard [118,122]. The

Nyquist plot in Figure 4–4 shows the ideal impedance of the WE without any

distortions compared to those affected by some artifacts, for the measurements with

and without the capacitor bridge. As ideal impedance a simplified Randles circuit

given by Rel – (Cdl / Rct) was employed, where ―–‖ and ―/‖ stand for series and

parallel connection, respectively.

In this case, the ideal impedance spectra appeared in the Nyquist plot as a semicircle

of diameter Rct, shifted from the origin along the x-axis by Rel. When the stray

capacitance between WE and RE was large (100 pF), a high frequency arc appeared

as predicted by Fletcher and Sadkowski and Diard [118,122]. This was merely an

instrumental artifact, the arc had no physicochemical meaning and its appearance

was due to the leaking current flowing into the RE branch toward the WE node. The

leaking current produced a voltage drop along ZRE which distorted the potential

sensed by the RE. Therefore, this artifact showed up only when there was RE-WE

stray capacitance and the RE possessed some non-negligible impedance. By

increasing the capacitive coupling between RE and CE, this artifact could be

suppressed. This was easily achieved by means of a capacitor bridge as explained in

Subsection 3.3.1 [119]. The bridge sank some current from the CE node towards the

RE node to balance the leakage current flowing in the reference branch. Surprisingly,

this improvement was neither suggested by Fletcher nor Sadkowski and Diard.

91

Instead, they recommended to keep all the capacitive coupling as small as possible

[118,122].

There was a drawback for having a large CE-RE coupling. This was visible when ZCE

was higher or comparable to ZWE. In Figure 4–4, the case of ZCE = ZWE with capacitor

bridge is presented. In this case, the impedance of the counter electrode influenced

the total impedance measured and an inductive loop followed the typical capacitive

loop [119]. Not surprisingly, a large CE with low impedance is also suggested by the

classic electrochemical manuals [123].

The following experiments were performed with a solution containing 10 mM

equimolar K3Fe(CN)6 and K4Fe(CN)6 in 1 M KCl as supporting electrolyte.

The effect of the capacitor bridge was visible also in a real electrochemical

experiment. Figure 4–5–a shows two EIS performed with the coaxial cell geometry.

The figure depicts the impedance of the system with and without the capacitor

bridge. When no bridge was employed, an arc was visible at high frequency. Instead,

when a 100 nF bridge was present, the high frequency arc disappeared and the high

frequency impedance tended towards a pure real value. The departure from the true

impedance appeared at frequencies above 20 kHz. It was not straightforward to

determine the point at which the artifact started. Therefore, to discard those points

which belong to the artifact in order to perform a best fit can be a doubtful

procedure.

In order to determine the size of the capacitor bridge, the knowledge of the

impedance of all the components of the cell and of all the capacitive couplings was

Figure 4–4: mathematical simulation of the impedance with and without artifacts. ZWE and Z with capacitor bridge completely overlap. (Replotted from [119])

92

required, which was not always the case. It was, instead, easier to perform an

empirical analysis. Figure 4–5–b reports the effect of different capacitor bridge

values. When the capacitor was too small, the high frequency part of the semicircle

intercepted the real axis at lower values as if the resistance of the electrolyte were

smaller. In this case, the semicircle appeared deformed. By increasing the

capacitance value, the intercept was shifted towards higher values and the semicircle

re-established its ideal shape. When the value of the bridge was too large instead,

the high frequency part was displaced from the real axis.

A value of 100 nF was chosen as an appropriate one and was, therefore, applied in all

the experiments of this thesis.

As reported in the third chapter (Subsection 3.3.2), the coaxial cell geometry had

the advantage that minimized the problem of reproducibility connected with the

position of the electrodes. In fact, once the RE was placed outside the zone where

the electric field was confined, its position was not relevant. To prove this point,

Figure 4–5: effect of the capacitor bridge. a) Nyquist plot of the impedance

performed with and without capacitor bridge; b) Effect of the size of the capacitor bridge on the impedance. (Replotted from [119])

93

three experiments with the coaxial cell geometry in a 10-fold diluted solution with a

100 nF capacitor bridge were performed. In these experiments, the position of the

reference electrode was changed. Lower electrolyte resistance strongly affected the

ohmic drop between WE and RE. It is usually suggested in the manuals to keep the

RE close to the WE to minimize the ohmic drop, especially when working with low

electrolyte conductivity [123]. Figure 4–6 shows three EIS spectra, first with the RE

placed in the proximity of the CE (approximately 1 mm distance), second at the

distance of approximately 1 cm, and third with the RE inside the platinum mesh of

the counter electrode.

As expected, the position of the RE had no influence on the measured impedance. In

fact, the curves with the RE next and far from the CE coincided. This confirmed that

the entire zone outside the CE was equipotential and, hence, that the current and the

electric field were successfully confined inside the CE mesh cylinder, leaving an

unperturbed equipotential volume outside. Interestingly, also the impedance

measured with the RE inside the CE shielding was perfectly overlapping with the

previous ones. This suggested that the quasi-spherical geometry was neither

sensitive to an insulator body placed in the middle of the electric field nor to the

different position of the RE.

In order to show the other advantages of the coaxial cell geometry, three EIS

measurements were performed with three different electrodes configurations:

coaxial, aligned, and triangular. Figure 4–7 reports the three measured spectra which

Figure 4–6: Nyquist plot of EIS performed with the coaxial cell geometry, with 100

nF capacitor bridge, and three different positions of the RE, once in the proximity of the CE, once far, and once inside the CE shielding. (Replotted from [119])

94

showed differences at all the frequencies. The coaxial and the aligned configuration

appeared to have similar Nyquist plot simply shifted along the real axis. However, the

spectrum was considerably different in the case of the triangular configuration. The

differences were highlighted by the results of the best fit reported in Table 4–1. The

fitting was performed using the following equivalent circuit:

diffctdlel

CPERCPER (4.1)

Where the double layer capacitance and the Warburg element were substituted with

constant-phase elements (CPE) in order to perform the fitting in the most ingenuous

way.

The resistance of the electrolyte changed for the three cases due to different

distances between WE and RE. Although in the coaxial cell geometry the two

electrodes were not placed in close proximity, the measured resistance of the

electrolyte was the lowest, as expected for a quasi-spherical geometry.

The electrodes‘ position strongly influenced the distribution of current lines and

potential, which gave a considerable difference in the charge transfer resistance for

the triangular configuration (Table 4–1). In fact, in this case Rct was 20 % lower than

with the other experiments. This should draw particular attention on the position of

the electrodes in the electrochemical experiments as already mentioned by

Nisancioglu [124].

Figure 4–7: Nyquist plot of EIS performed with different electrode configurations:

the coaxial geometry, triangular configuration, and with the RE in the middle. (Replotted from [119])

95

In the first section of this chapter, I showed the preliminary results concerning the

transfer function of the potentiostat and of the lock-in amplifier. Besides in the last

subsection, I described the problems connected with the stray capacitances and the

position of the electrodes in a three electrodes cell.

All the findings of this first section are applied in the next parts.

Table 4–1: best fit for the EIS spectra of Figure 4–7. (Adapted from [119])

Parameters Coaxial Aligned Triangular

Rele /ohm 46.7 52.8 53.3

Rct /ohm 64.4 63.9 49.2

Q (CPEdl) 1.37 10-6 1.11 10-6 7.23 10-7

n (CPEdl) 0.843 0.856 0.889

Q (CPEdiff) 3.65 10-4 3.63 10-4 3.65 10-4

n(CPEdiff) 0.484 0.484 0.490

96

4.2 IDEAL NONLINEAR SYSTEM: DIODE

In this subsection, I show the results concerning the Schottky diode as ideal

nonlinear system. In fact, the diode was stable, reproducible, and the

intermodulation could be predicted. This was used in order to study the dependence

of the capacitance on the potential which is given by the Mott-Schottky equation

reported in the second chapter (Equation (2.13)).

Initially, I report the characterization of the device through linear sweep voltammetry

in order to give a first qualitative analysis. Subsequently, the Mott-Schottky analysis

was employed to recover the flat-band potential of the diode. In the second part, this

value was compared with those calculated through the IDIS. This was performed

using both the oscilloscope and the lock-in setup as described in the third chapter

(Section 3.2). Besides, the same procedure was employed on a dummy cell

composed by passive elements, in order to evaluate the resolution limit of the

technique.

In the last part of this section, I used the uncompensated resistance correction (UR

correction) suggest in the second chapter (Subsection 2.3.7) in order to demonstrate

its liability. Additionally, I report some error considerations concerning this

procedure.

4.2.1 DIODE CHARACTERIZATION

A Schottky diode in reverse bias was used for the following experiments as described

in the third chapter (Subsection 3.4). The flat-band voltage of this diode was located

at circa -0.53 V. Initially, its characterization was carried out by linear sweep

voltammetry and Mott-Schottky analysis.

In Figure 4–8–a, the linear sweep voltammogramm of the diode at scan rate of 10

mV s-1 performed between 0 V and 2 V is reported. This voltage window corresponds

to the inverse region of the diode. The leakage current was equal to circa 400 nA at

0.5 V. In the same voltage range, the Mott-Schottky analysis was performed through

an impedance spectrum every 100 mV. This is shown in Figure 4–8–b. The parallel

capacitance was calculated from the imaginary part of the admittance at 100 kHz for

each potential. This was used to obtain the flat-band voltage and the dopant

concentration through the Mott-Schottky equation (Equation (2.13)).

97

From the linear regression of the Mott-Schottky plot, a flat-band potential of -0.536

± 0.005 V and a dopant concentration of 3.61 1017 ± 1015 cm-3 were calculated (εr =

11.68). The same results were achieved using lower frequencies (down to 1 kHz) and

restricting the potential range from 0.2 V to 2 V.

In the next subsection, I show how the same results are found with the IDIS.

4.2.2 IDIS OF THE DIODE: NONLINEAR CAPACITANCE

During an IDIS experiment two potential perturbations were sent to the system: the

stimulus and the probe. The former was spanned through a range of frequencies

while the latter was kept constant. In this case, the stimulus was between 0.1 and

100 Hz while the probe was 1 kHz. The experiments were performed at 0.5 V. From

the stimulus signal, the impedance spectrum of the system could be obtained.

Whereas, from the intermodulation sidebands, the nonlinear behavior of the diode

Figure 4–8: a) linear sweep voltammetry of the diode between 0 V and 2 V at 10

mV s-1; b) Mott-Schottky plot of the capacitance of the diode measured at 100 kHz with linear regression. (Replotted from [58])

98

was recovered. This behavior was connected with the potential dependence of the

capacity of the device.

Figure 4–9 reports the Nyquist plot of the impedance of the diode obtained through

IDIS together with the impedance measured with a commercial instrument at 0.5 V.

The two curves are very close, thus indicating the good quality of the data and that

the transimpedance HTI of the potentiostat was successfully corrected (Subsection

4.1.1 and 3.1.1). The parallel resistance measured by the IDIS impedance was

smaller than that coming from the commercial instrument because of the higher

amplitude of the stimulus perturbation. By fitting the impedance spectrum with a

capacitance parallel to a resistance, the values of 1.9 nF and 7.8 MΩ were obtained.

The frequency domain of the diode and of a dummy cell with similar impedance is

shown in Figure 4–10. The intermodulation sidebands which arose from the nonlinear

behavior of the system were visible only in the spectrum of the diode. For the

dummy cell, which was composed by linear elements, the sidebands, if present, were

completely buried under the noise level. This was in agreement with what discussed

in the second chapter (Subsection 2.3.1) about linear and nonlinear systems.

Besides, there was no second harmonic of the stimulus signal. This consolidates the

assumption that the leakage current was linear (Subsection 2.3.4).

Figure 4-9: Nyquist plot of the EIS of the diode performed with a commercial instrument and that recorded with the oscilloscope setup. (Replotted from [58])

99

Following the model for the intermodulation reported in the second chapter

(Subsection 2.3.4), Figure 4–11–a shows the Bode plot representation of the

differential conductance (dG) and of the differential susceptance (dB). They are

reported in their real and imaginary part, as a function of the stimulus frequency. As

discussed in the Subsection 2.3.4, the value of dG was imaginary, negative, and

increased with increasing stimulus frequency fω, while the value of dB was real,

negative, and constant with fω. In Figure 4–11–b, the differential admittance dY is

reported as a function of the stimulus frequency. dY was composed only by the

imaginary part, while the real part remained mostly near zero. Im(dY) increased at

higher stimulus frequencies, in accordance with the value of Im(dG) becoming larger.

