304
Magnetic Induction Tomography for Imaging Cerebral Stroke Massoud Zolgharni, BSc, MPhil A thesis submitted to Swansea University in fulfilment of the requirements for the Degree of Doctor of Philosophy Swansea University School of Medicine June 2010

Magnetic Induction Tomography for Imaging Cerebral Stroke

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Magnetic Induction Tomography for Imaging Cerebral Stroke

Magnetic Induction Tomography for

Imaging Cerebral Stroke

Massoud Zolgharni, BSc, MPhil

A thesis submitted to Swansea University in fulfilment of the

requirements for the Degree of

Doctor of Philosophy

Swansea University

School of Medicine

June 2010

Page 2: Magnetic Induction Tomography for Imaging Cerebral Stroke

ii

Summary

___________________________________

Magnetic induction tomography (MIT) is a relatively new contact-less soft-field

imaging modality which is used for imaging the internal distribution of the passive

electromagnetic properties of an object. The main purpose of this study is to

numerically investigate the feasibility of using MIT for imaging cerebral stroke in the

human brain. Stroke is a major cause of mortality and morbidity in Western society, and

conventional brain imaging techniques (e.g. MRI and CT) are expensive and,

sometimes, inaccessible for diagnosing stroke. It has been conjectured that MIT may

eventually be an attractive and low-cost alternative for rapid diagnostic imaging.

An anatomically realistic, multi-layer head geometry is used for the MIT simulations,

which is the first application of its kind in MIT. The finite element and finite difference

techniques are used for solving the forward problem and predicting the MIT signals. In

order to make the simulations more realistic, noise in the measuring hardware and two

examples of systematic errors are added to the simulated data. 3D images from the

conductivity distribution within the head are reconstructed by solving the ill-posed

inverse problem. To this end, a Tikhonov regularized minimization technique is used.

It is concluded that provided the noise in the hardware and the systematic errors in the

measuring protocol remain within a certain level, images of peripherally located strokes

may be reconstructed. However, apart from the unrealistic noise-free cases, imaging the

small volume strokes, located deep within the brain, is unfeasible. This problem may be

overcome by de-noising the MIT signals and employing other types of the inverse

solution techniques in the future. It is also shown that hemispherical MIT arrays

perform better than the conventional cylindrical arrays for imaging brain lesions.

Page 3: Magnetic Induction Tomography for Imaging Cerebral Stroke

iii

Declaration

This work has not previously been accepted in substance for any degree and is not being

concurrently submitted in candidature for any degree.

Signed .............................................................. (Massoud Zolgharni)

Date ................................................................

Statement 1

This thesis is the result of my own investigations, except where otherwise stated. Where

correction services have been used, the extent and nature of the correction is clearly

marked in a footnote(s).

Other sources are acknowledged by footnotes giving explicit references. A bibliography

is appended.

Signed .............................................................. (Massoud Zolgharni)

Date ................................................................

Statement 2

I hereby give consent for my thesis to be available for photocopying and for inter-

library loan, and for the title and summary to be made available to outside

organisations.

Signed .............................................................. (Massoud Zolgharni)

Date ................................................................

Page 4: Magnetic Induction Tomography for Imaging Cerebral Stroke

iv

To my parents

for their love and sacrifices...

Page 5: Magnetic Induction Tomography for Imaging Cerebral Stroke

v

Acknowledgements

___________________________________

First and foremost, I am deeply indebted to my experienced supervisors, Professor Huw

Griffiths from the School of Medicine and Dr. Paul D Ledger from the School of

Engineering, for their patience, guidance, advice, criticism, and insight throughout this

research.

I wish to thank all my colleagues at the South-Wales MIT group and the University of

Manchester with whom I had many fruitful discussions. I would also like to thank

Professor David S Holder for providing the UCL head mesh.

Critical reading and helpful suggestions by the PhD examiners, Professor Bill Lionheart

from the University of Manchester and Professor Kenneth Morgan from Swansea

University, is appreciated.

The financial support from the EPSRC and the ABM University Health Board is

gratefully acknowledged.

Special thanks to my friends with whom I shared my moments when there was no

forward or inverse problem to be solved.

My love to my family who have always encouraged and supported my efforts.

Finally, my deepest gratitude to the one, for the provision of joys, challenges, and grace

for growth.

Page 6: Magnetic Induction Tomography for Imaging Cerebral Stroke

vi

Contents

___________________________________

Part I: Preliminaries

Chapter 1. Introduction .......................................................................................... 1

1.1. MIT ......................................................................................................................... 1

1.2. MIT applications ..................................................................................................... 3

1.3. Motivation and objectives ...................................................................................... 3

1.4. Outline of the thesis ............................................................................................... 5

1.5. Contributions to knowledge ................................................................................... 6

1.6. LCOMIT ................................................................................................................... 8

Chapter 2. Principles of MIT .................................................................................... 9

2.1. Introduction ............................................................................................................ 9

2.2. Physical principles................................................................................................... 9

2.2.1. MIT problem ..................................................................................................................11

2.2.2. Eddy-current problem ....................................................................................................12

2.2.3. Solving the (forward) problem........................................................................................14

2.3. Working principles .................................................................................................14

2.3.1. The MIT signal ................................................................................................................16

2.3.2. SNR and operational frequency ......................................................................................17

2.4. MIT studies for brain imaging ................................................................................18

2.5. MK1 and MK2a systems .........................................................................................22

Chapter 3. Anatomy of Head and Stroke Pathology ............................................. 23

3.1. Introduction ...........................................................................................................23

3.2. Head anatomy .......................................................................................................23

3.2.1. Scalp ..............................................................................................................................24

3.2.2. Muscles .........................................................................................................................24

3.2.3. Skull ...............................................................................................................................25

3.2.4. Meninges .......................................................................................................................25

3.2.5. Brain ..............................................................................................................................27

3.2.6. Ventricles and cerebrospinal fluid ..................................................................................30

3.2.7. Spinal cord .....................................................................................................................30

3.2.8. Eyes and optic nerves .....................................................................................................31

3.2.9. Cavities ..........................................................................................................................32

3.2.10. Other tissues .............................................................................................................32

3.3. Stroke ....................................................................................................................33

Page 7: Magnetic Induction Tomography for Imaging Cerebral Stroke

vii

3.3.1. Classification of stroke ...................................................................................................33

3.3.2. Diagnosis and treatment ................................................................................................35

Chapter 4. Dielectric Properties of Head Tissues and Pathophysiology ................ 39

4.1. Introduction ...........................................................................................................39

4.2. Dielectric properties of head tissues ......................................................................40

4.2.1. Literature before 1996 ...................................................................................................41

4.2.2. The 1996 database .........................................................................................................41

4.2.3. Literature after 1996 ......................................................................................................43

4.3. Pathophysiology and conductivity change .............................................................59

4.4. Can MIT be applied for stroke imaging? ................................................................62

4.5. Adopted values for simulations and discussion .....................................................64

Part II: Discretized Forward and Inverse Problems

Chapter 5. Discrete Forward Problem ................................................................... 68

5.1. Introduction ...........................................................................................................68

5.1.1. BVP simplification ..........................................................................................................69

5.2. Finite element method ..........................................................................................71

5.2.1. FE discretization .............................................................................................................72

5.2.2. Comsol Multiphysics ......................................................................................................73

5.2.3. Model construction ........................................................................................................74

5.2.4. Solution technique .........................................................................................................76

5.3. Finite difference method .......................................................................................77

5.3.1. BVP ................................................................................................................................77

5.3.2. FD discretization ............................................................................................................79

5.3.3. Solution technique .........................................................................................................80

5.3.4. FD solver ........................................................................................................................81

5.4. Benchmark test ......................................................................................................82

