Anda di halaman 1dari 241

ELECTRICAL IMPEDANCE

TOMOGRAPHY AT LOW
FREQUENCIES

by
Johan Andoyo Effendi Noor
Supervised by
Professor H. G. L. Coster
and
Dr. T. C. Chilcott
A thesis submitted for the degree of
DOCTOR OF PHYLOSOPHY
in the Faculty of Science
UNIVERSITY OF NEW SOUTH WALES
2007
ii
CERTIFICATE OF ORIGINALITY
I hereby declare that this submission is my own work and to the best of
my knowledge it contains no materials previously published or written by
another person, or substantial proportions of material which have been
accepted for the award of any other degree or diploma at UNSW or any
other educational institution, except where due acknowledgement is
made in the thesis. Any contribution made to the research by others, with
whom I have worked at UNSW or elsewhere, is explicitly acknowledged
in the thesis. I also declare that the intellectual content of this thesis is the
product of my own work, except to the extent that assistance from others
in the project's design and conception or in style, presentation and
linguistic expression is acknowledged.
Signed ..............
Date 4 September 2007
iii
COPYRIGHT STATEMENT
I hereby grant the University of New South Wales or its agents the right
to archive and to make available my thesis or dissertation in whole or
part in the University libraries in all forms of media, now or here after
known, subject to the provisions of the Copyright Act 1968. I retain all
proprietary rights, such as patent rights. I also retain the right to use in
future works (such as articles or books) all or part of this thesis or
dissertation.
I also authorise University Microfilms to use the 350 word abstract of my
thesis in Dissertation Abstract International (this is applicable to doctoral
theses only).
I have either used no substantial portions of copyright material in my
thesis or I have obtained permission to use copyright material; where
permission has not been granted I have applied/will apply for a partial
restriction of the digital copy of my thesis or dissertation.'
Signed ...........................
Date 4 September 2007
AUTHENTICITY STATEMENT
I certify that the Library deposit digital copy is a direct equivalent of the
final officially approved version of my thesis. No emendation of content
has occurred and if there are any minor variations in formatting, they are
the result of the conversion to digital format.
Signed ...........................
Date 4 September 2007
iv
All praise due to ALLAH, God Almighty, who has granted me time and
capability to finish this work.
I dedicate this thesis to:
my beloved mother who did not live long enough to see this thesis,
my beloved father for all his discipline guidance,
my beloved wife and children for their love and warm company during all
the hardship,
and to all my teachers who had taught and guided me to reach this
achievement.
v
ACKNOWLEDGEMENTS
There is nothing I can do more than to express my most sincere
gratitude to Professor Hans G.L. Coster, my teacher and supervisor for his
support through out this lengthy project without which it is impossible for
me to complete this study. His fatherly supervision and advice have so
enormously been invaluable and strengthened me during my hard time
within the study period.
I am also indebted to my co-supervisor, Dr. Terry C. Chilcott, for
his encouragement and patience in transferring his expertise in the
BULFIS. His bright ideas have opened my eyes toward the impedance
tomography technology.
A deep appreciation should also go to Dr. Marko Vauhkonen, of
the Department of Applied Physics, University of Kuopio, Kuopio, Finland
for supplying the EIDORS algorithms used in this project. I also would like
to extend my thanks to Ken Jackson, Pritipal Baweja and Ping Luo for their
help on all mechanical and electronic work. All other members of
Biophysics Department must also receive my appreciation for all the
friendship during my stay in the department. My thank must also go to all
my colleagues: Suhrawardi Ilyas for his help in duplicating the thesis,
Alamsyah Juwono, Dr. D.J. Jack Santjojo, Dr. Adi Susilo and Prof
Budiono Mismail in Brawijaya University for all the supportive and
encouraging discussions.
vi
I also would like to thank all my friends in KPII and ISOC UNSW
for the warm friendship that make my life easy.
My gratitude and thank also go to Ms Jo Ronalds, the AusAID
Liaison Officer, for her assistance, especially on obtaining my award
extension.
Finally, I would like to thank the Australian Agency for International
Development (AusAID) for financing my study.
vii
ABSTRACT
Most EIT machine operates at high frequencies above 10 kHz.
Biological systems demonstrate dispersions of electrical impedance
characteristics at very low frequencies below 2 kHz due to the presence of
membrane surrounding the cells and diffusion polarisation effects.
A study was made on the feasibility of the use of low frequencies
in a range of 1.12 Hz to 4.55 kHz in EIT. One high frequency of 77.712
kHz similar to that normally used in common EIT was also used as a
comparison. The impedance measurements employed a four-terminal
method using the BULFIS, an ultra low frequency impedance spectrometer
and used conducting and insulating material as the objects/phantoms.
The results show that the conductance and capacitance of a
system containing metal object disperses at frequency range of 0.1 10
kHz, which is consistent to the electrical properties of a double layer
forming at the metal-electrolyte interface similar to the electrical properties
of a membrane.
The reconstructed images reveal that at low frequencies the
conducting and the insulating bodies were indistinguishable. They appear
differently at high frequencies above 4.55 kHz indicating that the use of
multi frequency instrumentation in EIT covering the very low frequency
range provides information that instrumentation restricted to frequencies
above 10 kHz does not supply. While the internal structure of the double
layers could not be delineated, the presence of the double layers could be
viii
readily detected by the behaviour of the images as the frequency was
varied. This has potential for EIT because it might allow the detection of
structures from the variation of the images with frequency. This variation
with frequency does not occur at the higher frequencies more usually used
for EIT.
ix
LIST OF SYMBOLS
c Dielectric permittivity
0
c Dielectric permittivity of free space
r
c Relative permittivity of material
| Phase difference
Complex admittivity
k
-1
,
D

Debye length
Electrochemical potential, permeability
u Phase angle
Space charge density, electrical resistivity
o Electrical conductivity
t Small volume
e Angular frequency
O Body volume
dO Boundary volume
Electrostatic potential
Zeta potential
A Cross-sectional area, amplifier gain
a Chemical activity
B Magnetic induction
C,c Capacitance
c
i
Concentration of species i
^ Set of complex number
D Electric displacement
d Thickness
E Electric field
f Frequency
f
i
Chemical activity coefficient
x
G,g Conductance
H Magnetic field
I,i Electrical current
J Electric current density, Jacobian matrix
j Imaginary unit
k Boltzmans constant
L Number of electrodes
M Molecular weight
N
A
Avogadros number
n Ionic concentration, normal vector
p pressure
q Ionic charge
R Electrical resistance
\ Set of real number
r radius
T Temperature
U(x), u Potential distribution inside an object
V Vector of measured voltages on all the electrodes
,
k k
l l
U V
Voltage on the lth electrode when kth current pattern is used
V
Volume
w Electrical work
X Electrical reactance
Y Admittance
Z Impedance
z
i
Ion valency

xi
LIST OF FIGURES

page
Figure 1.1. A renogram showing four images of radioactivity in
kidneys at various times after injection of
radiopharmaceutical. Reprinted from Elliott [3]
2
Figure 1.2. Some images obtained using (a) CT technique (human
thorax), (b) MR technique (human thorax) and (c)
ultrasound technique (right internal jugular vein). In (c)
CCA is common carotid artery; IJV, internal jugular
vein; TH, thyroid lobe. ((a) and (b) reprinted from Gaeta
[5], (c) from Tan and Gibson [6]).
3
Figure 1.3. A schematic diagram of an electrical impedance
tomography system.
6
Figure 1.4. The schematic diagram of the impedance camera as
used by Henderson and Webster [10] to make images of
thorax (adopted from Webster [12]).
9
Figure 1.5. a) An EIT thoracic image showing the change in
resistivity of the lung during breathing. The electrodes
arrangement for the imaging procedures is shown in b)
(taken from Brown [17]).
9
Figure 2.1. Current flows through a tissue. At high frequencies, the
cell membrane is short circuited making it invisible to
the current. Therefore, current proceeds in straight lines
(thin lines). In the other hand, at low frequencies, the
membrane impedance is very high that direct the current
to flow only through the extracellular fluid (in thick
lines).
25
Figure 2.2. A Maxwell-Wagner (MW) element comprises a parallel
combination of a resistor R and a capacitor C.
26
Figure 2.3. A series combination of two Maxwell-Wagner (MW)
elements representing two homogeneous layers.
27
Figure 2.4. Dispersions of (A) impedance magnitude and (B)
impedance phase with frequency. The solid line is for a
single layer and the dashed line is for a double layer
system. The impedance magnitudes of both systems are
almost identical, whereas the phase angle shows slight
differences in the frequency range 10 1000 Hz
(adopted from Coster et al. [5]).
29
Figure 2.5. Dispersions of (A) capacitance and (B) conductance
with frequency of the similar systems depicted in Figure
2.4. The solid line is for a single layer and the dashed
line is for a double layer system. The behaviour of the
single layer shows no dependency on the frequency. In
the contrary, conductance and capacitance
characteristics of double-layer system show big
dispersions towards high frequencies (reprinted from
30
xii
Coster et al. [5]).
Figure 2.6. Electrical R-C equivalent of a cell system depicted in
Figure 2.1. It includes the extracellular fluid. R
m
is the
membrane resistance, R
i
the intracellular resistance, R
e

the extracellular resistance and C
m
the membrane
capacitance (adopted from Rabbat [25]).
32
Figure 2.7. The (a) four- and (b) three-component representation of
cell, with R
2
~ 2R
m
and C ~ C
m
/2
33
Figure 2.8. The illustration of the three dispersion regions (o, |, )
of relative permittivity over the frequency range found
in biological system (reprinted from Coster et al. [5]).
36
Figure 2.9. Two-terminal method for impedance measurement. The
potential difference is measured using the electrodes
used to inject current.
40
Figure 2.10. Impedance measurement four-terminal method. The
current is injected through one pair of electrodes and the
second pair measures the voltage. The polarisation
impedance is eliminated since no current flows through
the voltage electrodes which are connected to a high
input impedance amplifier.
40
Figure 2.11. Illustration of the electrode-electrolyte interface
showing the Stern layer and the Gouy-Chapman layer of
diffuse space charge which extends from the Stern layer
into the bulk solution. The Stern and Gouy-Chapman
layers will be discussed later in the text.
42
Figure 2.12. The Helmholtz model of double layer (taken from [30].
M denotes the metallic electrode and L denotes the
aqueous electrolyte.
52
Figure 2.13. The Gouy-Chapman model of double layer (taken from
[30]). M denotes the metallic electrode and L denotes
the aqueous electrolyte. The thickness of the layer
equals the Debye length (k
1
), where the potential at the
region is called the zeta potential, .
56
Figure 2.14. The schematic illustration of the structure of the double
layer according the Stern theory.
59
Figure 3.1.
Typical electrodes used in the four-terminal measurements, where
the current electrode is separated from the voltage sensing electrode
by insulating material. My system used type (a) (see Chapter 4).
69
Figure 3.2. a). The adjacent configuration for EIT data collection
method. In adjacent method, equipotential lines exist in
a homogeneous medium. Currents are injected via
neighbouring electrodes and voltages picked up through
other adjacent electrode couples; b) the opposite
method, where currents are injected through
diametrically opposite electrodes and the voltages are
measured with respect to one reference electrode set
next to current electrode (redrawn from Webster [28])

69
xiii
Figure 3.3. a). In the cross method, current electrodes are separated
by large dimensions. b) The multi-reference method
uses one electrode to collect currents injected
simultaneously from other electrodes (redrawn from
Webster [28]). There are current and voltage electrodes
at each electrode location.
71
Figure 3.4. The adaptive current method uses all electrodes to send
currents simultaneously to the object and measures the
voltages with respect to the reference (redrawn from
Webster [28]). There are current and voltage electrodes
at each electrode location.
71
Figure 3.5. A situation used to derive Neumann boundary
conditions. J
s1
and J
s2
are the current density outside and
inside the body O, respectively. E
1
and E
2
are the
corresponding electric fields. The size of the small
cylindrical volume t has been exaggerated.
77
Figure 3.6. The finite element mesh used in the reconstruction
calculations. The electrodes positions are marked with
the green elements.
82
Figure 3.7. The FWHM method to determine the width of a
Gaussian function [64]. The width of the function is
given by the distance of the curves legs at half its
maximum.
93
Figure 4.1. Schematic diagram of BULFIS as illustrated in Coster et
al. [3]
103
Figure 4.2. Photograph of the BULFIS sitting next to the PC. 104
Figure 4.3. Photograph of the experimental chamber. 107
Figure 4.4. Photograph of the experimental chamber with the
amplifier. They sat inside a Faraday cage to eliminate
external interferences.
108
Figure 4.5. Small segment of the electrodes arrangement in the
chamber. Two copper plates were coupled to form a
current electrode. The gap between two current
electrodes was 5 mm. The silver voltage electrodes were
located 1 mm from its left and right current electrodes.
The drawing is not to scale.
108
Figure 4.6. Electrodes numbering arrangement. The electrodes were
numbered counter clock wisely with electrode #1 is
located on the right hand side and electrode #9 on the
left hand side of the chamber (top view).
109
Figure 4.7. Sketch of the observed signal interference pattern. (a)
The signal picked from the signal generator, and (b) the
current flowing through the electrode.
111
Figure 4.8. The flowchart of EIDORS. 114
Figure 5.1. The theoretical plots of the (a) conductance and (b)
capacitance of a 100 mM NaCl solution (see text for
explanation). Please note that the frequency axis is on a
log scale.
121
xiv
Figure 5.2. The experimental chamber. (a) Top view and (b) side
view. The sliding V-electrode holder allowed the
electrodes to be positioned anywhere in the chamber,
while the I-electrodes were fixed on the wall. L was the
distance between the current electrodes and l the voltage
electrodes separation.
123
Figure 5.3. Curves of impedance magnitude (a) and its phase (b) as
a function of frequency for two positions of the voltage-
electrodes: centre and 35 mm off-centre. The chamber
was filled with 50 mM NaCl. No other objects were
present in the chamber.
127
Figure 5.4 Curves of impedance magnitude (a) and its phase (b)
similar to Figure 5.3 when the chamber was filled with
100 mM NaCl.
128
Figure 5.5. Curves of impedance magnitude (a) and its phase (b)
similar to Figure 5.3 when the chamber was filled with
150 mM NaCl.
129
Figure 5.6. SNRs of the measurements over the frequency range. (a)
for the 50 mM solution, (b) for the 100 mM solution and
(c) for the 150 mM solution.
130
Figure 5.7. Dispersions of conductance (a) and capacitance (b) with
frequency for two voltage-electrodes positions: centre
and 35 mm off-centre. The chamber was filled with 50
mM NaCl. No other objects were present in the
chamber.
133
Figure 5.8. Dispersions of conductance (a) and capacitance (b) with
frequency similar to those of Figure 5.7 when the
chamber was filled with 100 mM NaCl.
134
Figure 5.9. Dispersions of conductance (a) and capacitance (b) with
frequency similar to those of Figure 5.7 with the
chamber was filled with 150 mM NaCl.
135
Figure 5.10. Dispersions of the impedance magnitude (a) and its
phase (b) with frequency for two voltage-electrodes
positions at the centre and 35 mm off-centre with The
chamber was filled with 50 mM NaCl. The left plots are
for the case when a circular stainless steel block placed
in the centre of the chamber. and the plots on the right
are for the case when no object was present in the
solution. The solid lines in the graph are the average
values of the three spectra taken for each electrode
position.
137
Figure 5.11. Dispersion of conductance (a) and capacitance (b) with
frequency obtained from the data of Figure 5.10. The
left graphs are for the case when a metallic object was
present and the right ones are when the object was
absent. The insets on the capacitance graphs are
expanded graphs of the dispersions over the frequency
range above 100 Hz.
138
xv
Figure 5.12. Dispersion of conductance (a) and capacitance (b) with
frequency for a circular stainless steel block suspended
in the centre of the chamber filled with 100 mM NaCl.
The left graphs are when the metallic object was present
and the right ones are when the object was absent. The
insets on the capacitance graphs are expanded views of
the dispersions over the frequency range above 100 Hz.
141
Figure 5.13. Dispersion of conductance (a) and capacitance (b) with
frequency for a circular stainless steel block suspended
in the centre of the chamber filled with 150 mM NaCl.
The left graphs are when the metallic object was present
and the right ones are when the object was absent. The
insets on the capacitance graphs are expanded views of
the dispersions over the frequency range above 100 Hz.
142
Figure 5.14. Dispersion of conductance (a) and capacitance (b) with
frequency for a circular perspex block suspended in the
centre of the chamber filled with 50 mM NaCl. The left
graphs are when an insulating object was present and
the right ones are when the object was absent. The insets
on the capacitance graphs are expanded views of the
dispersions over the frequency range above 100 Hz.
143
Figure 5.15. Dispersion of conductance (a) and capacitance (b) with
frequency for a circular perspex block placed at the
centre of the chamber filled with 100 mM NaCl. The
left graphs are when an insulating object was present
and the right ones are when the object was absent. The
insets on the capacitance graphs are expanded views of
the dispersions over the frequency range above 100 Hz.
144
Figure 5.16. Dispersion of conductance (a) and capacitance (b) with
frequency for a circular perspex block placed at the
centre of the chamber filled with 150 mM NaCl. The
left graphs are when an insulating object was present
and the right ones are when the object was absent. The
insets on the capacitance graphs are expanded views of
the dispersions over the frequency range above 100 Hz.
145
Figure 5.17. Dispersions of the impedance magnitude (a) and its
phase angle (b) with frequency for a circular aluminium
block placed in the centre of a circular chamber filled
with different NaCl solutions (left column) and when no
object was present (right column).
148
Figure 5.18. Dispersions of the conductance (a) and capacitance (b)
with frequency for a circular aluminium block
suspended in different NaCl solutions for a circular
chamber (left column) and when no object was present
(right column).
149
Figure 5.19. Dispersions of the conductance (a) and capacitance (b)
with frequency for an empty beaker glass suspended in
different NaCl solutions for a circular chamber (left
150
xvi
column) and when no object was present (right column).
Figure 5.20. Comparisons of the dispersions of the conductance (left
column) and capacitance (right column) with frequency
for aluminium and perspex block placed in a circular
chamber filled with (a) 50 mM, (b) 100 mM and (c) 150
mM NaCl solutions.
153
Figure 5.21. In the four-terminal measurement technique, the
currents entering and leaving the electrode through
points of unequal polarisation impedance creates the
potential polarisation [6].
156
Figure 6.1. Simulation of reconstruction of admittivity distributions
for a circular insulating object (40 mm C) immersed in
electrolyte at 1.12 Hz. The conductivity of the saline
was 1.1 S/m and its relative permittivity was 78. The
conductivity of the object was 0.0 S/m and its relative
permittivity was 2.
163
Figure 6.2. FWHMs for the simulation of (a) conductivity and (b)
susceptivity at 1.12 Hz. The FWHM for both images is
33% of the diameter of the imaging region (see text for
details) compared with 13.3% for the original
distributions. The dotted lines represent the actual
admittivity properties of the system.
165
Figure 6.3. Reconstructed conductivity distribution (top row) and
dielectric permittivity (capacitance) distribution (bottom
row) of a circular aluminium block (C 130 mm) at the
centre in 50 mM NaCl. The frequencies used (from left
to right) were 1.12 Hz, 15.56 Hz, 282.38 Hz, 4551.5 Hz
and 77,712 Hz
167
Figure 6.4. Reconstructed conductivity distribution (top row) and
dielectric permittivity (capacitance) distribution (bottom
row) of a circular aluminium block (C 130 mm) in the
centre in 100 mM NaCl. The frequencies used (from left
to right) were 1.12 Hz, 15.56 Hz, 282.38 Hz, 4551.5 Hz
and 77,712 Hz.
168
Figure 6.5. Reconstructed conductivity distribution (top row) and
dielectric permittivity (capacitance) distribution (bottom
row) of a circular aluminium block (C 130 mm) in the
centre in 150 mM NaCl. The frequencies used (from left
to right) were 1.12 Hz, 15.56 Hz, 282.38 Hz, 4551.5 Hz
and 77,712 Hz

169
Figure 6.6. Reconstructed conductivity distribution (top row) and
dielectric permittivity (capacitance) distribution (bottom
row) of a perspex cylinder (C 100 mm) in the centre in
50 mM NaCl. The frequencies used (from left to right)
were 1.12 Hz, 15.56 Hz, 282.38 Hz, 4551.5 Hz and
77,712 Hz.

170
xvii
Figure 6.7. Reconstructed conductivity distribution (top row) and
dielectric permittivity (capacitance) distribution (bottom
row) of a perspex cylinder (C 100 mm) in the centre in
100 mM NaCl. The frequencies used (from left to right)
were 1.12 Hz, 15.56 Hz, 282.38 Hz, 4551.5 Hz and
77,712 Hz.
171
Figure 6.8. Reconstructed conductivity distribution (top row) and
dielectric permittivity (capacitance) distribution (bottom
row) of a perspex cylinder (C 100 mm) in the centre in
150 mM NaCl. The frequencies used (from left to right)
were 1.12 Hz, 15.56 Hz, 282.38 Hz, 4551.5 Hz and
77,712 Hz.
171
Figure 6.9. Variations with frequency of peak values of
conductivities of (a) a circular aluminium block (C 130
mm) and (b) a perspex cylinder (C 130 mm); and peak
values of dielectric permittivities (capacitances) of (c) a
circular aluminium block (C 130 mm) and (d) a perspex
cylinder (C 130 mm).
173
Figure 6.10. Conductivities (left) and dielectric permittivities
(capacitances) (right) distributions of a circular
aluminium block (C 130 mm) located in the centre of
the chamber with solutions of: (a) 50 mM NaCl, (b) 100
mM NaCl and (c) 150 mM NaCl.
175
Figure 6.11. Conductivities (left) and dielectric permittivities
(capacitances) (right) distributions of a perspex cylinder
(C 100 mm) located in the centre of the chamber with
solutions of: (a) 50 mM NaCl, (b) 100 mM NaCl and (c)
150 mM NaCl.
176
Figure 7.1. The position of the object (stainless steel tubing) in the
chamber located off the centre near electrode #1 (the left
side of the tubing was 20 mm from the centre of the
chamber). The system used 16 electrodes evenly spaced
in the periphery of the chamber.
185
Figure 7.2. A stainless steel tubing (C 76mm) located off-centre
near electrode #1. The top row is conductance
distributions and bottom row capacitance distributions.
186
Figure 7.3. Cross-sectional curves (cross-section 1: along the centre
from electrode #1 on the right to electrode #9 on the
left) of the off-centre metal object near electrode #1 in a
150 mM NaCl solution. The vertical dashed lines show
the actual position and size of the metal object.

188
Figure 7.4. Variation of peak values of (a) conductance and (b)
capacitance of the metal object taken from Figure 7.3.
188
Figure 7.5. The position of the object (an empty beaker glass) in the
chamber located off the centre very close to electrode
#9.

189
xviii
Figure 7.6. The reconstructed images of an empty glass beaker (C
70mm) located off-centre near electrode #9. The top
row is conductance distributions and bottom row
capacitance distributions.
189
Figure 7.7. Cross-sectional curves (cross-section 1: along the centre
from electrode #1 on the right to electrode #9 on the
left) of the off-centre glass object near electrode #9 in a
150 mM NaCl solution. The vertical dashed lines show
the actual position and size of the metal object.
191
Figure 7.8. Variation of peak values of (a) conductance and (b)
capacitance of the metal object taken from Figure 7.7.
191
Figure 7.9. Positions of two hollow stainless steel tubings in the
chamber. Identical materials were used with separation
of 40 mm.
192
Figure 7.10. Images of two identical stainless steel tubing (C 76mm)
located off-centre. The top row is conductance
distributions and bottom row capacitance distributions.
193
Figure 7.11. Positions of two beaker glasses (C 70 mm) in the
chamber. The objects are 40 mm symmetrically
separated.
195
Figure 7.12. Images of two beaker glasses located off-centre. The top
row is the conductance distributions and bottom row
capacitance distributions.
195
Figure 7.13. Location the two different objects in the chamber. A
beaker glass was positioned on the left hand side and a
stainless steel tubing on the right hand side. The objects
were separated by 40 mm from each other.
196
Figure 7.14. Images of two objects: a stainless steel tubing (C
76mm) located off-centre near electrode #1 and a beaker
glass (C 70mm) located off-centre near electrode #9.
The top row is conductance distributions and the bottom
row capacitance distributions.
197
Figure 7.15. Location the polyurethane coated and uncoated objects
in the chamber.
199
Figure 7.16. Images a polyurethane-coated stainless steel tubing
located off-centre near electrode #9 (left column of each
set) and an uncoated stainless steel tubing located off-
centre near electrode #1 (right column of each set). The
left hand side set is conductance distributions and the
right hand side set is capacitance distributions.
200
xix
LIST OF TABLES

page
Table 2.1 Resistivity values of mammalian tissues (taken from Barber
and Brown [17]).
20
Table 5.1 Trend of the values of the measured conductance for different
concentrations of the surrounding NaCl solution.
152
Table 6.1 FWHM result of the simulation compared to those of Ider et al.
[1] and Wexler et al. [2].
164
Table 6.2a FWHM for aluminium (C 130 mm) and perspex cylinder (C
100 mm) calculated from the conductance images.
177
Table 6.2b FWHM for aluminium (C 130 mm) and perspex cylinder (C
100 mm) calculated from the capacitance images.
177
Table 6.3a Percentage of size difference between the original and
reconstructed objects for aluminium (C 130 mm) and perspex
cylinder (C 100 mm) calculated from half maximum of the
conductance images.
177
Table 6.3b Percentage of size difference between the original and
reconstructed objects for aluminium (C 130 mm) and perspex
cylinder (C 100 mm) calculated from half maximum of the
capacitance images.
178
xx
TABLE OF CONTENTS


page
TITLE PAGE i
ORIGINALITY STATEMENT OF ii
COPYRIGHT STATEMENT iii
DEDICATION iv
ACKNOWLEDGEMENTS v
ABSTRACT vii
LIST OF SYMBOLS ix
LIST OF FIGURES xi
LIST OF TABLES xix
TABLE OF CONTENTS xx
CHAPTER 1 INTRODUCTION
1.1 Medical Imaging 1
1.2 Electrical Impedance Imaging 4
1.3 History of Electrical Impedance Imaging 7
1.4 Medical Applications 10
1.5 Objectives of Thesis 11
1.6 Thesis Lay Out 12
References 14
CHAPTER 2 IMPEDANCE SPECTROSCOPY
2.1 Introduction 19
2.2 Impedance Measurements 21
2.2.1 Complex Mathematics Theory of Impedance 21
2.2.2 Maxwell-Wagner Theory of Tissue 24
2.2.3 Impedance Arising from Immobilised Charges 34
and Their Effects The Dispersion
2.2.4 Two- and Four-Terminal Measurements 37
xxi
2.3 The Modelling of Metal-Electrolyte Interface 41
2.3.1 The Debye-Hckel Theory 43
2.3.2 The Helmholtz Theory of Double Layer 51
2.3.3 The Gouy-Chapman Theory of Diffuse Layer 52
2.3.4 The Stern Model 58
References 63
CHAPTER 3 ELECTRICAL IMPEDANCE TOMOGRAPHY (EIT)
3.1 Introduction 66
3.2 The Mathematical Model 73
3.2.1 Governing Equations 73
3.2.2 Boundary Conditions 76
3.3 Numerical Implementation/Reconstruction Methods and 81
Algorithms
3.3.1 Mesh Design 87
3.3.2 Forward Solution 88
3.3.3 Inverse Problem 89
3.4 Determinations of the Spatial Resolution of the 92
Reconstructed Images
References 94

CHAPTER 4 MATERIALS AND METHODS
4.1 Data Acquisition System and Methods 101
4.2 Experimental Solutions and Objects 105
4.3 Experimental Chambers 106
4.4 Reconstruction Software 111
References 115
CHAPTER 5 ELECTRICAL PROPERTIES OF THE EXPERIMENTAL
CHAMBER
5.1 Introduction 116
xxii
5.2 Materials and Methods 122
5.2.1 The Experimental Chambers 122
5.2.2 Coating the Silver Voltage Electrodes 123
5.2.3 Impedance Measurements 124
5.3 Electrical Properties of the Chamber Systems 125
5.3.1 The Square Chamber System 125
5.3.2 The Circular Chamber System 146
5.4 Discussions 154
5.5 Conclusions 159
References 160
CHAPTER 6 COMPLEX ELECTRICAL IMPEDANCE
TOMOGRAPHY
6.1 Introduction 161
6.2 Materials and Methods 162
6.3 Simulated Reconstructions 163
6.4 Reconstruction Images from Impedance Measurements 166
6.5 Estimation of the Spatial Resolution 174
6.6 Discussions 179
6.7 Conclusions 180
References 181

CHAPTER 7 RECONSTRUCTIONS OF COMPLEX OBJECT
CONFIGURATIONS
7.1 Introduction 183
7.2 Materials and Methods 184
7.3 Reconstructed Images of a Single Off-Centre Object 185
7.4 Reconstructed Images of Double Objects 192
7.4.1 Metal-Metal 192
7.4.2 Glass-Glass 194
7.4.3 Metal-Glass 196
7.4.4 Coated-Uncoated Metal 198
xxiii
7.5 Discussions 201
7.6 Conclusion 202


CHAPTER 8 GENERAL DISCUSSIONS, CONCLUSIONS AND
FUTURE DIRECTIONS
8.1 General Discussions 203
8.2 General Conclusions 206
8.3 Future Directions 206
References 208

