Download Electric Fields for Surface Design and Chemical Analysis Christian Ulrich

Document related concepts

Nominal impedance wikipedia , lookup

Bio-MEMS wikipedia , lookup

Electroactive polymers wikipedia , lookup

Transcript
Linköping Studies in Science and Technology
Dissertation No. 1206
Electric Fields for Surface Design and
Chemical Analysis
Christian Ulrich
Department of Physics, Chemistry and Biology
Linköpings universitet, SE-581 83 Linköping, Sweden
Linköping 2008
The cover picture shows SPR response as a function of electrode potential, thickness variations of different surface gradients, and the variations in potential and
current density in a solution containing a bipolar electrode.
During the course of the research underlying this thesis, Christian Ulrich was
enrolled in Forum Scientium, a multidisciplinary doctoral programme at Linköping
University, Sweden.
c Copyright 2008 Christian Ulrich, unless otherwise noted.
All rights reserved.
Christian Ulrich
Electric Fields for Surface Design and Chemical Analysis
ISBN 978–91–7393–819–8
ISSN 0345–7524
Published online at www.ep.liu.se
Linköping Studies in Science and Technology
Dissertation No. 1206
Printed in Sweden by LiU-Tryck, Linköping 2008
Work. Finish. Publish.†
– Michael Faraday
(1791-1867)
†
Advice to the young William Crookes.
Abstract
This thesis deals with the use of electric fields for evaluation and control of chemical systems. An electric field can result in the flow of charge across an interface
between a metal and a solution, by means of chemical reactions. This interplay
between electricity and chemistry, i.e. electrochemistry, is a field of crucial importance both within research and industry. Applications based on electrochemical
principles encompass such diverse areas as batteries and fuel cells, pH electrodes,
and the glucose monitor used by people suffering from diabetes.
A major part of the present work concerns the use of static electric fields in
solutions containing a non-contacted metal surface. In such a setup it is possible
to control the extent of electrochemical reactions at different positions on the
metal. This allows the formation and evaluation of various types of gradients on
electrodes, via indirectly induced electrochemical reactions. This approach is a
new and simple way of forming for instance molecular gradients on conducting
surfaces. These are very advantageous in biomimetic research, because a gradient
contains a huge amount of discrete combinations of for example two molecules. The
basis for the technique is the use of bipolar electrochemistry. Briefly, a surface can
become a bipolar electrode (an electrode that acts as both anode and cathode)
when the electric field in the solution exceeds a certain threshold value, thereby
inducing redox reactions at both ends. In our experiments, the driving force
for these reactions will vary along the electrode surface. Since the result of an
electrochemical reaction can be the deposition or removal of material from an
electrode, bipolar electrochemistry can be used to create gradients of that material
on a surface. In order to gain a deeper understanding of these processes, the
potential and current density distributions at bipolar electrodes were investigated
with different methods. Especially the use of imaging techniques was important
for the visualization and analysis of the gradients. Using this knowledge, the
formation of more complex gradients was facilitated, and the results were further
compared to simulations based on simple conductivity models. These simulations
also provided us with means to predict the behavior of new and interesting setup
geometries for pattering applications.
The other major part is more application driven and deals with the use of alternating electric fields for chemical analysis, a technique known as electrochemical
impedance spectroscopy (EIS). In this work, EIS has been applied for the analyv
sis of engine oils and industrial cutting fluids. Emphasis was placed on practical
aspects of the measurement procedure, and on the evaluation of the results using
statistical methods. It was for example shown that it was possible to simultaneously determine the amount of different contaminants in low conducting solutions.
Generally, EIS is used to measure the impedance of a solution or a solid, often
as a function of the frequency of the alternating electric field. The impedance
of a system is closely correlated to its complex dielectric constant, and EIS can
therefor be used to examine many chemical and physical processes. It is further
well suited for characterizing low conducting media with little or no redox-active
species. The evaluation of impedance data is often a quite complex task, which
is why we have made use of statistical methods that drastically reduce the effort
and quickly reveal significant intrinsic parameters.
vi
Populärvetenskaplig
sammanfattning
Denna avhandling handlar i grunden om hur elektriska fält kan användas för att
utvärdera och kontrollera olika kemiska system. Området inom kemin som rör sambandet mellan elektricitet och kemiska föreningar kallas elektrokemi, och används
flitigt inom både industriella sammanhang och inom forskning. Det finns många
exempel på kommersiella produkter som är baserade på elektrokemiska principer,
till exempel batterier, pH-elektroder och glukosmätare som används av diabetiker.
En stor del av det här arbetet rör användandet av statiska elektriska fält i
en lösning innehållande en okontakterad metallyta. Med en sådan uppställning
är det möjligt att kontrollera omfattningen av elektrokemiska reaktioner på olika
positioner på ytan. På det här sättet är det möjligt att skapa och utvärdera olika
kemiska gradienter. Tillvägagångssättet bygger på bipolär elektrokemi, och har
här för första gången utnyttjats till att skapa en proteingradient på en bipolär
guldyta. En sådan ytgradient kommer att uppvisa en gradvis variation i en eller
flera egenskaper. En av de stora fördelarna är alltså att en gradient innehåller
en stor mängd kombinationer av till exempel två kemiska egenskaper. Metoden är
snabb och enkel, och gör det möjligt att skapa avancerade gradienter av molekyler,
som sedan skulle kunna användas inom forskning och utveckling av exempelvis
biosensorer.
En bipolär elektrod, det vill säga en elektrod som agerar som både anod och katod, kan skapas genom att placera en elektriskt ledande yta i ett parallellt elektriskt
fält. Detta kommer att ske under förutsättning att det elektriska fältet överstiger
ett visst tröskelvärde och därmed inducerar redoxreaktioner vid ytans båda ändar.
Drivkraften för dessa reaktioner kommer att variera utmed ytan, parallellt med
det elektriska fältet. Eftersom resultatet av en elektrokemisk reaktion kan vara
både borttagande och deponering av material på en elektrod, kan bipolär elektrokemi användas till att skapa gradienter av detta material. För att få en djupare
förståelse för fenomenet har mätningar gjorts för att utvärdera hur potentialen
och strömmen fördelar sig kring en bipolär elektrod. Särskilt kombinationen av
elektrokemi och avbildande optiska tekniker, huvudsakligen ellipsometri och ytplasmonresonans (SPR), har också varit en viktig del av arbetet. Detta eftersom
olika processer som sker vid en elektrod då kan avbildas och studeras i realtid.
vii
Resultaten har vidare jämförts med simuleringar baserade på grundläggande konduktivitetsmodeller.
Den andra stora delen av arbetet rör mer direkta analytiska applikationer
av elektrokemi, och speciellt användandet av varierande elektriska fält. Denna
mätteknik kallas för elektrokemisk impedansspektroskopi (EIS), och har här utnyttjats för analys av motoroljor och industriella skärvätskor. Tonvikt har lagts
vid att optimera mättekniken och att utvärdera resultaten med hjälp av statistiska
verktyg. Resultaten visar till exempel på möjligheten att samtidigt bestämma halter av olika föroreningar i lågledande lösningar. Generellt används EIS för att mäta
impedansen för både fasta och flytande ämnen, ofta som funktion av frekvensen
för det applicerade elektriska fältet. Denna impedans är tätt förknippad med systemets komplexa dielektricitetskonstant, och EIS kan användas till att undersöka
många kemiska och fysikaliska processer. Därmed är tekniken till exempel mycket
lämpad för att karaktärisera lågledande medier. Utvärderingen av impedansdata
kan dock vara mycket komplex, något som vi visat kan underlättas genom att
använda statistiska verktyg.
viii
List of Publications
This thesis is based on six papers (included at the end). In the text, they will be
referred to by their roman numerals (Papers I–VI).
Paper I
Formation of Molecular Gradients on Bipolar Electrodes
Christian Ulrich, Olof Andersson, Leif Nyholm, and Fredrik Björefors
Angewandte Chemie International Edition, 47:3034–3036, 2008.
Author’s contribution
Was responsible for the planning, and performed all experimental work together with Olof Andersson. Wrote the manuscript.
Paper II
Potential and Current Density Distributions at Electrodes Intended for Bipolar Patterning
Christian Ulrich, Olof Andersson, Leif Nyholm, and Fredrik Björefors
Submitted to Analytical Chemistry
Author’s contribution
Was responsible for the planning, and performed all experimental work and
simulations. Olof Andersson was responsible for the iSPR experiments.
Wrote the manuscript.
Paper III
Imaging SPR for Detection of Local Electrochemical Processes
on Patterned Surfaces
Olof Andersson, Christian Ulrich, Fredrik Björefors, and Bo Liedberg
Sensors and Actuators B, doi:10.1016/j.snb.2008.05.042, 2008.
Author’s contribution
Contributed during the planning, evaluation, and writing.
ix
Paper IV
Current Oscillations During Chronoamperometric and Cyclic
Voltammetric Measurements in Alkaline Cu(II)-citrate Solutions
Jonas Eskhult, Christian Ulrich, Fredrik Björefors, and Leif Nyholm
Electrochimica Acta, 53:2188–2197, 2008.
Author’s contribution
Was responsible for the ellipsometric experiments, and performed the
optical simulations. Contributed to the writing.
Paper V
Simultaneous Estimation of Soot and Diesel Contamination in
Engine Oil Using Electrochemical Impedance Spectroscopy
Christian Ulrich, Henrik Petersson, Hans Sundgren, Fredrik Björefors, and
Christina Krantz-Rülcker
Sensors and Actuators B, 127:613–618, 2007.
Author’s contribution
Responsible for the planning, performance, and evaluation of the experimental work, including the cell design. Wrote the manuscript.
Paper VI
Quality Evaluation of Industrial Cutting Fluids Using Electrochemical Impedance Spectroscopy
Christian Ulrich, Dan Louthander, Per Mårtensson, André Kluftinger,
Michael Gawronski, and Fredrik Björefors
In manuscript
Author’s contribution
Responsible for the planning, performance, and evaluation of the experimental work, including the cell design. Wrote the manuscript.
x
Acknowledgements
It is with a sense of sadness and sentimentality I write this acknowledgement,
because in thanking all the people who have helped me through the years, I realize
that my time at IFM is nearing its end. It has been an inspirational and rewarding
time, but now new things await.
I am forever grateful to my supervisor Fredrik Björefors. You have given me
invaluable help through the years, and further managed to provide both excellent supervision and support. My co-supervisor Ingemar Lundström is, by his
mere presence, a source of inspiration. Special thanks go to my former supervisor
Tina Krantz-Rülcker, who gave me the opportunity to explore the mysteries of
experimental research. My former co-supervisor Fredrik Winquist was also the supervisor on my diploma work, which started this journey. Thanks to Bo Liedberg
for welcoming me into his group. Hans Sundgren has been essential to many of
my undertakings. I could always knock on his door to discuss things, big as well
as small.
Bo Thunér and Ingemar Grahn are acknowledged for valuable assistance with
all things practical; Stefan Klintström for guidance whenever needed and for managing Furum Scientium; Agneta Askendal for the laboratory assistance; the administrative staff at IFM for all the help, and especially Susann Årnfelt, Kerstin
vestin, Pia Blomstedt, and Anna Maria Uhlin.
All the people I have collaborated with, who have contributed to this thesis.
In no particular order I would like to thank the following. Olof Andersson for the
successful team work, his positive attitude, and nice company; Henrik Petersson
for the data evaluations and general support; Jonas Eskhult and especially Leif
Nyholm for the successful cooperation and rewarding discussions; representatives
of the industrial partners through the years, Jaco Visser, Claes Frennfelt, Per
Mårtensson, Michael Gawronski, and André Kluftinger; Dan Louthander, for the
collaboration in the end, but foremost for all the Friday morning breakfasts at
Lanemos.
Without naming anyone in particular, I have very much valued the company
of my former room-mates, the coffee-group, the lunch-group, all the colleagues
within Forum Scientium, and all former and present members of S-SENCE and
the Sensor Science and Molecular Physics group.
xi
Last but not least, I wish to thank my friends and family for their never ending
love and support, which has made it possible for me to cope with all the hardships
of life, especially towards the end.
Christian Ulrich, Linköping, 2008.
xii
Contents
1 General introduction
1
2 Electrochemistry
2.1 Introduction to electrochemical reactions .
2.2 The electrical double layer . . . . . . . . .
2.3 Electrochemical cells and cell resistance .
2.4 Mass transfer and the diffusion layer . . .
2.5 Voltammetry . . . . . . . . . . . . . . . .
2.6 Electrochemical impedance spectroscopy .
2.6.1 Theory . . . . . . . . . . . . . . .
2.6.2 General applications . . . . . . . .
2.7 Bipolar electrochemistry . . . . . . . . . .
2.8 Limitations of electrochemistry . . . . . .
.
.
.
.
.
.
.
.
.
.
3
3
8
10
11
12
14
14
18
18
20
.
.
.
.
.
.
.
