Sei sulla pagina 1di 221

A

HIGH-SENSITIVITY

FLEXIBLE-EXCITATION

CAPACITANCE

TOMOGRAPHY

ELECTRICAL

SYSTEM

A thesis submitted to the University of Manchester


Institute of Science and Technology for the degree of
Doctor of Philosophy

1997

JOSE CARLOS GAMIO ROFFE


Department of Electrical Engineering and Electronics

LIST OF CONTENTS

ABSTRACT .............................................. vii


DECLARATION ............................................ ix
ACKNOWLEDGEMENTS ........................................ x

CHAPTER 1: Introduction ................................. 1


1.1

OVERVIEW OF ELECTRICAL CAPACITANCE TOMOGRAPHY ...... 1

1.2

MAIN AREAS OF IMPROVEMENT .......................... 3

1.3

AIMS AND OBJECTIVES ................................ 6

1.4

ORGANISATION OF THIS THESIS ........................ 6

CHAPTER 2: Theory of electrical capacitance


Tomography (ECT) ............................. 8
2.1

THE ECT SENSOR AS A SYSTEM OF CHARGED


CONDUCTORS ......................................... 8

2.2

ECT MEASUREMENT STRATEGIES ........................ 12

2.2.1 THE SINGLE-ELECTRODE EXCITATION METHOD ......... 12


2.2.2 QUALITATIVE IMAGE RECONSTRUCTION FOR
SINGLE-ELECTRODE EXCITATION: THE LINEAR
BACK-PROJECTION (LBP) ALGORITHM ................ 14
2.2.2.1 THE SENSITIVITY MAPS ....................... 14
2.2.2.2 THE NORMALISED MEASUREMENTS ................ 16
2.2.2.3 THE WEIGHTED BACK-PROJECTION OPERATION ..... 17
2.2.3 MULTIPLE-ELECTRODE EXCITATION METHODS .......... 17
2.2.3.1 RECONSTRUCTION WITH MULTIPLE-ELECTRODE
EXCITATION ................................. 22

2.2.3.2 OPTIMAL MULTIPLE-ELECTRODE EXCITATION


PATTERNS ................................... 25
2.3

ECT TRANSDUCERS ................................... 28

2.3.1 THE CHARGE-DISCHARGE CAPACITANCE TRANSDUCER


CIRCUIT ........................................ 29
2.3.2 AC-BASED ECT TRANSDUCERS ....................... 32
2.4

DISCUSSION AND CONCLUSIONS ........................ 35

2.4.1 SINGLE- VS. MULTIPLE-ELECTRODE EXCITATION ...... 35


2.4.2 CHARGE-DISCHARGE VS. AC-BASED TRANSDUCERS ...... 36

CHAPTER 3: Finite-element simulation of singleelectrode and parallel-field excitation ..... 38


3.1

INTRODUCTION ...................................... 39

3.2

FINITE-ELEMENT MODELLING OF ECT SENSORS............ 39

3.3

SINGLE-ELECTRODE EXCITATION ....................... 44

3.4

IDEAL SENSITIVITY MAPS .......................... 45

3.5

PARALLEL-FIELD EXCITATION ......................... 47

3.5.1 PARALLEL-FIELD GENERATION ...................... 47


3.5.2 SENSITIVITY MAPS FOR PARALLEL-FIELD
EXCITATION ..................................... 50
3.6

COMPARISON OF IMAGES OBTAINED USING SINGLEELECTRODE AND PARALLEL-FIELD EXCITATION ........... 52

3.7

DETERMINATION OF MUTUAL CAPACITANCES FROM


PARALLEL-FIELD CHARGE MEASUREMENTS ................ 56

3.8

CONCLUSIONS ....................................... 59

CHAPTER 4: Design of an AC-based multipleexcitation ECT system ....................... 62


4.1

INTRODUCTION ...................................... 62
ii

4.2

DESIRED SYSTEM CHARACTERISTICS .................... 64

4.3

GENERAL DESIGN STRATEGY ........................... 65

4.4

THE BASIC DETECTOR CIRCUIT ........................ 67

4.4.1 TRANSFER FUNCTION WITH FREQUENCY-DEPENDENT


OP-AMP GAIN .................................... 68
4.4.2 COMMENTS ON NOISE PERFORMANCE .................. 73
4.4.3 DETAILED CIRCUIT DESIGN ........................ 74
4.4.4 FINAL FREQUENCY RESPONSE ....................... 76
4.4.5 EFFECT OF CHANGES IN STRAY CAPACITANCE ......... 77
4.5

DESIGN OF THE MULTI-EXCITATION ECT TRANSDUCER


CHANNEL ........................................... 80

4.5.1 NOMINAL MEASUREMENT RANGES AND SENSITIVITIES ... 84


4.5.2 EFFECT OF CMOS SWITCH

ON RESISTANCE .......... 85

4.6

REFERENCE SINE-WAVE GENERATOR ..................... 91

4.7

SIGNAL CONDITIONING AND DATA CONVERSION SECTION ... 93

4.8

SYSTEM INTEGRATION AND CONSTRUCTION ............... 95

4.9

CALIBRATION OF SYSTEM ELECTRONICS ................ 101

4.9.1 CALIBRATION OF THE EXCITATION VOLTAGE


SOURCES ....................................... 101
4.9.2 DETECTOR CALIBRATION .......................... 103
4.10

EXPERIMENTAL EVALUATION OF THE SYSTEM ........... 106

4.10.1 CAPACITANCE MEASUREMENT COMPARATIVE TEST ..... 107


4.10.2 LINEARITY EVALUATION ......................... 108
4.10.3 INTRINSIC NOISE LEVEL TEST ................... 110
4.10.4 DYNAMIC PERFORMANCE OF THE SYSTEM ............ 112
4.11

IMAGE SAMPLES ................................... 115

4.12

CONCLUSIONS ..................................... 118

iii

CHAPTER 5: Comparative experimental evaluation of


single- and multiple-electrode excitation
methods..................................... 120
5.1

INTRODUCTION ..................................... 120

5.2

EXPERIMENTS WITH OPTIMUM MULTI-ELECTRODE


EXCITATION ....................................... 121

5.2.1 DETERMINATION OF THE OPTIMUM EXCITATION


VECTORS ....................................... 121
5.2.2 APPLYING THE OPTIMUM EXCITATION VECTORS ....... 122
5.2.3 EXPERIMENTAL PROCEDURE ........................ 127
5.3

EXPERIMENT RESULTS AND DISCUSSION ................ 130

5.3.1 EXPERIMENT No. 1 .............................. 131


5.3.2 EXPERIMENT No. 2 .............................. 136
5.3.3 EXPERIMENT No. 3 .............................. 139
5.3.4 EXPERIMENT No. 4 .............................. 142
5.3.5 EXPERIMENT No. 5 .............................. 145
5.3.6 EXPERIMENT No. 6 .............................. 148
5.4

CONCLUSIONS ...................................... 151

CHAPTER 6: Iterative linear back-projection imagereconstruction techniques .................. 153


6.1

INTRODUCTION ..................................... 153

6.2

PREVIOUS WORK .................................... 153

6.3

A RECONSTRUCTION ALGORITHM INSPIRED ON FEEDBACK CONTROL ..................................... 156

6.3.1 FIRST APPROACH: ITERATIVE LBP RECONSTRUCTION .. 156


6.3.2 AN ALGORITHM BASED ON A CONTROL-SYSTEM
ANALOGY ....................................... 163

iv

6.4

EXPERIMENTAL EVALUATION OF THE FEED-BACK


ALGORITHM ........................................ 169

6.4.1 RESULTS OF IMAGE RECONSTRUCTION TESTS ......... 169


6.4.2 DISCUSSION .................................... 178
6.5

CONCLUSIONS ...................................... 179

CHAPTER 7: Conclusions and further work ............... 180


7.1

CONCLUSIONS ...................................... 180

7.2

RECOMMENDATIONS FOR FUTURE WORK .................. 182

REFERENCES ............................................ 184

APPENDIX A: System design details and circuit


diagrams .................................. 194

APPENDIX B: Computer programs .........................


B.0
General .........................................
B.1
ECT system monitoring and control programs ......
B.2
Programs used in the optimum excitation
experiments .....................................
B.3
Simulation programs .............................
B.3.1 Programs to calculate the sensitivity maps
for single-electrode excitation ...............
B.3.2 Programs to calculate the sensitivity maps
for parallel-field excitation .................
B.3.3 Programs to simulate single-electrodeexcitation measurements .......................
B.3.4 Programs to simulate parallel-fieldexcitation measurements .......................
B.3.5 Programs to perform LBP image reconstruction ..
B.3.6 Iterative LBP image reconstruction program ....

201
201
201
204
205
206
206
207
208
208
209

APPENDIX C: Papers produced as a result of this


work ...................................... 210

vi

ABSTRACT

Several ways of improving the performance of electrical


capacitance tomography (ECT) systems are presented and
evaluated, including the use of alternative excitation
schemes, more sensitive and less noisy electronics, and more
accurate image reconstruction algorithms.
The design of a new electrical capacitance tomography (ECT)
data acquisition system is presented, having a number of
improvements over the one previously designed at UMIST. The
new system uses AC-based instead of charge-discharge
capacitance transducers, providing an increase in resolution
from 0.26 to 0.025 fF (peak value of noise level) and a
ten-fold improvement in stray-immunity. Thanks to the use of
AC amplifiers before demodulation the problem of drift is
practically eliminated. Phase-sensitive demodulation is
employed in order to be able to discriminate between the
effects of the conductive and capacitive components of the
unknown admittance. Each channel has its own demodulator thus
allowing parallel measurement. The latter, coupled with
high-frequency (500 kHz) operation, results in a potential
acquisition rate of more than 100 frames per second. Finally,
excitation signals can be applied to several electrodes at
the same time, and, thus, the system can be employed to
investigate the possibility of using multi-electrode
excitation patterns.
The use of parallel-field excitation (attempting to mimic
X-ray tomography) is explored employing finite-element
simulation techniques and found to be disadvantageous. It is
shown that, due to fundamental differences in the underlying
physics, an analogy cannot really be established between
X-ray tomography and parallel-field excitation ECT. On the
other hand, the use of optimal or adaptive excitation
vii

methods,
seeking
to
maximize
the
visibility
or
distinguishability of the permittivity distributions imaged,
were successfully tested against the conventional singleelectrode excitation method, and results are presented for
different permittivity distributions.
Finally, results are presented of an assessment of a new
iterative image reconstruction algorithm based on the quite
singular approach of viewing the reconstruction process as a
feed-back control system.

viii

DECLARATION

No portion of the work referred to in this thesis has been


submitted in support of an application for another degree or
qualification of this or any other university, or other
institute of learning.

ix

ACKNOWLEDGEMENTS

I would like to express my appreciation for their help,


encouragement and valuable advice to my supervisor Dr. R C
Waterfall, Professor M S Beck and Dr. W Q Yang.

I also acknowledge the financial support of the Mexican


Petroleum Institute and the National Council for Science and
Technology of Mexico (Conacyt).

Finally, I thank my wife Sara for her patience and support.

C H A P T E R

1 :

I N T R O D U C T I O N

1.1

OVERVIEW OF ELECTRICAL CAPACITANCE TOMOGRAPHY

Electrical capacitance tomography (or ECT) is one of a


relatively new breed of imaging techniques developed for
industrial process applications, collectively known as
process tomography [1,2]. The aim of all these methods, which
started to evolve in the mid 1980s, is to provide a
non-invasive, non-intrusive means to obtain cross-sectional
images of the interior of process vessels (figure 1.1), which
can be used to control and monitor process operations, or as
a model validation tool in process design.

Fig. 1.1

Process tomography system

Although the use of several tomographic modalities has been


explored, including ionising radiation, magnetic resonance
imaging, ultrasonic and optical techniques, electrical
methods based on impedance measurement have generally proved
more suitable for process tomography applications, being
fast, robust and relatively inexpensive. Electrical impedance
tomography (EIT) can be subdivided into resistance,
inductance and capacitance tomography, depending on the
physical quantity being measured.

Fig. 1.2

Electrical capacitance tomography system

Electrical capacitance tomography is aimed at industrial


processes involving non-conducting materials, or mixtures
where the continuous phase is non-conducting. In a
conventional ECT system (figure 1.2), the sensor takes the
form of a circular array of electrodes placed around an

insulating pipe and surrounded by a grounded screen (the


latter not shown in the figure for clarity). The data
acquisition unit contains capacitance transducers which are
used to determine the capacitance of all possible electrode
pair combinations, thus producing n(n-1) measurements, where
n is the number of electrodes. By means of a computer and a
suitable algorithm, these data must then be used to
reconstruct an image of the permittivity distribution inside
the
sensor,
which
directly
reflects
the
material
distribution.
The first attempts to do capacitance-based tomography were
carried out more or less at the same time (H 1986-88) both at
UMIST [3-5] and the Morgantown Energy Technology Centre (in
the USA) [6-8]. Later on, in 1991, the first real-time ECT
system was developed in a joint project by UMIST, Leeds
University and Schlumberger Cambridge Research [9,10]. ECT
has been applied, at an experimental level, to the on-line
visualisation of gas-oil flows [11,12], as well as imaging
combustion processes [13-17] and fluidised beds [18-20]. For
an excellent review article including numerous applications
and an extensive bibliography (80 references) see [42].

1.2

MAIN AREAS OF IMPROVEMENT

In what follows we identify some of the principal areas


subject to improvement in ECT, which will become the focus
and motivation of this work.
The main challenge encountered in the design of an ECT system
comes from the fact that the capacitances to be measured are
extremely small. For example, for a typical 12-electrode
empty sensor the various inter-electrode capacitances range
from 10 to 1,000 femtoFarads, and the expected full-scale
3

capacitance changes extend from about 5 to 80 femtoFarads,


for 10 cm long electrodes. Therefore, depending on the
particular permittivity distribution, the system may have to
measure accurately capacitance changes of a few tenths of a
femtoFarad. Furthermore, these very low inter-electrode
capacitances have to be measured in the presence of stray
capacitances to ground several order of magnitude larger
(H 100 picoFarads). There is, then, the need for better
high-sensitivity low-noise capacitance transducers, in order
to increase the signal to noise ratio (SNR) and the
resolution of the measurements. With the use of more
sensitive electronics, it would be possible to increase the
spatial and/or axial resolution by employing smaller
electrodes. Alternatively, the higher sensitivity can be
exploited for imaging lean flows.
Another problem area found
i n
E C T
i s
t h a t
inter-electrode capacitances
are much more sensitive to
changes in permittivity near
the
electrodes
than
to
changes occurring near the
centre of the sensor. This
is reflected on the shape of
the sensitivity maps, which
are
basically
graphs
of
Fig. 1.3
Sensitivity map
dC/d for a specific pair of
for opposite electrodes
electrodes. For instance, in
figure 1.3 we can see the
typical sensitivity map for opposite electrodes. It roughly
defines a channel of sensitivity across the sensor area.
However, we can see that the sensitivity is not constant
along this channel; the two peaks correspond to the
electrode positions and the sensitivity decreases towards the
centre. As a result of this non-uniform sensitivity effect,
4

it is much more difficult to see permittivity changes near


the centre than elsewhere in the sensor. The idea of somehow
increasing the sensitivity in the centre to achieve a uniform
sensing area sounds, therefore, very attractive. In the early
stages of this work, it was thought that the problem would be
solved or alleviated if a parallel electric field could be
created inside the sensor by applying specific voltages to
the electrodes
[21,22], although, after
a thorough
investigation of the matter, it was later found that this was
not entirely the case [23]. There are, however, other types
of multiple-electrode excitation that can be used to improve
the overall sensitivity of EIT systems in general, which have
been already successfully tried for resistance tomography
[30,31,33], but not for ECT.
Let us finally say a word about image reconstruction. To
perform an accurate quantitative reconstruction of the
permittivity distribution inside the sensor from the
capacitance measurements is a very complex task, which
mathematically belongs to the category of inverse problems
and normally involves computationally intensive iterative
procedures based on optimisation. However, by linearising the
problem, a qualitative reconstruction can be done using a
simple and fast algorithm known as linear back-projection
(LBP) [4], adapted from medical X-ray tomography. Because of
its simplicity and speed, virtually all ECT systems use the
LBP algorithm. Nevertheless, the quality of the images
obtained with LBP is rather poor, and there is clearly the
need for better reconstruction methods that can provide both
accuracy and speed.

1.3

AIMS AND OBJECTIVES

In view of the considerations presented in the previous


section, the general aims of this work are set as follows:
a) To investigate ways to increase the sensitivity of ECT
systems in the central area of the sensor, particularly
the use of parallel fields and multi-electrode excitation.
b) To develop new capacitance transducers for ECT use, having
more sensitivity and lower noise level than those
currently being employed.
c) To investigate improved reconstruction algorithms based
on iterative LBP methods.
In order to have experimental support, an ECT system will be
designed and built, incorporating the new high-sensitivity
transducers developed.

1.4

ORGANISATION OF THIS THESIS

In chapter 2, the physical situation occurring in an ECT


sensor
is
described,
according
to
the
laws
of
electromagnetism. The theoretical implications of using
various different excitation arrangements are presented, and
a comparison is made between them, showing the unsuitability
of parallel fields. Also, the previous work on ECT system
design in discussed here, and new ways of improving
capacitance transducer sensitivity using AC-based circuits
are proposed.
Chapter 3 presents a critical analysis of the idea of using
parallel-field excitation as a means to increase the
6

sensitivity in the centre of the sensor and reduce the


non-uniformity of the sensitivity maps, in an attempt to
mimic X-ray tomography. The analysis is based on finiteelement simulation using the software package PC-OPERA, and
includes a comparison between parallel-field and conventional
single-electrode excitation.
The design and testing of an AC-based high-sensitivity
capacitance transducer is presented in chapter 4. The design
of a multiple-excitation ECT system based on this transducer
is also described.
Chapter 5 presents results obtained with an ECT system built
after the design presented in chapter 4, including the
experimental evaluation of various single- and multipleelectrode excitation methods.
Chapter 6 discusses the implementation of a new iterative
reconstruction algorithm based on back-projection.
Finally, chapter 7 summarises the main achievements of this
project and gives suggestions about possible areas for future
work.

C H A P T E R

THEORY

2.1

OF

2 :

ELECTRICAL

CAPACITANCE

TOMOGRAPHY

THE ECT SENSOR AS A SYSTEM OF CHARGED CONDUCTORS

In ECT, a number of electrodes are installed around the pipe


or vessel to be imaged, surrounded by a grounded screen. This
is the basic ECT sensor configuration, shown in figure 2.1
for 12 electrodes. The mutual capacitance (defined below) of
the different electrode pairs depends on the permittivity
distribution inside the sensor. When a body is placed in the
sensor there will be a change in the mutual capacitances. The
principle of ECT is to measure the change in mutual
capacitance of the different electrode pair combinations and
then use these measurements to reconstruct an image of the
permittivity distribution in the cross section being
investigated by the sensor. The reconstruction process can be
done with a computer using a simple algorithm known as linear
back-projection (LBP) [4].

Fig. 2.1

Basic 12-electrode ECT sensor


8

From a physics point of view, the ECT sensor can be


considered as a special case of a system of charged
conductors separated by a dielectric medium [24-27], the
theory of which was first developed by Maxwell [28]. In our
particular case, with the sensor electrodes acting as the
charged conductors, the electrode charges Qi and the electrode
potentials Vi are related by the following set of linear
equations for an n-electrode sensor

Q1= c11V1+c12V2++c1nVn
Q2= c21V1+c22V2++c2nVn

Qn= cn1V1+cn2V2++cnnVn

(2.1)

where the coefficients cii are called the self-capacitance of


electrode i, while the others, cij, with i` j, are the mutual
capacitance of electrodes i and j.
Writing equation (2.1) in matrix form we have







Q1
Q2

Qn





=
=





c11
c21

cn1

c12
c22


cn2

c1n
c2n

cnn

V1
V2

Vn





(2.2)




or

Q =

C V

(2.3)

The matrix C is called the capacitance matrix of the system.


The self and mutual capacitances are sometimes called
coefficients
of
capacitance
and
coefficients
of
(electrostatic) induction respectively, and in the very old
books they are termed coefficients of capacity and
coefficients of influence. These coefficients depend only on
the geometry and the permittivity distribution of the system,
and have the following important properties:
a) The self-capacitances are always positive.
b) The mutual capacitances are always negative.
c) For every conductor we have
ci1 + ci2 + + cin e 0
d) For the mutual capacitances
cij = cji

(2.4)
(2.5)

The matrix C is called the capacitance matrix and completely


characterises the system of conductors. C is a non-linear
function of the system geometry and of the permittivity
distribution in the dielectric medium. In our case the
geometry is fixed, so any change in C will be due to a change
in the permittivity distribution.
An ECT sensor (or any system of charged conductors) can also
be modelled using a circuit theory approach, as a network of
component capacitances [29], as illustrated in figure 2.2
using a 4-electrode sensor. To do this, we re-write equation
(2.1) in terms of the voltage difference between the various
electrodes, to put it in the form

Q1= C1V1+C12(V1-V2)+C13(V1-V3)+
Q2= C2V2+C21(V2-V1)+C23(V2-V3)+

10

(2.6)

or, in compact form, for electrode i of n

Qi = CiVi +

j=1
(i j)

(2.7)

Cij(Vi-Vj)

where the component capacitance between electrode i and


ground is given by
Ci = ci1+ci2+ci3++cin
(2.8)
and the inter-electrode
electrodes i and j by

component

Cij = -cij

(i` j)

capacitance

between

(2.9)

Fig. 2.2 Equivalent circuit (based on the component


capacitances) of a 4-electrode ECT sensor

As mentioned earlier, in ECT we are concerned only with the


inter-electrode capacitances (i.e. the mutual capacitances),
11

since this are the ones that depend on the permittivity


distribution inside the sensing area of the sensor. We are
not interested in the component capacitances to ground (i.e.
between the electrodes and the screen) because, due to the
geometry of the sensor (assuming that the inter-electrode
gaps are small), they depend mainly on the permittivity
distribution in the annular region between the electrode ring
and the outer screen, outside the imaging area. We are not
interested in the self-capacitances either because they are
determined by the mutual capacitances and the component
capacitances to ground (see equation 2.8).
Consequently, we are only interested in the mutual
capacitances, only half of which are independent (because of
their reciprocity relationship). So, we can say that all the
information about any change in the permittivity distribution
inside the sensor will be contained in the variations of the
n(n-1) independent mutual capacitances, which form the lower
(or upper) triangular part of the capacitance matrix C.

2.2

ECT MEASUREMENT STRATEGIES

The value of the self and mutual capacitances can be found by


applying known potentials to the sensor electrodes and
measuring
the
electrode
charges.
In
practice,
the
determination on the electrode charges is normally done
indirectly by measuring the electrode currents (Q= i dt), and
the excitation potentials are applied to the electrodes in
the form of a periodic signal of known amplitude.
2.2.1

THE SINGLE-ELECTRODE EXCITATION METHOD

All ECT systems reported in the literature so far, including


those developed earlier at UMIST [10], use single-electrode
12

excitation to measure the mutual capacitances, with the


exception of the Morgantown system, which uses a special
bipolar excitation technique [8]. Let us consider a
12-electrode sensor (figure 2.1). Using the single-electrode
excitation method, the mutual capacitances are determined as
follows: First an excitation voltage is applied to electrode
1 while keeping all the others at zero potential and the
charge on electrodes 2 to 12 is measured. According to
equation (2.1), these measurements directly represent c2 1 to
c12 1 . Next, the excitation voltage is applied to electrode 2
while keeping all the others at zero and the charge on
electrodes 3 to 12 is measured, representing c3 2 to c12 2 . This
procedure is repeated, applying voltage to electrode n and
measuring the charge on electrodes (n+1) to 12, until, as a
final step, voltage is applied to electrode 11 and the charge
of electrode 12 is measured. In this way, the 66 independent
mutual capacitance values corresponding to the lower half of
the capacitance matrix are determined (the other 66 being
given by equation (2.5)), requiring 66 electrode charge
measurements.
From a hardware design point of view, single-electrode
excitation has the advantage of requiring only one voltage
source, which can be switched sequentially to the electrode
being used as a source.
The problem with this method is that, because the mutual
capacitances are so small, the electrode charges can also be
very small, and, as a result, the signal-to-noise ratio (SNR)
of the measurements tends to be rather poor, even when
low-noise measuring circuits are used. From equation (2.1),
it is clear that, if excitation potentials are applied to
more than one electrode, it is possible to obtain larger
electrode charges, although they would no longer be a direct
measure of any particular inter-electrode capacitance.

13

2.2.2

QUALITATIVE IMAGE RECONSTRUCTION FOR SINGLE-ELECTRODE


EXCITATION: THE LINEAR BACK-PROJECTION (LBP) ALGORITHM

In
single-electrode
excitation
ECT
systems,
image
reconstruction for two-component mixtures is done using the
linear back-projection (LBP) algorithm [3]. The basic idea of
this qualitative algorithm, which is an adaptation of a
method used in medical tomography, is to do a weighted
back-project or smearing of each one of the
n(n-1)
normalised measurements along its sensing zone, given by the
corresponding sensitivity map.
2.2.2.1

THE SENSITIVITY MAPS

Let us consider an n-electrode sensor and an image made of m


equal-area pixels. For each pair of electrodes i (source) and
j (detector), a capacitance sensitivity map can be defined by

Si j(k)=

Ci j(k) - Ci j emp


Ci j full - Ci j emp

Qi j(k) - Qi j emp


(2.10)

Qi j full - Qi j emp

where k is the pixel number (from 1 to m), Qi j(k) is the


charge induced
of pixel k is
rest of the
material, Qi j full

on electrode j by electrode i when the region


full of high-permittivity material while the
sensing area is full of low-permittivity
and Qi j emp are the charge induced on electrode

j by electrode i when the sensor is full of high- and


low-permittivity material, respectively, and Cij are the
corresponding mutual capacitances of electrodes i and j under
the same conditions.