APPENDIX A Reconstruction Source Codes

1
INTRODUCTION
1.1 Medical Imaging
One of the most important aspects in modern medicine is to obtain images
of the internal organs to be used in the diagnoses of diseases. The history of
medical imaging was initiated since the first human body image creation by
Wilhelm Conrad Rntgen in 1895 when he produced an image of his wifes hand
using the x-ray machine he invented. It was then revolutionised by its
employment in the Computerised Tomography technique, also known as CT-scan,
in the 70s introduced by Sir Godfrey Hounsfield [1], where body cross-sectional
images can be reconstructed with a very good resolution. The technique utilises
the properties of the x-ray beam, which are of short wavelength (0.1 to 1 ) and
not deflected by the object it passes through. A thin x-ray beam is rotated around
the body to be visualised and a diametrically opposing detector captures the
attenuated beam that emerges. The image is reconstructed from multiple
projections of the body.
The next technique, the Emission Computed Tomography (ECT), came
with the utilisation of radiopharmaceutical. Radioactive iodine, or other permitted
radionuclides, was administered to the patient and the count rate was measured
using a small collimated sodium iodide detector and maps of the count rate were
2
sketched. The automated process of this technique was invented by Benedict
Cassen et al. [2] in early 50s after he built his rectilinear scanner. ECT has
developed in two branches based on the type of radionuclide that is used [3].
Positron Emission Computed Tomography (PET) detects the two coincident 511-
keV annihilation photos from position emitters such as
11
C,
13
N,
15
O, and
18
F.
Single-Photon-Emission Computed Tomography (SPECT) detects gamma rays
emitted singly, and sequentially, by the radionuclide tracer. Images produced from
this method are given in Figure 1.1 showing good spatial resolution. The CT-scan
and ECT rely on ionizing radiation; therefore they carry a small but finite
potential detriment. This imposes a limit on their usage especially for pregnant
women and children.
Figure 1.1. A renogram showing four images of radioactivity in kidneys at
various times after injection of radiopharmaceutical. Reprinted from Elliott [3].
3
Other techniques use ultrasound (e.g. ultrasonography/USG) and magnetic
field and radio waves to access the inner workings of the body (e.g. magnetic
resonance imaging/MRI). The first technique was introduced by Ian Donald et al.
[4] in mid 50s. This technique is based on the sonar principle and employs sonic
beams having frequencies from 1 MHz to 20 MHz. The image is reconstructed
from the returning echoes of the sounds transmitted by an ultrasonic transducer.
Figure 1.2. Some images obtained using (a) CT technique (human thorax), (b) MR
technique (human thorax) and (c) ultrasound technique (right internal jugular vein). In
(c) CCA is common carotid artery; IJV, internal jugular vein; TH, thyroid lobe. ((a) and
(b) reprinted from Gaeta [5], (c) from Tan and Gibson [6]).
.
(a) (b)
(c)
4
MRI was pioneered by Paul C. Lauterbur [7] in the 1970s. This technique
uses magnetic properties of matter where in an applied magnetic field aligns the
magnetic dipole moments of the protons of hydrogen atoms that would otherwise
recess as a consequence of their spin angular momentum. The image is obtained
by displacing the equilibrium magnetisation vector with a radio frequency (RF)
pulse then observing the signal as the magnetisation vector returns to equilibrium.
Figure 1.2 shows images of a human organs produced by CT, MR and ultrasound
techniques with different quality. They demonstrate better spatial resolutions than
radiopharmaceutical technique does.
Although the above mentioned techniques carry some superiority in terms
of spatial resolutions, they also posses some disadvantages, except for USG: high
cost, invasive, and non-portable. One imaging method that offers low cost, non-
invasive, compact, and portable is the electrical impedance technique.
1.2 Electrical Impedance Imaging
Electrical Impedance Tomography (EIT) is a technique that employs
alternating electric currents I and voltages V for mapping the interior distribution
of the electrical properties, either real or complex, of the investigated media from
measurements over the surface. The main applications are in the fields of clinical
medicine and industrial process controls. Many medical problems could be easily
diagnosed by knowing the spatial distribution of electrical properties inside the
body.
5
The specific electrical properties of interest are the electrical conductivity
and permittivity. The electrical conductivity is a measure of the ease with which
electric charges passes through a material and is generally given by the real part of
admittance or reciprocal impedance, and the electrical permittivity is a measure
of how readily the charges within a material are stored and is given by the
imaginary part of the admittance divided by the angular frequency of the
alternating current. Conductivity, permittivity or both types of data have been
used to produce interior images of the objects.
In general, the EIT instrumentation injects small electrical alternating
currents (0.1 1.0 mA) into a system through a set of equidistantly spaced
electrodes over the surface. In its minimalist form the set is a pair of electrodes in
which case the instrument uses this pair to measure the voltage (magnitude and
phase) that develops between the electrodes (two-terminal measurements).
Alternatively a second pair of electrodes can be used to measure the voltage
response (four-terminal measurements). Either two- or four-terminal methods are
employed to acquire a set voltage responses over the surface. Measurements of the
voltage magnitude and its phase angle give information on the internal
distribution of electrical conductivity (o) and permittivity (c) in the body. The
acquired data is then transformed into images with the help of image
reconstruction algorithms. Every material possesses its own unique dielectric
property, as well as conductivity.
6
Figure 1.3. A schematic diagram of an electrical impedance tomography
system.
The general design of an EIT instrument is illustrated in Figure 1.3. It
comprises an electrode array attached to the surface of the body and which injects
currents and measures the voltages that develop across the object, a data
acquisition system (consisting of single or multiple current sources, one or more
voltmeters and a control unit) and a computing unit to reconstruct and display the
images. The use of high frequency alternating current (AC) avoids polarisation
effects that predominate when using direct current (DC) and diffusion polarisation
effects that manifest phase differences at low frequencies that are too large to be
attributed to dielectric permittivity phenomenon. Moreover, a DC current, when
applied to the human body, may generate electrostatic effects and heating near
skin-electrode contacts, which may damage the skin.
EIT imaging utilises a soft-field technique. It differs intrinsically from
the hard-field technique used, for instance, in the x-ray CT in which X-ray
beams are not deflected and only attenuated. This makes the spatial analysis of the
Current generator Control circuit
Voltmeters
Reconstruction
algorithm
Video
display unit
Object
Data acquisition
Computing unit
Electrode array
7
data a linear problem. On the other hand, the direction of the electrical current in
EIT depends on the conductivity and permittivity distributions in the object being
imaged. The voltage at a boundary point will depend on the electrical properties
of the whole body. This will give a non-linear relationship between the boundary
voltage measurements and the internal electrical parameters of the object.
Therefore, it is difficult to produce final reconstructed images of similar spatial
resolution to those reconstructed using CT or MRI. The spatial resolution of EIT
is also dependent of the number of possible independent measurements as well as
the quality (signal-to-noise ratio (SNR)) of the collected data [8].
Despite its limitations, EIT still offers benefits. It is compact, portable,
user friendly, non-invasive, non-ionising, low cost and fast, which attracts
researchers to develop such instrumentation. EIT has the potential to develop a
continuous bedside monitoring machine in intensive care units or in remote
medical facilities. Further EIT images yield properties of the tissue/organs that are
very different from X-ray adsorption properties. Thus, EIT can provide new
insights into bodily function not accessible to x-ray CT, despite its limitations in
spatial resolution.
1.3 History of the Electrical Impedance Imaging
EIT was first developed for use in medical applications by Henderson et
al. [9] in 1976 as an impedance camera for producing thoracic images. Henderson
and Webster [10] in 1978 used a frontal plane impedance camera to study
pulmonary oedema. The operation of the camera is shown in Figure 1.4. This
8
study revealed the difficulties of using a single-channel impedance measurement
system. The instrument used a 100 kHz voltage signal to drive 144 electrodes
positioned around the chest and measured the currents flowing in each electrode
sequentially. An image of conductivity revealed low conductivity regions
corresponding to the position of the lungs. This image was reconstructed
assuming that the current flowed in straight lines through the body. In the same
year 1978 Lytle and Dines [11] mapped the spatial conductance distribution
within the object by solving Laplaces equation and doing back projection
iterations from a set of voltage-current data.
Impedance Tomography was further pioneered by Barber et al. [13] and
Brown et al. [14] when they introduced their 16-electrode Applied Potential
Tomography (APT) machine to generate images of the human thorax. This is a
technique that initially assumes that the resistivity (reciprocal conductivity) is
uniform everywhere within the body. Voltages between all pairs of electrodes
were then calculated using the resistivities the values for which were modified
by the differences between the calculated and measured voltages. When the
measured and the calculated values differed by a certain factor for a certain pair of
electrodes, the calculated resistivity between these electrodes was multiplied by
that factor. Brown and Barber [15, 16] produced the first in vivo images of the
human thorax (see Figure 1.5).
9
Figure 1.4. The schematic diagram of the impedance camera as used by
Henderson and Webster [10] to make images of thorax (adopted from Webster
[12]).
(a) (b)
Figure 1.5. a) An EIT thoracic image showing the change in resistivity of the
lung during breathing. The electrodes arrangement for the imaging procedures is
shown in b) (taken from Brown [17]).
Voltage source Quantitative data
Computer processing
Current measurement
Multiplexer
D
Chest
Thorax
Back
10
Improvements in these techniques proceeded through the development of
hardware and software (i.e. the reconstruction algorithms) involving some 30
research groups worldwide. Gisser et al. [18] and Simske [19] of the Rensselaer
Polytechnic Institute introduced 32-electrode Adaptive Current Tomography
(ACT1), which operated by supplying a current pattern through all electrodes
which optimised the distinguishability of the object under investigation. The
image reconstruction used was the Newton One-Step Estimated Reconstruction
(NOSER) algorithm. The Rensselaer group is developing a third generation
instrumentation, called ACT3 [20], for use in real time 3-D breast imaging that
uses 30 kHz sinusoidal AC current. The Dartmouth group in the US specialises in
breast imaging. They employ a 32-electrode system working at multiple
frequencies from 10 kHz to 1 MHz [21, 22].
1.4 Medical Applications
Since its invention the impedance imaging technique has found its way to
more and more applications, especially in the medical area. Many possible
applications in medicine have been explored. The first system built was intended
to study the human thorax [13-16, 23]. Since then images of other parts of the
body have been produced and used for a number of clinical and research
applications [24].
The most valuable advantage of the EIT technique is its capability to
perform dynamic imaging, for example, of the respiratory system, gastric
activities and cardiac-related activities. Inhalation and exhalation of lungs produce
11
changes in the impedance distribution of the lung. Gastric processes can be
monitored as changes in acidity level similarly distinguishable as impedance
distributions [14, 25]. Eyuboglu et al. [26-28] used their EIT system to study
cardiac activities with a spatial resolution of 10% and distinguished thoracic
impedance variations associated with the cardiac cycle and correlated with an
electrocardiogram.
Another major clinical application of EIT is in human breast screening and
the detection of soft tissue lesions such as cysts and tumours. Any malignant cysts
and tumours have dielectric constants different from that of breast tissue [29, 30].
Further, the highly vascularized tissue wrapping the tumour has a low resistivity
contrasting with the high-resistivity avascular tissue of the tumour itself. The
presence of tumour could be identified by these differences.
EIT is also used in other clinical areas, such as cerebral hemodynamics
[31], limb plethysmography, pulmonary perfusion and ventilation [32-35], apnoea
monitoring, bladder volume monitoring, hyperthermia monitoring [36, 37], total
body water and body fat measurements [38], fracture healing monitoring and
brain imaging [39, 40].
1.5 Objectives of Thesis
The research project described in this thesis is aimed at investigating
complex AC impedance imaging at low frequencies using our Biophysics Ultra
Low Frequency Impedance Spectrometer (BULFIS). This spectrometer allows
12
four-terminal impedance measurements over a wide range of frequencies,
enabling a comparison of low and high frequency images.
Usually, EIT operates mainly at frequencies above 9.6 kHz (see e.g. Casas
et al. [41], Chauveau et al. [42], Li et al. [43] and Hampshire et al. [44]).
Although some instruments were capable of producing signals as low as 1 kHz
(e.g. Chauveau et al. [42]), studies have focused on frequencies >9.6 kHz. The
aim of this study was to explore frequencies <9.6 kHz for improved spatial
resolution and additional information in the phase differences between the
voltages and currents that may contribute to the resolution and have potential as a
parameter for therapeutic evaluations.
1.6 Thesis Lay-Out
This thesis comprises seven chapters with the first introduces Electrical
Impedance Tomography (EIT), its history and applications.
Chapter 2 outlines biological interactions with electricity and the basic
principles of impedance spectroscopy and measurement techniques. Some aspects
of the properties of metal-electrolyte interface are also discussed in this chapter.
The mathematical backgrounds of EIT and numerical analyses and
solutions are reviewed in Chapter 3, while Chapter 4 describes the materials and
experimental methods used in the project.
The measurements results were presented from Chapter 5, where the
measurements of the electrical properties of the chamber systems are reported.
13
Chapter 6 presents and discusses the reconstructed images of single object
positioned at the centre of the experimental chamber. And Chapter 7 presents and
discusses the reconstructed images of double objects (comprised of similar or
different materials) positioned off-centre of the experimental chamber.
Finally, Chapter 8 summarises the work and gives the future directions for
the project.
14
References
1. Bautz, W. and W. Kalender. (2005). Godfrey N. Hounsfield and his
influence on radiology. Radiologe, 45(4): 350-355.
2. Cassen, B., L. Curtis, and C.W. Reed. (1950). A sensitive directional
gamma-ray detector. Nucleonics: 78-81.
3. Elliott, A. (2005). Medical imaging. Nuclear Instruments and Methods in
Physics Research, A, 546(1): 1-13.
4. Donald, I., J. MacVicar, and T.G. Brown. (1958). Investigation of
abdominal masses by pulsed ultrasound. The Lancet, 271(7032): 1188-
1195.
5. Gaeta, M., S. Vinci, F. Minutoli, S. Mazziotti, G. Ascenti, I. Salamone, S.
Lamberto, and A. Blandino. (2002). CT and MRI findings of mucin-
containing tumors and pseudotumors of the thorax: pictorial review.
European Radiology, 12(1): 181-189.
6. Tan, P.L. and M. Gibson. (2006). Central venous catheters: The role of
radiology. Clinical Radiology, 61(1): 13-22.
7. Lauterbur, P.C. (1973). Image formation by induced local interactions:
Examples employing nuclear magnetic resonance. Nature, 242: 190-191.
8. Guardo, R., C. Boulay, B. Murray, and M. Bertrand. (1991). An
experimental study in electrical impedance tomography using
backprojection reconstruction. IEEE Transactions on Biomedical
Engineering, 38(7): 617-627.
9. Henderson, R.P., J.G. Webster, and D.K. Swanson. (1976). A thoracic
electrical impedance camera. Proceeding of the Annual Conference of
Engineering in Medicine and Biology, 18: 322.
10. Henderson, R.P. and J.G. Webster. (1978). An impedance camera for
spatially specific measurements of the thorax. IEEE Transactions on
Biomedical Engineering, 25(3): 250-254.
11. Lytle, R.J. and K.A. Dines. (1978). Impedance camera: a system for
determining the spatial variation of electrical conductivity. Lawrence
Livermore Lab.,Univ. California,Livermore,CA,USA.
15
12. Webster, J.G. (1990).Electrical Impedance Tomography. Bristol: IOP
Publishing Ltd.
13. Barber, D.C., B.H. Brown, and I.L. Freeston. (1983). Imaging spatial
distributions of resistivity using applied potential tomography. Electronics
Letters, 19: 933-935.
14. Brown, B.H., D.C. Barber, and A.D. Seagar. (1985). Applied potential
tomography: possible clinical applications. Clinical Physics and
Physiological Measurement, 6(2): 109-121.
15. Brown, B.H. and D.C. Barber. (1982). Applied Potential Tomography - A
new in vivo medical imaging technique. Proceeding of the Annual
Conference of Hospital Physicists' Association. Sheffield, UK.
16. Barber, D.C. and B.H. Brown. (1984). Applied potential tomography.
Journal of Physics E: Scientific Instrument, 17: 723-733.
17. Brown, B.H. (2001). Medical impedance tomography and process
impedance tomography: a brief review. Measurement Science and
Technology, 12(8): 991-996.
18. Gisser, D.G., D. Isaacson, and J.C. Newell. (1987). Current topics in
impedance imaging. Clinical Physics and Physiological Measurement,
8(Suppl. A): 39-46.
19. Simske, S.J. (1987). An adaptive current determination and one-step
reconstruction technique for a current tomography system, Masters
Thesis, in Graduate Faculty. Rensselaer Polytechnic Institute: Troy, New
York. p. 110.
20. Cook, R.D., G.J. Saulnier, D.G. Gisser, J.C. Goble, J.C. Newell, and D.
Isaacson. (1994). ACT3: a high-speed, high-precision electrical
impedance tomograph. IEEE Transactions on Biomedical Engineering,
41(8): 713-722.
21. Hartov, A., R.A. Mazzarese, F.R. Reiss, T.E. Kerner, K.S. Osterman, D.B.
Williams, and K.D. Paulsen. (2000). A multichannel continuously
selectable multifrequency electrical impedance spectroscopy. IEEE
Transaction on Biomedical Engineering, 47(1): 49-58.
16
22. Kerner, T.E., D.B. Williams, K.S. Osterman, F.R. Reiss, A. Hartov, and
K.D. Paulsen. (2000). Electrical impedance imaging at multiple
frequencies in phantoms. Journal of Physiological Measurement, 21(1):
66-77.
23. Barber, D.C. and B.H. Brown. (1989). Applied potential tomography.
Journal of British Interplanetary Society, 42(7): 391-393.
24. Dijkstra, A.M., B.H. Brown, A.D. Leathard, N.D. Harris, D.C. Barber, and
D.L. Edbrooke. (1993). Clinical applications of electrical impedance
tomography. Journal of Medical Engineering and Technology, 17(3): 89-
98.
25. Vaisman, N., N. Weintrob, A. Blumental, Z. Yosefsberg, and P. Vardi.
(1999). Gastric emptying in patients with type I diabetes mellitus. Annals
of the New York Academy of Sciences, 873: 506-511.
26. Eyuboglu, B.M., B.H. Brown, D.C. Barber, and A.D. Seagar. (1987).
Localisation of cardiac related impedance changes in the thorax. Clinical
Physics and Physiological Measurement, 8 Suppl A: 167-73.
27. Eyuboglu, B.M. and B.H. Brown. (1988). Methods of cardiac gating
applied potential tomography. Clinical Physics and Physiological
Measurement, 9 Suppl A: 43-8.
28. Eyuboglu, B.M., B.H. Brown, and D.C. Barber. (1998). Problems of
cardiac output determination from electrical impedance tomography
scans. Clinical Physics and Physiological Measurement, 9 Suppl A: 71-7.
29. Fricke, H. and S. Morse. (1926). The electric capacity of tumors of the
breast. Journal of Cancer Research: 340-376.
30. Blad, B., P. Wendel, M. Jnsson, and K. Lindstrm. (1999). An electrical
impedance index to distinguish between normal and cancerous tissues. J
Med Eng Technol, 23(2): 57-62.
31. Holder, D.S. (1992). Detection of cerebral ischaemia in the anaesthetised
rat by impedance measurement with scalp electrodes: implications for
non-invasive imaging of stroke by electrical impedance tomography.
Clinical Physics and Physiological Measurement, 13(1): 63-75.
17
32. Frerichs, I. (2000). Electrical impedance tomography (EIT) in applications
related to lung and ventilation: a review of experimental and clinical
activities. Physiological Measurement, 21(2): R1-21.
33. Frerichs, I., J. Hinz, P. Herrmann, G. Weisser, G. Hahn, T. Dudykevych,
M. Quintel, and G. Hellige. (2002). Detection of local lung air content by
electrical impedance tomography compared with electron beam CT.
Journal of Applied Physiology, 93(2): 660-666.
34. Adler, A., R. Guardo, and Y. Berthiaume. (1996). Impedance imaging of
lung ventilation: do we need to account for chest expansion? IEEE
Transaction on Biomedical Engineering, 43(4): 414-420.
35. Adler, A., R. Amyot, R. Guardo, J.H. Bates, and Y. Berthiaume. (1997).
Monitoring changes in lung air and liquid volumes with electrical
impedance tomography. Journal of Applied Physiology, 83(5): 1762-1767.
36. Conway, J. (1987). Electrical impedance tomography for thermal
monitoring of hyperthermia treatment: an assessment using in vitro and in
vivo measurements. Clinical Physics and Physiological Measurement, 8
Suppl A: 141-146.
37. Blad, B., L. Bertenstam, and B.R. Persson. (1990). Development of an
electrical impedance tomography system for noninvasive temperature
monitoring of hyperthermia treatments. Advances in Experimental
Medicine and Biology, 267: 235-243.
38. Baisch, F.J. (1994). Applied potential tomography shows differential
changes in fluid content of leg tissue layers in microgravity. Advances in
Space Research, 14(8): 359-364.
39. Holder, D.S. (1992). Electrical impedance tomography with cortical or
scalp electrodes during global cerebral ischaemia in the anaesthetised rat.
Clinical Investigator, 13(1): 87-98.
40. Holder, D.S. (1992). Electrical impedance tomography (EIT) of brain
function. Brain Topography, 5(2): 87-93.
41. Casas, O., J. Rosell, R. Bragos, A. Lozano, and P.J. Riu. (1996). A parallel
broadband real-time system for electrical impedance tomography.
Physiological Measurement, 17 Suppl 4A: A1-6.
18
42. Chauveau, N., B. Ayeva, B. Rigaud, and J.P. Morucci. (1996). A
multifrequency serial EIT system. Physiological Measurement, 17 Suppl
4A: A7-13.
43. Li, J.H., C. Joppek, and U. Faust. (1996). Fast EIT data acquisition system
with active electrodes and its application to cardiac imaging.
Physiological Measurement, 17 Suppl 4A: A25-32.
44. Hampshire, A.R., R.H. Smallwood, B.H. Brown, and R.A. Primhak.
(1995). Multifrequency and parametric EIT images of neonatal lungs.
Physiological Measurement, 16(3 Suppl A): A175-189.
2
IMPEDANCE SPECTROSCOPY
2.1 Introduction
Electrical impedance measurements have long been used to study
biological systems for investigating the fundamental electrical properties and
correlating these properties with tissue structure [1-7] as well as measuring
physiological events [7, 8]. Studies of the electrical properties of the tissue have
utilised the measurements of both real and reactive components of impedance
over a wide range of frequencies from a few mHz to more than 100 kHz [9] for
which the procedure is known as the Impedance Spectroscopy (IS) method. IS has
provided a non-invasive means of such characterisation procedures [5] and can be
carried out in situ, therefore IS can be applied in a wide range of biological
systems [10].
The earlier studies were carried out on various cell suspensions to identify
cell membrane properties [11-13]. It is significant that, in 1925, Fricke and Morse
[14] reported the first estimate of the thickness of a cell membrane giving a value
which was very close to that determined by Robertson [15] some decades later
using electron microscopy. However, due to instrumentation limitations and
safety reasons, physiological studies involving humans was commonly performed
20
at a single frequency above 20 kHz involving only measurements of impedance
magnitude [16].
Table 2.1. Resistivity values of mammalian tissues (taken from Barber and
Brown [17]).
Tissue Resistivity (Om) Conditions
Cerebrospinal fluid 0.65
Blood 1.5 50% haematocrit
Liver 3.5 In vivo perfused canine
tissue
Human arm 2.4 (longitudinal)
6.75 (transverse)
Skeletal muscle 1.25 (longitudinal)
18.00 (transverse)
Cardiac muscle 1.6 (longitudinal) Calculated interstitial
valve
4.24 (transverse) Calculated interstitial
valve
Neural tissue 5.8 Average results for
rabbit, cow and pig
Grey matter 2.84 Rabbit tissue
White matter 6.82 Rabbit tissue
Lung 7.2-23.6 Range from expiration to
inspiration
Fat 27.2 Average of many animal
results
Bone 166 Wet bovine bone
Cells are the main constituent of tissues. A cell comprises an outer
membrane, which separates the intracellular fluids and the cell nucleus from its
external environment. The intracellular fluids together with the nucleus fluids
make up about 95% of the tissue contents, which make the bulk tissue a good
electrical conductor. The membrane is composed of two layers of lipids into
which proteins are inserted. The lipid is an insulator, which makes the membrane
21
a poor electrical conductor. The composition of the biological tissues, which is
comprised of many cells and hence many cell membranes and biofluids has a
range of resistivity values, from 0.65 Om to 166 Om (see Table 2.1). This wide
range of resistivity allows electrical characterisation measurements to be made
with a good distinction between the body organs.
2.2 Impedance Measurements
2.2.1 Complex Mathematics Theory of Impedance
In general, impedance measurements are made by injecting an alternating
current (AC) of known angular frequency, e and small amplitude i
0
into the
system and measuring the amplitude v
0
and phase difference | of the response
voltage that develops across the system. The AC current has the form
0
sin( ) I I t e =

(2.1)
and the voltage response is
0
sin( ) V V t e | = +

(2.2)
Using the identities cos sin
j
e j
u
u u + and 1 j = (an imaginary unit),
the AC voltage and current can be expressed in phasor as
22
( ) ( )
0 0
and
j t j t
V V e I I e
e | e +
= =

(2.3)
The impedance, Z, can be written as a phasor, the magnitude and phase of which
are given by
0
0
j j Z
V V
Z e Z e
I I
| Z
= =

(2.4)
where
0
0
and
V
Z Z
I
| = Z = (2.5)
In Cartesian coordinates, the impedance is presented as a complex number
( )
0
0
cos sin
V
Z R jX j
I
| | = + = + (2.6)
where the real (Re(Z)) and the imaginary (Im(Z)) parts of the impedance describe
the resistance R and the reactance X respectively.
Please note that the time dependent term
j t
e
e
in each expression for the
AC voltage and current in Equation (2.3) cancel each other resulting in the time
independent relationship Equation (2.6).
It should be noted that impedance is also a specific form of the transfer
function of the system. Thus, if [ ] I s

and [ ] V s

denote Laplace transforms of the


sinusoidal functions of the AC current and electrical potential respectively, the
transfer function is
23
0
0
[ ]
TF( ) cos sin
[ ]
V V s s
s
I s I
| |
e
| |
= = +
|
\ .

(2.7)
where s = je is the Laplace complex variable. The solution to Equation (2.7)
yields the frequency domain definition of impedance similar to Equation (2.6)
( )
0
0
( ) TF( ) cos sin
V
Z j j
I
e e | | = + (2.8)
In the case of a homogeneous material, the system may be represented by
a conductance and a capacitance in parallel, thus
1
( ) Z
G j C
e
e
=
+
(2.9)
The parameters G and C describe the ability of the homogeneous material to
conduct and store electric charge respectively. For a piece of material of cross-
sectional area A and thickness x, these parameters are given by
and
A A
G C
x x
o c = = (2.10)
where o and c are the electrical conductivity and dielectric permittivity of the
material respectively. Thus, it can be deduced for such system that
24
1 1
cos and sin G C
Z Z
| |
e
= = (2.11)
In the case of applied DC signal, e.g. e = 0, the frequency dependent
reactance term in Equation (2.9) is zero making impedance equals to conductance
(reciprocal resistance).
2.2.2 Maxwell-Wagner Model of Tissue
For many years many researchers have shown a great interest in electrical
phenomena within biological tissue as well as interactions between biological
cells and electrical fields. Many investigations have been made since the late
1800s [16]. In the 1920s Fricke and Morse [14, 18] and Cole [19] worked on the
impedance of biological suspensions.
Basically, a tissue is made up of cells that are held together by an
extracellular matrix through which extracellular fluids flow. Figure 2.1 shows the
basic structure of organ tissue. Each cell is enveloped by membrane which
contains the intracellular fluids. When an electrical current is applied to the tissue
its flow through the cellular path will depend on its frequency. At low frequencies,
the membrane impedance is very high. This directs the current through the
surrounding fluids of low resistivity. At high frequencies the membrane
impedance is dominated by its capacitance, in the sense that the reactive term is
very low and shunts the membrane resistivity. In this instance the current flows
through the cells. This can be understood when one considers the equivalent
electrical R-C circuit depicted in Figure 2.6.
25
Figure 2.1. Current flows through a tissue. At high frequencies, the cell
membrane is short circuited making it invisible to the current. Therefore,
current proceeds in straight lines (thin lines). In the other hand, at low
frequencies, the membrane impedance is very high that direct the current to
flow only through the extracellular fluid (in thick lines).
The impedance measurements are best interpreted by using electrical
circuit theory. Figure 2.2 shows the simplest electric circuit of a parallel
combination of a resistor and a capacitor which represent the impedance of a
system. This configuration is known as a Maxwell-Wagner (MW) element. The
impedance of this circuit is usually expressed in terms of conductive, G, and
capacitive, C, elements in parallel, thus
1 1
( ) , where Z G
G j C R
e
e
= =
+
(2.12)
26
Figure 2.2. A Maxwell-Wagner (MW) element comprises a parallel
combination of a resistor R and a capacitor C.
In EIT applications Equation (2.8) applies for frequencies greater than 10kHz. It
can still apply to some biological systems for frequencies as low as 100Hz (e.g.
see Coster and Chilcott [20]).
For two homogeneous systems sandwiched in one, the equivalent electric
circuit is two MW elements in series (Figure 2.3), and the combinational
impedance is
1 1 2 2
1 1
( ) Z
G j C G j C
e
e e
= +
+ +
(2.13)
where
1
1
1
G
R
= and
2
2
1
G
R
= .
R C
27
Figure 2.3. A series combination of two Maxwell-Wagner (MW) elements
representing two homogeneous layers.
Tissues are made up of layers of cells with different conductance and
dielectric properties and, therefore, the Maxwell-Wagner model suits this tissue
representation. The Maxwell-Wagner model proves to be a very useful model to
describe impedance behaviour of biological [13, 21] and synthetic [22, 23]
membranes.
Given k layers of cells with individual conductance, G
1
, G
2
, , G
k
, and
capacitances, C
1
, C
2
, , C
k
, respectively. The total impedance of this model is
given by
1
1 1
( )
( ) ( )
k
i
i i
Z
G j C G j C
e
e e e e
=
=
+ +
_
(2.14)
where
1
i
i
G
R
= and G(e) and C(e) are the total conductance and capacitance at
frequency e. Each individual frequency-independent impedance dispersion
described by Equation (2.12) superimposes to form a single frequency-dependent
C
1
R
1 R
2
C
2
28
dispersion described by Equation (2.14) [5], each of which predominates near its
respective characteristic natural frequency of
i
i
i
G
C
e (2.15)
If the magnitude of the individual impedance dispersions are large, each
impedance dispersion can be identified within the combined impedance
dispersion. In many cases the dispersions of the impedance magnitude of a single
and two layer systems are barely indistinguishable as illustrated in Figure 2.4a. In
order to detect the presence of the second layer the dispersion of the impedance
phase must be measured (Figure 2.4b). Once the phase is known it is then
possible, through mathematical algorithms [24], to separate the conductance G(e)
and capacitance C(e) components [5, 6].
Figure 2.5 shows the conductance-frequency and capacitance-frequency
dependencies that are given by (derived from Equation (2.14) for k = 2)
2 2 2
1 2 1 2 1 2 2 1
2 2 2
1 2 1 2
( ) ( )
( )
( ) ( )
GG G G C G C G
G
G G C C
e
e
e
+ + +
=
+ + +
(2.16)
and
2 2 2
1 2 1 2 1 2 2 1
2 2 2
1 2 1 2
( ) ( )
( )
( ) ( )
C C C C C G C G
C
G G C C
e
e
e
+ + +
=
+ + +
(2.17)
29
Figure 2.4. Dispersions of (A) impedance magnitude and (B) impedance phase
with frequency. The solid line is for a single layer and the dashed line is for a
double layer system. The impedance magnitudes of both systems are almost
identical, whereas the phase angle shows slight differences in the frequency
range 10 1000 Hz (adopted from Coster et al. [5]).
30
Figure 2.5. Dispersions of (A) capacitance and (B) conductance with
frequency of the similar systems depicted in Figure 2.4. The solid line is for a
single layer and the dashed line is for a double layer system. The behaviour of
the single layer shows no dependency on the frequency. In the contrary,
conductance and capacitance characteristics of double-layer system show big
dispersions towards high frequencies (reprinted from Coster et al. [5]).
31
The subscripts 1 and 2 identify conductance and capacitance
components of the two layers. It is clear from Figure 2.5a that at the low
frequency limit (e 0) the capacitance value is maximum at
2 2
1 2 2 1
( 0) 2
1 2
( )
C G C G
C
G G
e
+
=
+
(2.18)
The capacitance decreases with increasing frequency and at the high
frequency limit (e ) approaches a minimum of
1 2
( )
1 2
C C
C
C C
e
=
+
(2.19)
i.e. a series combination of the two capacitances. It becomes obvious that adding
more layers will introduce further dispersion when the capacitance will decrease
to a value that is the series combinations of all capacitances.
In contrast, the conductance has a minimum value at very low frequencies
and maximum value at very high frequencies given by
1 2
( 0)
1 2
GG
G
G G
e
=
+
(2.20)
and
2 2
1 2 2 1
( ) 2
1 2
( )
C G C G
G
C C
e
+
=
+
(2.21)
respectively. Note that for a single element the conductance and capacitance are
independent of frequency. The introduction of the second element (layer)
32
produces dispersions of the conductance and capacitance with frequency. The
introduction of more elements (layers) increases the magnitude of the dispersions
and broadens it. To some extent, this explains the phenomenon of current flows
through layers of cells described earlier in Figure 2.1.
Figure 2.6. Electrical R-C equivalent of a cell system depicted in Figure 2.1. It
includes the extracellular fluid. R
m
is the membrane resistance, R
i
the
intracellular resistance, R
e
the extracellular resistance and C
m
the membrane
capacitance (adopted from Rabbat [25]).
In a similar fashion, Figure 2.6 provides a model to better understand the
electrical characteristics of a cell and its dependency on frequency. An ultra thin,
high resistive membrane of approximately 6 nm thickness surrounds the cell. It is
modelled electrically as a parallel combination of a resistor (R
m
) and a capacitor
C
m
R
i
R
e
R
m
R
m
C
m
33
(C
m
). The resistors R
i
and R
e
represent the intracellular the extracellular fluids
respectively. By combining the two membrane circuits in series, this model can be
represented by a four-component circuit (Figure 2.7a) which, with R
i
< R
m
[17],
can be simplified to a three-component circuit with R
2
~ 2R
m
and C ~ C
m
/2 [25] as
depicted in Figure 2.7b.
Figure 2.7. The (a) four- and (b) three-component representation of cell, with
R
2
~ 2R
m
and C ~ C
m
/2.
R
e
C
m
/2
(a) (b)
R
i
2R
m
C
R
2
R
1
34
2.2.3 Impedance Arising from Immobilised Charges and Their
Effects The o Dispersion
In addition to the impedance dispersions with frequency that can be
attributed to geometrical, conduction and dielectric properties of biological
systems and which can be modelled in a planar geometry using the Maxwell-
Wagner model are those that originate from structures containing immobilised
charges, e.g. membranes, and the processes that these structures support. These
additional dispersions manifest at substantially lower frequencies and are most
pronounced in measurements of the capacitance that are usually expressed in units
of relative permittivity. There are three dispersion regions of relative permittivity
in biological tissue, labelled o-, |- and -dispersions in the permittivity-frequency
plane [5, 26] (see Figure 2.8). The Nernst-Planck-Poisson electrodiffusion
equations for ions in an electrolyte containing two univalent species of opposite
charge can be used to explain the dispersions mechanism.
In an electrolyte system containing two univalent species of opposite
charge, the AC current density J at a location x is prescribed by the Nernst-Planck
equation
2 2
( ) ( ) ec
| | c c | |
= + + + + +
| |
c c
\ .
\ .

p n q D q D
J qD p n E P N j E
x x kT kT
(2.22)
where P and N represent the steady state concentration of positive and negative
ions respectively, p and n represent the AC perturbed concentration of positive
and negative ions respectively; E denotes the steady state electric field and E

35
denotes the AC perturbation of the field. In the presence of fixed or immobilised
positive and negative charges F

, the steady state and AC forms of Poissons


equation are
( ) and ( )
c c

c c
= =
c c

E q E q
N P F n p
x x
(2.23)
The first two term of Equation (2.22) gives rise to the diffusion effects
labelled as the o-dispersion. It occurs as a result of the frequency dependence of
the relative magnitudes of the AC diffusion and AC field-driven currents in the
AC Nernst-Planck equation (i.e. the magnitudes of the first two terms in Equation
(2.22) relative to the last two terms). For instance, at low frequencies the ionic
concentrations and gradients are created by the AC electric field making the
diffusion current larger than the field-driven current. This condition is similar to
those describing impedances of very large capacitors in which the reactive
component of the impedance decreases with increasing frequency. In some cases,
the capacitance of cell membrane at very low frequencies can have values many
folds larger than the dielectric capacitance and can also have negative values [27].
The negative values can also be interpreted as inductances, which occur as a result
of electric current loops (standing waves) in the electrolyte system [10]. However,
the capacitance will disperse towards the dielectric values at higher frequencies.
36
Figure 2.8. The illustration of the three dispersion regions (o, |, ) of relative
permittivity over the frequency range found in biological system (reprinted
from Coster et al. [5]).
The |-dispersion is attributable to the interfacial polarisation between the
phases (e.g. the bulk extracellular fluids and the membrane) that must satisfy the
boundary condition
bulk bulk memb memb
E E c c =