21
21
21
21
22
23
23
24
4 Electrode surface design and analysis
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2 Materials and methods . . . . . . . . . . . . . . . . . . . . . . . . .
4.3 Surface gradients . . . . . . . . . . . . . . . . . . . . . . . . . . . .
27
27
27
29
5 Alternating electric fields for
5.1 Introduction . . . . . . . . .
5.2 Practical considerations . .
5.3 Applications . . . . . . . . .
35
35
36
37
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
3 Imaging optical methods and electrochemistry
3.1 Introduction . . . . . . . . . . . . . . . . . . . . .
3.2 Surface plasmon resonance . . . . . . . . . . . . .
3.2.1 Theory . . . . . . . . . . . . . . . . . . .
3.2.2 SPR and electrochemistry . . . . . . . . .
3.3 Ellipsometry . . . . . . . . . . . . . . . . . . . .
3.3.1 Theory . . . . . . . . . . . . . . . . . . .
3.3.2 Ellipsometry and electrochemistry . . . .
chemical
. . . . . .
. . . . . .
. . . . . .
xiii
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
analysis
. . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . .
6 Future outlook
41
Bibliography
43
Paper I
51
Paper II
59
Paper III
71
Paper IV
79
Paper V
91
Paper VI
99
xiv
1
General introduction
In 1834, Michael Faraday published his two laws of electrolysis, relating the amount
of charge passed in a circuit to the extent of electrochemical reactions taking place
at the interface between an electrode and a solution. This fundamental discovery
might seem somewhat self-evident today, but one has to remember that Faraday’s
work was done in a time when electricity was poorly understood, and the nature
of the atom was still unknown. His contributions to this field credited him as
one of the founders of modern electrochemistry. Electrolysis was, however, only
a small part of his collected work. Faraday also introduced the idea of electric
fields into 19th century science. Furthermore, he defined the dielectric constant
as a result of his studies on capacitances. He is regarded as one of history’s
greatest experimentalists and, needless to say, his impact on modern science is
unmistakable.
A more recent pioneer was Jaroslav Heyrovsky, who has been called the father of electroanalytical chemistry. He received the Nobel Prize in chemistry in
1959 “for his discovery and development of the polarographic methods of analysis”. Polarography is an electroanalytical technique where a dropping mercury
electrode is used as the working electrode in voltammetric experiments. Basically,
a current response is recorded when a potential difference (and hence a resulting
electric field) is applied across the mercury/solution interface. This, for example,
facilitates the control and evaluation of the amount of electrochemical reactions at
1
2
General introduction
an electrode. Similar techniques can also be used to establish a significant electric
field in the entire solution between two electrodes. These types of electric fields
have been extensively used in the work described in this thesis.
Electric fields can be utilized to affect, control, and analyze different physical
and chemical parameters in a material or solution. This is the basis of electrochemistry, which deals with the interaction between electricity and chemical compounds.
What separates electrochemistry from many other techniques or methods is the
fact that it allows both the manipulation and probing of a sample at the same
time. Voltammetry is perhaps the most well known example, where information
about an analyte is obtained as it is oxidized or reduced at an electrode/solution
interface under potential control. Important parameters such as reaction kinetics,
adsorption processes, as well as mass transfer characteristics can be studied.
Many of the electrochemical techniques employed today are based on the use
of static electric fields. It is also possible to make use of alternating electric fields,
which for example allows the possibility to evaluate the impedance of the system
of interest. This technique is known as electrochemical impedance spectroscopy
(EIS), and is widely used in science and industrial applications. Impedance data often contains information on many different parameters of the investigated system,
and EIS has therefore been applied in many diverse areas. Some examples include
corrosion studies on pipelines and reinforcements in concrete, body impedance
measurements, and fundamental research related to sensor science and batteries.
The evaluation of the data can, however, be quite a complex task. In this work,
statistical methods were therefore used to more easily extract relevant information.
2
Electrochemistry
2.1
Introduction to electrochemical reactions
A very simple electrochemical setup can consist of a battery, two metal plates, and
a solution (i.e. an electrochemical cell). The battery is of course also an electrochemical cell, but is here treated as an arbitrary voltage source. The metals are
immersed in the solution and connected to the battery as in Figure 2.1, resulting
in a current in the circuit if the voltage is sufficiently high. The metals (and wires)
are in this case electronic conductors and charge is here transported by electrons.
In the solution, charge is carried by the movement of charged species. An imperative demand for the passage of a current in this circuit is that charge can cross
from the metal to the solution and vice versa. This is in most cases enabled by
the uptake or loss of electrons by species in the solution. These electrochemical
reactions are one of the most fundamental connections between electricity and
chemistry. It is also possible to make use of spontaneous electrochemical reactions
to convert chemical energy into electrical energy.
In electrochemistry, as in most other fields, a well defined terminology is essential. In his publication from 1834,1 Faraday (after discussions with colleagues)
defined many of the electrochemical terms we still use today. First of all, the
electronic conductors in contact with the solution were called electrodes. Typical
electrode materials today include solid metals (e.g. Pt and Au), carbon (graphite),
3
4
Electrochemistry
-
+
i
Figure 2.1: A simple electrochemical setup.
and semiconductors (e.g. indium-tin oxide and Si). The solution was called an electrolyte, and the most common electrolytes contain charged species such as H+ ,
Na+ , and Cl – , in order to increase the conductivity, in either water or organic solvents. Faraday defined the substances that can pass a current in a solution as ions,
and divided them into negative anions and positive cations. Further, he termed
the electrode at which oxidation occurs the anode and that where reduction occurs
the cathode. Oxidation (Red −−→ Ox + ne – ) is the removal of electrons from a
species and reduction (Ox + ne – −−→ Red) is the addition of electrons. A redox
reaction is consequently a reaction in which there is a transfer of electrons from
one species to another. The main concern in electrochemical systems is processes
and factors that affect the transport of charge across the interface between two
phases, e.g. between an electrode and an electrolyte.2 To chemists, Faraday is
perhaps best known for his work on electrolysis, which involves passing a current
through a solution and thereby inducing reactions at the metal/solution interfaces.
To proceed with a real example, Figure 2.2 shows an electrochemical setup
where two electrodes are immersed in an electrolyte. This electrolyte contains
1 M of Cl – and equal concentrations of ferri and ferrocyanide ions. One electrode is a piece of platinum and the other is a silver wire coated with solid silver
chloride. The composition of this cell allows two reactions to occur; Fe(CN)63 – +
–
4–
−−
*
−−
*
e– )
−
− Fe(CN)6 at the platinum electrode and AgCl + e – )
−
− Ag + Cl at the
silver electrode. These are called the half-reactions of the cell, and together they
will constitute the overall chemical reaction. If the power supply is disconnected,
the high impedance voltmeter will show the open-circuit of the cell. This potential, measured in volts (V), is related to the energy available in the cell to force
an external current in the circuit. By using an amperemeter, this current can be
recorded.
Associated with every electrochemical half-reaction is a potential, called the
2.1 Introduction to electrochemical reactions
5
Power
Power
supply
supply
Ai
V
V
Ag
Pt
AgCl
1M ClFe(CN)63-/4-
Figure 2.2: An electrochemical cell with two electrodes immersed in an electrolyte. A
power supply together with the voltmeter and amperemeter provides the possibility to
obtain current-potential curves.
standard reduction potential (denoted E 0 ). For the two reactions
4−
−−
*
Fe(CN)63− + e− )
−
− Fe(CN)6
and
−
−−
*
AgCl + e− )
−
− Ag + Cl
E 0 is 0.361 V and 0.222 V, respectively. These values are valid under certain
conditions (i.e. standardized temperature and pressure, and unit activity). In
practice, the values of the electrode potentials will depend on the temperature
and the activities of the species present in the solution. It is important to note
that the standard reduction potentials are defined with respect to something called
the standard hydrogen electrode, which will be further explained below. For nonstandard conditions, electrode potentials can be calculated using the Nernst equation, which includes two terms. The first is the standard reduction potential and
the second term involves the activities of the species in the reaction.
For the half-reaction
4−
−−
*
Fe(CN)63− + e− )
−
− Fe(CN)6
the Nernst equation giving the half-cell potential, E, is
E = E0 −
AF e(CN )4−
RT
ARed
RT
6
ln
= E0 −
ln
nF
AOx
F
AF e(CN )3−
(2.1)
6
where R = molar gas constant (8.31447 JK−1 mol−1 ), T = temperature (K), n =
number of electrons in the half-reaction, F = Faraday constant (96485.3 C), and
Ai = activity of species i. The activity can often be approximated by the concentration of the species. Solids and solvents are normally omitted from equation
because their activities are unity (or close to unity). When all activities are unity,
the logarithmic term in the equation is zero, thus giving E = E 0 .
6
Electrochemistry
For a complete cell, the voltage (Ecell ) is the difference between the potentials
of the cathode and anode:
Ecell = Ecathode − Eanode
(2.2)
If the cell voltage is positive, the net cell reaction is spontaneous in the forward
direction. If the cell voltage is negative, the reaction is spontaneous in the reverse
direction.
Going back to the example in Figure 2.2, AgCl and Ag are both solids, and
hence their activities are unity. The activity of Cl – can be found from the concentration in solution. The potential of the Ag/AgCl electrode will in this case be
0.235 V due to the concentration of chloride ions (1 M). The potential of the Pt
electrode will be E ≈ E 0 = 0.361 V , since the activities of the ferri/ferrocyanide
ions are approximately equal.
Often in electrochemical experiments, only one of the two half-reactions is
of interest. The electrode at which this reaction occurs is called the working
electrode.2–4 In order to study only one reaction, the other half of the cell is
standardized using a reference electrode, which is made up of phases having essentially constant chemical composition regardless of the experimental conditions.
The internationally accepted primary reference is the standard hydrogen electrode
(SHE), which has all components at unit activity:
Pt/H2 (p = 1 bar)/H+ (a = 1, aq)
where a slash represents a phase boundary. The potential of the SHE is defined
as being equal to zero, i.e. the standard reduction potential (E 0 ) for the reduction
of two protons to form hydrogen is 0.00 V. All electrochemical half reactions can
now be assigned a standard reduction potential with respect to the SHE. Potentials
are often measured and quoted with respect to other reference electrodes than the
SHE, since the latter is not very convenient from an experimental point of view.
A commonly used reference is the Ag/AgCl electrode shown in Figure 2.2, called
the silver-silver chloride electrode,
Ag/AgCl/KCl (saturated in water)
which for a saturated chloride solution has a potential of 0.197 V vs. SHE.
In an electrochemical experiment, the potential of the working electrode is said
to be observed or controlled with respect to the reference. This is equivalent to
observing or controlling the energy of the electrons within the working electrode.
The energy can be raised by driving the electrode to more negative potentials.
When the electrons reach a level sufficient to transfer into vacant electronic states
on species in the electrolyte, a flow of electrons from electrode to solution (a
reduction current) occurs, see Figure 2.3. Similarly, the energy of the electrons
can be lowered by imposing a more positive potential on the electrode. At some
2.1 Introduction to electrochemical reactions
Energy level
of electrons
Electrode
7
Electrode
Solution
Solution
e-
Vacant
MO
-
Reduction
Potential
Occupied
MO
+
Energy level
of electrons
-
Electrode
Electrode
Solution
Solution
Vacant
MO
Oxidation
Potential
+
A + e- → A-
Occupied
MO
A → A+ + e-
e-
Figure 2.3: Reduction (top) and oxidation (bottom) of species A in solution. Vacant
MO is the lowest vacant molecular orbital of species A and occupied MO is the highest
occupied.
point electrons on species in the electrolyte will find a more favorable energy level
on the electrode, resulting in a flow of electrons from solution to electrode, i.e.
an oxidation current. The potentials at which these processes occur are related
to the standard reduction potentials, E 0 , for each electrochemical reaction in the
system.
Continuing with the example in Figure 2.2, we are now ready to change the
potential of the Pt electrode with respect to the Ag/AgCl reference electrode.
If the potential is made sufficiently positive, electrons cross from the solution
phase into the electrode, resulting in the oxidation of Fe(CN)64 – to Fe(CN)63 – .
While this reaction occurs at the Pt electrode, AgCl is reduced to Ag, and Cl – is
liberated into the solution. The potential here will be nearly constant because the
composition of the Ag/AgCl/Cl – interface is almost unchanged with the passage
of modest currents, due to the high Cl – concentration. Thus, it works as an
excellent reference electrode in this setup. Indeed, the essential characteristic of
8
Electrochemistry
a reference electrode is that its potential remains practically constant with the
passage of small currents. Hence, when a potential is applied between the two
electrodes, nearly all of the potential change occurs at the Pt/solution interface.
The number of electrons that cross an interface is related stoichiometrically to
the extent of the chemical reaction, and is measured in terms of the total charge Q.