14

a) Location of electrodes

b) S13

c) S15

d) S17

Figure 2.3

Typical sensitivity maps for single-electrode


excitation

The sensitivity maps can be determined experimentally for a


particular sensor by probing the sensing area using a test
rod, although this is a very time-consuming task. A more
practical approach is to use computer simulation techniques
to model the sensor using the finite-element method (FEM).
The author chose the latter and, in his work, used PC-OPERA,
a
commercially
available
FEM
software
package
for
electromagnetic analysis and simulation. A more detailed
description of the procedures used to calculate the
sensitivity maps is given in chapter 3 and in appendix B.
15

Regardless of the method used, it is not necessary to


determine the sensitivity maps for all possible electrode
pair combinations, since, due to sensor symmetry, all the
sensitivity maps can be obtained by rotation from (for a
12-electrode sensor) the following basic set of 6: S12 , S13 ,
S14 , S15 , S16 and S17 .
Figure 2.3 shows typical sensitivity maps for several
electrode combinations, which were calculated for imaging oil
and gas mixtures ( high = 2.1 and low = 1), with m = 313. As
might be expected, the sensitivity maps show that each
electrode pair responds mainly to the material lying between
the electrodes, albeit in a very non-uniform way.
2.2.2.2

THE NORMALISED MEASUREMENTS

Prior to back-projection, the measurements obtained with each


electrode pair are normalised according to

Ci j meas - Ci j emp

i j= 

Ci j full - Ci j emp

Qi j meas - Qi j emp


(2.11)

Qi j full - Qi j emp

where i j is the normalised measurement (charge or


capacitance) corresponding to electrodes i (source) and j
(detector), Qi j meas is the measured charge induced on electrode
j by electrode i, Qi j full and Qi j emp are the charge induced on
electrode j by electrode i when the sensor is full of highand low-permittivity material, respectively, and Cij are the
corresponding mutual capacitances of electrodes i and j.
Normalised values are used so that the same software will
cope with systems having different electrode lengths, which
produce different absolute measurements but the same
normalised ones.
16

2.2.2.3

THE WEIGHTED BACK-PROJECTION OPERATION

Mathematically, for an n-electrode sensor and an m-pixel


image, the LBP algorithm calculates the grey level G(k) for
each pixel k as

n 1

ij

G(k)

Si j (k)

i 1 j i 1
n 1 n

(k 1..m)

(2.12)

Si j (k)
i 1 j i 1

where i j are the normalised measurements defined by equation


(2.11) and Si j are the sensitivity maps defined by equation
(2.10). The actual back-projection operation occurs in the
numerator of equation (2.12), while the quantity in the
denominator serves as a position-dependent weighting factor
used to compensate for the decrease in sensitivity towards
the centre of the sensor.
2.2.3

MULTIPLE-ELECTRODE EXCITATION METHODS

Although they have not been used in ECT, multiple-electrode


excitation techniques have been around for quite a while in
other EIT modes and are especially popular in electrical
resistance tomography (ERT) work [30,31], where they can be
found under various names like adaptive, multi-reference,
and

optimal

currents methods.

If there are n electrodes


reason why we should not
voltages to more than one
excitation voltage vectors

in an ECT sensor, there is no


simultaneously apply excitation
of them. In fact we can define
of the form

V = [V1 , V2 , ..., Vn ]
17

(2.13)

If we apply this excitation voltage vector to an ECT sensor


and then measure all the electrode charges, we can form
another vector

Q = [Q1 , Q2 , ..., Qn ]

(2.14)

Let us consider two different permittivity distributions

(p), corresponding to a known reference state such as an


empty sensor, and (p), corresponding to some unknown
material distribution, where p denotes a point inside the
sensor. When we apply a voltage vector V to these two
permittivity distributions we get two charge vectors Q and
Q. Let us now define our measurement signal as the change in
the electrode charges due to the permittivity distribution
changing from to , i.e.

Q

= Q- Q

(2.15)

We can then define a vector of measured signals m as

m = Q- Q = [Q1 ,

Following the ideas


distinguishability of
with V as

m Q2 =

of


Q- Q Q2 =

Q2

, ...,

Qn

Issacson [30],
with respect to

[Q1 ,

18

Q2

, ...,

(2.16)

we define the
when exciting

Qn

]Q2

(2.17)

additionally, from equation (2.3) we have

Q- Q Q2 =

(C- C )V Q2 =

(2.18)

QDVQ2

where C and C are the capacitance matrices corresponding to

and

,

respectively,

and

D = C- C

is

called

the

distinguishability matrix.
Equation

(2.17)

shows

that

is

an

indicator

of

the

magnitude of the detection signals (defined as the change in


electrode charge Q). Obviously, the larger the detection
signals the easier it will be to detect the change in
permittivity distribution from to . On the other hand,
equation (2.18) shows that

depends on the distingui-

shability matrix (which ultimately depends on

and ),
but, more importantly, also on the excitation vector V.

In other words, for given permittivity distributions

and

,

not all excitation vectors will produce the same signal


level, and there will be some excitation vectors that are the
best choice to distinguish between and , in the sense
that they will produce the largest detection signals, and,
therefore, the best signal-to-noise ratio (SNR). This is the
main justification for the use of multiple-electrode
excitation, since with single-electrode excitation the choice
of excitation vectors is limited to those having only one
non-zero element.
Extending to ECT the result obtained by Issacson in [30], we
can say that the best voltage vectors to distinguish
between and  are the eigenvectors of D having the
largest eigenvalue. In general, excitation patterns having
a low spatial frequency will yield measurements which are
more sensitive to changes in permittivity near the centre,
19

while high frequency patterns yield measurements sensitive


mainly to changes near the electrodes [30].
Now, although the multiple-electrode excitation method yields
measurements with the best SNR, there is a price to pay both
in terms of the system hardware and software. Firstly, the
hardware becomes more complex and expensive, since it must
now include n independent voltage sources. Secondly, unlike
with single-electrode excitation, what is measured is no
longer capacitance but the response of the sensor (in the
form of changes in electrode charge Q) to a set of
excitation vectors Vk, with k = 1, ..., L. Because of this, the
practical LBP algorithm, which is based on capacitance
measurements, cannot be used, at least not directly.
As it will soon become clear, it can be advantageous to
derive an expression for the measurements Q as a function of
the inter-electrode voltages, and to re-define the excitation
vectors in terms of the latter instead of the electrode
voltages themselves. We shall, for the sake of simplicity,
use a 4-electrode sensor (figure 2.2) to illustrate these
concepts, although the same ideas apply fully to sensors with
any number of electrodes.
From equation (2.7) we can arrive at the following system of
equations describing the sensor

Q1

C 1V1

C 12 (V1-V2)

C13

(V1-V3) +

C14

(V1-V4)

Q2

C 2V2

C 21 (V2-V1)

C23

(V2-V3) +

C24

(V2-V4)

=
Q4 =

C 3V3

+
C 4V4 +

C 31 (V3-V1)

+
C 41 (V4-V1) +

C32

(V3-V2) +
C42 (V4-V2) +

C34

Q3

(V3-V4)
C43 (V4-V3)

(2.19)

Now, because of the geometry of the sensor (assuming that the


inter-electrode gaps are small) we have Ci H 0, i = 1, .., n.
20

In other words, the change in the capacitances to ground is


negligible because they do not depend on the permittivity of
the material inside the sensing area, but only on that of the
material in the region between the electrodes and the outer
screen. Therefore, equation (2.19) becomes

=
Q2 =

C 12 (V1-V2)

+
C 21 (V2-V1) +

C13

Q3

C 31 (V3-V1)

Q4

C 41 (V4-V1)

Q1

(V1-V3) +
C23 (V2-V3) +

C14

(V1-V4)
C24 (V2-V4)

C32

(V3-V2) +

C34

(V3-V4)

C42

(V4-V2) +

C43

(V4-V3)

(2.20)

For our 4-electrode sensor (n = 4) we can then define the


following excitation vectors, of size n2-n:

U = [ (V1-V2),(V1-V3),(V1-V4),(V2-V1),(V2-V3),(V2-V4),
(V3-V1),(V3-V2),(V3-V4),(V4-V1),(V4-V2),(V4-V3) ]

(2.21)

In equation (2.20) we can see that our measurement signals,


considered as the change in charge Qi, do not depend on Vi,
the actual voltage on the measuring electrode, but only on
the voltage differences between the electrodes. However, in
practice the system cannot measure Qi directly, it has to
measure
Qi

Qi

and

Qi

separately

and

from

them

calculate

= Qi- Qi . From equation (2.7) we see that the voltage Vi

on the measuring electrode can have a considerable effect on


its charge Qi, since the capacitances to ground Ci (which in
an actual system include the capacitance of the cable
connecting the electrode to the instrument) are much larger
than the inter-electrode ones, whose effect will be obscured.
In order to avoid having to measure the electrode charge due
to Ci, the voltage Vi on the measuring electrode should be set
21

to zero. The advantage of defining the measurement signals Q


and the excitation vectors U in terms of the inter-electrode
voltages is that, once the most suitable vectors have been
chosen (according to some criterion), the voltage on the
measuring electrode can always be set to zero and the rest of
the electrode voltages adjusted accordingly to satisfy the
particular excitation vector being applied.
2.2.3.1

RECONSTRUCTION WITH MULTIPLE-ELECTRODE EXCITATION

The idea of multiple-electrode excitation brings about the


important question of how to reconstruct an image from the
knowledge of the applied excitation vectors and the vector of
measurements. For a quantitative reconstruction directly from
the charge measurements, we have to resort to iterative
algorithms like the those used in electrical resistance
tomography (ERT) [32,33], which are normally based on some
variant of Newtons method.
The formal development of iterative image reconstruction
algorithms based on optimisation is a vast and complex task
that could itself be the subject of another PhD thesis,
involving a considerable amount of advanced mathematics, and
is not within the scope of this work. The main objective of
this work, as far as multiple-electrode excitation is
concerned, is to design and build an actual system and use it
to confirm experimentally that multiple excitation can indeed
improve distinguishability and the SNR of the measurements
(chapters 4 and 5), a purpose which can be achieved without
having to produce any images. Nevertheless, without going
into details, we shall present a general description of some
of the methods that can be used for iterative image
reconstruction.
Let us assume that we apply a set of L = n - 1 linearly
independent inter-electrode voltage excitation vectors Uk,
22

k = 1, .., L. The excitation vectors are applied one by one


and, for each vector, the change in electrode charge

is
measured for every electrode. So, for each excitation vector
Uk we can form an n-dimensional vector of measurements m k like
Q

the one in equation (2.16). The total number of independent


measurements is, therefore, N = nL = n (n - 1), and they can be
stacked in a long vector

M = [m 1 ,m 2 , ...,m N ].

(2.22)

We know that M is a linear function of the excitation


vectors, which can be stacked in a vector E = [U1, U2, ..., Uk ],
and a non-linear function of the reference and unknown
permittivity distributions and , i.e. M = (E, , ).
As shown in figure 2.4, iterative algorithms start with an
initial guess of the unknown permittivity distribution, 0,
which

is

then

used

to

calculate

(E,

, 0)

using

finite-element model of the sensor. Then the iterative part


commences, by comparing the simulated measurements with the
actual ones, the difference between the two being calculated
as i = Q (E, , i) - M Q2, i = 0, 1, 2, ...; if i is smaller
than some specified tolerance then we can consider

i

as the

true permittivity distribution, otherwise this error is fed


to an optimiser (based on some Newton-like formula) that
produces a new estimated permittivity distribution i+1 ,
which is used to calculate the next simulated measurements.
The cycle is repeated until the error is within the accepted
tolerance. Despite requiring great computing power and being
relatively slow,iterative algorithms like this produce
quantitative reconstruction, unlike LBP. One serious problem,
however, is that the noise level of the measurements can
severely affect the algorithms convergence and accuracy. The
better SNR achieved through multiple-electrode excitation
would be particularly useful in alleviating this problem.
23

Fig. 2.4

General flow diagram of iterative reconstruction


algorithms
24

It really seems rather strange that iterative image


reconstruction algorithms based on optimisation theory, which
are mathematically sound and yield quantitative results, have
not been applied to the specific case of ECT, whereas in ERT
they are very popular and there has been a lot of research
into the subject, with the publication of many papers and
several PhD theses.
An alternative approach to image reconstruction for
multiple-electrode excitation ECT systems involves using
equation (2.20) to recover the inter-electrode capacitance
changes Cij from the measurements M and the inter-electrode
excitation voltages E. This can easily be done by solving n
systems of n - 1 linear equations each. Once we have the
inter-electrode capacitance changes

Cij

, which represent the

equivalent single-electrode excitation measurements, we can


use the standard LBP algorithm to perform a qualitative image
reconstruction.
2.2.3.2

OPTIMAL MULTIPLE-ELECTRODE EXCITATION PATTERNS

The best inter-electrode voltage vectors U

are the ones

that maximise the changes in electrode charge Q (equation


(2.20)), and they can be determined from the distinguishability matrix D = C-C .
Let us see, first, how the D matrix can be determined. For
i` j, we have Dij = cij- cij . For i=j, we have Dii = cii- cii , but
from equation (2.8)

cii = Ci -

j=1
(i j)

therefore

25

cij

(2.23)

Dii = Ci-

j=1
(i j)

cij - Ci +

j=1
(i j)

cij

(2.24)

Recalling that the capacitances to ground do not depend on


the permittivity distribution inside the imaging area, we
have Ci = Ci , so they cancel out and equation (2.24) becomes

Dii = -

j=1
(i j)

cij -

j=1
(i j)

cij

(2.25)

or

Dii = -

j=1
(i j)

Dij

(2.26)

From the foregoing discussion we can conclude that the matrix


D is completely determined by the mutual capacitances cij and
cij (with i` j) and does not depend on the self-capacitances
cii and cii . The mutual capacitances can easily be determined
using the single-electrode excitation method described in
section 2.2.1. In a practical situation, cij , the mutual
capacitances for the reference state (empty sensor), could
measured in advance and stored, so that it would only
necessary to measure the mutual capacitances corresponding
the unknown permittivity distribution  in order
determine D.

26

be
be
to
to

As mentioned earlier, the optimal excitation voltage vectors


are the eigenvectors of D corresponding to the largest
eigenvalues [30]. Many methods can be used to obtain the
eigensystem of the symmetric matrix D, like the QR algorithm,
singular value decomposition (SVD) and the Jacobi method. The
SVD method is preferred because it is stable and
straightforward to obtain [31]. Using SVD we factorise D as

D = X

YT

where X and Y are orthogonal matrices and

(2.27)

is a diagonal
matrix whose entries are the singular values. The
eigenvectors of D (i.e. the optimal excitation voltage
vectors) are given by the columns of X, while its eigenvalues
are equal to the singular values. The best excitation vector
is the eigenvector corresponding to the largest eigenvalue.
The last eigenvalue is always equal to zero, since the rank
of D is n - 1, and the corresponding excitation vector is not


used. If necessary, the electrode voltage vectors thus


obtained can be re-scaled, in order to fully exploit the
dynamic range of the voltage sources employed.
In this way we end up with an optimal set of orthogonal
n-dimensional
electrode-voltage
unit
vectors
Vk ,
k = 1, .., (n - 1), which will produce the largest measurements
for a particular permittivity distribution

compared with
other vectors of unit length. It was shown earlier that,
because of the need to always set the measuring electrode
voltage to zero, it is more convenient to work with
inter-electrode voltages. The optimal inter-electrode voltage
vectors Uk
can be obtained from Vk using equations (2.13)
and (2.21).

27

,

2.3

ECT TRANSDUCERS

The function of an ECT transducer is to measure the charge Q


on a detection electrode. On a single-electrode excitation
system, where only two electrodes are involved in each
measurement, this charge Q is also a measure of the
capacitance between the two electrodes, given by Cij = - Qi /Vj .
Among the desired characteristics of an ECT transducer are:
a) It must be stray-immune. This means that the capacitance
between the measuring electrode and ground should not have
an effect on the measurement. This is normally achieved by
ensuring that the detection electrode potential is
maintained at zero during measurement (thus making Vi = 0
in equation (2.7)).
b) Its range has to match the sensor and the characteristics
of the materials to be imaged. For example, for a typical
12-electrode sensor used to image oil ( r = 2.1) and gas
( r = 1) mixtures, the inter-electrode capacitances can lie
anywhere between 10 and 600 femtoFarads approximately.
c) It should be able to measure small inter-electrode
capacitance changes in the presence of large standing
values. Again for the same situation as in (b), we have
that, depending on the particular electrode pair, the
full-scale capacitance change can be as low as 15% of the
standing capacitance. The transducer must have some means
of balancing these standing capacitances.
d) The resolution must be high enough, and the noise level
low enough, to allow the detection of capacitance changes
of a few tenths of a femtoFarad.
e) If the system is to be used in fast real-time applications
like flow imaging, the transducer must have a fast dynamic
28

response. Considering a frame rate of, say, 100 frame/s,


the time available for collecting all the data for one
frame is 10 ms.
2.3.1

THE CHARGE-DISCHARGE CAPACITANCE TRANSDUCER CIRCUIT

The single-electrode excitation ECT system developed at UMIST


uses capacitance transducers based on the charge transfer
principle [9,34]. This type of transducer is stray-immune
(i.e. they are insensitive to the capacitances to ground) and
its main attractive is its simplicity and relatively low
cost.

Fig. 2.5 The charge-discharge capacitance measuring circuit

The charge transfer transducer circuit is shown in


figure 2.5. The device works by repeatedly charging and
discharging the unknown capacitance Cx through the combined
action of semiconductor switches S1 to S4. First, S1 and S2 are
29

closed (keeping S3 and S4 open) to charge Cx to voltage Vc, and


the charging current flows into the current-to-voltage
converter formed by operational amplifier (op-amp) A1 and its
feed-back resistor Rf, causing a negative output voltage. In
the second half of the operating cycle, S1 and S2 are open
while S3 and S4 are closed, thus discharging Cx to ground. The
discharge current flows out of the current-to-voltage
converter formed by op-amp A2 and its feed-back resistor Rf,
producing a positive output voltage. This charge-discharge
cycle repeats itself at a frequency f (up to 2 MHz) and the
successive charging and discharging current pulses are
averaged in the two current detectors, producing two dc
output voltages:

V1 = - f Vc Rf Cx + e1

(2.28)

V2 = f Vc Rf Cx + e2

(2.29)

where e1 and e2 are the output offset voltages of the current


to voltage converters. Since the detection electrode of Cx is
always connected to a virtual earth point, any stray
capacitance to ground Cs1 is always short-circuited and has no
effect on the measurement. And, because the excitation
electrode is always being driven by a low-impedance voltage
source, its stray capacitance Cs2 has no effect either.
The voltage difference V2 - V1 is taken as the output, giving

Vo = V2 - V1 = 2 f Vc Rf Cx + e2 - e1

(2.30)

this has the advantage of doubling the sensitivity and that


the offset signals e1 and e2 tend to cancel each other.
30

Capacitors C at the input of the op-amps ensure a stable


virtual earth during the fast charge and discharge of Cx.
The bandwidth of the circuit is set by Rf and Cf through
1
B =



(2.31)

2 Rf Cx
and the maximum bandwidth achievable with a particular op-amp
was calculated by Huang [35] as

Bmax = 0.1

A
#
# 
T(C+Cf)Rf
#

(2.32)

where A is the open-loop gain and T the open-loop time


constant of the op-amp.
The main limitation of the charge-discharge transducer is
that it has a lower signal-to-noise ratio than AC-based
circuits. For instance, in the system developed at UMIST,
which uses a sensor with 12 10-cm-long electrodes, the peak
noise level at the systems output is equivalent to an input
capacitance 0.26 femtoFarads, with a transducer bandwidth of
about 10 kHz. This figure essentially sets the resolution of
the system, and corresponds to a change of 2% in the gas void
fraction of an oil/gas mixture occurring in the middle of the
sensor (which is the least sensitive area) [9]. Clearly,
under these conditions, the measurements will not be too
reliable or accurate, since the signal-to-noise ratio is
equal to 1. If we want to use shorter electrodes in order to
reduce the averaging effect along the axial direction, or
increase the number of electrodes to, say, 16 in order to
31

increase the image spatial resolution, then the electrode


area would be reduced and more sensitive transducers would be
required to detect the smaller capacitances produced. The
same is true for imaging mixtures with low permittivity
contrast or lean flows.
Another drawback of the charge-discharge transducer is that
it is susceptible to the effects of conductance losses, i.e,
it is not phase-sensitive. This can be a problem in
applications involving water or in flame imaging, for
example.
Finally, this transducer suffers from charge injection caused
by the feed-through of gate control signals in the switching
semiconductor devices. This charge injection appears as an
offset voltage at the output, which is the main component of
e1 and e2 in equation (2.30). The effect is temperaturedependent and, for a temperature change of 15C, it is
equivalent to an input capacitance of up to 5 femtoFarads
[9]. These offsets cause a baseline drift and, in order to
maintain accuracy, they need to be periodically monitored and
compensated.
2.3.2

AC-BASED ECT TRANSDUCERS

The charge-discharge transducer effectively uses square-wave


excitation. Although this type of excitation signal has the
advantage of being easily implemented by switching between
two DC levels, it has been recognized that by employing
AC-based circuits, i.e., circuits based on sine-wave
excitation, a higher signal to noise ratio can be achieved
[9,35]. The main reason for this is that the use of
single-frequency excitation allows the use of narrow-band
filtering techniques based on phase-sensitive demodulators,
which can greatly reduce the noise bandwidth [45] and also
provide the means to discriminate between capacitive and
32

conductive effects. Additionally, switch-related problems


like charge injection and the generation of glitches are
eliminated. Likewise, drift is no longer such a big concern.
Although there are numerous AC-based methods to measure small
capacitance values (for a good review of many of them see
[35]), not all of them lend themselves well to a
multi-channel application like ECT. In particular, we are
interested in transducer circuits based on the use of an
operational amplifier as a current detector [36-40], an area
relatively new compared with other methods.
In order to be consistent with the concept of multipleelectrode excitation we shall consider the ECT transducer as
a charge rather than capacitance sensor. Figure 2.6 shows a
charge detector (or charge amplifier) [41] based on an
op-amp, which will be used as the basic component for the
design of a more sensitive ECT transducer in chapter 4.

Fig. 2.6

Basic charge detector

33

Essentially, the circuit consists of an operational amplifier


(op-amp) with capacitive feedback. The resistor R provides a
path for the op-amps DC bias current to avoid saturation of
the device due to that current charging Cf . It has negligible
effect on the op-amp output at the frequency of operation.
Assuming that amplifier has a high gain at the frequency of
operation, in the sinusoidal steady state, the circuit is
described by Vo (j) = - Zf Iin(j) = j Iin(j)/
) Cf where Iin(j)
is the input current. But Iin(t) = - dQ(t)/dt
Q(t)
, or in the
frequency

domain

Iin(j) = - j Q(j)

where

Q(j)

is

the

) f.
electrode charge. Therefore we have that Vo(j) = Q(j)/C
Thus, the circuit can be considered as an AC charge to
voltage (Q-V ) converter. Because of the feedback action of
the op-amp, this circuit has the important advantage of
keeping the measuring electrode at virtual earth, avoiding
the appearance of the comparatively large charge due to the
stray capacitance to ground (C1 in figure 2.6). In other
words, the circuit is stray immune.
This simple circuit measures the charge induced on the
detection electrode. If only one electrode is used for
excitation with a voltage Vexc (i.e. in the single-electrode
excitation method) we can get the capacitance between the
electrode pair from Cx = - Q /Vexc , and the circuit can be used
as a capacitance transducer with its output given by
Vo = - (Vexc /Cf ) Cx . The equivalent circuit for this case is
shown in figure 2.7. The capacitance to ground of the
detection electrode, CD , has no effect on Vo , since the
voltage across it is very close to zero, whereas that of the
excitation electrode, CE , is driven by a low-impedance
voltage source and so it does not affect Vo either. However,
if more than one electrode are used simultaneously for
(multi-electrode) excitation then we cannot say precisely
which capacitance is being measured, and it is more
convenient to view the circuit as a charge detector.
34

Fig. 2.7

The charge detector as a capacitance meter

A full analysis of the capabilities and limitations of the


proposed charge detector circuit will be presented in
chapter 4, which shows the detailed design of an AC-based
multiple-electrode excitation ECT system.