(2.24)
which with the application of the Nernst-Planck-Poisson Equations in a planar
(essentially one-dimensional) geometry and assuming a sufficiently high
frequency that renders the first two terms in Equation (2.8) negligible compared to
the last two terms, yield the Maxwell-Wagner model. The |-dispersion commonly
occurs in the frequency range of 10
2
-10
6
Hz [5, 6] which encompasses the range
commonly used in EIT.
37
The -dispersion occurs due to molecular dipole re-orientation in an AC field. The
net polarisation effect induced by dipole moment rotation will be reduced by
Brownian motion [5]. This yields the decreasing dielectric permittivity with the
increasing the frequency [6].
2.2.4 Two- and Four-Terminal Measurements
To measure the impedance of a system a pair of electrodes needs to be
attached to it. When metal electrodes are utilised, as is the common practice, a
metal-electrolyte interface is formed. In the case of EIT the electrolyte may be an
electrolyte paste or solution intentionally introduced. Complex electrochemical
phenomena can occur at the interface. For example, in the case of silver or copper
electrodes ions react with the electrolyte, usually with the chloride ions. For silver
electrodes, AgCl forms on the surface of the electrode and facilitates the exchange
of electric charge between the silver (electrons) and the electrolyte (chloride ions).
The Ag/AgCl electrode therefore provides a low ohmic contact with electrolytes.
A similar reaction applies to copper electrodes, except in this instance Cu
2
Cl
2
dissociates more readily into electrolyte reducing the effectiveness of the
electrode and its long-term viability.
In the case of an inert metal, e.g. gold electrodes, electrostatic forces
arising from the vastly differing dielectric properties of the metal and electrolyte
attract ions to the metal surface but they cannot react with metal. These forces
result in a charge distribution at the interface which resembles a parallel plate
capacitor and is known as the electric double layer. The charge distribution
38
manifests an electrical potential between the two phases. A current injected into
the system changes the charge distribution, resulting in a polarisation
impedance, whose magnitude is comparable to that of a biological bilayer [21].
The potential due to the polarisation impedance is dependent upon the metal-
electrolyte combination, current density and frequency of the injected current. The
polarisation impedance is also known as the contact impedance. Further detail is
given in Section 2.3.
The potential developed when a metal electrode is in contact with
electrolyte can be measured with respect to a standard (in this case the electrolyte)
and is referred to as a half-cell potential. When two identical metal electrodes are
used both electrodes will have the same half-cell potential with respect to the
same electrolyte whence the net potential measured should be zero. However, in
practice this is not the case since the conditions are most often not identical due to
surface irregularities or contaminants. The net potential differences, also known as
the DC offset or electrode potential, can vary from V to hundreds of mV
depending on the material [16].
In AC applications DC offset theoretically should not cause any problems
since the instrumentations can be AC-coupled. Nevertheless, the offset potential
tends to be unstable resulting in transient changes which obscure AC analysis.
Fortunately, this effect can be reduced by electroplating the electrode. By this
method, electrodes with very small and stable offset potential can be
manufactured. The silver-silver chloride (Ag/AgCl) electrode is the most familiar
example for this purpose.
39
Electrical impedance measurements are generally carried out by injecting a
known small amplitude AC current into the system under investigation and the
electrical potential response of the system including the phase difference between
the injected current and the potential are measured. Depending on the technique
being used the electrical potential differences (amplitude and phase with respect to
those of the current) developing across the system are either measured using the
same electrodes as those to inject the current, known as a two-terminal (2T)
measurement, or separate voltage electrodes, known as a four-terminal (4T)
measurement.
In 2T method (see Figure 2.9), the voltage measurements are carried out
using the same electrodes used to inject current. Although electrode polarisation
can be reduced by electroplating the electrode, it still is a problem at frequencies
below 1 kHz and a serious problem below 20 Hz [16] as at these frequencies the
electrode impedance, referred to as the electrode polarisation impedance, can be
many folds the sample impedance and dependent upon the electrochemistry of the
metal-body interface, current density and frequency [28, 29].
The 4T method employs four electrodes (Figure 2.10). One pair of
electrodes is used to supply current to the object and the other pair is used to
measure the potential difference that develops across the sample. The input
impedance of the amplifier is typically very high, say 10
12
O, which makes
currents flowing through the voltage-sensing electrodes considerably small
whence the electrical potential that is generated across the electrolyte-object
interface, even if the impedance of the interface is large, is similarly small and can
be neglected. With electroplating, the electrode polarisation impedance is
40
substantially reduced and its effect on measurements of the voltage response can
be eliminated entirely. Therefore, the impedance measured using the 4T method is
essentially the impedance of the sample.
Figure 2.9. Two-terminal method for impedance measurement. The potential
difference is measured using the electrodes used to inject current.
Figure 2.10. Impedance measurement four-terminal method. The current is
injected through one pair of electrodes and the second pair measures the
voltage. The polarisation impedance is eliminated since no current flows
through the voltage electrodes which are connected to a high input impedance
amplifier.
v
i
~
v
i
~
41
In practice, 4T technique reveals other parasitic impedances. For example,
Gaedt [10] has shown in his work in our laboratory that the size, shape, location
and orientation of the voltage-sensing electrodes influences the magnitude of the
phenomenological capacitance measurements at low frequencies. Best results are
obtained when the voltage electrodes are small, pointed and located as far as
possible from the current electrodes (see Chapter 5).
2.3 The Modelling of Metal-Electrolyte Interface
The origin of the electrical impedance properties of metal-electrolyte
interface has attracted many researchers since the early 1850s. The descriptions
of the origin of impedances given in Section 2.2.2 assumed that the layers
comprising the systems were represented by MW circuit elements describing
homogeneous electrical properties in one-dimension. More complex circuits are
required for multi-dimensional systems or Non-Maxwell-Wagner (NMW)
systems. The metal-electrolyte interface is an example of a multi-dimensional
system albeit on a micro-scale.
A one-dimensional representation of the interface is shown in Figure 2.11.
The absorbance of ions in the electrolyte to the metal-electrolyte interface is
balanced by the absorbance of electron charge carriers in the metal. If the metal
electrode is negatively charged, positively charged ions in the solution are
absorbed to the metal surface and negatively charged ions are repelled. But, the
positively charged ions also tend to diffuse randomly back into the solution due to
their thermal energy, creating a diffuse layer near the surface. The unequal
42
distributions of the two oppositely charged carriers at the surface and in the
diffuse layer together establish the electrical double layer.
Figure 2.11. Illustration of the electrode-electrolyte interface showing the
Stern layer and the Gouy-Chapman layer of diffuse space charge which extends
from the Stern layer into the bulk solution. The Stern and Gouy-Chapman
layers will be discussed later in the text.
Two models explain the electric double layer: the Gouy-Chapman model
and the Stern model, which draw upon the Debye-Hckel theory.
43
2.3.1 The Debye-Hckel Theory
The electrochemical potential,
i
, of an ion type i in an electrolyte solution
is defined by
0
ln
i i i i
kT a pV z q = + + + + (2.25)
where
0
i
is the chemical potential of the standard state,
k is the Boltzmanns constant,
T is the temperature,
a
i
is the chemical activity of the i
th
ion,
p is the pressure,
V is the volume,
z
i
is the ion valency, including the charge sign,
q is the ionic charge,
+ is the electrostatic potential.
The chemical activity of the i
th
ion, a
i
, is related to the ion concentration
by
i i i
a f c = (2.26)
where f
i
is the activity coefficient and c
i
is its concentration.
44
For an ideal condition where the concentration of the solution is ultra low
(infinite dilution), the distance between ions is far enough to neglect their
interactive electrostatic energy and the activity coefficient becomes unity for all
concentrations. Therefore, kT ln f
i
= 0 and Equation (2.25) becomes
0
(ideal)
ln
i i i i
kT c pV z q = + + + + (2.27)
remembering that ln ln( ) ln ln
i i i i i
a f c f c = = .
The movement (or the activity) of the ions is governed by both the
electrostatic forces between ions in the electrolyte solution and the thermal force.
Increasing the concentration leads to a decrease in the movement of ions since the
inter-ion distance becomes closer and results in the activity coefficient no longer
being unity. Under the influence of these forces, the ions are distributed unevenly
with the density exponentially decreasing with the distance from the metal
surface.
The Debye-Hckel theory assumes that an ion is a point charge surrounded
by a uniform ionic cloud. This ion also is assumed to continuously interact with
other ions in the vicinity. Consider a point whose electric potential is +, then the
electrical work, w, required in bringing in an ion of valency z from infinity to that
point is prescribed by
and w z q w z q
+ +
= + = + (2.28)
45
for positive and negative ions, respectively. According to the Boltzmann principle,
if the concentrations of those positive and negative ions at a very distant point
where the electric potential is zero are
0
n

and
0
n

, respectively, then at
equilibrium the average concentrations per unit volume of positive and negative
ions, n
+
and n

, respectively given by
/ 0 z q kT
n n e
+
+
+ +
= (2.29)
and
/ 0 z q kT
n n e

+

= (2.30)
at a point where the potential is +. This clearly suggests that a cloud of negative
ions predominates around the positive ions. The reverse condition holds for
negative ions.
The space charge density () at any point is then given by
n z q n z q
+ +
=
( ) ( )
/ / 0 0 z q kT z q kT
qz n e qz n e
+
+ +
+ +
= (2.31)
If both co-ions and counter ions have the same valency, i.e. z
+
= z

=z, and
the equilibrium concentrations are also equal (
0 0
n n n

= = ) then
( )
/ / zq kT zq kT
nzq e e
+ +
= (2.32)
46
Assuming that zq+/kT < 1, then the exponential terms can be replaced
by a series in which the second and higher order terms can be neglected yielding
1 1
q q
nzq
kT kT

+ + | |
~
|
\ .
2
2nzq
kT
1
~ (2.33)
If the ions that are present are of different species and valencies, i.e. z
+
=
z

, then Equation (2.31) becomes


2
2
i i i i
i i
q
q n z n z
kT
~ +
_ _
(2.34)
The first term in the right hand side of Equation (2.34) is equal to zero
since the solution is neutral (the concentrations of the co- and counter ions are
equal, therefore the sum over all available species equals zero). Therefore, we
have
2
2
i i
i
q
n z
kT
= +
_
(2.35)
This charge density satisfies the Poissons equation
47
2
0

cc
V + = (2.36)
where
2
\ is the Laplacian operator
2 2 2
2 2 2
x y z
c c c
= + +
c c c
,
c is the dielectric permittivity of the solution and
c
0
is the permittivity of free space.
Substituting Equation (2.35) into Equation (2.36) yields
2
2 2 2
0
i i
i
q
n z
kT
k
cc
V + = + = +
_
(2.37)
where
2
2
0
i i
i
q
n z
kT
k
cc
=
_
. (2.38)
Taking geometry of the ion as spherical, Equation (2.37) becomes
2 2 2
2
1
r
r r r
k
c c+ | |
V + = = +
|
c c
\ .
(2.39)
since radial symmetry yields 0
| u
c+ c+
= =
c c
.
In order to solve this equation we have to consider the boundary
conditions that require + 0 as r to yield
48
r
e
A
r
k
+ = (2.40)
The constant A is deduced using the Debye-Hckel assumption that the ion
is a point charge. In this case the potential at the centre of the ion (i.e. r 0) in
the absence of any other ions is given by
ion
0
i
z q
r cc
+ = (2.41)
Expanding Equation (2.40) gives a series
2
1
1 ( ) ...
2!
A
r r
r
k k
| |
+ = +
|
\ .
(2.42)
and taking r 0 we can take only the first term of the series to give,
ion
A
r
+ = (2.43)
Hence, from Equations (2.41) and (2.43) we have
0
i
z q
A
cc
= (2.44)
49
Substituting Equation (2.44) back into Equation (2.40) gives
0
r i
z q
e
r
k
cc

+ = (2.45)
where in fact + is the average potential at a point a distance r from the centre of
the ion which includes contributions from the ionic cloud, +
cloud
, and the point
charge at the centre of the ion, +
ion
. So,
cloud ion
+ = + + + (2.46)
and
cloud ion
+ = + +
0 0
r i i
z q z q
e
r r
k
cc cc

=
( )
0
1
r i
z q
e
r
k
cc

= (2.47)
For a dilute solution kr <1 applies and the exponential term 1
r
e r
L
L

~
reduces (2.47) to
cloud 1
0
i
z q
cc k

+ = (2.48)
Thus, the potential + is given by
50
1 1
0 0 0
1 1
i i i
z q z q z q
r r cc cc k cc k

| |
+ = =
|
\ .
(2.49)
The negative sign of the second term on the right hand side shows that the
ionic cloud has opposite charge to the ion charge. Here, while r is the distance
from the centre of the ion (or the radius in term of the spherical coordinate), the
quantity k
1
can be regarded as the effective radius of the surrounding ionic cloud.
Therefore, k
1
is considered as a characteristic length of the system and is known
as the Debye length,
D
=k
1
.
Looking at the interaction between a central point charge ion, qz
i
and its
surrounding ionic atmosphere with a potential
cloud
0
i
z qk
cc
+ = , the interaction
energy is defined by
0
1
2
i
i
z q
z q
k

cc
| |
A =
|
\ .
2 2
0
2
i
z q k
cc
= (2.50)
The quantity A is the difference in the electrochemical potential energy
associated to the addition or removal of an ion from the bulk real and ideal
solutions. Thus,
ln
i
kT f A = (2.51)
51
or
2 2
0
ln
2
i
i
z q
kT f
k
cc
= (2.52)
2 2
0
ln
2
i
i
z q
f
kT
k
cc
= (2.53)
This allows us to calculate the chemical activity coefficient at very low
concentrations.
2.3.2 Helmholtz Theory of Double Layer
In 1853 Helmholtz proposed that the double layer could be regarded as an
electrical parallel plate condenser with one plate being the surface of the metal
and the other plate a plane of ions in the solution which are attracted to the metal
surface by the resulting electrostatic force. Electroneutrality dictates that the
number of ions attracted to the surface must be exactly equal to the number of
surface charges. The thickness of the double layer l, therefore, is equal to the ionic
radius r
i
(see Figure 2.12). The Helmholtz theory, however, ignores the
dependencies of the double layer on the solution concentration and temperature.
52
Figure 2.12. The Helmholtz model of double layer (taken from [30]). M
denotes the metallic electrode and L denotes the aqueous electrolyte.
2.3.3 The Gouy-Chapman Theory of Diffuse-Layer
Gouy [31] in 1910 and Chapman [32] in 1913 proposed a model that deals
with the diffuse part of the double layer (see Figure 2.13). The model assumes that
the ions in the diffuse part are regarded as volume-less point charges distributed
according to Boltzmann distribution and the dielectric permittivity of the bulk
aqueous part is uniform everywhere. This model also neglects the discrete ionic
nature of the double layer (surface and diffuse parts). Instead, it approaches the
system as a continuous charge distribution. Assuming also that the surface is
infinite, flat and homogeneously charged (it is true since the curvature radius of
the electrode is, in practice, much larger than the Debye length of the system) then
the relation between the potential at any distance (x, y, z) from the surface and the
local charge density satisfies the Poissons equation
2
0

cc
V + = (2.54)
x l = r
i
+
0
53
where + is the potential at location (x, y, z) from the surface,
is the local charge density (coulombs per unit volume),
c is the dielectric permittivity of the solution and
c
0
is the permittivity of free space.
According to the Boltzmann distribution the ionic concentration is given
by
/ 0
i
z q kT
i i
n n e
+
= (2.55)
where
0
i
n is the concentration of ion species i in the bulk solution distant from the
surface where + = 0.
For z z z

= = and
0 0
n n n

= = , the charge density at a point where the
potential energy is + is given by
( ) zq n n
+
=
( )
/ / zq kT zq kT
zqn e e
+ +
=
2 sinh
zq
zqn
kT
+ | |
=
|
\ .
(2.56)
54
Since we are only interested in the flat surface and the potential variation
is in the normal co-ordinate to it, then substituting Equation (2.56) into Equation
(2.54) gives the Boltzmann-Poisson equation
2
2
0
2
sinh
zqn zq
x kT cc
c + + | |
=
|
c
\ .
(2.57)
To solve Equation (2.57) let
zq
y
kT
+
= ,
0
0
zq
y
kT
+
= ,
1/ 2
2 2
0
2z q n
kT
k
cc
| |
=
|
\ .
and
x k = . Equation (2.57) becomes
2
2
sinh( )
d y
y
d
= (2.58)
Taking into account two boundary conditions: 1) + = +
0
, (y = y
0
), when x
= 0 and 2) + 0,
d
dx
+
0 when x and integrating Equation (2.58) gives
2cosh 2 2sinh
2
dy y
y
d
| |
= =
|
\ .
(2.59)
Integrating once again and rearranging yields
( ) ( )
( ) ( )
0 0
0 0
/ 2 / 2
/ 2
/ 2 / 2
1 1
1 1
y y
y
y y
e e e
e
e e e

+ +
=
+
(2.60)
55
1
2ln
1
e
y
e

| | +
=
|

\ .
(2.61)
where
0
0
/ 2
/ 2
1
1
y
y
e
e

=
+
(2.62)
Using the original variables we have the solution of Equation (2.57) as
2 1
ln
1
x
x
kT e
qz e
k
k

| | +
+ =
|

\ .
(2.63)
where
0
0
/
/
1
1
zq kT
zq kT
e
e

+
+

=
+
(2.64)
and
1/ 2
2 2
0
2z q n
kT
k
cc
| |
=
|
\ .
(2.65)
If the surface potential is much less than the thermal energy,
0
1
2
zq
kT
+
<
(say, for instance,
kT
zq
= 25.6 mV at 25
o
C), the Debye-Hckel approximation
(using a series expansion) gives
0
x
e
k
+ = + (2.66)
confirming that at low surface potentials the potential falls exponentially with the
distance away from the charged surface boundary. At a location close to the
56
surface where the potential is high, however, the potential is falling at a greater
than exponential rate until the distance is equal to the Debye length,
D
, where it
decreases by an exponential factor (see Figure 2.13).
Figure 2.13. The Gouy-Chapman model of double layer (taken from [30]). M denotes
the metallic electrode and L denotes the aqueous electrolyte. The thickness of the
layer equals the Debye length (k
1
), where the potential at the region is called the zeta
potential, .
The surface charge density, o
0
, can be determined by taking the condition
that the system is electro-neutral, thus
2
0 0 2
0 0
d
dx dx
dx
o cc

+
= =
) )
0
0 x
d
dx
cc
=
+ | |
=
|
\ .
(2.67)
x = k
1
+
0
57
Applying the Poisson-Boltzmann distribution and integrating using the
second boundary condition, i.e. + 0,
d
dx
+
0 for x we reach an
expression
( )
1/ 2
0
0 0
8 sinh
2
zq
n kT
kT
T FF
1
1

=


( )
(2.68)
which at low potentials,
0
1
2
zq
kT
+
< , reduces to
0 0 0
o cc k = + (2.69)
showing that the capacitance of the diffuse part is equal to the capacitance of a
parallel plate with an effective separation of k
1
(i.e. the Debye length,
D
).
For a dilute aqueous solution of symmetrical electrolyte at 25
o
C (298 K)
1/ 2 1/ 2
1 0 0
2 2 3 2 2
2 2 10
D
A
kT kT
nz q N cz q
cc cc
k

| | | |
= = =
| |

\ . \ .
0.304
z c
| |
~
|
\ .
nm (2.70)
where N
A
is the Avogadro number and c = 10
3
nN
A
is the concentration of the
solution in mol/dm
3
.
58
Thus, the Gouy-Chapman approach to the double layer is valid only for
dilute solution where the diffuse layer must exceed the dimension of the ions. For
a concentrated solution where the ions are tightly packed the distribution of the
ions close to the charged surface does not satisfy the Boltzmann-Poisson
distribution. The discrepancies between the experimental and theoretical values
discovered by Grahame [33] are attributed to the assumption made in the Gouy-
Chapman model assumes that the ions in the aqueous phase are point charges.
2.3.4 The Stern Model
The difficulty encountered using the Gouy-Chapman model to describe the
double layer arises from the assumption that ions are point charges whence they
can approach the surface charge without limitation. The Stern model treats the
ions to be entities with volume of finite size so that they can approach the charged
surface to a distance as closest as its hydrated radius without being specifically
adsorbed. The Stern model considers the diffuse portion as two separate layers.
The first one is a layer closely attached to the charged surface with the thickness
of o (usually in the order of 0.1 nm) and identified as the Stern layer. The rest of
the diffuse part measured from o outward is specified as the Gouy layer, where the
Gouy-Chapman theory governs the distribution of the ions. Therefore, the Stern
potential +
o
is used as the inter-layer potential of the diffuse layer (instead of the
potential on the surface wall, +
0
), see Figure 2.14.
59
Figure 2.14. The schematic illustration of the structure of the double layer
according the Stern theory.
x
C
h
a
r
g
e

d
e
n
s
i
t
y
P
o
t
e
n
t
i
a
l
Stern
layer
+
+
+
+
+
+
-
Gouy-Chapman
layer
E
l
e
c
t
r
o
d
e
bulk
-
+
-
-
-
-
-
-
-
-
-
-
-
-
-
+
0
+
o
o
1
o
2
o
1/k
60
The decay in the electric potential in this region follows the sinh
(hyperbolic sin) function from +
o
to zero, where somewhere on the curve very
close to x = o lies the change of potential from the Stern layer to the bulk diffuse
region known as the zeta potential, . Please note that zeta potential, is not the
Stern potential +
o
, although they are often assumed to be similar.
The distribution of the adsorbed counter ions in the Stern layer is given by
Langmuir-type adsorption. The adsorption energy includes the electrical work,
qz+
o
, and the energy of adsorption of the counter ions, |. The surface charge
density of the Stern layer, o
1
, is given by
1 1
n zq o = (2.71)
where n
1
is the number of adsorbed ions per unit area.
Let N
1
be the number of adsorption sites per unit area, then the number of
unoccupied sites per unit area is N
1
n
1
. The number of available positions in the
solution is taken to be N
A
/M. Applying the Boltzmanns distribution gives
( ) /
1 1 1
/
q kT
A
n N n
e
n N M
o
| + +

= (2.72)
where N
A
is Avogadros number and M is the molecular weight of the solvent.
61
Combining Equations (2.71) and (2.72) we obtain the Stern equation
( )
1
1
1 /
zq
kT
A
N zq
N Mn e
o
|
o
+ +

=
+
(2.73)
The charge density within the whole solution is determined using the
Gauss law
( )
0
0
'
o
c c
o
o
= + + (2.74)
where c' is the dielectric permittivity of the medium between the surface and the
absorbed layer.
The charge density in the Gouy layer is given by
( )
1/ 2
2 0
8 sinh
2
zq
n kT
kT
o
o cc
+ | |
=
|
\ .
(2.75)
It is Equation (2.68) with +
0
replaced by +
o
.
From the requirement for the electro-neutrality of the double layer that
1 2
o o o = + (2.76)
the unknown quantities of o, o
1
, o
2
and +
o
can be calculated for a given +
0
.
62
For the case of dilute solution, where N
A
>Mn, the Stern equation can be
simplified to
1
1
zq
kT
A
N zqMn
e
N
o
|
o
+ +
= (2.77)
which is more commonly used than in the form of Equation (2.73).
Combining Equations (2.74), (2.75), (2.76) and (2.77) and maintaining the
electrical neutrality we have
( ) ( )
1/ 2
0 1
0 0
'
8 sinh 0
2
zq
kT
A
zq N zqMn
e n kT
N kT
o
|
o
o
c c
cc
o
+ +
+ | |
+ + =
|
\ .
(2.78)
The concentration difference between the Stern layer and the bulk (o
1
- o
2
)
is higher for dilute solution than that for more concentrated solution, therefore, +
o
will be larger. In other words, in more concentrated solutions the potential drop in
the Stern layer is larger than what occurs in dilute solution and the Gouy (diffuse)
layer is compressed. Thus, for the case of dilute solution the Stern model
approaches the Gouy model.
63
References
1. Coster, H.G.L. and U. Zimmermann. (1975). The mechanism of electrical
breakdown in the membranes of Valonia utricularis. Journal of Membrane
Biology, 22: 73-90.
2. Smith, J.R. (1977). Electrical characteristics of biological membranes in
different environments. PhD Thesis in Faculty of Science. University of
New South Wales. Sydney.
3. Ashcroft, R.G., H.G.L. Coster, and J.R. Smith. (1981). The molecular
organization of bimolecular lipid membranes. The dielectric structure of
the hydrophilic/hydrophobic interface. Biochimica et Biophysica Acta,
643: 191-204.
4. Karolis, C. (1993). The dielectric characterization of lipid bilayers. PhD
Thesis in Faculty of Science. University of New South Wales. Sydney.
5. Coster, H.G.L., T.C. Chilcott, and A.C.F. Coster. (1996). Impedance
spectroscopy of interfaces, membranes and ultrastructures.
Bioelectrochemistry and Bioenergetics, 40: 79-98.
6. Coster, H.G.L. and T.C. Chilcott. (1999). The characterization of
membranes and membrane surfaces using impedance spectroscopy, in
Surface chemistry and electrochemistry of membranes, Vol. 79, T.S.
Sorensen (Ed.). Marcel Dekker, Inc. New York. pp. 749-792.
7. Chilcott, T.C. and H.G.L. Coster. (1999). Electrical bioimpedance
methods: Electrical impedance tomography study of biological processes
in a single cell. Annals of the New York Academy of Sciences, 873: 269-
286.
8. Falk, F. and P. Fatt. (1964). Linear electrical properties of striated muscle
fibres observed with intracellular electrodes. Proceeding of the Royal
Society of London. Series B. Biological Sciences, 159: 606-651.
9. Schwan, H.P. (1992). Linear and nonlinear electrode polarization and
biological materials. Annals of Biomedical Engineering, 20: 269-288.
10. Gaedt, L.P. (1999). Impedance spectroscopy characterisation of synthetic
membranes and processes. PhD Thesis in Faculty of Engineering.
Univesity of New South Wales. Sydney, Australia.
64
11. Coster, H.G.L. (1969). Role of pH in the punch-through effect in the
electrical characteristics of Chara australis [Chara corallina]. Australian
Journal of Biological Sciences, 22(2): 365-74.
12. Coster, H.G.L. (1972). Role of pH in the punch-through effect in frog skin.
Biophysical Journal, 12(4): 447-448.
13. Chilcott, T.C. (1988). Admittance tomography of cells of Chara corallina:
a study of the electrical spatial structures associated with photosynthesis.
PhD Thesis in Faculty of Science. University of New South Wales.
Sydney, Australia.
14. Fricke, H. and S. Morse. (1925). The electric resistance and capacity of
blood for frequencies between 800 and 4.5 million cycles. Journal of
General Physiology, 9: 153-167.
15. Robertson, J.D. (1960). The molecular structure and contact relationships
of cell membrane. Progress in Biophysics, 10: 343-418.
16. Ackmann, J.J. and M.A. Seitz. (1984). Methods of complex impedance
measurements in biologic tissue. CRC Critical Reviews in Biomedical
Engineering, 11(4): 281-311.
17. Barber, D.C. and B.H. Brown. (1984). Applied potential tomography.
Journal of Physics E: Scientific Instrument, 17: 723-733.
18. Fricke, H. (1923). The electric capacity of cell suspensions. Physical
Review, 21: 708-709.
19. Cole, K.S. (1928). Electrical impedance of suspensions of Arbacia eggs.
Journal of General Physiology, 12: 37-54.
20. Coster, H.G.L. and T.C. Chilcott. (1999). The characterization of
membranes and membrane surfaces using impedance spectroscopy.
Surface Chemistry and Electrochemistry of Membranes, Ed. T.S.
Sorensen. Vol. 79. Marcel Dekker, Inc. New York. pp. 749-792.
21. Coster, H.G.L. and J.R. Smith. (1977). Low frequency impedance of Chara
corallina: simultaneous measurements of the separate plasmalemma and
tonoplast capacitance and conductance. Australian Journal of Plant
Physiology, 4: 667-764.
65
22. Hanai, T., K. Zhao, K. Asaka, and K. Asami. (1991). Dielectric theory of
concentration polarization, relaxation of capacitance and conductance for
electrolyte solutions with locally varying conductivity. Journal of
Membrane Science, 64: 153-161.
23. Coster, H.G.L., K.J. Kim, K. Dahlan, J.R. Smith, and C.J.D. Fell. (1992).
Characterization of ultrafiltration membranes by impedance spectroscopy.
I. Determination of the separate electrical parameters and porosity of the
skin and sublayers. Journal of Membrane Science, 66(1): 19-26.
24. Levy, E.C. (1959). Complex curve fitting. IEEE Transactions on
Automatic Control, 4: 37-43.
25. Rabbat, A. (1990). Tissue resistivity. The Adam Hilger series on
biomedical engineering, Ed. J.G. Webster. IOP Publishing Ltd. Bristol and
New York. pp. 8-20.
26. Schwan, H.P. (1957). Electrical properties of tissues and cell suspensions,
in Advances in Biological and Medical Physics, Vol. 5. Academic Press.
New York. pp. 147-209.
27. Chilcott, T.C. and H.G.L. Coster. (1989). Double fixed charge membrane:
a model for proteins in the cell membrane of Chara, in Plant Membrane
Transport, J. Dainty and E. Maree (Eds.). Elsevier. Amsterdam.
28. Geddes, L.A. (1972). Electrodes and the measurement of bioelectric
events. Wiley-Interscience. New York.
29. Ferris, C.D. (1974). Introduction to bioelectrodes. Plenum Press. New
York.
30. Antropov, L.I. (1972). Theoretical electrochemistry. Mir Publishers.
Moscow.
31. Gouy, G. (1910). Sur la constitution de la charge electrique a la surface
d'un electrolyte. J. Phys. Radium, 9: 457-468.
32. Chapman, D.L. (1913). A contribution to the theory of electrocapillarity.
Philosophical Magazine, 25: 475-481.
33. Grahame, D.C. (1947). The electrical double layer and the theory of
electrocapillarity. Chemical Reviews, 41: 441-501.
3
ELECTRICAL IMPEDANCE
TOMOGRAPHY (EIT)
3.1 Introduction
In general, electrical impedance tomographic technique utilises an array of
sensors located on the periphery of the object under investigation to capture
voltage signals emerging from the object arising from a current injected into the
object from a certain location (see Figure 1.3 in Chapter 1). The procedures are
repeated for different current electrode locations until the whole body has been
covered. The collected data contain information of the internal body structure
beneath the sensing zone. By means of image reconstruction algorithms the data
are then converted into useful images. There are three types of tomographic
imaging: static, multi-frequency, and dynamic imaging [1], where their
employments are dependent upon the objects to be investigated.
Static imaging technique is common for use in imaging static (slow
changing) organs (e.g. tumours and bone). This technique generates images from
a distribution of absolute impedances at certain, usually one or two, fixed
frequencies. The choice of measurement frequencies is application-dependent
because some different tissues cannot be discriminated when using two fixed
frequencies [2]; and Cherepenin et al. [3] have experienced that the reconstruction
67
of static images is difficult due to the uncertainty of the body shape and electrode
spacing. This limitation, together with low spatial resolution, has restricted the
implementations of EIT in clinical practice.
Increasing evidence that dispersions in the electrical property spectrum of
tissues below 10 MHz provide useful diagnostics for tissue pathology and/or
pathophysiology have led to the development of multi-frequency systems
operating in a range of frequencies that can potentially overcome the limitations
of the static imaging technique. For example, Hartov et al. developed a system
operating in a frequency range of 1 MHz [4] and later in frequency range of 10
MHz [5]. However, the complexity associated with designing a multichannel
system has limited widespread development of these ideas.
The Hartov et al. systems operate in the static spatial domain, whereas
some diagnostics require real time imaging that requires a dynamic technique.
Dynamic imaging creates images from variations in impedance in real time. This
enables the recording of immediate changes of organ physiological events, such as
respiratory [6] and cardiac activities [7]. For example, Hahn et al. [8] has
performed dynamic simulations on a simple resistor phantom of the human thorax
system; and Kim et al. [9, 10] has monitored abruptly changing resistivity
distributions on-line arising from movements of a plastic rod in the phantom from
one side to the other. However, as mentioned previously, dynamic images
sacrifice spatial resolution necessary in clinical application of EIT requiring early
detection of, for example, small tumours whose growth rate is slow.
Basically, there are two modes of measurements in EIT, the I- and V-
modes. The hardware systems that operated in I-mode [4, 11-13] inject a known
68
sinusoidal current of constant amplitude into the body and measure the magnitude
and phase of voltage differences that result. The hardware systems operating in V-
mode [4, 14, 15] applies a known sinusoidal voltage of constant amplitude to the
system across pairs of driving electrodes and measures the magnitude and phase
of the resulting currents through other pairs of electrode. In both modes, the
system applies and/or measures voltages or currents between any combinations of
electrodes either simultaneously or sequentially [16].
Unfortunately, systems that employ two-terminal techniques in either the
I-mode or V-mode, cannot carry out an ideal measurement at an active electrode
because of the voltage drop across the contact impedance [17, 18]. For example,
in the I-mode the measured voltage is inaccurate due to the contribution of an
unknown voltage drop across the electrode-skin contact layer. Inaccuracies also
arise for similar reasons using the V-mode [17, 19].
Some four-terminal measurement strategies have been introduced, e.g.
adjacent [20-22], opposite [23], cross [24], multi-reference [25] and adaptive [26,
27] methods. The V- and I-electrodes arrangements are shown in Figure 3.1. The
adjacent method (see Figure 3.2a), also known as the complete electrode method,
is the most widely used. This method was also used through out the course of my
thesis (see Chapter 4).
69
(a) (b) (c)
Figure 3.1. Typical electrodes used in the four-terminal measurements, where
the current electrode is separated from the voltage sensing electrode by
insulating material. My system used type (a) (see Chapter 4).
(a) (b)
Figure 3.2. a). The adjacent configuration for EIT data collection method. In
adjacent method, equipotential lines exist in a homogeneous medium. Currents
are injected via neighbouring electrodes and voltages picked up through other
adjacent electrode couples; b) the opposite method, where currents are injected
through diametrically opposite electrodes and the voltages are measured with
respect to one reference electrode set next to current electrode (redrawn from
Webster [28]).
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
I
V
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
I-electrode
V-electrode
insulator
70
In the adjacent method, the current is injected through two neighbouring
(one next to the other) electrodes and the voltage differences are measured using
all other adjacent pairs. This is repeated for all electrodes. From the 16 voltage
measurements, however, there are only 13 independent measurements (i.e.
measurements that do not involve current injecting electrodes). Therefore, for
instance, for a 16-electrode system a total of 1613 = 208 independent voltage
measurements are obtained. However, due to the reciprocity principle which states
that the voltage measured by a pair of electrodes when a current is injected
through another electrode pair is identical to that measured when voltage
measurement and current injection electrode pairs are reversed, only half, i.e. 104
measurements are independent. The current density for this method is not uniform
throughout the body as the current flows are concentrated in the area near the
injecting electrodes. This reduces the sensitivity of the voltage measurements to
the spatial conductivity variations in the middle of the object as well as at points
far from the driving electrodes.
The second strategy uses opposite method in which current is injected
through diametrically opposed electrodes (see Figure 3.2b). The voltages are
measured with respect to one reference electrode next to the current electrode.
This method gives the same number of independent measurements as the adjacent
method but gives a more uniform current density and hence better sensitivity than
the adjacent method. A combination of these two methods is the cross method
(see Figure 3.3a) in which one electrode, e.g. electrode 1, is used as the current
reference and the current source is applied successively to electrodes 3, 5, 7, ...,
15. For each current pair the voltages are measured with respect to electrode 2,
71
which acts as the voltage reference, for all the other electrodes except the current
injection electrodes. Another set of data is obtained by changing the reference
electrode of both current and voltage. The number of independent measurements
is the same as in the previous two methods.
(a) (b)
Figure 3.3. a). In the cross method, current electrodes are separated by large
dimensions. b) The multi-reference method uses one electrode to collect currents
injected simultaneously from other electrodes (redrawn from Webster [28]).
There are current and voltage electrodes at each electrode location.
Figure 3.4. The adaptive current method uses all electrodes to send currents
simultaneously to the object and measures the voltages with respect to the
reference (redrawn from Webster [28]). There are current and voltage electrodes
at each electrode location.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
I = cos 0
I = cos 180
72
The previous strategies use a pair of electrodes to feed currents and
another pair to sense the potential differences and leave the remaining electrodes
open circuited. Hua et al. [25] introduced a multi-reference method of data
collection strategy, in which they varied the current amplitude in each of many
current sources from 0 to 5 mA. For a 16-electrode system depicted in Figure 3.3b
the currents are simultaneously supplied to the system through electrodes 2
through 16 and exit via the grounded electrode 1 (the reference electrode). The
voltages are measured from electrodes 2 through 16 with respect to the reference.
Once this cycle completes the reference electrode is shifted to electrode 2 and
currents are injected via the remaining electrodes. The voltages are measured with
respect to the new reference.
Simske [27] and Zhu et al. [26] introduced a system utilizing an adaptive
current strategy. In the adaptive method (see Figure 3.4), independent currents are
sent to the object through all electrodes simultaneously and the corresponding
voltages are read with respect to the reference electrode. An algorithm is used to
adjust the current magnitudes following the particular conductivity distribution of
the object until the best current pattern is obtained. This method gives better
distinguishability, resolution and signal-to-noise ratio than a non adaptive system
with an expense of providing independently programmable current generator for
each electrode in the system.
As mentioned in Chapter 1.2 EIT images are not so good in spatial
resolution compared to other techniques. Nevertheless, high quality EIT images
are not impossible to produce. In doing so, measurements must be carried out with
a good accuracy using precision instrumentation that provides repeatable
73
measurements. Metherall et al. [29] obtained spatial resolution of about 10% of
the total cross-sectional diameter for a centrally located object, while Cook et al.
[11] reported a 14% resolution.
3.2 The Mathematical Model
The mathematical approach to solve the EIT problems was started by
Calderon [30] in 1980 where the uniqueness of the EIT inverse problem solution
was addressed. This was followed by many other researchers [31-39], who
proposed and studied the uniqueness of the recovery of the internal conductivity
distribution of the body from peripheral measurements for the isotropic case. If
the conductivity is anisotropic the recovery is non-unique [40].
3.2.1 Governing Equations
In an inhomogeneous medium the Maxwells equations are written as
B
E
t
c
V =
c
(3.1)
and
D
H J
t
c
V = +
c
, (3.2)
where E is electric field, H magnetic field, B magnetic induction, D electric
displacement and J electric current density. Also, in a linear isotropic medium the
following relations apply
74
D E c =
B H = (3.3)
J E o =
where c is the permittivity, permeability and o conductivity of the medium. This
isotropic condition is generally used to approximate the anisotropic biological
events in EIT.
Using relations (3.3) with the assumption that the injected currents are
time-harmonic with angular frequency e, i.e.
i t
E Ee
e
=