Charge is expressed in units of coulombs (C), where 1 C is equivalent to 6.24 · 1018
elementary charges. The relationship between charge and amount of reactions is
given by Faraday’s law. This law states that the passage of 96485.3 C causes 1
equivalent of reaction (e.g. consumption of 1 mol of reactant or generation of 1
mol of product in a one-electron reaction). The current, i, is the rate of flow of
charges (or electrons), where 1 ampere (A) is equivalent to 1 C/s. If the current
is plotted as a function of the potential, a current-potential (i vs. E) curve is
obtained. Such curves are very informative about the nature of both the solution
and the electrodes, and about the reactions that occur at the interfaces.
2.2
The electrical double layer
Charge transfer reactions take place in a region very close to the electrode surface.
As described above, the driving force for these reactions is the potential difference
between the electrode and electrolyte. However, other processes are also important
in this region due to the polarization of the electrodes. If the solution in a setup like
the one in Figure 2.1 only contains an inert supporting electrolyte (e.g. KNO3 ),
no charge transfer reaction will take place if the applied potential difference is
moderate. The left electrode is in this case made negative by the battery and
will hence have an excess of electrons at the surface. This electrode will thus
attract positive ions and dipoles present in the solution (see Figure 2.4). If the
potential (and thereby the charge) of the electrode is changed, the composition
of the solution side will also change to maintain electroneutrality. In this respect,
the electrode/solution interface behaves like a capacitor, which is governed by the
equation
q
=C
(2.3)
E
where q is the charge stored on the capacitor, E is the potential across the capacitor
and C is the capacitance (in farads, F). When the electrode potential is changed,
charge will accumulate until q satisfies Equation 2.3. The resulting movement
of charged species in the solution will give rise to a current, called the charging
current. A transient current can thus flow in the circuit without any charge transfer
reactions at the interface. This is a very important effect that has to be considered
in electrochemical experiments whenever the potential of an electrode is changed.
The whole array of charged species at the metal/solution interface is called
the electrical double layer.2 The interface can be characterized by a double-layer
capacitance, Cdl , which typically is in the range of 10–40 µF/cm2 . Cdl is, however,
2.2 The electrical double layer
9
Electrode
Solution
Figure 2.4: Illustration of a negatively charged electrode and the corresponding abundance of positive charges in the solution.
unlike real capacitors often a function of potential. Normally the existence of
the double-layer capacitance and the presence of a charging current cannot be
neglected in electrochemical experiments. The charging current can in reality be
much larger than the charge transfer current if the concentrations of electroactive
species are very low, or if the potential change is very rapid.
The solution side of the double layer is itself made up of at least two layers. The
one closest to the electrode is called the inner layer and contains solvent molecules
and sometimes other species (ions or molecules) that are said to be specifically
adsorbed (Figure 2.5). This inner layer is also called the compact or Helmholtz
layer. The locus of the electrical centers of the specifically adsorbed ions defines
the inner Helmholtz plane (IHP), which is at a distance x1 from the electrode
surface. The inner layer extends out from the electrode to the outer Helmholtz
plane (OHP), which is defined as passing through the locus of centers of the nearest
solvated ions, at a distance x2 from the surface. The solvated ions interact with
the charged metal only through long-range electrostatic forces. They are said
to be non-specifically adsorbed and are distributed in a region called the diffuse
layer, which extends from the OHP into the bulk of the solution. The thickness
of the diffuse layer is determined by the potential difference and the total ionic
concentration, and for concentrations greater than 10−2 M, the thickness is less
than ∼100 Å.
The double layer capacitance is a very important parameter when evaluating
and modeling results from electrochemical experiments. It can give important
information on the structure of the double layer, and also on the composition of
the electrolyte. If an electrode surface is modified by a layer of some material, the
thickness and dielectric properties of that material will greatly affect Cdl .
10
Electrochemistry
IHP OHP
Electrode
-
Diffuse layer
+
+
+
Solvated
cation
+
+
+
= Solvent molecule
= Specifically adsorbed
ion or molecule
x1 x2
Figure 2.5: Schematic representation of the double layer for a negatively charged electrode. IHP is the inner Helmholtz plane and OHP is the outer.
2.3
Electrochemical cells and cell resistance
In the electrolyte, charge is transported by the movement of ions (anions in one
direction and cations in the other). The total resistance between the electrodes
will thus depend on the charge, mobility, and concentration of the species present
in the electrolyte. This solution resistance, Rs , will behave as a true resistance
over a wide range of conditions. If a current, i, is passed through a solution,
a potential drop equal to iRs will arise. When controlling the potential of an
electrode (with respect to a reference electrode), this potential drop often has to
be taken into account. The actual potential of the electrode will be less than the
applied potential by and amount equal to iRs . This is shown by Equation 2.4
Eappl (vs. SHE) = E(vs. SHE) + iRs
(2.4)
where Eappl is the voltage applied by the external power supply and E is the
potential of the working electrode. If both the current and the solution resistance
are reasonably small, the potential drop can often be disregarded. However, in
situations where this is not true, the problem can be minimized by an alteration
of the electrochemical setup.
When a working electrode and a reference electrode are used to perform electrochemical measurements, the setup is called a two-electrode cell (see Figure 2.6A).
The potential of the working electrode will in this case be given by Equation 2.4.
Many electrochemical instruments offer the possibility to use a third electrode,
called a counter electrode, resulting in a three-electrode cell. As can be seen in
Figure 2.6B, the current will now be passed between the working and the counter
2.4 Mass transfer and the diffusion layer
(A)
11
(B)
Power
supply
Ai
WE
Power
supply
RE
Eappl
V
V
WE
V
V
Ai
CE
RE
EWE vs. RE
Figure 2.6: A two-electrode cell (A) and a three-electrode cell (B). WE is the working
electrode, RE the reference electrode, and CE the counter electrode.
electrodes. The behavior of the working electrode is not affected by the electrochemical properties of the counter electrode and it often consists of an inert
conducting material. A high input impedance device is used to measure the potential difference between the reference and working electrodes, and the current
passed in this circuit will hence be negligible. The result is thus a very low potential drop in the solution, and the measured voltage will be representative of the
true potential of the working electrode.
The existence of a potential drop can however be useful in many situations.
Large potential drops can for instance be used to perform electrophoresis, and gel
separations of proteins and DNA. We have utilized the potential drop in a solution
containing an isolated conducting surface, to form reaction gradients. This will be
further discussed in section 4.3.
2.4
Mass transfer and the diffusion layer
As previously discussed, the electrolyte is far from static during electrochemical
experiments. The majority of processes occurring here are based on the movement
of different species. This movement of material from one location in solution to
another is called mass transfer and can arise from the following:2
1. Migration – movement of charged species in an electric field.
2. Diffusion – movement of species as a result of a concentration gradient.
3. Convection – movement of species caused by stirring or by a density (or
heat) gradient.
In electrochemical experiments, it is often desirable to keep one or more of these
contributions to mass transfer to a minimum. For example, by the addition of
a supporting electrolyte the migration of the redox species can be made negligi-
12
Electrochemistry
COx
COx*
1
2
δN
x
Figure 2.7: Concentration profiles at two different electrode potentials: (1) where
∗
COx (x = 0) is about half of COx
and (2) where COx (x = 0) ≈ 0.
ble. The diffusion, however, is more complicated to control, and it often has a
considerable impact on electrochemical experiments.
In an experiment where species Ox is reduced at a cathode, the concentration
at the electrode surface, COx (x = 0), will become smaller than the bulk concen∗
. This will give rise to a concentration profile that extends into the
tration, COx
solution near the electrode by an amount δN , the Nernst diffusion layer thickness.3
The magnitude of δN is defined by the intersection of the tangent to the concen∗
tration profile with the horizontal line COx
(Figure 2.7). δN increases with time
until it reaches a constant value, termed the stationary diffusion layer thickness.
If the solution is unstirred, δN of up to ∼0.5 mm is possible, and reaching steady
state can take minutes. In a stirred solution, δN is well defined and much smaller
(in the order of 1 µm), and steady state will usually be reached within seconds.
If the potential is sufficiently high, COx (x = 0) will decrease to an extremely
small value and, provided that the chemical reactions are rapid compared to the
mass transfer processes, the current will be limited by the rate at which new species
can reach the electrode. Increasing the potential further will hence not affect the
current, since the diffusion is a function of the concentration gradient, only.
2.5
Voltammetry
There are many ways to use electrochemistry as an analytical tool. One widely
used method is to let the potential of a working electrode vary linearly with time
and record the current. A result from such a linear potential sweep can be seen in
Figure 2.8. In this example, the potential of a working electrode was swept between E1 and E2 in an electrolyte containing only the reduced form of an analyte.
2.5 Voltammetry
13
Current
0
E1
E0
E2
Potential
Figure 2.8: Current potential curve resulting from a linear potential scan.
In the beginning of the sweep (near E1 ) no reaction can occur, but as the potential gets closer to E 0 an oxidation starts, resulting in the current increase. The
surface concentration of the reduced analyte now begins to decrease, leading to a
flux of species to the surface. When the potential has moved past E 0 , the surface
concentration will be almost zero, and the flux of redox species will reach a maximum value. This rate of mass transfer can however not be maintained, because
the diffusion layer continues to grow. The result will be a current-potential curve
with a peak, positioned somewhat positive of E 0 for the analyte. This method is
referred to as linear sweep voltammetry, and is just one of many electroanalytical
techniques. Voltammetry2–5 is actually a collection of techniques where current
is measured as a function of applied potential. Two of these techniques will be
further described, i.e. cyclic voltammetry and pulse voltammetry.
Cyclic voltammetry is an extension of linear sweep voltammetry, and is often
used to study the redox behavior of electroactive species. The word cyclic comes
from the fact that when the potential has been swept between E1 and E2 , the direction of the sweep is reversed. The resulting triangular waveform can be seen in
Figure 2.9A. The potentials E1 and E2 are chosen to incorporate the redox behavior of the analyte, and is often centered around its E 0 -value. Another important
parameter is the sweep rate (or scan rate), normally ranging from 10 mV/s to a
few V/s. The current-potential curve obtained is called a cyclic voltammogram.
Figure 2.9B shows the result of a measurement in a solution containing a reversible
redox couple. During the reversed scan, the redox behavior of the species oxidized
during the forward scan can be evaluated. The concentration of an analyte can be
determined, since the current is proportional to this concentration. Also, the value
of E 0 can be estimated by the position of the oxidation and reduction peaks. In
fact, several important parameters such as reaction kinetics, surface adsorption,
and mass transfer are frequently studied with cyclic voltammetry.
In pulse voltammetry, the potential is instead changed in a stepwise manner.
There are many pulse parameters that can be changed, naturally giving different
14
Electrochemistry
(A)
Potential
E2
E1
Time
Current
(B)
Oxidation
0
Reduction
E1
E0
E2
Potential
Figure 2.9: (A) Triangular potential waveform applied to the working electrode in cyclic
voltammetry. (B) A schematic cyclic voltammogram for a reversible redox couple.
current responses. This technique provides the same general information as cyclic
voltammetry, but is especially known for low limits of detection, and better sensitivity and resolution. This is due to the fact that the influence of the charging
current can be minimized using an appropriate sampling of the current. Also,
short potential pulses will counteract the build-up of thick diffusion layers, and
redox species close to the electrode will consequently not be depleted. Both of
these effects will result in a better signal to noise ratio.
2.6
2.6.1
Electrochemical impedance spectroscopy
Theory
Electrochemical impedance spectroscopy (EIS)2, 3, 6 is a powerful tool for examining many chemical and physical processes in solutions as well as solid materials.
It is a non-destructive technique capable of separating different contributions to
the overall electrochemical processes. The basis for this is that the impedances
associated with different processes have varying frequency-dependencies. EIS can
therefore be used to study electrode kinetics, adsorption rates, corrosion processes,
battery properties, ageing of sensors, and many more.
Impedance (Z) is generally defined as the total opposition a device offers to the
flow of an alternating current at a given frequency. Impedance can be measured in
2.6 Electrochemical impedance spectroscopy
15
E(t) = E0sin(ωt)
I(t) = I0sin(ωt+θ)
Time
Phase
shift
Figure 2.10: Sinusoidal potential excitation and current response.
many different ways. Usually, a sinusoidal potential or current excitation signal is
applied to a device and the resulting current or potential is recorded, respectively.
If a potential excitation signal (E(t)) is used, expressed as a function of time,
it will have the form
E(t) = E0 sin (ωt)
(2.5)
where E0 is the amplitude of the signal and ω (rad/s) is the radial frequency. The
resulting current can be expressed as
I(t) = I0 sin (ωt + θ)
(2.6)
where It is the amplitude and θ is the phase difference between the potential and
current, see Figure 2.10. For a pure resistance, the phase difference is zero, and
for a pure capacitor it is 90◦ .