2.4
2.4.1

DISCUSSION AND CONCLUSIONS


SINGLE- VS. MULTIPLE-ELECTRODE EXCITATION

It has been shown earlier that multiple-electrode excitation


using optimal inter-electrode voltage vectors U can produce
larger detection signals

and, thus, increase the SNR of


measurements compared with single-electrode excitation. This,
Q

35

however, comes at the cost of more complicated and expensive


hardware, since n independent voltage sources are required.
Another point to consider is parallel measurement. In
single-electrode excitation systems all electrodes but one
are kept at zero volts, and they can all be put to measure
simultaneously (i.e. parallel measurement). This is no longer
achievable with multiple-electrode excitation, because, for
any particular application of an inter-electrode voltage
vector U, generally only a few electrodes will be at zero
potential (i.e. available for measurement) and, therefore,
each inter-electrode voltage vector will actually have to be
applied several times, until all electrodes have had a chance
to be at zero volts and be measured. This, of course, takes
time and has the effect of slowing down the system.
Considering the foregoing, it is the view of the author that,
although multiple-electrode excitation can indeed be very
useful in special applications where the absolute maximum
sensitivity is desired, for industrial process applications
where low-cost and speed of operation are important,
single-electrode excitation ECT systems, with its straightforward approach to both hardware design and image
reconstruction (LBP algorithm), would be a more sensible
choice.
2.4.2

CHARGE-DISCHARGE VS. AC-BASED TRANSDUCERS

The use of AC-based ECT transducers result in better noise


performance, and hence, better resolution and SNR. Later on,
in chapter 4, a system based on this type of transducer will
be designed and the results of its experimental evaluation
presented, showing that a 10-fold improvement in SNR over the
charge-discharge system is possible in practice. Once again,
there is a price to pay, and it comes in the form of more
complex and expensive electronics, mainly due to the fact
36

that one demodulator per channel is required in order to have


parallel measurement capability for fast operation (see
chapter 4).
Despite its higher cost, AC-based transducers are a good
choice in application that require higher resolution and SNR,
like when working with low-contrast mixtures or with lean
flows, or in systems where the electrodes are small, either
because they are short (say less that 10 cm) for better axial
resolution, or because there are a large number of them (more
that 12, say 16) for improved spatial resolution.

37

C H A P T E R

3 :

FINITE-ELEMENT

SIMULATION

PARALLEL-FIELD

EXCITATION

3.1

OF

SINGLE-ELECTRODE

AND

INTRODUCTION

Of
the
many
possible
multiple-electrode
excitation
arrangements, so-called parallel-field excitation deserves
special attention. At the beginning of this work, some ECT
researchers shared the idea that the problem of low
sensitivity in the
centre of the sensor
and the
non-uniformity of the sensitivity maps (figure 2.3) had
something to do with the uneven distribution of electric
force lines which occurs when single-electrode excitation is
employed [21,22]. This type of excitation results in the
electric field being very strong near the excitation
electrode, rapidly weakening as we move away (as shown in
figure 3.4). It was thought that the ideal situation would
rather be to have a parallel electric field uniformly
distributed across the entire sensing area. By so trying to
mimic X-ray tomography, it was believed that increased
sensitivity in the central region would be achieved and also
that the quality of the reconstructed images could be
improved.
In this chapter, we show how parallel-field excitation can be
realised by applying specific excitation voltages to all
electrodes in the sensor. We present the results of
simulation experiments carried out to determine what effects
would the use of parallel-field excitation have, both on the
shape of the sensitivity maps and on the reconstructed
38

images, compared to single-electrode excitation. Finally, we


consider the question of whether or not parallel field
excitation can provide a complete set of independent
measurements required to determine the sensor mutual
capacitances.

3.2

FINITE-ELEMENT MODELLING OF ECT SENSORS

In order to investigate the effects of parallel-field


excitation, simulation experiments were carried out using
finite-element (FE) models of the sensor. PC-OPERA, a
commercially available software package for 2-dimensional
electromagnetic field analysis based on the finite-element
method (FEM), was used to perform the simulation. Note that,
because of the two extra cylindrical guarding electrodes
(grounded) used in the actual sensor on each side of the
sensing electrodes in the axial direction (figure 3.1), 2-D
simulation can be used to model the sensor [4], albeit we are
restricted to work with 2-dimensional material distributions.

Fig. 3.1

Side view of ECT sensor showing guard electrodes


39

We are going to use the FEM package to solve the following


basic problem:
Given a number of conducting bodies with known applied
potentials placed in a dielectric medium with a known
permittivity distribution, find the resultant electric
field distribution and the self and mutual capacitances
characterising the system
Essentially what PC-OPERA does is to find the value of the
electric potential at a large number of points or nodes in a
mesh of contiguous triangular elements used to represent the
actual physical system. The potential data can then be used
to calculate the electric field vectors (by E = -  V ) or
other parameters of interest.
Given a relative permittivity distribution

(x,y), PC-OPERA

finds the potential distribution V(x,y) by numerically


solving the following partial differential equation (where o
is the free-space permittivity)

 [ o (x,y)  V(x,y)] = 0

(3.1)

subject to the corresponding Dirichlet conditions (known


potentials on the boundary).
Working with PC-OPERA involves the following three steps:
1) First, in the preprocessing stage, a model of the problem
is generated using an interactive pre-and-postprocessor
program. In this phase two files are generated, one with
extension "MES" which contains the geometry of the mesh,
and another with extension "OP2" which contains the
boundary conditions, material properties, etc.

40

2) Second, the static analysis program is used to generate


the solution (i.e. the potential distribution). The input
to this program are the two files created in the previous
step, while its output is a solution file with extension
"ST".
3) In the third and final step the solution is viewed and the
parameters of interest (i.e. electrode charge or
capacitance) are calculated, using once again the pre-andpostprocessor program.

The main parameter of interest, in our case, is the detection


electrode charge Q, which can be calculated using Gauss Law:

Dds
s

(3.2)

however, since we are working in two dimensions, we will not


integrate over a closed surface, but over a closed line
around the electrode. It will not be a surface integral but
a line integral, which is evaluated using one of the
pre-and-postprocessor commands. And, of course, Q then
represents the charge per unit length.
Once the charge is known, the capacitance per unit length can
be easily found (for single-electrode excitation) as
C = - Q /Vexc , where Vexc is the voltage on the excitation
electrode.
PC-OPERA can be run interactively or in an off-line mode.
Using this option, the three basic steps mentioned earlier
are automatically executed in sequence. First, a series of
optional pre-and-postprocessor commands (contained in a
command input file) are executed on the specified input
model, then the solution program is run, and finally another
41

set of pre-and-postprocessor commands (contained in another


command input file) is executed on the solution file to
calculate the parameters of interest.
For our work, two FE models of a 12-electrode ECT sensor were
created as shown in figures 3.2 and 3.3. In this type of
2-dimensional
problems
the
absolute
dimensions
are
irrelevant, so we state the geometric characteristics of the
model in terms of the inner radius R of the pipe (i.e. the
radius of the imaging area). In this way, the thickness of
the pipe wall is 0.1R, and the distance between the external
side of the pipe and the outer screen is 0.2R. The relative
permittivity of the region between the pipe and the screen is
set to 1 (air), while that of the pipe is set to 2.5
(perspex). The electrode angle is 26, with inter-electrode
gaps of 4.

a) Model geometry

Figure 3.2

b) FE mesh (5881 elements)

Basic FE sensor model

42

a) Model geometry

Figure 3.3

b) FE mesh (9912 elements)

FE sensor model with polar grid

One of the models (figure 3.3) includes a polar grid composed


of 313 equal-area regions and is used in the determination of
the sensitivity maps required for the LBP algorithm. The
permittivity of each one of this regions can be independently
set to any value, hence allowing the simulation of arbitrary
material distributions. A polar grid was chosen instead of
square one because of its particular symmetry, which allows
the model behaviour to be orientation-independent.
By setting the proper boundary conditions, the application of
any excitation voltage vector can be simulated.

43

3.3

SINGLE-ELECTRODE EXCITATION

Figure 3.4 shows the simulation results for single-electrode


excitation. The excitation voltage is applied to electrode
one and the figure shows the equipotential lines. Clearly,
with this type of excitation the electric field distribution
is quite uneven, with the field concentrated near the
excitation electrode.

Fig. 3.4

Equipotential lines for single-electrode


excitation

In order to calculate the sensitivity maps (as defined by


equation (2.10) in chapter 2), a program was written in
QuickBasic 4.5, which iteratively runs PC-OPERA off-line
using the 313-region polar-grid sensor model of figure 3.3
with the electrode potentials set for single-electrode
excitation. At the start of the ith iteration, the QuickBasic
program generates a command input file that will be used by
PC-OPERA to set the relative permittivity of in-pipe region
i equal to 2.1 ( oil ) while that of all the others will be set
44

to 1. Then the program calls PC-OPERA off-line, which finds


the solution and calculates the charge on all electrodes
using another command input file previously written for this
purpose. The process is repeated 313 times (for each one of
the in-pipe regions). In this manner, all the sensitivity
maps associated with electrode 1 are obtained (i.e. S1 1 to
S1 12 ). The full program can be found in appendix B. The
sensitivity maps associated with the electrodes 2 to 12 are
obtained by rotation of those calculated for electrode 1.
Finally we end up with 144 sensitivity maps, of which only 66
are really needed for use with the LBP image reconstruction
algorithm (one for each measurement, see sections 2.2.1 and
2.2.2 in chapter 2).
Typical examples of sensitivity maps for single-electrode
excitation obtained in the way described above are shown in
figure 2.3 (chapter 2). It can be observed that the detection
areas of the sensor form clearly defined channels between
the detection and excitation electrodes. This is a desirable
feature in a tomography sensor, since each detector looks
only to a specific area. However, the single-electrode
excitation ECT sensor departs from the ideal situation in
that:
a) The

channels are not straight, and

b) Their height is not constant, that is to say, the


response of the detectors is lower in the middle of the
channel and higher near the electrodes.

3.4

IDEAL SENSITIVITY MAPS

In an ideal situation, we would like to see something similar


to the case of parallel-beam X-ray tomography, where each
45

detector is sensitive over a


channel of constant height.

very

narrow

and

straight

For an ECT sensor, though, the sensitivity channels could


not be narrowed too much without having to reduce the
electrode width to such an extent that measurement signals
would become undetectable. However, it might seem natural to
think that if multiple-electrode excitation is used so as to
create a parallel electric field inside the sensor, at least
sensitivity maps forming straight and uniform channels
could be obtained, probably something similar to figure 3.5.

a) Location of electrodes

b) Electrode 6

Figure 3.5
Ideal
ECT
sensor
sensitivity
maps
intuitively
expected
with
parallel-field
excitation (field parallel
to the x axis)

c) Electrode 5

46

In order to verify the previous conjecture, in the next


sections we shall describe how a parallel field can be
created inside an ECT sensor, and the actual characteristics
of the resulting sensitivity maps, which were calculated by
FE simulation, will also be reported.

3.5 PARALLEL-FIELD EXCITATION


3.5.1

PARALLEL-FIELD GENERATION

A parallel field can be approximated inside an n-electrode


ECT sensor by applying electrode potentials according to

Vi =

sin (i - )

n
MAX [ sin(i - )]

(3.3)

i=1

where, referring to figure 3.6, i indicates the electrode


number (1 to n), is the angle between the field direction
and the y axis, E is a voltage constant (determined by
hardware constraints), and i is the angular position of the
centre of the ith electrode, given by
360

i =  ( i - 0.5 )

(3.4)

By using this voltage distribution, the potential difference


between pairs of electrodes facing each other in the
direction of the field is made proportional to the separation
between their centres.

47

Fig. 3.6

Parallel electric field inside an ECT sensor

For example, to produce a parallel field along the y axis


( = 0) on a 12-electrode sensor, with E = 15 V, the voltages
shown on table 3.1 would have to be used.

Table 3.1

Electrode potentials for parallel-field


excitation

,e,
 ELECTRODE  VOLTAGE Q ELECTRODE  VOLTAGE 
<k<$

1

4.02
Q
7

-4.02

<k<$

2

10.98
Q
8
 -10.98

<k<$

3

15.00
Q
9
 -15.00

<k<$

4

15.00
Q
10
 -15.00

<k<$

5

10.98
Q
11
 -10.98

<k<$

6

4.02
Q
12

-4.02

4h4
48

The generation of a parallel field inside the sensor using


the method described above was confirmed by FE simulation as
shown on figure 3.7.

Fig. 3.7

Equipotential lines for parallel-field


excitation using the voltages of Table 3.1

By shifting the potentials one electrode position, the field


can be rotated and different projections (in the sense of
conventional X-ray computed tomography) can be defined, each
one associated with a particular direction of the field. For
a 12-electrode sensor, rotating the field in this way means
that we can define six different projections (numbered 1
to 6), corresponding to equal to 0, 30, 60, 90, 120
and 150. Note that the remaining six projections (180,
210, 240, 270, 300 and 330) do not contribute any
additional information since they are just a sign-changed
repetition of the first ones. For every projection, the
signal from each one of the twelve electrodes can be
measured, giving a total of 6 12 = 72 measurements.
49

3.5.2

SENSITIVITY MAPS FOR PARALLEL-FIELD EXCITATION

The sensitivity map for parallel-field excitation is defined


as follows, for projection i and electrode j:

Si j(k)=

Qi j(k) - Qi j emp



(3.5)

Qi j full - Qi j emp

were k = 1, .., 313 is the region (or pixel) number, Qi j(k) is


the charge of electrode j in projection i when region k is
full of high-permittivity material and the rest of the
sensing area is full of low-permittivity material, while Qi j full
and Qi j emp are the charge of electrode j in projection i when
the sensor is full of high- and low-permittivity material,
respectively.
The sensitivity maps were calculated for projection 1 using
the same QuickBasic 4.5 program described in section 3.3,
which runs of PC-OPERA iteratively, but this time with the
electrode potentials set for parallel-field according to
table 3.1, and are shown in figure 3.8. The 12 sensitivity
maps for projection 1 were then rotated to obtain those for
projections 2 to 6, giving a total of 72 sensitivity maps.
Unfortunately, it can be seen in figure 3.8 that the actual
sensitivity maps for parallel-field excitation bear no
resemblance whatsoever with the ideal ones of figure 3.5.
Not only that, but the sensing areas for each electrode are
much less clearly defined than in the single-electrode
excitation maps, and, in any case, the sensitivity in the
middle is still very poor compared to that near the detection
electrode, all of which seems to indicate that the
parallel-field sensor is less convenient for doing tomography.
50

a) Location of electrodes

b) Electrode 4

c) Electrode 5

d) Electrode 6

Figure 3.8 Typical sensitivity maps for parallel-field


excitation, shown for projection 4 (field
parallel to the x axis). The maps for the
remaining electrodes can be obtained by
reflection relative to the x and y axes

The
simulation
results
shown
experimentally by Yang et al [23].

51

above

were

confirmed

3.6

COMPARISON OF IMAGES OBTAINED USING SINGLE-ELECTRODE AND


PARALLEL-FIELD EXCITATION

In order to test the suitability of parallel-field excitation


for tomographic imaging, FE simulation experiments were
carried out. In these experiments, the presence of an object
inside the sensor was simulated and the charge of the
electrodes was calculated using PC-OPERA, for both singleelectrode and parallel-field excitation.
To do these experiments, two programs were written in
QuickBasic 4.5, one for single-electrode and another for
parallel-field excitation, following the same strategy as in
the program that calculates the sensitivity maps (see
section 3.3), i.e., to iteratively run PC-OPERA under the
control of the QuickBasic programs. The 313-region FE model
of figure 3.3 was used and the permittivity of the regions
was set to simulate test objects with relative permittivity
of 2.1 ( high ) in a background with relative permittivity of
1 ( low ).

For

each

type

of

excitation,

two

different

permittivity distributions were tried: one corresponding to


an object in the centre and another corresponding to an
object half-way between the centre and the wall, as shown in
figure 3.9. Therefore, four simulations were done. The
complete programs can be found in appendix B.
Unlike in the simulations to calculate the sensitivity maps,
this time the permittivity distribution remains the same
through the whole simulation, and the command input file
generated by the QuickBasic program at the start of each
iteration is now used to change the electrode potentials in
order to change the excitation electrode (single-electrode
excitation) or set the various projections (parallel-field
excitation). For single-electrode excitation there are 11
iterations, corresponding to electrodes 1 to 11 being used
for excitation and, on each iteration, the charge on the
52

detection electrodes (according to the protocol described in


section 2.2.1 of chapter 2) is computed by PC-OPERA, yielding
a total of 66 charge measurements.

a) Object between the centre


and the wall

Figure 3.9

b) Object in the centre

Simulated test objects

For parallel-field excitation there are 6 iterations,


corresponding to projections 1 to 6 and, on each iteration,
PC-OPERA calculates the charge on all 12 electrodes, giving
72 charge measurements.
From the simulated charge measurements, the reconstructed
images shown in figure 3.10 were produced using LBP. For
single-electrode excitation, the standard LBP algorithm was
employed (with high = 2.1 and low = 1), as described in section
53

2.2.2 (chapter 2). For parallel-field excitation, an adapted


version of the algorithm was used, given by

GPF (k)

12

i j Sij (k)

i 1 j 1
6
12

(k 1..313)

(3.6)

Sij (k)
i 1 j 1

where GPF (k) is the grey level corresponding to each of the


313 in-pipe regions in our model, i and j are the projection
and measuring electrode numbers, Si j are the sensitivity maps
given

by

equation

(3.5),

and

 i j are the normalised

electrode-charge measurements defined as

Qi j meas - Qi j emp

 i j= 

(3.7)

Qi j full - Qi j emp

where Qi j meas is the charge measured on electrode j in


projection i (with the test object present), while Qi j full and
Qi j emp are the charge of electrode j in projection i when the
sensor is full of high- and low-permittivity material,
respectively. This algorithm is based on 6 projections, taken
at 0, 30, 60, 90, 120 and 150, with measurements taken
from each of the 12 electrodes for every projection, thus
giving a total of 72 measurements.

54

a) Object between the centre


and the wall, true image

b) Object in the centre,


true image

c) Object between the centre


and the wall, S.E. excit.

d) Object in the centre,


S.E. excit.

e) Object between the centre


and the wall, P.F. excit.

f) Object in the centre


P.F. excit.

Figure 3.10 Comparison of


reconstructed images
using
single-electrode (S.E.) and parallel-field (P.F.) excitation
55

In figure 3.10 one can see that the reconstructed images for
single-electrode excitation, despite showing the spreading
effect inherent to LBP, at least are an approximate
representation of the test objects, with the peak values
occurring very close to the test objects positions.
On the other hand, the images for parallel-field excitation
appear almost totally wrong, with the objects sort of
pulled outwards to the periphery. From the observation of
the sensitivity maps of figure 3.8 one might expect this
effect, since the maps are almost flat and uniform in the
middle part and therefore the sensor is quite unable to
distinguish or resolve anything in that area (figure 3.10-f).
Sensitivity is, however, very high near the electrodes and
thus the sensor can give us some information on this area.
These results confirm that parallel-field excitation is not
a very good choice for tomographic applications. In the next
section we will approach the question from a different angle,
and will consider the issue of whether or not the mutual
capacitances can be obtained from the parallel-field charge
measurements (which would let us use the standard LBP
algorithm to reconstruct an image).

3.7

DETERMINATION OF MUTUAL CAPACITANCES FROM PARALLELFIELD CHARGE MEASUREMENTS

As mentioned in chapter 2, the value of the self and mutual


capacitances of an ECT sensor can be found by applying known
potentials to the electrodes and measuring the induced
electrode charges.
For example, for an n-electrode sensor, let us consider the
problem of determining the self- and mutual capacitances
associated with electrode i (i.e. Ci 1 to Ci n ). This can be
56

done by applying n independent (and preferably orthogonal)


excitation voltage vectors Vk of the form

Vk = [Vk 1 , Vk 2 , ..., Vk n ]

(k = 1, .., n)

(3.8)

For each excitation vector Vk we must measure the charge on


electrode i, Qi k . Then, taking the equation for Qi from
equation (2.1) in chapter 2, we can assemble a system of n
simultaneous linear equations as follows

Qi 1 = ci 1 V1 1 + ci 2 V1 2 + + ci n V1 n
Qi 2 = ci 1 V2 1 + ci 2 V2 2 + + ci n V2 n

Qi n = ci 1 Vn 1 + ci 2 Vn 2 + + ci n Vn n

(3.9)

or, in matrix form






V1 1
V2 1

Vn 1

V1 2 V1 n
V2 2 V2 n

Vn 2 Vn n











ci 1
ci 2

ci n


 =








Qi 1
Qi 2

Qi n






(3.10)

By solving this system of equations of the form A x = b we can


find all the self- and mutual capacitances of electrode i. Of
course, we can only solve the system if the excitation matrix
A (the coefficient matrix) is invertible (i.e. non-singular),
57

and this is by no means guaranteed for all sets of excitation


voltage vectors. It is also desirable that the matrix A be
well conditioned (with a condition number close to 1), and,
ideally, it should be made of an orthogonal set of voltage
vectors.
Let us now consider single-electrode excitation to begin
with. In this case, if we denote the excitation voltage by V,
the excitation matrix A takes the form






= 







V
0
0
0
0
0
0
0
0
0
0
0

0
V
0
0
0
0
0
0
0
0
0
0

0
0
V
0
0
0
0
0
0
0
0
0

0
0
0
V
0
0
0
0
0
0
0
0

0
0
0
0
V
0
0
0
0
0
0
0

0
0
0
0
0
V
0
0
0
0
0
0

0
0
0
0
0
0
V
0
0
0
0
0

0
0
0
0
0
0
0
V
0
0
0
0

0
0
0
0
0
0
0
0
V
0
0
0

0
0
0
0
0
0
0
0
0
V
0
0

0
0
0
0
0
0
0
0
0
0
V
0

0
0
0
0
0
0
0
0
0
0
0
V














(3.11)

This is the simplest set of orthogonal vectors and it results


in an excitation matrix with a condition number of 1. This
particular excitation matrix is very convenient because the
solution of equations (3.9) and (3.10) then becomes a trivial
matter, with each capacitance given directly by one of the
electrode charge measurements.
Let us turn our attention now to parallel-field excitation.
In this case, the basic excitation voltage vector is given by
equation (3.3), which, for = 0 and E = 15 V, results in the
values of table 3.1. Up to 11 additional vectors can be
produced by shifting the basic vector components. Each one of

58

this 12 vectors defines a projection with a particular


field direction. This voltage vectors result in the following
excitation matrix A

4.02

10.98

15.00

15.00

10.98

4.02

-4.02

4.02

10.98

15.00

15.00

10.98

4.02

-10.98

-4.02

4.02

10.98

15.00

15.00

10.98

4.02

-15.00 -10.98

-4.02

4.02

10.98

15.00

15.00

10.98

4.02

-15.00 -15.00 -10.98

-4.02

4.02

10.98

15.00

15.00

10.98

4.02

-10.98 -15.00 -15.00 -10.98

-4.02

4.02

10.98

15.00

15.00

10.98

4.02

-4.02

-4.02 -10.98 -15.00 -15.00 -10.98

-4.02

4.02

10.98

15.00

15.00

10.98

4.02

-4.02 -10.98 -15.00 -15.00 -10.98

-4.02

4.02

10.98

15.00

15.00

10.98

-4.02 -10.98 -15.00 -15.00 -10.98

-4.02

4.02

10.98

15.00

15.00

-4.02 -10.98 -15.00 -15.00 -10.98

-4.02

4.02

10.98

15.00

-4.02 -10.98 -15.00 -15.00 -10.98

-4.02

4.02

10.98

-4.02 -10.98 -15.00 -15.00 -10.98

-4.02

4.02

4.02
10.98

4.02

15.00

10.98

4.02

15.00

15.00

10.98

4.02

10.98

15.00

15.00

10.98

4.02

-4.02 -10.98 -15.00 -15.00 -10.98

-4.04

-4.02 -10.98 -15.00 -15.00 -10.98


-4.02 -10.98 -15.00 -15.00
-4.02 -10.98 -15.00
-4.02 -10.98
(3.11)

It turns out that this type of matrix is non-invertible (i.e.


singular) and, therefore, can not be used to solve equations
(3.9) and (3.10) in order to find the self- and mutual
capacitances of a particular electrode. Moreover, the rank of
this excitation matrix was found equal to just 2, which means
that only two of the vectors can be considered independent.
This is not surprising, since a parallel field in any
direction can be specified in terms of only two orthogonal
components.

3.8

CONCLUSIONS

The main conclusions of this chapter can be summarised as


follows

59

a) Using FE simulation techniques, we found that, with regard


to the problems of non-uniformity of the sensitivity maps
and lack of sensitivity in the centre the sensor, the
situation is in fact worse with parallel-field excitation
than it is with single-electrode excitation.
b) It was also found that, as a result of (a), the LBP images
produced using parallel-field excitation are clearly of an
inferior quality compared to those obtained using singleelectrode excitation.
c) Finally, we showed that a complete set of independent
measurements required to determine the sensor mutual
capacitances can not be obtained using only parallel-field
excitation. A similar conclusion is expressed with respect
to electrical resistance tomography (ERT) in [43] and
[44].
In short, parallel-field excitation results in rather flat
sensitivity maps, making it more difficult to determine the
spatial position of features or anomalies inside the sensor,
and thus making this type of excitation a poor choice for
tomographic imaging applications.
The flatness of the sensitivity maps suggests that
parallel-field excitation could perhaps be exploited
advantageously in the development of distribution-independent
concentration sensors, a field of great industrial relevance.
This, however, falls outside the scope of this thesis, and is
left as an area of further work.
The fact is that there is a fundamental difference between
parallel-field excitation ECT and parallel-beam X-ray
tomography. In the latter, at any one time the response of
each detector is affected only by one source, whereas in the
former, despite the parallel-field, the charge on the
60

detection electrode is influenced by the voltages applied to


all the other (source) electrodes, as is evident from
equation (2.1) given in chapter 2. This is the main reason
why parallel-field excitation ECT cannot be considered
analogous to parallel-beam X-ray tomography.