Figure 4–10: frequency domain of the current of the diode and of the dummy cell

with a stimulus frequency of 10 Hz and a probe frequency of 1 kHz. (Replotted from [58])

100

From Equations (2.53) and (2.54) the flat-band voltage and the dopant level of the

diode could be calculated. These were equal to -0.535 ± 0.01 V and 3.92 1017 ± 8

1015 cm-3, respectively. Table 4–2 presents a summary of the results obtained with

Figure 4–11: IDIS spectrum of the diode. a) Bode representation of the

differentials; b) Bode representation of the differential admittance dY. (Replotted from [58])

Table 4–2: Flat-band potential, Ufb, and dopant concentration, ND, of the diode

calculated with the IDIS measured by the oscilloscope setup and the lock-in setup,

with the Mott-Schottky (M-S) analysis, and reported in the datasheet. (Adapted

from [58])

Ufb

(V)

ND

(1017 cm-3)

Tabulated data -0.53 not reported

M-S analysis -0.536 ± 0.005 3.61 ± 0.01

Oscilloscope setup -0.535 ± 0.01 3.92 ± 0.08

Lock-in setup -0.527 ± 0.006 3.88 ± 0.01

101

the Mott-Schottky analysis and with the two setups of the IDIS. They showed to be

all in good agreement.

As seen from Figure 4–9 the dummy cell did not show any intermodulation

sidebands. Therefore, it was used to estimate the resolution limit of the IDIS. This

limit was connected with the noise level of the instrumental setups. Figure 4–12

shows the results of the IDIS for the oscilloscope setup and for the lock-in setup

measured on the dummy cell. The value of dY was less than 1% of the measured

value in the case of the diode. Additionally, it was clear that the lock-in setup worked

better than the oscilloscope setup especially at low frequencies. In this case, the limit

of detectability of the IDIS was equal to circa 20 nS V-1.

So far the experiments were performed only with a diode which presented only a

nonlinear capacitive branch. In the next subsection, I show how the addition of a

passive element to the system influenced the intermodulation.

4.2.3 UNCOMPENSATED RESISTANCE CORRECTION

In the first reported experiments with the diode, the system did not possess any

element in series which could influence the intermodulation signals. This is not the

usual case for electrochemical systems where the nonlinear parts of the system,

faradaic and capacitive branches, are in series with the resistance of the electrolyte.

In this case, the intermodulation is altered by the electrolyte presence.

Figure 4–12: comparison of the resolution limit for the oscilloscope and for the lock-in setups. (Replotted from [58])

102

Following Equations (2.81) and (2.82) in the second chapter, the unperturbed

differential admittance and anti-differential admittance could be recovered from an

experiment with a non-negligible series resistance. In order to control the accuracy of

such a process, the IDIS was measured on the Schottky diode with and without a

28.7 kΩ resistor placed in series (R – diode). This corresponded to circa one third of

the impedance measured at the probe frequency and mimics the resistance of the

electrolyte in a standard electrochemical experiment. In particular, the ratio one to

three was chosen to fit the following experiments with the redox couple.

Figure 4–13 shows the differentials with and without resistance correction. Contrary

to Figure 4–11, where only the real part of dB was nonzero, in the case without UR

correction also the real part of dG departed from zero. Small deviations were present

also on the imaginary parts which changed slope. The resistance placed in series

phase-shifted and reduced the intermodulation sidebands, which mixed the features

of the differentials.

The differentials with UR correction resembled those visible in Figure 4–11,

confirming that the correction was successful. As additional proof, the flat-band

potential was calculated with the Mott-Schottky analysis and with an IDIS performed

on the diode without additional resistor. The results are presented in Table 4–3. The

flat-band potential was derived also from the corrected differentials and the value

was in good agreement with the formers as further proof of the goodness of the

correction. This was calculated substituting in the Equation (2.55) the interfacial

potential, which was equal to the applied one minus the ohmic drop.

Figure 4–13: effect of the UR correction on the system R – diode.

103

Knowing the proper value of the uncompensated resistance to correct was not always

straightforward. In order to estimate the effect of an improper correction the

Table 4–3: Flat-band potential, Ufb, calculated with the Mott-Schottky (M-S)

analysis and through the IDIS with the uncompensated resistance correction for a

diode with and without a resistor in series.

Ufb

(V)

M-S analysis -0.557 ± 0.005

IDIS diode -0.557 ± 0.007

IDIS R – diode with UR correction -0.554 ± 0.005

IDIS R – diode with URover correction -0.573 ± 0.005

IDIS R – diode with URunder correction -0.568 ± 0.005

Figure 4–14: IDIS spectrum of the R – diode with mis-estimated resistance. a) effect on dG; b) effect on dB.

104

procedure was repeated overestimating and underestimating the resistance of 20 %.

The results on the differentials are reported in Figure 4–14. The biggest effects were

visible on the real parts of the differentials. Re(dG) which should lay on zero, moved

positively or negatively with an overestimated and underestimated correction,

respectively. Re(dB), instead, moved always toward less negative values. On the

imaginary parts the effects were less marked and Im(dB) showed a marginal

deviation only at high frequencies, while Im(dG) was completely insensitive to the

magnitude of the UR correction. The Ufb calculated by the mis-estimated resistance

are also shown in Table 4–3. The Ufb were in both cases more negative that the real

one, which was in accordance with the fact that Re(dB) moved in both cases toward

less negative values.

In this section, I presented the results of the intermodulation on a diode. This was

taken as ideal nonlinear system. In the next subsection instead, I show what

happened in the case of a real electrochemical system.

105

4.3 REAL ELECTROCHEMICAL SYSTEM: REDOX COUPLE

In this section, I show the results of the intermodulation on a real electrochemical

system. A well-known and fast redox couple was chosen as target. The goal of the

experiments was to show whether the IDIS could be employed to recover, for

instance, the symmetry of barrier and the diffusion coefficients of the redox couple.

Besides, the results demonstrated that also the double layer responded to the

intermodulation and that it was possible to estimate the time constant of changing of

the capacitive current upon a potential variation.

First, I characterized the system through cyclic voltammetry and EIS. The resistance

of the electrolyte and the value of the mass transport limitation recovered from the

fitting of the EIS spectrum served in the second subsection. In the second

subsection, the results of the intermodulated differential spectroscopy were reported

together with the effect of the UR correction.

Finally, in the third subsection, I show the results of the best fit of the IDIS

spectrum. These follow what reported in the second chapter in the Subsection 2.3.5

and 2.3.6. The results demonstrated the possibility to employ the IDIS to obtain

information on the kinetics, mass transport, and double layer. These pieces of

information were not directly achievable through EIS. In fact, in order to quantify the

kinetic and mass transport parameters several impedance spectra were necessary.

These were usually obtained by changing either the potential or the composition of

the electrolyte. Therefore, some assumptions on the non-dependence of the

parameters on these factors were necessary.

In particular in the last subsection, I investigated several parameters which affected

the final results. For instance, the frequency extent in which the fitting was

performed and the resistance of the electrolyte employed in the UR correction greatly

influenced the quality of the fitting but had very small influence on the derivation of

the kinetic and mass transport parameters.

4.3.1 CHARACTERIZATION OF THE REDOX COUPLE

The experiments reported in this section were performed on a 125 μm diameter

platinum disk electrode in a 10 mM equimolar solution of K3Fe(CN)6 and K4Fe(CN)6

with 0.5 M phosphate buffer at pH 7 as supporting electrolyte. The coaxial cell with a

100 nF capacitor was employed as described in the Subsection 3.3.2 and 4.1.3.

Figure 4–15–a shows the cyclic voltammetry of the redox couple. The curve

resembled the typical sigmoid given by a reversible system. Because of the larger

diffusion coefficient of the oxidized species, the cathodic peak was greater than the

anodic. The existence of these peaks guaranteed that, despite the small dimension of

106

the working electrode (125 um), the assumption of semi-infinite diffusion profile was

sound.

The Nyquist plot of the EIS and its fit are reported in Figure 4–15–b. As expected,

the impedance spectrum resembles that shown in the second chapter (Subsection

2.2.2.2) in the Figure 2–1 where both kinetics and mass transport had a significant

role. Though there were some differences. First, the semicircle was deformed, which

was due to the distribution of time constants. Second, the Warburg impedance

deviated from the 45° straight line. This was due to the small dimension of the disk

electrode. At low frequency a radial more than a semi-infinite diffusion profile took

place. The frequency at which the use of a semi-infinite profile became unsatisfactory

was around 10 Hz.

As equivalent circuit for the best fit, the Randles model with semi-infinite diffusion

profile [116] was used. This was composed by the double layer capacitance Cdl in

parallel with the faradaic branch which was made of the charge transfer resistance Rct

Figure 4–15: characterization of the redox couple. a) cyclic voltammetry; b) EIS at the OCP with best fit.

107

and of the Warburg element ZW. These were in series with the resistance of the

electrolyte Rel. The equivalent circuit was summarized by:

Wctdlel ZRCR (4.2)

where ―–‖ and ―/‖ stand for series and parallel connection, respectively. The

normalized result of the fit was: Rel = 0.116 Ω cm2, Cdl = 25.4 μF cm-2, Rct = 0.291 Ω

cm2, and ζ = 19.9 Ω s-0.5 cm2 with a final Χ2 of 0.001. From Rct through Equations

(2.28) and (2.29) a k0 of 0.1 cm s-1 was derived which was in agreement with what

reported in literature [125].

Later on, the value of Rel was employed in the fitting of the IDIS spectrum and ζ was

used in combination with the symmetry of mass transport δ to calculate the diffusion

coefficients of the redox species.

4.3.2 IDIS OF THE REDOX COUPLE

In this subsection, I report the results of the IDIS performed on the redox couple.

This was carried out with a probe frequency of 20 kHz while the stimulus was

spanned from 1 to 15000 Hz at the OCP. The lower frequency limit for the stimulus

was dictated by the impossibility of employing the equations for semi-infinite

diffusion profile at low frequencies. As noted in the characterization (Figure 4–15–b),

10 Hz represented the lower boundary for this model in these conditions.

Figure 4–16: frequency domain of the current with a stimulus frequency of 10 Hz

and a probe frequency of 20 kHz.

108

The Fourier transform of the current around the stimulus and probe frequency is

reported in Figure 4–16 for a stimulus of 10 Hz. At low frequencies the stimulus

signal was visible together with the first two harmonics. These were circa two orders

of magnitude lower than the fundamental, which guaranteed small influences in the

EIS and that the small signal approximation was suitable (Subsection 2.2.2.2).

Interestingly, the third harmonic was greater than the second, which could give some

insight on the fact that symmetry of barrier should be close to 0.5. At high

frequencies, the sidebands are visible. As for the stimulus, also on the sidebands the

third order signals appeared and they were slightly greater than the second ones.

Interestingly, the sidebands were almost an order of magnitude higher than the

second harmonic, which was in contradiction with the causality factors introduced by

Bosch et al. [39]. In fact, in this case the ratio between the first sideband and the

second harmonic should be two. Additionally, the sidebands were also higher than

the power line noise at 50 Hz.

Figure 4–17 shows the differential immittance Bode plot where dG and dB are

reported in their real and imaginary part, as a function of the stimulus frequency with

and without UR correction. The UR correction was calculated starting from the

resistance of the electrolyte found with the best fit of the EIS (Figure 4–15–b).

As for the case of the diode, the uncompensated resistance affected primarily the

real and the imaginary part of dG and dB. Re(dG) shifted toward negative values,

while Re(dB) moved toward more positive values. Furthermore, both showed a

substantial changing in their inclination.

With the help of the mathematical simulation performed in the second chapter

(Subsection 2.3.6), it was possible to obtain some qualitative information from Figure

4–17–b which could be employed as starting point for the next fitting. From the

position of Re(dG), should be greater than 0.5. Re(dB) was nonzero, which was

connected with the derivative of the capacitance Δdl of the double layer and, although

Im(dB) seemed very flat, with the time constant ηdl of the double layer.

From the fact that the Re(dG) was negative and bent toward zero and that Im(dG)

was positive, it was deduced that the symmetry of mass transport δ was aligned with

and, therefore, it should be greater than 0.5.