5.4.1. FE vs. FD vs. Analytical....................................................................................................85

5.5. Screen implementation in FD model ......................................................................88

5.5.1. Screen for MK1 and MK2a systems.................................................................................89

5.5.2. Simulated vs. measured data .........................................................................................93

5.6. Discussion ..............................................................................................................96

Chapter 6. Discrete Inverse Problem ..................................................................... 99

6.1. Introduction ...........................................................................................................99

6.2. Time-difference imaging ......................................................................................101

6.2.1. Non-linearity ................................................................................................................ 101

6.2.2. Underdetermined ........................................................................................................ 103

6.2.3. Ill-posedness ................................................................................................................ 105

6.3. Sensitivity analysis ...............................................................................................106

Page 8: Magnetic Induction Tomography for Imaging Cerebral Stroke

viii

6.3.1. Sensitivity computation................................................................................................ 106

6.3.2. Ill-conditioning ............................................................................................................. 112

6.4. Truncated singular value decomposition .............................................................115

6.5. Tikhonov regularization .......................................................................................119

6.5.1. Standard Tikhonov ....................................................................................................... 119

6.5.2. Generalized Tikhonov................................................................................................... 122

6.5.3. Neighbouring matrix .................................................................................................... 123

6.6. Error estimation and regularization parameter ...................................................126

6.6.1. GCV ............................................................................................................................. 127

6.6.2. L-curve ......................................................................................................................... 129

6.7. Frequency-difference imaging .............................................................................131

6.8. Absolute imaging .................................................................................................134

6.8.1. Iterative non-linear solution ......................................................................................... 134

6.8.2. Newton-type methods ................................................................................................. 137

6.8.3. Absolute imaging using phase changes ......................................................................... 138

6.8.4. Three-layer geometry .................................................................................................. 147

6.8.5. Edge-preserving ........................................................................................................... 148

6.9. Imaging centrally located features.......................................................................150

6.9.1. Differential imaging...................................................................................................... 150

6.9.2. Absolute imaging ......................................................................................................... 152

6.10. Other iterative methods ......................................................................................153

6.11. Inverse crime .......................................................................................................154

6.12. Discussion ............................................................................................................155

Part III: Stroke Imaging

Chapter 7. Simulated Stroke Imaging Using the MK1 System ............................ 158

7.1. Introduction .........................................................................................................158

7.2. Head model .........................................................................................................158

7.2.1. UCL head mesh ............................................................................................................ 158

7.2.2. Stroke domains ............................................................................................................ 160

7.2.3. FE mesh construction ................................................................................................... 160

7.2.4. FD mesh construction .................................................................................................. 163

7.2.5. Dielectric properties..................................................................................................... 163

7.3. Forward problem .................................................................................................164

7.3.1. FE mesh convergence................................................................................................... 164

7.3.2. MIT signals (FE results) ................................................................................................. 167

7.3.3. FD vs. FE ...................................................................................................................... 169

7.4. Time-difference imaging ......................................................................................172

7.5. Frequency-difference imaging .............................................................................182

7.6. Specific absorption rate .......................................................................................184

Page 9: Magnetic Induction Tomography for Imaging Cerebral Stroke

ix

7.7. Discussion ............................................................................................................187

Chapter 8. Customized Coil Arrays, Systematic Errors, and the MK2b System.... 189

8.1. Introduction .........................................................................................................189

8.2. Forward and inverse problems ............................................................................191

8.3. Hemispherical and annular arrays .......................................................................192

8.3.1. Hemispherical (Helmet) array ....................................................................................... 192

8.3.2. Annular array ............................................................................................................... 193

8.3.3. Forward problem ......................................................................................................... 193

8.3.4. Sensitivity analysis ....................................................................................................... 194

8.3.5. Image display ............................................................................................................... 196

8.3.6. Image reconstruction ................................................................................................... 197

8.4. Simulated errors ..................................................................................................203

8.4.1. Phase noise .................................................................................................................. 203

8.4.2. Systematic errors ......................................................................................................... 205

8.5. MK2b system .......................................................................................................210

8.5.1. Concept of MK2b array ................................................................................................ 210

8.5.2. Dimensions .................................................................................................................. 211

8.5.3. Number of coils............................................................................................................ 212

8.5.4. Other coil arrangements and quality measurement ...................................................... 214

8.5.5. MK2b construction....................................................................................................... 219

8.6. Discussion ............................................................................................................220

Part IV: Experiments and Conclusions

Chapter 9. Experiments ...................................................................................... 224

9.1. Introduction .........................................................................................................224

9.2. Experimental setup ..............................................................................................224

9.3. Simulations ..........................................................................................................227

9.4. Signal analysis ......................................................................................................227

9.5. Image reconstruction ...........................................................................................230

9.5.1. Differential imaging...................................................................................................... 231

9.5.2. Absolute imaging ......................................................................................................... 234

9.6. Further discussion ................................................................................................235

9.6.1. Discretization error ...................................................................................................... 237

9.6.2. Phase noise .................................................................................................................. 238

9.6.3. Conductivity value error ............................................................................................... 238

9.6.4. Systematic errors ......................................................................................................... 240

Chapter 10. Conclusions and Future Work ............................................................ 243

10.1. Introduction .........................................................................................................243

10.2. Remarks and conclusions .....................................................................................243

Page 10: Magnetic Induction Tomography for Imaging Cerebral Stroke

x

10.2.1. Chapter 1 ................................................................................................................ 243

10.2.2. Chapter 2 ................................................................................................................ 243

10.2.3. Chapter 3 ................................................................................................................ 244

10.2.4. Chapter 4 ................................................................................................................ 244

10.2.5. Chapter 5 ................................................................................................................ 244

10.2.6. Chapter 6 ................................................................................................................ 244

10.2.7. Chapter 7 ................................................................................................................ 245

10.2.8. Chapter 8 ................................................................................................................ 246

10.2.9. Chapter 9 ................................................................................................................ 247

10.3. Discussion and future work .................................................................................247

10.3.1. Immediate follow up research ................................................................................. 247

10.3.2. Forward and inverse problems ................................................................................ 249

10.3.3. Practical measurements and systematic errors ........................................................ 250

10.3.4. Computational vs. Measurement Accuracy .............................................................. 251

10.3.5. Sensor array design ................................................................................................. 253

10.3.6. Image resolution ..................................................................................................... 254

Part V: Appendices and Bibliography

Appendix A. System Description ............................................................................... 257

A.1. MK1 system ..............................................................................................................257

A.2. MK2a system.............................................................................................................258

Appendix B. Dielectric Theory and Terminology ....................................................... 260

Appendix C. Supplements to chapter 6 ..................................................................... 265

C.1. GCV method ..............................................................................................................265

C.2. Steepest descent .......................................................................................................267

C.3. Landweber method ...................................................................................................267

C.4. Conjugate gradient method.......................................................................................268

C.5. ART and SIRT .............................................................................................................268

C.6. Noise .........................................................................................................................269

References 270

Page 11: Magnetic Induction Tomography for Imaging Cerebral Stroke

xi

List of tables

___________________________________

Table 3.1. Performance of CT and MRI for diagnosing stroke; values in percentage [83]. .......................36

Table 4.1. Weighted average dielectric properties assigned to heterogeneous or minor organs in the

head (frequency of 10 MHz).........................................................................................................65

Table 6.1. Conductivities (in S m-1) of white matter and blood at 1 and 10 MHz, and conductivity change

between two frequencies. ......................................................................................................... 133

Table 7.1. Conductivity (S m-1) and relative permittivity values assigned to each tissue type at the

frequencies of 1 and 10 MHz. The mass density of tissues (kg m-3), adopted from [205], are to be

used for the power absorption calculations later in section 7.6. a filled with air. ......................... 164