61

C H A P T E R

DESIGN

4.1

OF

AN

4 :

AC-BASED

MULTIPLE

EXCITATION

ECT

SYSTEM

INTRODUCTION

This chapter reports the design of an AC-based multipleelectrode excitation ECT system which was built in order to
a) carry out an experimental evaluation of this excitation
method,
b) be able to compare in practice the effect of different
sets of multi-electrode excitation vectors, and
c) evaluate the performance of the AC-based capacitance and
charge transducer presented in chapter 2.
A number of AC-based ECT transducer designs using an
operational amplifier (op-amp) have been reported [37,38,40],
similar to the basic detector circuit that we intend to use,
shown in figure 2.6 (chapter 2).
In [37] and [38], Yang et al proposed a circuit employing
resistive feedback. This has the advantage that the
capacitance measurement sensitivity becomes proportional to
the frequency and, therefore, the former can be raised by
increasing the latter. However, this configuration resembles
a differentiator, resulting in amplification of the highfrequency components of amplifier noise, and also can have
instability problems [65,66]. To achieve stray-immunity, this
62

circuit relies on the fact that the excitation electrode is


connected to a low-impedance voltage source, while the
detection electrode is connected to what can be considered a
virtual earth point, namely the inverting input of an op-amp
having a very high open-loop gain. However, the effect of the
reduction in op-amp gain with frequency is not investigated
or considered, nor is the effect of the switches used to
connect the electrodes to the either the op-amp input or the
excitation voltage source.
In [40], Pickup et al, of the Manchester Metropolitan
University, describe a similar ECT transducer circuit.
Although this time the frequency dependent op-amp gain is
taken into account, the effect of the switches is still being
ignored. The approach followed by Pickup et al is based on
operating the circuit at a relatively narrow peak in its
frequency response in order to achieve a high capacitance
sensitivity. This, however, makes the circuit quite sensitive
to changes in the parameters due to drift, etcetera (i.e.
stability is not very good). Moreover, the practical
application of this particular circuit is rather limited
because it is not stray-immune.
We intend to design an op-amp-based AC ECT transducer that
will address the weak points of the circuits described above.
As we shall see later on in section 4.4, it will be a strayimmune low-noise circuit and its frequency response will
exhibit a well defined pass-band. We shall thoroughly
investigate the effects of both non-ideal op-amp gain and the
analogue switches used to connect the electrodes.
We shall begin by setting the overall design objectives at
the system level and then follow a top-down design strategy,
moving from the general to the particular.

63

4.2

DESIRED SYSTEM CHARACTERISTICS

Generally speaking, we want an ECT system that, in singleelectrode excitation mode, can run as fast as the system
developed earlier at UMIST [9,11], while providing a higher
capacitance sensitivity (i.e. a lower noise level) through
the use of AC-based measurement techniques, plus offering the
capability to apply arbitrary multi-electrode excitation
patterns to the sensor. To be more precise, the principal
characteristics of the system are listed below (see also
section 2.3 of chapter 2):
a) Measurement range:
Inter-electrode capacitance (single-electrode excitation):
0 - 1,000 fF
AC electric charge (multiple-electrode excitation):
0 - 10,000 fCoulpeak
b) Full

standing-value compensation:

It should always be possible to zero the output, in


order to measure small changes in the input variable.
c) Resolution:
Capacitance: 0.1 fF.
Electrode charge: 1 fCoul.
d) Intrinsic noise level:
The target is a one order of magnitude improvement over
the previous UMIST system figure of 0.26 fF [9], i.e.,
H 0.026 fF (peak value of input-referred noise level).
e) Stray-immunity:
The measurements must be insensitive to stray capacitances
to ground up to H 150 pF, resulting from the cables
64

connecting the sensor to the electronics. The degree of


insensitivity should be at least the same as for the
previous UMIST system based on the charge-discharge
transducer,
i.e.,
the
ratio
of
stray-capacitance
sensitivity to measured-capacitance sensitivity should be
less that 0.0002 [34,35].
f) Speed:
The system must be capable of capturing all the
measurement data for one image frame in 10 ms (i.e., as
fast as the previous UMIST system [9]).
g) Fully software-controlled circuit setting.
h) Number of electrodes: 12 (expandable to 16).
i) Modular construction for easy expansion.
j) Frequency of operation: 500 kHz.
The system should operate at a relatively high frequency
(within the limitations
imposed by the
hardware
components) in order to achieve a high measurement speed,
and to minimise the effect of parallel loss conductance in
the measured capacitance.

4.3

GENERAL DESIGN STRATEGY

The block diagram of the proposed ECT system is shown in


figure 4.1. The chosen design approach was to have each
sensor electrode connected to an independent ECT transducer
channel, capable of working either as a programmable AC
voltage source (excitation mode) or as a charge detector
(detection mode). Within each channel, the detection section
65

is based on the charge amplifier circuit presented in section


2.3.2
of chapter 2, and every channel includes its own
demodulator in order to allow parallel measurement.
A direct-digital-synthesis (DDS) signal generator is used to
generate stable low-distortion 500 kHz excitation and
demodulation reference sine-waves with adjustable phase
shift.
The demodulated detection signal from any channel can be
selected via a multiplexer and sent to a common signal
conditioning circuit, where a standing-value compensation
voltage, produced by a digital-to-analogue converter (DAC),
is subtracted from it. The compensated signal (which is now
a measure of Q) is then sent to a programmable-gain
amplifier (PGA), where it is scaled in order to fully exploit
the dynamic range of the analogue-to-digital converter (ADC).

Figure 4.1

AC-based multi-electrode excitation ECT system

66

Finally, the measurements in digital form are sent to the


host computer (a conventional PC), where the image
reconstruction process takes place. The computer, in turn,
provides the digital signals required to control the
operation of the system (including converters, multiplexer,
PGA, transducer settings, etc.).
Communication between the data acquisition system and the
computer is done using a parallel input-output plug-in card
(Amplicon PC14-AT).

4.4

THE BASIC DETECTOR CIRCUIT

Our detector circuit is based on the charge amplifier


described earlier in chapter 2, shown again in figure 4.2.

Figure 4.2

Basic detector circuit

67

The circuit measures the charge Q induced on the detection


electrode as Vo = Q /Cf . Or, alternatively, if there is only
one excitation electrode (i.e. single-electrode excitation),
as shown in the figure inside the dotted line, then the
circuit measures the inter-electrode capacitance Cx as
Vo = - (Vexc /Cf )Cx .
The previous equations for Vo were derived for

(1 /Rf Cf ) and

assuming an ideal operational amplifier (op-amp), having


constant infinite gain, infinite input impedance, and zero
output impedance. A real op-amp, however, will depart from
these ideal characteristics. In particular, although the gain
is normally very high for DC signals, it shows a marked
frequency dependence, becoming smaller at higher frequencies.
We
shall,
therefore,
examine
the
effect
that
the
frequency-dependent gain has on our detector circuit.
4.4.1 TRANSFER FUNCTION WITH FREQUENCY-DEPENDENT OP-AMP GAIN
We shall analyze the circuit as a capacitance meter, i.e.,
including the excitation voltage source. Assuming that the
output impedance of the latter is low, the excitation
electrode stray capacitance to ground Cs will have no effect
and can be ignored (this will be discussed further in section
4.5.2). There are many different ways of analyzing feedback
circuits. Our analysis will be based on the fact that the
closed-loop gain A C of an op-amp feedback circuit can be
always be put in the form [46,47]

A C (s) =

Vo (s)
Vexc (s)

A I (s)

1
1 +
A(s) (s)

68

(4.1)

where s is the Laplace variable, A I (s) is the closed-loop


gain that would be obtained with an ideal op-amp, A(s) is the
real op-amp open-loop gain, and (s) is the feedback factor
(the fraction of the output voltage that is fed back to the
input of the op-amp).
First we find A I (s). For an ideal op-amp the voltage across
Cs will be zero, so Cs can be ignored, leaving a classical
inverting amplifier with Zf = (1 /Rf + s Cf )- 1 and Zin = 1 /s Cx .
Thus we have

A I (s) = -

Zf

s Rf Cf

= -

(4.2)

1 + s Rf Cf

Zin

Now, the typical op-amp can be modelled as a single-pole


amplifier, with open-loop gain given by

A(s) =

Vo
(V+ - V_ )

Ao
1 + (s / c )

(4.3)
1

s
+

Ao

Ao

c is the -3 dB cut-off
frequency. Recalling that u = Ao c (the op-amps unity-gain
bandwidth) and considering that Ao 1 (typically e 10,000),
where Ao is the DC open-loop gain and

equation (4.3) can be simplified to

69

A(s) =

u /s

(4.4)

The feedback factor is found by inspection of figure 4.2.


If we let Vexc = 0 to eliminate its contribution and assuming
that the amplifier has a high input impedance, we can see
that the fraction of Vo that appears at the op-amp input is
given by the voltage divider formed by Zf and CxQ Cs . Thus

1
(s) =

s (Cx + Cs )
1
Zf +

1 + s Rf Cf

(4.5)

1 + s Rf (Cx + Cf + Cs )

s (Cx + Cs )

Substituting equations (4.2), (4.4) and (4.5) back into


(4.1), after some algebraic manipulations we come to the
transfer function of the basic detector circuit, given by

G(s) = A C (s) =

Cx
Cf

s Rf Cf
s2

Rf (Cx + Cf + Cs )

(4.6)
1

+ s R f Cf +

70

+ 1

Equation (4.6) can be put in the form

G(s) = -

Cx

s Rf Cf

Cf

where

(4.7)

s
1 +

s
1 +

1 and 2 are the pole-frequencies, given by the roots

of the denominator of equation (4.6) as

1 = -s1 =

Rf Cf +

2
1
Rf (Cx + Cf + Cs )
#
- # Rf Cf +
+ 4
u
u

Rf (Cx + Cf + Cs )

(4.8)

and

2 = -s2 =

Rf Cf +

2
1
Rf (Cx + Cf + Cs )
#
+ # Rf Cf +
+ 4
u
u

Rf (Cx + Cf + Cs )

71

(4.9)

For

2 1

u (1 /Rf Cf ),

and

equations

(4.8)

and

(4.9)

approximate to

1 =

(4.10)
Rf Cf

and

2 =

u Cf

(4.11)

Cx + Cf + Cs

and if, furthermore, Cx (Cf + Cs ), equation (4.11) approximates


to

2 =

u Cf

(4.12)

Cf + Cs

To obtain the circuits transfer function as a charge


measurement device we divide both sides of equation (4.7) by
-Cx to get

Vo (s)
Q(s)

s Rf Cf

=
Cf

(4.13)

s
1 +

s
1 +

72

Figure 4.3 shows the magnitude Bode plot corresponding to


equations (4.7), (4.10) and (4.12).

Figure 4.3

4.4.2

Bode plot of capacitance transducer response

COMMENTS ON NOISE PERFORMANCE

The noise performance of the basic detector is limited by the


characteristics of the op-amp. As will be seen later (section
4.5), through the use of phase-sensitive demodulation the
circuits bandwidth will effectively be restricted to a
narrow band fm f/2, where fm is the frequency of operation.
Assuming Cx small, at these frequencies the noise gain of the
circuit is approximated by Gn = 1 + (Cs /Cf ).

73

The total output noise over the bandwidth fm f/2 will be

Vn = Gn

# [en2 + 4 K T R f + (in /m Cp )2 ] f

(4.14)

where en and in are the op-amps equivalent input noise

voltage and current, K is Boltzmanns constant (1.38 10-23

Joules /K), T is the temperature in K,


Cp = CsQ Cf .

m = 2 fm , and

Observing the form of equation (4.14), some comments can be


made about the noise performance of the circuit. Firstly, we
see that the noise gain Gn increases with the stray
capacitance, and will generally be much larger that the
signal gain ( = - Cx /Cf ), thus the need to use low-noise
op-amps (see next section). Secondly, since the total noise
voltage depends on f, bandwidth-narrowing techniques, like
the use of phase-sensitive detection, can greatly reduce the
noise level.
4.4.3

DETAILED CIRCUIT DESIGN

The charge detector design involves determining the right


values for u , Cf and Rf. The amplifier of choice in our
design is Analog Devices high-speed low-noise op-amp AD829.
It was decided to use this op-amp because it offers not only
a high bandwidth, with u = 2 (750 MHz) = 4.71 109 rad/s
(typical), but also a very low equivalent input noise voltage
of just 1.7 nV/ Hz.
From

equations

(4.7)

and

(4.13)

we

can

see

that

the

sensitivity or gain of the circuit is inversely proportional


74

to the feedback capacitance Cf , which must, therefore, be


chosen as small as practically possible, while still being
large compared with the stray capacitances associated with
the physical construction of the circuit (due to printed
circuit tracks, component leads, etc.), so that these will
not have a significant effect. For our work, we used
Cf = 22 pF
and
employed
high-stability
silvered
mica
capacitors. Considering a typical value for Cs of about

u = 4.71 109 rad/s, using equation (4.12) gives


an upper -3 dB frequency 2 = 850 Mrad/s or f2 = 135 MHz, which
100 pF and

quite safely exceeds


measurement frequency.

the

500 kHz

specified

for

the

Rf , on the other hand, should be chosen high enough to ensure


that the lower -3 dB frequency lies below the excitation
frequency, but not as high as to cause an excessive DC offset
at the op-amp output due to the input bias current. We chose
a value of 220 k, which, using the AD829, causes a DC offset
of about 2 V at the output of the circuit. This, however,
should not be a problem with regard to op-amp saturation
because the AC output component is expected to be quite small
anyway. Using equation (4.10), this value of Rf combined with
Cf = 22 pF

results

in

lower

-3 dB

frequency

1 = 207 10 rad/s or f1 = 33 kHz, which is more than a decade


3

below the excitation frequency of 500 kHz.


The

selected

value

for

results

Cf

in

basic

charge

sensitivity SQ = 1/Cf = 1/(22 pF) = 0.0455 mV/fCoul. Moreover,


considering

Vexc = 10 Vpeak ,

we

have

basic

capacitance

sensitivity SC = - (Vexc /Cf ) = - (10 Vpeak /22 pF) = - 0.455 mVpeak /fF.
Clearly, these sensitivity values are rather low for the
expected input levels, and there is a need for further
amplification.

75

4.4.4

FINAL FREQUENCY RESPONSE

The design of the detector circuit was done using the


approximate formulas in equations (4.10) and (4.12). Figure
4.4 shows the magnitude of the circuits frequency response,
calculated employing the full transfer function of equation
(4.6) with s = j , using the design values of u , Rf and Cf ,
and assuming Cx = 100 fF and Cs = 100 pF, in order to see how
close it comes to the approximate design.

Figure 4.4

Detector circuit frequency response

Figure 4.4 shows a mid-band gain of -47 dB, with f1 = 31.7 kHz
and

f2 = 130 MHz,

quite

close

to

the

values

obtained

previously in section 4.4.3 using the simplified equations


(4.10) and (4.12).
76

4.4.5

EFFECT OF CHANGES IN STRAY CAPACITANCE

By looking at the transfer function of the circuit, in


particular equation (4.12), it is clear that any variation in
the stray capacitance Cs will have the effect of shifting the
upper 3-dB frequency up or down. Provided that the frequency
of operation is sufficiently far away from the upper 3-dB
point, any shifting on the latter caused by stray capacitance
variations of say 100%, is expected to have very little
effect on the circuit response.
In what follows, we will determine how insensitive the
circuit actually is to changes in Cs. Since, as we shall see
later on in section 4.5, the system will be using phasesensitive detection (PSD), we must take into account the
effect of Cs on both the magnitude and phase of the op-amp
output, Vo (j). Using equation (4.7) with s = j and recalling
that G(s) = Vo (s)/Vexc , we can obtain the following expressions
for the magnitude

Vo  and the phase angle o the op-amp

output voltage:

 Vo 

V exc C x
Cf

 Rf Cf
1 (/ 1)2

1 (/ 2)2

o = (3 /2) - tan-1( /1 ) - tan-1( /2 )

where

(4.15)

(4.16)

1 and 2 are the pole-frequencies, given by equations

(4.10) and (4.12).


77

This signal has to go through additional amplification stages


before being fed to the PSD. Therefore, the magnitude and
phase of the signal at the PSD input are

Vin PSD  = K1 Vo 

(4.17)

and

in PSD = o + 1

where K1 and

(4.18)

1 are the gain and phase shift due to the

additional amplification stages.


The output of the PSD is then given by

Vo PSD =

K2 K1 Vo  cos[r - (o +

where K2 is a constant and

1 )]

(4.19)

r is the phase of the demodulation

reference signal [45].


If we adjust the phase of the demodulation reference signal
so that r = + 1 , where

 = o

(4.20)
Cs = Co

78

is the phase of the signal at the output of the op-amp in the


basic detector circuit when the stray capacitance Cs is equal
to its nominal value Co , then the final output after the
PSD becomes

Vo PSD =

K2 K1 Vo  cos(-

o )

(4.21)

We now define the relative stray-capacitance sensitivity as

R =

Scs
Scx

 Vo PSD

 Cs

 Cs

 Vo PSD

 Cx


 Cx

[Vo  cos(-

o )]
(4.22)

[Vo  cos(-

o )]

Using the designed circuit parameters given in section 4.4.3,


R was estimated numerically using finite differences for a
range of Cs from 80 pF to 200 pF, considering a frequency of
operation of 500 kHz and a nominal stray capacitance Co of
100 pF. As it turns out that Scs is directly proportional to
Cx , the worst case was assumed and the calculations were done
using Cx = 1 pF (the maximum consistent with the proposed
measurement range of the system). The results, shown in
figure 4.5, indicate that changes in stray capacitance Cs
within the range specified above have a negligible effect on
the system output compared with that of variations in the
measured capacitance Cx . When Cs is equal to Co (= 100 pF), we
have

R = 1.12 10-7.


79

Figure 4.5

4.5

Relative stray-capacitance sensitivity

DESIGN OF THE MULTI-EXCITATION ECT TRANSDUCER CHANNEL

Based on the based detector circuit already discussed, we


shall now embark on the design of one of the transducer
channel blocks shown in figure 4.1. Some of the desired
features of this transducer channel are presented next.
a) It should have two alternative operating modes, excitation
and detection, so that with one of this measuring circuits
connected to each electrode in an ECT sensor, any
particular electrode can be used either for excitation or
detection.
b) In the excitation mode, it should provide an adjustablelevel high-frequency (500 kHz) AC signal with low output
impedance and constant phase shift. The latter requirement
arises because when several electrodes are used for
excitation all the applied signals must be in-phase.
80

c) In the detection mode, it should have high-enough gain and


low-enough noise to amplify the very low signals induced
on the detection electrodes (H 0 - 10,000 fCoul). Its
output must be a high-level signal.
d) It should employ phase-sensitive detection (PSD) in the
demodulator stage, in order to have discrimination between
the effects of the capacitive and conductive components of
the measured impedance, as well as to achieve high noise
rejection. Although the use of PSD brings about the
potential ability to measure the conductive component as
well by applying a 90 phase shift to the demodulation
reference signal, in this work we limit our attention to
the measurement of the reactive component only.
e) It should have several measurement ranges in order to be
versatile enough to cope with the wide variation in
capacitance (or charge) encountered in ECT. This can be
realised by having variable AC-gain control.
h) Finally, the operating mode, excitation level and AC-gain
should be digitally controlled, so that many of this
transducer circuits can be integrated into a multiexcitation ECT system.
Figure 4.6 shows the block diagram of the complete transducer
channel. One of these transducers is needed for each
electrode. It contains both excitation and detection
circuitry, either of which can be connected to the electrode.
At least two of these transducers are needed to make a
measurement. To measure capacitance between a pair of
electrodes, the excitation section of one transducer circuit
is connected to one electrode while the detection section of
the other transducer is connected to the other electrode. If
we want to measure electrode charge instead, then more than

81

one transducer can be used to apply excitation signals to


several electrodes.

Figure 4.6

AC-based multi-excitation ECT transducer

Referring to figure 4.6, the mode of operation is set by


means of an array of CMOS analogue switches (DG413). Three of
them form the excitation switch, arranged in a "T"
configuration in order to improve off-isolation and prevent
any unwanted signal-coupling within the same channel.
Detection is controlled by a single SPST switch.
The excitation level can be adjusted by passing the 10 Vpeak
500 kHz excitation reference signal through a four-quadrant
analogue multiplier (AD734), whose other input is an
adjustable DC voltage coming from a bipolar-output 12-bit

82

digital to analogue converter (DAC813). This arrangement


allows the peak voltage of the excitation signal to be swung
from -10 V to 10 V with great precision and no change in
phase shift.
In the detection mode, after the Q - V converter the signal
goes to a gain-of-20 amplifier, followed by an AC
programmable-gain amplifier (PGA) based on four cascaded
gain-of-5 stages, with gain settings GAC-PGA of 1, 5, 25 and
125. The AC-PGA is required because of the large variation in
ECT sensor capacitances. Low-noise op-amps (AD829) were used
in these amplification stages.
Finally, the signal from the AC-PGA goes to a phase-sensitive
(or synchronous) demodulator (PSD), which yields a DC voltage
proportional to the measured capacitance (or charge). The PSD
can be thought of as a very stable narrow-band filter tuned
to fref (the frequency of the reference and excitation
sine-waves) and having extremely high

Q [45]. It provides

a high degree of noise rejection and also makes the


transducer insensitive to any conductive component that may
exist in parallel with the capacitance being measured. Our
PSD was built using an analogue multiplier (AD734) followed
by a low-pass filter (LPF), as shown in figure 4.7, with its
output given by Vo (PSD) = (Vi (PSD) / 2 ) GLPF , where Vi (PDS) is its
input signal and GLPF is the pass-band gain of the LPF. An
active 10 kHz second-order Butterworth LPF was used
(GLPF = 1.58), thus giving Vo (PSD) = 0.79Vi (PDS) . The PSD will
only accept signals within the frequency band fref - fLPF to
fref + fLPF , which means, for fref = 500 kHz, 490 to 510 kHz.
Lowering the LPF cut-off frequency reduces the PSD input
acceptance, thus improving the signal to noise ratio (SNR),
but also increases the output settling time, slowing down the
system. The LPF cut-off frequency of 10 kHz was chosen as an
adequate compromise between speed and SNR. The specifications
of the multiplier used in our design limit the amplitude of
83

the PSD input voltage to 10 V, therefore the useful output


voltage range goes from 0 to 0.79 10 = 7.9 V.

Figure 4.7

Phase-sensitive detector

The complete circuit diagram of the ECT transducer, including


the necessary interface circuits for address decoding and
command data storage (not shown in figure 4.6 for simplicity)
can be found in appendix A.
4.5.1

NOMINAL MEASUREMENT RANGES AND SENSITIVITIES

The theoretical total charge sensitivity of the whole


transducer
is
SQ = SQ (bas) 20 GAC-PGA 0.79,
where
SQ (bas) = 0.0455 mV/fCoul

is

the

sensitivity

of

the

basic

detector circuit and GAC-PGA is the gain of the AC-PGA. The


capacitance sensitivity is given by SC = Vexc SQ = 10 SQ .
Considering a full-scale output voltage of Vo (max) = 7.9 V, as
mentioned before, the nominal measurement ranges for each
gain setting of the AC-PGA are found to be as shown in
table 4.1.
84

Table 4.1


 G

AC-PGA

1


5


 25

 125


So,

,



<

<

<

<

4

for

ECT transducer nominal measurement ranges and


sensitivities

,

 S E N S I T I V I T Y 
<
,
$
SQ (mV/fCoul) SC (mV/fF) 
<
<
$
0.72
7.2



<
<
$
3.6
36



<
<
$
18
180



<
<
$
90
900



4
4


R A N G E
Q (fCoul)
11,000
2,200
440
88

,

<

<

<

<

4

GAC-PGA = 125

C (fF)
1,100
220
44
8.8

(the

most

sensitive

setting),

for

example, we now have a charge sensitivity SQ = 90 mV/fCoul,


and

(for

Vexc = 10 Vpeak )

capacitance

sensitivity

SC = 900 mV/fF, quite acceptable figures, even for the lower


end of the expected range of input charge or capacitance
levels.
4.5.2

THE EFFECT OF CMOS SWITCH

ON RESISTANCE

We shall now discuss the effect that the non-zero on


resistance of the actual CMOS analogue switches has on the
circuit performance. If we take into account the switch
resistance, we have the basic detector equivalent circuit
shown in figure 4.8, where Vexc is the excitation voltage
source, V is the actual voltage on the source electrode, R
is the on resistance CMOS analogue switches, Cs is the
stray capacitance, due mainly to the screened cables
connecting the sensor, and Cx is the capacitance between the
source and detection electrodes. We shall assume an ideal
op-amp and disregard the feed-back resistor in parallel with
Cf (which is used in practice to avoid op-amp saturation and
has the effect of limiting the low-frequency response). The
85

analogue switches employed (DG413) have an on resistance


specification of 35 (typical). Cx will typically be less
than 1 pF, and Cs about 100 pF.