In the next subsection, I show whether these findings were in agreement with the

results of the fitting of the IDIS spectrum.

109

4.3.3 FITTING OF THE DIFFERENTIALS

The findings concerning the intermodulation spectrum reported in the previous

subsection were only qualitative. These were made from the knowledge of the

mathematical simulation of the differentials performed in the second chapter

(Subsection 2.3.6). In this subsection, starting from the model proposed in

Subsection 2.3.5, the IDIS was fitted.

The best fit of the IDIS spectrum was performed on the curves of Figure 4–17–a

from 10 to 15000 Hz. The result is reported in Figure 4–18. The fitting of dG was

very good and both the real and the imaginary part were in good agreement with the

experimental data. The fitting of dB was less accurate especially on the imaginary

part at high frequencies.

In order to better visualize the fitting, the UR correction was applied. The curves

were not well-fitted at high frequencies and the discrepancies became noticeable

above 1 kHz.

Figure 4–17: effect of the UR correction on the IDIS of the redox couple. a) differentials without UR correction; b) differentials with UR correction.

110

To understand which parameters controlled the accuracy of the fitting in the high

frequency region, this was repeated changing the upper boundary frequency from

0.5 to 15 kHz. The results are reported in Table 4–4 together with Χ2. As predicted

previously (Subsection 4.3.2) and visible in Table 4–4, the symmetry of the barrier

was higher than 0.5 and it was aligned with the symmetry of the mass transport δ

which was also higher than 0.5. Moreover, the time constant of the double layer ηdl

was experimentally observed. This was associated with the variation of the double

layer Δdl.

The values associated with the faradaic current were rather insensitive to the

frequency extent of the fitting. On the other hand, the parameters of the capacitive

current changed. Although Δdl varied only of 10 %, meaning that it was a good

result, ηdl changed considerably.

Figure 4–18: best fit of the differentials with UR correction. a) fitting of dG; b) fitting of dB.

111

The fact that the best fit was less and less satisfactory at high frequencies was

evident also from the variation of Χ2. This was five times lower when the procedure

was performed between 10 and 500 Hz than when it was performed up to 15 kHz.

However, the kinetic and mass transport parameters and δ were only minimally

affected by the extent of the fitted curves. In this regard, the employed model

seemed to suit well the experimental data. Both and δ varied less than 0.2 % in

respect to the frequency range. Table 4–5 reports the diffusion coefficients of the

redox couple calculated from δ. These are close to what reported in literature for a 1

M KCl solution [126]. The variation in DRed and DOx was approximately 0.5 %. This

showed that in order to accurately recover the kinetic and mass transport parameters

it was not important to explore high frequencies. Moreover, the results demonstrated

that the kinetic and mass transport parameters were not responsible of the failure of

the best fit in the high frequency region.

On the other hand, the parameters connected with the capacitive current, that is the

variation of the double layer capacitance Δdl and the time constant of the charging of

the interface ηdl, were more sensitive to the upper frequency at which the fitting was

Table 4–4: results of the fitting as function of the upper boundary frequency.

Upper

Frequency

(kHz)

δ Δdl

(nF V-1)

ηdl

(s) Χ2

0.5 0.506 0.513 2.75 13.4 0.0073

1 0.506 0.513 2.78 15.4 0.0069

6 0.506 0.513 2.87 10.0 0.0170

10 0.506 0.514 2.95 7.54 0.0296

15 0.506 0.514 3.01 5.99 0.0401

Table 4–5: diffusion coefficients recovered from the fitting.

Upper

Frequency

(kHz)

DRed

(cm2 s-1)

DOx

(cm2 s-1)

0.5 6.77 10-6 7.50 10-6

1 6.77 10-6 7.50 10-6

6 6.75 10-6 7.52 10-6

10 6.74 10-6 7.53 10-6

15 6.74 10-6 7.54 10-6

112

performed. Especially ηdl showed a large variation. This demonstrated two things: the

double layer parameters were responsible for a good fitting of the high frequency

region and the model employed for the capacitive current was not accurate enough.

In particular, the variation in time of the concentrations of the redox species were

neglected in the second order expansion in the Subsection 2.3.5. Apparently, this did

not influence the derivation of the parameters of the faradaic current, but it was

more important in the case of the capacitive current.

Notably, the a priori separation of faradaic and capacitive current held also in the

results of the fitting. In fact, although the goodness of the best fit was influenced by

the extent of the investigated frequencies, the kinetic and mass transport parameters

were insensitive to this factor and their values were mostly connected with the fitting

of the low frequency region. The fitting procedure was extremely robust under this

aspect. On the contrary, the parameters connected to the capacitive current were

dependent on the extent of the investigated frequencies. Furthermore, the variability

of these parameters did not affect the results of the faradaic current.

Apart for the influence of the frequency range on the parameters recovered, the

fitting procedure was also affected by the resistance of the electrolyte. In order to

understand how the value of the resistance of the electrolyte influences and δ and

therefore DRed and DOx, the best fit was repeated changing the value for the UR

correction of 10 %. The results are reported in Table 4–6 together with those with

the resistance of the electrolyte derived from the EIS fit (Rel = 0.116 Ω cm2). It was

evident that an improper value of the resistance of the electrolyte strongly affected

Χ2. This rose of more than ten times in the case of a 10 % lower resistance of the

electrolyte and of more than twenty times instead for a 10 % higher resistance.

Although the quality of the best fit was strongly influenced by the UR correction,

and δ showed only minimal variations. In particular, and δ changed of 0.5 % and

DRed and DOx of 1 % in these cases.

Table 4–6: faradaic parameters recovered from the fitting as function of the

different electrolyte resistance used in the UR correction.

Electrolyte

resistance

( cm2)

α δ DRed

(cm2 s-1)

DOx

(cm2 s-1) Χ2

0.104 0.504 0.512 6.80 7.47 0.55

0.116* 0.506 0.514 6.74 7.54 0.04

0.123 0.508 0.516 6.70 7.58 0.83

*: value of the resistance of the electrolyte derived from the EIS fit.

113

So far only the influence of the resistance of the electrolyte and of the frequency was

investigated. However, DRed and DOx also depended on another parameter which was

recovered from the EIS: ζ. This, as already mentioned, represented the mass

transport limitation and had a strong influence on the derivation of DRed and DOx. In

fact, a variation of ζ of 10 % changed the values of the two diffusion coefficients of

20 %. Such a strong effect was due to the quadratic relationship in the equation of

the Warburg impedance and represented the main source of uncertainty for the

calculation of the diffusion coefficients through the IDIS.

In this section, the intermodulation was employed to perform the IDIS. From the

fitting of the IDIS spectrum some valuable information on the electrochemical system

were recovered. In particular, this showed to be a robust technique to obtain the

symmetry of barrier and the diffusion coefficients of the redox couple. In the next

section, the intermodulation is applied to the gas-evolving electrode, a time variant

system.

114

4.4 TIME VARIANT SYSTEM: GAS-EVOLVING ELECTRODE

I present in this section the part of the experiments concerning the gas-evolving

electrode understood as a time variant system. In these experiments the concept of

intermodulation was employed not to recover an intermodulated differential

immittance spectrum as done for the diode and the redox couple, but to study the

effect of the gas bubble formation, expansion, and departure on the impedance of

the electrode. In this case, the system was treated as a time variant system where

the scheduling parameter was the variation of the surface area of the electrode upon

bubble formation and departure as explained in the second chapter (Section 2.4).

In the first subsection, I show the characterization of the system with linear sweep

voltammetry and EIS. The EIS was performed before any bubbles were evolved. This

represented the starting point for the treatment. In fact, in this condition, the

resistance of the electrolyte is related to the total surface area of the electrode

normally available for the electrochemical reaction.

Later on, in the second part, I describe the problem as a time variant system.

Through digital filtering, the variables of the system: current, admittance and surface

area, are shown as function of time. This represented a first intuitive approach for

the further analysis performed in the last subsection.

In the third subsection, the concept of intermodulation and Fourier analysis were

used as developed in the second chapter (Section 2.4). Two results were noteworthy

in this part. The first one was the clear impedance spectrum of the system recovered

during bubble evolution. As reported in literature [79–83], this was not a trivial

problem. The second result was connected with the normalized impedance ZA the

transfer function described in the second chapter (Section 2.4). ZA permitted to

understand the influence of the bubble evolution on the impedance and the effect of

the current density.

4.4.1 CHARACTERIZATION OF THE OXYGEN EVOLUTION REACTION

All the experiments of this section were performed on a cavity microelectrode filled

with RuO2, a well-known catalyst for oxygen evolution in a 0.1 M KOH solution.

Figure 4–19 shows a linear sweep voltammetry performed on the cavity

microelectrode at 5 mV s-1. The curve was flat until 1.4 VRHE and then started to rise

due to water oxidation. Over 1.5 VRHE the first sign of the electrochemical noise was

visible and above 1.6 VRHE the current was strongly affected by the bubble evolution.

At 1.7 VRHE the slope of the voltammogram started to flatten.

115

Figure 4–20 reports the Bode plot of the impedance Z recorded at 1.480 VRHE. This

potential was sufficiently high to produce O2, but not high enough to generate gas

bubbles under potentiostatic control. The absolute value of the impedance in Figure

4–20–a was nearly flat from 50 kHz to circa 1 kHz and the phase-shift was close to

zero in this interval. In this region the influence of the resistance of the electrolyte

was visible. The value of 10 kHz was employed in the next step to track the variation

of the resistance of the electrolyte. The value of the resistance at 10 kHz was taken

as representative of the whole uncovered area of the electrode for later

normalization.

In order to visualize the reaction of oxygen evolution, the Nyquist plot of the EIS is

reported in in Figure 4–20–b. The spectrum shows a small semicircle at high

frequencies and then it ramps vertically at low frequency. This resembled a multistep

reaction with adsorbed intermediates [127] in which the first semicircle belongs to

the one step of the reaction and the part at the low frequencies represents the

beginning of a second larger semicircle which corresponds to another reaction step.

In the next subsection, the gas evolution was pushed to higher overpotentials and

gas bubbles generated at the electrode surface. In this case, the system was

interpreted as a periodic steady state. The variables as current and overpotentials

varied with time in a periodic fashion. The system was a time variant one. In the

next subsections, the features of such a system are unveiled.

Figure 4–19: linear sweep voltammetry of the cavity microelectrode filled with RuO2.

116

4.4.2 BUBBLE EVOLUTION AS A TIME VARIANT SYSTEM

When the current density was high enough, some oxygen bubbles formed at the

electrode surface. This was due to the rising of the supersaturation of dissolved gas

in the proximity of the electrode. Initially, the bubbles formed stochastically, but at

some high current densities their behavior became periodic. In fact, the formation,

growth and departure of the bubbles followed a periodic pattern. As explained in the

first chapter (Subsection 1.2.4) this gave rise to electrochemical noise which was

characterized by the variation in time of the current. Here I show the first approach

to this problem. The study was conducted for three different current densities which

were calculated dividing the current by the geometrical surface area.

In this subsection, I report the transient in time of the current i(t), of the real part of

the admittance Re(Yω(t)), and of surface area a(t) of the electrode. As example,

Figure 4–20: impedance of the system without bubble generation. a) Bode plot; b) Nyquist plot.

117

Figure 4–21 shows the transient recorded during bubble evolution with an apparent

current density of 33 mA cm-2. The gas bubbles formed and departed periodically

from the electrode and the noise frequency ωn was circa 1.1 Hz. The admittance was

measured at a frequency ω of 80 Hz and the surface area was tracked through a high

frequency Ω of 10 kHz.

As described in the second chapter (Section 2.4), the high frequency admittance was

used as scheduling parameter to track the variation of the surface area of the

electrode. By knowing the conductivity in the same conditions but without bubble

evolution, it was possible to transform the admittance to the area. The normalization

factor was obtained from the EIS measured at potential sufficiently high to achieve

oxygen evolution, but not high enough to generate bubbles at the electrode. This

factor was gained from the EIS shown in Figure 4–20.

The surface area transients and the Re(Yω) transient were recovered simulating

digitally the work of a lock-in amplifier reported in the third chapter (Subsections

3.1.2.3 and 3.2.1). The current was demodulated at both frequencies Ω and ω. This

was achieved by band-passing current and potential around Ω and ω with a second

order digital filter with a bandwidth of 10 Hz. Later on, current and potential were

mixed, normalized and low-pass filtered. The same low-pass filter was applied to the

current directly in order to clean the current transient of high frequencies signals.