In the time domain, the relation between the system properties and the response is, however, very complex. If many capacitive and inductive elements
are present, a solution of a system of differential equations is required. By using
Fourier transformation of the signals into the frequency domain, the mathematical
treatment is significantly simplified. The Fourier transforms of the voltage, E(jω),
and current, I(jω), become E0 π and I0 π · exp(θj), respectively. The impedance
function can thus be defined as
Z(jω) =
E(jω)
I(jω)
(2.7)
and Z(jω) can be now be related to different circuit elements. For a resistance it
is simply equal to R, and for a capacitance and an inductance, Z(jω) is 1/(Cωj)
16
Electrochemistry
Im(Z)
Z’’
|Z|
θ
Z’
Re(Z)
Figure 2.11: Vector representation of the complex impedance.
and Lωj, respectively. Usually, Z(jω) is for simplicity written as just Z(ω). Now,
the impedance of a system with multiple elements can be calculated in the same
way as for multiple resistors.
Since the complex quantity Z(ω) contains information on both the magnitude
and the phase of the impedance, it is convenient to express it as a complex number.
Z = Re(Z) + jIm(Z) = Z 0 + jZ 00
(2.8)
The impedance is easily depicted in a plot of the imaginary part of the impedance
vs. the real part, as in Figure 2.11. The impedance is here represented by a
vector of length |Z|, and the angle between this vector and the x-axis is the phase
shift θ (=argZ). The real and imaginary parts of the impedance are termed the
resistance (R) and the reactance (X), respectively. It is worth noticing that the
original time variations of the applied and measured signals have disappeared in
the representation of the impedance. Some important relationships are
X
(2.9)
R
To display the variation of the impedance with frequency, a Bode plot is often
used. Here, log |Z| and θ are both plotted against log ω or log f . Another representation is a Nyquist plot, where X is plotted against R for different values of ω.
Examples of both these types of plots can for a series and parallel RC circuit be
seen in Figures 2.12 and 2.13, respectively.
Sometimes it is mathematically convenient to use the reciprocal of the impedance,
in which case
|Z| =
p
R2 + X 2
and
θ = arctan
1
1
=
= Y = G + jB
(2.10)
Z
R + jX
where Y is the admittance, G the conductance, and B the susceptance. The unit
of impedance is ohm (Ω), and the unit of admittance is siemen (S). Two more
2.6 Electrochemical impedance spectroscopy
17
-Im(Z)
ω
R
C
∞
Re(Z)
-θ
log |Z|
log ω
log ω
Figure 2.12: A series RC circuit (top left) and the resulting Nyquist plot (top right).
Bottom left and right are the corresponding Bode plots.
-Im(Z)
C
∞
ω
R
Re(Z)
-θ
log |Z|
log ω
log ω
Figure 2.13: A parallel RC circuit (top left) and the resulting Nyquist plot (top right).
Bottom left and right are the corresponding Bode plots.
18
Electrochemistry
quantities related to impedance are the modulus function M = jωCc Z = M 0 +
jM 00 , and the complex dielectric constant ε = M −1 = Y /(jωCc ) = ε0 + jε00 . Here,
Cc = ε0 Ac /l is the capacitance of the empty cell of electrode area Ac and electrode
separation length l. ε0 is the dielectric permittivity of free space, 8.854·10−12 F/m.
All four of these functions are valuable in EIS due to their different dependence
on, and weighting with, frequency.
An electrochemical cell can often be modeled as an equivalent circuit using
electrical components. Examples of circuit elements include a solution resistance,
a double layer capacitance, a charge transfer resistance, and an impedance term
representing mass transfer. Using such an equivalent circuit, impedance data can
now be evaluated in order to determine the physical and chemical properties of
the electrochemical system.
2.6.2
General applications
During the last couple of decades, EIS has been used for many different applications. Here, a brief summary of some important examples will be given. For a
thorough insight into the theory and applications of EIS, see the excellent text by
Barsoukov and Macdonald.6
The characterization of the interfacial region is very important in optimization
of electroanalytical sensors, and EIS is very useful for evaluating the processes occurring at the electrode/solution interface. The sinusoidal excitation will give rise
to an alternating charging and discharging of the double layer, as well as influencing
redox reactions and mass transfer. Hence, a lot of information may be obtained
since these processes can have different frequency-dependencies. For example,
Janek et al. used EIS to investigate the interfacial impedance of self-assembled
monolayers (SAMs) on gold surfaces.7 Surface modifications with DNA have also
been studied, for example the electron transfer through monolayers of thiol-labeled
DNA duplexes,8 as well as oligonucleotide-DNA interactions.9 Further, EIS has
been used for surface characterization of conductive polymer-modified electrodes,10
and solid polymer electrolytes.11
Corrosion research is an enormous field and EIS has been extensively used to
gain information on the corrosion processes at metals and metal-coated surfaces.12
Also, reaction mechanisms can be deduced, as shown by Keddam et al. who studied
iron dissolution in acidic media.13, 14
2.7
Bipolar electrochemistry
If the word bipolar is used to describe an object, the meaning will be that is has
two poles. In an electrochemical setup the poles are the anode and cathode. A
bipolar electrode is hence a single electrode, acting as both anode and cathode.
A requisite for this is that an oxidation and a reduction must take place at this
2.7 Bipolar electrochemistry
19
Voltage source
E
Reduction
Oxidation
i1
Oxidation
Reduction
i2
Feeder electrode
Feeder electrode
i1+i2
Bipolar electrode
Figure 2.14: Experimental setup showing the current paths when electrochemical reactions are taking place on the bipolar electrode. A variable voltage source is used to
control the total current between two feeder electrodes.
electrode simultaneously. Furthermore, these two reactions must be separated in
space to give rise to the two poles. The driving force for these reactions is an
electric field in the solution parallel or perpendicular to the electrode surface.
In Figure 2.14, an example of a bipolar setup is shown. Here, a voltage (or
current) source and two feeder electrodes are used to establish an electric field
in the solution. It should be noted that the isolated conducting surface has no
physical connection to the power source, and its potential is hence floating. The
potential difference between a point on the equipotential surface and the solution
will now vary laterally along the surface. This potential difference can induce
charge transfer reactions at both ends of the surface, provided that the electric
field exceeds a certain threshold value. At this point, the surface becomes a bipolar
electrode. A potential gradient across the electrode is thus induced by the electric
field in the solution. In practice, the bipolar electrode will act as an additional,
less resistive, path for the total current.
As apposed to a regular electrochemical setup, it is the applied external electric
field that is used to control the interfacial potentials at the electrode of interest.
The overall potential drop across the cell will now be non-linear when a floating
substrate acts as a bipolar electrode, and the current density distribution will also
be affected. Since the potential difference between a point on the bipolar electrode
and the solution will vary laterally along the surface, the rates of the reactions will
vary accordingly. Thus, electrochemical reaction gradients are present on bipolar
electrodes. The use of bipolar electrochemistry can be very advantageous, for
instance when connecting stacks of individual cells, as a result of the reduced
internal resistance.
Bipolar electrochemistry is a quite established field. For example, in 1973
Eardley et al.15 investigated the conductivity of fixed arrays of particles dispersed
in a continuous electrolyte medium. In recent years several new and interest-
20
Electrochemistry
ing applications have been reported. The phenomenon has been used in battery
applications,16, 17 in proton exchange membrane based fuel cells,18 for energy storage and load leveling,19, 20 as well as to carry out electrochemical reactions in
poorly conducting media.21 Further, Bradley et al. have in several publications
studied the use of bipolar electrochemistry for the directional growth of copper
wires between micrometer-sized particles,22 and the electrodeposition of palladium
catalysts.23, 24 Moreover, Said et al.25 used bipolar electrochemistry to visualize
the magnitude and direction of an electric field in an electrolyte. This type of
electrochemistry has also been used as an external electric field driven in-channel
detection technique in microfluidic PDMS channels.26 Finally, the use of bipolar
electrochemistry provides the possibility of studying chemical reactions without
any physical contact to the electrode, for instance in electrochemiluminescence
applications.27–29
Considerable theoretical contributions to the understanding of bipolar electrochemistry have been made by Duval et al.30–35 Here, simulations were made of
the electrochemical behavior of parallel plate electrodes in a lateral electric field,
in conjunction with electrokinetic experiments. It was shown how the potential
and current distribution can be evaluated for different setups.
2.8
Limitations of electrochemistry
Even though electrochemical techniques are very versatile, voltammetry and EIS
are more widely utilized as analytical tools within research than in day-to-day
use. This is related to some of the inherent limitations with electrochemical methods. First, there is the issue of stability of the electrode surface. The activity
and morphology of the electrodes frequently change over time, often as a result
of the adsorption of species from the solution. This is, however, something that
can be circumvented by the use of disposable electrodes, or by careful polishing
and cleaning procedures. Second, the limit of detection is normally in the micromolar range. Although this is sufficient for many applications, electrochemistry
cannot compete with other high-performance techniques. Third, is the issue of
selectivity. It is sometimes difficult to separate the response from a single analyte
from the total current. For instance, if two compounds are to be distinguishable
in voltammetry, their standard potentials must be sufficiently separated.
Finally, a rather important limitation is the fact that regular electrochemistry
offers no lateral resolution of the reactions and processes at electrode surfaces.
Apart from scanning techniques such as scanning electrochemical microscopy, a
collective current response from the entire surface is always obtained. However,
this limitation can be overcome by combining electrochemistry with imaging optical methods, such as ellipsometry and surface plasmon resonance.
3
Imaging optical methods and
electrochemistry
3.1
Introduction
One of the major advantages of electrochemistry is that it is easily combined with
other techniques. Successful combinations have been made with, for example,
ultraviolet and visible spectroscopy, vibrational spectroscopy, acoustic resonance
(i.e. quartz crystal microbalance), mass spectrometry, chromatography, and electrophoresis. The combination of imaging optical techniques and electrochemical
measurements is a very interesting area, because it allows the detection of electrochemical processes with lateral resolution in the micrometer range. This chapter
focuses on the combination of electrochemistry with imaging surface plasmon resonance and imaging ellipsometry.
3.2
3.2.1
Surface plasmon resonance
Theory
Surface plasmon resonance (SPR) spectroscopy is an optical method for measuring
the refractive index of very thin layers of material on a metal. SPR is a very
21
22
Imaging optical methods and electrochemistry
Evanescent field
Glass prism
Metal film
Ambient
Figure 3.1: Schematic illustration of the Kretschmann setup.
sensitive technique for real-time, label-free monitoring of events taking place at a
surface/liquid interface, and has been exploited for biosensing36 for over 20 years.
A surface plasmon is a p-polarized surface bound electromagnetic wave propagating at the interface between a plasma and a dielectricum. Since a metal has
nearly free electrons, it can be regarded as a plasma. Associated with the surface plasmon is an evanescent field, which is a non-propagating wave that decays
exponentially in the direction orthogonal to the direction of the interface. A surface plasmon can, under certain conditions, be excited by photons. In Figure 3.1
a common SPR setup, called the Kretschmann configuration, is shown. Here, a
light beam is totally reflected at the glass/metal interface, exciting a surface plasmon and an evanescent field. If the metal film is thin enough, the evanescent
field will penetrate into the ambient on the other side. The optical properties
of the interface region will affect the excitation resonance condition. Thereby, a
link is established between the ambient and the reflected light. Examples of SPR
biosensing applications include the study of biomolecular interactions and binding
properties, environmental monitoring, and medical diagnostics.37
Imaging SPR38 (also known as SPR microscopy) is attained when the detector
is a replaced by a charge-coupled device (CCD) camera. The lateral resolution,
typically in the µm range, is a function of the wavelength of the light, the objective,
and the CCD resolution. See Brockman et al.39 for an extensive review on the
theory and applications of imaging SPR.
3.2.2
SPR and electrochemistry
If the metal substrate in SPR also acts as a working electrode in an electrochemical experiment, these techniques can be employed simultaneously. Many
electrochemical reactions will change the optical properties of the electrode surface and the nearby solution, and hence change the SPR response. This fact has
been used for several different investigations of surface interactions. Iwasaki et
al.40 used the combination of SPR and cyclic voltammetry to investigate oxide
3.3 Ellipsometry
23
film formation and ion adsorption on gold electrodes. They also showed that it
was possible to separate the diffusion and adsorption processes. Further, Xiang et
al.41 showed that SPR in conjunction with scanning electrochemical microscopy
could be used to determine local variations in thin film thicknesses. The setup was
also used to study conformational changes of cytochrome c molecules attached to
a surface. Electropolymerization is a convenient way of forming a polymer film
on an electrode. This has been investigated by Knoll and co-workers for poly(3,4ethylenedioxythiophene) (PEDOT) films of different thicknesses.42, 43 Finally, Ertl
and co-workers used imaging SPR to obtain images of the potential distribution
at electrodes in order to study spatiotemporal patterns.44 The same setup was
also used to study Turing-type patterns on electrode surfaces.45
In this work Papers I, II, and III are based on the use of imaging SPR and
electrochemistry. It is especially the visualization of electrochemical reactions with
lateral resolution that has been explored.
3.3
3.3.1
Ellipsometry
Theory
Ellipsometry is an optical technique for characterization of surfaces, thin films, and
multilayers.46 The basic idea is to analyze the polarization changes when light is
reflected by a surface or film. Ellipsometry is non-destructive, making it suitable
for real-time in situ measurements. Furthermore, it is essentially insensitive to drift
in light intensity, since it is the polarization change that is evaluated. Ellipsometry
is very sensitive to surface changes, and can have sub-Ångström resolution in film
thickness determinations.