Table 7.2. Series of meshes for the convergence tests; NS: not solved due to memory restrictions. ..... 165

Table 7.3. Series of FD meshes for the convergence tests. ................................................................... 170

Page 12: Magnetic Induction Tomography for Imaging Cerebral Stroke

xii

List of illustrations

___________________________________

Figure 2.1. Schematic view of boundary value problem in MIT; C: conductive domain, N: non-

conducting region (free space). ....................................................................................................12

Figure 2.2. Block diagram of a typical MIT system (after Peyton et al [30]). ............................................14

Figure 2.3. Phasor diagram representing primary (Vp) and secondary ( V) voltages. ..............................16

Figure 3.1. Skeletal muscles: (a) structure of a skeletal muscle fibre (after Saladin [66]), (b) muscles of

head and neck (after Mader [65]). ...............................................................................................24

Figure 3.2. (a) simplified side view of human skull (after Villarreal [67]). (b) illustration of a typical long

bone showing the location of cortical and cancellous bones (picture adapted from [68]); skull

bones have a similar internal structure. .......................................................................................25

Figure 3.3. Meninges; (a) meningeal membranes surrounding brain, (b) frontal section of head to show

the meninges (after Seely [71]). ...................................................................................................26

Figure 3.4. Human brain; cerebrum, seen here in longitudinal section, is the largest part of brain in

humans (after Mader [65])...........................................................................................................27

Figure 3.5. Cerebrum: (a) lateral view of right hemisphere with insula exposed; colours in this figure

distinguish the lobes of cerebral hemispheres (after Shier et al [72]), (b) coronal section of brain

showing cortex, white matter, and corpus callosum (after Scanlon and Sanders [73]). ..................28

Figure 3.6. Illustration of the brainstem, diencephalon and cerebellum (after Shier et al [72]). ..............29

Figure 3.7. Spinal cord: (a) illustration of entire length, (b) drawing of a segment (after Seely [71])........31

Figure 3.8. Illustration of eye balls and optic nerves (after Mader [65]). .................................................31

Figure 3.9. Air-filled cavities in head; (a) nasal cavity, (b) ear canal (after Mader [65]). ...........................32

Figure 3.10. Stroke; ischaemia due to embolism (a) and thrombosis (b), haemorrhage due to rupture of a

blood vessel (c) (after Broderick and Santilli [77]). ........................................................................34

Figure 3.11. Diagram showing different types of haemorrhage based on its location within head. .........34

Figure 3.12. Examples of intra- and extra-axial haemorrhages (after Caplan [80]). .................................35

Figure 3.13. Stroke images: (a) CT scan showing a large cerebral infarct, (b) MRI scan showing a left

cerebral infarct, (c) CT scan showing a haemorrhage in deep portion of brain, (d) MRI scan showing

a brain haemorrhage similar to that in (c) (after Caplan [79]). ......................................................37

Figure 3.14. CT scan of a haemorrhage worsened by thrombolytic therapy (after Harold and Adams [85]).

....................................................................................................................................................38

Figure 4.1. Dielectric properties of muscle against frequency; longitudinal and transverse directions. ...52

Figure 4.2. Dielectric properties of grey matter against frequency. ........................................................53

Figure 4.3. Dielectric properties of white matter against frequency. ......................................................54

Figure 4.4. Dielectric properties of cancellous bone against frequency. .................................................55

Figure 4.5. Dielectric properties of cortical bone against frequency. ......................................................56

Figure 4.6. Dielectric properties of CSF against frequency. .....................................................................57

Figure 4.7. Dielectric properties of blood against frequency. .................................................................58

Figure 4.8. Current flow through cells at both low and high frequencies. Due to ischaemia and

subsequent oedema, brain cells swell resulting in impedance increase at low frequencies. ECS:

extra-cellular space (low-frequency and high-frequency diagrams are redrawn from Holder [116]

and Holder [118], respectively). ...................................................................................................59

Figure 4.9. Current flow through cells. Due to haemorrhage, increase in conductive blood volume results

in impedance decrease. BV: blood volume (diagram redrawn from Holder [118]). ........................61

Page 13: Magnetic Induction Tomography for Imaging Cerebral Stroke

xiii

Figure 4.10. CT scan showing haemorrhage growth in a patient with basal ganglion intra-cerebral

haemorrhage (after Adeoye and Broderick [122]). .......................................................................62

Figure 4.11. Dielectric properties of head tissues used for simulations. .................................................66

Figure 5.1. A 2D representation of the grid showing the electrical connections between a node and its 4

neighbours in the grid (6 neighbours in 3D). The admittances Y and vector voltage generator

components e are known [153]. ...................................................................................................79

Figure 5.2. Exciter coil and conductive sphere. ......................................................................................82

Figure 5.3. Typical FE meshes generated for benchmark test; a coarse (a), and a fine (b) mesh. Note that

the presence of coil, sphere, and layer of infinite elements (separate subdomains) may result in

generation of non-uniform meshes, e.g. in (a). .............................................................................83

Figure 5.4. Error in the numerically computed eddy current inside the conducting sphere versus the

number of degrees of freedom; results for first- and second-order elements plotted in broken red

and solid blue lines, respectively. .................................................................................................84

Figure 5.5. Cross section of sphere showing r. .......................................................................................85

Figure 5.6. Plots of current density |J| along r for = 1 S m-1 (b) and = 100 S m-1 (c). ..........................86

Figure 5.7. FE results: real, imaginary, and norm of current density along r for = 1 S m-1 (left) and =

100 S m

-1 (right). ..........................................................................................................................87

Figure 5.8. Maps of induced current densities in A m

-2 on a cross-section of sphere for FE results when

= 1 and 100 S m-1. ........................................................................................................................87

Figure 5.9. FE models for computing A-field for MK1 and MK2a systems. ..............................................90

Figure 5.10. Plots of |A| for 4 meshes along a test line. .........................................................................91

Figure 5.11. Induced secondary voltage at 16 sensor coils for Ex1 activated (FE mesh comprised 108,811

tetrahedral elements). .................................................................................................................92

Figure 5.12. Plan view of the MIT array showing the position of the sample and coils within the

electromagnetic screen. Red (outer) and black (inner) coils are exciters and sensors, respectively.

....................................................................................................................................................93

Figure 5.13. Values of Im( V/Vp) in percentage against conductivity: (a) FE and FD results, (b) measured

data where straight fitted line is also shown (after Griffiths et al [8]). ...........................................94

Figure 5.14. Values of Re( V/Vp) in percentage against conductivity: (a) FE and FD results, (b) measured

data (after Griffiths et al [8]). .......................................................................................................95

Figure 5.15. Total values of conduction and displacement current densities induced throughout the

cylindrical vessel. .........................................................................................................................96

Figure 6.1. One geometrical example of the model (to scale): (a) diagram of the 3D model, (b) x-y and y-z

cross-sectional planes for image representations. ...................................................................... 100

Figure 6.2. Diagram illustrating the conductivity distributions for two states in x-y planes. .................. 101

Figure 6.3. Diagram illustrating the conductivity distribution for two inhomogeneous states. .............. 103

Figure 6.4. Perturbation method for sensitivity calculation; highlighted voxels represent the perturbed

elements (pixels). ...................................................................................................................... 106

Figure 6.5. Comparison of sensitivity matrices computed based on perturbation and E.E methods.

Percentage values correspond to conductivity change in the perturbed voxel with respect to the

background. .............................................................................................................................. 109

Figure 6.6. Comparison of non-linear and linear models; is contrast between added inclusion and

background. .............................................................................................................................. 111

Figure 6.7. Comparison of non-linear and linear models; is contrast between added inclusion and

background. .............................................................................................................................. 111

Figure 6.8. Singular spectrum plotted on a logarithmic scale, normalized to the largest singular value.