The result is reported in Figure 4–21.

One problem was connected with the difficulty to change digitally the phase of the

signal. For this reason the Im(Yω) was not calculated. This required either the current

Figure 4–21: time transient of the current i(t), available surface area a(t), and real part of the admittance Re(Y(t)) during bubble evolution.

118

or the potential to be placed out of phase with the original signal. Moreover, to

handle the signal in time the transimpedance of the potentiostat had to be

considered through convolution.

The current i(t) was not a pure sinusoidal signal and, noise aside, resembled a

sawtooth wave. A similar shape was already reported for gas-evolving reactions [98].

The minima of the curve represent the time at which the bubble reached the maximal

size before departing. At this point, the surface was left free and the current could

rise rapidly. At the maxima of the current the bubble started to nucleate and the

growth was manifested by the slow decrease of the current. Similar behavior was

observed for a(t) which had a very similar shape as i(t).

All three transients were in phase. This was expected for i(t) and a(t). However, the

behavior of Re(Yω) was more complicated. In fact Re(Yω) was only the projection of

Yω(t) and this in general has no defined phase-shift in regard to a(t). Yω(t) was a

circular or elliptical polarized wave with the time coordinate as axis. The wave

oscillated both in the real and in the imaginary plane of the admittance. The ratio

between the real and the imaginary projection was correlated with the phase of the

polarized wave.

The curves show different levels of noise which came mostly from signals at higher

frequencies. The cleanest transient was a(t). In fact, no coherent signal was injected

at frequencies higher than Ω. Both Re(Yω) and i(t), instead suffered from the Ω signal

and in the case of i(t) also from the ω signal.

In Table 4–7, the parameters recovered by the current and surface area transients

are reported for the three apparent current densities. The geometrical area A0 of the

electrode was 78.54 10-6 cm-2. One could notice that the available surface area was

in average between 79 % and 87 % of this value, meaning that the electrode was

never completely free from the bubbles‘ influence.

Furthermore, the available surface area never reached A0 and approached 88 % of

this value as maximum. In fact, the electrode, because of its porosity, contained at

least one gas cavity. This cavity behaved as a stable and instantaneous source of

bubble nucleation, as explained by the non-classical type of nucleation [90]. Besides,

one should consider that the measured resistance of the electrolyte was not only

influenced by the surface area of the electrode, but also by the cross section of the

electrolyte in proximity of the electrode. As soon as a bubble departed from the

surface, this was for a very short time free (as short as permitted by the gas cavity),

but the bubble still occupied some volume in the electrolyte next to the electrode.

119

From the available surface area the actual local current was recovered. This was

considerably higher than the apparent one. It is interesting to compare the relative

variation of the current and of the surface area. At lower current densities there is a

large discrepancy, whereas at the highest current densities the two variations are

very close. This is connected with the fact that the screening effect of the bubble is

predominant at high currents densities [100,101].

This treatment had the advantage of showing the transients as function of time and

therefore of being more intuitive. The disadvantage was connected with the

impossibility of using proper filters. Even if a real lock-in amplifier was used, some

problems with the frequency range would rise. In fact, a phase-sensitive detector

cannot work when for example ω approaches Ω or when ω approaches the noise

frequency ωn.

This problem was circumvented in the next subsection, where the use of a transfer

function permitted to skip the time representation of the transients and to recover

the relation between the impedance and the electrode coverage.

4.4.3 NORMALIZED IMPEDANCE

As seen in the previous part, the time representation of the transients gave only a

first idea of the system through averaged values. Besides, the development of such

approach had some limitations. In this subsection, I demonstrate how to overcome

these shortcomings. This was possible through Fourier analysis and introducing a

new transfer function ZA. This was explained in the second chapter (Section 2.4) and

represented the impedance of the system normalized by the available surface area of

the electrode not covered by the gas phase.

Figure 4–22 shows the Fourier domain of the current used to derive Figure 4–21. The

frequency range was divided into three parts. At the lowest frequency, the spectrum

Table 4–7: parameters recovered from the current i(t) and area a(t) transients for

different apparent current densities.

Apparent

current

density

(mA cm-2)

Noise

frequency

(Hz)

Local

current

density

(mA cm-2)

Variation of

current

density

(%)

Available

surface

area

(10-6 cm2)

Variation of

surface area

(%)

23 1.0 27 0.5 68 1.3

33 1.1 40 1.0 67 1.7

44 1.2 59 2.1 62 2.3

120

contained the electrochemical noise. The time representation of this section produces

i(t) in Figure 4–21. As already observed, the current resembled a sawtooth wave

which contained odd and even harmonics of the fundamental frequency ωn. This

frequency was circa 1.1 Hz and the harmonics extended up to 10 Hz. This was in

contradiction with the assumption of the electrochemical noise being a pure

sinusoidal wave and it affected the intermodulation. In fact, as shown later, the

harmonics overlap with the sidebands at low frequencies disturbing the

measurements.

In the middle part of the spectrum, there were the signals connected with the

stimulus frequency ω at 80 Hz. One could notice the intermodulation of the

electrochemical noise with the stimulus signal, which showed several sidebands. The

demodulation of these sidebands provided the sawtooth-wave shape of Re(Yω) in

Figure 4–21.

The last part of the frequency domain of Figure 4–22 shows the high frequency probe

signals. These were used to track the variation of the electrolyte conductivity which

was connected with the occupation of the surface area of the electrode by the gas

bubbles. As for the case of the stimulus, the sidebands of the electrochemical noise

were present. These represented the intermodulation of the bubble dynamics with

the high frequency signal. The demodulation of this frequency region generated the

transient of a(t) in Figure 4–21.

Although not shown in Figure 4–22, also the stimulus intermodulated with the probe.

However, the electrolyte conductivity was a rather linear term and these

intermodulation sidebands were up to two orders of magnitude smaller than those

coming from the electrochemical noise.

Following the treatment reported in the second chapter (Section 2.4), a new transfer

function ZA was introduced. This was the impedance normalized for the variation of

the surface area a(t). The trend of the absolute value and of the phase of ZA

described the effect of the bubble dynamics on the oxygen evolution reaction. This

together with the value of the average impedance Zm elucidated the complex

dynamics of the reaction.

The study was conducted for three different current densities. These current densities

were high enough to locally supersaturate the solution, which led to the formation of

gas bubbles. After starting in a point, the bubbles went on to nucleate always at the

same spot in the cavity microelectrode usually because a single gas cavity was

present. In the range of current density which was investigated, the bubble formation

and departure were periodic. The periodicity was between 1 and 1.2 Hz as reported

in Table 4–7.

121

It was clear from the frequency representation of Figure 4–22 that to recover the

impedance in such a condition was not a trivial task. This was possible only through

precise Fourier analysis in which the peaks of current and potential were carefully

identified. Both FRA and PSD technique might fail in such a condition, either because

of the impossibility to resolve the current modulation, as in the case of the FRA, or

because the signal could not be demodulated properly when the impedance

frequency approached the noise frequency, as discussed previously (Subsection

3.1.2.3, 3.2.1 and 4.4.2) for the case of the lock-in amplifier.

Figure 4–23 shows the Nyquist plot of Zm for the three experiments. All curves

resembled a more or less deformed semicircle. This gave a first insight about the

dynamics of the reaction. No clear diffusion profile was visible in the plot meaning

that the reaction was always controlled by the electrokinetics. Similar findings were

reported also by Gabrielli for hydrogen evolution under high current densities [114].

The reason was that the bubble eruption did not allow the formation of a continuous

Nernst diffusion layer and mass transport of reagents and products never affected

the reaction.

The semicircle corresponded to a multistep reaction as already reported in literature

for oxygen evolution reaction also for lower current densities [128]. The size of the

semicircle shrank with the rise of the current density. This was in accordance with

the fact that electrokinetics limitation decreased when the current rose.

Figure 4–22: frequency domain of the current used to derive Figure 4-21.

122

Zm represented the impedance averaged in time of the system. Although the bubble

formed, grew, and departed changing at any time the surface area of the electrode,

Zm considered a mean electrodic surface. On the other hand, ZA was the impedance

normalized by the true surface area free from the bubbles.

Figure 4–24 shows the comparison of ZA and Zm for the three current densities. As

expected from Figure 4–23, the absolute value of Zm was a sigmoid with the part at

high frequency flat. In this part the impedance was dominated by the resistance of

the electrolyte and ZA was very close to Zm. This was in agreement with the fact that

the bubble blocked the electrodic area and reduced the cross section of electrolyte in

contact with the electrode increasing the ohmic overpotential. At circa 100 Hz, Zm

rose due to the impedance of electrokinetics and started flattening around 1 Hz as

expected from the closing of the deformed semicircles of Figure 4–23. However, ZA

remained flat until 10 Hz and at 1 Hz rapidly rose. At low frequencies, ZA was

different for the different current densities. While at 23 and 33 mA cm-2 ZA tended

toward Zm (Figure 4–24–a and –b), at the higher current density of 44 mA cm-2 ZA

rose higher than Zm (Figure 4–24–c). This was in agreement with what reported in

literature. In fact, at high current densities the primary effect of the bubble was of

decreasing the surface area of the electrode influencing both activation and ohmic

overpotentials [100,101]. Whereas, at low current densities the effect on the

activation overpotential is not so marked.

Figure 4–23: averaged impedance Zm for three different current densities.

123

Figure 4–25 shows the phase-shift of ZA measured at the three different current

densities and the phase of Zm at 23 mA cm-2 as comparison. The quality of the data

for ZA between 1 and 10 Hz was too low to recover a clear and reliable phase. The

reason was that in that frequency region there were several harmonics of the

electrochemical noise as one could notice from Figure 4–22. These interfered with the

Figure 4–24: modulus of ZA and Zm for the three different current densities. a) 23 mA cm-2; a) 33 mA cm-2; a) 44 mA cm-2;

124

intermodulation sidebands of the stimulus. Although this had low influence on the

absolute values: the data are only minimally scattered, the phase was more sensitive

to disturbances. Therefore, the points between 1 and 10 Hz were discarded.

Furthermore, the phase of ZA was shifted from 180° to 0°. As explained previously,

at high frequency the influence of a bubble on the electrode was merely that of

blocking the surface area and of reducing the cross section of the electrolyte

increasing the ohmic and the activation overpotentials. When the bubble grew, the

blocking effect rose and the surface a(t) of the electrode available for the

electrochemical reaction shrank. Therefore, when the bubble grew, a(t) decreased

and the current decreased too, which meant that the impedance rose. Changing the

phase of ZA of 180° was equivalent to consider the size of the bubble, rather than the

available surface area a(t).

The phase-shift of ZA showed the same behavior for all the current densities. At high

frequency the phase was zero, which meant that the growth of bubbles and

impedance were in phase: when the bubble became larger the impedance became

larger too. Between 100 and 10 Hz the phase decreased and, although not

completely visible because of the missing points, it reached a minimum at around 1

Hz. This minimum was circa -90°. In this frequency range, impedance and bubble

dynamics were completely out of phase. When one was increasing the other was

decreasing and vice versa.

Figure 4–26 exemplifies this point. The time variant impedance (t) and the bubble

size b(t) are represented as sinusoidal waves. At the time t1, (t) had a minimum

Figure 4–25: phase-shift of ZA at different current densities compared to the phase of Zm at 23 mA cm-2.

125

and b(t) was decreasing at the maximum rate. This corresponded to the time when

the bubble was leaving the surface of the electrode. In fact, this event was abrupt

and the electrode surface was passing from being occupied by a large bubble to be

completely free in a short time span. Given the violence of the event this was also

associated with a large stirring in the vicinity of the electrode surface. On the other

side, at the point t2, the bubble was growing at its maximum rate which was one

more time associated with large stirring of the solution. In the meantime the

impedance was facing a maximum and then decreased because of the enhancement

of convection.

Coming back to Figure 4–25, at low frequencies the phase-shift of ZA tended toward

zero. This was in accordance with the causality principle. In fact, at zero frequency,

current, and therefore impedance, had to be in phase with the bubble formation. The

phase of ZA unveiled that the bubble formation, growth, and departure influenced,

through the mass transport, the oxygen evolution reaction. The only difference laid in

the way the modulus of ZA behaved at low frequencies as seen in Figure 4–24.