The electric field, E, associated with a wave of light, can be divided into two
components, Ep and Es . Ep is the p-polarization (or TM-polarization) and is the
component of the electric field that lies in the plane of incidence. Es is the spolarization (or TE-polarization) and its component is perpendicular to the plane
of incidence. It is the correlation between Ep and Es that determines the state of
polarization of the light.
In order to perform an ellipsometric measurement, an incident light beam with
known polarization is required. After reflection at a sample, the change in the
polarization of the light is determined. In Figure 3.2, a common setup, the PCSA
(polarizer compensator sample analyzer) ellipsometer, can be seen. The light
source is often a laser, emitting monochromatic, collimated light. The polarizer
and analyzer are both linear polarization filters, and the compensator is a quarter
wave plate, with which an arbitrary elliptical polarization can be obtained by
inducing a phase shift between the components of the light. As the beam travels
through the polarizer, it becomes linearly polarized. The compensator, which
is fixed at ±45◦ , will then produce elliptically polarized light. The reflection
24
Imaging optical methods and electrochemistry
Light source
Detector
Rotating polarizer
Fixed compensator
Rotating analyzer
Sample
Figure 3.2: Schematic representation of a PCSA ellipsometer.
at the sample will induce an additional phase shift, denoted 4, along with a
relative amplitude change, denoted tan Ψ. The beam then passes through the
analyzer before reaching the detector, where the intensity of the transmitted light
is measured. The strategy is to adjust the optical components so that the light is
extinguished at the detector. This will be possible if the light is linearly polarized
after the reflection at the surface, which means that it can be quenched by the
analyzer at a certain angle. By rotating the polarizer and the analyzer until zero
intensity is detected, 4 and tan Ψ can be determined using the angles of the optical
components. Now, using the two parameters together with a model of the system,
different surface characteristics can be estimated. For instance, the thickness of
a film on the surface can be calculated with the McCrackin algorithm,47 if the
refractive indices of the substrate and the film are known.
Imaging ellipsometry48 is an interesting development of ellipsometry, where the
regular detector is replaced by a CCD camera and a focusing objective. This allows
the simultaneous determination of for instance film thicknesses over the entire
surface area observed with the CCD. The lateral resolution will be determined by
the wavelength of the light, the CCD resolution, and the objective. Typical values
for the resolution are in the µm range. Imaging ellipsometry has for example been
used for biosensor applications, as shown by Jin et al.,49 where antigen-antibody
interactions were visualized.
3.3.2
Ellipsometry and electrochemistry
In combination with electrochemistry, ellipsometry is perhaps most widely used to
evaluate film formation and growth on electrode surfaces. This in situ combination
was pioneered by Tronstad50 in 1931, and has since gained popularity, see for
instance Kruger51 and Kruger and Calvert.52 Here, the thickness and optical
properties of passive iron films were evaluated, as the electrochemical potential
was either controlled or monitored.
3.3 Ellipsometry
25
In a review by Christensen and Hamnett,53 the in situ evaluation of electrodeposited polymer films was discussed. The influence of potential, and hence the
film thickness, on the ellipsometric angles and reflected intensity was shown. Another example is the in situ study of a self-assembled ferrocenylalkylthiol monolayer
on a gold electrode performed by Abrantes et al.54 They showed that the monolayer thickness increased 1 to 2 Å as the ferrocene was oxidized to ferricinium, due
to a reversible orientation change. Further, Yu and Jin55 used imaging ellipsometry to investigate the effect of the potential of a gold electrode on the adsorption
of fibrinogen.
In this thesis, imaging ellipsometry was used in Paper I to obtain thickness
maps of a protein gradient, formed via bipolar patterning. It proved to be a
very convenient way of visualizing the results from different preparation steps. In
Paper IV, ellipsometry was used in situ during the electrodeposition of nanostructured Cu and Cu2 O materials. This gave valuable information on the deposition
process, and the structure of the resulting film.
26
Imaging optical methods and electrochemistry
4
Electrode surface design and
analysis
4.1
Introduction
Enormous efforts have been put into the modification of electrodes to alter or
produce specific functionalities. Such modifications can be very advantageous, and
examples of application areas include fuel cells and batteries, the protection from
corrosion, and the use of modified electrodes as analytical sensors. Some 30 years
ago, interest arose regarding the modification of electrode surfaces by covalent
attachment of monolayers of different species.2, 56 Later, thicker polymeric films
and inorganic layers started to be used, as well as conducting polymers and organic
metals. This chapter focuses on the use of electrochemistry for both the design
and evaluation of modified electrodes.
4.2
Materials and methods
In this section, a summary of some important methods and materials used to
modify electrodes will be given, with some emphasis on self-assembly of alkanebased thiols on gold surfaces.
27
28
Electrode surface design and analysis
Polymers. A polymer is a macromolecule built up by repeating monomer units.
There is an abundance of different polymers available, and new ones can readily be synthesized. Polymer films can be formed on an electrode from a solution of dissolved polymer by, e.g., cast or dip coating, spin coating, covalent
attachment or electrodeposition. If the solution contains the monomer, polymerization on the surface can be induced via different means, e.g., thermal,
photochemical, or electrochemical. Conducting polymers contain partially
delocalized electrons, and are typically classified as semiconductors. Electroactive polymers contain electroactive components linked to the polymer
backbone, while ion-exchange polymers (polyelectrolytes) can attract ions
from the solution to charged sites in the film via ion-exchange processes.
Some electrochemical applications include work by Smela et al.,57 who used
a conducting polymer (polypyrrole) to fabricate microactuators. Further,
the ellipsometric characterization of electropolymerized polypyrrole was performed by for instance Kim et al.58 For an extensive review on conducting
polymers, see Simonet and Rault-Berthelot.59
Inorganic films. The most common inorganic material formed on electrodes are
metal oxides. By anodization of a metal electrode, an oxide film can easily
be grown, where the thickness is influenced by the potential and the time.
Interesting work in this area has been done by Iwasaki et al.40 and Stevenson
et al.60 In Paper IV, electrodeposition was used to form nanostructured Cu
and Cu2 O layers. Electrochemistry was here combined with a gravimetric
technique (quartz crystal microbalance, QCM) and ellipsometry to investigate the potential and local pH dependence on the nature of the deposited
materials.
Self-assembled monolayers (SAMs). The formation of a self-assembled monolayer is a spontaneous process, leading to a molecular film with an often
large degree of order, as a result of the lateral interactions between the adsorbed molecules. One of the most investigated systems is the self-assembly
of organosulfur (thiol) compounds. In 1983, Nuzzo and Allara showed that
gold surfaces could be easily functionalized by the spontaneous self-assembly
of organic disulfides.61 The tail-group of a thiol can be chosen to display different physical and chemical properties at the ambient interface. SAMs can
also be used as linkers for the attachment of biomolecules to surfaces, which
is usually done in order to prepare electrochemical sensors. The most developed application is a surface modified with an enzyme, such as glucose
oxidase.62 Further, patterns of SAMs can be formed by microcontact printing,63 giving linewidths ranging from 30 nm to 500 µm. Informative reviews
on the subject are given by Ulman64 and Whitesides and co-workers.65
Extensive work has been done in order to characterize and optimize self-
4.3 Surface gradients
29
assembly of thiol based monolayers. The most commonly used substrate
is gold, but SAMs can also be formed on other metals.66 Self-assembly
from a solution containing two different thiol species is also possible. Such
co-adsorbtion will result in SAMs having various degrees of phase separation.67–73 One important application of SAMs is the formation of protein
resistant surfaces. Li et al.74 concluded that SAMs of oligo(ethylene glycol) thiols with different lengths displayed different surface morphology and
thereby different protein resistance.
A SAM on a surface will have a thickness of the same order of magnitude
as the diffuse layer. This means that the total capacitance of a modified
electrode will be greatly influenced by the dielectric properties of the surface
film. Therefore, electrochemical techniques are well suited for studying thin
self-assembled systems. An early example is an investigation by Porter et
al.,75 focusing on the influence of alkanethiol chain length on heterogeneous
electron transfer and interfacial capacitance. Voltammetry has further been
used to characterize a mixed monolayer of thiol analogues of cholesterol and
fatty acids on gold,76 as well as to examine the chemistry of the bound alkanethiol head group.77 In the latter work, it was also shown that the adsorbed
thiols could be desorbed from the surface by both oxidation and reduction.
In a recent study, Chan and Yousaf 78 demonstrated that the ligand activity of an electroactive monolayer could be controlled by the electrochemical
potential.
The degree of organization of a SAM on electrode surfaces has a huge impact on its optical and electrical properties. Different packing densities and
porosities will for instance affect the ability of the SAMs to impede charge
transfer reactions. This was investigated in Paper III, where a pattern
of microcontact printed SAMs of thiocholesterol and 1-hexadecanethiol was
used to investigate the possibility of detecting local electrochemical processes
in situ on chemically modified electrodes.
4.3
Surface gradients
A surface with a gradually changing chemical or physical property is said to contain
a surface gradient. Many types of gradients are found in nature, for instance
concentration gradients across cell membranes and biomolecular gradients guiding
the motility of cells. The mimicking of these in vivo gradients in the lab can
be a valuable tool to better understand biological processes. Also, by having
gradually changing properties on a biosensor surface, it is possible to perform highthroughput and cost effective analysis. Since all electroanalytical devices based on
biomolecular interactions have to be thoroughly optimized and investigated, the
access to a surface gradient of the interacting species would be very advantageous.
30
Electrode surface design and analysis
i1+i2
Reduction
Oxidation
i1
Oxidation
Reduction
i2
Cathodic side
Figure 4.1: Illustration of alkanethiol desorption from the cathodic side of a bipolar
electrode.
Two good examples of the use of molecular gradients are the study of surface
tension effects79, 80 and protein adsorption.81, 82 Recent reviews on the subject are
given by Kim et al.,83 Morgenthaler et al.,84 and Genzer and Bhat.85
Several gradient-forming approaches have been demonstrated to date; these
approaches are based, for example, on diffusion,86, 87 electric fields88 ,89 microfluidic systems,90 and immersion procedures.91, 92 Bohn and co-workers proposed an
electrochemical method to form gradients on very thin metal substrates (about
50 nm) by using in-plane electrochemical potentials.93–105 In that method, a potential gradient was generated across the surface by passing a current through
the resistive film using a bipotentiostat. Gradients with varying widths and positions could thus be created on the surface. This method has also been used
to deposit copper gradients in nanoporous alumina membranes,106 to manipulate
droplet transportation by the spatiotemporal control of a wetting gradient,107 and
to electrodeposit polymer films in various patterns.108
Based on the use of bipolar electrochemistry, we have developed a new technique to produce surface gradients, which we call bipolar patterning. Here, we
use the fact that the driving force for electrochemical reactions will vary laterally
across a bipolar electrode. If the result of such reactions on either end of the
electrode is the adsorption or desorption of a specific molecule, a gradient of that
molecule can be created on the surface. An example of this is seen in Figure 4.1.
In Paper I, this technique was used to form a biomimetic gradient on a gold
surface. Briefly, self-assembled monolayers and protein immobilization procedures
were used to realize a protein gradient. The first step involved the reductive desorption of a SAM based on methoxy terminated poly(ethylene glycol)-containing
alkanethiols. After backfilling with carboxyl terminated poly(ethylene glycol)-
4.3 Surface gradients
31
(A)
Thickness (nm)
5
3
4
~0.6 nm
3
~2.4 nm
2
2
1
1
0
0
0.1
0.2
0.3
0.4
0.5
x (mm)
0.6
0.7
0.8
0.9
Thickness (nm)
(B)
4.5
4
3.5
3
2.5
0
1
0.2
0.4
x (mm)
0.6
0.8
0
0.5
m)
y (m
Figure 4.2: (A) Line profiles obtained from imaging ellipsometry measurements, in
which the thicknesses of the gradients after the different preparation steps are shown.
Line 1 shows the result of the desorption of HS−C2 H4 −(O−C2 H4 )6 −OCH3 , line 2 shows
that obtained after backfilling with HS−C2 H4 −(O−C2 H4 )8 −COOH, and line 3 represents the resulting protein gradient. (B) Thickness map of the protein gradient (the solid
line shows the region from which the line profile was taken). See Paper I for further
details.
containing alkanethiols and activation of the carboxyl groups, the surface was
incubated in a lysozyme solution. By employing imaging ellipsometry, the results
of the preparation steps and the protein gradient could be evaluated (Figure 4.2).
The resulting protein gradients formed with this specific setup were about 0.5–1
mm wide. Bipolar patterning is a fast and straightforward technique with many
possibilities for improvements, for instance the generation of gradients of different
geometries. The substrate can basically consist of any conducting material and
the technique does not require access to advanced laboratory equipment.
Gradients with widths in the millimeter range are sufficient for many applications, but sometimes wider gradients and other spatial geometries are required. A
clear knowledge of the effects of the electric field in a solution containing a conducting substrate will make it possible to optimize this gradient forming technique.