Also shown some arbitrary singular vector maps; index = 1, 3, 60, 136, 223. .............................. 114

Figure 6.9. Levels of truncation of singular values. ............................................................................... 116

Page 14: Magnetic Induction Tomography for Imaging Cerebral Stroke

xiv

Figure 6.10. Reconstructed images using TSVD; (a) columns from left: target and images with different

levels of truncation, rows from top: x-y and y-z cross-sectional planes, colour-map shows

conductivity change in S m-1

, (b) 1D representation of target and reconstructed conductivity

changes along a line in x-y plane. ............................................................................................... 117

Figure 6.11. As figure 6.10, but with 60 dB noise added to simulated data. .......................................... 119

Figure 6.12. Modified singular values (si + /si) for 3 different levels of regularization. ......................... 121

Figure 6.13. Reconstructed images using standard Tikhonov regularization with 60 dB noise added to

simulated data. Columns from left: target and images with different levels of regularization

indicated by value of . Rows from top: x-y and y-z cross-sectional planes. Colour-map shows

conductivity change in S m-1. ...................................................................................................... 122

Figure 6.14. Correlation between neighbouring elements with common face (a), common node (b), and

a combination of faces and edges (c) in a 2D image. Values shown in voxels correspond to a central

finite difference approximation of Laplacian operator. ............................................................... 123

Figure 6.15. As in figure 6.13, but reconstructed using (6.25) with R-face and xref = 0. .......................... 124

Figure 6.16. Comparison of standard and generalized Tikhonov: 1D plots showing reconstructed

conductivity changes ( = 1 10-18) along cross-sectional lines in y-direction (a) and x-direction (b).

.................................................................................................................................................. 125

Figure 6.17. Regularization parameter selection using GCV method for three noise levels. .................. 128

Figure 6.18. Reconstructed images by standard Tikhonov method and using suggested value of by GCV

method. Colour-bar in S m-1. ...................................................................................................... 129

Figure 6.19. Plots of L-curve for selecting the regularization parameter for three noise levels. ............. 130

Figure 6.20. Regularization parameter selected by GCV (a) and L-curve (b) methods for three noise levels

(40, 50, and 60 dB). .................................................................................................................... 131

Figure 6.21. Frequency-difference imaging; (a): target, (b) and (c): reconstructed images using

generalized Tikhonov method (1 m added phase noise, R-face, xref = 0, = 1 10-9), but plotted

with different colour-bars in S m-1. ............................................................................................. 133

Figure 6.22. Flowchart for iterative non-linear MIT image reconstruction. ........................................... 136

Figure 6.23. Simulated target; coil array is placed at three levels along height of cylindrical tank. Colour-

bar in S m-1................................................................................................................................. 138

Figure 6.24. Absolute imaging using (6.39) with R-face, 0 = 1 S m-1

, l = 0.5 and 1 = 1 10-8

. First row:

plots of ||rk||2 and ||ek||2 for iteration procedure. Second row: same plots as in first row but

magnified at iteration 52. Also shown reconstructed images at iterations 52, 53, and 54. Upper

limit of colour-bars (S m-1) is set to maximum conductivity value for each image. ....................... 140

Figure 6.25. Absolute imaging using (6.39) with R-face, 0 = 1 S m-1, l = 0.5 and 1 = 1.8 10-20. Errors

||rk||2 and ||ek||

2 are plotted against iterations. Reconstructed images correspond to iterations

1, 2, and 8. Colour-bar (S m-1) limits are set to maximum conductivity for each image. ............... 142

Figure 6.26. Absolute imaging using (6.39) with R-face, 0 = 1 S m-1, l = 0.1 and 1 = 1 10-8. Errors ||rk||2

and ||ek||2 are plotted against iterations. Reconstructed images correspond to iterations 10, 18,

and 19. Colour-bar (S m-1) limits are set to maximum conductivity value in each image. ............. 143

Figure 6.27. Plots of ||rk||2 and ||ek||2 for absolute imaging using (6.39) with R-face, 0 = 1 S m-1, l = 0.9

and 1 = 1 10-8. ......................................................................................................................... 144

Figure 6.28. Absolute imaging using (6.39) with R-face, 0 = 1 S m-1, l = 0.5 and 1 = 1 10-8. with 1 m

additive phase noise. Errors ||rk||2 and ||ek||2 are plotted against iterations. Reconstructed

images correspond to iterations 7, 10, and 14. Colour-bar (S m-1) limits are set to maximum

conductivity value in each image................................................................................................ 145

Figure 6.29. Absolute imaging using (6.39) with R-face, 0 = 1 S m-1

, l = 0.5 and 1 = 1 10-8

. with 10 m

additive noise. Errors ||rk||2 and ||ek||2 are plotted against iterations. Reconstructed images

correspond to iterations 3, 6, and 7. Colour-bar (S m

-1) limits are set to maximum conductivity

value in each image. .................................................................................................................. 146

Page 15: Magnetic Induction Tomography for Imaging Cerebral Stroke

xv

Figure 6.30. Absolute imaging of a 3-layer model including CSF, white matter, and blood. Colour-bar in S

m-1............................................................................................................................................. 147

Figure 6.31. Surface plots of x-y plane in images for noise free (upper row) and 5 m noise (lower row).

Left column is for images reconstructed using R-face, and right column is for edge-preserving

regularization matrix.................................................................................................................. 149

Figure 6.32. Time-difference imaging. Left: target, Middle and Right: reconstructed images using

generalized Tikhonov method (R-face, xref = 0) and suggested by L-curve and GCV methods,

respectively. Colour-bars in S m

-1. .............................................................................................. 151

Figure 6.33. Frequency-difference imaging. Left: target, Middle and Right: reconstructed images using

generalized Tikhonov method (R-face, xref

=

0) and different values. Colour-bars in S

m

-1. ......... 151

Figure 6.34. Absolute imaging using (6.39) with R-face, 0 = 1 S m-1

, l = 0.5 and 1 = 1 10-8

, using noise-

free data and with 1 m additive noise. Colour-bar (S m

-1) limits are set to maximum conductivity

value in each image. .................................................................................................................. 152

Figure 7.1. UCL head mesh, triangular surface elements of skin and internal compartments. ............... 159

Figure 7.2. Transverse planes showing internal tissues separated by different colours. ........................ 159

Figure 7.3. Stroke regions in the right hemisphere of brain; side view represents the vertical level of the

LP which was similar for all three strokes. .................................................................................. 160

Figure 7.4. FE mesh construction; details given in text. ........................................................................ 162

Figure 7.5. Level of coil array relative to head and LP stroke (red = exciter, blue = sensor). .................. 162

Figure 7.6. FD mesh with two cuts illustrating the internal structure, including the LP stroke. .............. 163

Figure 7.7. Convergence of the induced voltage, for activated Ex1, in sensors Se6, Se9, and Se13 for first-

and second-order elements plotted in broken red and solid blue lines, respectively. .................. 166

Figure 7.8. Typical results. (a) Projection on the transverse plane of magnetic field lines with Ex13

activated with a current of 1 A. The colour of a field line indicates the magnitude of the magnetic

field at that point; the colour-bar indicates the field strength in A m−1. (b) Map of induced current

density (A m−2) in a transverse slice through the normal brain, 25 mm above the centre of the coils.

(c) as (b) but with the LP stroke present. .................................................................................... 166

Figure 7.9. Profiles of for three strokes, simulated at frequency of 10 MHz and with Ex1 activated.