In this last part of the chapter, I showed how to employ the intermodulation and

Fourier analysis to the study of periodic electrochemical noise as that produced by

gas bubble evolution. Besides, the problem of having reliable EIS spectra in these

conditions was circumvented.

It was shown that the bubble evolution had a positive effect on the electrochemical

oxidation of water because of the enhancement of mass transport. This effect was

Figure 4–26: schematic representation of the impedance 𝑍(t) and of the bubble

size b(t).

126

present at all the current densities, but at the highest density the decreasing of the

surface area of the electrode overcame it.

127

4.5 FINAL REMARKS

In this chapter, I described and discussed the findings of this thesis. In the first

section I showed which were the major problems encountered in an impedance

measurement. These were connected both with the instrumental setup and with the

electrochemical cell. In particular, I reported the results of the transimpedance of the

potentiostat and how this affected the EIS. Furthermore, I showed the influences of

the proper design of the electrochemical cell on the impedance. This demonstrated

the importance of the capacitor bridge and of the wise placement of the electrodes.

In the first part I also reported the transfer function of the lock-in amplifier. This was

particularly important for the use of the lock-in setup in the IDIS.

In the second part the IDIS was applied to a diode. This was taken as ideal nonlinear

system and represented the benchmark of the technique. Both instrumental setups of

the IDIS were tested with this system and both showed good performances.

From the IDIS spectrum it was possible to recover the flat band potential and the

doping level of the device. These were in agreement with what was calculated

through the Mott-Schottky analysis. Besides, the UR correction described in the

second chapter was tested on the diode. It proved to work properly and this allowed

its employment for the following experiments.

After the diode the IDIS was applied, in the third section, to a real electrochemical

system. This was composed by a redox couple in solution as ideal reversible system.

The UR correction and the model developed in the second chapter were applied to

this system. It was possible to recover the parameters of the faradaic and capacitive

current. The symmetry of the barrier and of the mass transport, the variation and the

time constant of the double layer capacitance were all unveiled by the results.

The model proved to be robust in regard to the faradaic parameters and the results

were insensitive to the variation of the fitting procedure. On the other hand, the

capacitive parameters were more affected by the high frequency part of the spectrum

and suffered more from the fitting procedure.

In the last section I applied the intermodulation to the study of the electrochemical

noise. This was generated by the periodic formation of gas bubbles at the electrode

upon strong oxygen evolution. Following the principles reported in the second

chapter, the system was analyzed as a time variant system. The surface area of the

electrode was employed as scheduling parameter. This was recovered by the high

frequency part of the impedance spectrum and later used to normalize the

impedance. The normalized impedance was used to understand the effect of the

bubble dynamics on the oxygen evolution reaction. Two features were discovered.

First, the bubbles enhanced the mass transport of the products. This feature was

common to all the experiments. Second, in the experiment performed at higher

128

current density the normalized impedance rose higher than the mean impedance.

This was correlated with the fact that the bubbles partially blocked the surface of the

electrode reducing the area available to sustain the electrochemical reaction affecting

the activation overpotential, which overcame the positive effect of the increased

mass transport.

The next chapter is the conclusion of this thesis. It summarizes all the work and

suggests further development and applications for the IDIS.

129

5 CONCLUSION

In this chapter, I first give a short summary of the work treated in this thesis and

then describe its main achievement, followed by some further developments.

In the first chapter, I reported a short overview of the state of the art regarding the

nonlinear analysis (Subsection 1.2.2), the electrochemical noise (Subsection 1.2.3),

and the gas-evolving electrodes (Subsection 1.2.4).

In the second chapter, I described the mathematical framework which was developed

for the Intermodulated Differential Immittance Spectroscopy (IDIS). Two examples,

the diode (Subsection 2.4.4) and the redox couple (Section 2.5), were used to

illustrate the application and the modus operandi of the technique. The results

concerning these experiments were reported in the fourth chapter (Section 4.2 and

4.3).

In the last part of the second chapter, I showed how the intermodulation was used to

investigate a time variant system (Section 2.6). The case of a gas-evolving electrode

was analyzed. The results of this model were described in the last part of the fourth

chapter (Section 4.4).

The experimental part was described in the third chapter. There, the instrumental

setups developed to perform the IDIS are discussed (Section 3.2). Furthermore, I

explained the major sources of disturbances in working with AC signals. These can be

given by the instrumentation (Section 3.1) or by the electrochemical cell (Section

3.3).

In the following section, I report the main contributions achieved in this work and I

suggest some further developments.

130

5.1 MAIN CONTRIBUTIONS

The aim of this doctoral thesis was to introduce, characterize, and apply a new

electrochemical technique, IDIS, designed to implement the electrochemical

impedance spectroscopy.

The IDIS was based on the phenomenon of the intermodulation. This phenomenon

appears when two periodic stimuli with different frequencies interact in a nonlinear

system creating an amplitude modulation.

Three new transfer functions were derived. These were the differential conductance,

the differential susceptance, and the differential admittance which represented the

variation of the equivalent quantities with the potential. They were correlated with

the nonlinearity of the system and provided some insights into its properties.

The IDIS represented the merging of the electrochemical frequency modulation

(EFM) proposed by Bosch and Bogaerts [38,39] and the modulation of interface

capacitance transfer function (MICTF) technique proposed by Keddam and Takenouti

[54–57]. In fact, the EFM and the MICTF technique were designed to study only the

faradaic current and only the capacitive one, respectively. Furthermore, the transfer

functions and the mathematical framework developed in this thesis overcame the

limitations of these techniques and of several nonlinear approaches.

The IDIS was first applied to the study of an ideal nonlinear system, a Schottky

diode. In this case, the flat-band potential and the doping level of the device were

recovered. These were in good agreement with the tabulated data and with the

results coming from the Mott-Schottky analysis. Interestingly, the Mott-Schottky

analysis requires a series of impedance spectroscopy performed at different

potentials, whereas in the IDIS only a single experiment was necessary.

These experiments showed the reliability of the technique and of the

instrumentation. Furthermore, in a counter test performed on a dummy cell, no

intermodulation was detected. This experiment was used to quantify the error level

for the technique which was less than 1%.

Subsequently, the IDIS was applied to the study of an ideal electrochemical system,

the electrochemical reaction of a redox couple, in order to quantify the parameters of

the faradaic and of the capacitive current. The faradaic current is controlled by the

electrokinetics and by the mass transport. With the IDIS, it was possible to recover

the symmetry factor of the electrochemical reaction and the diffusion coefficients of

the redox couple. Apart from that, also the variation of the double layer capacitance

upon potential change and its time constant were experimentally measured. All these

quantities were not directly accessible via impedance spectroscopy and mostly

overlooked in the nonlinear analysis.

131

This achievement was possible via the mathematical framework developed for the

analysis of the differential immittance spectra. The system was modeled on the basis

of the a priori separation of the faradaic and capacitive current. Under this

assumption, it was possible to divide the faradaic current into electrokinetic and mass

transport contributions.

This model was one of the few examples where the nonlinearities of the double layer

were considered. Furthermore, a procedure to eliminate the effect of the electrolyte

resistance was suggested and proved.

The model was implemented as a best fit algorithm. This procedure was extremely

robust in regard to the faradaic parameters. In fact, these suffered only marginally

from the different factors of the fitting. However, these factors did affect the

capacitive parameters which were successfully recovered, but showed large

dependence on the fitting procedure.

The intermodulation was also applied to the study of the macrokinetic effect of gas

bubbles on the oxygen evolution reaction. In this case, the periodic bubble formation,

growth, and departure from the electrode surface produced the electrochemical noise

observed in the current transient. Through precise Fourier analysis, it was possible to

record the impedance spectrum of the reaction also in the conditions of

electrochemical noise. The impedance showed that, stirring the solution near the

electrode, the bubbles cancelled the mass transport limitation. Furthermore, the

system was modeled upon a periodic nonlinear time variant model. The surface area

of the electrode was taken as scheduling parameter. In fact, under bubble evolution,

this was periodically blocked by gas phase. The impedance normalized by the real

surface area of the electrode was recovered. This showed the effect of the bubble

evolution on the electrochemical reaction. The effect was different for different

current densities. In fact, at low current densities, the main result was the stirring of

the solution and the cancelling of the mass transport limitation, whereas, at high

current densities, this effect was overcome by the obstructing of the electrode

surface.

Besides, two instrumental setups to perform the IDIS were developed. In the first

one, a lock-in amplifier was used to increase the resolution of the technique. In the

second one, the spectra were recovered by Fourier analysis. This second setup

offered more flexibility in the experiments. In fact, it had fewer limitations in the

choice of the investigated frequency, whereas, with the lock-in amplifier, there was a

practical limit in the ratio of the stimulus to the probe frequency.

Some strategies to perform the IDIS were discussed. In order to achieve high quality

results, several points had to be taken into account. First, the potentiostat and the

lock-in amplifier imposed their fingerprints on the signals. In this work, the way to

132

recover and correct these fingerprints was discussed and reported. Second, the

measurement of AC perturbation suffered from artifacts, especially at high

frequencies. These could be corrected by the use of a capacitor bridge. Besides, a

proper geometry for the electrochemical cell was suggested. This geometry

guaranteed several advantages such as low resistance of the electrolyte, high

reproducibility, and low distortions.

In the next section, I suggest some developments based on the work reported in this

thesis.

133

5.2 FURTHER DEVELOPMENT

There are several implementations for the IDIS. For instance, a multisine excitation

could be associated with the intermodulation. This excitation is composed by several

harmonic perturbations sent simultaneously to the system. An example of such an

approach is the odd random phase multisine [129]. The multisine improves the

experimental efficiency of the technique, because more frequencies can be

investigated at once. Besides, it offers a sophisticated analysis of the errors which is

of great importance in a second order analysis. In this case, the signals to seek are

usually very small compared with the first order terms. It is also interesting to extend

the concept of differentials transfer function to the second harmonics. These contain

the same information of the intermodulation sidebands, but in a different

combination.

There are various examples where the IDIS could be employed successfully. For

instance, it can be applied to the study of mass transport in lithium ion batteries. In

this case, it could be possible to recover the diffusion coefficients of the ions into the

host electrode. Another possible application is the study of multistep reactions. The

feature of the IDIS of characterizing the capacitive current could give some insight

into those reactions which undergo adsorption. For instance, it could be possible to

recover the reaction path in the hydrogen evolution reaction.

Because the IDIS can recover the time constant of the interfacial capacitance, it

could be used in the study of transport and the trapping of carriers in

semiconductors. In fact, in this case the capacitance is frequency dependent.

So far, the stimulus frequency was spanned whereas the probe was kept constant. It

is interesting to change both frequencies and to allow the stimulus to cross the

probe. Stimulus and probe are completely exchangeable: the sidebands appear

around the highest frequency signal. When stimulus and probe are close, a quartet of

peaks appears in the frequency domain.

A fascinating development is the generalization of intermodulated differential transfer

function. In fact, one of the electric perturbations can be substituted with another

signal. There are several possible combinations. For instance, the intermodulation

can be coupled with the electrohydrodynamic impedance. In the electrohydrodynamic

impedance, a rotation speed perturbation of the electrode is employed [130]. This

affects the mass transport at the interface. In this case, the stimulus can be given by

the rotation speed perturbation and the probe remains a potential signal. With this

configuration, it is possible to investigate the influence of mass transport on the

impedance. In particular, the effect of the diffusion of electroactive species on the

double layer should be visible.

134

A second possible combination for the intermodulation can be with the intensity-

modulated photocurrent spectroscopy (IMPS) [131]. This technique employs a

sinusoidal modulation of the intensity of the incident light in order to investigate the

response of photoactive materials. In this case, the light modulation could act as

stimulus while the probe is a high frequency potential signal. This could be used to

investigate the changes in the capacitance of the semiconductor upon light

excitation.

135

APPENDIX

A. MASS TRANSPORT OPERATOR

In this appendix, I report how to derive the operator mi used to link the

concentration of the electroactive species in solution to the faradaic current. This

operator was derived by the work of Rangarajan who casted an unified formalism ―to

arrive at the results for even the most complex models without tears‖ [68]. In his

work, he first employed a matrix system which allowed the use of some operators to

deal with the mass transport. The advantage is that one can cast all the equations of

a system without the need to specify immediately the law of diffusion. One can

simply add or modify it at the very end. Here, a brief summary of the operators used

by Rangarajan in the simplest case is reported.