Figure 4.8

Effect of switch resistance

First we shall determine V. Since Vexc actually comes from the
output of a high-bandwidth op-amp in a voltage follower
configuration, its output resistance can be neglected
compared with 2R. For an ideal op-amp, point a will be a
virtual earth and, therefore, the resistance R of the switch
connected to the detection electrode will be effectively in
parallel with the stray capacitance Cs . Under these
conditions, for the typical values quoted above, the
impedance Z seen by output of the excitation switch will be
dominated by the source electrode stray capacitance (Cs ). The
equivalent circuit of figure 4.9 can, therefore, be used to
calculate the voltage V on the source electrode.
86

Figure 4.9 Equivalent circuit used to determine the


voltage V on the source electrode

The circuit of figure 4.9 is a simple first-order RC lowpass filter. With the values given above, the cut-off
frequency fc will be 22.3 MHz. This is sufficiently higher
than the excitation frequency (500 kHz) to guarantee that
reasonably large changes in Cs will have a very small effect
on V. For example, using the figures quoted earlier, it was
calculated that a change in Cs from 80 pF to 150 pF will
cause a change in the amplitude and phase of V of 0.039% and
0.88 respectively. The magnitude and phase of V are given
by

V 

Vexc
1 (2  R Cs )

87

(4.23)
2

and

V = - tan-1 (2 R Cs )

(4.24)

If Cx << Cs , the input current to the op-amp, Iin , is given by


the following expression

Iin (j) =

j VCx
1 + j R Cs

(4.25)

The op-amp output voltage is Vo = - Zf Iin , where Zf is the


impedance provided by the feed-back capacitor Cf . Therefore,
we have

Vo = -

Iin
j Cf

= V

- Cx /Cf
1 + j R Cs

(4.26)

The magnitude and phase of Vo are given by

 Vo(j) 

 V (j) C x /Cf
1 ( R Cs )2

88

(4.27)

and

o () =  + V () - tan-1 ( R Cs )

Substituting equations (4.23) and (4.24 for

(4.28)

V and V we

come to

 Vo 

Vexc Cx /Cf
1 ( 2  RC s )2

1 ( R Cs )2

(4.29)

and

o () =  - tan-1 (2 R Cs ) - tan-1 ( R Cs )

(4.30)

The presence of the non-zero switch resistance means that,


even considering an ideal op-amp, the effect of the stray
capacitance on the detection electrode can not be immediately
neglected, since the voltage across it will not be zero.
Having found expressions for Vo and o which include the

on resistance, we can then use


equation (4.22) to see what the effects are on the circuits
effects of CMOS switch

relative sensitivity to stray capacitance variations, R. For


the typical values of Cs (H 100 p), the break frequencies in
equations (4.29) and (4.30) will be much higher that the
excitation frequency, and we would not expect the circuit to
89

be affected significantly by changes in Cs due to mechanical


vibration, varying cable length, etcetera.
Using equations (4.29) and (4.30) for

Vo and o , and

employing the designed circuit parameters given in section


4.4.3, the relative stray-capacitance sensitivity R (given by
equation 4.22) was estimated numerically using finite
differences for a range of Cs from 80 p to 200 p, considering
a frequency of operation of 500 kHz and a

nominal stray

capacitance Co of 100 p. As it turns out that Scs is directly


proportional to Cx , the worst case was assumed and the
calculations were done using Cx = 1 p (the maximum consistent
with the proposed measurement range of the system). The
results, shown in figure 4.10, indicate that changes in stray
capacitance Cs within the specified range have a negligible
effect on the system output compared with that of variations
in the measured capacitance Cx .

Figure 4.10

Relative stray-capacitance sensitivity

90

When Cs is equal to Co (= 100 p), we have

R = 6.06 10-6.

This is the amount of stray-capacitance sensitivity due


exclusively to the non-zero on resistance of the CMOS
switches, and it is interesting to see that it is about one
order of magnitude larger than the figure obtained in section
4.4.5 for the amount of stray-capacitance sensitivity caused
solely by the non-ideal characteristics of the op-amp used in
the basic detector circuit. This suggests that the dominant
source of stray-capacitance sensitivity are the CMOS
switches. In any case, however, the sensitivity of the
AC-based transducer to stray-capacitance variations is indeed
extremely small, being more than one order of magnitude lower
than that of the charge/discharge transducer [34,35].

4.6

REFERENCE SINE-WAVE GENERATOR

The ECT transducer presented in the previous section needs


stable low-distortion 500 kHz excitation and demodulation
reference sine-waves, whose phase shift has to be adjusted so
that the whole transducer circuit responds only to
capacitance, and not to the orthogonal signal contribution
resulting from any parallel conductive component.
For this purpose, a dual-output signal generator was designed
using two AD7008 direct-digital-synthesis (DDS) chips,
recently introduced by Analog Devices [48]. This device
allows accurate software control of signal phase-shift with
0.09 resolution.
The block diagram of the reference-signal generator is shown
in figure 4.11. Each AD7008 chip produces an approximation to
a sine-wave made of a large number of "steps." The number of
steps per cycle depends on the desired output frequency fo
and is equal to fCLK /fo, where fCLK is the clock frequency. In
91

our case we have fCLK = 50MHz and fo = 500kHz, therefore each


cycle is made of 100 steps. The chips allow digital control
of frequency, amplitude and phase with a very high
resolution, although in our application we are only
interested in phase control.

Figure 4.11

Reference-signal generator

The raw DDS outputs are passed through a low-pass filter to


remove the higher order harmonics associated with the
"steps", leaving two low-distortion 10Vpeak sine-waves which
are then buffered before being sent to the transducers.
On power-up, the frequency data must be loaded into each DDS
using a 32-bit word F according to fo = F fCLK / 232. This
gives a frequency resolution of about 0.01Hz. The phase of
92

any of the two sine waves can be adjusted within the range
0 - 360 using a 12-bit word, thus resulting in a phase
resolution of 360/4096

H 0.09, which allows for a very fine

control.
The complete circuit diagram of the signal generator is given
in appendix A.

4.7

SIGNAL CONDITIONING AND DATA CONVERSION SECTION

Figure 4.12 shows the block diagram of the signal


conditioning and data conversion section of our ECT system.
The output from each of the transducer circuits is sent in
pseudo-differential form (i.e., although the transducer
outputs are single-ended, individual signal return wires are
used) to a
couple of 8-to-1 differential multiplexers
(DG527) configured to work together as a 16-to-1. The output
of the multiplexer is transformed to single-ended form by an
instrumentation amplifier (INA114). The idea of this
pseudo-differential arrangement is to exploit the high
common-mode rejection ratio of the INA114 to eliminate any
common-mode noise or interference picked up on the wires
carrying the measurement signals from the various transducer
channels. Twisted-pair wires were used for this purpose. As
mentioned in section 4.5, the transducer circuits have a
full-scale output of 7.9 V, therefore, the gain of the INA114
was set to 1.26, in order to achieve a more convenient
full-scale value of 10 V.
The standing-value compensation voltage is generated by a
16-bit digital-to-analogue converter (DAC712) and is
subtracted from the measurement signal at the input of a
differential binary-step PGA (gain settings 1 to 64),
implemented using two programmable-gain instrumentation
93

94

amplifiers (PGA206). Using the programmable gain, the


compensated signal (representing the change in charge or
capacitance) is scaled-adjusted and fed to a 12-bit 3 s
sampling analogue-to-digital converter (ADS7800).
The digital data go through a set of 74HC365 buffers (not
shown in the figure) before, finally, being sent in parallel
to the plug-in digital interface card (Amplicon PC14-AT) in
the image reconstruction computer.

4.8

SYSTEM INTEGRATION AND CONSTRUCTION

Following the previously presented design, a complete ECT


data acquisition system was built using printed circuit
boards. It includes:

12 transducer modules.
1 DDS signal generator module with dual 500 kHz sine-wave
output.

1 signal conditioning and data conversion module including


multiplexer, standing value compensation circuitry, PGA,
ADC and output data buffers.

1 parallel input module containing buffers for the control


signals sent by the image reconstruction computer via the
plug-in digital interface card (Amplicon PC14-AT).
All the prototype cards conform to the Eurocard format and
were mounted on an IEC 297-3 compatible sub-rack. Special
attention was given to low-noise high-frequency design and
construction techniques. Low-noise (1.7 nV/ Hz) high-speed
op-amps (AD829) were employed in the detection sections, a
massive ground plane was used on the component side of the
PCBs, and decoupling capacitors were extensively used.
95

Separate wires were used to carry the power supply and ground
to each module. To avoid unwanted coupling, screened cable
was used to distribute the high frequency excitation and
demodulation reference signals. Screening between cards was
also employed to avoid inter-channel coupling. The latter was
accomplished by enclosing each card in its own case, thus
forming the individual modules. Inter-connection between the
modules was done using wire-wrap DIN 41612 connectors.
Figures 4.13 to 4.16 show photographs of the complete ECT
system, as well as one of the transducer modules. Figure 4.13
shows the complete ECT system, including a twin-plane sensor
(of which only one plane is actually being used), the ECT
data acquisition system designed and built by the author, and
the image reconstruction computer showing an image of the
interior of the sensor, which contains a plastic tube full of
polypropylene beads as a test object.
Figure 4.14 shows a view if the interior of the data
acquisition unit seen from the back, showing the power supply
(surrounded by a screen made of perforated steel sheet), the
wired back-plane and the various modules.
Figure 4.15 shows one of the transducer modules, showing the
solder side of the printed circuit board (PCB). On the
component side, the PCB is completely surrounded by a
metallic screen, to avoid unwanted coupling between adjacent
boards. Figure 4.16 shows a transducer module with the screen
removed, showing the component side of the PCB. As can be
observed, it was necessary to use a number of jumper wires in
order to avoid placing PCB tracks on the component side,
which would affect the integrity of the ground plane. Note
the coaxial cable (next to the bottom rail) used to route the
high-frequency signal.

96

A computer program was written in QuickBasic 4.5 to control


and monitor the system, including image reconstruction and
display functions, and is given in appendix B.

4.9

CALIBRATION OF SYSTEM ELECTRONICS

It was decided from the start that calibration of the system


electronics would be done by software and, therefore, no
potentiometers or trimmers (which are prone to mechanical
misadjustment) were included in the hardware. The calibration
procedure itself basically consists of determining a set of
zero-offset and gain correction constants for both the
excitation and the detection circuits, which are then
included in the control software. All the calibration
procedures and, in general, all measurements, were done after
a 30-minute warm-up period.
4.9.1

CALIBRATION OF THE EXCITATION VOLTAGE SOURCES

For each transducer channel, the programmable voltage source


was calibrated to compensate for zero-offset and gain
variations, so that the response of the excitation sources
would be the same for all channels. A three-point calibration
was realised. The actual output V produced by the voltage
sources is assumed to be linearly related to the
corresponding command data D sent by the control computer, in
the way depicted in figure 4.17. This is expressed
mathematically as

V =

m1 D + b

m2 D + b

e0

for

for

D <0

where m1 , m2 and b are constants.


101

(4.31)

Figure 4.17

Three-point calibration curve for excitation


voltage sources

Solving for D on equation (4.31) we find an expression for


the right data to send in order to get a calibrated output:

D =

V - b

for

m1

eb
(4.32)

V - b

for

m2

V < b

The actual output voltage data sent to the data acquisition


unit by the computer is calculated by the control software
using equation (4.32), using default values of 0 for b and 1
for m1 and m2 if the system has not been calibrated.
102

In
practice,
calibration
essentially
involves
the
determination (by measurement) of the constants m1 , m2 and b.
Initially, with the system uncalibrated, commands are sent to
the instrument to set the output voltage to -10 V, 0 and
+10 V, and the actual output is measured, yielding Vmin , b and
Vmax (see figure 4.17). m1 and m2 are then given by

m1 =

Vmax - b

(4.33)

10

and

m2 =

b - Vmin

(4.34)

10

This values of m1 , m2 and b, obtained through the calibration


measurements, are then incorporated into the control software
and used instead of the default values, thus completing the
calibration. This whole procedure, of course, has to be done
for each voltage source (one per channel).
Due to normal variability, each voltage source will have a
different value of Vmax and Vmin (see figure 4.17). For system
consistency and uniformity, therefore, the amplitude of the
output voltage (which ideally should be 10 V) for all the
sources will be limited by the lowest Vmax (or the highest
Vmin ). In our case, the output voltage range was set to
9.6 V.
4.9.2

DETECTOR CALIBRATION

For a rigorous calibration of the detection circuits,


very-low-value capacitance standards [49-51] are required.
103

Nevertheless, due to the unavailability of these rather hard


to find devices, the system was calibrated by comparison
against a Hewlett-Packard 4192A impedance analyzer (IA),
having a capacitance resolution of 0.1 fF.
A two-point calibration was performed, with reference points
at zero and at a value Cr near full-scale, as shown in figure
4.18. This was done for every channel and for the various
ranges (gain settings) within each channel.

Figure 4.18

Two-point detector calibration curve

The uncalibrated measurements Craw are assumed to be linearly


related to the true input capacitance Cinp in the way depicted
in figure 4.18, which is expressed mathematically as

Craw = m Cinp + b
104

(4.35)

where b (the ordinate at the origin) is the uncalibrated


measurement obtained with zero input capacitance (i.e., open
circuit), and m is the slope of the straight line in figure
4.18, given by

m =

Cm - b

(4.36)

Cr

where Cm is the uncalibrated reading corresponding to Cr .


Solving equation (4.35) for Cinp (the true input capacitance),
we obtain

Cinp =

Craw - b

(4.37)

So, for each range in each channel, the zero-offset


calibration constant b and the gain calibration constant m
were determined, and then equation (4.37) was used in the
control software to calculate the true input capacitance,
thus accomplishing calibration of the detector circuits.
The zero-offset constants were determined first, and this was
done by setting the uncalibrated instrument to measure with
the inputs not connected ( i.e., in open-circuit conditions,
Cinp = 0 ), and recording the corresponding readings. This was
done for all channels and ranges.
Next the gain constant were determined. To produce different
values of Cr , we employed an 8-electrode ECT sensor, using
different electrode-pair combinations. The electrode-pair
105

whose capacitance was closest to, but less than the


full-scale value for a particular range (see table 4.1) was
used to calibrate that range. The capacitance of the chosen
electrode-pair was first measured using the impedance
analyzer (HP-4192A) in order to get a value for Cr , and then
it was measured using the uncalibrated ECT data-acquisition
unit, to give Cm . Equation (4.36) was then used to determine
m, with the values obtained previously for b. This was done
for all ranges except for the most sensitive one
(GAC-PGA = 125), because there was not any electrode-pair whose
capacitance was less than the full-scale value for this
range, of just 8.8 fF. This range, however, was never used in
practice because it turned out to be too sensitive.
At last, the standing-value compensation DAC was also
calibrated. Offset correction is already included in the
calibration constant b. To calibrate the DAC gain, the ECT
system was set to the highest measuring range ( 1,100 fF,
GAC-PGA = 1 ) and a capacitance input Cref close to full-scale was
applied to the instrument. Then the compensation data Cdata
sent to the system by the computer was increased gradually
starting from zero, until the system output was nulled. Under
these conditions we know that the actual effective
compensation
signal
C comp = C ref .
The
quotient
mdac = (Ccomp /Cdata ) = (Cref /Cdata ) was then included in the control
software as a standing-value-compensation calibration factor.

4.10 EXPERIMENTAL EVALUATION OF THE SYSTEM


After having calibrated the system electronics as described
in section 4.9, a number of tests were carried out in order
to verify its performance.

106

4.10.1

CAPACITANCE MEASUREMENT COMPARATIVE TEST

In the first test, we took an empty ECT sensor having eight


10-cm-long electrodes and measured the capacitance between
electrode 1 (used for excitation) and each one of the others
(keeping the unused electrodes floating), using first our ECT
system (unit under test, U.U.T.) and then the HP-4192A
impedance analyzer (REFERENCE). When measuring with the ECT
system, channel 1 was used for excitation and channel 2 for
detection. The results are given in table 4.2 and in figures
4.19 and 4.20.

Table 4.2





























Comparative test of ECT system










,
,
$
DETECTION 
REFERENCE
U.U.T.


ELECTRODE  (femtoFarads)  (femtoFarads) 
<
<
$
2
881.3
881.0



<
<
$
3
81.8
81.6



<
<
$
4
39.8
40.0



<
<
$
5
30.3
30.5



<
<
$
6
38.1
38.2



<
<
$
7
75.2
75.0



<
<
$
8
832.5
832.0



4
4

COMPARISON TEST BETWEEN ECT SYSTEM AND
HP IMPEDANCE ANALYZER
4/10/96
CAPACITANCE BETWEEN ELECTRODE PAIRS
8-ELECTRODE SENSOR
EXCITATION ELECTRODE: 1
UNUSED ELECTRODES: FLOATING
EXCITATION: CHANNEL 1
DETECTION: CHANNEL 2

107

Figure 4.19 ECT system vs IA

Figure 4.20 ECT system vs IA


(non-adjacent measurements)

Figure 4.19 shows all the inter-electrode capacitances of


electrode 1, while figure 4.20 shows the capacitances between
non-adjacent electrodes only, on a magnified scale. The
results indicate that there is close agreement between the
ECT system and the impedance analyzer readings.

4.10.2

LINEARITY EVALUATION

The system linearity was tested using the HP-4192A impedance


analyzer as the reference instrument, and employing the same
8-electrode ECT transducer mentioned in the previous section
to provide the capacitance stimulus. The test consisted of
making a number of capacitance measurements between a pair of
opposite electrodes (1 and 5), using both the ECT system
(U.U.T.) and the impedance analyzer (REFERENCE) on each
measurement. The first set of measurements were made with the
sensor empty and, after that, in order to generate an
evenly-spaced range of test capacitances, the inter-electrode
capacitance was increased before making each new set of

108

measurements by pouring a certain amount


material (ceramic beads) into the sensor.

of

dielectric

The results are shown in table 4.3 and figure 4.21, while
figure 4.22 shows the deviation of the measured values from
the best straight line.

Table 4.3


























System linearity test results







$
U.U.T.

(femtoFarads) 
$
30.3

$
42.7

$
54.3

$
62.7

$
74.1

$
84.5

$
94.8



COMPARISON TEST BETWEEN ECT SYSTEM


AND HP IMPEDANCE ANALYZER 3/10/96
EXCITATION: CHANNEL 1
DETECTION: CHANNEL 2
AC GAIN=5 / DC GAIN=2
REFERENCE
(femtoFarads)
30.3
42.7
54.3
62.8
74.3
84.7
95.1

,


<

<

<

<

<

<

<

4

109

Figure 4.21 Linearity test


results

Figure 4.22 Deviation


from best straight line

We can see from these results that the ECT system readings
are very close to the reference values. The correlation
coefficient between the reference and measured values is
0.99999, and the root-mean-square deviation from the best
straight line is 0.037 fF.
4.10.3

INTRINSIC NOISE LEVEL TEST

The total input-referred noise level at the output of the


system was determined. The most sensitive of the calibrated
ranges ( 44 fF, GAC-PGA = 25 ) was selected for this test because
it makes use of all the circuitry including the amplification
stages used in the other two calibrated ranges, and is,
accordingly, the one with the most sources of noise. This can
thus be considered the worst case.
Essentially, the test consisted of looking at the variation
in the output readings obtained when a fixed test capacitance
was connected to the system input. Channel 1 and 2 were used
for excitation and detection respectively, and a pair of
opposite electrodes (1 and 7) in an empty 12-electrode ECT
sensor provided a fixed test capacitance of about 11 fF (the
110

actual capacitance value is not too relevant, since we were


only interested in the system output variation in this test).
Blocks of about 4,000 output readings were examined. In order
to actually see some variation at the system output, the gain
of the binary-step DC-PGA used before the analogue-to-digital
converter (ADC), by default set to 1, had to be increased to
4, so that the noise signal at the input of the ADC would be
larger than the quantisation threshold.
The result is shown in figure 4.23. The peak value of the
intrinsic system noise was found to be equivalent to an input
capacitance of 0.025 fF, at the 99.7% (3) confidence level,
more than ten times smaller than that of the previous UMIST
ECT system based on charge/discharge transducers [9]. As an
aside, notice that the effect of ADC quantisation is clearly
distinguishable in figure 4.23.

Figure 4.23

ECT system noise level

111

The systems noise level is determined mainly by the low-pass


filter used at the output of the phase-sensitive detectors
(see section 4.5). By reducing the filter bandwidth the noise
level can be decreased, at the expense of increasing the
settling time of the transducers. Thus, we are faced with the
all too familiar and often unavoidable trade-off between
noise level (accuracy) and speed. This design uses secondorder Butterworth filters with a bandwidth of 10 kHz, which
result in a theoretical transducer settling time of 188 s
(to 0.01%).
4.10.4

DYNAMIC PERFORMANCE OF THE SYSTEM

Considering a 12-electrode sensor, and using the conventional


single-electrode excitation method as per the protocol
described in section 2.2.1 (chapter 2), the acquisition time
per frame is given by TO = 11 T1 + 66 (T2 + T3 + TADC ) [9], where T1
is the settling time of the transducers, T2 the settling time
of the intermediate amplifiers, T3 the time of communication
between the acquisition system and the computer during a
measurement and TADC the ADC conversion time. To meet the
desired system specifications (section 4.2) a TO of no more
than 10 ms is required.
The actual step response of the transducer circuit, which
depends mainly on the characteristics of the output low-pass
filter in the phase-sensitive demodulator, was measured using
a Tektronix TDS-360 digital storage oscilloscope and is shown
in figure 4.24. We can see that there seems to be agreement
between this curve and the 188 s settling time (to 0.01%)
predicted by theory for a 10 kHz second-order Butterworth
response. To include a safety margin, we consider a
transducer settling time T1 = 200 s. This represents less
than one half of the figure for the earlier UMIST system,
which has a T1 of 500 s (to 0.1%) [9].

112

Figure 4.24

Figure 4.25

Step response of ECT transducer

Step response of intermediate amplifiers


113

The step response of the intermediate DC amplifiers was also


measured using the digital storage oscilloscope and is shown
in figure 4.25. Here we can see that the settling time T2
consists mainly of the rise time, and is about 21 s. Again,
to have a safety margin, we consider T2 = 25 s. This is
about the same as for the previous UMIST system (20 s) [9].
The ADC conversion time TADC is 3 s (ADS7800 specification),
which compares favourably with the value for the previous
UMIST system of 5 s [9].
The time of communication with the computer, T3 , depends on
the particular computer hardware and software used. As was
mentioned in section 4.8, the software to control and monitor
the system was written using QuickBasic 4.5. This was done so
because this programming language is very easy to use, and it
offers features like step-by-step program execution and usercontrollable program interruption/resumption, which are very
convenient during system development. However, being a slow
interpreted language, it results in a rather sluggish program
execution, which, in turn, results in a T3 of 3 ms, using a
66 MHz 486-DX2 PC. Nevertheless, although it was not done due
to time constraints, by using a faster computer and a fast
compiled language like C, it should be quite possible to
optimise the software for speed and bring T3 down to about
20 s, as is indeed the case for the previous UMIST system
[9].
So, if we consider T1 = 200 s, T2 = 25 s, T3 = 20 s and
TADC = 5 s,

we

obtain TO = 11 T1 + 66 (T2 + T3 + TADC ) = 5.5 ms,

compared with 8.5 ms for the previous UMIST ECT system [9].

114

4.11

IMAGE SAMPLES

Despite the fact that the main points of this work, namely,
the development of more sensitive low-noise ECT transducers
and the use of multiple-electrode excitation techniques to
obtain optimum measurements, can be presented without
resorting to any imaging experiments, some sample images are
given below anyway, for illustrative purposes and to show
that the system does actually work satisfactorily. The
images, shown in figures 4.26 to 4.29, were obtained with our
system using the conventional single-electrode excitation
method
and
simple
linear
back-projection
(LBP)
reconstruction, without any post-processing operations
(thresholding, filtering, etc.) except histogram stretching
for contrast enhancement. The 12-electrode sensor shown in
figure 4.13 was used, having an internal diameter of 102 mm.

Figure 4.26
Reconstructed image of a thin perspex bar
(whose cross section occupies 0.4% of the sensing area)
placed in the centre of the sensor, obtained using LBP
115

Figure 4.26 shows the reconstructed image of a thin perspex


( r = 2.5) bar 6 mm in diameter (occupying approximately 0.4%
of the sensing area) placed in the centre of the sensor,
where the sensitivity is lowest. The image presents the
spreading effect that normally results from the crude LBP
algorithm. For this object location and relative permittivity
value, the previous UMIST system was limited (because of its
higher intrinsic noise level) to detecting objects occupying
no less than 2% of the sensing area [9].

Figure 4.27
LBP image of a thin perspex bar (occupying
0.4% of the sensing area) placed half-way between the
centre and the wall of the sensor

Figures 4.27 and 4.28 show the reconstructed images of the


same perspex bar as it is shifted towards the periphery of
the sensing area, where the sensitivity is highest. Figure
4.27 corresponds to the bar placed half-way between the
116

centre and the wall of the sensor, while for figure 4.28 the
bar was adjacent to the wall. Finally, figure 4.29 was
obtained using a 1-pint drinking glass as the test object,
and it clearly shows the annular shape of its cross-section,
albeit blurred by the spreading effect inherent in LBP
reconstruction.

Figure 4.28
LBP image of a thin perspex bar (occupying
0.4% of the sensing area) placed right next to the wall of
the sensor

117

Figure 4.29

4.12

Cross-section of a drinking glass

CONCLUSIONS

The development of a new electrical capacitance tomography


(ECT) system, having a number of improvements over the one
initially designed in UMIST [9], has been presented here.
Instead of the charge-discharge type, the new system uses
AC-based capacitance
transducers, featuring
low-noise
amplifiers and narrow-band filtering techniques in the form
of phase-sensitive detectors, thus achieving a reduction of
intrinsic noise level from 0.26 to 0.025 fF (peak value of
noise level referred to input). This increase in accuracy is
not gained at the expense of speed, though, since the ACbased transducers are more than twice as fast as the chargedischarge ones, as shown by their settling times (200 s to
0.01% vs. 500 s to 0.1%). Additionally, the relative
sensitivity to stray capacitance variations (6 10-6)
118

represents a reduction of more than one order of magnitude


compared with the charge-discharge circuit. The increased
detection capability of this system brings about a whole new
range of possibilities, like the use of a larger number (say
16) of smaller electrodes to increase spatial resolution,
using shorter electrodes for better axial resolution, and
imaging lean-phase flows and low-contrast mixtures, for
example.
The
use
of
phase-sensitive
demodulation
permits
discrimination between the effects of the real (conductive)
and imaginary (capacitive) components of the unknown
admittance. Each channel has its own demodulator, thus
allowing parallel measurement. The latter, coupled with
high-frequency (500 kHz) operation, means that an acquisition
rate of 100 frames per second is possible. The high-stability
excitation and demodulation reference sine-waves are
generated using a novel direct-digital-synthesis (DDS) chip
(AD7008), recently introduced by Analog Devices.
Additionally, arbitrary excitation signals can be applied to
many individual electrodes at the same time. Thanks to this
feature, the system can be employed to investigate the use of
multi-electrode excitation arrangements, including the
grouped-electrode technique suggested by Reinecke and Mewes
[52], and especially the so-called optimum measurement
strategies, so far only tried for resistance tomography
[30,31], which can lead to higher detection signals,
resulting in an even greater detection capability. The use of
this latter multi-electrode excitation method will be the
subject of the next chapter.