Also shown is the phase angle for normal head relative to empty space (values are divided by 20

for convenience of display). ....................................................................................................... 167

Figure 7.10. The 256 values of magnitude of sorted in descending order for each of three strokes,

simulated at frequency of 10 MHz. ............................................................................................ 168

Figure 7.11. The 256 values of magnitude of sorted in descending order for LP stroke with different

blood percentages. .................................................................................................................... 169

Figure 7.12. Secondary voltage in 16 sensors for Ex1 activated, in normal head at frequency of 10 MHz.

Numbers 7.5, 6, 5, 4, 3, and 2.5 indicate grid sizes used in FD model. Converged FE results are

plotted in broken line for comparison. ....................................................................................... 171

Figure 7.13. Discrepancy between FE and FD (2.5 mm grid size) results in figure 7.12. ......................... 171

Figure 7.14. Regularization parameter selection using GCV method for three phase noise levels (0, 1 m ,

17 m ) added to phase changes for LP stroke. ............................................................................ 174

Figure 7.15. Columns left to right: simulated LP target, and reconstructed images with 0, 1 m , 17 m

additive noise. Rows from top to bottom: transverse, coronal, and sagittal cross-section planes.

Colour-bars in S m

-1. is selected by GCV method. ..................................................................... 174

Figure 7.16. Regularization parameter selection using L-curve method for three phase noise levels (0, 1

m , 17 m ) added to phase changes for LP stroke....................................................................... 176

Figure 7.17. As figure 7.15, but selected by L-curve method. ............................................................ 176

Figure 7.18. Reconstructed images of LP stroke with 1 m phase noise added to phase changes, and for

different values of . Upper limit in each colour-bar (S m-1) is set to maximum value in the

corresponding image. ................................................................................................................ 177

Page 16: Magnetic Induction Tomography for Imaging Cerebral Stroke

xvi

Figure 7.19. As figure 7.15, but selected empirically. ........................................................................ 179

Figure 7.20. As figure 7.19, but for SP stroke. ...................................................................................... 180

Figure 7.21. As figure 7.19, but for SD stroke. ...................................................................................... 180

Figure 7.22. (a): profiles of noise-free for SD stroke (solid blue line) and those with 1 m added phase

noise (broken red line). (b): difference between two plots in (a). ................................................ 181

Figure 7.23. Regularization parameter selection using GCV and L-curve method for noise-free frequency-

difference data for LP stroke. ..................................................................................................... 182

Figure 7.24. Columns left to right: simulated LP target, and reconstructed images from noise-free data

with selected using GCV and L-curve methods. Rows from top to bottom: transverse, coronal,

and sagittal cross-section planes. Colour-bars in S m

-1. ............................................................... 183

Figure 7.25. SAR distribution in normal head for Ex1 activated at 10 MHz; (a) skin surface, (b) slice at

through centres of coils. Colour-bar indicates logarithm to the base 10 of value in W kg-1........... 185

Figure 8.1. Helmet MIT arrays: (a) side view of hemispheres H100 and H120 in position over head, (b)

coil positions on hemisphere H100, magnified in relation to (a). Angles of elevation of coil centres

in each ring are indicated with respect to lowest ring. Each of first three rows from bottom

includes 16 coils, and the upper row includes 8 coils. Exciters and sensors are shown as red and

black coils, respectively. Origin of coordinate system is defined as centre of helmet. .................. 192

Figure 8.2. Annular MIT array. Three levels of array for each plane of measurement are indicated as Z1,

Z2, and Z3 which are level with first lower rows of H100 (see figure 8.1b). Two pairs of coils for

each plane of measurement are shown. Exciters and sensors are shown as red and black coils,

respectively. Coordinate system is that in Helmet arrays. ........................................................... 193

Figure 8.3. Singular spectrum for three MIT arrays, normalised to the largest singular value and plotted

on a logarithmic scale. ............................................................................................................... 194

Figure 8.4. Total sensitivity plots for three MIT arrays. Rows from top to bottom: transverse, coronal,

and sagittal cross-sections. Colour-bar in radian.m.S-1. ............................................................... 195

Figure 8.5. A reconstructed image of LP stroke, shown with different colour-bars in S m-1. The values

given above each image are upper limits in the colour-bars. ...................................................... 196

Figure 8.6. Integration volumes, shown as blue regions, for computing Ehead, Ebrain, and Eupper. ............. 197

Figure 8.7. Error Ehead in reconstructed images plotted in percentage as a function of , for three MIT

arrays. ....................................................................................................................................... 198

Figure 8.8. Columns left to right: simulated LP target, and reconstructed images from phase changes

with 1 m added noise for annular (three planes), H100, H120, and annular (with one plane) MIT

arrays; was optimised based on Ehead. Maximum conductivity change occurred in each images is

given as ( )max. Rows from top to bottom: transverse, coronal, and sagittal planes. Colour-bar in

S m

-1. ......................................................................................................................................... 199

Figure 8.9. As figure 8.7, but for Eupper (a) and Ehead (b). ........................................................................ 200

Figure 8.10. As figure 8.8, but reconstructed using values of minimising Eupper. Also fifth column from

left shows image reconstructed for H100 and using a high level of regularization. ...................... 201

Figure 8.11. Columns left to right: simulated SP target, reconstructed SP image, simulated SD target,

reconstructed SD image, simulated normal brain, reconstructed normal image. Rows from top to

bottom: transverse, coronal, and sagittal cross-sections. Images reconstructed using phase

changes with 1 m added noise to H100 data, and optimised for Ebrain. Colour-bar in S m

-1. ...... 202

Figure 8.12. Value of error, Enoise, in reconstructed images using H100 (λ = 2.5x10-9) for different levels of

added Gaussian phase noise (standard deviation). Three simulations (different colours) for each

noise level are shown along with the fitted straight line. ............................................................ 204

Figure 8.13. Columns left to right: simulated LP target, and reconstructed images for H100 with different

levels of added phase noise ( = 2.5 10-9). Rows from top to bottom: transverse, coronal, and

sagittal cross-sections. Colour-bar in S m

-1. ................................................................................. 204

Page 17: Magnetic Induction Tomography for Imaging Cerebral Stroke

xvii

Figure 8.14. Exaggerated geometrical errors: light pink and dark blue colours show the 'true' and

'erroneous' boundaries, respectively. (a) displacement in y-direction. (b) contraction. ............... 206

Figure 8.15. Columns left to right: simulated LP target, and reconstructed images with different

boundary errors in x-direction. Rows from top to bottom: transverse, coronal, and sagittal planes.

.................................................................................................................................................. 207

Figure 8.16. As figure 8.15, but for y- and z-direction displacements.................................................... 207

Figure 8.17. As figure 8.15, but for contraction and expansion scaling errors. ...................................... 208

Figure 8.18. Image error, Eboundary, for different displacements of head or percentage scaling. ............. 209

Figure 8.19. Conceptual diagram of MK2b system (not to scale). ......................................................... 211

Figure 8.20. Standard measurement indexes for head (after ARTUR [218]). ......................................... 211

Figure 8.21. Top view of three coil arrays; red and black coils are exciters and sensor coils, respectively.

.................................................................................................................................................. 212

Figure 8.22. Columns left to right: simulated LP target, and reconstructed images for 30-coil, 46-coil, 60-

coil, and 46-coil+screen arrays; reconstructed with 1 m noise, and optimised for Ehead. Rows

from top to bottom: transverse, coronal, and sagittal cross-sections. Colour-bar in Sm-1

. ........... 213

Figure 8.23. Examples of separated coil arrangements favoured from the point of view of the electronics

layout; red coils are exciters....................................................................................................... 215

Figure 8.24. Columns left to right: simulated LP target, and images reconstructed from noise-free data

and using different values of for asymmetric arrangement shown in figure 8.23 (left array). Rows

from top to bottom: transverse, coronal, and sagittal cross-sections. Colour-bar in Sm-1. ........... 215

Figure 8.25. Six selected coil arrangements for the 46 coil configuration; red coils are exciters. D6 is the

arrangement used previously in figure 8.21 (middle array). ........................................................ 216

Figure 8.26. Quality measures for 6 designs, D1-D6, normalized to Qm = 100128. ................................ 217

Figure 8.27. Columns left to right: reconstructed images for D1 to D6 designs, respectively, where was

optimised for Ehead. Rows from top to bottom: transverse, coronal, and sagittal cross-sections.