First, the variation of the concentration cj is linked to the flux J:

kk,jj Jm~cd (A.1)

In Rangarajan formulation, j,k represents the operator that link the variation of the

concentration to the flux. This operator contains the information about the kind of

diffusion. The second step is to link the flux J of the species j to the faradaic current

iF:

Fjj diJ (A.2)

One has to take care of the indexes j, k, and l according to the size and to the

direction of the vectors and of the matrixes. With a semi-infinite diffusion the

operator in its Fourier representation becomes:

dRe

Ox

Dj

10

0Dj

1

m~ (A.3)

It is seen as operator because it depends on the frequency ω at which the flux

oscillates. The vector ε is given by:

136

nF

10

0nF

1

(A.4)

The utility of this vector can be seen in more complicated cases when, for instance,

there is simultaneous production, consumption, and diffusion of different species.

Combining Equation (A.1) and (A.2) and plugging them into Equation (A.3) and (A.4)

one obtains:

F

F

dRe

Ox

dRe

Ox

di

di

nF

10

0nF

1

Dj

10

0Dj

1

dc

dc (A.5)

Which leads to:

F

dRe

F

Ox

dRe

Ox

diDjnF

1

diDjnF

1

dc

dc (A.6)

Although this formalism is extremely powerful, a downgraded version is employed in

the main text. Instead of using an operator to link the concentration to the flux

and a second vector ε to link the flux to the faradaic current, the two passages are

collected into a single operator. Furthermore, the matrix formulation is abandoned

for a more trivial list of diffusion operator mRed and mOx, one for every electroactive

species. According to this simplification Equation (A.6) is converted into:

F

dRe

F

Ox

FdRe

FOx

diDjnF

1

diDjnF

1

dim

dim (A.7)

Which is the final expression for the operator m used in the main text within

Rangarajan notation. There is a great loss in respect to the generality for this

formalism which cannot be easily applied to more complicated cases. However, this

simplification considerably streamlines all the calculations.

137

B. INTERMODULATED DIFFERENTIAL IMMITTANCE SPECTROSCOPY IN

IMPEDANCE FORMAT

In this appendix, I show how to derive the differential impedance from the

differential admittance. The differential admittance is defined as the derivative of the

admittance Y(u) on the potential u:

uYdu

d,dY

(A.8)

Where the subscription indicates the frequency at which the impedance is measured.

In the same way, the differential impedance is given by the derivative of the

impedance Z(u) on the potential u.

uZdu

d,dZ

(A.9)

Substituting Y(u) into Equation (A.9) and deriving:

uY

,dY,dZ

2

(A.10)

This equation links the differential impedance with the admittance and the differential

admittance calculated at the same frequency.

138

C. RELAXATION OF THE A PRIORI SEPARATION OF FARADAIC AND CAPACITIVE

CURRENT

In this appendix, I report the general case of intermodulation in an electrochemical

system without the assumption of a priori separation of faradaic and capacitive

current.

The current during intermodulation is given by:

ωXHΩXωΩXJωΩI TT (A.11)

Where J represents the Jacobian, X the vector of variables, and H the Hessian. This

equation substitutes Equation (2.21) of the main text. J and X are defined as:

Ox,0oRed,Ox,0oRed,

T

c

i

c

i

u

i

c

i

c

i

u

iJ

(A.12)

And

Ox,0oRed,Ox,0oRed,

T CCUCCUX (A.13)

The Hessian is composed by:

2

Ox,0

2

oRed,oOx,

2

oOx,

2

Ox,0

2

oOx,oRed,

2

2

Red,0

2

oRed,

2

oRed,

2

oOx,

2

oRed,

2

2

22

Ox,0

F

2

Red,0

F

2

Ox,0

2

oRed,

22

Ox,0

F

2

Red,0

F

2

2

F

2

c

i

cc

i

uc

i00

uc

i

cc

i

c

i

uc

i00

uc

i

cu

i

cu

i

u

i00

uu

i

00000uc

i

00000uc

i

cu

i

cu

i

uu

i

cu

i

cu

i

u

i

H

(A.14)

139

The mass transport operator used in the main text (Equation (2.11)) cannot be

employed in this treatment. In fact, in this condition not only the Fick‘s law, but also

migration has to be considered. Therefore, the equations become extremely complex.

140

BIBLIOGRAPHY

[1] D.D. Macdonald, Reflections on the history of electrochemical impedance

spectroscopy, Electrochimica Acta. 51 (2006) 1376–1388.

[2] H.W. Bode, Network analysis and feedback amplifier design., New York,

1949.

[3] E. Warburg, Ueber das Verhalten sogenannter unpolarisirbarer Elektroden

gegen Wechselstrom, Annalen Der Physik. 303 (1899) 493–499.

[4] D.C. Grahame, The Electrical Double Layer and the Theory of

Electrocapillarity., Chemical Reviews. 41 (1947) 441–501.

[5] M. Senda, P. Delahay, Rectification by the Electrical Double Layer and

Adsorption Kinetics, J. Am. Chem. Soc. 83 (1961) 3763–3766.

[6] R. Neeb, Intermodulations-Polarographie, Naturwissenschaften. 49 (1962)

447–447.

[7] D.J. Kooyman, M. Sluyters-Rehbach, J.H. Sluyters, Linearization of the

current/voltage characteristic of an electrode process. application to relaxation

methods, Electrochimica Acta. 11 (1966) 1197–1204.

[8] K. Darowicki, Fundamental-harmonic impedance of first-order electrode

reactions, Electrochimica Acta. 39 (1994) 2757–2762.

[9] J.-P. Diard, B. Le Gorrec, C. Montella, Impedance measurement errors due to

non-linearities—I. Low frequency impedance measurements, Electrochimica Acta. 39

(1994) 539–546.

[10] L. Mészáros, G. Mészáros, B. Lengyel, Application of Harmonic Analysis in the

Measuring Technique of Corrosion, Journal of The Electrochemical Society. 141

(1994) 2068–2071.

[11] K. Darowicki, Corrosion rate measurements by non-linear electrochemical

impedance spectroscopy, Corrosion Science. 37 (1995) 913–925.

[12] K. Darowicki, Frequency dispersion of harmonic components of the current of

an electrode process, Journal of Electroanalytical Chemistry. 394 (1995) 81–86.

[13] K. Darowicki, The amplitude analysis of impedance spectra, Electrochimica

Acta. 40 (1995) 439–445.

[14] K. Darowicki, Theoretical description of fundamental-harmonic impedance of

a two-step electrode reaction, Electrochimica Acta. 40 (1995) 767–774.

[15] J.-P. Diard, B.L. Gorrec, C. Montella, Comment on ―Application of Harmonic

Analysis in the Measuring Technique of Corrosion‖ [J. Electrochem. Soc., 141, 2068],

Journal of The Electrochemical Society. 142 (1995) 3612–3612.

141

[16] D.K. Wong, D.R. MacFarlane, Harmonic impedance spectroscopy. Theory and

experimental results for reversible and quasi-reversible redox systems, The Journal

of Physical Chemistry. 99 (1995) 2134–2142.

[17] J.-P. Diard, B. Le Gorrec, C. Montella, Deviation from the polarization

resistance due to non-linearity I - theoretical formulation, Journal of Electroanalytical

Chemistry. 432 (1997) 27–39.

[18] J.-P. Diard, B. Le Gorrec, C. Montella, Deviation of the polarization resistance

due to non-linearity II. Application to electrochemical reactions, Journal of

Electroanalytical Chemistry. 432 (1997) 41–52.

[19] J.-P. Diard, B. Le Gorrec, C. Montella, Deviation of the polarization resistance

due to non-linearity. III—Polarization resistance determination from non-linear

impedance measurements, Journal of Electroanalytical Chemistry. 432 (1997) 53–62.

[20] J.-P. Diard, B. Le Gorrec, C. Montella, Non-linear impedance for a two-step

electrode reaction with an intermediate adsorbed species, Electrochimica Acta. 42

(1997) 1053–1072.

[21] D.A. Harrington, Theory of electrochemical impedance of surface reactions:

second-harmonic and large-amplitude response, Canadian Journal of Chemistry. 75

(1997) 1508–1517.

[22] K. Darowicki, J. Majewska, The effect of a polyharmonic structure of the

perturbation signal on the results of harmonic analysis of the current of a first-order

electrode reaction, Electrochimica Acta. 44 (1998) 483–490.

[23] J.-P. Diard, B. Le Gorrec, C. Montella, Corrosion rate measurements by non-

linear electrochemical impedance spectroscopy. Comments on the paper by K.

Darowicki, Corros. Sci. 37, 913 (1995), Corrosion Science. 40 (1998) 495–508.

[24] K. Darowicki, J. Majewska, Harmonic Analysis Of Electrochemical and

Corrosion Systems-A Review, Corrosion Reviews. 17 (1999) 383–400.

[25] E. Van Gheem, R. Pintelon, A. Hubin, J. Schoukens, P. Verboven, O. Blajiev,

et al., Electrochemical impedance spectroscopy in the presence of non-linear

distortions and non-stationary behaviour: Part II. Application to crystallographic

pitting corrosion of aluminium, Electrochimica Acta. 51 (2006) 1443–1452.

[26] J.R. Wilson, D.T. Schwartz, S.B. Adler, Nonlinear electrochemical impedance

spectroscopy for solid oxide fuel cell cathode materials, Electrochimica Acta. 51

(2006) 1389–1402.

[27] M. Kiel, O. Bohlen, D.U. Sauer, Harmonic analysis for identification of

nonlinearities in impedance spectroscopy, Electrochimica Acta. 53 (2008) 7367–

7374.

[28] V.F. Lvovich, M.F. Smiechowski, Non-linear impedance analysis of industrial

lubricants, Electrochimica Acta. 53 (2008) 7375–7385.

142

[29] H. Xiao, S.B. Lalvani, A Linear Model of Alternating Voltage-Induced

Corrosion, Journal of The Electrochemical Society. 155 (2008) C69–C74.

[30] A.T. DeMartini, A. Unemoto, T. Kawada, S. Adler, Nonlinear Analysis of the

Oxygen Surface Reaction and Thermodynamic Behavior of La, in: ECS, 2009: pp. 47–

67.

[31] T. Kadyk, R. Hanke-Rauschenbach, K. Sundmacher, Nonlinear frequency

response analysis of PEM fuel cells for diagnosis of dehydration, flooding and CO-

poisoning, Journal of Electroanalytical Chemistry. 630 (2009) 19–27.

[32] W. Lai, Fourier analysis of complex impedance (amplitude and phase) in

nonlinear systems: A case study of diodes, Electrochimica Acta. 55 (2010) 5511–

5518.

[33] C. Kreller, M. Drake, S.B. Adler, H.-Y. Chen, H.-C. Yu, K. Thornton, et al.,

Modeling SOFC Cathodes Based on 3-D Representations of Electrode Microstructure,

in: 2011: pp. 815–822.

[34] M.F. Smiechowski, V.F. Lvovich, S. Srikanthan, R.L. Silverstein, Non-linear

impedance characterization of blood cells-derived microparticle biomarkers

suspensions, Electrochimica Acta. 56 (2011) 7763–7771.

[35] N. Xu, J. Riley, Nonlinear analysis of a classical system: The double‐layer

capacitor, Electrochemistry Communications. 13 (2011) 1077–1081.

[36] C. Montella, Combined effects of Tafel kinetics and Ohmic potential drop on

the nonlinear responses of electrochemical systems to low-frequency sinusoidal

perturbation of electrode potential – New approach using the Lambert W-function,

Journal of Electroanalytical Chemistry. 672 (2012) 17–27.

[37] N. Xu, D.J. Riley, Nonlinear analysis of a classical system: The Faradaic

process, Electrochimica Acta. 94 (2013) 206–213.

[38] R.W. Bosch, W.F. Bogaerts, Instantaneous Corrosion Rate Measurement with

Small-Amplitude Potential Intermodulation Techniques: Corrosion, Corrosion. 52

(1996) 204–212.

[39] R.W. Bosch, J. Hubrecht, W.F. Bogaerts, B.C. Syrett, Electrochemical

Frequency Modulation: A New Electrochemical Technique for Online Corrosion

Monitoring: Corrosion, Corrosion. 57 (2001) 60–70.