The distributions of the potential and current densities in a bipolar setup can be
evaluated by rather simple methods involving the use of ordinary reference electrodes. Also, the possibilities offered by simple conductivity models was evaluated
to perform simulations of new and exciting geometries. In Paper II, the potential
Electrode surface design and analysis
1.4
20
1.2
10
1
0
0.8
-10
0.6
-20
0.4
-20
-10
0
10
20
Position (mm)
30
40
-30
1.6
30
(B)
1.4
1.2
20
10
1
0
0.8
-10
0.6
-20
0.4
-20
-10
0
10
20
Position (mm)
30
Solution potential (mV)
30
(A)
Relative current density
Relative current density
1.6
Solution potential vs. surface (mV)
32
-30
40
Figure 4.3: Solution potentials and relative current densities for different positions with
respect to the bipolar electrode (0.5 mm above the surface). The shaded areas represent
the position of the surface. (A) Experimental results as 1 mA was passed in a solution
consisting of 10 mM [Fe(CN)6 ] 4 – and 10 mM [Fe(CN)6 ] 3 – in 500 mM KNO3 . (B)
Simulated results using COMSOL Multiphysics. A simple conductive media DC model
of the experimental setup was utilized. See Paper II for further details.
and relative current density distributions in a solution containing a bipolar electrode was investigated, see Figure 4.3A. It is clear that the presence of a bipolar
electrode has changed both distributions, which would otherwise be linear. As can
be seen from the current density, the amount of electrochemical reactions varies
laterally across the surface. In Figure 4.3B, results from a simulation of the setup
used to obtain the experimental results are shown. The important result is that
a good qualitative agreement was obtained between the measured and simulated
values. This demonstrates the value of the model to predict important parameters
in bipolar setups.
Further, the use of imaging SPR provided the possibility to evaluate the amount
of reactions occurring at the bipolar electrode. Since the extent of these reactions
are governed by the potential difference between the bipolar electrode and the
solution, the SPR response will be a function of this potential difference. By
comparing the SPR responses with those obtained in a regular three-electrode
setup (including a standard reference electrode), the potential distribution very
close to the surface can be determined (Figure 4.4).
To demonstrate the relevance of bipolar patterning for biomimetic and biosensor applications, a spherical electric field was used to selectively remove thiols
from a surface. Here, one feeder electrode was a platinum rod, placed above the
monolayer. The resulting effective layer thickness is shown in Figure 4.5. This
method could now be further used to create an array of such gradient regions on
a single substrate.
4.3 Surface gradients
33
0.05
(A)
(B)
10 mA
1
5 mA
0.5
1 mA
0
0
5 mA (in KNO3)
i (mA)
∆RTM/RTE
1.5
-0.5
-1
-1.5
-0.05
2
2.5
3
3.5
Position (mm)
4
-0.1
0.1
0.3
E vs. Ag/AgCl (V)
0.5
d (nm)
Figure 4.4: (A) The SPR response for different currents passed through the electrolyte,
simultaneously showing the reduction (left side) and oxidation (right side). The electrolyte consisted of 10 mM [Fe(CN)6 ] 4 – and 10 mM [Fe(CN)6 ] 3 – in 500 mM KNO3 .
(B) SPR response for different potentials when the sensor surface acted as a working
electrode in a three-electrode setup. The solid line was calculated using the Nernst equation. The scan rate used to obtain the cyclic voltammogram was 50 mV/s. See Paper II
for further details.
2.5
2
1.5
0.5
y (mm)
0
0
0.5
1
x (mm)
1.5
2
Figure 4.5: Results from imaging null-ellipsometry measurements, showing the effective
layer thickness, d, of the patterned thiol layer. The inset shows a larger image and the
region from where the thickness map was taken. See Paper II for further details.
34
Electrode surface design and analysis
5
Alternating electric fields for
chemical analysis
5.1
Introduction
It is often desirable to monitor the state, or quality, of a liquid. This is sometimes
a difficult task when relying on techniques based on static electric fields, due to
the often varying conductivity of the samples and the complexity of the solution.
These problems can be circumvented by using EIS, which is based on alternating
electric fields. We have utilized EIS in two different applications; the simultaneous
determination of soot and diesel contamination in engine oil, and the determination of concentration and pH in an industrial cutting fluid. Both of these liquids
contain very low amounts of electroactive species, and they can further have very
different conductivities. Evaluations of these can however still be performed with
EIS due to the possibility to separate different processes based on their frequencydependencies. Since this technique is very sensitive, it is of crucial importance to
use a carefully designed measurement setup.
35
36
5.2
Alternating electric fields for chemical analysis
Practical considerations
When performing EIS measurements, several practical aspects have to be considered. A lot can be gained by using a correctly optimized experimental setup. There
are several commercially available test fixtures and cells, but the prices are often
quite high. Sometimes it is hence worth the time and effort to design and optimize
an application-specific measurement setup. When evaluating a solid sample in order to gain information on the bulk properties, the role of the electrodes is simply
to provide electrical contact to the material. For liquids, electrochemical reactions
and other surface interactions can take place at the electrode/solution interfaces.
In these cases a parallel plate setup is often used, where the plates consist of an
inert metal. The electrode area and separation can be varied depending on the
application. Listed below are some other important parameters that need to be
considered.
System. Most often a potentiostat is used together with a frequency response
analyzer (FRA) for EIS measurements. When potentiostatic control is not
required, as is the case for most materials impedance tests, a self-contained
impedance analyzer can be used instead. Many different systems are commercially available, with varying specifications regarding sensitivity and frequency range.
Terminal configuration. There are several connection configurations available,
and the choice depends on the expected impedance range and the frequencies
used. Most modern equipments have four terminals; high and low potential
and high and low current. It is between the high and low current terminals
that the excitation signal is applied, and the resulting signal is measured
between the high and low potential terminals. If the two high and the two
low terminals are shorted, a two-terminal configuration is obtained, which
is sufficient for many high impedance measurements. A configuration providing a very wide measurement range is the four-terminal configuration.
This ensures that only the potential drop across the cell is measured, and
eliminates the effects of the cables connected to the high and low current
terminals.
Cables. The cables for connecting the cell to the measuring equipment need to
be carefully designed. Preferably a short, static, and well shielded cable
assembly should be used. It is also very important to make sure that all
connections are good in order to not introduce resistive loads outside the
sample volume. Otherwise these will be included in the total impedance.
Shielding. For sensitive measurements, it is always recommended to shield the
cell from external interference by placing it in a Faraday cage. This is true
5.3 Applications
37
Coaxial
cable
Hc
Hp
Shielding box
Triaxial
cable
Electrode 2
Sample
Lp
~200µm
Electrode 1
Lc
Figure 5.1: Schematic representation of the measurement set-up showing the impedance
meter, the four-terminal cable configuration, the shielding box, and the electrodes. Hc:
High current, Hp: High potential, Lp: Low potential, Lc: Low current.
for both low and high impedances. The cage should be connected to the
shield of the instrument input terminals.
Temperature. The impedance is generally a function of temperature. In order
to compare results from different samples, it is very important to be able to
control or at least to measure the temperature of the sample.
A schematic representation of the setup used in Papers V and VI can be
seen in Figure 5.1. In both cases, the electrodes were separated by approximately
200 µm, and consisted of stainless steel in a parallel plate cell configuration. A
four-terminal configuration was employed together with a coaxial/triaxial cable
assembly, to be able to perform measurements over a large impedance range.
5.3
Applications
General applications of EIS were given in section 2.6. Here, the focus will be on
the chemical analysis of lubricants and cutting fluids. These liquids are abundant in industrial applications and there is presently a huge demand for on-line
measurement techniques.
Smiechowski and Lvovich have made several interesting EIS studies on industrial lubricants. Some examples include the detection of water leaks in engine
oil,109 the analysis of colloidal dispersions,110 and the investigation of the relationship between lubricant chemical composition and EIS data.111 Further, Allahar
et al.112 studied the impedance of steels in new and degraded jet engine oil, while
Wang and Lee113 focused on the glycol contamination in regular engine oil. Wang
and co-workers have also employed fast scan voltammetry to determine the condition of similar samples.114–117 Here, the maximum current output was measured
and correlated to degradation processes in the oil. The obvious drawback was the
use of only one parameter to evaluate the condition of the sample.
The quality of the engine oil greatly affects the performance of an engine, and
the need for on-board sensors for real-time monitoring is increasing. In Paper V,
38
Alternating electric fields for chemical analysis
8
1S1D
4S1D
3S3D
1S5D
4S5D
7
6
Diesel (%)
5
4
3
2
1
0
-1
1
1.5
2
2.5
3
Soot (%)
3.5
4
4.5
5
Figure 5.2: The result from soot and diesel predictions made with impedance data and
PLS modeling. The cross-hair markers represent the true concentrations of soot and
diesel for the five samples. The samples are labeled according to their levels of soot (S)
and diesel (D). See Paper V for further details.
the simultaneous estimation of soot and diesel in engine oil was shown to be
feasible. In these experiments, a carefully planned measurement cell was employed,
consisting of two stainless steel electrodes in a parallel plate configuration. Real
engine oil samples with varying amounts of soot were spiked with diesel and then
evaluated using a frequency range of 600 kHz to 20 Hz. Due to the low conductivity
of the samples, the voltage amplitude was set to 20 Vrms . The data was then
evaluated with multivariate data analysis, namely partial least squares (PLS).
The use of statistical tools for the interpretation of impedance data can be very
informative, since it quickly provides information on the correlation between the
data and the sought parameters. It also allows the classification of the samples into
categories. In Figure 5.2, the results of predictions made with two PLS models are
shown, one for soot and one for diesel. As can be seen, samples with different soot
concentrations are easily classified, with good accuracy. The diesel variations were
harder to predict. The analysis would most likely be improved if lower frequencies
could have been used.
One drawback of the cell used in Paper V was the lack of temperature control.
Therefore, a new cell was designed, including heaters and a temperature sensor.
To simplify the sample application procedure, it was further designed as a flow
cell. In Paper VI, this cell was used for quality evaluation of industrial cutting
liquids. In this study, the frequency range was expanded (1 MHz–10 mHz), and
the amplitude was much lower (40 mVrms ). The correlation between EIS data
5.3 Applications
39
1
Nitrite
Oil
pH
Correlation
0.5
θ
|Z|
0
-0.5
-1
10
6
-2
6
10
10
Frequency (Hz)
10
-2
Figure 5.3: Correlation of EIS data to cutting fluid (oil) concentration, pH, and nitrite
level. The left part represents |Z| and the right part θ. See Paper VI for further details.
11
pH
Concentration (% (w/w))
(A)
15
10
(B)
9.5
8
5
Samples
Samples
Figure 5.4: PLS training data for the cutting fluid concentration (A) and the pH level
(B). Real values (no markers) are also shown. See Paper VI for further details.
and the cutting fluid concentration, the pH, and the nitrite level is shown in
Figure 5.3. It is evident that the concentration and the pH correlate well with
the measured variables, and more importantly, that both parameters have different
frequency dependencies. Interesting is also the fact that when the correlation for
one parameter is zero, the correlation for the other is often quite high. This should
provide the means to separate their individual contributions to the impedance.
The nitrite level, however, correlated poorly to the impedance and was therefore
not further evaluated. In Figure 5.4, the PLS training data for the concentration
and pH is shown. The figures indicate the possibility of predicting both parameters
simultaneously. If more samples were evaluated, model stability would most likely
improve. This would also permit validations to be performed. However, the next
step should in our opinion be an on-line analysis of real samples.
40
Alternating electric fields for chemical analysis
6
Future outlook
One of the major conclusions reached during the work underlying this thesis, is that
electric fields in their various forms are excellent tools for surface design. Static
and alternating, they also allow the analysis of surface and interfacial properties,
as well as chemical analysis of the bulk of a solution.
The use of bipolar electrochemistry for the formation of surface gradients is
a new and exciting method. In fact, in the last decade or so, many interesting
applications of bipolar electrochemistry have been reported. It is especially the
lack of an electrical contact to the electrode that makes this technique quite unique.
We have shown that bipolar patterning can be used to create surface gradients
of many different materials. Interestingly, the width, position, and geometry of
the gradients can be controlled. The possibility to predict the bipolar behavior by
simulations also allows the optimization of the experimental setup. This knowledge
can be used to form new and exciting gradients, with varying functionalities and
geometries. Ongoing work hence deals with bipolar patterning with spherical
electric fields. Here, one of the feeder electrodes is placed above a substrate,
inducing a circular reaction region on the bipolar electrode. With this, it is possible
to create several local gradients on one surface. This approach could for instance
be used to form an array of circular gradient spots on a single substrate.
There are several ways to increase the widths of the gradients. If the bipolar
electrode, for instance, is placed at an angle relative to the feeder electrodes,
41
42
Future outlook
the gradient will be widened. Another interesting concept is to use two bipolar
electrodes, connected to an external voltage source. We have also worked with the
electrodeposition of polymers, and initial experiments show the possibility to form
gradients in the degree of doping of polypyrrole.