Colour-bar in S m-1 with upper limit of 0.23. ............................................................................... 218

Figure 8.28. As figure 8.27, but with 1 m added noise. ....................................................................... 219

Figure 8.29. Photographs of MK2b system at early stages of construction. Left: screen and coil formers

before mounting. Right: array (screen, formers, plastic shell) prior to adding the electronics and

coil windings. ............................................................................................................................. 220

Figure 9.1. Conceptual diagram of targets in case-A and case-B. Upper row shows x-y cross-sections, and

lower row shows vertical cross-sections along broken lines shown in upper row. Distances show

displacement of inclusion relative to centre of tank. .................................................................. 225

Figure 9.2. Photographs of experimental setup showing inclusion within tank (a) and tank within MK2a

array (b). ................................................................................................................................... 226

Figure 9.3. Planes of measurement: (a) levels of 9 planes (P1-P9) relative to tank, (b) & (c) highest and

lowers positions of tank within array, respectively. .................................................................... 226

Figure 9.4. Phase changes in degrees for uniform tank: (a) simulated data for tank of 19.2 cm diameter,

(b) measured data for tank of 19.5 cm diameter, (c) difference between (a) and (b). .................. 228

Figure 9.5. Phase differences caused by addition of inclusion for case-A, relative to uniform tank: (a)

simulated, (b) measured. ........................................................................................................... 229

Figure 9.6. As figure 9.5, but for case-B. .............................................................................................. 230

Figure 9.7. Transverse slices cutting through the tank used for image representations. ....................... 231

Figure 9.8. Case-A: Columns from left: level of cross-section along tank, target on 1 cm grid, differential

images from simulated, measured data. .................................................................................... 232

Figure 9.9. As figure 9.8, but for case-B. .............................................................................................. 233

Figure 9.10. Absolute imaging of uniform tank of 0.205 S m-1 conductivity. Transverse slides cutting

halfway up the tank. Rows from top to bottom: images reconstructed from noiseless simulated

Page 18: Magnetic Induction Tomography for Imaging Cerebral Stroke

xviii

data, simulated data with 3 m additive noise, and measured data, respectively. Colour-bar in S m-

1. ............................................................................................................................................... 235

Figure 9.11. Comparison of measured and simulated phase changes for uniform tank of 0.205 S m-1

conductivity. Blue, red, and green colours represent Ex1-Se8, Ex2-Se9, and Ex12-Se6 coil pairs,

respectively. For each coil pair (each colour), 9 points are shown which correspond to 9 planes of

measurement for that particular coil pair. The line of identity is also drawn. .............................. 236

Figure 9.12. Discrepancy between the FD and FE phase changes for the uniform tank and middle plane

of measurement. ....................................................................................................................... 237

Figure 9.13. Discrepancy between the noise-free and noisy simulated phase changes for the uniform

tank. .......................................................................................................................................... 238

Figure 9.14. Values of r.m.s. for the phase error as a function of conductivity; phase error is computed as

the discrepancy between the measured and scaled simulated phase changes. ........................... 239

Figure 9.15. Discrepancy between the measured phase changes for the uniform tank and the simulated

ones after the conductivity adjustment. ..................................................................................... 240

Figure 9.16. Discrepancy between the simulated phase changes for the uniform with adjusted

conductivity when placed at the centre of sensor array and when displaced for 2.8 mm. ........... 241

Figure A.1. An image of frond-end of the 16-channel MK1 system [33]. ............................................... 257

Figure A.2. Diagram of the MK1 configuration; dimensions in mm: (a) the coil array inside the

electromagnetic screen; labels represent the coil numbering, (b) detail of an individual coil

module; the coil windings are shown in black (sensor) and red (exciter). .................................... 258

Figure A.3. An image of frond-end of the 14-channel MK2a system. .................................................... 259

Figure A.4. Geometrical specifications of the system; dimensions in mm: (a) exciter and sensor coils sit

on red and blue broken lines, respectively, (b) details of an individual module. .......................... 259

Figure B.1. Normalized permittivity ( ' - )/( s - ), loss factor ''/( s - ), and conductivity 0 ''/( s -

) for a single time constant relaxation plotted against f/fr (after Barnes and Greenebaum [91]).

.................................................................................................................................................. 262

Figure B.2. Dispersion regions , , and (after Grimnes and Martinsen [113]). .................................. 264

Page 19: Magnetic Induction Tomography for Imaging Cerebral Stroke

xix

Abbreviations

___________________________________

A&E Accident and emergency

ABC Absorbing boundary conditions

ART Algebraic reconstruction technique

BVP Boundary value problem

CG Conjugate gradient

CNS Central nervous system

CSF Cerebrospinal fluid

CT Computed tomography

CVA Cerebro-vascular accident

ECT Electrical capacitance tomography

EIT Electrical impedance tomography

FD Finite difference

FDM Finite difference method

FDTD Finite difference time domain

FE Finite element

FEM Finite element method

FGMRES Flexible generalised minimum residual

GCV Generalized cross validation

GSVD Generalized singular value decomposition

Page 20: Magnetic Induction Tomography for Imaging Cerebral Stroke

xx

IBC Impedance boundary conditions

LS Least squares

MIT Magnetic induction tomography

MRI Magnetic resonance imaging

PML Perfectly matched layer

RF Radio frequency

RMS Root mean square

SAR Specific absorption rate

SIRT Simultaneous iterative reconstruction technique

SNR Signal to noise ratio

SOR Successive over relation

SSOR Symmetric successive over relation

SVD Singular value decomposition

TLM Transmission line matrix

TSVD Truncated singular value decomposition

Page 21: Magnetic Induction Tomography for Imaging Cerebral Stroke

Part I

Preliminaries

Page 22: Magnetic Induction Tomography for Imaging Cerebral Stroke

Chapter 1.

Introduction

1.1. MIT

Magnetic Induction Tomography (MIT) is a relatively new non-contact imaging

modality which is used for imaging the internal distribution of the passive

electromagnetic properties (conductivity, permittivity, and permeability) of an object by

measuring the mutual inductances between pairs of coils placed around its periphery

[1]. MIT relies on coupling an alternating magnetic field, the primary field, to the

region under investigation to induce interrogating eddy currents, which depend on the

internal electrical property distribution of the region. These generate a secondary

magnetic field. The secondary field is then measured with a sensor coil.

Note that the primary and secondary fields are established simultaneously, but here are

defined separately for the convenience of explanation. By having an array of exciter

coils to be activated in turn for generating the primary field, and an array of sensor coils,

a set of projections can be acquired to form an MIT data set from which the internal

distribution of the electrical properties may be inferred, i.e. reconstruct an image.

MIT is one member of the larger group of electrical tomographic techniques which are

used for non-invasive imaging of dielectric properties of materials. Related techniques

include:

Electrical Impedance Tomography (EIT): estimating the conductivity or

permittivity distribution of the subject from surface electrical measurements.

Page 23: Magnetic Induction Tomography for Imaging Cerebral Stroke

2 Introduction

Typically, conducting electrodes are attached to the skin (surface) of the subject

and small alternating currents are applied to some or all of the electrodes. The

resulting electrical potentials are measured to form the EIT data set (e.g. [2]).