119

C H A P T E R

COMPARATIVE

EXPERIMENTAL

MULTIPLE-ELECTRODE

5.1

5 :

EVALUATION

EXCITATION

OF

SINGLE-

AND

METHODS

INTRODUCTION

In this chapter, we show how multi-electrode excitation,


employing optimum excitation patterns [30,31], can be used
in an ECT system in order to achieve an increase in detection
signal levels.
So called optimum or adaptive excitation methods were
originally proposed by Isaacson [30] for conductivity imaging
and have been successfully implemented by Hua [31] and many
others using electrical resistance tomography (ERT). Although
it has not been tried before, the same basic idea can equally
well be applied to permittivity imaging using ECT, as we
shall see.
In the following sections, we report how the increase in
detection signals gained through the use of optimum
multi-electrode
excitation
patterns
(compared
with
conventional single-electrode excitation) was determined
experimentally for a number of different permittivity
distributions, using the multi-excitation ECT system
developed by the author (described in chapter 4).

120

5.2

EXPERIMENTS WITH OPTIMUM MULTI-ELECTRODE EXCITATION

The experiments were carried out using a 12-electrode ECT


sensor with 10 cm long electrodes and an internal diameter of
102 mm (shown in figure 4.13 in chapter 4, it is a twin-plane
sensor but only one plane was used). The experiments
involved:
a) Determining the optimum excitation patterns to detect the
presence of a particular test object (or permittivity
distribution) in an otherwise empty sensor.
b) Applying these optimum excitation patterns to the sensor
and then measuring the changes in the electrode charges Q
caused by the presence of the object inside the sensor
(i.e., the optimum detection signals).
c) Comparing these optimum measurements with the detection
signals Q obtained using the conventional singleelectrode excitation method.
5.2.1

DETERMINATION OF THE OPTIMUM EXCITATION VECTORS

As explained in chapter 2 (sections 2.2.3 and 2.2.3.2), the


optimum excitation vectors to distinguish between two
permittivity distributions and , in the sense that they
are the ones who yield the largest detection signals Q, are
given by the eigenvectors of the distinguishability matrix D
corresponding to the largest eigenvalues [30]. In what
follows, we shall consider the reference permittivity
distribution to be that of the empty sensor.
So the first thing we need to do is to determine the
distinguishability matrix D. As shown in section 2.2.3.2
(chapter 2), provided that the inter-electrode gaps are
small, we can assume that the electrode capacitances to
121

ground are independent of the permittivity distribution


inside the sensing area, and then the matrix D is fully
determined by the mutual capacitances cij and cij (with i` j),
corresponding to the empty sensor and the presence of a test
permittivity
distribution
, respectively. In our
experiments, this mutual capacitances were measured using the
single-electrode excitation method (described in section
2.2.1). The elements of the matrix D were then calculated
using Dij = cij - cij for i` j, and equation (2.26)

Dii = -

Dij
j=1
(ij)

for i=j.
As the next step, singular value decomposition (SVD) was
employed in order to obtain the eigensystem of the symmetric
matrix D. The svd function of the software package MATLAB was
used for this purpose. As mentioned in section 2.2.3.2
(chapter 2), SVD factorizes D as D = X  Y T, and the
eigenvectors of D (i.e. the basic optimal excitation voltage
vectors) are given by the columns of X. The vector
corresponding to the last eigenvalue is not used, and thus,
for a 12-electrode system, we end up with 11 orthogonal
12-dimensional optimal excitation unit vectors.
5.2.2

APPLYING THE OPTIMUM EXCITATION PATTERNS

In practice, due to the need to always keep the electrode


used for detection at zero volts, when we apply the
excitation vectors they have to be level-shifted up or down
by the proper amount. If all the electrode voltages are
increased or decreased by the same amount, the interelectrode voltages will remain unchanged and, provided that
122

the electrode capacitances to ground are independent of the


permittivity inside the sensor (i.e., small inter-electrode
gaps), the detection signals Q will not be affected by so
level-shifting the excitation vectors, as explained in
chapter 2 (section 2.2.3, see equations 2.19 to 2.20).
Moreover, since the original optimum vectors obtained by SVD
are all unit vectors, they also need to be re-scaled after
the level-shifting operation in order to ensure that the
dynamic range of the system is fully exploited and to avoid
possible detector saturation problems.
The full experimental results will be given in following
sections, but, just to illustrate the previous points, let us
consider the first experiment as an example. In this
particular experiment, the test object was a circular perspex
( r = 2.5) bar, 47 mm in diameter, placed in the centre of the
sensor, as shown in figure 5.1.

Fig. 5.1 Permittivity distribution for experiment No. 1:


perspex bar (bar = 47 mm) in the centre of the sensor
(sensor = 102 mm)
123

For this permittivity distribution, the first one of the


basic optimum excitation unit vectors obtained as described
in the previous section (which, by the way, is the best of
the whole set) corresponds to a sinusoidal voltage
distribution, and is shown in figure 5.2.

Fig. 5.2 The first optimum excitation vector for the


permittivity distribution of figure 5.1, obtained by
SVD of the D matrix

Using QuickBasic 4.5, we wrote program that will take this


basic vector and, for each particular electrode, will
level-shift the vector up or down so as to make that
electrode s voltage equal to zero (thus allowing it to be
used for detection), and then, taking into account the
measurement ranges of the ECT system and the specific
permittivity distribution being tested, will stretch or
squash the level-shifted vector in order to use as much as
possible of the 9.6 V that can be supplied by the excitation
124

sources, but without saturating the detection circuits. Since


this has to be done for each electrode, we end up with 12
excitation patterns per basic vector. This excitation
patterns represent the same vector of inter-electrode
voltages, only scaled up or down. The program also determines
the right instrument settings (AC gain, DC gain and
compensation) to be used with each excitation pattern. So,
considering the basic vector of figure 5.2, for example, we
end up with the 12 excitation patterns shown in figure 5.3.

a) excitation pattern No.1

b) excitation pattern No.2

c) excitation pattern No.3

d) excitation pattern No.4

Fig. 5.3

Actual excitation patterns used when applying the


optimum excitation vector of figure 5.2
(continues on the following page)

125

e) excitation pattern No.5

f) excitation pattern No.6

g) excitation pattern No.7

h) excitation pattern No.8

i) excitation pattern No.9

j) excitation pattern No.10

k) excitation pattern No.11

l) excitation pattern No.12

Fig. 5.3

Actual excitation patterns used when applying the


optimum excitation vector of figure 5.2
(continued from the previous page)
126

So, as we can see in figure 5.3, each one of the 12


excitation patterns derived from the basic unit vector of
figure 5.2 corresponds to each one of the 12 electrodes being
used for detection.
The same level-shifting and scaling procedure that was
conducted with the first of the 11 basic optimum excitation
vectors has to be carried out with each one of the remaining
10, and we finally end up with a set of 132 excitation
patterns (12 per vector), which allow us to make the same
number of independent measurements of the change in electrode
charge Q (our detection signal). This measurements are used
to form a vector of optimum measurements M similar to the
one in equation (2.22) (chapter 2, section 2.2.3.1).
It has to pointed out that the excitation vectors obtained as
described in section 5.2.1 are optimum compared with other
unit excitation vectors. In our experiments they have been
scaled up or down to make the best use of the system s
dynamic range and to avoid saturation. Therefore, our
experiments will include the effects of the practical
implementation of this method. This, however, is unavoidable
in a real-life application. Of course, one could compare the
performance of optimum excitation with that of singleelectrode methods by using software simulation, employing
only unit vectors in both cases and without having to
re-scale the vectors at all, but we believe that this would
be of little use in practice, and would not be as significant
as a real-life comparison like the one we have done.
5.2.3

EXPERIMENTAL PROCEDURE

For each object or permittivity distribution tested, the


following experimental procedure was carried out:
a) Initially, the ECT system was warmed-up for 30 minutes.
127

b) With the sensor empty, the 132 mutual capacitances cij


were measured using the conventional single-electrode
excitation method, and then stored.
c) The object or permittivity distribution to be tested was
put in place inside the sensor and, using the conventional
single-electrode excitation method again, the 132 mutual
capacitances cij were measured and then stored. The object
was not removed from the sensor after this measurements
but was left inside, so that the optimum measurements
could be carried out using the same permittivity
distribution, once the corresponding excitation patterns
had been determined.
d) The distinguishability matrix D was calculated from the
cij and cij values and then factorized using SVD to obtain
the basic set of 11 optimum excitation unit vectors, as
described in section 5.2.1.
e) Using a computer program written in QuickBasic 4.5 for
this purpose, the actual optimum excitation voltage
patterns to be applied were derived from the basic set of
optimum unit vectors calculated in (d), as described in
section 5.2.2. The system settings (AC gain, DC gain and
compensation level) to be used when applying each one of
the derived excitation patterns were also determined by
the program. The 132 calculated excitation patterns and
its corresponding system settings were stored in a
computer file, to be used in the next step of the
experiment.
f) With the aid of another QuickBasic 4.5 computer program,
which was written to control the ECT system and employs
the data file generated in (e) (containing the 132 optimum
excitation patterns and their corresponding system
settings), the optimum measurements were conducted. First,
128

with the object still inside the sensor, each one of the
132 optimum excitation patterns was applied and the
corresponding electrode charge was measured, stacking the
measurements in a vector Q . Next, the test object was
removed, leaving the sensor empty, and the process was
repeated, generating a new set of 132 electrode-charge
measurements, which were stacked in a vector Q . The
second set of measurements was then subtracted from the
first, to obtain the vector of optimum measurements:

= Q - Q = [Q1 ,

g) The corresponding

Q2 , ..., Q132 ]

(5.1)

Q measurements for single-electrode

excitation were obtained from the knowledge of the matrix


D = C - C , determined in (d). They were calculated, for
i` j, as

Qij = Dij Vexc

(5.2)

where i and j represent the detection and excitation


electrodes
respectively,
while
Vexc = 9.6 V
is
the
excitation voltage level. The resulting 132 Q values were
stacked in a vector of single-electrode excitation
measurements M SE .
h) Finally, the euclidean length (equal to the 2-norm) of the
vectors M

and M SE was calculated and used as a measure of

the overall detection signal level obtained with each of


the two excitation methods, which could then be compared.

129

5.3

EXPERIMENT RESULTS AND DISCUSSION

Following the procedure just described, six experiments were


performed using different permittivity distributions, in
order to see the effect that each one has on the form of the
optimum excitation patterns, and to measure the magnitude of
the actual increases in detection signal level. The following
results show, for each experiment:
a) The permittivity distribution used.
b) The resulting set of basic optimum excitation unit
vectors. The vectors are shown graphically and are
presented in order, starting with the best (the one that
yields the largest detection signals) at the front of the
figure, to the worse at the back. This ordering is
directly related to the magnitude of the corresponding
singular values (or eigenvalues) of the matrix D [30].
c) The

vector

of

optimum

measurements

M .

The

132

measurements are shown in graphical form, grouped by


excitation vector, i.e., the first 12 measurements are the
ones associated with the first ( best ) excitation vector,
the next 12 correspond to the second vector, etc. Within
each group of 12, the measurements are ordered by
electrode number (from 1 to 12).
d) The vector of single-electrode excitation measurements
M SE . The 132 measurements are shown graphically, grouped
by excitation electrode, i.e., the first 11 measurements
correspond to electrode 1 being used for excitation, the
next 11 correspond to electrode 2 being used for
excitation, etc. Within each group of 11, the measurements
are in order of ascending electrode number (excluding the
number of the electrode used for excitation).

130

e) A comparative table of the size of the two measurement


vectors (based on the value of their 2-norm).
f) A brief discussion and interpretation of the results.

5.3.1

EXPERIMENT No. 1

Fig. 5.4 Permittivity distribution for experiment No. 1:


perspex bar (bar = 47 mm, r = 2.5) in the centre of the
sensor (sensor = 102 mm)

131

Fig. 5.5

Fig. 5.6

Basic set of optimum excitation unit vectors for


experiment No. 1

Experiment No. 1: vector of 132 Q measurements


obtained with optimum excitation, M
132

Fig. 5.7 Experiment No. 1: vector of 132 Q measurements


obtained with single-electrode excitation, M SE

Table 5.1

Experiment no. 1: comparison between optimum (M )


and single-electrode (M SE) measurements

,,
VALUE
UNITS
 PARAMETER



<<$
851
fCoul

Q M Q2



<<$

Q M SE Q2



730
fCoul
<<$





Q M Q2






1.17

N/A


Q M SE Q2







<<$




SIGNAL
17
%




INCREASE
44

133

The theoretical
optimum excitation vectors
for the
permittivity distribution of figure 5.4 (a concentric disk)
were obtained analytically by Isaacson [30]. They were found
to be based on sine and cosine patterns of increasing spatial
frequency, as shown in figure 5.8, with the low-spatialfrequency ones yielding the largest measurements.

Fig. 5.8 Theoretically predicted basic set of optimum


excitation unit vectors for the permittivity distribution
of experiment No. 1

In our experiment, the resulting optimum vectors, shown in


figure 5.5, are also sine- and cosine-like, but their order
is not quite as predicted by the theory. For example, the
tenth and eleventh vectors are sine and cosine patterns with
double the spatial frequency of the first and second ones
(which are the best of the whole set), and, according to
134

theory, they should occupy the third and fourth positions. We


believe that the small differences between the theoretical
vectors and those obtained experimentally can be attributed
to inaccuracies in the positioning of the test bar, and to
the fact that the particular sensor employed had radial
inter-electrode guards projecting inwards from the outer
screen, as shown in figure 5.9. The use of these radial
guards, which are sometimes employed in order to reduce the
very large standing capacitance between adjacent electrodes
(compared with that of the other electrode combinations),
means that the electrode capacitances to ground will actually
show some dependence on the permittivity distribution within
the sensing area, and, therefore, the procedure used to
determine the distinguishability matrix D would not be
strictly valid, since (as explained in section 5.2.1) it
relies on the assumption that the capacitances to ground do
not change.

Fig. 5.9 ECT sensor with projected radial guards (normally


grounded as well as the outer screen), which result in the
electrode capacitances to ground showing some dependence on
the permittivity distribution inside the sensor

135

Even though the previous considerations could mean that the


optimum excitation vectors actually used in the experiments
are really only
approximately
optimum, the resulting
measurements are still quite encouraging. By looking at
figures 5.6 and 5.7, we can see that the detection signals
measured with optimum excitation are clearly larger that
those obtained using single-electrode excitation, with the
first two excitation vectors producing a very strong response
indeed. This is confirmed by the data shown in table 5.1,
which shows that there is an overall signal increase of 17%
when using the calculated optimum excitation vectors.

5.3.2

EXPERIMENT No. 2

Fig. 5.10 Permittivity distribution for experiment No. 2:


perspex bar (bar = 47 mm, r = 2.5) next to the wall of the
sensor, between electrodes 6 and 7 (sensor = 102 mm)

136

Fig. 5.11

Fig. 5.12

Basic set of optimum excitation unit vectors for


experiment No. 2

Experiment No. 2: vector of 132 Q measurements


obtained with optimum excitation, M
137

Fig. 5.13 Experiment No. 2: vector of 132 Q measurements


obtained with single-electrode excitation, M SE

Table 5.2

Experiment no. 2: comparison between optimum (M )


and single-electrode (M SE) measurements

,,
VALUE
UNITS
 PARAMETER



<<$
2,369
fCoul

Q M Q2



<<$

Q M SE Q2



2,049
fCoul
<<$





Q M Q2






1.16

N/A


Q M SE Q2







<<$




SIGNAL
16
%




INCREASE
44

138

By looking at figure 5.11, we can observe that the first


optimum excitation vectors all involve the electrodes that
are close to the test bar, as seems logical to expect. Here
too, the detection signals obtained using optimum excitation
are larger than the ones measured using single-electrode
excitation, as shown in figures 5.12 and 5.13, and the
overall signal increase achieved is 16% (table 5.2).

5.3.3

EXPERIMENT No. 3

Fig. 5.14 Permittivity distribution for experiment No. 3:


two polypropylene bars (bar = 30 mm, r H 2.25) next to the
wall of the sensor, in front of electrodes 4 and 10
(sensor = 102 mm)

139

Fig. 5.15

Fig. 5.16

Basic set of optimum excitation unit vectors for


experiment No. 3

Experiment No. 3: vector of 132 Q measurements


obtained with optimum excitation, M
140

Fig. 5.17 Experiment No. 3: vector of 132 Q measurements


obtained with single-electrode excitation, M SE

Table 5.3

Experiment no. 3: comparison between optimum (M )


and single-electrode (M SE) measurements

,,
VALUE
UNITS
 PARAMETER



<<$
1,949
fCoul

Q M Q2



<<$

Q M SE Q2



1,755
fCoul
<<$





Q M Q2






1.11

N/A


Q M SE Q2







<<$




SIGNAL
11
%




INCREASE
44

141

In this experiment, we see, once more, that the first optimum


excitation vectors involve the electrodes that are close to
the test bars, as shown in figure 5.15. In particular, the
first vector presents a large peak on electrode 10 and a
smaller one on electrode 4, while the second vector shows a
large peak on electrode 4 and a small one on electrode 10.
This seems to suggest that the optimum excitation vectors
tend to concentrate the electric field in the areas where the
permittivity changes occur. Again the detection signals are
larger when using the optimum excitation vectors, as seen in
figures 5.16 and 5.17, and we have an overall signal increase
of 11% (table 5.3).

5.3.4

EXPERIMENT No. 4

Fig. 5.18 Permittivity distribution for experiment No. 4:


two polypropylene bars (bar = 30 mm, r H 2.25) next to the
wall of the sensor, in front of electrodes 5 and 8
(sensor = 102 mm)

142

Fig. 5.19

Fig. 5.20

Basic set of optimum excitation unit vectors for


experiment No. 4

Experiment No. 4: vector of 132 Q measurements


obtained with optimum excitation, M
143

Fig. 5.21 Experiment No. 4: vector of 132 Q measurements


obtained with single-electrode excitation, M SE

Table 5.4

Experiment no. 4: comparison between optimum (M )


and single-electrode (M SE) measurements

,,
VALUE
UNITS
 PARAMETER



<<$
2,435
fCoul

Q M Q2



<<$

Q M SE Q2



1,782
fCoul
<<$





Q M Q2






1.37

N/A


Q M SE Q2







<<$




SIGNAL
37
%




INCREASE
44

144

In this experiment, the optimum excitation vectors (shown in


figure 5.19) appear pretty much the same as in the previous
one, except that the peaks in the first two vectors have been
shifted, reflecting the new positions of the test bars.
Figures 5.20 and 5.21 show that the measurements made using
optimum excitation are even larger this time, compared with
the ones resulting from single-electrode excitation. The
resultant overall signal increase is 37%, as seen in
table 5.4.

5.3.5

EXPERIMENT No. 5

Fig. 5.22 Permittivity distribution for experiment No. 5:


50% stratified flow was simulated using polypropylene beads

145

Fig. 5.23

Fig. 5.24

Basic set of optimum excitation unit vectors for


experiment No. 5

Experiment No. 5: vector of 132 Q measurements


obtained with optimum excitation, M
146

Fig. 5.25 Experiment No. 5: vector of 132 Q measurements


obtained with single-electrode excitation, M SE

Table 5.5

Experiment no. 5: comparison between optimum (M )


and single-electrode (M SE) measurements

,,
VALUE
UNITS
 PARAMETER



<<$
5,678
fCoul

Q M Q2



<<$

Q M SE Q2



3,553
fCoul
<<$





Q M Q2






1.60

N/A


Q M SE Q2







<<$




SIGNAL
60
%




INCREASE
44

147

Once again, we can see in figure 5.23 that, for the first few
optimum excitation vectors, the excitation activity is
concentrated on the electrodes that are close to the area
where the permittivity has changed with respect to the empty
state. Figures 5.24 and 5.25 show that there is a relatively
large increase in the detection signals obtained using
optimum excitation relative to those measured employing
single-electrode excitation. The overall increase in the
level of the detection signals is 60% in this case
(table 5.5).

5.3.6

EXPERIMENT No. 6

Fig. 5.26 Permittivity distribution for experiment No. 6:


annular flow was simulated using polypropylene beads
(interface = 49 mm, sensor = 102 mm)

148

Fig. 5.27

Fig. 5.28

Basic set of optimum excitation unit vectors for


experiment No. 6

Experiment No. 6: vector of 132 Q measurements


obtained with optimum excitation, M
149

Fig. 5.29 Experiment No. 6: vector of 132 Q measurements


obtained with single-electrode excitation, M SE

Table 5.6

Experiment no. 6: comparison between optimum (M )


and single-electrode (M SE) measurements

,,
VALUE
UNITS
 PARAMETER



<<$
12,045
fCoul

Q M Q2



<<$

Q M SE Q2



5,359
fCoul
<<$





Q M Q2






2.25

N/A


Q M SE Q2







<<$




SIGNAL
125
%




INCREASE
44

150

As show in figure 5.27, for this permittivity distribution,


which is concentrated in the periphery of the sensor, the
first few of the optimum excitation vectors (the
best
within the set) are dominated by high spatial frequencies,
with the low-spatial-frequency ones coming last. The
situation is opposite to that of experiment No.1 (a circular
bar in the centre of the sensor), where it was found that the
best
vectors were the ones with the lowest spatial
frequency. This agrees with the theoretical prediction made
by Isaacson [30], who suggested that suggested that low
spatial frequency excitation vectors yield detection signals
that are most sensitive to changes in permittivity near the
centre of the imaging area, while high spatial frequency
excitation vectors yield measurements which respond mainly to
changes near the periphery. The asymmetric characteristic of
the first vector in figure 5.27 was probably caused by
symmetry errors incurred in the actual set up of the annular
permittivity distribution, which was implemented by putting
a plastic pipe inside the sensor and surrounding it with
polypropylene beads.
As seen in figures 5.28 and 5.29, the degree of detection
signal increase obtained through the use of optimum
excitation was quite significant in this particular
experiment, reaching a level of 125% (table 5.6) over the
single-electrode excitation measurements.

5.4

CONCLUSIONS

Experiments were carried out to evaluate the application of


optimum excitation techniques to ECT, using several different
permittivity distributions. This excitation method was
originally proposed [30] for electrical resistance tomography
(ERT) and has not been tried before on ECT. The results of
151

the experiments confirmed that the use of optimum excitation


patterns does increase the level of the detection signals,
and thus the signal to noise ratio (SNR), compared to those
achieved using the standard single-electrode excitation
method. For the permittivity distributions tested, the amount
of signal increase ranged from 11 to 125 per cent. This
figures do not seem like a huge improvement (a 10-fold boost
would be much more welcome, of course), but what they are
really telling us is that the single-electrode excitation
method is not such a bad choice after all.
The experiments showed that the optimum excitation vectors
tend to concentrate the electric field in the areas where the
permittivity changes take place. They also confirmed that
excitation vectors with low spatial frequency are better for
detecting permittivity changes near the centre of the sensor,
while those with high spatial frequency are better for
detecting changes in permittivity occurring in the periphery.
Although the use of optimal excitation techniques clearly
results in increased detection signals, in general the
magnitude of this increase will probably not be large enough
to justify the additional expense associated with the more
complex hardware required for optimal excitation (with one
adjustable voltage source per electrode), except in very
special applications demanding the ultimate detection
capability.

152

C H A P T E R

ITERATIVE

6.1

LINEAR

6 :

BACK-PROJECTION

IMAGE

RECONSTRUCTION

INTRODUCTION

In this chapter, we shall present a new image reconstruction


method for ECT [53], which uses the LBP operation in an
iterative way, and is based on the concept (admittedly quite
unconventional) of establishing an analogy between the image
reconstruction procedure and the action of an automatic feedback control system. This work was realised in collaboration
with Dr. W.Q. Yang, of UMISTs Process Tomography Group, and
the results obtained are pretty encouraging, despite the
rather unusual approach followed.