[40] S.S. Abdel-Rehim, K.F. Khaled, N.S. Abd-Elshafi, Electrochemical frequency

modulation as a new technique for monitoring corrosion inhibition of iron in acid

media by new thiourea derivative, Electrochimica Acta. 51 (2006) 3269–3277.

[41] E. Kuş, F. Mansfeld, An evaluation of the electrochemical frequency

modulation (EFM) technique, Corrosion Science. 48 (2006) 965–979.

143

[42] L. Han, S. Song, A measurement system based on electrochemical frequency

modulation technique for monitoring the early corrosion of mild steel in seawater,

Corrosion Science. 50 (2008) 1551–1557.

[43] K.F. Khaled, Application of electrochemical frequency modulation for

monitoring corrosion and corrosion inhibition of iron by some indole derivatives in

molar hydrochloric acid, Materials Chemistry and Physics. 112 (2008) 290–300.

[44] M.A. Amin, S.S. Abd El Rehim, H.T.M. Abdel-Fatah, Electrochemical frequency

modulation and inductively coupled plasma atomic emission spectroscopy methods

for monitoring corrosion rates and inhibition of low alloy steel corrosion in HCl

solutions and a test for validity of the Tafel extrapolation method, Corrosion Science.

51 (2009) 882–894.

[45] K.F. Khaled, Evaluation of electrochemical frequency modulation as a new

technique for monitoring corrosion and corrosion inhibition of carbon steel in

perchloric acid using hydrazine carbodithioic acid derivatives, Journal of Applied

Electrochemistry. 39 (2009) 429–438.

[46] A. Rauf, W.F. Bogaerts, Monitoring of crevice corrosion with the

electrochemical frequency modulation technique, Electrochimica Acta. 54 (2009)

7357–7363.

[47] A. Rauf, W.F. Bogaerts, Employing electrochemical frequency modulation for

pitting corrosion, Corrosion Science. 52 (2010) 2773–2785.

[48] N.A. Al-Mobarak, K.F. Khaled, M.N.H. Hamed, K.M. Abdel-Azim, Employing

electrochemical frequency modulation for studying corrosion and corrosion inhibition

of copper in sodium chloride solutions, Arabian Journal of Chemistry. 4 (2011) 185–

193.

[49] A. Rauf, E. Mahdi, Evaluating Corrosion Inhibitors with the Help of

Electrochemical Measurements Including Electrochemical Frequency Modulation, Int.

J. Electrochem. Sci. 7 (2012) 4673–4685.

[50] A. Rauf, E. Mahdi, Reliability and Usefulness of Causality Factors of

Electrochemical Frequency Modulation, Int. J. Electrochem. Sci. 7 (2012) 10108–

10120.

[51] P. Beese, H. Venzlaff, J. Srinivasan, J. Garrelfs, M. Stratmann, K.J.J.

Mayrhofer, Monitoring of anaerobic microbially influenced corrosion via

electrochemical frequency modulation, Electrochimica Acta. 105 (2013) 239–247.

[52] M.N. El-Haddad, Chitosan as a green inhibitor for copper corrosion in acidic

medium, International Journal of Biological Macromolecules. 55 (2013) 142–149.

[53] S. Junaedi, A. Al-Amiery, A. Kadihum, A. Kadhum, A. Mohamad, Inhibition

Effects of a Synthesized Novel 4-Aminoantipyrine Derivative on the Corrosion of Mild

144

Steel in Hydrochloric Acid Solution together with Quantum Chemical Studies,

International Journal of Molecular Sciences. 14 (2013) 11915–11928.

[54] R. Antaño-Lopez, M. Keddam, H. Takenouti, A new experimental approach to

the time-constants of electrochemical impedance: frequency response of the double

layer capacitance, Electrochimica Acta. 46 (2001) 3611–3617.

[55] R. Antaño-Lopez, M. Keddam, H. Takenouti, Interface capacitance at mercury

and iron electrodes in the presence of organic compound, Corrosion Engineering,

Science and Technology. 39 (2004) 59–64.

[56] H. Cachet, M. Keddam, H. Takenouti, R. Antaño-Lopez, T. Stergiopoulos, P.

Falaras, Capacitance probe of the electron displacement in a dye sensitised solar cell

by an intermodulation technique: a quantitative model, Electrochimica Acta. 49

(2004) 2541–2549.

[57] E.R. Larios-Durán, R. Antaño-Lopez, M. Keddam, Y. Meas, H. Takenouti, V.

Vivier, Dynamics of double-layer by AC Modulation of the Interfacial Capacitance and

Associated Transfer Functions: IMPEDANCE SPECTROSCOPY AND TRANSFER

FUNCTIONS, Electrochimica Acta. 55 (2010) 6292–6298.

[58] A. Battistel, F. La Mantia, Nonlinear Analysis: The Intermodulated Differential

Immittance Spectroscopy, Analytical Chemistry. 85 (2013) 6799–6805.

[59] C.R. Kreller, S. Adler, Insight into Rate Limiting Mechanisms of Porous La, in:

ECS, 2009: pp. 2743–2752.

[60] T.J. McDonald, S. Adler, (Invited) Theory and Application of Nonlinear

Electrochemical Impedance Spectroscopy, ECS Transactions. 45 (2012) 429–439.

[61] S.K. Rangarajan, High amplitude periodic signal theory: I. Potential

controlled, Journal of Electroanalytical Chemistry and Interfacial Electrochemistry. 56

(1974) 55–71.

[62] S.K. Rangarajan, Non-linear relaxation methods: I. An operator formalism,

Journal of Electroanalytical Chemistry and Interfacial Electrochemistry. 56 (1974) 1–

25.

[63] S.K. Rangarajan, Non-linear relaxation methods: II. A unified approach to

sinusoidal and modulation techniques, Journal of Electroanalytical Chemistry and

Interfacial Electrochemistry. 56 (1974) 27–53.

[64] M. Sluyters-Rehbach, J. Struys, J.H. Sluyters, On second order effects in a

glavanic cell: Part II. A unified treatment of the theoretical principles of the second

order techniques for the study of electrode processes and for electrochemical

analysis, Journal of Electroanalytical Chemistry and Interfacial Electrochemistry. 100

(1979) 607–624.

145

[65] J. Struijs, M. Sluyters-Rehbach, J.H. Sluyters, On second-order effects in a

galvanic cell: Part IV. A theory of demodulation polarography, Journal of

Electroanalytical Chemistry and Interfacial Electrochemistry. 143 (1983) 37–59.

[66] R.S. Hansen, K.G. Baikerikar, Frequency dependence of capacitance at the

desorption potentials of organic molecules, Journal of Electroanalytical Chemistry and

Interfacial Electrochemistry. 82 (1977) 403–412.

[67] E.J.F. Dickinson, R.G. Compton, How well does simple RC circuit analysis

describe diffuse double layer capacitance at smooth micro- and nanoelectrodes?,

Journal of Electroanalytical Chemistry. 655 (2011) 23–31.

[68] S.K. Rangarajan, A unified approach to linear electrochemical systems: I. The

formalism, Journal of Electroanalytical Chemistry and Interfacial Electrochemistry. 55

(1974) 297–327.

[69] S.K. Rangarajan, A unified approach to linear electrochemical systems: II.

Phenomenological coupling, Journal of Electroanalytical Chemistry and Interfacial

Electrochemistry. 55 (1974) 329–335.

[70] S.K. Rangarajan, A unified theory of linear electrochemical systems: III. The

hierarchy of special cases, Journal of Electroanalytical Chemistry and Interfacial

Electrochemistry. 55 (1974) 337–361.

[71] S.K. Rangarajan, A unified theory of linear electrochemical systems: IV.

Electron transfer through adsorbed modes, Journal of Electroanalytical Chemistry and

Interfacial Electrochemistry. 55 (1974) 363–374.

[72] J.S. Morgan, XXX.—The periodic evolution of carbon monoxide, Journal of the

Chemical Society, Transactions. 109 (1916) 274–283.

[73] P.G. Bowers, R.M. Noyes, Chemical oscillations and instabilities. 51. Gas

evolution oscillators. 1. Some new experimental examples, Journal of the American

Chemical Society. 105 (1983) 2572–2574.

[74] K.W. Smith, R.M. Noyes, Gas evolution oscillators. 3. A computational model

of the Morgan reaction, The Journal of Physical Chemistry. 87 (1983) 1520–1524.

[75] K.W. Smith, R.M. Noyes, P.G. Bowers, Chemical oscillations and instabilities.

Part 52. Gas evolution oscillators. 2. A reexamination of formic acid dehydration, The

Journal of Physical Chemistry. 87 (1983) 1514–1519.

[76] R.M. Noyes, Gas-evolution oscillators. 4. Characteristics of the steady state

for gas evolution, The Journal of Physical Chemistry. 88 (1984) 2827–2833.

[77] G.C. Barker, Noise connected with electrode processes, Journal of

Electroanalytical Chemistry and Interfacial Electrochemistry. 21 (1969) 127–136.

[78] G.C. Barker, Flicker noise connected with the hydrogen evolution reaction on

mercury, Journal of Electroanalytical Chemistry and Interfacial Electrochemistry. 39

(1972) 484–488.

146

[79] L. Bai, D.A. Harrington, B.E. Conway, Behavior of overpotential—deposited

species in Faradaic reactions—II. ac Impedance measurements on H2 evolution

kinetics at activated and unactivated Pt cathodes, Electrochimica Acta. 32 (1987)

1713–1731.

[80] Y. Choquette, L. Brossard, A. Lasia, H. Menard, Study of the Kinetics of

Hydrogen Evolution Reaction on Raney Nickel Composite‐Coated Electrode by AC

Impedance Technique, Journal of The Electrochemical Society. 137 (1990) 1723–

1730.

[81] Y. Choquette, L. Brossard, A. Lasia, H. Ménard, Investigation of hydrogen

evolution on Raney-Nickel composite-coated electrodes, Electrochimica Acta. 35

(1990) 1251–1256.

[82] P. Ekdunge, K. Jüttner, G. Kreysa, T. Kessler, M. Ebert, W.J. Lorenz,

Electrochemical Impedance Study on the Kinetics of Hydrogen Evolution at

Amorphous Metals in Alkaline Solution, Journal of The Electrochemical Society. 138

(1991) 2660–2668.

[83] I. Herraiz-Cardona, E. Ortega, L. Vázquez-Gómez, V. Pérez-Herranz, Double-

template fabrication of three-dimensional porous nickel electrodes for hydrogen

evolution reaction, International Journal of Hydrogen Energy. 37 (2012) 2147–2156.

[84] P.M. Senik, A periodic steady-state process in autonomous, essentially

nonlinear systems, Soviet Applied Mechanics. 4 (1968) 38–42.

[85] J. Aprille, T.J., T.N. Trick, Steady-state analysis of nonlinear circuits with

periodic inputs, Proceedings of the IEEE. 60 (1972) 108–114.

[86] L.O. Chua, A. Ushida, Algorithms for computing almost periodic steady-state

response of nonlinear systems to multiple input frequencies, IEEE Transactions on

Circuits and Systems. 28 (1981) 953–971.

[87] A. Semlyen, A. Medina, Computation of the periodic steady state in systems

with nonlinear components using a hybrid time and frequency domain methodology,

IEEE Transactions on Power Systems. 10 (1995) 1498–1504.

[88] J. Feder, K.C. Russell, J. Lothe, G.M. Pound, Homogeneous nucleation and

growth of droplets in vapours, Advances in Physics. 15 (1966) 111–178.

[89] A. Volanschi, W. Olthuis, P. Bergveld, Gas bubbles electrolytically generated

at microcavity electrodes used for the measurement of the dynamic surface tension

in liquids, Sensors and Actuators A: Physical. 52 (1996) 18–22.

[90] S.F. Jones, G.M. Evans, K.P. Galvin, Bubble nucleation from gas cavities — a

review, Advances in Colloid and Interface Science. 80 (1999) 27–50.

[91] S.F. Jones, G.M. Evans, K.P. Galvin, The cycle of bubble production from a

gas cavity in a supersaturated solution, Advances in Colloid and Interface Science. 80

(1999) 51–84.

147

[92] A.R. Zeradjanin, E. Ventosa, A.S. Bondarenko, W. Schuhmann, Evaluation of

the Catalytic Performance of Gas-Evolving Electrodes using Local Electrochemical

Noise Measurements, ChemSusChem. 5 (2012) 1905–1911.

[93] H. Vogt, On the supersaturation of gas in the concentration boundary layer of

gas evolving electrodes, Electrochimica Acta. 25 (1980) 527–531.