Electrical Resistance Tomography (ERT): a geophysical technique for imaging

the conductivity of sub-surface structures from electrical measurements made at

the surface or by the electrodes in one or more boreholes. It is closely related to

the EIT. In contrast, however, ERT is essentially a direct current method. For

instance, in imaging the content of process vessels and pipelines, the electrodes

make electrical contact with the fluid inside the vessel (e.g. [3]).

Electrical Capacitance Tomography (ECT): determination of the dielectric

permittivity distribution in the interior of an object from external capacitance

measurements. The measurement electrodes must be sufficiently large to give a

measureable change in capacitance. ECT has been applied to, for instance,

imaging of flow systems such as oil-gas and oil-water flows (e.g. [4]).

Induced Polarization Tomography (IPT): a geophysical technique which, similar

to ECT, is used for the determination of spatial distribution of the capacitance

characteristics of the soil (e.g. [5]).

Metal detectors: a kind of manually scanned MIT which are used for the

detection of foreign bodies in food, land mines, weapons, and treasure (e.g. [6]).

Electromagnetic geophysics tomography: used for the reconstruction of the

conductivity and permittivity profiles of the geophysical medium under

investigation (e.g. [7]).

Compared to EIT and ERT, MIT has the advantage that it does not require direct contact

with the object but operates through an air gap. The contact-less nature of MIT means

that errors due to the electrode contact impedances and the associated image artefacts

can be avoided completely. Furthermore, the magnetic field easily penetrates the

barriers posed by the electrically resistive tissues such as skull bones and fat in medical

imaging. Unlike ECT, MIT is sensitive to all three passive electromagnetic properties;

MIT signals contain information that can be used to reconstruct conductivity,

permittivity, and permeability [8,9].

Page 24: Magnetic Induction Tomography for Imaging Cerebral Stroke

3 Introduction

1.2. MIT applications

There has been considerable interest shown over the past decade in the MIT imaging

technique and its application to various problems. If there is a contrast or change in

dielectric properties of the object of interest, MIT may potentially be used as an imaging

tool [10]. MIT has possible applications in fields such as:

Medical imaging: in medical applications, the non-invasive measurement of the

impedance of biological tissues can potentially yield data of diagnostic relevance

(e.g. [1]).

Industrial application: using conductivity images in monitoring of industrial

processes such as molten steel flow visualization in continuous casting (e.g.

[11]).

Non-destructive testing: eddy-current testing is very well known in non-

destructive testing which, for instance, is used for crack detection in metallic

objects (e.g. [12]).

1.3. Motivation and objectives

The main aim of this study was to numerically investigate the feasibility and sensitivity

of MIT imaging for the detection of cerebral stroke. Cerebral stroke is a major cause of

mortality and morbidity in Western society that is likely to rise with the ageing

population. The cost of aftercare places a significant burden on the UK‟s health and

social care budget. Conventional brain imaging techniques (e.g. MRI and CT) are

expensive and sometimes inaccessible, and MIT may eventually be an attractive and

low-cost (< £10k) candidate for rapid diagnostic imaging.

The major challenge for using MIT imaging in medical applications, however, is the

fact that the induced MIT signals are small. Biological tissues have very low

conductivities ( < 2 S m

-1) and their interaction with the magnetic field is weak, leading

to extremely small MIT signals. Consequently, the noise contaminating the MIT data

due to the hardware measurement errors can limit the accuracy of MIT medical

imaging. Therefore, a realistic feasibility study can be of crucial importance for each

MIT application and, to this end, the ability to accurately model the MIT signals first is

vital to the process of hardware design.

Page 25: Magnetic Induction Tomography for Imaging Cerebral Stroke

4 Introduction

Prior to commencing this research, all the numerical and experimental studies on brain

imaging used models or setups, consisting of simple geometries, which were unlikely to

accurately estimate the MIT signals in in-vivo imaging. An overview of the history of

MIT employment for medical imaging is provided in the next chapter. Therefore, in

order to assess and optimize the sensitivity of the MIT systems for clinical imaging,

more realistic simulations are required. The key questions relating to the use of MIT

technology for imaging the stroke were:

What is magnitude of the MIT signals produced by the head and stroke, and can

they be measured using the existing MIT systems?

Can an image of the stroke be reconstructed?

What is the design for the MIT coil array in order to obtain the most sensitivity

for stroke imaging?

The image reconstruction process in MIT is highly sensitive to the error in the data; a

low level of noise in the signal can cause large errors in the calculated conductivity

distribution. Consequently, contamination of the measured MIT signal with the inherent

noise in the hardware can introduce considerable artefacts in the reconstructed images.

Hence, efficient image reconstruction methods that can reduce the effect of the noise on

the images are required for the MIT technique. Therefore, the main objectives of this

study were to:

Develop a realistic head model of a typical human head in terms of geometry

and dielectric properties of different head tissues.

Simulate the MIT problem using an approximate numerical method in order to

estimate the MIT signals due to the occurrence of the stroke in the brain.

Investigate and adopt the image reconstruction algorithms for the intended

application.

Investigate different MIT coil arrays in order to improve the sensitivity for

stroke imaging.

Page 26: Magnetic Induction Tomography for Imaging Cerebral Stroke

5 Introduction

1.4. Outline of the thesis

This thesis comprises ten chapters and three appendices. In the first section of every

chapter, there is an introduction on the subject of that chapter, followed by the main

body:

In chapter 1, the MIT imaging technique is introduced and a brief account of the

research study and this thesis is provided.

In chapter 2, the working principle of a typical MIT system is explained and

physical concepts are discussed. The eddy-current problem in MIT is described.

An overview of MIT studies for medical imaging reported by different research

groups is also provided.

Chapter 3 briefly describes the anatomy of the head and its tissues and

introduces the different types of stroke.

A literature survey of dielectric properties of the head tissues at different

frequencies is carried out in chapter 4, and the effect of stroke on the dielectric

properties is explained.

The discretized MIT problem is dealt with in chapter 5. Two approaches for the

numerical approximation of the MIT problem (i.e. finite element and finite

difference) are explained. A benchmark test is conducted to assess the accuracy

of the employed numerical techniques.

Chapter 6 is devoted to the image reconstruction process. It discusses different

single-step and iterative image reconstruction algorithms. Images of

geometrically simplified models are reconstructed as evaluation tests prior to

stroke imaging using the anatomically realistic head model.

MIT simulations for a 16-channel cylindrical system are presented in chapter 7.

The MIT problem is solved for strokes with different sizes and sites in the brain

and 3D images are reconstructed from the simulated data.

In chapter 8, the performance of customised MIT systems for stroke imaging is

investigated. Different numbers and sizes of the exciter/sensor coils are

evaluated in order to enhance the sensitivity of the system. Details of the

adopted design are also provided. Chapter 8 also describes examples of

systematic errors which are likely to occur during in-vivo imaging, and evaluates

their effect on the images.

Page 27: Magnetic Induction Tomography for Imaging Cerebral Stroke

6 Introduction

Preliminary imaging studies based on the measured MIT data is provided in

chapter 9.

Chapter 10 concludes the study while some suggestions for further development

are also provided.

A bibliography is placed at the end of this thesis, and is preceded by the appendices.

Appendix A provides a brief description of two existing MIT measurement systems,

constructed by the South-Wales MIT group. Appendix B explains the basic concepts of

dielectric theory and terminology. And finally, Appendix C provides some insight into

some of the image reconstruction techniques.

1.5. Contributions to knowledge

Considering the novel elements of the research undertaken, the following contributions

to knowledge are claimed:

An up to date review of the dielectric properties values for the human head

tissues published since 1996.

Extension of an existing EIT head geometry for MIT applications.

Comparison of the finite element and finite difference models for the MIT

problem.

Numerical simulation of MIT signals due to the presence of stroke in a realistic

head model.