and
i t
B Be
e
=

, the
Maxwells Equations (3.1) and (3.2) can be expressed as
E i H e V = (3.4)
and
H J i E ec V = + . (3.5)
It is known that in any bioelectric phenomena current sources, J
s
, are
present. Hence, the current density comprises of two components, J = J
o
+ J
s
, with
J
o
is the ohmic current. Having this condition, Equations (3.4) and (3.5) can be
rewritten as
E i H e V = (3.6)
and
( )
s
H i E J o ec V = + + . (3.7)
75
after taking J
o
= oE. These are the full form of the Maxwells equations [41].
In EIT these equations are simplified by assuming a static condition for a
linear, isotropic medium. So,
E u = V (3.8)
and
s
H E J V = + (3.9)
where u is the electric potential and = o + iec is the complex admittivity. Taking
the divergence of both sides of Equation (3.9) and substituting Equation (3.8) into
Equation (3.9) gives,
s
E J V = V (3.10)
and since J
s
= 0 inside the body O, we finally have the equivalent of Poissons
equation for the EIT as
( ) 0, u x V V = eO (3.11)
Here x is a point inside O and is the admittivity of a block of
homogeneous material that is proportional to the reciprocal of its impedance.
76
3.2.2 Boundary Conditions
In order to solve the problems in EIT, a boundary condition must be
found. Consider the situation illustrated in Figure 3.5. A very small cylindrical
volume t is placed on the surface of the object O such that its top and bottom
sides are in parallel to the object boundary. Integrating the Equation (3.10) over
the volume t,
s
E d J d
t t
t t V = V
) )
(3.12)
and applying the divergence theorem, we have
s
S S
E ndS J ndS =
) )
(3.13)
where dS is the boundary of t and n is the unit normal. If then the volume t
shrinks to zero (e.g. t 0) its top and bottom sides overlap and sit on the surface
and the relation
inside
outside
s
E n J n = (3.14)
is true, since J
s
= 0 inside the object and E = 0 outside the object. When currents
are applied to the electrodes on the surface cO of the body they produce a current
density on the surface whose inward pointing normal component is denoted by j.
77
Thus, using the relation E = Vu (i.e. Equation (3.8)) the boundary
condition
s
u
J n j
n

c
=
c
(3.15)
on the boundary cO is obtained, where j is the injected current.
Figure 3.5. A situation used to derive Neumann boundary conditions. J
s1
and
J
s2
are the current density outside and inside the body O, respectively. E
1
and
E
2
are the corresponding electric fields. The size of the small cylindrical
volume t has been exaggerated.
One possible model for EIT uses Equation (3.11) and Equation (3.15) and
applies the conservation of charge condition 0 j
cO
=
)
and the condition 0 u
cO
=
)
,
which amounts to choosing a ground or reference voltage. This model is the
commonly used continuum model, which assumes that the injected current j is a
spatial continuous function, that is
O
J
s2
J
s1
E
2
E
1
n
t
78
( ) cos( ) j C k u u = (3.16)
where C is a constant and 2 / L u t = / , 1, , L = / . , 1, , / 2 k L = . . Unfortunately,
this model overestimates the resistivity values by as much as 25% when using
experimental data [42] as the effect of the electrodes are ignored and because we
do not know the injected current j.
In practice we only inject known currents I
/
through the electrodes. So,
the injected current density is given by
, 1, 2,...,
0 elsewhere
I
x e L
e
j

e =

/
/
/
/
(3.17)
where e
/
is the electrode contact area, I
/
is the injected current into the
th
/
electrode and L the number of electrodes. This model is known as the gap model,
which although it gives a slight improvement over the continuum model it is still
inadequate and overestimates the resistivities inside the body because it ignores
the shunting effect of the electrodes due to the highly conductive electrode
material and their contact impedances [42].
The shunt model takes into account the shunting effect of the electrodes
(i.e. the highly conductive metal electrodes applied to body circumference shunt
some of the current through the electrodes instead of through the body), that is,
the voltages are constant across each electrode. Therefore, the integral of the
79
current density over the electrode is equal to the total current flowing into it. Thus
the boundary condition has a more reliable condition
, 1, 2,...,
e
u
ds I L
n

c
= =
c
)
/
/
/ (3.18)
where I
/
is the injected current into the
th
/ electrode. The shunting effect of
electrodes imposes the constraint that u is constant on each of the electrodes. In
practice these values are the measured voltage on the electrodes. Therefore,
on u U e =
/ /
(3.19)
where U
/
is the measured voltage. Unfortunately, the shunt model does not
produce good estimation as it ignores the contact impedance of the voltage
electrodes [42] that may arise from the electrode geometry as well as the
electrodeselectrolyte interface. Even when the four-terminal method is in use the
contact impedance remains non-zero (see Chapter 5.3 for discussion).
Another model that takes into account the presence of the electrode contact
impedance and the shunting effect of the electrodes is the complete electrode
model. Therefore, this model is found to be the most accurate [42, 43].
Introducing contact impedance of the
th
/ electrode, z
/
, into (3.19) and recalling
the Poissons Equation (3.11) we have
80
, 1, 2,...,
u
u z U x e L
n

c
+ = e =
c
/ / /
/ (3.20)
And for the shunting effect we have Equation (3.18),
, 1, 2,...,
e
u
ds I L
n

c
= =
c
)
/
/
/
This model also includes the boundary condition 0
u
n

c
=
c
in the gaps
(spaces) between electrodes. Thus, the following conditions
1
0
L
I
=
=
_ /
/
(3.21)
and
1
0
L
U
=
=
_ /
/
(3.22)
must also be considered to produce the best and unique result. Somersalo [43]
found this model has proven to give measurement accuracy of more than 0.1%.
81
3.3 Numerical Implementation/Reconstruction Methods
and Algorithms
The first attempt on the work to reconstruct impedance imaging was done
by Henderson and co-workers in late 1970s [44, 45] using an impedance camera
they built for thoracic measurements (see Figure 1.4). The instrument measured
current that flows into the system upon application of a constant voltage source
via a large voltage electrode placed on one side of the object. An array of 100
electrodes, surrounded by 44 guarding electrodes, were placed on the other side of
the body and used to measure the current. From these measurements 100 specific
impedances (Z
i
= V/I
i
) were calculated and the impedance image was then
deduced.
In EIT, the reconstruction procedures are started by setting an initial
homogenous impedance distribution and peripheral potentials were estimated
assuming that constant currents are injected. The resulted potentials are then
compared to the actual measured peripheral values. The new values of the
boundary potentials are used to calculate new distribution of the internal
impedance. The new impedance distribution is then returned to the algorithm to
recalculate new boundary potentials. The process continues and alternating
iterations performed until a convergent solution is reached or a number of
iterations have been completed. The procedures to calculate peripheral potentials
from internal impedance distributions is called the forward solution; and the
procedure to yield internal impedance distribution from boundary potentials is
called the inverse problem solution.
82
Figure 3.6. The finite element mesh used in the reconstruction calculations.
The electrodes positions are marked with the green elements.
The forward problem for a simple case of a homogenous conductivity
distribution in a circular geometry can be solved analytically. However, the real
situation of the inverse problem is far more complicated and numerical methods
are the only choice. The most commonly used numerical method for solving the
partial differential equations (PDEs) in EIT is the finite element method (FEM)
[19, 46-48]. The FEM is a suitable method for solving PDEs with complex
geometries and non-trivial boundary conditions both for two- and three-
dimensional cases [49-51].
The FEM procedure starts with the formulations of the discontinuous
Galerkin-type variational (weak) equations of the problems [52]. The continuous
form of the problem is turned into a discrete formulation using the method of
finite elements in that the approximation and discretization are constructed
entirely using nodal information [53]. This attribute is extremely attractive as it
permits reduced restriction in the discretization of the problem domain. The area
83
O to be solved is divided into small elements with geometries of rectangles,
triangles or any other more complicated forms. This study uses triangles as shown
in Figure 3.6. The vertices of the triangles are called the nodes. The potential of
the area is approximated by the sum of the electric potentials of the elements
1
( ) ( )
N
h
i i
i
u x x o |
=
=
_
(3.23)
where x is the nodal coordinate within the mesh, |
i
(x) are the basis functions of
the finite-dimension subspace Q
h
, N is the number of nodes in the finite element
mesh. The potentials on the electrodes are approximated as
1
1
L
h
j j
j
U n |

=
=
_
(3.24)
where L is the number of electrodes, n
1
= (1, -1, 0, , 0)
T
, n
2
= (1, 0, -1, , 0)
T
e
1 L
1 , etc. are matrices representing the current injection patterns. The choice of n
j
is made to ensure that the condition (3.22) is satisfied. o
i
and|
j
in both equations
are the coefficients to be determined.
It has been shown by Somersalo, Cheney and Isaacson [43] that for any
(v,V), v e H
1
(O), V e
1 L
C
( ) ( ) ( )
1
, , ,
L
s
B u U v V I V
=
=
_ / /
/
(3.25)
84
Where u and v are the potentials distribution inside the object, U is the calculated
voltages on the electrodes, V is the measured voltages on electrodes, H
1
(O) is the
associated Sobolev space and ( ) ( ) ( )
, , ,
s
B u U v V is the variational (weak) form of
the complete electrode model, defined as (for the full derivations see Renardy and
Rogers [54])
( ) ( ) ( ) ( )( )
1
1
, , ,
L
s
e
B u U v V u v dx u U v V dS
z

O
=
= V V +
_
) )
/
/ /
/
/
(3.26)
Now, by inserting Equations (3.23) and (3.24) into Equation (3.26) and by
choosing v = |
i
and V = n
j
we obtain a matrix equation
Ab f = (3.27)
from which b can be solved,
1
b A f

= . (3.28)
Here b
o
|
(
=
(

e
1 N L +
, where
1 2
( , ,..., )
T
N
o o o o = and
1 2 1
( , ,..., )
T
L
| | | |

= are
the coefficients to be determined; and
0
f
I
(
=
(

, where 0 = (0, , 0)
T
e
1 N
and
1 2 1 3 1
( , ,..., )
T
L
I I I I I I I =

e
( 1) 1 L
are the injected currents. The sparse
block matrix A e
( 1) ( 1) N L N L + +
takes the form
85
T
B C
A
C D
(
=
(

(3.29)
where
1
1
( , ) ,
L
i j i j
e
B i j dx dS
z
| | ||
O
=
= V V +
_
) )
/
/
/
(3.30)
i, j = 1, 2, , N
1 1
1 1
1 1
( , ) ,
j
i i
e e
j
C i j dS dS
z z
| |
+
+
| |
=
|
|
\ .
) )
(3.31)
i = 1, 2, , N, j = 1, 2, , L-1
1
1
( , ) ( ) ( )
L
i j
e
D i j n n dS
z
=
=
_
)
/
/ /
/
/
1
1
1
1
1 1
,
, , 1, 2,..., 1
,
j
j
e
i j
z
i j L
e
e
i j
z z
+
+

= =

+ =

(3.32)

where |e
j
|is the contact area of the electrode j and z is the electrode contact
impedance. Therefore, Equation (3.27) can be rewritten in a matrix form as
0
T
B C
C D I
o
|
( ( (
=
( ( (

(3.33)
86
The first N coefficients in b in Equation (3.28) give the approximate
solution u
h
for the nodes (nodal potentials) and the last L1 coefficients the
referenced potentials on the electrodes. The approximated potentials
h
U
/
on the
th
/ electrode are calculated from Equation (3.24) to yield
1
1
L
h
U |
=
=
_ /
/
2 1
h
U | =
3 2
h
U | = (3.34)
:
:
1
h
L L
U |

=
which can be expressed in a matrix form as
h
U | = (3.35)
where e
( 1) L L
1 is a sparse matrix of
1 1 1
1
1
1
| |
|

|
| =
|
|
|

\ .

(3.36)
87
Therefore, the potential differences of the electrode pairs U

=
1 2 1 3 1
( , ,..., )
T
L
U U U U U U is calculated from
1 T
U A f

=

(3.37)
where
( 1)
(0 )
L N L +
= e

.
3.3.1 Mesh design
As stated earlier, the image reconstruction area is divided into a mesh of
many elements. The construction of the mesh must be devised very carefully
because the ill-conditioning of the reconstruction is very much affected by the
number, size, shape, and the position of the mesh elements [55]. The effect of the
size and position can be determined by monitoring the sensitivity of boundary
measurements to those elements near the boundary where electrodes are located
[56]. In this sense a better resolution can be produced in this region, which
requires that the boundary mesh elements must have smaller size. However, if the
size is too small smaller than the smallest distinguishable object at that location
then solving the inverse problem will be more difficult.
The number of elements in the mesh is also crucial. If the mesh number is
more than the number of independent measurements, then the reconstruction
problem will be underestimated. The number of elements is also proportional to
the reconstruction time.
88
3.3.2 Forward Solution
In the forward solution process it is important to map the changes in the
measured voltages on the electrodes that are caused by small changes in the
conductivity of the elements in the mesh [57]. The Jacobian (sensitivity) matrix of
the forward operator with respect to the conductivity T is calculated with the aid
of physical modeling and FEM discretisation. The n
th
element of the Jacobian
matrix is denoted by
k
kn
n
U
J
o
c
c
/
/
= and is calculated as
n
k
k
n
U
u u
o
A
c
= V V
c
)
/
/
(3.38)
where u
A
and u
k
are the voltage distributions when the A
th
and k
th
current patterns
were used and A
n
is the triangular element with respect to which the derivative is
calculated. The matrix takes the form [57]
1 1
1 1
1
1 1
2 2
1
1 1
1
1
N
N
K K
N
K K
L L
N
U U
U U
J
U U
U U
o o
o o
o o
o o
( c c
(
c c
(
(
c c
(
c c
(
(
(
=
( c c
(
c c
(
(
(
c c
(
(
c c

. .

. .

(3.39)
89
thus, the Jacobian is denoted as U' (o
0
) = J.
In the case of complex admittivity the Jacobian is composed of four
matrices which resemble the sensitivities of the real and imaginary parts of the
measurements with respect to real or imaginary admittivity perturbations, i.e.
r
rr
r
U
J

c
=
c
,
r
ri
i
U
J

c
=
c
,
i
ir
r
U
J

c
=
c
and
i
ii
i
U
J

c
=
c
, where V
r
and V
i
are the real and
imaginary parts of the measurements while
r
= o and
i
= ec
0
c
r
the conductivity
and permittivity vectors. The complex formulation of the forward problem is
therefore [58]
rr ri r r
ir ii i i
J J U
J J U

( ( (
=
( ( (

(3.40)
The forward calculations are performed by using the Cholesky method or
preconditioned conjugate gradient if
1 N

e and the LU method or biconjugate


gradient for the case if
1 N

e .
3.3.3 Inverse Problem
The inverse problem deals with the recovery of the unique distribution of
internal conductivity of an object when the peripheral potentials are known. A
deterministic model for EIT can be written as
( ; ) / ( ) V U I I G o o = = (3.41)
90
where
L K
V

e on the left is a vector of the measured voltages on the electrodes


in which K is the number of current patterns, ( ; )
L K
U I o

e is the (discrete)
model between parameters
1 N
o

e and V and ( ) o

e
L K
G is a vector of the
conductance distribution. In this case we assume that ( ; ) ( ) U I U o o . Since it is
an ill-posed problem, this model depends nonlinearly on the conductivity o and
linearly on the current I. The matrix G(o) is the part of the inverse of the matrix A
in FEM calculations. In reality, the physical model of Equation (3.41) is a discrete
version of the inverse of the so-called Dirichlet-to-Neumann (DN) map.
Since the data depends nonlinearly on o a nonlinear iterative method such
as the Gauss-Newton (GN) algorithm has to be used in order to obtain the
solution. However, in EIT this is often avoided for practical reasons, such in fast
imaging, because such iterative procedures require lengthy calculations.
There are two different approaches to the solution in the nonlinear case.
The first is to linearise the model U(o) at o
0
and compute only the difference oo.
The second approach use iterative methods, such as the GN algorithm. In the first
approach the forward solution is needed only for the computation of the Jacobian.
Since two separate measurements are needed, absolute values for the conductivity
distribution cannot be obtained. In the GN algorithm the forward problem has to
be solved in each iteration. However, in both of these approaches the Jacobian of
the discrete map U(o) is required.
In the first approach where the model is linearised and only the difference
oo is calculated. U(o) is approximated using the first order Taylor series at o
0
, i.e.
91
( )
2
0 0 0 0
( ) ( ) '( )( ) U U U O o o o o o o o = + + (3.42)
where
0
'( )
LK N
U o

e is the Jacobian of ( )
LK
U o e calculated at o
0
. By
eliminating the second and higher order terms of the expansion, Equation (3.42)
can be simplified by substituting
U J o oo ~ (3.43)
where oU = U(o) U(o
0
) and oo = o o
0
and J = U'(o
0
). Thus, in this non-
iterative approach we need to solve
min U J
oo
o oo (3.44)
where oU is the difference between two successive measurement sets. The
solution of oo is

J U oo o = (3.45)
which needs to be regularised due to the ill-posedness of the problem.
In the second approach where the iterative method is used, the solution is
mainly based on the Gauss-Newton algorithm with certain regularisation and
weighting matrices. A Tikhonov-type regularisation [59] similar to the variable
regularisation parameter proposed by Lavenberg [60] and Marquardt [61] is
92
usually employed to handle the ill-conditioned Jacobian. The algorithm for the
i+1th iteration is
1 i i i
o o oo
+
= + (3.46)
where
( ) ( ) ( )
1
( )
T T
i i i i i
J J I J U U oo o o