[94] H. Vogt, The incremental ohmic resistance caused by bubbles adhering to an

electrode, Journal of Applied Electrochemistry. 13 (1983) 87–88.

[95] J. Dukovic, C.W. Tobias, The Influence of Attached Bubbles on Potential Drop

and Current Distribution at Gas‐Evolving Electrodes, Journal of The Electrochemical

Society. 134 (1987) 331–343.

[96] J.A. Leistra, P.J. Sides, Hyperpolarization of gas evolving electrodes—I.

Aqueous electrolysis, Electrochimica Acta. 32 (1987) 1489–1494.

[97] J.A. Leistra, P.J. Sides, Voltage Components at Gas Evolving Electrodes,

Journal of The Electrochemical Society. 134 (1987) 2442–2446.

[98] C. Gabrielli, F. Huet, M. Keddam, A. Macias, A. Sahar, Potential drops due to

an attached bubble on a gas-evolving electrode, Journal of Applied Electrochemistry.

19 (1989) 617–629.

[99] C. Gabrielli, F. Huet, M. Keddam, A. Sahar, Investigation of water electrolysis

by spectral analysis. I. Influence of the current density, Journal of Applied

Electrochemistry. 19 (1989) 683–696.

[100] C. Gabrielli, F. Huet, R.P. Nogueira, Fluctuations of concentration

overpotential generated at gas-evolving electrodes, Electrochimica Acta. 50 (2005)

3726–3736.

[101] J.M. Silva, R.P. Nogueira, L. de Miranda, F. Huet, Hydrogen Absorption

Estimation on Pd Electrodes from Electrochemical Noise Measurements in Single-

Compartment Cells, Journal of The Electrochemical Society. 148 (2001) E241–E247.

[102] S. Trasatti, Electrocatalysis: understanding the success of DSA®,

Electrochimica Acta. 45 (2000) 2377–2385.

[103] H.B. Beer, The Invention and Industrial Development of Metal Anodes,

Journal of The Electrochemical Society. 127 (1980) 303C–307C.

[104] H. Bouazaze, J. Fransaer, F. Huet, P. Rousseau, V. Vivier, Electrolyte-

resistance change due to an insulating sphere in contact with a disk electrode,

Electrochimica Acta. 55 (2010) 1645–1655.

[105] S. Shibata, The Concentration of Molecular Hydrogen on the Platinum

Cathode, Bulletin of the Chemical Society of Japan. 36 (1963) 53–57.

[106] H. Vogt, The Concentration Overpotential of Gas Evolving Electrodes as a

Multiple Problem of Mass Transfer, Journal of The Electrochemical Society. 137

(1990) 1179–1184.

148

[107] K. Stephan, H. Vogt, A model for correlating mass transfer data at gas

evolving electrodes, Electrochimica Acta. 24 (1979) 11–18.

[108] H. Vogt, Superposition of microconvective and macroconvective mass

transfer at gas-evolving electrodes—a theoretical attempt, Electrochimica Acta. 32

(1987) 633–636.

[109] L. Müller, M. Krenz, K. Rübner, On the relation between the transport of

electrochemically evolved Cl2 and H2 into the electrolyte bulk by convective diffusion

and by gas bubbles, Electrochimica Acta. 34 (1989) 305–308.

[110] D.P. Sutija, C.W. Tobias, Mass‐Transport Enhancement by Rising Bubble

Curtains, Journal of The Electrochemical Society. 141 (1994) 2599–2607.

[111] P. Boissonneau, P. Byrne, An experimental investigation of bubble-induced

free convection in a small electrochemical cell, Journal of Applied Electrochemistry.

30 (2000) 767–775.

[112] J. Eigeldinger, H. Vogt, The bubble coverage of gas-evolving electrodes in a

flowing electrolyte, Electrochimica Acta. 45 (2000) 4449–4456.

[113] C. Gabrielli, F. Huet, R.P. Nogueira, Electrochemical Noise Measurements of

Coalescence and Gas-Oscillator Phenomena on Gas-Evolving Electrodes, Journal of

The Electrochemical Society. 149 (2002) E71–E77.

[114] C. Gabrielli, F. Huet, R.P. Nogueira, Electrochemical impedance of H< sub>

2</sub>-evolving Pt electrode under bubble-induced and forced convections in

alkaline solutions, Electrochimica Acta. 47 (2002) 2043–2048.

[115] C. Gabrielli, F. Huet, M. Keddam, J.F. Lizee, Measurement time versus

accuracy trade-off analyzed for electrochemical impedance measurements by means

of sine, white noise and step signals, Journal of Electroanalytical Chemistry and

Interfacial Electrochemistry. 138 (1982) 201–208.

[116] B.E. Conway, J.O. Bockris, R.E. White, Modern aspects of electrochemistry,

Plenum, New York; London, 1999.

[117] P.R. Unwin, Instrumentation and electroanalytical chemistry, Wiley-VCH,

Weinheim, 2003.

[118] S. Fletcher, The two-terminal equivalent network of a three-terminal

electrochemical cell, Electrochemistry Communications. 3 (2001) 692–696.

[119] A. Battistel, M. Fan, J. Stojadinović, F. La Mantia, Analysis and mitigation of

the artefacts in electrochemical impedance spectroscopy due to three-electrode

geometry, Electrochimica Acta. 135 (2014) 133–138.

[120] M. Nebel, T. Erichsen, W. Schuhmann, Constant-distance mode SECM as a

tool to visualize local electrocatalytic activity of oxygen reduction catalysts, Beilstein

Journal of Nanotechnology. 5 (2014) 141–151.

149

[121] A.S. Bondarenko, G.A. Ragoisha, Inverse Problem In Potentiodynamic

Electrochemical Impedance, in: A.L. Pomerant s ev (Ed.), Progress in Chemometrics

Research, Nova Science Publishers, New York, 2005: pp. 89–102.

[122] A. Sadkowski, J.-P. Diard, On the Fletcher‘s two-terminal equivalent network

of a three-terminal electrochemical cell, Electrochimica Acta. 55 (2010) 1907–1911.

[123] A.L. Bard, Electrochemical methods: Fundamentals and applications, 2nd ed.,

John Wiley & Sons, New York, 2001.

[124] K. Nisancioglu, The Error in Polarization Resistance and Capacitance

Measurements Resulting from Nonuniform Ohmic Potential Drop to Flush-Mounted

Probes, Corrosion. 43 (1987) 258–265.

[125] D.R. Lide, CRC Handbook of Chemistry and Physics 2004-2005: A Ready-

Reference Book of Chemical and Physical Data, CRC press, 2004.

[126] Handbook of electrochemistry, 1st ed, Elsevier, Amsterdam ; Boston, 2007.

[127] D.A. Harrington, B.E. Conway, ac Impedance of Faradaic reactions involving

electrosorbed intermediates—I. Kinetic theory, Electrochimica Acta. 32 (1987) 1703–

1712.

[128] M.E. Lyons, M.P. Brandon, The oxygen evolution reaction on passive oxide

covered transition metal electrodes in aqueous alkaline solution. Part 1—nickel, Int J

Electrochem Sci. 3 (2008) 1386–1424.

[129] Y. Van Ingelgem, E. Tourwé, O. Blajiev, R. Pintelon, A. Hubin, Advantages of

Odd Random Phase Multisine Electrochemical Impedance Measurements,

Electroanalysis. 21 (2009) 730–739.

[130] S. Bruckenstein, M.I. Bellavance, B. Miller, The Electrochemical Response of a

Disk Electrode to Angular Velocity Steps, Journal of The Electrochemical Society. 120

(1973) 1351–1356.

[131] L.M. Peter, J. Li, R. Peat, H.J. Lewerenz, J. Stumper, Frequency response

analysis of intensity modulated photocurrents at semiconductor electrodes,

Electrochimica Acta. 35 (1990) 1657–1664.

150

LIST OF PUBLICATIONS

PATENT

―Battery for lithium extraction from seawater and brine.‖ DE102012212770.4

PUBLISHED PEER-REVIEWED ARTICLES

1. Battistel Alberto, Mu Fan, Jelena Stojadinović, and Fabio La Mantia. 2014.

―Analysis and Mitigation of the Artefacts in Electrochemical Impedance

Spectroscopy due to Three-Electrode Geometry.‖ Electrochimica Acta 135:

133–38. (Section 3.3 and Subsection 4.1.3)

2. Battistel Alberto, and Fabio La Mantia. 2013. ―Nonlinear Analysis: The

Intermodulated Differential Immittance Spectroscopy.‖ Analytical Chemistry

85 (14): 6799–6805. (Subsections 2.3.1, 2.3.2 and 2.3.4 and Sections

3.2 and 4.2)

3. Hüsken Nina, Magdalena Gębala, Alberto Battistel, Fabio La Mantia,

Wolfgang Schuhmann, and Nils Metzler-Nolte. 2012. ―Impact of Single

Basepair Mismatches on Electron-Transfer Processes at Fc-PNA⋅DNA Modified

Gold Surfaces.‖ ChemPhysChem 13 (1): 131–39.

4. Pasta Mauro, Alberto Battistel, and Fabio La Mantia. 2012. ―Lead–lead

Fluoride Reference Electrode.‖ Electrochemistry Communications 20 (0): 145–

48.

5. Pasta Mauro, Alberto Battistel, and Fabio La Mantia. 2012. ―Batteries for

Lithium Recovery from Brines.‖ Energy & Environmental Science 5 (11):

9487–91.

ACCEPTED WORK

1. Rafael Trócoli, Alberto Battistel, Fabio La Mantia. ―Selectivity of lithium

recovery process based on LiFePO4.― Accepted in Chemistry - A European

Journal (2014). DOI: 10.1002/chem.201403535

WORKS IN PREPARATION

1. Alberto Battistel, Fabio La Mantia. ―Intermodulated differential immittance

spectroscopy of a redox couple in solution: the model.‖ (Subsections 2.3.5

and 2.3.6)

151

2. Alberto Battistel, Fabio La Mantia. ―Intermodulated differential immittance

spectroscopy of a redox couple in solution: the experimental results.‖

(Subsection 2.3.7 and Section 4.3)

3. Rosalba A. Rincón, Alberto Battistel, Edgar Ventosa, Xingxing Chen,

Michaela Nebel, Wolfgang Schuhmann. ―Using cavity microelectrodes for

direct electrochemical noise studies of oxygen evolving catalysts.‖

4. Alberto Battistel, Rosalba A. Rincón, Fabio La Mantia.―Effect of the dynamics

of gas bubble formation on the macrokinetics.‖ (Sections 2.4 and 4.4)

5. Jelena Stojadinović, Mu Fan, Battistel Alberto, and Fabio La Mantia.

―Analysis and mitigation of the artefacts in electrochemical impedance

spectroscopy in the four-electrode configuration.‖

6. Xingxing Chen, Artjom Maljusch, Rosalba A. Rincón, Alberto Battistel,

Aliaksandr S. Bandarenka, Wolfgang Schuhmanna. ―Characterization of gas

evolving electrodes using scanning electrochemical microscopy.‖

TALKS AT INTERNATIONAL CONFERENCES

1. 09/2013 International Workshop on Impedance Spectroscopy (IWIS 2013),

Chemnitz, Germany.

Title: ―Non-linear analysis: studying the kinetics of Fe (II) / Fe (III) cyanide

complex through the Intermodulated Differential Immittance Spectroscopy.‖

2. 06/2013 9th International Symposium on Electrochemical Impedance

Spectroscopy (EIS 2013), Okinawa, Japan.

Title: ―Intermodulation Technique: a study on the mechanism of hydrogen

evolution reaction.‖

3. 09/2012 Electrochemistry 2012 München, Munich, Germany.

Title: ―Differential Photo-Impedance Spectroscopy―

POSTERS AT INTERNATIONAL CONFERENCES

1. Alberto Battistel, Fabio La Mantia, International Workshop on Impedance

Spectroscopy (IWIS 2013), Chemnitz, Germany. Germany (2013).

Title: ―Intermodulated differential immittance spectroscopy: physical

fundamentals and instrumental setup.‖

2. Mauro Pasta, Alberto Battistel, Fabio La Mantia, 63rd Annual Meeting of the

International Society of Electrochemistry, Prague, Czech Republic (2012).

Title: ―Lead-lead Fluoride Reference Electrode.‖