In this thesis, EIS has been extensively employed for quality determination in
industrial applications. We have explored the combination of EIS and statistical
tools, which proved to be an easy way to quickly obtain information on the state of
a sample. The evaluation of EIS data to gain insight on intrinsic parameters, can
however be a quite complex task. It would be very interesting to use one of the
softwares available to parameterize data and extract physical quantities. Finally,
the implementation of an on-line test setup would really put the method to the
test. This would also require optimization of the setup and electronics, together
with a model based on the interactions present.
Bibliography
[1] M. Faraday. Experimental researches in electricity. Seventh series. Philosophical
Transactions of the Royal Society of London, 124:77–122, 1834.
[2] A.J. Bard and L.R. Faulkner. Electrochemical Methods — Fundamentals and
Applications. John Wiley & Sons, Inc., 2nd edition, 2000.
[3] C.H. Hamann, A. Hamnett, and W. Vielstich. Electrochemistry. Wiley-VCH, 1998.
[4] C.M.A. Brett and A.M.O. Brett. Electrochemistry – Principles, Methods, and
Applications. Oxford University Press, 1993.
[5] D.C. Harris. Quantitative Chemical Analysis. W. H. Freeman and Company, 5th
edition, 1998.
[6] E. Barsoukov and R.J. Macdonald, editors. Impedance Spectroscopy — Theory,
Experiment, and Applications. John Wiley & sons, Inc., 2nd edition, 2005.
[7] R.P. Janek, W.R. Fawcett, and A. Ulman. Impedance spectroscopy of selfassembled monolayers on Au(111): evidence for complex double-layer structure
in aqueous NaClO4 at the potential of zero charge. Journal of Physical Chemistry
B, 101:8550–8558, 1997.
[8] Y.-T. Long, C.-Z. Li, H.-B. Kraatz, and J.S. Lee. AC impedance spectroscopy of
native dna and m-dna. Biophysical Journal, 84:3218–3225, 2003.
[9] A. Bardea, F. Patolsky, A. Dagan, and I. Willner. Sensing and amplification of
oligonucleotide-dna interactions by means of impedance spectroscopy: a route to a
tay-sachs sensor. Chemical Communications, 1:21–22, 1999.
[10] B.J. Johnson and S.-M. Park. Electrochemistry of conductive polymer. Journal of
the Electrochemical Society, 143:1269–1276, 1996.
[11] D.K. Cha and S.-M. Park. Electrochemical charaterization of polyethylene glycols
as solid polymer electrolytes. Journal of Electroanalytical Chemistry, 459:135–144,
1998.
[12] B. Pejcic and R. De Marco. Impedance spectroscopy: over 35 years of electrochemical sensor optimization. Electrochimica Acta, 51:6217–6229, 2006.
[13] M. Keddam, O.R. Mattos, and H. Takenouti. Reaction model for iron dissolution
studied by electrode impedance, I. experimental results and reaction model. Journal
of the Electrochemical Society, 128:257–266, 1981.
[14] M. Keddam, O.R. Mattos, and H. Takenouti. Reaction model for iron dissolution
studied by electrode impedance, II. determination of the reaction model. Journal
of the Electrochemical Society, 128:266–274, 1981.
43
44
Bibliography
[15] D.C. Eardley, D. Handley, and S.P.S Andrew. Bipolar electrolysis with intra phase
conduction in two phase media. Electrochimica Acta, 18:839–848, 1973.
[16] H. Karami, M.F. Mousavi, and M. Shamsipur. A novel dry bipolar rechargeable
battery based on polyaniline. Journal of Power Sources, 124:303–308, 2003.
[17] K. Wiesener, D. Ohms, G. Benczúr-Ürmössy, M. Berthold, and F. Haschka. High
power metal hydride bipolar battery. Journal of Power Sources, 84:248–258, 1999.
[18] V. Mehta and J.S. Cooper. Review and analysis of pem fuel cell design and manufacturing. Journal of Power Sources, 114:32–53, 2003.
[19] F.C. Walsh. Electrochemical technology for environmental treatment and clean
energy conversion. Pure and Applied Chemistry, 73:1819–1837, 2001.
[20] A. Price, S. Bartley, S. Male, and G. Cooley. A novel approach to utility scale
energy storage. Power Engineering Journal, 13:122–129, 1999.
[21] M. Fleischmann, J. Ghoroghchian, D. Rolison, and S. Pons. Electrochemical behavior of dispersions of sperical ultramicroelectrodes. Journal of Physical Chemistry,
90:6392–6400, 1986.
[22] J.-C. Bradley, H.-M. Chen, J. Crawford, J. Eckert, K. Ernazarova, T. Kurzeja,
M. Lin, M. McGee, W. Nadler, and S.G. Stephens. Creating electrical contacts
between metal particles using directed electrochemical growth. Nature, 389:268–
271, 1997.
[23] J.-C. Bradley and Z. Ma. Contactless electrodeposition of palladium catalysts.
Angewandte Chemie International Edition, 38:1663–1666, 1999.
[24] J.-C. Bradley, S. Babu, and P. Ndungu. Contactless tip-selective electrodeposition
of palladium onto carbon nanotubes and nanofibers. Fullerenes, Nanotubes, and
Carbon Nanostructures, 13:227–237, 2005.
[25] E. Said, N.D. Robinson, D. Nilsson, P.-O. Svensson, and M. Berggren. Visualizing
the electric field in electrolytes using electrochroism from a conjugated polymer.
Electrochemical and Solid-State Letters, 8:H12–H16, 2005.
[26] O. Ordeig, N. Godino, J. del Campo, F.X. Munoz, F. Nikolajeff, and
L. Nyholm. On-chip electric field driven electrochemical detection using a
poly(dimethylsiloxane) microchannel with gold microband electrodes. Analytical
Chemistry, 80:3622–3632, 2008.
[27] A. Arora, J.C.T. Eijkel, W.E. Morf, and A. Manz. A wireless electrochemiluminescence detector applied to direct and indirect detection for electrophoresis on a
microfabricated glass device. Analytical Chemistry, 73:3282–3288, 2001.
[28] W. Zhan, J. Alvarez, and R.M. Crooks. Electrochemical sensing in microfluidic
systems using electrogenerated chemiluminescence as a photonic reporter of redox
reactions. Journal of the American Chemical Society, 124:13265–13270, 2002.
[29] K.-F. Chow, F. Mavré, and R.M. Crooks. Wireless electrochemical dna microarray
sensor. Journal of the American Chemical Society, 130:7544–7545, 2008.
[30] J. Duval, J.M. Kleijn, and H.P. van Leeuwen. Bipolar electrode behaviour of the
aluminium surface in a lateral electric field. Journal of Electroanalytical Chemistry,
505:1–11, 2001.
Bibliography
45
[31] J.F.L. Duval, G.K. Huijs, W.F. Threels, J. Lyklema, and H.P. van Leeuwen.
Faradaic depolarization in the electrokinetics of the metal-electrolyte solution interface. Journal of Colloid and Interface Science, 206:95–106, 2003.
[32] J.F.L. Duval, M. Minor, J. Cecilia, and H.P. van Leeuwen. Coupling of lateral electric field and transversal faradaic processes at the conductor/electrolyte solution
interface. Journal of Physical Chemistry B, 107:4143–4155, 2003.
[33] J.F.L. Duval, H.P. van Leeuwen, J. Cecilia, and J. Galceran. Rigorous analysis of reversible faradaic depolarization processes in the electrokinetics of the
metal/electrolyte solution interface. Journal of Physical Chemistry B, 107:6782–
6800, 2003.
[34] J.F.L Duval. Electrokinetics of the amphifunctional metal/electrolyte solution interface in the presence of a redox couple. Journal of Colloid and Interface Science,
269:211–223, 2004.
[35] J.F.L. Duval, J. Buffle, and H.P. van Leeuwen. Quasi-reversible faradaic depolarization processes in the electrokinetics of the metal/solution interface. Journal of
Physical Chemistry B, 110:6081–6094, 2006.
[36] B. Liedberg, C. Nylander, and I Lundström. Surface plasmon resonance for gas
detection and biosensing. Sensors and Actuators, 4:299–304, 1983.
[37] J. Homola. Surface Plasmon Resonance Based Sensors, volume 4 of Springer Series
on Chemical Sensors and Biosensors. Springer-Verlag, Berlin Heidelberg, 2006.
[38] B. Rothenhäusler and W. Knoll. Surface-plasmon microscopy. Nature, 332:615–
617, 1988.
[39] J.M. Brockman, B.P. Nelson, and R.M. Corn. Surface plasmon resonance imaging
measurements of ultrathin organic films. Annual Review of Physical Chemistry,
51:41–63, 2000.
[40] Y. Iwasaki, T. Horiuchi, M. Morita, and O. Niwa. Time differential surface plasmon resonance measurements applied for electrochemical analysis. Electroanalysis,
9:1239–1241, 1997.
[41] J. Xiang, J. Guo, and F. Zhou. Scanning electrochemical microscopy combined
with surface plasmon resonance: studies of localized film thickness variations and
molecular conformation changes. Analytical Chemistry, 78:1418–1424, 2006.
[42] C. Xia, R.C. Advincula, A. Baba, and W. Knoll. In situ investigations of the electrodeposition and electrochromic properties of poly(3,4-ethylenedioxythiophene)
ultrathin films by electrochemical-surface plasmon spectroscopy.
Langmuir,
18:3555–3560, 2002.
[43] A. Baba, J. Lübben, K. Tamada, and W. Knoll. Optical properties of ultrathin
poly(3,4-ethylenedioxythiophene) films at several doping levels studied by in situ
electrochemical surface plasmon resonance spectroscopy. Langmuir, 19:9058–9064,
2003.
[44] G. Flätgen, K. Krischer, B. Pettinger, K. Doblhofer, H. Junkes, and G. Ertl. Twodimensional imaging of potential waves in electrochemical systems by surface plasmon microscopy. Science, 269:668–671, 1995.
46
Bibliography
[45] Y.-J. Li, J. Oslonovitch, N. Mazouz, F. Plenge, K. Krischer, and G. Ertl. Turingtype patterns on electrode surfaces. Science, 291:2395–2398, 2001.
[46] R.M. Azzam and N.M Bashara. Ellipsometry and Polarized Light. North-Holland,
1988.
[47] F.L. McCrackin. A fortran program for analysis of ellipsometer measurements.
NBS Technical Note, 479, 1969.
[48] A.-H. Liu, P.C. Wayner, and J.L. Plawsky. Image scanning ellipsometry for measuring nonuniform film thickness profiles. Applied Optics, 33:1223–1229, 1994.
[49] G. Jin, P. Tengvall, I. Lunström, and H. Arwin. A biosensor concept based on
imaging ellipsometry for visualization of biomolecular interactions. Analytical Biochemistry, 232:69–72, 1995.
[50] L. Tronstad. Optische untersuchungen zur frage der passivität des eisens und stahls.
Det Kgl Norske Videnskabers Selskabs Skrifter, 1931.
[51] J. Kruger. Optical studies of the formation and breakdown of passive films formed
on iron single crystal surfaces in inorganic inhibitor solutions. Journal of the Electrochemical Society, 110:654–663, 1963.
[52] J. Kruger and J.P. Calvert. Ellipsometric-potentiostatic studies of iron passivity.
Journal of the Electrochemical Society, 114:43–49, 1967.
[53] P. Christensen and A. Hamnett. In-situ techniques in electrochemistry – ellipsometry and ftir. Electrochimica Acta, 45:2443–2459, 2000.
[54] L.M. Abrantes, M. Kalaji, and A.S. Viana. In situ ellipsometric study of redox
induced orientation of a short chained ferrocenylalkylthiol monolayer self-assembled
on gold. Russian Journal of Electrochemistry, 38:39–43, 2002.
[55] Y. Yu and G. Jin. Influence of electrostatic interaction on fibrinogen adsorption
on gold studied by imaging ellipsometry combined with electrochemical methods.
Journal of Colloid and Interface Science, 283:477–481, 2005.
[56] R.W. Murray. Chemically modified electrodes. Accounts of Chemical Research,
13:135–141, 1980.
[57] E. Smela, O. Inganäs, and I. Lundström. Controlled folding of micrometer-size
structures. Science, 268:1735–1738, 1995.
[58] Y.-T. Kim, R.W. Collins, K. Vedam, and D.L. Allara. Real time spectroscopic
ellipsometry: in situ characterisation of pyrrole electropolymerization. Journal of
the Electrochemical Society, 138:3266–3275, 1991.
[59] J. Simonet and J. Rault-Berthelot. Electrochemistry: a technique to form, to
modify and to charaterize organic conducting polymers. Progress In Solid State
Chemistry, 21:1–48, 1991.
[60] K.J. Stevenson, G.J. Hurtt, and J.T. Hupp. High resolution assembly of patterned
metal oxide thin films via microtransfer molding and electrochemical deposition
techniques. Electrochemical and Solid-State Letters, 2:175–177, 1999.