6.2

PREVIOUS WORK

The problem of ECT image reconstruction is one that has not


yet been fully resolved. The first reconstruction method to
be developed (and still the most widely employed) was the
linear back-projection (LBP) algorithm [4], already described
in chapter 2. This is a simple and fast (one-step) procedure
which is easy to implement, but does not take into account
the non-linear relationship between the permittivity
distribution and the inter-electrode capacitances, yielding
poorly defined qualitative images having a considerable
blurring effect. Attempts have been made to improve the LBP
images by using a thresholding operation [9], however, the
153

results are not always satisfactory, since the right


threshold level depends on the specific permittivity
distribution to be imaged.
In general, the way to obtain improved quantitative images is
through the use of iterative algorithms. The main
disadvantage of these reconstruction methods is that they are
inherently slow, both because of their iterative nature and
due to the fact that the forward problem has to be solved
(which is a computationally intensive task). A number of
iterative methods have been published. They all rely on the
calculation (by solving the forward problem) of the interelectrode capacitances associated with the image, which are
then compared with the actual measured values. If the
difference between the two is excessive then a new improved
image is produced and the whole process is repeated, until
the capacitance error is below a predetermined tolerance. The
difference between the various iterative methods lies in how
the successive improved images are generated.
In the method reported by Chen et al [54], the standard LBP
image is used as the first approximation. The corresponding
permittivity distribution is then used to solve the forward
problem and calculate the corresponding inter-electrode
capacitances. If the difference between the calculated and
measured capacitances is acceptably small the procedure ends,
otherwise the permittivity distribution information is used
to obtain an improved set of sensitivity maps, which are then
used to produce a new improved image, and the whole procedure
is repeated.
In [52], Reinecke et al describe an LBP-based iterative
scheme. It begins with the standard LBP image, which is used
to obtain the corresponding inter-electrode capacitances.
These are subtracted from the measured capacitances to
produce the capacitance error. The error is itself
154

effectively back-projected to generate a correction image,


which is then added to the current image to obtain a new
improved one. Then the whole process can be repeated again.
Research in Norway has led to so-called model-based
reconstruction (MOR) [55]. In this method, the permittivity
distribution (i.e., the image) is represented by a set of
parameters. Now the aim is no longer to obtain the right
image, but to find the correct set of parameters (the ones
that correspond to the true permittivity distribution). The
initial image is determined using a procedure based on the
standard LBP algorithm, and the initial set of parameters are
derived from this image. Next, the forward problem is solved
using finite-element techniques and the capacitances
corresponding to the initial image are calculated. These
capacitances are then subtracted from the measured ones to
determine the capacitance error. If the sum-of-squares
capacitance error exceeds a pre-defined value, it is then fed
to an optimiser which will produce an improved set of
parameters (corresponding to a new improved image). Like in
the other iterative methods, the objective is to minimize the
capacitance error by repeatedly applying the previous
procedure. The algorithm ends when the capacitance error
falls below the predetermined tolerance.
Quite apart from the methods discussed so far, a radically
different way to tackle ECT image reconstruction has emerged
recently, based on the use of artificial neural networks
[56-58], which looks very promising indeed. Although the
neural network would initially need a lengthy training
process, once this is completed the actual reconstruction
could be potentially very fast. We believe, however, that
these methods still need to mature a bit more.
A good review of various image reconstruction methods for ECT
can be found in [59].
155

6.3

A RECONSTRUCTION ALGORITHM INSPIRED ON FEED-BACK CONTROL

6.3.1

FIRST APPROACH: ITERATIVE LBP RECONSTRUCTION

We will develop our method beginning with the iterative LBP


scheme depicted in figure 6.1. We shall assume a 12-electrode
ECT sensor. The initial image is obtained using normal LBP as
described in section 2.2.2.3 of chapter 2. Any negative grey
levels, corresponding to a relative permittivity less than
one (a physically impossible thing), are truncated to zero.
Then, referring to the figure, the image grey levels G
(ranging from 0 to 1) are converted to permittivity values in
the range low to high , which are then used to solve the
forward

problem

in

order

to

determine

the

calculated

capacitances CC . These are subtracted from the measured


capacitances and normalized. Using the sensitivity map
information, the 66 normalized capacitance errors E are then
back-projected and the resulting error image is added to
the current permittivity distribution image, thus generating
a new and improved permittivity image (any negative grey
levels are, once again, truncated to zero). This improved
image can now take the place of the initial image and the
whole process can be repeated. This can be done many times
until the capacitance errors are within acceptable limits.
The author implemented this method in QuickBasic 4.5, using
the finite-element package PC-OPERA to solve the forward
problem. Tests were done using simulated measurements, also
obtained using PC-OPERA, employing the polar-grid sensor
model described in chapter 3 (section 3.2). As shown in
figure 6.2, a small test object with relative permittivity of
2.1 ( high ) was simulated in a background with relative

permittivity of 1 ( low ), placed roughly half-way between the

centre and the edge of the imaging area. The resulting


reconstructed images for 10 successive iterations are also
shown in the figure, along with the ideal image.
156

Figure 6.1

Iterative LBP image reconstruction algorithm

157

A progressive, but rather slow, improvement of the


reconstructed images is readily observable. Note that, as the
iteration process goes on, the peak in the image gets
narrower and its magnitude increases.

a) simulated test object

b) Location of electrodes

c) Initial LBP reconstruction

d) 1st iteration

Figure 6.2 Reconstructed images obtained using the


iterative LBP algorithm shown in figure 6.1
(continued on the following page)

158

e) 2nd iteration

f) 3rd iteration

g) 4th iteration

h) 5th iteration

i) 6th iteration

j) 7th iteration

Figure 6.2 Reconstructed images obtained using the


iterative LBP algorithm shown in figure 6.1
(continues from the previous page and is continued on the
following page)
159

k) 8th iteration

l) 9th iteration

m) 10th iteration

n) True image

Figure 6.2 Reconstructed images obtained using the


iterative LBP algorithm shown in figure 6.1
(continued from the previous page)

Going one step further, we now tried to apply some gain to


the capacitance errors that are used to correct and improve
the images, in order to see if this could speed up the
correction process. So we did the same simulated experiment
as before, but this time the normalized capacitance errors
were multiplied by 2 prior to being back-projected. Indeed,
the results of this experiment, shown in figure 6.3, indicate
160

some (albeit not too much) improvement compared with the


previous case, where the error gain was equal to one, in
that the peak in the images is higher and slightly narrower.

a) Initial LBP reconstruction

b) 1st iteration

c) 2nd iteration

d) 3rd iteration

Figure 6.3 Reconstructed images obtained using the


iterative LBP algorithm shown in figure 6.1 when a gain
of 2 is applied to the normalized capacitance errors E
(continued on the following page)

161

e) 4th iteration

f) 5th iteration

g) 6th iteration

h) 7th iteration

i) 8th iteration

j) 9th iteration

Figure 6.3 Reconstructed images obtained using the


iterative LBP algorithm shown in figure 6.1 when a gain
of 2 is applied to the normalized capacitance errors E
(continues from the previous page and is continued on the
following one)
162

k) 10th iteration

n) True image

Figure 6.3 Reconstructed images obtained using the


iterative LBP algorithm shown in figure 6.1 when a gain
of 2 is applied to the normalized capacitance errors E
(continues from the previous page)

The preceding results were interpreted as an indication that,


perhaps, the process of image refinement could be optimised
through careful manipulation of the capacitance errors, much
in the same way that a feed-back control system can be
improved by using the right compensator or controller [60].
6.3.2

AN ALGORITHM BASED ON A CONTROL-SYSTEM ANALOGY

So, as a first step towards the realisation of our new


algorithm, we will put the iterative LBP algorithm discussed
above in a form that resembles the block diagram of a feedback control system, trying to establish an analogy between
the two. To begin with, it should be quite clear that the
iterative LBP algorithm of figure 6.1 can be represented by

163

the block-diagram of figure 6.4., where the meaning of the


symbols is as follows:

the measured normalized capacitances

c(n)

the current calculated normalized capacitances

 c(n)

the current normalized capacitance errors

G(n)

the current

G(n)

the current image

G(n-1)

the previous image

(n)

the current permittivity distribution (calculated


from the previous image)

LBP

linear back-projection

SPF

solve forward problem

G
D

error image

grey-level to permittivity linear transformation


unit delay.

Figure 6.4

Block-diagram representation of the iterative


LBP algorithm shown in figure 6.1

164

Figure 6.4 shows how the current image, G(n), is calculated


by adding the previous image and the error image, which, in
turn, is obtained by linear back-projection of the normalized
capacitance errors. The latter are computed as the difference
between the measured normalized capacitances and the
calculated ones, which are determined by solving the forward
problem using the permittivity distribution obtained from the
previous image, G(n-1).
The system of figure 6.4 can be re-arranged and put in the
unity feed-back configuration of figure 6.5.

Figure 6.5 The iterative LBP algorithm put in a form


analogous to a unity feed-back control system

In the configuration of figure 6.5, the section enclosed by


the dotted line will be our process, and the goal of the
system is to achieve an output ( c , the calculated
capacitances) equal to the reference input or set point
(the measured capacitances m ), while generating the
165

reconstructed images as part of the process. Put in this way,


the reconstruction algorithm has the same structure as a
sampled-data multi-variable process control system. As shown
in figure 6.6, we can then include a compensator in the loop,
in the form of a proportional-integral-derivative (PID)
controller [60] (the type often used to control non-linear
industrial processes), in an attempt to optimize the system.
By optimize we mean make its response as fast as possible,
while maintaining accuracy and stability (i.e., convergence).

Figure 6.6 Use of a PID controller to optimize the


iterative LBP reconstruction algorithm (the process is
the section of figure 6.5 enclosed by the dotted line)

The output of a PID controller is given, as a function of its


error signal input, by the sum of three terms: the first is
proportional to the error signal itself, the second is
proportional to the integral of the error signal, and the
third is proportional to the derivative of the error signal.
In our case, we employ the sampled-data version, which,
166

referring to figure 6.6, can be expressed by the following


recurrence equation

(n) = P  (n) + I  (k) + D [  (n) -  (n-1) ]

(6.1)

k=1

where the sequence

represents the controller output, while

 is the sequence of the normalized capacitance errors (the

controller input). The three parameters P, I and D, are


called the proportional, integral and derivative control
gains, respectively, and are normally adjusted by trial and
error [60]. The proportional gain P has the effect of
speeding up the controller, but if it is too high can cause
instability. The integral gain I makes the steady-state error
go to zero, but it adds inertia to the system, and hence
can slow down its response or cause overshooting. Finally,
the derivative gain D is used to provide speed, but it will
not respond to a constant error signal (since its derivative
would be zero). The process of adjusting the PID parameters
in order to obtain a satisfactory feed-back system is known
as tuning the controller.
In our work, we used a variable tuning strategy. For the
first three iterations (i.e., for n d 3), we used P = 2, I = 0
and D = 0.5. This was done so in an attempt to quickly reduce
the large initial capacitance error. After that, for n > 3,
P = 1, I = 0.5 and D = 0 were used, in order to gradually
reduce the remaining error. These values were found
empirically
to
give
satisfactory
results
(through
experimentation
using
data
measured
for
different
permittivity distributions).
167

Finally, let us say a word about the forward problem solver.


This is a key point in the performance of any ECT iterative
image reconstruction algorithm. The solution of forward
problem is a computationally intensive task which, depending
on the way it is done, can take an excessively long time.
Roughly speaking, the conventional way to calculate the
inter-electrode capacitances involves three steps [61]:
a) Solve Poissons equation using the finite-element method
(FEM), in order to find the potential distribution inside
the sensor.
b) From the potential distribution the electric field can be
found, and then Gauss Law can be applied to each electrode
in order to find the electrode charges.
c) The capacitances can then be calculated as charge divided
by voltage (the excitation voltage).
In our algorithm, however, we exploited an improved forward
problem solver recently developed [62,63], which enables the
capacitance matrix to be extracted directly from the FEM
representation of the sensor without the need to calculate
intermediate field solutions. Consequently, the number and
complexity of the matrix operations required is greatly
reduced, resulting in a much faster execution.
An implementation of the iterative feed-back control based
image reconstruction algorithm described in the this section
was realised by Dr. W Q Yang on a PC computer using C++
language. Thanks in part to the improved forward problem
solver used, the algorithm can run at a speed of 12
iterations per second on a 200 MHz Pentium computer. This
implementation was used by the author to carry out evaluation
tests in order to establish the algorithms performance (see
the following section).
168

Figure 6.7 Geometry of the FE sensor model used, with


1,200 triangular elements inside the sensing area

6.4

EXPERIMENTAL EVALUATION OF THE

FEED-BACK ALGORITHM

The feed-back control based iterative algorithm was tested


by the author using real data from different permittivity
distributions. Using a 12-electrode sensor connected to the
ECT data acquisition system described in chapter 4, a set of
66 normalized capacitance values was acquired for each
permittivity distribution. The algorithm was then used
off-line to do the image reconstruction.
6.4.1

RESULTS OF IMAGE RECONSTRUCTION TESTS

The results of six tests are presented next. For the first
test, corresponding to a concentric disk of high permittivity
material, we show how the reconstructed images improve and
the normalized capacitance errors decrease as the iteration
process progresses. For the other five tests only the final
image obtained after 19 iterations is given, alongside the
LBP image and the true image, for comparison.
169

Figure 6.10 Test No. 1: reduction in the RMS value of the


normalized capacitance errors with the number of iterations

Figures 6.9 and 6.10 illustrate the convergence towards the


true image. The RMS value of the normalized errors, plotted
in figure 6.10 for every iteration, was calculated as

(6.2)

where

 i are the differences between the measured and

calculated normalized capacitances for each one of the 66


electrode pair combinations.

172

6.4.2

DISCUSSION

Overall, the results obtained are quite encouraging. The


blurring associated with LBP is reduced considerably, and the
structure and shape of the test permittivity distributions is
much clearer in the images generated with the iterative
algorithm compared with simple LBP. On average, the RMS
normalized capacitance error was reduced by 51%, for the
permittivity distributions tested.
Figure 6.8 shows how the iterative algorithm overcomes one of
the main problems found with LBP alone, namely, the
difficulty to make a good reconstruction of objects located
near the centre of the sensor. In figure 6.8 (b) we can see
that simple LBP results in a very spread image which
underestimated the permittivity of the test object, whereas
the iterated image in (c) shows much better agreement with
the true image. Figures 6.9 and 6.10 illustrate the
convergence towards the true image.
Figure 6.12 is particularly interesting, because it
illustrates another typical problem associated with LBP. In
the LBP image, the bar near the centre of the sensor is
practically obscured by the one near the wall, in what could
be considered a shielding effect, while in the iterative
algorithm image the grey level of the bar near the centre has
been restored and the shape of both bars is much more clearly
defined.
Figure 6.14 shows another good example of the effectiveness
of the iterative algorithm. In this case we are dealing with
a slightly more complex permittivity distribution, which, due
to the strong blurring, is hard to identify in the LBP image.
However, in the iterative algorithm image the structure of
the distribution is clearly recognizable.

178

6.5

CONCLUSIONS

A novel iterative image reconstruction algorithm for ECT has


been developed, based on the application of ideas borrowed
from control theory. The algorithm searches for a minimum in
the space described by the difference between the measured
capacitances and those associated with the reconstructed
image (i.e., the capacitance error). It exploits a new
procedure for the solution of the forward problem, based on
a modified Cholesky factorization [62,63], reducing the
computational burden and allowing a fast determination of the
calculated capacitances.
Imaging
tests
conducted
using
several
permittivity
distributions show that, after 10 iterations, the new
algorithm compares favourably with the standard LBP. After 19
iterations, the improvement over LBP is quite substantial,
with a 51% reduction (on average) in the normalized
capacitance error. Thanks to the use of a new fast forward
problem solver, the algorithm can run at a speed of 12
iterations per second on a 200 MHz Pentium computer.
It should be a very interesting matter to establish the
relationship between the control-theory based algorithm
described here and more formal approaches (mathematically
speaking) based on optimization theory, like the modified
Newton-Raphson algorithm often used in electrical resistance
tomography [32]. This, however, is a subject that requires a
complex mathematical analysis which goes beyond the scope of
this work.

179

C H A P T E R

7 :

CONCLUSIONS

FURTHER

7.1

AND

WORK

CONCLUSIONS

In this final chapter, we shall look at our work in


retrospect and try to sum up what has been accomplished. It
was said in the beginning that our general objectives were,
firstly, to investigate ways to improve the detection
capability of ECT systems, in particular its sensitivity to
permittivity changes near the centre of the sensor, and
secondly, to try to perfect the image reconstruction
procedure.
In order to improve the detection ability, a high-sensitivity
low-noise ECT transducer using high-frequency (500 kHz) AC
(sine-wave) excitation and phase-sensitive detection was
developed, featuring a total input-referred noise level of
0.025 fF (at the 3 confidence level) for a 10 kHz
bandwidth, which represents a ten-fold improvement over the
charge/discharge transducer previously designed at UMIST.
The use of parallel-field excitation, which initially was
considered to be a good choice as a means to increase the
sensitivity in the central area (in an attempt to create an
analogy with X-ray tomography), was studied in depth using
finite-element simulation (chapter 3). Nevertheless, it was
established that it yields quite unsatisfactory images due to
the rather flat sensitivity maps associated with this type
of excitation. It was further established that, through the
180

use of parallel-field excitation alone, it is impossible to


obtain a complete set of independent measurements required to
determine the inter-electrode capacitances.
Having found parallel field excitation unsatisfactory, we
then turned our attention to the use of other types of
multi-electrode excitation arrangements, in particular
optimal or adaptive excitation methods aimed at producing the
largest possible detection signals for a particular
permittivity distribution (i.e., maximizing its visibility or
distinguishability).
Using our AC-based transducer, a multi-electrode excitation
ECT system was designed and built in order to test the
practical application of optimal excitation techniques.
Through a series of experiments with different permittivity
distributions, it was confirmed that the use of optimal
excitation patterns does yield larger measurements compared
with the traditional single-electrode excitation method. The
magnitude of the signal increase achieved, though, is not
considered large enough (except for special applications
demanding the ultimate detection capability) to justify the
additional cost associated with optimal excitation designs,
which require the use of one adjustable voltage source per
electrode. What this really means, however, is that singleelectrode excitation actually produces measurements which are
nearly as good as the best that can be obtained.
Finally, concerning the issue of image reconstruction, an
iterative algorithm has been developed based on the quite
singular approach of viewing the reconstruction process as a
feed-back control system (chapter 6), with quite encouraging
results.

181

7.2

RECOMMENDATIONS FOR FUTURE WORK

The more sensitive low-noise AC-based transducers developed


in this work are a first step towards the production of new
ECT systems featuring smaller electrodes. If the electrodes
are made narrower, more of them can be placed around the pipe
or vessel, which would result in higher imaging resolution.
The feasibility of systems using 16 or 20 electrodes should,
therefore, be investigated (initially through simulation). On
the other hand, if the electrodes are made shorter, then the
axial resolution would be increased (i.e., the width of the
slice imaged would be smaller).
However, we believe that a bolder move forward, also brought
about by the possibility of using smaller electrodes, would
be the development of fully 3-dimensional (3-D) ECT systems,
as has been done with resistance tomography [64]. Imagine,
for example, a sensor formed by, say, four contiguous rings
of 12 short (square) electrodes each, as shown in figure 7.1.

Figure 7.1

Side view of a

182

3-D

ECT sensor

Using a sensor like that of figure 7.1 would result in a much


more accurate reconstruction of the permittivity distribution
inside, since most distributions found in practice are 3-D in
nature, and not 2-D as assumed when using a conventional ECT
sensor. To develop a 3-D ECT system, the sensitivity
distribution corresponding to each pair of electrodes would
have to be determined, possibly through 3-D finite-element
simulation. The inter-electrode capacitance measurements
(there would be 1,128 of them) could be carried out using the
single- electrode excitation method, and image reconstruction
could be done employing 3-D LBP. Of course, the system would
be considerably slower than a 2-D one, and thus would have to
be restricted to slowly changing processes.
Finally, one point that is long overdue is the development of
an iterative image reconstruction algorithm for ECT based on
conventional optimisation theory (for example, using the
Newton-Raphson method, as is done in resistance tomography).
It should be quite interesting to compare the performance of
such
an
algorithm
with
the
control-theory based
reconstruction method presented in chapter 6. Furthermore,
the ability of this latter algorithm to reconstruct highcontrast permittivity distributions (water/gas mixtures, for
example) needs to be assessed.

183

R E F E R E N C E S

1. Beck M S and Williams R A, Process tomography: a European


innovation and its applications, Measurement Science and
Technology, 1996, 7 (3), pp. 215-224.
2. Williams R A and Beck M S (ed), Process Tomography Principles Techniques and Applications, ButterworthHeinemann, 1995.
3. Beck M S, Plaskowski A and Green R G, Imaging for
measurement of two-phase flow, in Veret C (ed), Flow
Visualization IV
(1987). Proceedings of the 4th
International Symposium on Flow Visualization, 26-28
August 1986 (Hemisphere Publishing Corporation), pp.
585-588.
4. Xie C G, Plaskowski A and Beck M S, 8-electrode
capacitance system for two-component flow identification.
Part 1: Tomographic flow imaging, IEE Proceedings A, 1989,
136 (4), pp. 173-183.
5. Xie C G, Plaskowski A and Beck M S, 8-electrode
capacitance system for two-component flow identification.
Part 2: Flow regime identification, IEE Proceedings A,
1989, 136 (4), pp. 184-191.
6. Fasching G E and Smith N S, High resolution capacitance
imaging
system,
Technical
Note
DOE/METC-88/4083
(DE88010277), US Department of Energy, Office of Fossil
Energy, Morgantown Energy Technology Center, 1988.

184

7. Fasching G E and Smith N S, Tree-dimensional capacitance


imaging
system,
Technical
Note
DOE/METC-90/4097
(DE90000470), US Department of Energy, Office of Fossil
Energy, Morgantown Energy Technology Center, 1990.
8. Fasching G E and Smith N S, A capacitive system for
three-dimensional imaging of fluidized beds, Review of
Scientific Instruments, 1991, 62 (9), pp. 2243-2251.
9. Huang S M, Xie C G, Thorn R, Snowden D and Beck M S,
Design of sensor electronics for electrical capacitance
tomography, IEE Proceedings G, 1992, 139 (1), pp. 83-88.
10.

Xie C G, Huang S M, Hoyle B S, Thorn R, Lenn C,


Snowden D and Beck M S, Electrical capacitance
tomography for
flow imaging: system
model for
development of image reconstruction algorithms and
design of primary sensors, IEE Proceedings G, 1992, 139
(1), pp. 89-98.

11.

Xie C G, Huang S M, Hoyle B S, Lenn C, and Beck M S,


Transputer-based electrical capacitance tomography for
real-time imaging of oil-field pipelines, in Beck M S,
Campogrande E, Morris M, Williams R A and Waterfall R C
(ed), Tomographic Techniques for Process Design and
Operation (1993). Proceedings of the European Concerted
Action on Process Tomography, Manchester, UK, 26-29
March 1992 (Computational Mechanics Publications),
pp. 333-346.

12.

Yang W Q, Stott A L, Beck M S, and Xie C G, Development


of capacitance tomographic imaging systems for oil
pipeline measurements, Review of Scientific Instruments,
1995, 66 (8), pp. 4326-4322.

185

13.

He R, Beck C M, Waterfall R C and Beck M S, Applying


capacitance
tomography
to
combustion
phenomena,
Proceedings of the European Concerted Action on Process
Tomography 93, Karlsruhe, Germany, 27 March 1993,
pp. 299-301.

14.

He R, Beck C M, Waterfall R C and Beck M S, Development


of capacitance measurements towards tomographic imaging
of flames, in Grattan K T V and Augousti A T (ed),
Sensors VI: Technology, Systems and Applications (1993),
IOP Publishing, pp. 365-368.

15.

He R, Beck C M, Waterfall R C and Beck M S,


Finite-element modelling and experimental study of
combustion phenomena using capacitance measurement,
Proceedings of the European Concerted Action on Process
Tomography 94, Oporto, Portugal, 24-26 March 1994,
pp. 367-376.

16.

He R, Xie C G, Waterfall R C, Beck M S and Beck C M,


Engine flame imaging using electrical capacitance
tomography,
Electronics
Letters,
1994,
30
(7),
pp. 559-560.

17.

He R, Waterfall R C, Beck C M, Xie C G and Beck M S,


Electrical impedance techniques to study combustion
phenomena, Proceedings of the European Concerted Action
on Process Tomography 95, Bergen, Norway, 6-8 April
1995, pp. 102-109.

18.

Wang S J, Dyakowski T and Williams R A, Measurement of


fluidization dynamics in a fluidized bed using
capacitance tomography, Proceedings of 1995 IChemE
Research Event, 5-6 January, Edinburgh, UK, pp. 728-730.

186

19.

Wang S J, Dyakowski T, Xie C G, Williams R A and


Beck M S, Real time capacitance imaging of bubble
formation at the distributor of a fluidized bed,
Chemical Engineering, 1995, 56, pp. 95-100.

20.

Wang S J, Dyakowski T, Xie C G, Williams R A and


Beck M S, Tomographic imaging of phase distribution and
spatial correlation of the fluidization dynamics in a
fluidized bed, Proceedings of PARTEC 95, Nurnberg,
Germany, pp. 115-124.

21.

Yu Z Z, Lyon G M, Zeiabak S A, Tan H P, Peyton A J and


Beck
M
S,
Towards
optimising
the
sensitivity
distribution of electrical tomography sensors, in
Augousti A T (ed), Sensors and their Applications VII
(1995). Proceedings of the Seventh Conference on Sensors
and their Applications, Dublin, Ireland, 10-13 September
1995 (IOP Publishing), pp. 278-283.

22.

Yang W Q, Xie C G, Gamio J C and Beck M S, Design of a


capacitance tomographic imaging sensor with uniform
electric field, Proceedings of the European Concerted
Action on Process Tomography 95, Bergen, Norway, 6-8
April 1995, pp. 266-276.

23.

Yang W Q, Spink D M, Gamio J C and Beck M S, Sensitivity


distributions of capacitance tomography sensors with
parallel field excitation, Measurement Science and
Technology, 1997, 8 (5), pp. 562-569.

24.

Haus H A and Melcher J R, Electromagnetic Fields and


Energy, Prentice-Hall, 1989

25.

Kip A F, Fundamentals of Electricity and Magnetism,


McGraw-Hill, second edition 1969, pp. 100-101.

187

26.

Page L and Adams N I, Principles of Electricity, Van


Nostrand, fourth edition 1969, pp. 53-56.

27.

Pidduck F B, A Treatise on Electricity, Cambridge


University Press, London, second edition 1925, pp.75-77.

28.

Maxwell J C, A Treatise on Electricity and Magnetism


(Vol. I), Clarendon Press, 1873, pp. 88-97.

29.