= + (3.47)
where o is the weighted regularisation matrix, I is the identity matrix, U is the
measured voltage vector and U(o
i
) is the voltage vector obtained from the forward
solution of the i
th
iteration. The similar applies to the complex conductance, .
3.4 Determinations of the Spatial Resolution of the
Reconstructed Images
The reconstructed images dimension (i.e. the width) is determined using
the full-width at half-maximum (FWHM) method [29, 62-64]. The FWHM
method is commonly used to measure the width of an object in a picture when it
does not have sharp edges.
Figure 3.7 shows the FWHM of a normal Gaussian function. The width is
given by the distance between points on the curve at which the function reaches
half of its maximum value.
93
Figure 3.7 The FWHM method to determine the width of a Gaussian
function [64]. The width of the function is given by the distance of the
curves legs at half its maximum.
The spatial resolution of the recovered images is presented as percentage
of the ratio of the FWHM to the diameter of the imaging region,
(%) 100%
FWHM
w
d
= (3.48)
where d is the diameter of the imaging region (i.e. the chamber).
94
References
1. Boone, K., D. Barber, and B. Brown. (1997). Imaging with electricity:
report of the European Concerted Action on Impedance Tomography.
Journal of Medical Engineering and Technology, 21(6): 201-232.
2. Riu, P.J., J. Rosell, A. Lozano, and R. Palls-Areny. (1992). A broadband
system for multifrequency static imaging in electrical impedance
tomography. Clinical Physics and Physiological Measurement, 13 Suppl
A: 61-65.
3. Cherepenin, V., A. Karpov, A. Korjenevsky, V. Kornienko, Y. Kultiasov,
A. Mazaletskaya, and D. Mazourov. (2002). Preliminary static EIT images
of the thorax in health and disease. Physiological Measurement, 23(1): 33-
41.
4. Hartov, A., R.A. Mazzarese, F.R. Reiss, T.E. Kerner, K.S. Osterman, D.B.
Williams, and K.D. Paulsen. (2000). A multichannel continuously
selectable multifrequency electrical impedance spectroscopy. IEEE
Transactions on Biomedical Engineering, 47(1): 49-58.
5. Halter, R., A. Hartov, and K.D. Paulsen. (2004). Design and
implementation of a high frequency electrical impedance tomography
system. Physiological Measurement, 25(1): 379-390.
6. Woo, E.J., P. Hua, J.G. Webster, and W.J. Tompkins. (1992). Measuring
lung resistivity using electrical impedance tomography. IEEE Transactions
on Biomedical Engineering, 39(7): 756-760.
7. Hoetink, A.E., J.C. Faes Th, J.T. Marcus, H.J.J. Kerkkamp, and R.M.
Heethaar. (2002). Imaging of thoracic blood volume changes during the
heart cycle with electrical impedance using a linear spot-electrode array.
IEEE Transactions on Medical Imaging, 21(6): 653-661.
8. Hahn, G., M. Beer, I. Frerichs, T. Dudykevych, T. Schroder, and G.
Hellige. (2000). A simple method to check the dynamic performance of
electrical impedance tomography systems. Physiological Measurement,
21(1): 53-60.
95
9. Kim, K.Y., B.S. Kim, M.C. Kim, S. Kim, D. Isaacson, and J.C. Newell.
(2005). Dynamic electrical impedance imaging with the interacting
multiple model scheme. Physiological Measurement, 26(2): S217-233.
10. Kim, K.Y., B.S. Kim, M.C. Kim, and S. Kim. (2004). Dynamic inverse
obstacle problems with electrical impedance tomography. Mathematics
and Computers in Simulation, 66(4): 399-408.
11. Cook, R.D., G.J. Saulnier, D.G. Gisser, J.C. Goble, J.C. Newell, and D.
Isaacson. (1994). ACT3: a high-speed, high-precision electrical
impedance tomograph. IEEE Transactions on Biomedical Engineering,
41(8): 713-722.
12. Edic, P.M. (1994). The implementation of a real-time electrical impedance
tomograph. in Rensselaer Polytech. Inst.,Troy,NY,USA. FIELD URL:.
13. Taktak, A., A. Spencer, P. Record, R. Gadd, and P. Rolfe. (1996).
Feasibility of neonatal lung imaging using electrical impedance
tomography. Early Human Development, 44(2): 131-138.
14. Koukourlis, C.S., G.A. Kyriacou, and J.N. Sahalos. (1995). A 32-electrode
data collection system for electrical impedance tomography. IEEE
Transactions on Biomedical Engineering, 42(6): 632-636.
15. Smith, R.W., I.L. Freeston, and B.H. Brown. (1995). A real-time electrical
impedance tomography system for clinical use--design and preliminary
results. IEEE Transactions on Biomedical Engineering, 42(2): 133-140.
16. Kerner, T.E., D.B. Williams, K.S. Osterman, F.R. Reiss, A. Hartov, and
K.D. Paulsen. (2000). Electrical impedance imaging at multiple
frequencies in phantoms. Physiological Measurement, 21(1): 67-77.
17. Kerner, T.E., A. Hartov, K.S. Osterman, C. DeLorenzo, and K.D. Paulsen.
(2001). An improved data acquisition method for electrical impedance
tomography. Physiological Measurement, 22(1): 31-38.
18. Heikkinen, L.M., T. Vilhunen, R.M. West, and M. Vauhkonen. (2002).
Simultaneous reconstruction of electrode contact impedances and internal
electrical properties: II. Laboratory experiments. Measurement Science
and Technology, 13(12): 1855-1861.
96
19. Hartov, A., T.E. Kerner, and K.D. Paulsen. (2001). Simulation of error
propagation in finite element image reconstruction for electrical
impedance tomography. Measurement Science and Technology, 12(8):
1040-1049.
20. Gadd, R., P. Record, and P. Rolfe. (1992). Electrical impedance
tomography. A sensitivity region reconstruction algorithm using adjacent
drive current injection strategy. Clinical Physics and Physiological
Measurement, 13 Suppl A: 101-105.
21. Kolehmainen, V., M. Vauhkonen, P.A. Karjalainen, and J.P. Kaipio.
(1997). Assessment of errors in static electrical impedance tomography
with adjacent and trigonometric current patterns. Physiological
Measurement, 18(4): 289-303.
22. Schuessler, T.F. and J.H. Bates. (1998). Current patterns and electrode
types for single-source electrical impedance tomography of the thorax.
Annals of Biomedical Engineering, 26(2): 253-259.
23. De Simone, B.C., R. Siciliano, A. Pachi, C. Cametti, and F. De Luca.
(2002). Electrical impedance tomography via filtered-back projection of
fan current distribution: a numerical simulation. Bioelectromagnetics,
23(7): 516-521.
24. Yerworth, R.J., R.H. Bayford, G. Cusick, M. Conway, and D.S. Holder.
(2002). Design and performance of the UCLH mark 1b 64 channel
electrical impedance tomography (EIT) system, optimized for imaging
brain function. Physiological Measurement, 23(1): 149-158.
25. Hua, P., J.G. Webster, and W.J. Tompkins. (1987). Effect of measurement
method on noise handling and image quality of EIT imaging. Proceeding
of Annual International Conference of IEEE Enhineering in Medicine and
Biology Society, 9: 1429-1430.
26. Zhu, Q.S., C.N. McLeod, C.W. Denyer, F.J. Lidgey, and W.R. Lionheart.
(1994). Development of a real-time adaptive current tomograph.
Physiological Measurement, 15 Suppl 2a: A37-43.
27. Simske, S.J. (1987). An adaptive current determination and one-step
reconstruction technique for a current tomography system, Masters
97
Thesis. MSc Thesis in Graduate Faculty. Rensselaer Polytechnic Institute.
Troy, New York.
28. Webster, J.G. (1990). Electrical Impedance Tomography. IOP Publishing
Ltd. Bristol.
29. Metherall, P., D.C. Barber, R.H. Smallwood, and B.H. Brown. (1996).
Three-dimensional electrical impedance tomography. Nature, 380(6574):
509-512.
30. Calderon, A.P. (1980). On an inverse boundary value problem. Seminar
on numerical analysis and its applications to continuum physics. Rio de
Janeiro: Brazilian Mathematical Society: 65-73.
31. Friedman, A. and V. Isakov. (1989). On the uniquness in the inverse
conductivity problem with one measurement. Indiana University
Mathematics Journal, 38: 563-579.
32. Isakov, V. (1988). On uniqueness of recovery of a discontinuous
conductivity coefficient. Communications on Pure and Applied
Mathematics, 41: 865-877.
33. Kim, H. and J.K. Seo. (1996). Unique determination of a collection of a
finite number of cracks from two boundary measurements. SIAM Journal
on Mathematical Analysis, 27: 1336-1340.
34. Kohn, R. and M. Vogelius. (1984). Determining conductivity by boundary
measurements. Communications on Pure and Applied Mathematics, 37:
289-298.
35. Nachman, A.I. (1996). Global uniqueness for a two-dimensional inverse
boundary value problem. Annals of Mathematics, 143: 71-96.
36. Ramm, A.G. (1988). A simple proof of the uniqueness theorem in
impedance tomography. Applied Mathematics Letter, 1: 287-290.
37. Sun, Z. (1990). The inverse conductivity problem in two dimensions.
Journal of Differential Equations, 87: 227-255.
38. Sylvester, J. and G. Uhlmann. (1987). A global uniqueness theorem for an
inverse boundary value problem. Annals of Mathematics, 125: 153-169.
98
39. Sylvester, J. and G. Uhlmann. (1986). A uniqueness theorem for inverse
boundary value problem in electricalprospection. Communications on
Pure and Applied Mathematics, 39: 91-112.
40. Sylvester, J. (1990). An anisotropic inverse boundary value problem.
Communications on Pure and Applied Mathematics, 43: 201-232.
41. Somersalo, E., D. Isaacson, and M. Cheney. (1992). A linearized inverse
boundary value problem for Maxwell's equations. Journal of
Computational and Applied Mathematics, 42(1): 123-136.
42. Cheng, K.S., D. Isaacson, J.C. Newell, and D.G. Gisser. (1989). Electrode
models for electric current computed tomography. IEEE Transaction on
Biomedical Engineering, 36(9): 918-924.
43. Somersalo, E., M. Cheney, and D. Isaacson. (1992). Existence and
uniqueness for electrode models for electric current computed
tomography. SIAM Journal on Applied Mathematics, 52: 1023-1040.
44. Henderson, R.P., J.G. Webster, and D.K. Swanson. (1976). A thoracic
electrical impedance camera. Proceeding of the Annual Conference of
Engineering in Medicine and Biology, 18: 322.
45. Henderson, R.P. and J.G. Webster. (1978). An impedance camera for
spatially specific measurements of the thorax. IEEE Transactions on
Biomedical Engineering, 25(3): 250-254.
46. Babuska, I., R. Tempone, and G.E. Zouraris. (2005). Solving elliptic
boundary value problems with uncertain coefficients by the finite element
method: the stochastic. Computer Methods in Applied Mechanics and
Engineering, 194(12): 1251-1294.
47. Cao, W., W. Huang, and R.D. Russell. (1999). Anr-Adaptive Finite
Element Method Based upon Moving Mesh PDEs. Journal of
Computational Physics, 149(2): 221-244.
48. Mofid, M., A. Vafai, and K. Farahani. (1999). Finite element solution of
Dirichlet's nonlinear partial differential equation with mixed boundary
conditions. Computer Methods in Applied Mechanics and Engineering,
169(1): 81-88.
99
49. Bonovas, P.M., G.A. Kyriacou, and J.N. Sahalos. (2001). A realistic three
dimensional FEM of the human head. Physiological Measurement, 22(1):
65-76.
50. Bayford, R.H., A. Gibson, A. Tizzard, T. Tidswell, and D.S. Holder.
(2001). Solving the forward problem in electrical impedance tomography
for the human head using IDEAS (integrated design engineering analysis
software), a finite element modelling tool. Physiological Measurement,
22(1): 55-64.
51. Vauhkonen, P.J., M. Vauhkonen, T. Savolainen, and J.P. Kaipio. (1999).
Three-dimensional electrical impedance tomography based on the
complete electrode model. IEEE Transactions on Biomedical Engineering,
46(9): 1150-1160.
52. Tereshchenko, V.Y. (1996). A relation between the formulations of the
boundary element method. Journal of Applied Mathematics and
Mechanics, 60(4): 615-620.
53. Chen, J.-S., S. Yoon, and C.-T. Wu. (2002). Non-linear version of
stabilized conforming nodal integration for Galerkin mesh-free methods.
International Journal for Numerical Methods in Engineering, 53: 2587-
2615.
54. Renardy, M. and R.C. Rogers. (1993). An introduction to partial
differential equations. Texts in Applied Mathematics 13, ed. F. John, et al.
Springer-Verlag. New York.
55. Gisser, D.G., D. Isaacson, and J.C. Newell. (1990). Electric current
computed tomography and eigenvalues. SIAM Journal on Applied
Mathematics, 50: 1623-1634.
56. Isaacson, D. (1986). Distinguishability of conductivities by electric current
computed tomography. IEEE Transaction on Medical Imaging, MI-5(2):
92-95.
57. Vauhkonen, M. (1997). Electrical impedance tomography and prior
information. PhD Thesis in Department of Applied Physics. University of
Kuopio. Kuopio, Finland.
100
58. Polydorides, N. and W.R.B. Lionheart. (2002). A Matlab toolkit for three-
dimensional electrical impedance tomography: a contribution to the
Electrical Impedance and Diffuse Optical Reconstruction Software
project. Measurement Science and Technology, 13(12): 1871-1883.
59. Tikhonov, A.N. and V.Y. Arsenin. (1977). Solutions of ill-posed problems.
Scripta Series in Mathematics. Vh Winston. Washington D.C.
60. Lavenberg, K. (1944). A method for the solution of certain problem in
least squares. SIAM Journal on Numerical Analysis, 16: 588-604.
61. Marquardt, D. (1963). An algorithm for least squares estimation of
nonlinear parameters. SIAM Journal on Applied Mathematics, 11: 431-
441.
62. Wexler, A., B. Fry, and M.R. Neuman. (1985). Impedance-computed
tomography algorithm and system. Applied Optics, 24: 3985-3992.
63. Ider, Y.Z., B.M. Eyuboglu, M. Kuzuoglu, K. Leblebicioglu, U. Baysal,
B.K. Caglar, and O. Birgu. (1995). A method for comparative evaluation
of EIT algorithms using a standard test. Physiological Measurement, 16
Suppl. 3A: 227-236.
64. Murugan, R.M. (1999). An improved electrical impedance tomography
(EIT) algorithm for the detection and diagnosis of early stages breast
cancer. PhD Thesis in Department of Electrical and Computer
Engineering. University of Manitoba. Winnipeg, Canada.
4
MATERIALS AND METHODS
4.1 Data Acquisition System and Methods
Traditionally, AC impedance measurements have been carried out using a
bridge with manual balancing. However, for a heterogenous system, where each
constituent material has characteristic dielectric and conductance properties, the
impedance disperses with the frequency. For example, the impedance of the ionic
double layer at the metal-solution interface (see Chapter 2.3) disperses by five or
more orders of magnitude over the frequencies range between 10 mHz 10 kHz.
Determination of the null point, in the bridge method, is very hard and tedious at
low frequencies and especially at frequencies approaching 10 mHz where the
period is 100 s. Therefore, this manual method, which is useful at high
frequencies, becomes impractical to use at very low frequencies.
In 1975, Bell et al. [1] introduced their digital system to overcome this
burden. This instrumentation was further developed by Chilcott [2] and Coster et
al. [3] and is known as the Biophysics Ultra Low Frequency Spectrometer
(BULFIS). This project employed BULFIS to collect data.
BULFIS is controlled via a PC using a program written in the ZBASIC
language. The basic block diagram of the BULFIS is shown in Figure 4.1. It
consists of differential amplifiers (input impedance ~10
12
O and gain = 50),
102
Analog-to-Digital Converters (ADCs) and Digital-to-Analog Converters (DACs).
The digital sine wave tables are generated by the computer and stored in memory
(65k x 16 bit RAM) on the signal generator board. The clock on this board then
downloads digital data to a DAC to generate a raw analog signal, which is then
filtered to remove any remaining notches/glitches that distort the signal before
application to the unknown impedance system via a standard impedance. Figure
4.2 shows the BULFIS sitting next to the PC.
Some 65k digital samples of each AC signal cycle are acquired
simultaneously from both the standard impedance and the unknown impedance
system via two separate data acquisition boards, each of which utilises either a
high-speed (12 bit, 10 MHz) or high-resolution (16 bit, 500 kHz) ADCs. A single
clock programmed via the computer controls all signal generations and data
acquisition sequences. The data is stored in on-board RAM (65k x 16 bit RAM)
prior to transfer to the computer for analysis. Then, an optimised least-square
fitting routine is used to calculate the magnitude, phase, offset and the signal-to-
noise ratio (SNR) of the measured AC signals. The SNR is an important quantity
that reflects the quality (e.g. noise level and distortion) of the measurements.
103
Figure 4.1. Schematic diagram of BULFIS as illustrated in Coster et al. [3].
standard
impedance
RAM
from PC
system
clock
Filter
DAC
A
A
A
DAC
RAM RAM
ADC
control latches
from
PC
to
PC
to
PC
Four-terminal configuration
DC offset
ADC
104
Figure 4.2. Photograph of the BULFIS sitting next to the PC.
The data acquisition procedure is repeated if the SNRs are not large
enough. The software automatically compensates for DC offset (e.g. resulting
from poor electrode coating) and adjusts the amplitude and filtering to maximise
SNRs. The magnitudes and phases (and the known impedance of the standard) are
used to calculate the conductance and the capacitance of the unknown system. The
BULFIS operates in the frequency range of 0.006 Hz 100 kHz and is capable of
providing very high resolution of 0.002% in magnitude and 0.001
o
in the phase
difference. The measurements through out the course of the project were carried
out in the frequency range of 1.11757 Hz 77,712 Hz.
The sets of data for image reconstruction in this project were collected
using the adjacent/neighbouring method. The current was injected into the system
through a pair of adjacent current electrodes, starting with electrodes number 1
BULFIS
105
and 2 (see Figure 4.6). Voltages were then measured between pairs of voltage
electrodes 1-2, 2-3, 3-4, , and 16-1. In the next step the current was injected via
electrodes 2-3 and similar voltage measurements were carried out using the
voltage electrodes. The procedure was complete after voltage measurements were
acquired using the electrode pair 16-1 for current injection. Most of the
measurements were done automatically using a programmable multiplexer to
switch the current between the various current electrode pairs and switching
between the voltage sampling electrodes. However, a number of measurements
were also performed by manually connecting the current and voltage electrodes.
This latter manual procedure was very tedious and time consuming but was done
to verify the integrity of the results as the impedance of the stray capacitance of
the multiplexer hardware dominated measurements at high frequencies.
4.2 Experimental Solution and Objects
A laboratory grade NaCl (Rhone-Poulenc, product code SL944) was used
in the experiment. A stock of 2000 mM was prepared by dissolving the NaCl
powder into Milli-Q water (59.3 S/cm) and later diluted into the desirable
concentrations of 50 mM, 100 mM and 150 mM. The conductivities of the saline
solutions were measured using a labCHEM-901CP Conductivity, TDS, pH, mV
and Temperature-meter with a k=1.0/ATC conductivity cell probe (TPS Pty. Ltd.,
Brisbane, Australia). The conductivities were 7.02 mS/cm, 10.67mS/cm and 15.12
mS/cm for 50 mM, 100 mM and 150 mM solutions respectively. The chamber
was filled with 2 litres of saline in the absence of an object. When the object to be
106
imaged was present the saline volume was reduced so that the level was similar to
that level when no object was present.
The objects used as phantoms were cylindrical blocks of aluminium, 250-
ml beaker glasses (empty and filled) and perspex pipes. The aluminium was used
to imitate conducting organs and the glasses and Perspex pipes were to imitate the
insulating organs. The objects were put in the centre and off-centre (close to
electrodes numbers 1 and 9). This was done to investigate the effects of the
phantom location on the reconstructed images and the sensitivity of the EIDORS
algorithms on the variation of the phantom location. Silicone grease (RS
Component, product no. 494-124) was applied to the bottom of the phantoms in
order to make sure that there was no leakage of solution underneath, that may act
as a short circuit,.
4.3 The Experimental Chambers
The experimental systems employed were a square chamber and a circular
chamber.
The square system was comprised of a square perspex chamber of inner
dimension of 150 mm (w) x 150 mm (l) x 20 mm (h) with two flat current
electrodes made from copper attached to the chamber walls. The voltage
electrodes were made of pure silver wire (99.99%) of 1 mm in diameter and
inserted into a pair of movable holders, which sat on a movable slide on top of the
chamber as depicted in Figure 5.2. The set up was similar to that used by
Ackmann and Seitz [4].
107
The circular system was made from of cylindrical perspex tubing of inner
diameter of 300 mm and height of 350 mm (see Figure 4.3) with electrodes
equally spaced around the periphery. The current electrodes were made of 1 mm
thick copper plates, 25 mm wide and 300 mm long. Silver wires of 1 mm in
diameter were used for the voltage electrodes. The arrangement of the electrodes
and labelling are depicted in Figure 4.5 and Figure 4.6, respectively.
Figure 4.3. Photograph of the experimental chamber.
108
Figure 4.4. Photograph of the experimental chamber with the amplifier.
They sat inside a Faraday cage to eliminate external interferences.
Figure 4.5. Small segment of the electrodes arrangement in the chamber.
Two copper plates were coupled to form a current electrode. The gap
between two current electrodes was 5 mm. The silver voltage electrodes
were located 1 mm from its left and right current electrodes. The drawing is
not to scale.
Current electrode Current electrode
Voltage electrode Voltage electrode
1 mm
5 mm
25 mm
amplifier
chamber
the Faraday cage
109
Figure 4.6. Electrodes numbering arrangement. The electrodes were
numbered counter clock wisely with electrode #1 is located on the right hand
side and electrode #9 on the left hand side of the chamber (top view).
The silver wires were coated with silver-chloride (AgCl) to form Ag/AgCl
electrodes. This Ag/AgCl coating reduced the contact impedance with the solution
and, thus, minimised differences in the DC electrode potentials between pairs of
electrodes that needed to be offset via the amplifier, hardware and software (see
Figure 4.1).
Since the electrodes were arranged in a parallel formation, they could act
as antennas and pick up any external disturbances, e.g. from the 50 Hz power
lines, which will reduce SNRs. To minimise external interference, the chamber
and the amplifier unit of the BULFIS were placed inside a Faraday cage, which
was earthed.
One possible cause of poor voltage SNRs encountered during the
measurements was electrolysis that occurred at the copper electrode-solution
interface. The BULFIS is capable of generating a maximum AC voltage of 4 volts
1
5
6
4
9
2
3 7
8
13
10
12
11
14
16
15
110
peak to peak. Since the copper-solution interface has a high impedance the electric
field generated through the application of the AC voltage was sufficiently large to
induce electrolysis.
The copper electrodes act as an anode during half a cycle and as a cathode
during the remaining half. During the anodic half-cycle the Cu atoms react with
the Cl

ions in the solution to produce a precipitate of copper(I) chloride (CuCl)


[5, 6] that was found deposited on the chamber base. The reaction resulted in a
current that was larger in magnitude than that during the cathodic half-cycle. This
phenomenon contributed to the distortion to the sinusoidal signals at low
frequencies (see bottom peaks in Figure 4.7) that were detected as low SNRs by
the impedance spectrometer [7].. In order to improve the SNRs, the voltage
amplitude on the system was reduced from 4V peak-to-peak to a maximum of 500
mV peak-to-peak. This was done by programming a lower cut-off frequency for
the low-pass filtering on the signal generator board of the spectrometer.
111
Figure 4.7. Sketch of the observed signal interference pattern. (a) The signal
picked from the signal generator, and (b) the current flowing through the
electrode.
4.4 Reconstruction Software
The numerical processes described in Chapter 3.3 were implemented in
computer programs and routines written in MATLAB

version 5.11 and called the


Electrical Impedance and Diffuse Optical Reconstruction Software (EIDORS) that
was supplied by Dr. Marko Vauhkonen of the Department of Applied Physics,
University of Kuopio, Kuopio, Finland. Some modifications to the software were
made to use the software in conjunction to our hardware. The software utilises an
iteration procedure of Tikhonov-type (Equation (3.47)) to solve the inverse
problem. It also includes the electrode impedances in the calculation, fixed to a
112
certain uniform value [8] although the impedance disperses with frequency (as
described in Chapter 2). The flowchart in Figure 4.8 shows the reconstruction
procedures performed by EIDORS.
The EIDORS routine firstly loads the mesh data file containing parameters
of the finite element meshes for further calculation. It then loads data including
the frequency, impedance magnitude, impedance phase, conductance, capacitance,
and the set of electrodes used to make the measurements. Once the data have been
loaded, the program creates the adjacent current pattern, FEM and regularization
matrices by means of subroutines Cur r ent . m, FemMat r i x. m and
MakeRegmat r i x. m, respectively. Full listing of the EIDORS routines can be
found in Appendix A.
Next, EIDORS will set the first estimate of homogenous impedance of 0.5
mS/m and then compute the forward solution of the peripheral potentials using
For war dSol ut i on1. m subroutine after updating the FEM matrix
(Updat eFemMat r i x. m).
EIDORS iterations to find the internal admittance (reciprocal of
impedance) distribution now start using the initial values. The sensitivity Jacobian
matrix is calculated using Jacobi an. m subroutine. The internal real admittance
distribution is calculated from Equation (3.47). The program will plot the real and
imaginary parts of the resulted admittance and then use them to resume the
iteration procedures until the specified iteration number (ii in the program) is
reached when EIDORS halts the reconstruction procedures. The final image data
(including the mesh data, nodal and peripheral voltages, current pattern matrix,
etc.) are then saved in a .mat file type for further analysis.
113
Some modifications were made to the original scripts to accommodate the
data acquisition technique employed for this thesis which is found to be different
from that employed by Dr Vauhkonen when dealing with the complex input data
handling. The modifications were done especially to the For war dSol ut i on. m
routine renamed For war dSol ut i on1. m and the main routine.
114
Start
Load meshdata
Read Node number & co-ordinates and Element
number & co-ordinates from meshdata
Load .lis & ei3 data
Read Z
meas
, phase, U
meas
Create FEM matrix
Create regularization matrix
Set a homogenous system and calculate
peripheral voltages (U
ref
)
(Forward solution)
Calculate first estimation of
admittance (Y
0
)
(Inverse problem solution)
Find peripheral voltages (U
ref
)
Calculate estimation of admittance (Y
ii
)
Plot Re(Y
ii
) & Im(Y
ii
) images
ii = iter?
Set iteration number = iter
ii = 1
Save Re(Y
ii
) & Im(Y
ii
)
ii = ii + 1
yes
no
Figure 4.8. The flowchart of EIDORS.
115
References
1. Bell, D.J., H.G.L. Coster, and J.R. Smith. (1975). A computer based, four
terminal impedance measuring system for low frequencies. Journal of
Physics E: Scientific Instrument, 8: 66-70.
2. Chilcott, T.C. (1988). Admittance tomography of cells of Chara corallina:
a study of the electrical spatial structures associated with photosynthesis.
PhD Thesis in Faculty of Science. University of New South Wales.
Sydney, Australia.
3. Coster, H.G.L., T.C. Chilcott, and A.C.F. Coster. (1996). Impedance
spectroscopy of interfaces, membranes and ultrastructures.
Bioelectrochemistry and Bioenergetics, 40: 79-98.
4. Ackmann, J.J. and M.A. Seitz. (1984). Methods of complex impedance
measurements in biologic tissue. CRC Critical Reviews in Biomedical
Engineering, 11(4): 281-311.
5. Skoog, D.A. (1985). Principles of Instrumental Analysis. 3
rd
ed. Saunders
College Pub. Philadelphia.
6. Bard, A.J. and L.R. Faulkner. (2000). Electrochemical Methods:
Fundamentals and Applications. John WIley & Sons. New York.
7. Barber, D.C. and B.H. Brown. (1984). Applied potential tomography.
Journal of Physics E: Scientific Instrument, 17: 723-733.
8. Vauhkonen, M. (1997). Electrical impedance tomography and prior
information. PhD Thesis in Department of Applied Physics. University of
Kuopio. Kuopio, Finland.
5
ELECTRICAL PROPERTIES OF
THE EXPERIMENTAL
CHAMBER
5.1 Introduction
It has been common to employ metal electrodes for injecting electrical
currents and measuring voltage in electronic systems. Electrical Impedance
Tomography (EIT) also employs such metal electrodes. As was discussed in
Chapter 2, when an electrode is in contact with electrolyte solutions an ionic
double layer is established at the interface. In addition there will also be a contact
or electrode potential established. The latter is dependent on the composition and
concentration of the electrolyte and the composition of the electrode material. For
two electrodes of similar material in contact with the same electrolyte, the double
layers and the electrode-solution potential will, ideally, be identical. The result
would be a zero potential difference between the two electrodes. However, this
condition is affected by surface irregularity and possibly contaminants (e.g.
corrosions of the electrode material). To suppress the effects of double layer
potentials on electrodes, it is therefore essential to achieve similar surface
chemistries by appropriate a coating of the surfaces. For silver electrodes used in
these experiments this was achieved using electrolysis to coat them with silver
117
chloride. The Ag/AgCl electrode has a well defined electrode potential which
depends on the concentration of chloride ions in the electrolyte. The presence of
this layer on both of the measuring electrodes produces nearly identical electrode
potentials with respect to the same electrolyte which would cancel yielding each
otbher and hence produce a net zero potential difference between the electrodes.
When the system is comprised of only the electrolyte (no phantom objects
in the solution) the impedance can be derived from the Nernst-Planck equation.
For a homogeneous, isotropic material the most general equation for the current
density, J
k
, of an ion species k and the electric field gradient are
2 2
with
k k k k
k k k w k k k k
k
C z q D C E E q
J z qD v z qC z C
x kT x c
c c
= + + =
c c
_
(5.1)
where q is the electronic charge, c the dielectric permittivity, E the electric field
and z
k
, C
k
and D
k
are the valency (including the sign), concentration and diffusion
constant of the ion species k, respectively. The last term in Equation (5.1) is due
to the present of water fluxes with the bulk velocity of v
w
which in our system are
not present.
The total concentrations of the cation and anion of a univalent-univalent
electrolyte comprise the DC concentrations (i.e. N and P , respectively) and the
AC perturbed concentrations of the anion and cation (i.e. n
~
and p
~
, respectively),
so that = +
N
C N n and = +
P
C P p . By substituting +

E E for the total electric
field and + for the charge density into the Poissons equation (e.g. Equation
2.28) in effect the DC components cancel leaving only the AC terms yielding
118

c
c
=
c
E
x
(5.2)
where ( ) q p n = .
If we consider an ideal electrolyte the condition D
P
= D
N
= D is satisfied.
It is assumed that
0 0 0
0
P N w
V = V = V = , 0 E = and v
w
= 0 applies at the
boundary.
0
k
V is the standard chemical potential at position x for the reference
concentration of species k. For this electrolyte Coster and Chilcott have shown
in [1] that the AC charge density must satisfy the following conditions
2
2 2 2
( )
0 where and
o ec
k k o
e
+ +
V =
j q D P N
D kT
(5.3)
whose solution for the symmetrical electrode configuration depicted in Figure 5.2
is
sinh
c
x k = (5.4)
where
c
is constant in x.
Integrating Equation (5.2) twice gives the AC field
cosh

k
ck
= +

c
c
E x E (5.5)
119
where E
c
is constant in x; and potential difference
2
2
sinh
2 2 2
k
ck
| | | |
= + + = +
`
| |
\ . \ . )

c
c
l l l
V l E (5.6)
where presumably
1
2
x l = are the positions of the voltage electrodes. The total
ionic current is
2
2
( ) ( ) o o ec o
k
V
+ = V + = + +



P N
E
J J qD p n E j E (5.7)
Adding the displacement current to the ionic current gives the total current
2
2
( ) ( ) ec o ec o ec
k
V
= + + = + = +
`
)


P N c
E
J J J j E j E j E (5.8)
which as expected is also constant in x. Assuming that the current electrodes at the
boundaries (x = L/2) are blocking electrodes
1
then Equations (5.5) and (5.7) give
cosh
2
c
c
j L
E
e k
ok
= (5.9)
The admittance, y, per unit area is the defined as

1
A blocking electrode is an electrode at which no electrochemical reactions occur and, therefore,
admits no Faradaic currents.
120
1
o ec
oo
ec
+
= =

+
`
)

J j
y
V
l
j
(5.10)
where
2sinh
2
cosh
2
k
o
k
k

l
L
l
(5.11)
The conductance and capacitance (per unit area) at high frequencies are o/l
and c/l, respectively, whereas for DC condition (e = 0) they are, respectively, 0
and c/lo. Figure 5.1 depicts the theoretical conductance and capacitance curves for
a 100 mM NaCl solution (o = 1.1 S/m, c = 78c
0
, k
1
= 0.96 nm [2]) using the
experimental set up as described in Figure 5.2 with L = 15 cm and l = 9 cm. The
conductance proceeds to zero at low frequencies and reaches a constant value of
12.22 S/m
2
at a frequency about 6 Hz. In contrast, the capacitance drops from 7.99
F/m
2
at low frequencies to 7.67 10
9
F/m
2
also at a frequency about 6 Hz.
In this chapter I present a study on the electrical properties of a system that
utilises parallel electrodes (see Figure 5.2) as well as non-parallel electrodes
arranged in the circular formation used in EIT (see Figures 4.3 and 4.5).
Measurements of the impedance magnitude and phase of the system, with and
without objects in the chamber, were undertaken using the four-terminal method
over a range of frequencies (200 mHz 90 kHz). Three full spectra of impedance
were acquired for each electrode configuration. The measurements of the
electrolyte-only system were aimed to mimic the homogeneous system used as the
starting point in the EIT reconstruction procedures. And, therefore, it is important
121
to observe the capability and performance of the BULFIS machine I employed
during the course of the project.
(a)
(a)
(b)
Figure 5.1. The theoretical plots of the (a) conductance and (b) capacitance
of a 100 mM NaCl solution (see text for explanation). Please note that the
frequency axis is on a log scale.
10
-2
10
0
10
2
10
4
0
5
10
15
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
frequency (Hz)
10
-2
10
0
10
2
10
4
-1
0
1
2
3
4
5
6
7
8
9
10
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
frequency (Hz)
122
5.2 Materials and Methods
5.2.1 The Experimental Chambers
The experimental systems employed a square chamber and a circular
chamber.
The square system comprised a square perspex chamber of inner
dimension of 150 mm (w) 150 mm (l) 20 mm (h). It was equipped with a pair
of flat current electrodes made from copper fitted to the chamber walls. The
voltage electrodes were silver wires of 1 mm in diameter and inserted into a pair
of movable holders, which sat on a movable slide on top of the chamber as
depicted in Figure 5.2. The set up was similar to that used by Ackmann and Seitz
[3].
The circular system consisted of a circular chamber with the inner
diameter of 300 mm and height of 350 mm (see Figure 4.3) and electrode
arrangement as depicted in Figure 4.5.
The square chamber was filled with 50 mM, 100 mM and 150 mM NaCl
solutions to the level just below the voltage electrode holders (about 8 mm high),
so it required about 180 ml of solution. The circular chamber was filled with the
same solution and concentrations up to 30 mm high.
123
Figure 5.2. The experimental chamber. (a) Top view and (b) side view. The
sliding V-electrode holder allowed the electrodes to be positioned anywhere
in the chamber, while the I-electrodes were fixed on the wall. L was the
distance between the current electrodes and l the voltage electrodes
separation.
5.2.2 Coating the Silver Voltage Electrodes
Prior to the measurements, the silver electrodes were coated with silver
chloride to make Ag/AgCl electrodes. The surfaces of the silver electrodes were
cleaned with ethanol before being coated with silver chloride to remove any oil or
grease. The coating was done by applying a ~ +2.5 volts from a battery to the
electrodes (the positive terminal of the battery was connected to the silver wire
and the negative terminal to the copper plate) for about 2 minutes. The polarity
was reversed for a brief time (about 510 seconds) to remove any loose and
150 mm
150 mm
current
electrodes
voltage
electrodes
20 mm
solution
level
solution
(b)
(a)
sliding holder
L
150
150
0
0
l
124
improper silver chloride coating and then returned to the normal polarity for
another 2 minutes. In this way, a rigid coating of uniform thickness was obtained.
5.2.3 Impedance Measurements
The impedance measurements were carried out using a four-terminal
method covering a frequency range of 200 mHz up to 90 kHz. Three full spectra
were acquired. In the square system the voltage electrodes were positioned at the
centre (75 mm from the 0 edge) and 35 mm off-centre (110 mm from the 0 edge)
with the separation fixed at 90 mm (at 30 mm and 120 mm from the 0 edge), i.e. l
= 90 mm in Figure 5.2(b). The current electrodes were attached to the chamber
wall with a separation of 150 mm (L in Figure 5.2). The choice of only one off-
centre voltage-electrode position was hased on the assumption that the chamber
was symmetrical, and thus, similar results would have been obtained when
positioned off-centre on the other side of the chamber (i.e. at 40 mm from the 0
edge).
125
5.3 Electrical Properties of the Chamber Systems
5.3.1 The Square Chamber System
A good manner to present the impedance data is by plotting the measured
impedance against the log of the frequency [4]. Figure 5.3 shows plots of the
impedance magnitude and phase of the chamber system when filled with 50 mM
NaCl solution as a function of frequency. The solid lines in the graph join the
average value of the three measurements for each electrode position. There was no
object in the solution for these measurements; therefore, the data reflect the
impedances of the NaCl solution, geometry of the chamber and the position of the
electrodes.
The magnitudes of the impedance at the lowest frequency range from 153
O (35 mm off-centre) to 161 O (centered). The magnitudes of the impedance for
the situation when the electrodes were 35 mm off-centre can be seen (Fig 5.3) to
exhibit a dependency on frequency not exhibited when electrodes were located at
the centre. However, the magnitude converges at high frequencies toward the
value obtained when the voltage electrodes were located at the centre of the
chamber. Further, the phases of the impedance exhibited a dependence on the
position of the electrodes that did not converge at high frequencies.
The dispersions of the magnitudes indicate that the solution is not a pure
resistance. This was also supported by measurements of phase, which would be
zero if the solution were indeed purely resistive. In fact, the phase angles were
frequency dependent at lower and higher regions of the frequency range. The
126
magnitude and phase dispersions reveal dependence on the position of the voltage
electrodes.
For higher concentrations (e.g. 100 mM and 150 mM) the curves of the
impedance magnitudes and the phase angles are depicted in Figures 5.4 and 5.5,
respectively. In contrast to measurements obtained for the 50mM solution, the
magnitude of the impedance appeared relatively constant throughout the whole
frequency range. However, the phase angle exhibited a dependency on electrode
position that was more pronounced at the lower and higher regions of the
frequency range.
127
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
140
145
150
155
160
165
170
175
180
i
m
p
e
d
a
n
c
e

(
o
h
m
s
)
freq (Hz)
35 mm off-centre
centre
(a)
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-1
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
i
m
p
e
d
a
n
c
e

p
h
a
s
e

(
d
e
g
r
e
e
s
)
freq (Hz)
35 mm off-centre
centre
(b)
Figure 5.3. Curves of impedance magnitude (a) and its phase (b) as a function of
frequency for two positions of the voltage-electrodes: centre and 35 mm off-
centre. The chamber was filled with 50 mM NaCl. No other objects were present
in the chamber.
128
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
105
110
115
120
i
m
p
e
d
a
n
c
e

(
o
h
m
s
)
freq (Hz)
35 mm off-centre
centre
(a)
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-1
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
i
m
p
e
d
a
n
c
e

p
h
a
s
e

(
d
e
g
r
e
e
s
)
freq (Hz)
35 mm off-centre
centre
(b)
Figure 5.4. Curves of impedance magnitude (a) and its phase (b) similar to Figure
5.3 when the chamber was filled with 100 mM NaCl.
129
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
65
70
75
80
85
90
i
m
p
e
d
a
n
c
e

(
o
h
m
s
)
freq (Hz)
35 mm off-centre
centre
(a)
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-1
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
i
m
p
e
d
a
n
c
e

p
h
a
s
e

(
d
e
g
r
e
e
s
)
freq (Hz)
35 mm off-centre
centre
(b)
Figure 5.5. Curves of impedance magnitude (a) and its phase (b) similar to Figure
5.3 when the chamber was filled with 150 mM NaCl.
130
It is also important to look at the accuracy of the measurements. The
degree of the accuracy gives the confidence level of the results of the
measurements. The accuracy here will be expressed as the signal-to-noise ratio
(SNR) of the signals sampled to determine the impedance. For these experiments
the SNRs of all measurements and all concentrations of solution are shown in
Figure 5.6. The SNR values vary within the frequency frame with the best SNR
being at the frequency of 500 Hz.
(c)
(a) (b)
Figure 5.6. SNRs of the measurements over
the frequency range. (a) for the 50 mM
solution, (b) for the 100 mM solution and (c)
for the 150 mM solution.
10
0
10
5
30
40
50
60
70
s
a
m
p
l
e