[61] R.G. Nuzzo and D.L. Allara. Adsorption of bifunctional organic disulfides on gold
surfaces. Journal of the American Chemical Society, 105:4481–4483, 1983.
Bibliography
47
[62] J. Wang. Glucose biosensors: 40 years of advances and challenges. Electroanalysis,
12:983–988, 2001.
[63] Y. Xia and G.M. Whitesides. Soft lithography. Annual Review of Materials Science,
28:153–184, 1998.
[64] A. Ulman. Formation and structure of self-assembled monolayers. Chemical Reviews, 96:1533–1554, 1996.
[65] J.C. Love, L.A. Estroff, J.K. Kriebel, R.G. Nuzzo, and G.M. Whitesides. Selfassembled monolayers of thiolates on metals as a form of nanotechnology. Chemical
Reviews, 105:1103–1169, 2005.
[66] P.E. Laibinis and G.M. Whitesides. ω-treminated alkanethiolate monolayers on surfaces of copper, silver, and gold have similar wettabilities. Journal of the American
Chemical Society, 114:1990–1995, 1992.
[67] S.J. Stranick, A.N. Parikh, Y.-T. Tao, D.L. Allara, and P.S. Weiss. Phase separation of mixed-composition self-assembled monolayers into nanometer scale molecular domains. Journal of Physical Chemistry, 98:7636–7646, 1994.
[68] C.D. Bain, J. Evall, and G.M. Whitesides. Formation of monolayers by the coadsorption of thiols on gold: variation in the head group, tail group, and solvent.
Journal of the American Chemical Society, 111:7155–7164, 1989.
[69] K. Tamada, M. Hara, H. Sasabe, and W. Knoll. Surface phase behavior of nalkanethiol self-assembled monolayers adsorbed on Au(111): an atomic force microscopy study. Langmuir, 13:1558–1566, 1997.
[70] G. Pace, A. Petijean, M.-N. Lalloz-Vogel, J. Harrowfield, J.-M. Lehn, and Samorı̀.
Subnanometer-resolved patterning of bicomponent self-assembled monolayers on
Au(111). Angewandte Chemie International Edition, 47:1–6, 2008.
[71] S. Chen, L. Li, C.L. Boozer, and S. Jiang. Controlled chemical and structural properties of mixed self-assembled monolayers of alkanethiols on Au(111). Langmuir,
16:9287–9293, 2000.
[72] Y. Okabe, U. Akiba, and M. Fujihira. Chemical force microscopy of -CH3 and
-COOH terminal groups in mixed self-assembled monolayers by pulsed-force-mode
atomic force microscopy. Applied Surface Science, 157:398–404, 2000.
[73] J.P. Folkers, P.E. Laibinis, G.M. Whitesides, and J. Deutch. Phase behavior of
two-component self-assembled monolayers of alkanethiolates on gold. The Journal
of Physical Chemistry, 98:563–571, 1994.
[74] L. Li, S. Chen, J. Zheng, B.D. Ratner, and S. Jiang. Protein adsorption on
oligo(ethylene glycol)-terminated alkanethiolate self-assembled monolayers: the
molecular basis for nonfouling behavior. Journal of Physical Chemistry B,
109:2934–2941, 2005.
[75] M.D. Porter, T.B. Bright, D.L. Allara, and C.E.D. Chidsey. Spontaneously organized molecular assemblies. 4. structural characterization of n-alkyl thiol monolayers on gold by optical ellipsometry, infrared spectroscopy, and electrochemistry.
Journal of the American Chemical Society, 109:3559–3568, 1987.
48
Bibliography
[76] Z. Yang, I. Engquist, B. Liedberg, and J.-M. Kauffmann. Electrochemical characterisation of mixed monolayer assemblies of thiol analogues of cholesterol and fatty
acids on gold. Journal of Electroanalytical Chemistry, 430:189–195, 1997.
[77] C.A. Widrig, C. Chung, and M.D. Porter. The electrochemical desorption of nalkanethiol monolayers from polycrystalline Au and Ag electrodes. Journal of Electroanalytical Chemistry, 310:335–359, 1991.
[78] E.W.L. Chan and M.N. Yousaf. Immobilization of ligands with precise control
of density to electroactive surfaces. Journal of the American Chemical Society,
128:15542–15546, 2006.
[79] M.K. Chaudhury and G.M. Whitesides. How to make water run uphill. Science,
256:1539–1541, 1992.
[80] B.S. Gallardo, V.K. Gupta, F.D. Eagerton, L.I. Jong, V.S. Craig, R.R. Shah, and
N.L. Abbott. Electrochemical principles for active control of liquids on submillimeter scales. Science, 283:57–60, 1999.
[81] M. Mrksich, J.R. Grunwell, and G.M. Whitesides. Biospecific adsorption of carbonic anhydrase to self-assembled monolayers of alkanethiolates that present benzenesulfonamide groups on gold. Journal of the American Chemical Society,
117:12009–12010, 1995.
[82] C.D. Tidwell, S.I. Ertel, B.D. Ratner, B.J. Tarasevich, S. Atre, and D.L. Allara.
Endothelial cell growth and protein adsorption on terminally functionalized, selfassembled monolayers of alkanethiolates on gold. Langmuir, 13:3404–3413, 1997.
[83] M.S. Kim, G. Khang, and H.B. Lee. Gradient polymer surfaces for biomedical
applications. Progress in Polymer Science, 33:138–164, 2008.
[84] S. Morgenthaler, C. Zink, and N.D. Spencer. Surface-chemical and -morphological
gradients. Soft Matter, 4:419–434, 2008.
[85] J. Genzer and R.R. Bhat. Surface-bound soft matter gradients. Langmuir, 24:2294–
2317, 2008.
[86] B. Liedberg and P. Tengvall. Molecular gradients of omega-substituted alkanethiols
on gold - preparation and characterization. Langmuir, 11(10):3821–3827, 1995.
[87] M. Riepl, M. Östblom, I. Lundström, S.C.T. Svensson, A.W. Deiner van der Gon,
M. Schäferling, and B. Liedberg. Molecular gradients: An efficient approach for
optimizing the surface properties of biomaterials and biochips. Langmuir, 21:1042–
1050, 2005.
[88] K.Y.C. Lee, J.F. Klingler, and H.M. Mcconnell. Electric-field-induced concentration gradients in lipid monolayers. Science, 263(5147):655–658, 1994.
[89] J.T. Groves, S.G. Boxer, and H.M. McConnell. Electric field-induced critical demixing in lipid bilayer membranes. Proceedings of the National Academy of Sciences
of The United States of America, 95(3):935–938, 1998.
[90] N.L. Jeon, S.K.W. Dertinger, D.T. Chiu, I.S. Choi, A.D. Stroock, and G.M. Whitesides. Generation of solution and surface gradients using microfluidic systems.
Langmuir, 16(22):8311–8316, 2000.
Bibliography
49
[91] S. Morgenthaler, S. Lee, S. Zucher, and N.D. Spencer. A simple, reproducible approach to the preparation of surface-chemical gradients. Langmuir, 19(25):10459–
10462, 2003.
[92] S.M. Morgenthaler, S. Lee, and N.D. Spencer. Submicrometer structure of surfacechemical gradients prepared by a two-step immersion method. Langmuir, 22:2706–
2711, 2006.
[93] R.H. Terrill, K.M. Balss, Y. Zhang, and P.W. Bohn. Dynamic monolayer gradients:
active spatiotemporal control of alkanethiol coatings on thin gold films. Journal of
the American Chemical Society, 122(5):988–989, 2000.
[94] K.M. Balss, B.D. Coleman, C.H. Lansford, R T. Haasch, and P.W. Bohn. Active spatiotemporal control of electrochemical reactions by coupling to in-plane
potential gradients. Journal of Physical Chemistry B, 105:8970–8978, 2001.
[95] S.T. Plummer and P.W. Bohn. Spatial dispersion in electrochemically generated surface composition gradients visualized with covalently bound fluorescent
nanospheres. Langmuir, 18(10):4142–4149, 2002.
[96] K.M. Balss, G.A. Fried, and P.W. Bohn. Chemically selective force mapping of
electrochemically generated two-component w-substituted alkanethiol monolayer
gradients by pulsed-force-mode atomic force microscopy. Journal of the Electrochemical Society, 149:C450–C455, 2002.
[97] K.M. Balss, T.C. Kuo, and P.W. Bohn. Direct chemical mapping of electrochemically generated spatial composition gradients on thin gold films with surfaceenhanced raman spectroscopy. Journal of Physical Chemistry B, 107(4):994–1000,
2003.
[98] S.T. Plummer, Q. Wang, P.W. Bohn, R. Stockton, and M.A. Schwartz. Electrochemically derived gradients of the extracellular matrix protein fibronectin on gold.
Langmuir, 19(18):7528–7536, 2003.
[99] Q. Wang and P.W. Bohn. Active spatiotemporal control of arg-gly-asp-containing
tetradecapeptide organomercaptans on gold with in-plane electrochemical potential
gradients. Journal of Physical Chemistry B, 107(46):12578–12584, 2003.
[100] B.D. Coleman, N. Finnegan, and P.W. Bohn. Sharply defined lateral composition
gradients of copper on gold by spatiotemporal control of the in-plane electrochemical potential distribution. Thin Solid Films, 467(1-2):121–126, 2004.
[101] B.D. Coleman, N. Finnegan, and P.W. Bohn. Factors influencing the interfacial
width of copper gradients on gold produced by spatiotemporal control of the inplane electrochemical potential distribution: electrode geometry and plating solution composition. Journal of Electroanalytical Chemistry, 571(2):139–148, 2004.
[102] X.J. Wang and P.W. Bohn. Anisotropic in-plane gradients of poly(acrylic acid)
formed by electropolymerization with spatiotemporal control of the electrochemical
potential. Journal of the American Chemical Society, 126(21):6825–6832, 2004.
[103] X.J. Wang, R.T. Haasch, and P.W. Bohn. Anisotropic hydrogel thickness gradient films derivatized to yield three-dimensional composite materials. Langmuir,
21(18):8452–8459, 2005.
50
Bibliography
[104] X.J. Wang, H.L. Tu, P.V. Braun, and P.W. Bohn. Length scale heterogeneity
in lateral gradients of poly(n-isopropylacrylamide) polymer brushes prepared by
surface-initiated atom transfer radical polymerization coupled with in-plane electrochemical potential gradients. Langmuir, 22(2):817–823, 2006.
[105] X. Wang and P.W. Bohn. Spatiotemporal coltrolled formation of two-component
counterpropagating lateral graft density gradients of mixed polymer brushes on
planar Au surfaces. Advanced Materials, 19:515–520, 2007.
[106] T. Sehayek, A. Vaskevich, and I. Rubinstein. Preparation of graded materials by
laterally controlled template synthesis. Journal of the American Chemical Society,
125(16):4718–4719, 2003.
[107] R. Yamada and H. Tada. Manipulation of droplets by dynamically controlled
wetting gradients. Langmuir, 21(10):4254–4256, 2005.
[108] E.L. Ratcliff and A.C. Hillier. Directed electrodeposition of polymer films using
spatially controllable electric field gradients. Langmuir, 23:9905–9910, 2007.
[109] M.F. Smiechowski and V.F. Lvovich. Electrochemical monitoring of watersurfactant interactions in industrial lubricants. Journal of Electroanalytical Chemistry, 534:171–180, 2002.
[110] M.F. Smiechowski and V.F. Lvovich. Characterization of non-aqueous dispersions
of carbon black nanoparticles by electrochemical impedance spectroscopy. Journal
of Electroanalytical Chemistry, 577(1):67–78, 2005.
[111] V.F. Lvovich and M.F. Smiechowski. Impedance characterization of industrial
lubricants. Electrochimica Acta, 51:1487–1496, 2006.
[112] K.N. Allahar, D.P. Butt, M.E. Orazem, H.A. Chin, G. Danko, W. Ogden, and
R.E. Yungk. Impedance of steels in new and degraded ester based lubricating oil.
Electrochimica Acta, 51:1497–1504, 2006.
[113] S.S. Wang and H.-S. Lee. The application of a.c. impedance technique for detecting
glycol contamination in engine oil. Sensors and Actuators B, 40:193–197, 1997.
[114] S.S. Wang, H.-S. Lee, and D.J. Smolenski. The developement of in situ electrochemical oil-condition sensors. Sensors and Actuators B, 17:179–185, 1994.
[115] H.-S. Lee, S.S. Wang, D.J. Smolenski, M.B. Viola, and E.E. Klusendorf. In situ
monitoring of high-temperature degraded engine oil condition with microsensors.
Sensors and Actuators B, 20:49–54, 1994.
[116] S.S. Wang and H.-S. Lee. An electrochemical sensor for distinguishing two-strokeengine oils. Sensors and Actuators B, 40:199–203, 1997.
[117] S.S. Wang. Road tests of oil condition sensor and sensing technique. Sensors and
Actuators B, 73:106–111, 2001.