Campbell A
and
Childs E C,
The
Measurement
of
Inductance, Capacitance, and Frequency, Macmillan, 1935,
pp. 275-276.

30.

Isaacson, D., Distinguishability of conductivities by


electric current computed tomography , IEEE Trans. Med.
Imaging, vol. MI-5, pp. 91-95, 1986.

31.

Hua, P., E. J. Woo, J. G. Webster and W. J. Tompkins,


Improved methods to determine optimal currents in
electrical impedance tomography , IEEE Trans. Med.
Imaging, vol. 11, No. 4, pp. 488-495, 1992.

32.

Yorkey, T. J., J. G. Webster and W. J. Tompkins,


Comparing reconstruction algorithms for electrical
impedance tomography , IEEE Trans. Biomed. Eng., vol.
BME-34, pp. 843-851, 1987.

33.

Hua, P., E. J. Woo, J. G. Webster and W. J. Tompkins,


Iterative reconstruction methods using regularization
and optimal current patterns in electrical impedance
tomography , IEEE Trans. Med. Imaging, vol. 10, No. 4,
pp. 621-628, 1991.

188

34.

Huang S M, Green R G, Plaskowski A ans Beck M S, A high


frequency stray-immune capacitance transducer based on
the charge transfer principle, IEEE Transactions on
Instrumentation
and
Measurement,
1988,
37
(3),
pp. 368-373.

35.

Huang S M, Capacitance Tranducers for Concentration


Measurement in Multi-component Flow Processes, PhD
Dissertation, University of Manchester, UK, 1986.

36.

Maroli D, Sardini E and Taroni A, Measurement of small


capacitance
variations,
IEEE
Transactions
on
Instrumentation
and
Measurement,
1991,
40
(2),
pp. 426-428.

37.

Yang W Q and Stott A L, Low value capacitance


measurements for process tomography, in Beck M S,
Campogrande E, Morris M, Williams R A and Waterfall R C
(ed), Tomographic Techniques for Process Design and
Operation (1993). Proceedings of the European Concerted
Action on Process Tomography, Manchester, UK, 26-29
March 1992 (Computational Mechanics Publications),
pp. 107-114.

38.

Yang W Q, Stott A L and Beck M S, High frequency and


high resolution capacitance measuring circuit for
process tomography, IEE Proceedings - Circuits, Devices
ans Systems, 1994, 141 (3), pp. 215-219.

39.

Kuroda S, A simple stray-free capacitance meter using an


operational
amplifier,
IEEE
Transactions
on
Instrumentation
and
Measurement,
1983,
32
(4),
pp. 512-513.

189

40.

Pickup E, Deloughry R and Hartley T, Measurement of


small capacitance changes for use in tomographic
imaging, Proceedings of the European Concerted Action on
Process Tomography 95, Bergen, Norway, 6-8 April 1995,
pp. 256-265.

41.

Graeme J G, Tobey G E and Huelsman L P (eds.),


Operational Amplifiers, 1971, McGraw-Hill, pp. 233-235.

42.

Reinecke N and Mewes D, Recent developments and


industrial/research
applications
of
capacitance
tomography, Measurement Science and Technology, 1996, 7
(3), pp. 233-246.

43.

Bates R H T, McKinnon G C and Seagar A D, A limitation


on
systems
for
imaging
electrical
conductivity
distributions ,
IEEE
Transactions
on
Biomedical
Engineering, vol. BME-27, pp. 418-420, 1980.

44.

Barber D C and Seagar A D, Fast reconstruction of


resistance images, Clinical Physics and Physiological
Measurement, vol. 8, supplement A, 1987, pp. 47-54.

45.

Blair D P and Sydenham P H, Phase sensitive detection as


a means to recover signals buried in noise, Journal of
Physics E: Scientific Instruments, 1975, volume 8,
pp. 621-627.

46.

Clayton G B and Newby B W G, Operational Amplifiers,


third edition 1992, Newnes (Butterworth-Heinemann),
p. 28.

47.

Graeme J, Feedback models reduce op-amp circuits to


voltage dividers, EDN, June 20, 1991, pp. 139-158.

190

48.

Analog Devices,
February 1995.

AD7008 CMOS DDS Modulator (data sheet),

49.

Snow C, A standard of small capacitance, Journal of


Research of the National Bureau of Standards, vol. 42,
March 1949, pp. 287-308.

50.

Moon C and Sparks C M, Standards for low values of


direct capacitance, Journal of Research of the National
Bureau of Standards, vol. 41, November 1948, pp. 287-308

51.

A.M. Thompson, The cylindrical cross-capacitor as a


calculable standard , IEE Proceedings, 106B, 1959,
pp 307-310.

52.

Reinecke N and Mewes D, Resolution enhancement for


multi-electrode capacitance sensors, Proceedings of the
European Concerted Action on Process Tomography 94,
Oporto, Portugal, 24-26 March 1994, pp. 50-61.

53.

Yang W Q, Gamio J C and Beck M S, A fast iterative image


reconstruction algorithm for capacitance tomography, in
Augousti A T and White N M (eds), Sensors and their
Applications VIII (1997). Proceedings of the Eighth
Conference on Sensors and their Applications, Glasgow,
Scotland, 7-10 September 1997 (IOP Publishing),
pp. 47-52.

54.

Chen Q, Hoyle B S and Strangeways H J, Electrical field


interaction and an enhanced reconstruction algorithm in
capacitance process tomography, in Beck M S, Campogrande
E, Morris M, Williams R A and Waterfall R C (ed),
Tomographic Techniques for Process Design and Operation
(1993). Proceedings of the European Concerted Action on
Process Tomography, Manchester, UK, 26-29 March 1992
(Computational Mechanics Publications), pp. 205-212.
191

55.

Isaksen and Nordtvedt J E, A new reconstruction


algorithm for process tomography, Measurement Science
and Technology, 1993, 4, pp. 1464-1475.

56.

Nooralahiyan A Y, Hoyle B S and Bailey N J, Pattern


association and feature extraction in electrical
capacitance tomography, Proceedings of the European
Concerted Action on Process Tomography 94, Oporto,
Portugal, 24-26 March 1994, pp. 266-275.

57.

Duggan P M
and
York T A,
Tomographic
image
reconstruction
using
RAM-based
neural
networks,
Proceedings of the European Concerted Action on Process
Tomography 95, Bergen, Norway, 6-8 April 1995,
pp. 411-419.

58.

Nooralahiyan A Y, Hoyle B S and Bailey N J, Performance


of neural network with noise and parameter variation in
electrical capacitance tomography, Proceedings of the
European Concerted Action on Process Tomography 95,
Bergen, Norway, 6-8 April 1995, pp. 420-424.

59.

Isaksen , A rewiew of reconstruction techniques for


capacitance
tomography,
Measurement
Science
and
Technology, 1996, 7 (3), pp. 325-337.

60.

Chen T C, Analog and Digital Control System Design:


Transfer-function, State-space and Algebraic Methods,
Saunders College Publishing, 1993.

61.

Khan S H, Xie C G, and Abdullah F, Computer modelling of


process tomography sensors and systems, in Williams R A
and Beck M S (eds), Process Tomography - Principles
Techniques and Applications, Butterworth-Heinemann,
1995.

192

62.

Spink D M and Noras J M, Developments and improvements


in the solution of the forward problem in capacitance
and impedance tomography, Proceedings of the Engineering
Foundation Conference Frontiers in Industrial Process
Tomography II, Delft, The Netherlands, 9-12 April 1997,
pp. 205-210.

63.

Spink D
M, Direct finite-element
solution for
capacitance, conductance and inductance in static linear
problems, International Journal of Computation and
Mathematics in Electrical and Electronic Engineering,
1996, 15 (3), p. 70.

64.

Pinheiro P A T, Loh W W, Wang M, Mann R, Forrest A E and


Dickin F J, Three-dimensional electrical resistance
tomography in a stirred mixing vessel, Proceedings of
the Engineering Foundation Conference Frontiers in
Industrial
Process
Tomography
II,
Delft,
The
Netherlands, 9-12 April 1997, pp. 251-256.

65.

Millman J, Microelectronics: Digital and Analog Circuits


and Systems, International Student Edition, McGraw-Hill,
1979, p. 580

66.

Horowitz P and Hill W, The Art of Electronics, Second


Edition, Cambridge University Press, 1989, p. 224

193

A P P E N D I X

SYSTEM

A.1

DESIGN

A :

DETAILS

AND

CIRCUIT

DIAGRAMS

GENERAL

The system was built on an IEC 297-3 compatible sub-rack,


using 100 160 mm PCBs conforming to the Eurocard format.
Connections to the boards were done using mixed-contact
DIN-41612 connectors. The system architecture is depicted in
figure A.1. It consists of 12 transducer cards, a data
conversion card, a signal generator card and a bi-directional
buffer card (not shown in the figure for simplicity). Control
data is sent to the boards on a common control bus (which
includes strobe, address and control command data lines).
Communication with the host computer is realised via an
Amplicon PC14-AT plug-in digital I/O card. This card has two
82C55 programmable peripheral interface (PPI) chips with
three 8-bit ports each (PA, PB and PC), which can be
individually configured for input or output.
PA1 and PB1 were used to output a 16-bit control command data
word (although the 16-bits are not always needed). The upper
byte was output through PB1 and the lower byte through PA1.
PA2 and PB2 were used as inputs, to read the measurement data
(i.e., the ADC output). PB2 was used for the upper byte and
PA2 for the lower byte.
PC1 was used to output the 6-bit address data (b7 - b2), the
reset (RST) signal (b1), and the negated strobe (/STR) signal
(b0). PC2 was used as input to read the busy signal from the
ADC through bit 0.
194

Figure A.1

195

System

architecture

A.2

ADDRESS TABLE

The addresses were assigned as follows:

TABLE A.1







































ADDR
00
01
02
03
04
05
06
07




32
33
34
35
36
37
38

,

<

<

<

<

<

<

<

<

<



<

<

<

<

<

<

<

4

ADDRESS ASSIGNMENTS
CONTROL

DATA

EXCITATION REFERENCE DATA (DDS1)


EXC. REF. TRANSFER CONTROL REGISTER (DDS1)
DEMODULATION REFERENCE DATA (DDS2)
DEM. REF. TRANSFER CONTROL REGISTER (DDS2)
CHANNEL 1: EXCITATION LEVEL
CHANNEL 1: MODE (BIT 4) + RANGE (BITS 3 - 0)
CHANNEL 2: EXCITATION LEVEL
CHANNEL 2: MODE (BIT 4) + RANGE (BITS 3 - 0)




CHANNEL 15: EXCITATION LEVEL
CHANNEL 15: MODE (BIT 4) + RANGE (BITS 3 - 0)
CHANNEL 16: EXCITATION LEVEL
CHANNEL 16: MODE (BIT 4) + RANGE (BITS 3 - 0)
STANDING VALUE COMPENSATION
DC GAIN (BITS 7 - 4) + CHAN. SELECT (BITS 3 ANALOGUE TO DIGITAL CONVERTER

196



$

$

$

$

$

$

$

$

$



$

$

$

$

$

$
0) 
$



A.3

CONTROL BUS CONNECTOR PIN ASSIGNMENT

The pin assignment for the DIN-41612 connectors used to carry


the control signals to the boards is as follows:

TABLE A.2 CONTROL SIGNAL PIN ASSIGNMENTS


ON THE DIN-41612 CONNECTOR























12
13
14
15
16
17
18
19
20
21


A

<
 RST
<
 RST (15)
<
 A2 (4)
<
 A5 (7)
<
 D2 (10)
<
 D5 (14)
<
 D8 (18)
<
 D11 (21)
<

<
 D15 (25)
4

,

<

<

<

<

<

<

<

<

<

<

4

B
STR
STR (13)
A1 (3)
A4 (6)
D1 (9)
D4 (12)
D7 (17)
D10 (20)
D12 (22)
D14 (24)

,
C

<

<
 GND (1)
<
 A0 (2)
<
 A3 (5)
<
 D0 (8)
<
 D3 (11)
<
 D6 (16)
<
 D9 (19)
<

<
 D13 (23)
4



$

$

$

$

$

$

$

$

$

$



Note: The number in parenthesis indicates the corresponding


pin number on the DB-25 external connector used to
connect the data acquisition unit to the host computer.

197

A.4

CIRCUIT DIAGRAMS

The complete circuit diagrams of the multi-excitation ECT


transducer and the DDS dual sine-wave generator are included
next.

198

A P P E N D I X

B :

C O M P U T E R

P R O G R A M S

B.0

GENERAL

To avoid having an excessively large number of pages, and in


the interest of practicality, the actual programs themselves
are included in two accompanying 3" diskettes (1.44 MB, DOS
formatted), attached to the back cover. A brief explanation
of each program is given in the following sections. All the
programs were written in QuickBasic version 4.5.

B.1

ECT
SYSTEM
MONITORING
AND
CONTROL
PROGRAMS:
"IMG16LX.BAS", "IMG16LY.BAS" AND "MULTI5B.BAS"

This programs are used to control the ECT system electronics,


and to acquire and display capacitance ("IMG16LX.BAS" and
"IMG16LY.BAS")
or
electrode
charge
("MULTI5B.BAS")
measurement data. The excitation level, operating mode
(excitation or detection) and measuring range (AC gain) can
be individually set for each and every channel. Any specific
channel can be selected for analogue-to-digital conversion.
The offset (i.e. standing value) compensation and DC gain can
also be adjusted. Because of the large size of the array
required to store the sensitivity data, this programs have to
be
run
using
the
"/AH"
option.
"IMG16LX.BAS"
and
"IMG16LY.BAS" differ in the way the log command works, and
in that "IMG16LX.BAS" includes a routine to measure the
distinguishability matrix D, used in the optimal excitation
experiments.
201

Figure B.1

Basic flow diagram of the ECT system monitoring


and control program

The basic flow diagram of the program is shown in figure


The program starts with an initialization routine which
the value of a number of constants an initializes
variables used in the program. Among other things,
routine does the following tasks:

B.1.
sets
the
this

a) Reads the detector gain and zero-offset calibration


constants (see chapter 4, section 4.9.2) from the files
"GAINDETW.DAT" and "ZERODETW.DAT"
202

b) Sets the addresses and configures the I/O ports in the


Amplicon PC14-AT plug-in card used as interface between
the computer and the ECT system electronics. The PC14-AT
base address is set to 310 (hexadecimal).
c) Sets the excitation sources calibration constants, by
calling the subroutine "CALDACS", which contains the
corresponding values.
Following initialization, there are a pair of routines to
configure the DDS chips used to generate the excitation and
demodulation reference sine-waves. At this point the
frequency data is loaded onto the DDS chips.
Next, the main user interface is displayed on the computer
screen. It shows the status of all settings on each channel
and on the data conversion card, as well as the measured
data, which is displayed in both analogue and digital form.
At this stage, the program enters a keyboard-scanning and
data-acquisition loop, as shown in figure B.1. If one of the
several command keys is pressed by the user, the program
jumps to the proper subroutine, otherwise it just reads a new
measurement from the ADC. Commands are included to change any
one of the system settings, as well as for imaging (using the
single-electrode excitation method). A list of the available
commands is available on-line by pressing [ H ] to show the

help screen. The only command not listed is log data to


disk, activated by pressing [ L ].
In "IMG16LX.BAS", once data logging has been activated (while
in imaging mode), blocks of 66 capacitance measurements are
written to the file "LOG.DAT" until the imaging mode is
terminated. In "IMG16LY.BAS", on the other hand, when data
logging is activated the user is prompted for a test
identification number N, and a set of 66 10-sample-averaged
203

normalized capacitance measurements are acquired and written


to the file "NCAP-N.DAT". Normalized capacitance data
acquired in this way were used to test the feed-back
control based reconstruction algorithm (chapter 6).
Imaging can be carried out with (pressing [ I ]) or without
(pressing [ Q ]) initial sensor calibration. If imaging with
initial calibration is selected, then the program goes
through a sensor calibration procedure where Clow and Chigh are
measured, and the proper standing value compensation and gain
settings (AC and DC gains) are determined for each electrode
pair combination employed. If imaging without initial
calibration is desired, the calibration data and settings are
read from the file "IMGCAL1.DAT". The sensitivity map
information required in order to calculate the image values
using the LBP algorithm is read from the file "RMW.DAT". To
display the image, the program needs the coordinates of the
various pixels, which are read from the files "XCOOR.TXT" and
"YCOOR.TXT".

B.2

PROGRAMS USED IN THE OPTIMUM EXCITATION EXPERIMENTS

In the optimum excitation experiments described in chapter 5,


a modified version of the monitoring and control program was
used. In this modified version ("IMG16LX.BAS") there is a

test routine, invoked by pressing [ T ], which will ask for


a test identification number N and then measure the
capacitance matrices corresponding to both the empty sensor
and the permittivity distribution being tested, as well as
the distinguishability matrix D, and put them in the files
"CMTX0-N.DAT", "CMTX1-N.DAT" and "DMTX-N.DAT", respectively.
Using the svd MATLAB command, singular value decomposition
was applied to the matrix D in order to obtain the optimum
204

excitation (unit) vectors, which were written to the file


"BMTX-N.TXT".
Having determined the theoretical optimum excitation vectors,
another program, "PATTERNS.BAS", was used to find the actual
excitation voltage patterns to apply in practice, taking into
account the limitations of the real ECT system (excitation
and detection ranges). The program also determined the proper
system settings to use during the optimum excitation
measurements (AC gain, DC gain and compensation). The
excitation patterns along with their respective measurement
settings are put in the file "MLTDAT-N.TXT".
Finally, the charge measuring program "MULTI5B.BAS" was used
in order to apply the excitation voltage patterns and carry
out the corresponding electrode charge measurements, both for
the permittivity distribution being tested and for the empty
sensor. The routine that performs this procedure is invoked
by pressing [ U ]. It reads the excitation voltage data and
the system settings from the file "MLTDAT-N.TXT", applies the
patterns, and measures the charges. The resulting electrode
charge measurements with and without the test permittivity
distribution present are stored in the files "OPTIM1-N.TXT"
and "OPTIM0-N.TXT", respectively.

B.3

SIMULATION PROGRAMS

A number of investigations were realised using simulation


based on a finite-element (FE) model of a 12-electrode ECT
sensor. The simulations were carried out employing the FE
software package PC-OPERA, made by Vector Fields, Ltd.. The
sensor model geometry is shown in figure 3.3 (chapter 3). The
model information is contained in the files "CGR80.OP2" and
"CGR80.MES". There are 313 in-pipe regions (pixels).
205

B.3.1

PROGRAMS TO CALCULATE THE SENSITIVITY MAPS FOR SINGLEELECTRODE EXCITATION

To obtain the sensitivity maps for the single-electrode


excitation case, first of all, the model ("CGR80.OP2") was
modified by setting the potential of electrode one equal to
some excitation voltage, say 10 volts. The actual value is
not important, since it is eventually cancelled out in the
calculations. We then used the program "MAP4.BAS". Using this
program, PC-OPERA is run off-line in order to successively
simulate a change in permittivity in each one of the 313
in-pipe regions of the model. In each case, the relative
change in capacitance for each electrode pair is calculated
using the command input file "CGR80C1S.CMI" in the postprocessing stage. "MAP4.BAS" puts all the results in a long
file. The data needs to be separated into the various
sensitivity maps. This was done with "EXTRACT3.BAS". After
this, we have files containing all the sensitivity maps
associated projection 1. The sensitivity maps corresponding
to the remaining projections were obtained by rotation using
the program "SENSROT.BAS". The sensitivity map files are
named "DQJCI.DAT", where J is the (detection) electrode
number and I is the projection (i.e., the excitation
electrode) number. "GN4.DAT" contains a list of all the inpipe region numbers.
B.3.2

PROGRAMS TO CALCULATE THE


PARALLEL-FIELD EXCITATION

SENSITIVITY

MAPS

FOR

In the case of parallel-field excitation, first of all, we


set the electrode potentials to produce a parallel-field in
the direction of projection 1, using, for example, the
voltages given in table 3.1 (chapter 3). Then we used
"MAP4.BAS",
but
this
time
the
command
input
file
"CGR80C1.CMI" was utilized in the post-processing stage of
PC-OPERA. Next, we used "EXTRACT3.BAS" to separate the
206

results, obtaining 12 files with the sensitivity maps


associated with projection 1. Following this, the program
"SENSROT2.BAS" was used to generate, by rotation, the
sensitivity
maps
corresponding
to
the
remaining
5
projections, resulting in a set of 72 sensitivity map files
named "DQJPPI.DAT", where J is the (detection) electrode
number and I is the projection number.
B.3.3

PROGRAMS TO
MEASUREMENTS

SIMULATE

SINGLE-ELECTRODE-EXCITATION

Having previously set the desired permittivity distribution


in the sensor model ("CGR80.OP2"), the program "SIMREC2.BAS"
was employed in order to obtain simulated measurements using
single-electrode excitation. This program runs PC-OPERA offline 12 successive times, each time calculating the
measurements for a particular projection. In the preprocessing stage, PC-OPERA uses the command input file
"SPOTSEN.CMI", where N is the projection number, to set the
electrode voltages according to the projection being
simulated. Similarly, in the post-processing stage, the
command input file "GR80CNS.CMI" is used to calculate the
normalized measurements corresponding to each projection. At
the end of each cycle, a file of projection measurements is
created, named "PROJN.DAT". The final result is a set of 12
files containing the normalized measurement data for the same
number of projections.
Projection 12 is not actually required. Furthermore, half of
the 132 measurements contained in the remaining 11
projections are redundant, and are not needed for LBP image
reconstruction. The program "EXTRACT9.BAS" removes the
redundant data and leaves 11 projection files having each one
the right number of measurements.

207

B.3.4

PROGRAMS
TO
MEASUREMENTS

SIMULATE

PARALLEL-FIELD-EXCITATION

The
program
to
simulate
parallel-field-excitation
measurements, "SIMREC3.BAS",
is similar to the one just
described for single-electrode excitation. However, in this
case PC-OPERA is run only N = 6 times (once for each
projection), and the command input files used in the pre- and
post-processing stages are "ZPOT0N.CMI" and "CGR80CN.CMI",
respectively. The normalized measurements are stored in files
"PROJN.DAT". In these files, the measurements are mixed with
irrelevant information, which is removed using the program
"EXTRAC10.BAS".
B.3.5

PROGRAM TO PERFORM LBP IMAGE RECONSTRUCTION (USING


SINGLE-ELECTRODE EXCITATION)

The LBP algorithm, expressed by equation (2.12) in chapter 2,


can also be implemented as a simple multiplication of a
reconstruction matrix R times a (column) vector of normalized
measurements . We can form a matrix with the sensitivity
maps, by using them as the matrix columns. We take into
account the weighting factors (i.e., the denominator of
equation 2.12) by dividing each element in every row of the
matrix by the sum of all the elements in that row. The result
is the reconstruction matrix R (in our case 66 313).
Based on the sensitivity maps calculated using the programs
of section B.3.1, we produced a reconstruction matrix and
stored it in the file "RMW.DAT". The program "RECNS.BAS" uses
this matrix and the vector of normalized measurements
contained in the file "MEAS.DAT" to perform LBP image
reconstruction. It requires the coordinates of the image
pixels, which are stored in the files "XCOOR.TXT" and
"YCOOR.TXT". The image values are put in the file "IMAG.DAT".

208

B.3.6

ITERATIVE LBP IMAGE RECONSTRUCTION PROGRAM

"ITER2.BAS" is the iterative LBP image reconstruction program


written by the author to carry out an initial assessment of
the feed-back control algorithm (chapter 6). The vector of
measured normalized capacitances is stored in the file
"CM.DAT" and the initial image is in "IMG0.DAT". In the
procedure to obtain the calculated measurements, the PC-OPERA
command input files "VITERN.CMI" and "GR80INS.CMI" (where N
is the projection number) are used to set the electrode
voltages and calculate the electrode charges, respectively.
The file "FILYNG.CMI" is used as a template to generate the
command
input
file
which
updates
the
permittivity
distribution in the FE model. The resulting successive images
are put in the files "IMGI.DAT", where I is the iteration
number.

209

A P P E N D I X

C :

PAPERS PRODUCED AS A RESULT OF THIS WORK

1. Yang W Q, Xie C G, Gamio J C and Beck M S, Design of a


capacitance tomographic imaging sensor with uniform
electric field, Proceedings of the European Concerted
Action on Process Tomography 95, Bergen, Norway, 6-8
April 1995, pp. 266-276.
2. Yang W Q, Spink D M, Gamio J C and Beck M S, Sensitivity
distributions of capacitance tomography sensors with
parallel field excitation, Measurement Science and
Technology, 1997, 8 (5), pp. 562-569.
3. Gamio J C, Yang W Q, Waterfall R C and Beck M S, A high
signal-to-noise ratio capacitance transducer for use in
a multiple-excitation tomography system, Proceedings of
the Engineering Foundation Conference Frontiers in
Industrial Process Tomography II, Delft, The Netherlands,
9-12 April 1997, pp. 229-233.
4. Yang W Q, Gamio J C and Beck M S, A fast iterative image
reconstruction algorithm for capacitance tomography, in
Augousti A T and White N M (eds), Sensors and their
Applications VIII (1997). Proceedings of the Eighth
Conference on Sensors and their Applications, Glasgow,
Scotland, 7-10 September 1997 (IOP Publishing), pp. 47-52.
5. Gamio J C and Waterfall R C, An improved second-generation
electrical capacitance tomography system, in Augousti A
T and White N M (eds), Sensors and their Applications VIII
(1997). Proceedings of the Eighth Conference on Sensors
and their Applications, Glasgow, Scotland, 7-10 September
1997 (IOP Publishing), pp. 163-168.
210

Potrebbero piacerti anche