S
N
R

(
d
B
)
freq (Hz)
35 mm off-centre
centre
10
0
10
5
30
40
50
60
70
s
a
m
p
l
e

S
N
R

(
d
B
)
freq (Hz)
35 mm off-centre
centre
10
0
10
5
30
40
50
60
70
s
a
m
p
l
e

S
N
R

(
d
B
)
freq (Hz)
35 mm off-centre
centre
131
An alternative representation of the electrical characteristics of a system is
to express impedance (Z) in terms of the conductance and capacitance as a
function of the frequency. The conductance, Re[1/Z]/A and capacitance
Im[1/Z]/e/A were calculated from the measured impedance magnitudes and phase
angles, the angular frequency e and the bulk area of the current injecting
electrodes, The effects of the micro-roughness of the current injecting electrodes
were not considered .
Figure 5.7 shows the conductance and capacitance curves for the chamber
system filled with 50 mM NaCl solution whose impedance curves are presented in
Figure 5.3. The conductance is, in fact, the reciprocal of the impedance magnitude
multiplied by the cosine of its phase angle. Since the phase angles were very
small, cos | was close to unity over most of the frequency range. Hence, it would
be predicted that for this system the dispersions of the conductance should
similarly, but inversely, be dependent on the frequency as per the impedance
magnitude. Figure 5.7 shows that this is the case.
Interestingly, the capacitance disperses at frequencies below 50 Hz. This is
consistent with the theoretical predictions derived earlier in Section 5.1 where the
capacitance at low frequencies ( / l c o ) is larger than the dielectric geometrical
capacitance / l c . Indeed, the dispersion increases with decreasing frequency at
the lower end of the frequency range (i.e. 200 mHz). The magnitude of the
dispersion can be seen to increase with the degree to which voltage electrodes
were located from the centre of the chamber. Upon zooming into the frequency
range of 200 Hz 100 kHz (the insets in Figure 5.7(b)) the dispersions remain
clearly visible below 10 kHz. The separation of the capacitance curves for the two
132
voltage electrode positions is very readily distinguished. The difference in the
sign of the dispersions indicates some geometrical asymmetry in the electrode
configuration. This clearly shows that the capacitance is dependent on the position
of the voltage electrodes as well as on the frequency. This is a direct consequence
of the fact that the phase angle at low frequencies varied with frequency and the
capacitance is deduced from the impedance magnitude and sin|. The latter
quantity is small and sensitive to variation in |. The dispersion in capacitance at
low frequencies, that in some cases yieldeda negative value, was attributable to
the occurrence of the electrodiffusive standing waves between current
electrodes (see also Chapter 2.2.2).
Results for solution concentrations of 100 mM and 150 mM demonstrated
similar trends (see Figures 5.8 and 5.9) but with the conductance exhibiting a
general independence of frequency compared to that for 50 mM. The capacitance
dispersions, however, retained a similar, frequency dependent, behaviour.
133
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
5.5
6
6.5
7
x 10
-3
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
35 mm off-centre
centre
(a)
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-2
0
2
x 10
-5
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
35 mm off-centre
centre
(b)
Figure 5.7. Dispersions of conductance (a) and capacitance (b) with frequency
for two voltage-electrodes positions: centre and 35 mm off-centre. The chamber
was filled with 50 mM NaCl. No other objects were present in the chamber.
10
2
10
3
10
4
10
5
-4
-3
-2
-1
0
1
x 10
-8
centre
35 mm off-centre
134
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
8.5
8.6
8.7
8.8
8.9
9
9.1
9.2
9.3
9.4
9.5
x 10
-3
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
35 mm off-centre
centre
(a)
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-2
0
2
4
x 10
-5
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
35 mm off-centre
centre
(b)
Figure 5.8. Dispersions of conductance (a) and capacitance (b) with frequency
similar to those of Figure 5.7 when the chamber was filled with 100 mM NaCl.
10
2
10
3
10
4
10
5
-5
0
5
10
15
20
x 10
-9
centre
35 mm off-centre
135
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
0.011
0.0115
0.012
0.0125
0.013
0.0135
0.014
0.0145
0.015
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
35 mm off-centre
centre
(a)
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-4
-2
0
2
4
6
8
10
x 10
-5
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
35 mm off-centre
centre
(b)
Figure 5.9. Dispersions of conductance (a) and capacitance (b) with frequency
similar to those of Figure 5.7 with the chamber was filled with 150 mM NaCl.
10
2
10
3
10
4
10
5
-4
-2
0
2
4
6
8
x 10
-8
centre
35 mm off-centre
136
Since a double layer will form at the interface of a metal in contact with an
electrolyte we can expect changes in the electrical characteristics when a metallic
object is placed in the chamber. Such changes would be expected to be similar at
low frequencies to those for an insulating object as the double layer effectively
insulates the whole object at very low frequencies.. Differences between the
metallic and insulating objects would be expected at high frequencies where the
impedance of the double layer (for the metal object) decreases rapidly with
increasing frequency.
The magnitudes of the measured impedances of a circular metallic object
suspended at the centre of the chamber filled with a 50 mM NaCl solution are
depicted in Figure 5.10 as a function of frequency and compared with magnitudes
in the absence of the object. The dimensions of the suspended circular stainless
steel block were 25 mm in diameter and 8 mm thick (i.e. it occupied ~2.2% of the
chamber space). The voltage electrodes were located at the centre and 35 mm off-
centre (see Figure 5.2). The dispersions observed were similar to those for the
electrolyte alone (see Figure 5.3) at low frequencies but are distinctly different in
the frequency range between 1 kHz and 10 kHz. While the starting points of the
dispersions were spread over a wide range (162 O 178 O) at low frequencies
(<300 Hz), they converged at high frequencies (>10 kHz).
Figure 5.10(b) shows the phase angle of the impedance. The values of
phase angle at low frequencies were also similar to those obtained for the
electrolyte alone and remained so up until a frequency ~ 50 Hz when the
dispersions dipped sharply to reach minima at a frequency of ~ 2000 Hz. The
dispersions then climbed steeply to about 1
o
at the other end of the frequency
137
range. This was not a feature in impedance spectra of the electrolyte alone
without the object.
(a)
(b)
Figure 5.10. Dispersions of the impedance magnitude (a) and its phase (b) with
frequency for two voltage-electrodes positions at the centre and 35 mm off-
centre with The chamber was filled with 50 mM NaCl. The left plots are for the
case when a circular stainless steel block placed in the centre of the chamber.
and the plots on the right are for the case when no object was present in the
solution. The solid lines in the graph are the average values of the three spectra
taken for each electrode position.
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
140
145
150
155
160
165
170
175
180
185
i
m
p
e
d
a
n
c
e

(
o
h
m
s
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
140
145
150
155
160
165
170
175
180
185
i
m
p
e
d
a
n
c
e

(
o
h
m
s
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-6
-5
-4
-3
-2
-1
0
1
2
3
4
i
m
p
e
d
a
n
c
e

p
h
a
s
e

(
d
e
g
r
e
e
s
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-6
-5
-4
-3
-2
-1
0
1
2
3
4
i
m
p
e
d
a
n
c
e

p
h
a
s
e

(
d
e
g
r
e
e
s
)
freq (Hz)
35 mm off-centre
centre
138
(a)
(b)
Figure 5.11. Dispersion of conductance (a) and capacitance (b) with frequency
obtained from the data of Figure 5.10. The left graphs are for the case when a
metallic object was present and the right ones are when the object was absent.
The insets on the capacitance graphs are expanded graphs of the dispersions
over the frequency range above 100 Hz.
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
5.5
6
6.5
7
x 10
-3
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
5.5
6
6.5
7
x 10
-3
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-4
-3
-2
-1
0
1
x 10
-5
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-4
-3
-2
-1
0
1
x 10
-5
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
2
10
3
10
4
10
5
-5
0
5
10
x 10
-8
35 mm off-centre
centre
10
2
10
3
10
4
10
5
-5
0
5
10
x 10
-8
centre
35 mm off-centre
139
Figure 5.11 shows that the conductance of the system containing a circular
stainless steel block disperses at frequencies between 1 kHz to 10 kHz. The
conductance at low frequencies is less than that of the electrolyte alone (right
column of Figure 5.11) at 6.5 S/m, consistent with the presence of a low
conducting double layer formed between metal and electrolyte. Effectively the
metal object is insulated by the ionic double layer. As the frequency increases
the contribution of the capacitive reactance increasingly revealed the conductive
contributions from the metal, which are substantially higher than that of the
double layer and higher than electrolyte alone. Similarly the capacitive dispersions
in the presence of the object (insets in Figure 5.11(b)) also featured a dispersion
between 100 Hz and 10 kHz that did not appear in dispersions of the electrolyte
alone. The capacitance was positive in the presence of the object consistent with
the presence of the double layer at its surface. Within this new domain of
capacitance dispersion, there was now a positive capacitance in the interval 100
Hz to 10 kHz, for both positions of the electrodes. Note that this capacitance was
greater for the electrodes in the central position, a fact consistent with there being
more charge storage in the double layer centrally where current lines were more
normal to the conducting block, than there was in the region of electrolyte at the
edges of the conducting block, where current flow was at an angle to the double
layer.
The graphs of these electrical characteristics for the same object but
immersed in 100 mM and 150 mM concentrations of NaCl are presented in Figure
5.12 and Figure 5.13, respectively. Dispersions of conductance and capacitance
behaved similarly to that of 50 mM. Note that the magnitude of the dispersions of
140
the capacitance increased with increasing concentration consistent with the double
layer thickness decreasing with increasing concentration as predicted by the
Debye theory (see Chapter 2.6)
The results for an insulating object (e.g. a circular perspex block of 25 mm
in diameter) are shown in Figures 5.14 5.16. The dispersions of capacitances
with frequency were very similar to those for the electrolyte alone for a
concentration of 50 mM, opposite to the trend observed for the metal object. The
conductance values were smaller than those of electrolyte alone consistent with
the displacement of electrolyte in the presence of the object. This decrease in the
bulk conductance in the presence of the insulating object is the only indicator of
the presence of the object at 50 mM concentration. At higher electrolyte
concentrations (Figures 5.15 and 5.16) the capacitance dispersions in the presence
of the object were distinctly different at low frequencies and in the middle
frequency range (see insets to Figures 5.15 and 5.16). The difference was also
larger at the centre than off-centre indicating that structural information may be
derived from the capacitance dispersions.
141
(a)
(b)
Figure 5.12. Dispersion of conductance (a) and capacitance (b) with frequency
for a circular stainless steel block suspended in the centre of the chamber filled
with 100 mM NaCl. The left graphs are when the metallic object was present and
the right ones are when the object was absent. The insets on the capacitance
graphs are expanded views of the dispersions over the frequency range above
100 Hz.
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
8
8.5
9
9.5
10
10.5
11
x 10
-3
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
8
8.5
9
9.5
10
10.5
11
x 10
-3
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-2
0
2
4
x 10
-5
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-2
0
2
4
x 10
-5
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
2
10
3
10
4
10
5
0
2
4
6
x 10
-8
centre
35 mm off-centre
10
2
10
3
10
4
10
5
0
2
4
6
x 10
-8
35 mm off-centre
centre
142
(a)
(b)
Figure 5.13. Dispersion of conductance (a) and capacitance (b) with frequency for a
circular stainless steel block suspended in the centre of the chamber filled with 150
mM NaCl. The left graphs are when the metallic object was present and the right
ones are when the object was absent. The insets on the capacitance graphs are
expanded views of the dispersions over the frequency range above 100 Hz.
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
0.011
0.0115
0.012
0.0125
0.013
0.0135
0.014
0.0145
0.015
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
0.011
0.0115
0.012
0.0125
0.013
0.0135
0.014
0.0145
0.015
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-4
-2
0
2
4
6
8
10
x 10
-5
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-4
-2
0
2
4
6
8
10
x 10
-5
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
2
10
3
10
4
10
5
-2
0
2
4
6
8
x 10
-8
centre
35 mm off-centre
10
2
10
3
10
4
10
5
-2
0
2
4
6
8
x 10
-8
centre
35 mm off-centre
143
(a)
(b)
Figure 5.14. Dispersion of conductance (a) and capacitance (b) with frequency
for a circular perspex block suspended in the centre of the chamber filled with 50
mM NaCl. The left graphs are when an insulating object was present and the
right ones are when the object was absent. The insets on the capacitance graphs
are expanded views of the dispersions over the frequency range above 100 Hz.
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
5.6
5.8
6
6.2
6.4
6.6
x 10
-3
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
5.6
5.8
6
6.2
6.4
6.6
x 10
-3
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-3
-2.5
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
x 10
-5
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-3
-2.5
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
x 10
-5
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
2
10
3
10
4
10
5
-4
-3
-2
-1
0
1
x 10
-8
centre
35 mm off-centre
10
2
10
3
10
4
10
5
-4
-3
-2
-1
0
1
x 10
-8
35 mm off-centre
centre
144
(a)
(b)
Figure 5.15. Dispersion of conductance (a) and capacitance (b) with frequency
for a circular perspex block placed at the centre of the chamber filled with 100
mM NaCl. The left graphs are when an insulating object was present and the
right ones are when the object was absent. The insets on the capacitance graphs
are expanded views of the dispersions over the frequency range above 100 Hz.
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
8
8.5
9
9.5
x 10
-3
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
8
8.5
9
9.5
x 10
-3
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-2
0
2
4
x 10
-5
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-2
0
2
4
x 10
-5
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
2
10
3
10
4
10
5
-2
-1
0
1
2
3
4
x 10
-8
centre
35 mm off-centre
10
2
10
3
10
4
10
5
-2
-1
0
1
2
3
4
x 10
-8
centre
35 mm off-centre
145
(a)
(b)
Figure 5.16. Dispersion of conductance (a) and capacitance (b) with frequency
for a circular perspex block placed at the centre of the chamber filled with 150
mM NaCl. The left graphs are when an insulating object was present and the
right ones are when the object was absent. The insets on the capacitance graphs
are expanded views of the dispersions over the frequency range above 100 Hz.
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
0.011
0.0115
0.012
0.0125
0.013
0.0135
0.014
0.0145
0.015
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
0.011
0.0115
0.012
0.0125
0.013
0.0135
0.014
0.0145
0.015
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-4
-2
0
2
4
6
8
10
12
x 10
-5
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-4
-2
0
2
4
6
8
10
12
x 10
-5
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
35 mm off-centre
centre
10
2
10
3
10
4
10
5
-4
-2
0
2
4
6
x 10
-8
centre
35 mm off-centre
10
2
10
3
10
4
10
5
-4
-2
0
2
4
6
x 10
-8
centre
35 mm off-centre
146
5.3.2 The Circular Chamber System
Figures 5.17 and 5.18 show impedance measurements obtained using a
circular chamber filled with NaCl solution and a metal object (aluminium block C
50 mm and 35 mm thick) suspended at the centre of the chamber. The impedance
curves (Figure 5.17(a)), as expected, show the presence of dispersions at low
frequencies (f < 1000 Hz) that were similar to those obtained using the square
chamber. The steepness of the dispersions slope was dependent on the
concentration of the surrounding solution. Low concentration (c = 50 mM) gave
the steepest slope, which flattened as the concentration increased. A high
concentration (c =. 150 mM) gave the shallowest slope. Note also that the
dispersion region widened with increasing concentration. The region extended
from 0.25 Hz up to 300 Hz for 50 mM, up to 600 Hz for 100 mM and up to 1000
Hz for 150 mM. The behaviour of the conductance reflected that of the impedance
magnitude since the conductance is approximately the reciprocal of the magnitude
given the small phase angles that were measured over the entire frequency range.
Nonetheless, although the phase angles are small they provide important
information about the object as reflected in the capacitance spectra.
The capacitance curves (Figure 5.18(b)) reveal dispersions over the
frequency region below 20 Hz. This behaviour was also observed for the same
object using the square chamber. However, differences in the capacitance
dispersions, in the presence and absence of the metal object were most evident in
the frequency range 100 4000 Hz (see insets to Figure 5.18) where dispersions
arising from the ionic double layer are commonly observed (see Gaedt et al. [5]
page 174). This is also the frequency range where the conductance (in the
147
presence and absence of the object) disperses. This also is the expected behaviour
for a metal object that forms a double layer with an electrolyte.
The electrical characteristics of the system when an insulator (e.g. an
empty glass beaker of 50 mm in diameter) was suspended at the same location as
the metal object are depicted in Figure 5.19. At low frequencies the behaviours are
similar to the metal system, but differences in the conductance dispersions appear
in the presence and absence of the object. And the conductance in the presence of
the insulating object remained smaller than that in the absence of the object over
the whole frequency range.
148
(a)
(b)
Figure 5.17. Dispersions of the impedance magnitude (a) and its phase angle (b)
with frequency for a circular aluminium block placed in the centre of a circular
chamber filled with different NaCl solutions (left column) and when no object
was present (right column).
10
0
10
5
0
20
40
60
80
100
i
m
p
e
d
a
n
c
e

(
o
h
m
s
)
freq (Hz)
50 mM
100 mM
150 mM
10
0
10
5
0
20
40
60
80
100
i
m
p
e
d
a
n
c
e

(
o
h
m
s
)
freq (Hz)
50 mM
100 mM
150 mM
10
0
10
5
-10
-8
-6
-4
-2
0
2
i
m
p
e
d
a
n
c
e

p
h
a
s
e

(
d
e
g
r
e
e
s
)
freq (Hz)
50 mM
100 mM
150 mM
10
0
10
5
-10
-8
-6
-4
-2
0
2
i
m
p
e
d
a
n
c
e

p
h
a
s
e

(
d
e
g
r
e
e
s
)
freq (Hz)
50 mM
100 mM
150 mM
149
(a)
(b)
Figure 5.18. Dispersions of the conductance (a) and capacitance (b) with frequency
for a circular aluminium block suspended in different NaCl solutions for a circular
chamber (left column) and when no object was present (right column).
10
0
10
5
0.01
0.02
0.03
0.04
0.05
0.06
0.07
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
50 mM
100 mM
150 mM
10
0
10
5
0.01
0.02
0.03
0.04
0.05
0.06
0.07
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
50 mM
100 mM
150 mM
10
0
10
5
-0.5
0
0.5
1
1.5
2
2.5
3
x 10
-3
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
10
0
10
5
-0.5
0
0.5
1
1.5
2
2.5
3
x 10
-3
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
10
2
10
3
10
4
10
5
0
1
2
3
x 10
-6
50 mM
100 mM
150 mM
10
2
10
3
10
4
10
5
0
1
2
3
x 10
-6
50 mM
100 mM
150 mM
150
(a)
(b)
Figure 5.19. Dispersions of the conductance (a) and capacitance (b) with frequency
for an empty beaker glass suspended in different NaCl solutions for a circular
chamber (left column) and when no object was present (right column).
10
0
10
5
0.01
0.02
0.03
0.04
0.05
0.06
0.07
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
50 mM
100 mM
150 mM
10
0
10
5
0.01
0.02
0.03
0.04
0.05
0.06
0.07
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
50 mM
100 mM
150 mM
10
0
10
5
-0.5
0
0.5
1
1.5
2
2.5
3
x 10
-3
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
10
0
10
5
-0.5
0
0.5
1
1.5
2
2.5
3
x 10
-3
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
10
2
10
3
10
4
10
5
-5
0
5
10
15
x 10
-7
50 mM
100 mM
150 mM
10
2
10
3
10
4
10
5
-5
0
5
10
15
x 10
-7
50 mM
100 mM
150 mM
151
The comparisons between the metal (aluminium) and insulator (glass)
systems are displayed in Figure 5.20. The graphs reveal the dispersions at low
frequencies for both systems. They originate from the interfaces between the
electrolyte and the objects. Both conductance dispersions commence from the
same low value at the lower end of the frequency range, indicating that the two
different objects were not distinguishable from impedance measurements in this
frequency range. However, at higher frequencies the conductance of the system
containing the metal object were much larger than those of the system containing
the insulating object, consistent with the conductivity of the metal being
substantially larger than that of the insulator.
The inability to distinguish between dispersions obtained for the two
objects at low frequencies is understandable since the double layer on the
interface acts as an almost perfect insulator wrapped around the metal object
yielding conductance properties similar to those of the insulating object at low
frequencies. As the frequency increases the conductance dispersions diverge as
the impedance of the double layer decreases with increasing frequency (see
Chapter 2.6). The conductance at high frequencies is then derived from the bulk
electrolyte and the metal for the metal object, or the electrolyte and the insulating
object (and for the latter case will yield an overall conductance which is lower at
high frequencies). Note also that the capacitance dispersions in the presence of the
metal object were generally larger than those dispersions in the presence of the
insulator. This can be attributed to the presence of the double layer at the metal-
electrolyte interface which in its simplest form can be modelled as a parallel plate
capacitor.
152
When comparing the effect of solution concentrations on conductance
characteristics (see the left column of Figure 5.20) it is noted that the starting
points of the dispersions increase as the concentration increases. Similar trends
apply to the final values of the conductance (when the dispersions reach the
plateau). Both trends are summarised in Table 5.1 although at a concentration of
150 mM the conductance dispersions had not plateaued. The difference between
the minimum and maximum values reflects that the thickness of the double layer
(i.e. the Debye length) was larger at low concentrations resulting in a smaller
double layer capacitance and conductance than at higher concentrations.
Table 5.1. Trend of the values of the measured conductance for different
concentrations of the surrounding NaCl solution.
Concentration Start Glass Aluminium Difference
(mM) (mS.m
-2
) (mS.m
-2
) (mS.m
-2
) (mS.m
-2
)
50 15.5 19 21 3
100 28 32 38 6
150 40 50 60 10
The dielectric geometrical capacitance properties of the double layer are
revealed in the frequency range of 100 Hz 4000 Hz where the capacitance
values for the metal object were much larger than those values for the glass object.
The frequency at which the capacitance curves converge increased with
increasing concentration, again consistent with Debye equation for the width of
the double layer.
153
(a)
(b)
(c)
Figure 5.20. Comparisons of the dispersions of the conductance (left column)
and capacitance (right column) with frequency for aluminium and perspex block
placed in a circular chamber filled with (a) 50 mM, (b) 100 mM and (c) 150 mM
NaCl solutions.
10
0
10
5
0.01
0.02
0.03
0.04
0.05
0.06
0.07
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
metal
glass
10
0
10
5
0.01
0.02
0.03
0.04
0.05
0.06
0.07
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
metal
glass
10
0
10
5
0.01
0.02
0.03
0.04
0.05
0.06
0.07
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
2
)
freq (Hz)
metal
glass
10
0
10
5
-5
0
5
10
15
20
x 10
-4
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
10
0
10
5
-5
0
5
10
15
20
x 10
-4
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
10
0
10
5
-5
0
5
10
15
20
x 10
-4
c
a
p
a
c
i
t
a
n
c
e

(
F
/
m
2
)
freq (Hz)
10
2
10
3
10
4
10
5
0
1
2
3
x 10
-6
metal
glass
10
2
10
3
10
4
10
5
0
1
2
3
x 10
-6
metal
glass
10
2
10
3
10
4
10
5
0
1
2
3
x 10
-6
metal
glass
50 mM
100 mM
150 mM
154
5.4 Discussions
Four-terminal measurements of the NaCl electrolyte alone revealed that
the conductance and capacitance dispersed with frequency. This can be partially
attributed to the fact that the electrodes used to inject the current were partially
blocking.
The conductance and capacitance per unit area of a blocking electrode at
very low frequencies are given by (from Equation (5.10))
( 0) 0 and ( 0)
c
e e
o
= = = = G C
l
(5.12)
where l is the voltage electrode spacing (see Figure 5.2) and o , which is given by
Equation (5.11) simplifies to
( ) / 2
2
1

o

~ <
L l
e
l
(5.13)
where L is the distance between the current electrodes and is the Debye length
of the solution.
Only at high frequencies do the conductance and capacitance
measurements yield electrical properties of the electrolyte,
( 0) and ( 0)
o c
e e > = > = G C
l l
(5.14)
155
Equations (5.12) and (5.14) reveal substantial dispersions in the
conductance from zero to o/l as well as the capacitance from c/lo to c/l for
increasing frequency.
Although Ag/AgCl electrodes were used as voltage electrodes in my
experiments, remnant effects arising from non-blocking electrodes will be
expected from using copper for the current electrodes. Chapter 4.2 reveals that
although copper reacts with chloride to form CuCl
2
unlike AgCl, which is
adsorbed to the surface of a silver electrode, CuCl
2
precipitates in solution,
accumulating at the bottom of the chamber.
Another phenomenon attributable to the dispersion characteristic of an
electrolyte at low frequencies is the spatial distortion of the current flow patterns
by the voltage electrodes. The extreme case arises for blocking electrodes when a
double layer forms at the interface of the two phases. At low frequencies the
double layer diverts current around the electrode, thus the measurement is not that
for a spatially uniform current density. In contrast, at high frequencies, current
passes through the voltage electrodes (see Figure 5.21) creating another non-
uniform current density pattern. Thus the measurements of voltage using such
electrodes yield dispersions of the impedance with frequency that would not be
expected from ideal point electrodes that do not distort the local current density.
Gaedt [7] has shown that the geometrical orientation of the current
electrodes had no substantial effects on the overall impedance measurements.
However, he observed that the spacing of the voltage electrodes and their
geometry influenced the measurements of the impedance.
156
Figure 5.21. In the four-terminal measurement technique, the currents entering
and leaving the electrode through points of unequal polarisation impedance
creates the potential polarisation [6].
Remnants of these two phenomenological effects were observed
principally as the dispersions of capacitance (also known as the o-low dispersion
[1]) when measuring the impedance of only the electrolyte and when objects were
immersed in the electrolytes.
When a metal object was present, the dispersions of the capacitance were
larger than those of the electrolyte alone, indicating additional contributions from
the object. Figures 5.11, 5.12 and 5.13 show that the conductance at low
frequencies was below those of the electrolyte, only in the presence of the metal
object. This indicates that the object was enveloped by a high impedance layer
that electrically isolates the object from the solution. However, at high frequencies
the conductance exceeded those of the electrolyte alone indicating that the effect
of the layer is removed and the metal conductance contributes to the measured
conductance. Such trends are consistent with the layer being an ionic double layer
that forms at the interface between a metal and electrolyte, which is modelled as a
157
parallel plate capacitor of thickness given by the Debye length in its simplest
form.
At low frequencies the capacitive reactance dominates in the measured
impedance but approaches zero at high frequencies where its contribution to the
measured impedance becomes negligible. This dispersion is most pronounced
between 50 Hz 10 kHz and absent in capacitance dispersions of the electrolyte
alone. This suggests that dispersions below 50 Hz are principally due to the
remnants of the phenomenological effects of partially blocking electrodes and the
effect of standing wave.
The system containing an insulator object (a circular perspex block in this
case) behaved relatively similarly to that of the electrolyte alone (see Figures 5.14,
5.15 and 5.16). Conductance measurements in the presence of the insulating
object were less than those measurements of the electrolyte alone, which can be
understood because the perspex conductance is much lower than that of the NaCl
solution. The presence of the object basically blocked the passage of the current
thus requiring the current to flow around the object. The larger current paths and
the replacements of electrolyte with the insulating object yielded conductance
values smaller than those obtained in the absence of the object.
The capacitance dispersions in the presence and absence of the object
revealed little difference at 50 mM. However, at higher concentrations differences
were observed that were attributed to the presence of the insulating object. But
these differences were smaller than those observed for the metal object. This was
due to the fact that there was no frequency dependent ionic double layer formed at
158
the insulating object-solution interface and in any case no current could pass
through any such layer because the object itself was an insulator.
In the square chamber different dispersions were obtained for voltage
electrodes located at the centre an off-centre. The dispersions obtained with
electrodes off-centre more closely resembled those of the electrolyte alone. This
was expected because the objects did not extend all the way across the
experimental chamber, leaving predominantly areas of electrolyte between the
voltage electrodes. The centre and off-centre measurements, therefore, provide
information on the size of the object. Presumably, further off-centre
measurements could potentially provide more detailed information of the size,
shape and location of the objects.
The measurements using the circular chamber yielded similar results as
square chamber using central electrodes. Chapter 7 deals with using off-centre
electrodes.
The metal and glass materials were very easily distinguishable at low
frequencies from their differing electrical properties. This is the most important
observation since other researchers in the area of EIT have overlooked this low
frequency domain and have generally used frequencies above 10 kHz.
159
5.5 Conclusions
Impedance measurements using voltage electrodes strategically placed in
square and circular chambers were performed using the BULFIS impedance
spectrometer. These measurements were able to distinguish between conducting
and insulating objects at frequencies well below the commonly used frequencies
in EIT imaging (usually > 10 kHz).
This study reveals the potential of low frequency impedance
measurements and especially capacitance measurements to determine the size,
shape, location and electrical properties of objects. Chapter 7 investigates this
potential further using a common EIT algorithm.
160
References
1. Coster, H.G.L. and T.C. Chilcott. (1999). The characterization of
membranes and membrane surfaces using impedance spectroscopy, in
Surface chemistry and electrochemistry of membranes, Vol. 79, T.S.
Sorensen (Ed.). Marcel Dekker, Inc. New York. pp. 749-792.
2. Sheetz, M. (2002). Cell as a machine: Membrane asymmetry and surface
potential. Internet source, accessed 3 May 2006. Available from:
http://www.columbia.edu/cu/biology/courses/w3150/lecture10.pdf.
3. Ackmann, J.J. and M.A. Seitz. (1984). Methods of complex impedance
measurements in biologic tissue. CRC Critical Reviews in Biomedical
Engineering, 11(4): 281-311.
4. Coster, H.G.L., T.C. Chilcott, and A.C.F. Coster. (1996). Impedance
spectroscopy of interfaces, membranes and ultrastructures.
Bioelectrochemistry and Bioenergetics, 40: 79-98.
5. Gaedt, L., T.C. Chilcott, M. Chan, T. Nantawisarakul, A.G. Fane, and
H.G.L. Coster. (2002). Electrical impedance spectroscopy characterisation
of conducting membranes II. Experimental. Journal of Membrane Science,
195(2): 169-180.
6. Schwan, H.P. (1992). Linear and nonlinear electrode polarization and
biological materials. Annals of Biomedical Engineering, 20: 269-288.
7. Gaedt, L.P. (1999). Impedance spectroscopy characterisation of synthetic
membranes and processes, PhD Thesis. in Faculty of Engineering.
Univesity of New South Wales. Sydney, Australia.
6
COMPLEX ELECTRICAL
IMPEDANCE TOMOGRAPHY
6.1 Introduction
Electrical Impedance Tomography (EIT) aims to recover the spatial
distributions of the electrical properties within the system under investigations
from impedance measurements made via peripheral electrodes distributed around
the system.
In this chapter I present the reconstructed images from simulated data as
well as data from experiments. The dimensions (e.g. width) of the reconstructed
images are compared to those of the object in order to assess the accuracy of the
reconstruction algorithms. The widths of the objects were determined using the
Full-Width at Half-Maximum (FWHM) method described in Chapter 3. Images
reconstructed from experimental data are also presented and compared with
images from simulation results.
162
6.2 Materials and Methods
The first part of this chapter demonstrates image reconstructions from
peripheral complex impedances (e.g. the conductivity and scaled permittivity).
The simulation was aimed to test the capability of the EIDORS (Electrical
Impedance and Diffuse Optical Reconstruction Software) to perform image
reconstructions from simulated data.
The simulations were based on a circular insulating object placed at the
centre of the chamber (the diameter of the chamber in the simulation was 300
mm) filled with saline. The saline conductivity (o
soln
) was set to 1.1 S/m and its
relative permittivity (c
r,soln
) 78. The conductivity (o
obj
) and relative permittivity
(c
r,obj
) of the object were 0 S/m and 2, respectively. At a frequency of 1.12 Hz, the
susceptivity (i.e. ec
0
c
r
) of the electrolyte and object were 0.47710
8
S/m and
0.01210
8
S/m, respectively (see Figures 6.1a and b). The reconstructed images
are presented in Figures 6.1c and d.
In the experiments, the data were collected using electrodes configured in
the adjacent current pattern (see Figure 4.5). Current was injected via two
neighbouring electrodes and the voltage responses were measured using adjacent
pairs of electrodes located on the circumference of the cylindrical chamber.
Sodium chloride (NaCl) solutions of concentrations of 50 mM, 100 mM and 150
mM were used as electrolyte. The objects used were a perspex cylinder (of 100
mm in diameter) and an aluminium block (of 130 mm in diameter). The object
was placed at the centre of the chamber of 300 mm in diameter and immersed in
the electrolytes. Impedance measurements were made at 1.12 Hz, 15.56 Hz,
163
282.38 Hz, 4551.5 Hz and 77,712 Hz. The impedance data were then used to
reconstruct the images using the EIDORS.
6.3 Simulated Reconstructions
(a) Actual conductivity
distribution.
(b) Actual susceptivity
distribution.
(c) Reconstruction of
conductivity distribution.
(d) Reconstruction of susceptivity
distribution.
Figure 6.1. Simulation of reconstruction of admittivity distributions for a circular insulating object
(40 mm C) immersed in electrolyte at 1.12 Hz. The conductivity of the saline was 1.1 S/m and its
relative permittivity was 78. The conductivity of the object was 0.0 S/m and its relative
permittivity was 2.
Figures 6.1c and d depict results of simulating an insulating object of 40
mm in diameter placed at the centre of the chamber of 300 mm diameter. The
reconstructions were run for five iterations, which converged at the third iteration.
0.7
0.8
0.9
1
1.1
0.3
0.35
0.4
0.45
S/m
S/m
10
8
S/m
10
8
S/m
164
The electrical characteristics of the original system (i.e. the electrolyte and object)
has been described in Section 6.2. The algorithm predictions approach the
conductivity and susceptibility of the electrolyte only at distances remote from the
object and underestimate the conductivity and susceptibility of the object in the
vicinity of the object.
Figure 6.2 depicts the spatial resolution graphs of the simulations shown in
Figure 6.1. It illustrates the degree to which the algorithm underestimates the
properties of the object and overestimates the size of the object. The width of the
reconstructed object was measured using the full-width at half-maximum
(FWHM) method discussed in Chapter 3.4. The FWHM for the simulations in this
work is compared to others and summarised in Table 6.1. The FWHM of the
EIDORS is slightly greater than those of Ider et al. [1] and Wexler et al. [2], who
both used a 30 mm diameter object inside a 300 mm diameter chamber. This
indicates that EIDORS algorithms underestimate the recovered electrical values,
while on the other hand they overestimate the size of the recovered object.
Table 6.1. The Full-Width at Half-Maximum (FWHM) result of the simulation
compared to those of Ider et al. [1] and Wexler et al [2].
Ider et al. Wexler et al. This work (EIDORS)
Object position FWHM (%) FWHM (%) FWHM (%)
Central 27 24 33
165
(a)
(b)
(a)
(b)
Figure 6.2. FWHMs for the simulation of (a) conductivity and (b)
susceptivity at 1.12 Hz. The FWHM for both images is 33% of the diameter
of the imaging region (see text for details) compared with 13.3% for the
original distributions. The dotted lines represent the actual admittivity
properties of the system.
-15 -10 -5 0 5 10 15
-0.2
0
0.2
0.4
0.6
0.8
1
1.2
1.4
c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
)
diameter (cm)
-15 -10 -5 0 5 10 15
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
5
x 10
-9
r
e
a
c
t
a
n
c
e

(
S
/
m
)
diameter (cm)
s
u
s
c
e
p
t
a
n
c
e

(
S
/
m
)

c
o
n
d
u
c
t
a
n
c
e

(
S
/
m
)

166
6.4 Reconstruction Images from Impedance
Measurements
The reconstructed images obtained from data measured from a system
with a 130 mm diameter cylindrical aluminium block suspended in the NaCl
saline are shown in Figure 6.3 through Figure 6.5. For simplicity of labelling, the
frequencies used will be denoted as F1, F2, F3, F4 and F5, where F1 = 1.12 Hz,
F2 = 15.56 Hz, F3 = 282.38 Hz, F4 = 4551.5 Hz and F5 = 77,712 Hz,
respectively. Further, the susceptivity (i.e. ec
0
c
r
) is replaced by dielectric
permittivity (i.e. c
0
c
r
) enabling a direct comparison of dielectric images obtained
at these frequencies. The images were obtained from individual reconstruction
process and then plotted on the same scale for all five frequencies. The
conductance images in Figure 6.3 (top row) were from an electrolyte of 50 mM
NaCl. The object is identified by a lower conductance regime than its surrounding
solution at frequencies F1, F4 and F5. Such is consistent with previous
measurements and theoretical analyses (see Chapter 5.3) revealing that the double
layer of high impedance at these frequencies, diverges current around the object.
This is inconsistent with previous analyses revealing that at higher frequencies the
effect of the double layer diminishes whence the object (aluminium) would be
expected to have a substantially higher conductance than the solution. That the
reconstruction imaging predicts the opposite at high frequencies indicates a
weakness in the algorithm used for reconstruction that does not have sufficient
spatial resolution to accommodate a nanometer-scale structure such as a double
layer; i.e. the mesh size is much larger than the thickness of the double layer.
167
Figure 6.3. Reconstructed conductivity distribution (top row) and dielectric
permittivity (capacitance) distribution (bottom row) of a circular aluminium
block (C 130 mm) at the centre in 50 mM NaCl. The frequencies used (from
left to right) were 1.12 Hz, 15.56 Hz, 282.38 Hz, 4551.5 Hz and 77,712 Hz.
The dielectric permittivity (capacitance) distribution images do not show
any significant features to reveal the presence of the object, except at F5, which is
in the frequency regime of present EIT imaging. Nonetheless capacitance images
do predict negative capacitances consistent with previous measurements of the
electrolyte and the electrolyte plus object (see Chapter 5.3.2).
Al
130 mm
300 mm
NaCl solution
10
3
Sm
10
6
Fm
F1 F2 F3 F4 F5
168

Figure 6.4. Reconstructed conductivity distribution (top row) and dielectric
permittivity (capacitance) distribution (bottom row) of a circular aluminium
block (C 130 mm) in the centre in 100 mM NaCl. The frequencies used
(from left to right) were 1.12 Hz, 15.56 Hz, 282.38 Hz, 4551.5 Hz and
77,712 Hz.
Figures 6.4 and 6.5 reveal the emergence of features in the dielectric
permittivity (capacitance) images for higher concentrations that reveal the
presence of the object at low frequencies (except at frequency F3). Interestingly
the capacitance images at the higher electrolyte concentrations register a higher
dielectric permittivity (capacitance) for the object than for the electrolyte,
consistent with the dielectric properties of these phases. Even so, the conductance
images still predict a lower conductance for the object than for electrolyte at high
frequencies, which remains inconsistent with the previous measurements and
theories (see Chapter 5.3.2).
10
3
Sm
10
6
Fm
F1 F2 F3 F4 F5
169
Figure 6.5. Reconstructed conductivity distribution (top row) and dielectric
permittivity (capacitance) distribution (bottom row) of a circular aluminium
block (C 130 mm) in the centre in 150 mM NaCl. The frequencies used
(from left to right) were 1.12 Hz, 15.56 Hz, 282.38 Hz, 4551.5 Hz and
77,712 Hz.
The reconstructed images for the perspex cylinder (C 100 mm) immersed
in 50 mM NaCl show the expected bulk conductance characteristics of an
insulator (i.e. a conductance regime that is lower than that of the surrounding
electrolyte). The conductance and capacitance images at low frequencies (e.g. F1
and F2) do not reveal the circular shape of the object (see Figure 6.6) and its size
begins to emerge at higher frequencies (e.g. F3 and higher). The images at F5
strongly reveal the shape, size and location of the object. Even so, the capacitance
image predicts that the capacitive properties of the perspex (c
r
~ 3) are higher than
those of the surrounding electrolyte (c
r
~ 80) opposite to what would be expected,
having given consideration to the dielectric properties of the two phases. These
10
-3
Sm
10
-6
Fm
F1 F2 F3 F4 F5
170
trends appear to be independent of the concentration of the electrolyte (see
Figures 6.7 and 6.8).
Figure 6.6. Reconstructed conductivity distribution (top row) and dielectric
permittivity (capacitance) distribution (bottom row) of a perspex cylinder (C
100 mm) in the centre in 50 mM NaCl. The frequencies used (from left to
right) were 1.12 Hz, 15.56 Hz, 282.38 Hz, 4551.5 Hz and 77,712 Hz.
Perspex
cylinder
100 mm
300 mm
NaCl solution
10
-3
Sm
10
-6
Fm
F1 F2 F3 F4 F5
171
Figure 6.7. Reconstructed conductivity distribution (top row) and dielectric
permittivity (capacitance) distribution (bottom row) of a perspex cylinder (C
100 mm) in the centre in 100 mM NaCl. The frequencies used (from left to
right) were 1.12 Hz, 15.56 Hz, 282.38 Hz, 4551.5 Hz and 77,712 Hz.
Figure 6.8. Reconstructed conductivity distribution (top row) and dielectric
permittivity (capacitance) distribution (bottom row) of a perspex cylinder (C
100 mm) in the centre in 150 mM NaCl. The frequencies used (from left to
right) were 1.12 Hz, 15.56 Hz, 282.38 Hz, 4551.5 Hz and 77,712 Hz.
10
-3
Sm
10
-3
Sm
10
-6
Fm
10
-6
Fm
F1 F2 F3 F4 F5
F1 F2 F3 F4 F5
172
In order to inspect the positive and the negative magnitude peaks of the
conductance and capacitance, the plots of those peaks are given in Figure 6.9. It is
clearly indicated that for the metallic object F4 gives the best contrast in the
images with negative peaks for conductance and positive peaks for capacitance
(Figures 6.9(a) and 6.9(c), respectively) except for the solution concentration of
50 mM, which did not provide consistent results. Similar images and trends were
also observed for an insulating object (Figures 6.9(b) and 6.9(d)), but the contrast
is seen at F3.
173
(a) (b)
(c) (d)
Figure 6.9. Variations with frequency of peak values of conductivities of (a) a
circular aluminium block (C 130 mm) and (b) a perspex cylinder (C 130 mm);
and peak values of dielectric permittivities (capacitances) of (c) a circular
aluminium block (C 130 mm) and (d) a perspex cylinder (C 130 mm).
10
0
10
2
10
4
-70
-60
-50
-40
-30
-20
-10
0
10
p
e
a
k

c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
2
)
log (freq)
NaCl 50 mM
NaCl 100 mM
NaCl 150 mM
10
0
10
2
10
4
-40
-20
0
20
40
60
p
e
a
k

c
a
p
a
c
i
t
a
n
c
e

(
u
F
/
m
2
)
log (freq)
NaCl 50 mM
NaCl 100 mM
NaCl 150 mM
10
0
10
2
10
4
-70
-60
-50
-40
-30
-20
-10
0
10
p
e
a
k

c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
2
)
log (freq)
NaCl 50 mM
NaCl 100 mM
NaCl 150 mM
10
0
10
2
10
4
-40
-20
0
20
40
60
p
e
a
k

c
a
p
a
c
i
t
a
n
c
e

(
u
F
/
m
2
)
log (freq)
NaCl 50 mM
NaCl 100 mM
NaCl 150 mM
p
e
a
k

c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
)

p
e
a
k

c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
)

p
e
a
k

c
a
p
a
c
i
t
a
n
c
e

(

F
/
m
)
p
e
a
k

c
a
p
a
c
i
t
a
n
c
e

(

F
/
m
)
174
6.5 Estimation of the Spatial Resolution
The spatial resolutions of the reconstructed images shown in Section 6.4
were determined using the FWHM method. All images were obtained for an
object (of both metal and plastic) placed at the centre of the chamber. Thus, the
geometry was symmetrical and conductance and capacitance values of pixels
located the same distance from the centre could be averaged and plotted as a
function of that distance.
The bell-shape curves of the cross-sectional values of each individual
measurement are depicted in Figures 6.10 and 6.11 and the FWHM values
determined from conductance curves are recorded in Table 6.2a and those from
capacitance curves in Table 6.2b. The FWHM calculates the percentage of the
diameter (width) of the curves at their half maximum values compared to the
diameter of the chamber (see Chapter 3.4). For instance, at 1.12 Hz in 50 mM
NaCl the width of aluminium object reflected 31% of the chamber diameter. Table
6.3 lists the resolution expressed as the ratio of the difference between recovered
and actual sizes. Images that appear bigger than the actual object have a value of
more than 100%; and those that appear smaller have a value less than 100%.
175
(a)
(b)
(c)
Figure 6.10. Conductivities (left) and dielectric permittivities (capacitances)
(right) distributions of a circular aluminium block (C 130 mm) located in the
centre of the chamber with solutions of: (a) 50 mM NaCl, (b) 100 mM NaCl
and (c) 150 mM NaCl.
-15 -10 -5 0 5 10 15
-70
-60
-50
-40
-30
-20
-10
0
10
20
c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
2
)
diameter (cm)
1.1176 Hz
15.5645 Hz
282.384 Hz
4,551.5 Hz
77,712 Hz
-15 -10 -5 0 5 10 15
-40
-30
-20
-10
0
10
20
30
40
50
60
70
c
a
p
a
c
i
t
a
n
c
e

(
u
F
/
m
2
)
diameter (cm)
1.1176 Hz
15.5645 Hz
282.384 Hz
4,551.5 Hz
77,712 Hz
-15 -10 -5 0 5 10 15
-70
-60
-50
-40
-30
-20
-10
0
10
20
c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
2
)
diameter (cm)
1.1176 Hz
15.5645 Hz
282.384 Hz
4,551.5 Hz
77,712 Hz
-15 -10 -5 0 5 10 15
-40
-30
-20
-10
0
10
20
30
40
50
60
70
c
a
p
a
c
i
t
a
n
c
e

(
u
F
/
m
2
)
diameter (cm)
1.1176 Hz
15.5645 Hz
282.384 Hz
4,551.5 Hz
77,712 Hz
-15 -10 -5 0 5 10 15
-70
-60
-50
-40
-30
-20
-10
0
10
20
c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
2
)
diameter (cm)
1.1176 Hz
15.5645 Hz
282.384 Hz
4,551.5 Hz
77,712 Hz
-15 -10 -5 0 5 10 15
-40
-30
-20
-10
0
10
20
30
40
50
60
70
c
a
p
a
c
i
t
a
n
c
e

(
u
F
/
m
2
)
diameter (cm)
1.1176 Hz
15.5645 Hz
282.384 Hz
4,551.5 Hz
77,712 Hz
c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
)

c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
)

c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
)

c
a
p
a
c
i
t
a
n
c
e

(

F
/
m
)

c
a
p
a
c
i
t
a
n
c
e

(

F
/
m
)

c
a
p
a
c
i
t
a
n
c
e

(

F
/
m
)

176
(a)
(b)
(c)
Figure 6.11. Conductivities (left) and dielectric permittivities (capacitances)
(right) distributions of a perspex cylinder (C 100 mm) located in the centre of
the chamber with solutions of: (a) 50 mM NaCl, (b) 100 mM NaCl and (c) 150
mM NaCl.
-15 -10 -5 0 5 10 15
-70
-60
-50
-40
-30
-20
-10
0
10
20
c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
2
)
diameter (cm)
1.1176 Hz
15.5645 Hz
282.384 Hz
4,551.5 Hz
77,712 Hz
-15 -10 -5 0 5 10 15
-40
-30
-20
-10
0
10
20
30
40
50
60
70
c
a
p
a
c
i
t
a
n
c
e

(
u
F
/
m
2
)
diameter (cm)
1.1176 Hz
15.5645 Hz
282.384 Hz
4,551.5 Hz
77,712 Hz
-15 -10 -5 0 5 10 15
-70
-60
-50
-40
-30
-20
-10
0
10
20
c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
2
)
diameter (cm)
1.1176 Hz
15.5645 Hz
282.384 Hz
4,551.5 Hz
77,712 Hz
-15 -10 -5 0 5 10 15
-40
-30
-20
-10
0
10
20
30
40
50
60
70
c
a
p
a
c
i
t
a
n
c
e

(
u
F
/
m
2
)
diameter (cm)
1.1176 Hz
15.5645 Hz
282.384 Hz
4,551.5 Hz
77,712 Hz
-15 -10 -5 0 5 10 15
-70
-60
-50
-40
-30
-20
-10
0
10
20
c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
2
)
diameter (cm)
1.1176 Hz
15.5645 Hz
282.384 Hz
4,551.5 Hz
77,712 Hz
-15 -10 -5 0 5 10 15
-40
-30
-20
-10
0
10
20
30
40
50
60
70
c
a
p
a
c
i
t
a
n
c
e

(
u
F
/
m
2
)
diameter (cm)
1.1176 Hz
15.5645 Hz
282.384 Hz
4,551.5 Hz
77,712 Hz
c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
)

c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
)

c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
)

c
a
p
a
c
i
t
a
n
c
e

(

F
/
m
)

c
a
p
a
c
i
t
a
n
c
e

(

F
/
m
)

c
a
p
a
c
i
t
a
n
c
e

(

F
/
m
)

177
Table 6.2a. FWHM for aluminium (C 130 mm) and perspex cylinder (C 100
mm) calculated from the conductance images.
FWHM (%) for aluminium FWHM (%) for perspex cylinder
Freq
(Hz)
50 mM 100 mM 150 mM 50 mM 100 mM 150 mM
1.12 31 29 26 71 71 73
15.56 71 30 26 72 71 71
282.38 43 26 28 37 30 49
4551.5 49 32 33 41 46 39
77712 44 29 29 38 37 36
Table 6.2b. FWHM for aluminium (C 130 mm) and perspex cylinder (C 100
mm) calculated from the capacitance images.
FWHM (%) for aluminium FWHM (%) for perspex cylinder
Freq (Hz) 50 mM 100 mM 150 mM 50 mM 100 mM 150 mM
1.12 85 35 28 75 77 76
15.56 82 35 29 79 73 75
282.38 50 36 37 40 41 39
4551.5 62 39 31 84 79 35
77712 56 37 41 47 46 43
Table 6.3a. Percentage of size difference between the original and reconstructed
objects for aluminium (C 130 mm) and perspex cylinder (C 100 mm) calculated
from half maximum of the conductance images.
Difference (%) for aluminium Difference (%) for perspex cylinder
Freq (Hz) 50 mM 100 mM 150 mM 50 mM 100 mM 150 mM
1.12 105 67 60 212 214 218
15.56 164 69 60 216 213 214
282.38 98 60 64 111 89 148
4551.5 114 74 75 123 138 118
77712 101 68 66 113 112 108
178
Table 6.3b. Percentage of size difference between the original and reconstructed
objects for aluminium (C 130 mm) and perspex cylinder (C 100 mm) calculated
from half maximum of the capacitance images.
Difference (%) for aluminium Difference (%) for perspex cylinder Freq (Hz)
50 mM 100 mM 150 mM 50 mM 100 mM 150 mM
1.12 189 80 51 224 231 229
15.56 188 82 56 236 220 224
282.38 115 83 84 119 123 118
4551.5 144 89 58 251 238 106
77712 129 86 86 140 137 130
Tables 6.3a and 6.3b reveal that the resolution of the reconstruction
algorithm improves with increasing frequency when viewing a perspex
(insulating) object. The resolution when viewing the metal object also improves
with increasing frequency and concentration. It also is significantly higher over
the whole frequency range with capacitance images providing the best resolution
in this instance.
179
6.6 Discussions
The reconstructed images of a metallic object suspended in an electrolyte
(Figures 6.3 6.5) was seen as an insulator because the impedance of the double
layer formed on the metal-electrolyte interface is very high at low frequencies.
The images increasingly reflected the properties of a metal object at higher
concentrations correlating with a decrease in the difference between the
conductance of the metal and that of the surrounding electrolyte. This was because
the impedance of the double layer decreases with increasing concentration (as the
Debye length is inversely proportional to the square root of the concentration, e.g.
Equation 2.66). Nonetheless, the conductance images at the highest frequency
reflect the properties of the double layer rather than those of the metal with the
capacitance images providing the highest resolution of the size, shape and position
of the object.
In contrast, reconstruction images for an insulating object failed to produce
good images in terms of the size, shape, and position of the object, except at the
highest frequency F5 (= 77,712 Hz). However, the conductance images were
better resolved at F4 (= 4,551.5 Hz) than at F5 and the capacitance images appear
to resolve the object at F3 (= 282.38 Hz) and F5.
Comparing the concentration variations for both sets of images the
concentration of 150 mM yielded the best resolutions. A comparison of the best
reconstruction images for the metal and insulator, i.e. at the highest frequencies
and when the objects were immersed in 150 mM electrolyte, revealed little
180
difference. So, relying on images produced from data over only one frequency
regime, the regime used by present EIT researchers, did not provide sufficient
information for distinguishing between a conductor and an insulator. The
additional images derived from data over lower frequency regimes did yield
differences albeit that present reconstruction algorithms were proven inadequate
in producing an accurate electrical characterization such as presented in Chapter
5. This illustrates the benefits of extending the EIT frequency range to lower
frequencies and including capacitance measurements, which yielded the best
resolution for the metal object.
In addition, the algorithms always underestimate the values of the
reconstructed capacitance by ~1/10 of the measured values, for instance as
obtained in Chapter 5 (see Figures 5.18, 6.3, 6.4 and 6.5 for comparison). This
was also shown in the simulation of an insulating object earlier in Section 6.3.
6.7 Conclusions
From the results presented in this chapter it can concluded that:
1. The EIDORS reconstruction algorithms yielded images of a conducting and
insulating objects in frequency regimes (> 9.6 kHz) presently used by EIT
researchers [3-6] that could not characterise the electrical properties of the
objects [7-9]. As indicated in the earlier results in Chapter 5, the dispersions
of both conductance and capacitance take place at frequencies lower than 2
kHz.
181
2. Lower frequency regimes yielded differences in electrical properties of the
two objects, consistent with the observed conductance and capacitance
dispersions.
3. Capacitance-based images provided the best estimates of the size, shape and
location of the metal (conducting) object.
4. Low frequency impedance measurements of the magnitude and phase
provide additional and significant insight into the properties of objects being
viewed using EIT. This confirms that our BULFIS machine has potential for
EIT data acquisitions.
References
1. Ider, Y.Z., B.M. Eyuboglu, M. Kuzuoglu, K. Leblebicioglu, U. Baysal,
B.K. Caglar, and O. Birgu. (1995). A method for comparative evaluation
of EIT algorithms using a standard test. Physiological Measurement, 16
Suppl. 3A: 227-236.
2. Wexler, A., B. Fry, and M.R. Neuman. (1985). Impedance-computed
tomography algorithm and system. Applied Optics, 24: 3985-3992.
3. Casas, O., J. Rosell, R. Bragos, A. Lozano, and P.J. Riu. (1996). A parallel
broadband real-time system for electrical impedance tomography.
Physiological Measurement, 17 Suppl 4A: A1-6.
4. Chauveau, N., B. Ayeva, B. Rigaud, and J.P. Morucci. (1996). A
multifrequency serial EIT system. Physiological Measurement, 17 Suppl
4A: A7-13.
5. Li, J.H., C. Joppek, and U. Faust. (1996). Fast EIT data acquisition system
with active electrodes and its application to cardiac imaging.
Physiological Measurement, 17 Suppl 4A: A25-32.
182
6. Hampshire, A.R., R.H. Smallwood, B.H. Brown, and R.A. Primhak.
(1995). Multifrequency and parametric EIT images of neonatal lungs.
Physiological Measurement, 16(3 Suppl A): A175-189.
7. Chilcott, T.C. and H.G.L. Coster. (1999). Electrical bioimpedance
methods: Electrical impedance tomography study of biological processes
in a single cell. Annals of the New York Academy of Sciences, 873: 269-
286.
8. Chilcott, T.C., H.G.L. Coster, and E.P. George. (1995). A novel method for
the characterization of the double fixed charge (bipolar) membrane using
impedance spectroscopy. Journal of Membrane Science, 108(1-2): 185-97.
9. Coster, H.G.L. (2003). Planar lipid bilayers (BLMs) and their
applications: Dielectric and electrical properties of lipid bilayers in
relation to their structure. Membrane Science and Technology Series, 7:
75-108.
7
RECONSTRUCTIONS OF
COMPLEX OBJECT
CONFIGURATIONS
7.1 Introduction
In this chapter I present reconstructed images obtained from experimental
data for two types of objects in various configurations within the chamber: e.g.
off-centre single object, double identical objects and double non-identical objects.
The spatial resolution is again calculated using the Full-Width at Half-Maximum
(FWHM) method described in Chapter 3.4.
It was concluded in Chapter 6 that the electrolyte concentration of 150
mM gives the best outcomes, therefore, this concentration was used for the
experiments described throughout this chapter. This concentration is also in the
physiological range of biological fluids.
184
7.2 Materials and Methods
The experimental data were collected from sets of experiments for
different object configurations. The first object was a stainless steel tubing ( 76
mm) placed off the centre of the chamber near electrode #1 (Figure 7.1). The
second object was a glass beaker (capacity 250 ml and 70 mm) placed close to
electrode #9. Later, two identical objects (in one instance metal and the other
glass) were positioned just off-centre, with one being close to electrode #1 and the
other one close to electrode #9. Then the same configuration was used but with a
glass object and a metal object. Lastly the reconstructions of a polyurethane
coated and uncoated stainless steel tubing were compared. The data were
collected using the same chamber and adjacent method to that described in
Chapter 6.
The solution used was sodium chloride (NaCl) with a concentration of 150
mM. To ensure that no solution penetrated beneath the objects a generous amount
of silicon grease was applied to the base of the object. Five signal frequencies
were used, i.e. 1.12 Hz (F1), 15.56 Hz (F2), 282.38 Hz (F3), 4551.5 Hz (F4) and
77712 Hz (F5). The impedance and its phase angle data were then used to
reconstruct the conductance and capacitance distributions using the program
EIDORS.
185
7.3 Reconstructed Images of a Single Off-Centre Object
In this section the images of a single object located just off the centre of
the chamber were reconstructed. The first set of images was obtained from a
stainless tubing (C 76 mm) that was located off-centre close to electrode #1 as
illustrated in Figure 7.1. The hollow tubing was chosen since there was not much
difference in the imaging of a hollow and solid metal (data not shown) since metal
is a very good conductor and stainless steel tubing was less expensive and more
readily available than solid stainless steel circular bars of the dimensions required.
Figure 7.1. The position of the object (stainless steel tubing) in the chamber
located off the centre near electrode #1 (the left side of the tubing was 20
mm from the centre of the chamber). The system used 16 electrodes evenly
spaced in the periphery of the chamber.
The recovered images are presented in Figure 7.2 showing that
conductance and capacitance distributions at low frequencies (e.g. F1, F2 and F3)
do not clearly reveal the metallic object at the actual location inside the system.
However, there are features in the capacitance distributions at F1 and F2 that
suggest its presence but its shape, size and location is best revealed at frequency
Electrode #1 Electrode #9
Electrode #5
Electrode #13
186
F4. The distributions at F5 tend to underestimate the degree to which the object is
off-centre. Note that conductance distributions show the traces of the measuring
electrodes, especially at frequencies F1, F2 and F4.
Figure 7.2. A stainless steel tubing (C 76mm) located off-centre near
electrode #1. The top row is conductance distributions and bottom row
capacitance distributions.
The conductance distributions look similar to those obtained previously
for a solid metal object (see Chapter 6.4) where the conductance images do not
clearly reveal the presence of the object at low frequencies (F1, F2 and F3). Thus,
these distributions seem to predict the conductance of the metal object is of the
order of that of the surrounding electrolyte although metal is the more highly
conducting material. This discrepancy can be explained by the presence of the
double layer enveloping the metal, whose impedance is very high at low
F1 F2 F3 F4 F5
10
-6
F/m
10
-3
S/m
187
frequencies. The object is revealed at F4 with a good estimate on its size (130% of
the original size) and shape, but not its location which is toward the centre of the
chamber, while the actual position is midway between the centre and the electrode
(see Figure 7.1 above). The frequency F5 does not resolve the object.
Similar trends are also shown by the capacitance images. Lower
frequencies F1 and F2 reveal a feature in the middle of the chamber that from a
prior knowledge of the system, i.e. the actual location of the object was in the
centre of the right hemisphere of the chamber, it can only be assumed to be an
artefact generated by the reconstruction algorithm. F3 reveals no object at all. The
object was observed at F4 with a good estimation on its size (110% of its original
size), shape and location. Although the reconstructed image is biased towards the
centre as per the conductance images. Again, F5 shows no object at all.
The cross sectional slice (taken from electrode #1 to electrode #9) of the
reconstructed images are given in Figure 7.3. The FWHMs were calculated only
for F4 and F5, which were 33% and 37%, respectively for the conductance
images, and 29% and 44%, for the capacitance images. All reconstructed objects
appeared bigger (by 1.2 to 1.7 times) than the original size of the object. The peak
values of the conductance and capacitance of the reconstructed image are
presented in Figure 7.4. It is clearly revealed that F4 gave the best contrast for the
conductance and capacitance images. This is consistent with the characteristics
obtained previously in Chapter 6.4 (see Figures 6.9(a) and (c), although here the
capacitance image gave a high peak at F5.
188
Figure 7.3. Cross-sectional curves (cross-section 1: along the centre from
electrode #1 on the right to electrode #9 on the left) of the off-centre metal
object near electrode #1 in a 150 mM NaCl solution. The vertical dashed
lines show the actual position and size of the metal object.
(a) (b)
Figure 7.4. Variation of peak values of (a) conductance and (b) capacitance
of the metal object taken from Figure 7.3.
-15 -10 -5 0 5 10 15
-20
-15
-10
-5
0
5
10
c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
2
)
diameter (cm)
1.1176 Hz
15.5645 Hz
282.384 Hz
4,551.5 Hz
77,712 Hz
-15 -10 -5 0 5 10 15
-10
0
10
20
30
40
50
c
a
p
a
c
i
t
a
n
c
e

(
u
F
/
m
2
)
diameter (cm)
1.1176 Hz
15.5645 Hz
282.384 Hz
4,551.5 Hz
77,712 Hz
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-20
-15
-10
-5
0
5
p
e
a
k

c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
2
)
frequency (Hz)
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
0
5
10
15
20
25
30
35
40
45
50
p
e
a
k

c
a
p
a
c
i
t
a
n
c
e

(
u
F
/
m
2
)
frequency (Hz)
c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
)

c
a
p
a
c
i
t
a
n
c
e

(

F
/
m
)

p
e
a
k

c
a
p
a
c
i
t
a
n
c
e

(

F
/
m
)
p
e
a
k

c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
)

189
Another configuration is shown in Figure 7.5 where an empty glass beaker
of 250 ml capacity (C 70 mm) was placed off centre near electrode #9, the left
side of the glass was 22 mm from the electrode.
Figure 7.5. The position of the object (an empty beaker glass) in the
chamber located off the centre very close to electrode #9.
Figure 7.6. The reconstructed images of an empty glass beaker (C 70mm)
located off-centre near electrode #9. The top row is conductance
distributions and bottom row capacitance distributions.
Electrode #1 Electrode #9
Electrode #5
Electrode #13
10
-3
S/m
10
-6
F/m
F1 F2 F3 F4 F5
190
Figure 7.6 depicts the reconstructed images of the conductance and
capacitance distributions that detect the presence of the beaker at frequencies F2,
F3 and F4 but neither its size nor its location. The capacitance image at F1 seems
to provide the best estimate of its location. However, there is limited contrast
between capacitances identifying the object and the electrolyte. The object was
washed away at F5, but not at F2. However, both types of images show a strong
artefact at the centre of the imaging area.
The cross sectional slice of the reconstructed images are given in Figure
7.7. For the conductance images the FWHM was calculated only for F4 that gives
the strongest visibility and gave 16% of the diameter of the chamber. From
capacitance images the FWHMs were determined for F1, F3 and F4, which give
25%, 15% and 15% of the diameter of the chamber, respectively. These were
taken from the top of the curves on the left side of the graphs (i.e. not the
shallower peaks on the conductance images or the peaks on the capacitance
images). The peak values of the conductance and capacitance images are depicted
in Figure 7.7 showing consistency with the previous results obtained presented in
Chapter 6.4 (see Figures 6.9(b) and (d)) where frequency F4 gave the best
contrast. However, the reconstructed object in Figure 7.6 appeared smaller
(between 57% and 90% of the actual size) than its original size for they were
plotted on the same scale.
From the above reconstructions the metallic object was visualised as
bigger (~120% to 170%) than its actual size, while in contrast, the glass object
was visualised as smaller (~57% to 90%) than its actual size.
191
Figure 7.7. Cross-sectional curves (cross-section 1: along the centre from
electrode #1 on the right to electrode #9 on the left) of the off-centre glass
object near electrode #9 in a 150 mM NaCl solution. The vertical dashed
lines show the actual position and size of the metal object.
(a) (b)
Figure 7.8. Variation of peak values of (a) conductance and (b) capacitance
of the metal object taken from Figure 7.7.
-15 -10 -5 0 5 10 15
-15
-10
-5
0
5
10
c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
2
)
diameter (cm)
1.1176 Hz
15.5645 Hz
282.384 Hz
4,551.5 Hz
77,712 Hz
-15 -10 -5 0 5 10 15
-5
0
5
10
c
a
p
a
c
i
t
a
n
c
e

(
u
F
/
m
2
)
diameter (cm)
1.1176 Hz
15.5645 Hz
282.384 Hz
4,551.5 Hz
77,712 Hz
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-1
0
1
2
3
4
5
p
e
a
k

c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
2
)
frequency (Hz)
10
-1
10
0
10
1
10
2
10
3
10
4
10
5
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
p
e
a
k

c
a
p
a
c
i
t
a
n
c
e

(
u
F
/
m
2
)
frequency (Hz)
c
a
p
a
c
i
t
a
n
c
e

(

F
/
m
)

c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
)

p
e
a
k

c
o
n
d
u
c
t
a
n
c
e

(
m
S
/
m
)

P
e
a
k

c
a
p
a
c
i
t
a
n
c
e

(

F
/
m
)
192
7.4 Reconstructed Images of Double Objects
7.4.1 Metal-Metal
Figure 7.9. Positions of two hollow stainless steel tubings in the chamber.
Identical materials were used with separation of 40 mm.
The configuration of two identical stainless tubing (C 76mm) as depicted
in Figure 7.9 was also investigated. The solution used was 150 mM sodium
chloride. The reconstructed images of conductance and capacitance distributions
are presented in Figure 7.10. and generally reveal the presence of the two
conducting objects. At frequency F2 they are recognised as two inhomogeneities
of low conductivity (compared to the surrounding solution) although not identical.
The low conductivity was attributed to the presence of low conducting double
layer at the interface between the electrolyte and metal (see Chapters 5.4 and 6.4)
The impedance of the double layer decreases with increasing frequency, which is
indicated by the change of the object colour from dark blue (low) at F2 to light
Electrode #1 Electrode #9
Electrode #5
Electrode #13
193
blue at F3 and then to red (high) at F4. The later (i.e. the red at F4) identifies the
objects as regions of high conductance. This transition is consistent with the
impedances of the double layers decreasing as the frequency rises. This trend does
not extend to higher frequencies where present EIT measurements are made,
which is illustrated in Figure 7.10 where the images of the objects disperse into
the yellow background. Therefore, the images of the objects characteristically
flipped from being apparent insulators to becoming conductors at a frequency
somewhere between F3 and F4, as was the case for a single metal object in the
results presented in Chapter 6.4. In contrast, the capacitance characteristics of the
images flipped from low to high between F2 and F3, again consistent with the
earlier characterisations of metal objects (see Chapters 5.4 and 6.4).
Figure 7.10. Images of two identical stainless steel tubing (C 76mm)
located off-centre. The top row is conductance distributions and bottom row
capacitance distributions.
10
-3
S/m
10
-6
F/m
F1 F2 F3 F4 F5
194
The capacitance distribution at higher frequency (i.e. at F4) does yield
distinct images of the objects. Both types of distributions reveal the size, shape
and location of the objects at F2, F3 and F4, even the capacitance images provide
a better location with distinct separation (except at F3) between the two objects.
7.4.2 Glass-Glass
In this experiment two identical empty glass beaker objects (C 70 mm)
replaced the stainless steel as illustrated in Figure 7.11. The reconstructed
distributions are shown in Figure 7.12. In these reconstructions the images of the
objects are evident at F1, F2 and F4 but not at F3 and F5. This characteristic was
expected since the objects were insulators and any change of the impedance of
double layers
1
at the surface of the objects would not affect the total impedance
(see Chapter 6.4). The sizes and shapes of the insulating objects were difficult to
determine from the images.
1
The double layers at the surface of an insulator such as glass are more diffuse (lower
capacitance) compared with those at the metal-electrolyte interface and in any case do not
affect the conduction into and through the object which is an insulator..
195
Figure 7.11. Positions of two beaker glasses (C 70 mm) in the chamber.
The objects are 40 mm symmetrically separated.
Figure 7.12. Images of two beaker glasses located off-centre. The top row is
the conductance distributions and bottom row capacitance distributions.
10
-3
S/m
10
-6
F/m
Electrode #1 Electrode #9
Electrode #5
Electrode #13
F1 F2 F3 F4 F5
196
7.4.3 Metal-Glass
The configuration of mixed objects was also investigated. A stainless steel
tubing (C 76mm) and an empty beaker glass (C 70mm) were placed in the
chamber as illustrated in Figure 7.13. The resulting reconstructed images are
presented in Figure 7.14.
Figure 7.13. Location the two different objects in the chamber. A beaker
glass was positioned on the left hand side and a stainless steel tubing on the
right hand side. The objects were separated by 40 mm from each other.
The images of the different types of objects look similar at low frequencies
(e.g. F1 and F2). This is because the impedance of the double layer at metal-
electrolyte interface is very high at these frequencies, yielding an image typical of
an insulator. The reconstructions clearly distinguish between the objects at the
higher frequency of F4 where the impedance of the double layer is orders of
magnitude smaller.
Electrode #1 Electrode #9
Electrode #5
Electrode #13
glass
metal
197
The capacitance images at all frequencies except F3 reveal the position of
the objects, whereas the conductance images are less definitive at lower
frequencies. The capacitance images at F1, F2, F4 and F5 further distinguish
between the conducting and insulating objects. The conducting images at F4 and
F5 reveal a high conductance region corresponding to the known position of the
conducting object but not a low conductance region for the insulator. This further
highlights the benefits of capacitance characterisations at frequencies below those
normally used in EIT.
Figure 7.14. Images of two objects: a stainless steel tubing (C 76mm)
located off-centre near electrode #1 and a beaker glass (C 70mm) located
off-centre near electrode #9. The top row is conductance distributions and
the bottom row capacitance distributions.
10
-3
S/m
10
-6
F/m
F1 F2 F3 F4 F5
198
7.4.4 Coated - Uncoated Metal
In order to investigate further the effects of double layer insulation at low
frequencies two experiments were carried out. The first experiment was conducted
on a stainless steel tube (C 76mm) similar to experiment described in section 7.3
(e.g. Figure 7.1). Thus, the object would be expected to be wrapped by the double
layer which would insulate the obnject in the low frequency regime. In the second
experiment the stainless tube was coated with a thin layer of polyurethane plastic
(sprayed painted with Wattyl Estapol Gloss paint) and placed on the other side of
the chamber (near electrode #9). The reconstructed images obtained from these
measurements are shown in Figure 7.16.
The conductance images do reveal low conductance regions in the
locations of the objects at F1, F2, F3 and F4. At F2 the coated and uncoated
objects look similar in size and shape, however at F3 and F4 the size of the
uncoated object was 103% recovered, in contrast to the coated object that was
only 60% and 50% recovered at F3 and F4, respectively. This suggests that the
impedance of the double layer on the uncoated metal object was orders of
magnitude higher than the impedance of the plastic coating. At F5 the object
disappeared from the imaging region indicating that the double layer and plastic
coating impedances are comparable. This transition shows that at frequencies
between F4 and F5 the effect of the impedances of the double layer and plastic
coating greatly decreased. At the higher frequencies above F5 it is expected that
the object would be identified by its highly conducting metallic properties.
The capacitance images of the uncoated object do not reveal any object at
low frequencies (i.e. F1, F2 and F3). The features shown in the middle of the
199
imaging area are the artefacts from the algorithm (see Section 7.3). They contrast
with the images of the coated object, where, although it is difficult to detect, the
presence of the object is identified by the dark blue colour and from prior
knowledge of the object location. At frequency F4, however, only the images of
the uncoated object revealed the object with its correct size, shape and location. At
F5 both objects could not be discerned.
Figure 7.15. Location the polyurethane coated and uncoated objects in the
chamber.
Electrode #1 Electrode #9
Electrode #5
Electrode #13
uncoated
coated
200
F1
F2
F3
F4
F5
Figure 7.16. Images a polyurethane-coated stainless steel tubing located off-
centre near electrode #9 (left column of each set) and an uncoated stainless
steel tubing located off-centre near electrode #1 (right column of each set).
The left hand side set is conductance distributions and the right hand side set
is capacitance distributions.
Conductance Capacitance
Uncoated Coated Uncoated Coated
10
-3
S/m 10
-6
F/m
201
7.5 Discussions
When the object was made of metal, it appeared as an insulator at low
frequencies and a conductor at high frequencies. This was because an electrical
double layer formed on the surface of the metal in contact with the electrolyte,
which has properties similar to that of a parallel plate capacitor. The impedance of
the layer is very high at low frequencies and low at high frequencies. Whilst for
the object made of insulating material, this phenomenon can be understood as the
object was not conducting any currents at all. The two different types of objects
(conducting and insulating) look similar and were identified as insulating objects
at low frequencies. At high frequencies the two objects appear as two different
objects (see Figures 7.2, 7.6, 7.10 and 7.12).
The dispersions of the complex electrical impedance of the system
containing metal object do occur at low frequencies. The conductance of the
metallic object appeared very low at low frequencies and increased with
increasing frequency. The frequencies between F3 (= 282.38 Hz) and F4 (=
4,551.5 Hz) were where most of the impedance transitions occurred from a high
value (corresponding to the impedance of the double layer) to a low value
(corresponding to the impedance of the surrounding electrolyte). This feature also
exists when the two dissimilar objects were present in the system (Figure 7.14).
The reconstructed images show that the algorithm approached the size of
the two dissimilar objects differently at high frequencies. The algorithm
overestimates the size of metallic object (see Figures 7.2 and 7.10 for metallic
objects only and Figure 7.14 when conducting and non-conducting objects were
present in the system).
202
The dominance of the double layer impedance was found stronger than
that of the polyurethane plastic coating (see Figure 7.16). The double layer
maintained its dominance up to the frequency F4, in contrast to the plastic coating
whose dominance decreased well below F4.. However, the results obtained with
the polymer coated metal object confirmed that the an insulating film (double
layer) surrounds the metallic object even when it is not coated with polymer.
7.6 Conclusions
From all the experiments that have been carried out it is clear that low
frequencies are, indeed, capable of providing more useful information on the
position, size and shape as well as the electrical behaviour of the object being
imaged than at high frequencies. This was especially so of images derived from
capacitance distributions.
Images at the frequency F4 (i.e. 4,551.5 Hz) provided little information
about the objects, whereas those at higher frequencies yielded the location and,
with images obtained below the frequency, provided a means of distinguishing
between the electrical properties of the objects.
8
GENERAL DISCUSSIONS,
CONCLUSIONS AND FUTURE
WORK
8.1 General Discussions
Medical imaging has been playing a significant role in clinical medicine.
Amongst other methods EIT drew great attention from a score of researchers in
the world. Despite its limitation in term of spatial resolution compared to other
techniques, EIT possesses a lot of advantages. EIT instrumentation and
technology is inexpensive, portable, non-invasive, non-ionising and fast.
Most available EIT instrumentations operate at frequencies above 9.6 kHz
and measure only the impedance magnitudes. Researchers in our laboratory [1-12]
have shown that biological tissues exhibit dispersions of their electrical
characteristics (i.e. conductance and capacitance) at very low frequencies. These
observations were obtained by measuring the magnitude of the impedance as well
as its phase angle using a four-terminal impedance measurement method.
This thesis aimed at explaining the usefulness of the low frequency range
(i.e. < 9.6 k Hz).
In Chapter 5 the characteristics of the chamber were presented. The results
revealed that the geometry and position of the electrodes in the absence of any
204
object affected the impedance measurements. Other aspects affecting the
impedance measurements were the electrode materials and the magnitude of the
injecting current. For example, the copper electrodes used for current electrodes
distorted the current in the low frequency range (typically below 15 Hz). This
phenomenon was also encountered by an earlier researcher in our laboratory [13].
The magnitude of the injecting current in our system was reduced at these
frequencies (see Chapter 4) to minimise the distortion which was attributable to
either water splitting (electrolysis) or precipitation of CuCl
2
. Dispersions of the
capacitance with frequency were also observed at very low frequencies due to
electrochemical events within the electrolyte as predicted by the Nernst-Planck-
Poisson theory (see Chapter 5.1) which includes the establishment of standing
electrochemical waves that result from blocking electrodes.
The placement of a conducting (metal) object into the chamber manifested
dispersions in the conductance and capacitance spectra (see Chapter 5). The
conductances were very low and lower than those of the electrolyte alone system
at low frequencies and dispersed at frequencies between 1 kHz to 10 kHz to
values that were higher than those of the electrolyte alone. On the other hand, the
capacitive dispersions occurred over the frequency range of 0.1 10 kHz. The
dispersions were consistent with the presence of the double layer at the metal-
electrolyte interface whose impedance is characterised by a very high value at low
frequencies that disperses to a low value at high frequencies and varies with the
concentration of the electrolyte. The conductance increases with the increasing
concentration. This is consistent with the theoretical prediction of the double layer
thickness varying with the concentration (see Chapter 2.6).
205
The reconstructed images of a single metallic object centrally located in
the experimental chamber presented in Chapter 6 reflected the electrical properties
of the double layer formed at the interface of metal and electrolyte, which
disperses with frequency in similar manner to that of a membrane [9, 10]. At
higher frequencies the capacitive properties of the double layer dominated and
leads to current flows through the metallic object. Such phenomena was
characterised using one set of four terminals (Chapter 5) and the phenomenon was
evident in images reconstructed from other characterisation using multiple sets of
electrodes distributed around the system (Chapters 6 and 7). However, images at
low frequencies were not completely consistent with the impedance measurements
because the pixels used in the images were not sufficiently small to model a
nanometer-thick double layer. Nonetheless, the images at low frequencies, and
especially those derived from capacitance distributions, provided best estimates of
the size, position and shape of the object as well as distinguishing between
conducting and insulating objects. This was especially illustrated when a metal
and a glass object were imaged simultaneously in the chamber utilising multiple
sets of electrodes (see Chapter 7). At low frequencies the two different objects
were both recognised as insulating bodies, whereas at frequencies greater than 4.5
kHz the two objects were distinguishable as two different entities with different
conductance and capacitance properties.
The quality of the images and the ability to discriminate between objects
of differing conductance properties is generally improved by the use of low
frequencies. More importantly, the use of EIT data acquisition over multiple
frequencies covering the low frequency range up to 10 kHz might provide a
206
means of discriminating between objects with different dielectric/conductivity
properties provides some advantages in EIT.
8.2 General Conclusions
In conclusion, it can be said that the use of multi frequency
instrumentation in EIT covering the very low frequency range provides structural,
positional, conduction and dielectric information of objects that instrumentation
restricted to frequencies > 10 kHz does not provide.
The results presented in this thesis show that objects of different electrical
properties (in this case, metallic and insulating objects) appeared differently in the
frequency range used (1.12 Hz up to 77,712 Hz). It is, therefore, promising to use
the EIT on human body, the system contains poorly conducting objects such as
bone, highly conducting interstitial fluids, blood and organs contained within
sacks or envelopes of relatively poorly conducting material, etc. The exploration
of such body systems was outside the scope of this thesis but the work described
herein has revealed the potential to gain additional information by multiple
frequencies EIT at frequencies below that commonly used in EIT.
8.3 Future Work
This thesis showed that impedance measurements of magnitude and phase
at very low frequencies provided 2D images of conductance and capacitive
properties of objects utilising an iterative EIT image reconstruction algorithm,
207
namely, the EIDORS algorithm. The metal object used in this study was chosen
because the double layer that forms at the interface between a metal and an
electrolyte has electrical properties that have a frequency dependence not unlike
that of a cell membrane or membranes that surround organs. This study
demonstrated the potential of such low frequency impedance measurements in
possible body imaging. However, the study also revealed practical limitations
such as the inability of the algorithm to model object as small as a double layer or
a membrane. Future work would need to develop better algorithms/models to
overcome this shortfall. Further methods need to be developed to shorten the data
acquisition procedures which with the present impedance spectrometer were time
consuming. The impedance spectrometer used in this study was not specifically
designed for EIT but was rather built for exacting impedance spectroscopy studies
of 1D molecular films and membranes. It would be in principle possible to make
measurements at several frequencies simultaneously by synthesizing a complex
waveform containing the frequency components of interest as the current source.
The response data could then be decomposed into the responses at each of the
known frequencies present to yield the impedance magnitude and phase at each of
these frequencies. Such a system would need to use an optimized mix of
sinusoidal signals as the impedance (and hence response) at the different
frequencies varies over a wide range. The development of such a multi-frequency
EIT instrument would be a major undertaking but the results obtained in the
present investigation would suggest this would yield worthwhile outcomes. The
optimum choice of the electrodes material also needs to be investigated.
208
Other investigations would include the use of multiple current sources and
extending measurements and analyses to low frequency 3D EIT. The work carried
out for this thesis justifies the feasibility of undertaking this future work. The
image reconstruction computations could also be enhanced by the use of
programmable logic arrays which would speed up the computations.
References
1. Ashcroft, R.G., H.G.L. Coster, and J.R. Smith, Location and effect of
adsorbed chemicals on the dielectric substructure of membranes by ultra
low frequency spectrometry, in U.S. 1980, (Unisearch Ltd., Australia). Us.
p. 7 pp.
2. Ashcroft, R.G., H.G.L. Coster, and J.R. Smith, The molecular
organization of bimolecular lipid membranes. The dielectric structure of
the hydrophilic/hydrophobic interface. Biochimica et Biophysica Acta,
1981. 643: p. 191-204.
3. Ashcroft, R.G., et al., Perturbations to lipid bilayers by spectroscopic
probes as determined by dielectric measurements. Biochimica et
Biophysica Acta, 1980. 602(2): p. 299-308.
4. Chilcott, T.C., Admittance tomography of cells of Chara corallina: a study
of the electrical spatial structures associated with photosynthesis, PhD
Thesis, in Faculty of Science. 1988, University of New South Wales:
Sydney, Australia.
5. Chilcott, T.C. and H.G.L. Coster, Electrical impedance tomography study
of biological processes in a single cell. Annals of the New York Academy
of Sciences, 1999. 873(Electrical Bioimpedance Methods): p. 269-286.
6. Chilcott, T.C., et al., Novel method for in vivo studies of plasmodesmata in
Chara corallina using impedance spectroscopy. Comptes Rendus de
l'Academie des Sciences, Serie III: Sciences de la Vie, 1996. 319(1): p. 17-
27.
209
7. Coster, H.G.L., The physics of cell membranes. Journal of Biological
Physics, 2003. 29(4): p. 363-399.
8. Coster, H.G.L., Dielectric and electrical properties of lipid bilayers in
relation to their structure. Membrane Science and Technology Series,
2003. 7(Planar Lipid BiLayers (BLMs) and Their Applications): p. 75-
108.
9. Coster, H.G.L. and T.C. Chilcott, The characterization of membranes and
membrane surfaces using impedance spectroscopy, in Surface chemistry
and electrochemistry of membranes, T.S. Sorensen, Editor. 1999, Marcel
Dekker, Inc: New York. p. 749-792.
10. Coster, H.G.L., T.C. Chilcott, and A.C.F. Coster, Impedance spectroscopy
of interfaces, membranes and ultrastructures. Bioelectrochemistry and
Bioenergetics, 1996. 40: p. 79-98.
11. Coster, H.G.L. and J.R. Smith, Low frequency impedance of Chara
corallina: simultaneous measurements of the separate plasmalemma and
tonoplast capacitance and conductance. Australian Journal of Plant
Physiology, 1977. 4: p. 667-764.
12. Smith, J.R., Electrical characteristics of biological membranes in different
environments, PhD Thesis, in Faculty of Science. 1977, University of New
South Wales: Sydney.
13. Gaedt, L.P., Impedance spectroscopy characterisation of synthetic
membranes and processes, PhD Thesis, in Faculty of Science. 1999,
Univesity of New South Wales: Sydney, Australia.
14. Ider, Y.Z., et al., A method for comparative evaluation of EIT algorithms
using a standard test. Physiological Measurement, 1995. 16 Suppl. 3A: p.
227-236.
15. Wexler, A., B. Fry, and M.R. Neuman, Impedance-computed tomography
algorithm and system. Applied Optics, 1985. 24: p. 3985-3992.
APPENDIX A
RECONSTRUCTIONS SOURCE CODES
A.1 The Main Routine
%EidorsCompl.m is to reconstruct complex images from real measurements
% M. Vauhkonen 20.5.2000,
% University of Kuopio, Department of Applied Physics, PO Box 1627,
% FIN-70211 Kuopio, Finland, email: Marko.Vauhkonen@uku.fi
%
% Modified by Johan Noor, 28.04.2004
% Dept. of Biophysics, University of New South Wales
% Sydney NSW 2052, Australia
%
% This routine uses .lis and .ei3 data files.
%
clear all
tic
load meshdata7 % Load the meshdata for two different meshes.
NNode1=max(size(Node1)); %The number of nodes
NElement1=max(size(Element1)); %The number of element
NNode2=max(size(Node2)); %The number of nodes
NElement2=max(size(Element2)); %The number of elements
g1=reshape([Node1.Coordinate],2,NNode1)';
H1=reshape([Element1.Topology],3,NElement1)';
g2=reshape([Node2.Coordinate],2,NNode2)';
H2=reshape([Element2.Topology],3,NElement2)';
% Loading .LIS and .EI3 data files.
load d0a1050a.LIS % contents: hours freq SNRv DCv G C SNRc DCc
load d0a1050b.LIS
dlist=[d0a1050a;d0a1050b];
load d0a1050a.EI3 % contents: freq Z Zphase ACv ACc
load d0a1050b.EI3
dei3=[d0a1050a;d0a1050b]; % append file b to file a.
frek=dei3(1:5,1);
A-2
Z=dei3(:,2);
Z=reshape(Z,5,max(size(dei3))/5);
Z1=Z(1,:)'; Z2=Z(2,:)'; Z3=Z(3,:)'; Z4=Z(4,:)'; Z5=Z(5,:)';
Zph=dei3(:,3);
Zph=reshape(Zph,5,max(size(dei3))/5);
Zph1=Zph(1,:)'; Zph2=Zph(2,:)'; Zph3=Zph(3,:)'; Zph4=Zph(4,:)';
Zph5=Zph(5,:)';
ACv=dei3(:,4);
ACv=reshape(ACv,5,max(size(dei3))/5);
ACv1=ACv(1,:)'; ACv2=ACv(2,:)'; ACv3=ACv(3,:)'; ACv4=ACv(4,:)';
ACv5=ACv(5,:)';
L=16; % The number of electrodes.
z=0.005*ones(L,1); % Contact impedances.
[II1,T]=Current(L,NNode2,'adj'); % Adjacent current pattern.
MeasPatt=T;
on=ones(NElement1,1);
[Agrad,Kb,M,S,C]=FemMatrix(Node2,Element2,z);
Agrad1=Agrad*Ind2; % Group some of the elements for the inverse
% computations
%Regularisation parameter and matrix
alpha=0.000005;
R=MakeRegmatrix(Element1);
R=[R,zeros(size(R));zeros(size(R)),R];
%% PROCEDURE TO SOLVE THE INVERSE PROBLEM %%
% Approximate the best homogenous 1/adm.
A=UpdateFemMatrix(Agrad,Kb,M,S,.5*(1+i)*ones(NElement2,1));
% The system matrix.
Urefh=ForwardSolution1(NNode2,NElement2,A,C,T,MeasPatt,'comp');
Urefhel=Urefh.Electrode;
Urefhela=Urefhel(1:L,:);
Urefhelb=Urefhel(L+1:end,:);
iter=5; % set the number of iterations
% =================================================
%% DATA #1 ==> freq = 1.11757 Hz
freq=frek(1);
meas=ACv1;
phase=Zph1;
disp(['Now doing frequency ',num2str(frek(1)),' Hz .... '])
for ii=1:max(size(phase))
if phase(ii) < -45
phase(ii)=-phase(ii);
end
end
A-3
Uela=[];Uelb=[];
for ii=1:max(size(meas))
Uela=[Uela;meas(ii)*cos(phase(ii)*pi/180)]; % real part
Uelb=[Uelb;meas(ii)*sin(phase(ii)*pi/180)]; % imaginary part
end
Uela=reshape(Uela,L,L);
Uelb=reshape(Uelb,L,L);
Uel=[Uela;Uelb];
% First estimation
J=Jacobian(Node2,Element2,sum(Agrad1')',Urefh.Current,Urefh.MeasField,[]
,'comp');
adm0=1/600*ones(2,1)+(J'*J)\(J'*(Uel(:)-Urefh.Electrode(:)));
adm=adm0(1)*ones(NElement1,1)+i*adm0(2)*ones(NElement1,1);
admbig=Ind2*adm;
A=UpdateFemMatrix(Agrad,Kb,M,S,admbig); % The system matrix.
Uref=ForwardSolution1(NNode2,NElement2,A,C,T,MeasPatt,'comp');
Urefel=Uref.Electrode(:);
J=Jacobian(Node2,Element2,Agrad1,Uref.Current,Uref.MeasField,[],'comp');
adm=[real(adm);imag(adm)];
adm15=[];
remm15=[]; %real max min
immm15=[]; %imag max min
for ii=1:iter
adm=adm+(J'*J+alpha*R'*R)\(J'*(Uel(:)-Urefel)-alpha*R'*R*adm);
adm=reshape(adm,max(size(Element1)),2);
adm=adm(:,1)+i*adm(:,2);
adm15=[adm15,adm];
readm=real(adm);imadm=imag(adm);
Minimag15=min(imadm(:));Maximag15=max(imadm(:));
Minreal15=min(readm(:));Maxreal15=max(readm(:));
remm15=[remm15;max(readm(:));min(readm(:))];
immm15=[immm15;max(imadm(:));min(imadm(:))];
admbig=Ind2*adm;
A=UpdateFemMatrix(Agrad,Kb,M,S,admbig); % The system matrix.
Uref=ForwardSolution1(NNode2,NElement2,A,C,T,MeasPatt,'comp');
Urefel=Uref.Electrode(:);
J=Jacobian(Node2,Element2,Agrad1,Uref.Current,Uref.MeasField,...
[],'comp');
figure,Plotinvsol(real(adm),g1,H1),
title(['Conductance (mS) for freq ',num2str(freq),...
' Hz after ',num2str(ii),' iterations']);colorbar,drawnow
figure,Plotinvsol(imag(adm),g1,H1),
title(['Capacitance (mF) for freq ',num2str(freq),...
' Hz after ',num2str(ii),' iterations']);colorbar,drawnow
adm=[real(adm);imag(adm)];
end
remm15=reshape(remm15,2,ii);
immm15=reshape(immm15,2,ii);
save d0a105015c7.mat adm15 ii freq remm15 immm15
A-4
A.2 The Forward Solution Function
function [U,p,r]=ForwardSolution1(NNode,NElement,A,C,T,MeasPattern,...
style,p,r);
%ForwardSolution1 Solves the potential distribution in 2D EIT
% Function [U,p,r]=ForwardSolution1(MeshData,rho);
% computes the internal voltages for the injected currents (U.Current)
% and for the "measurement field" (U.MeasField) and the voltages
% on the electrodes (U.Electrode).
%
% INPUT
%
% NNode = the number of nodes
% NElement = the number of elements
% A = the system matrix
% C = the reference matrix
% T = current patterns
% MeasPattern = measurement pattern, default \sum_{l=1}^L=0.
% style = 'comp' for complex reconstruction and 'real' for real
% p = the permutation vector (optional)
% r = reversed permutation (optional)
%
% OUTPUT
%
% U = potential data structure, includes potetial corresponding to
% the actual current injections (U.Current), for the measurement
pattern (U.MeasField)and voltages on the electrodes (U.Electrode)
% p = the permutation vector (optional)
% r = reversed permutation (optional)
% M. Vauhkonen 13.8.1999
% University of Kuopio, Department of Applied Physics,
% PO Box 1627, FIN-70211, Kuopio, Finland, email:Marko.Vauhkonen@uku.fi
% Ordering added 15.11.1999, modified 28.3.2000 by M. Vauhkonen
%
% Modified by Johan Noor, 28.04.2004
% Dept. of Biophysics, University of New South Wales
% Sydney NSW 2052, Australia
L=max(size(A))-NNode+1; % The number of electrodes
if style=='comp'
if ~isempty(MeasPattern)
II1=sparse([zeros(L,NNode),C]);
II1=MeasPattern'*II1;
II1=sparse([II1,zeros(L,NNode+L-1);zeros(L,NNode+L-1),II1]);
II1=II1';
else
II1=sparse([zeros(L,NNode),C,zeros(L,NNode+L-1);...
zeros(L,2*NNode+L-1),C]);
MeasPattern=eye(max(size(C)));
II1=II1';
A-5
end
II=sparse([[zeros(NNode,size(T,2));C'*T];...
zeros(NNode+L-1,size(T,2))]);
A=[real(A),-imag(A);imag(A),real(A)];
UU=A'\II1; % Voltages for the "measurement field" and for the
% current patterns.
UU=full([UU,A\II]);
%The "measurement field" data
U.MeasField=[UU(1:NNode,1:size(II1,2));...
UU(NNode+L:2*NNode+L-1,1:size(II1,2))];
%Voltages on the electrodes
if ~isempty(MeasPattern)
U.Electrode=[MeasPattern'*C,zeros(size(C));zeros(size(C)),...
MeasPattern'*C]*[UU(NNode+1:NNode+L-1,...
size(II1,2)+1:size(UU,2)); UU(2*NNode+L:size(UU,1),...
size(II1,2)+1:size(UU,2))];
else
U.Electrode=[C,zeros(size(C));zeros(size(C)),C]*...
[UU(NNode+1:NNode+L-1,size(II1,2)+1:size(UU,2));...
UU(2*NNode+L:size(UU,1),size(II1,2)+1:size(UU,2))];
end
U.Current=[UU(1:NNode,size(II1,2)+1:size(UU,2));...
UU(NNode+L:2*NNode+L-1,size(II1,2)+1:size(UU,2))];
elseif style=='real'
II1=sparse([zeros(L,NNode),C]);
if ~isempty(MeasPattern)
II1=MeasPattern'*II1;
II1=II1';
else
MeasPattern=eye(max(size(C)));
II1=II1';
end
II=[zeros(NNode,size(T,2));C'*T];
if nargin<8
p=symmmd(A);
r(p)=1:max(size(p));
end
R=chol(A(p,p));
UU=R\(R'\[II1(p,:),II(p,:)]);
UU=full(UU(r,:));
%The "measurement field" data
U.MeasField=UU(1:NNode,1:size(II1,2));
%Voltages on the electrodes
U.Electrode=MeasPattern'*C*UU(NNode+1:size(A,1),...
size(II1,2)+1:size(UU,2));
U.Current=UU(1:NNode,size(II1,2)+1:size(UU,2));
end
A-6
A.3 The FEM Matrix Construction Function
function [Agrad,Kb,M,S,C]=FemMatrix(Node,Element,z);
%FemMatrix Computes the blocks of the system matrix for 2D EIT with
% linear and quadratic basis
% Function [Agrad,Kb,M,S,C]=FemMatrix(Node,Element,z);
% computes the matrices needed in the finite element
% approximation of the 2D EIT forward problem.
%
% INPUT
%
% Node = nodal data structure
% Element = element data structure
% z = a vector of (complex) contact impedances
%
% OUTPUT
%
% Agrad = the gradient part of the system matrix
% Kb,M and S = other blocks of the system matrix
% C = voltage reference matrix
% M. Vauhkonen 11.5.1994, modified from the version of J.P. Kaipio
% 25.4.1994. Modified 5.9.1994 by M. Vauhkonen for EIT.
% Modified 13.8.1999 and 23.3.2000 for the EIDORS by M. Vauhkonen,
% University of Kuopio, Department of Applied Physics, PO Box 1627,
% FIN-70211 Kuopio, Finland, email: Marko.Vauhkonen@uku.fi
Nel=max(size(z)); %The number of electrodes.
NNode=max(size(Node)); %The number of nodes
NElement=max(size(Element)); %The number of elements
M=sparse(NNode,Nel);
Kb=sparse(NNode,NNode);
Agrad=sparse(NNode^2,NElement);
s=zeros(Nel,1);
g=reshape([Node.Coordinate],2,NNode)'; %Nodes coordinate
for ii=1:NElement
A=sparse(NNode,NNode);
ind=(Element(ii).Topology); % Indices of the 3 nodes/vertices of
% the ii'th element
gg=g(ind,:); % A 3x2 or 6x2 matrix of triangle
% vertices in (x,y) coord.
if max(size(gg))==3 % First order basis
grint=grinprodgaus(gg,1);
else
grint=grinprodgausquad(gg,1); % Second order basis
end
if any([Element(ii).Face{:,3}]), % Checks if the triangle ii is
% a triangle that is under
% the electrode.
[In,Jn,InE]=find([Element(ii).Face{:,3}]);
bind=Element(ii).Face{Jn,1}; % Nodes on the boundary
ab=g(bind(:),:);
A-7
if max(size(bind))==2 % First order basis
bb1=bound1([ab]);Bb1=zeros(max(size(ind)),1);
bb2=bound2([ab]);Bb2=zeros(max(size(ind)));
s(InE)=s(InE)+1/z(InE)*2*bb1; % 2*bb1 = length of the
% electrode.
eind=[find(bind(1)==ind),find(bind(2)==ind)];
else % Second order basis
bb1=boundquad1([ab]);Bb1=zeros(max(size(ind)),1);
bb2=boundquad2([ab]);Bb2=zeros(max(size(ind)));
s(InE)=s(InE)+1/z(InE)*electrlen([ab]);
eind=[find(bind(1)==ind),find(bind(2)==ind),find(bind(3)==ind)];
end
Bb1(eind)=bb1;
M(ind,InE)=M(ind,InE)-1/z(InE)*Bb1;
Bb2(eind,eind)=bb2;
A(ind,ind)=grint;
Agrad(:,ii)=A(:);
Kb(ind,ind)=Kb(ind,ind)+1/z(InE)*Bb2;
else % The triangle isn't under the electrode.
A(ind,ind) = grint;
Agrad(:,ii)=A(:);
end
end
S=sparse(diag(s));
[II1,C]=Current(Nel,NNode,'adj');
C=C(:,1:Nel-1); % For the voltage reference
C=sparse(C(:,1:Nel-1));
S=C'*S*C;
M=M*C;
A-8
A.4 The Jacobian Function
function [J]=Jacobian(Node,Element,Agrad,U,U0,rho,style);
%Jacobian Computes the Jacobian for 2D EIT when elementwise basis
% is used
% Function [J]=Jacobian(g,H,Agrad,U,U0,rho,style);
% computes the Jacobian for 2D EIT when elementwise basis is used.
%
% INPUT
%
% Node = nodal data structure
% Element = element data structure
% Agrad = \int_{Element(ii) \nabla\phi_i\cdot\nabla\phi_j
% U = voltages corresponding to the injected currents
% U0 = voltages corresponding to the measurement field
% rho = resistivity or admittivity vector
% style = either 'real' for reconstructing resistivity or 'comp'
% for reconstructin admittivity
%
% OUTPUT
%
% J = Jacobian
% M. Vauhkonen 13.8.1999,
% University of Kuopio, Department of Applied Physics, PO Box 1627,
% FIN-70211 Kuopio, Finland, email: Marko.Vauhkonen@uku.fi
NNode=max(size(Node));
NElement=max(size(Element));
if nargin<3
Agrad=sparse(NNode^2,NElement); % Gradients of the basis functions
% integrated over each element.
for ii=1:NElement
ind=Element(ii).Topology;
gg=reshape([Node(ind).Coordinate],2,max(size(ind)))';
if max(size(ind))==3
anis=grinprodgaus(gg,1);
else
anis=grinprodgausquad(gg,1);
end
Aa=sparse(NNode,NNode);
Aa(ind,ind)=anis;
Agrad(:,ii)=Aa(:);
end
J=Agrad;
elseif nargin<6
J=zeros(size(U,2)*size(U0,2),size(Agrad,2));
for ii=1:size(Agrad,2);
A-9
JJ=U0.'*reshape(Agrad(:,ii),NNode,NNode)*U;
J(:,ii)=JJ(:);
end
else
if style=='comp'
J=zeros(size(U,2)*size(U0,2),2*size(Agrad,2));
for ii=1:size(Agrad,2);
Agrad1=reshape(Agrad(:,ii),NNode,NNode);
AgU1=Agrad1*U(1:NNode,:);
AgU2=Agrad1*U(NNode+1:2*NNode,:);
JJ=-U0.'*[AgU1;AgU2];
J(:,ii)=JJ(:);
JJ=-U0.'*[-AgU2;AgU1];
J(:,ii+size(Agrad,2))=JJ(:);
end
elseif style=='real'
J=zeros(size(U,2)*size(U0,2),size(Agrad,2));
for ii=1:size(Agrad,2);
JJ=U0.'*1/rho(ii)^2*reshape(Agrad(:,ii),NNode,NNode)*U;
J(:,ii)=JJ(:);
end
end
end

Anda mungkin juga menyukai