Reconstruction of 3D images of stroke using the realistic numerical model.

Improvement of the geometrical design of MIT system for stroke imaging.

Investigation of the effect of systematic errors on the stroke images.

Preliminary MIT imaging experiments using practical measured data for simple

geometries.

During the course of this PhD study, outcomes of the research have been published in

three journal articles, and presented at several relevant national and international

conferences:

Page 28: Magnetic Induction Tomography for Imaging Cerebral Stroke

7 Introduction

Journal articles:

1. M. Zolgharni, P.D. Ledger, D.W. Armitage, D.S. Holder, H. Griffiths, Imaging

cerebral haemorrhage with magnetic induction tomography: numerical

modelling, Physiol. Meas. (2009) 30:187-200

2. M. Zolgharni, P.D. Ledger, H. Griffiths, Forward modelling of magnetic

induction tomography: a sensitivity study for detecting haemorrhagic cerebral

stroke, Med. Biol. Eng. Comput. (2009) 47:1301-1313

3. M. Zolgharni, H. Griffiths, P.D. Ledger, Frequency-difference MIT imaging of

cerebral haemorrhage with a hemispherical coil array: numerical modelling,

Physiol. Meas. (2010) in press

Conference proceedings:

4. M. Zolgharni, P. D. Ledger, H. Griffiths, D. S. Holder, Numerical Modelling of

the Cerebral Stroke in Human Brain, 16th UK Conf. on Computational

Mechanics (ACME), Newcastle upon Tyne 2008, UK , pp.139-142.

5. M. Zolgharni, P.D. Ledger, D.W. Armitage, H. Griffiths, D.S. Holder,

Detection of haemorrhagic cerebral stroke by magnetic induction tomography:

FE and TLM numerical modelling, EIT 2008, Dartmouth NH, USA, June 16-18.

6. Y. Maimaitijiang, S. Watson, M.A. Roula, M. Zolgharni, H. Griffiths, R.J.

Williams, An Iterative Absolute Image Reconstruction Algorithm for Magnetic

Induction Tomography, EIT 2008, Dartmouth NH, USA, June 16-18.

7. M. Zolgharni, H. Griffiths, D.S. Holder, Imaging haemorrhagic cerebral stroke

by frequency-difference magnetic induction tomography: numerical modelling,

2008 IFMBE Proc. 22 2464-7 ISBN 978-3-540-89207-6.

8. B. Dekdouk, M.H. Pham, D.W. Armitage, C. Ktistis, M. Zolgharni, A.J.

Peyton, A feasibility study on the delectability of edema using magnetic

induction tomography using an analytical model, 2008 IFMBE Proc. 22:736-9

ISBN 978-3-540-89207-6.

9. M. Zolgharni, P.D. Ledger, H. Griffiths, Imaging cerebral haemorrhage with

MIT: frequency-difference imaging with a customised coil array, EIT 2009,

Manchester, UK, June 16-19.

Page 29: Magnetic Induction Tomography for Imaging Cerebral Stroke

8 Introduction

10. M. Zolgharni, P.D. Ledger, H. Griffiths, High-Contrast Frequency-Difference

Imaging for Magnetic Induction Tomography, EIT 2009, Manchester, UK, June

16-19.

11. S. Watson, H.C. Wee, R. Patz, R.J. Williams, M. Zolgharni, H. Griffiths,

Detection of peripheral haemorrhagic cerebral stroke by magnetic induction

tomography: phantom measurements, EIT 2009, Manchester, UK, June 16-19.

12. M. Zolgharni, P.D. Ledger, H. Griffiths, Difference imaging of cerebral stroke

in the human brain using edge finite element simulation of magnetic induction

tomography, CMBE 2009, pp.288-91, Swansea, UK, ISBN 978-0-9562914-0-0

13. H. Griffiths, M. Zolgharni, P.D. Ledger, S. Watson, The Cardiff Mk2b MIT

head array: optimising the coil configuration, EIT 2010, Gainesville FL, USA,

April 4-8

1.6. LCOMIT

This study forms part of a larger 3.5-year EPSRC1-funded collaborative project,

LCOMIT2

, for imaging low-conductivity objects involving academic partners at

Swansea University [13], University of Manchester [14], University of Glamorgan [15],

University College London, ABM University Health Board, and industrial/NHS

partners at Philips research laboratories in Aachen, Germany [16]. The work at Swansea

University is devoted to the application of MIT for stroke imaging, while the work by

other partners involves the development of hardware and software aspects of MIT in

general and for two other specific industrial applications, involving a three further

industrial partners.

End of chapter 1.

1 Engineering and Physical Sciences Research Council 2 LCOMIT: Low conductivity magnetic induction tomography

Page 30: Magnetic Induction Tomography for Imaging Cerebral Stroke

Chapter 2.

Principles of MIT

2.1. Introduction

The purpose of this chapter is to outline the MIT imaging technique. First the physical

concepts and governing equations of the forward and inverse boundary value problems

are explained. Next, the working principles and the constituent parts of a typical MIT

system are discussed. An overview of the history of employing MIT for brain imaging

applications is also provided.

2.2. Physical principles

MIT imaging, and electrical tomography techniques in general, consist of solving

boundary value problems that are governed by Maxwell's equations. The corresponding

'forward' and 'inverse' problems are defined as:

The purpose of solving the forward boundary value problem is to compute the

electromagnetic fields for a given geometry, distribution of dielectric properties,

operating frequency, and source current, subjected to known boundary

conditions.

In contrast, the purpose of solving the inverse boundary value problem is to

recover the distribution of dielectric properties from the measured fields on

(parts of) an enclosed surface surrounding the region of interest.

Page 31: Magnetic Induction Tomography for Imaging Cerebral Stroke

10 Principles of MIT

For time-harmonic fields varying with an angular frequency of (s-1

), the governing

Maxwell's equations are expressed by:

(2.1)

(2.2)

(2.3)

(2.4)

In the above equations, the field quantities E (V.m-1

) and H (A.m-1

) represent the

electric field and magnetic fields, respectively, D (C.m-2

) and B (T) represent the

electric and magnetic displacements, and j2 = -1. The term s (C

m

-3) is the charge

density. The current density J (A.m-2

) has two components:

(2.5)

where Jc represents the conduction current density induced by the presence of the

electric field, Js is a source current density that induces the electromagnetic fields (here

assumed divergence free throughout), and (S.m-1

) is the conductivity of the

constituting medium.

The auxiliary constitutive relationships between the field quantities and electric and

magnetic displacements provide the additional constraints needed to complete the

equations (2.3) and (2.4). These equations characterize a given material on a

macroscopic level in terms of the two quantities, permittivity (F.m-1

), and permeability

(H.m-1

) as:

(2.6)

(2.7)

where 0 = 4 ×10-7

H m

-1 and 0 = 8.85×10

-12 F

m

-1 are the permeability and

permittivity of free space, respectively. Also, r and r are the relative permeability and

permittivity of the constituting medium, respectively. For isotropic linear materials,

which are of interest in this study, the electrical properties ( , , ) are scalar quantities

(material properties of the tissues in the head are discussed in chapter 4).

Page 32: Magnetic Induction Tomography for Imaging Cerebral Stroke

11 Principles of MIT

Forward problem:

For a bounded domain , the forward problem is: Find E and H such that:

(2.8)

where 3, = D N with D N = , GD and GN are general Dirichlet and

Neumann boundary conditions, vector n is a unique outward normal at each boundary

or interface, Js is a given current source such that Js = 0 in , ( , , ) are given

distributions of the medium parameters, [...] represents the jump in the quantity, and

finally i j denotes the interface between the regions of different materials [17].

Inverse problem:

Find the distributions of ( , , ) such that (2.8) holds for given E, H.

2.2.1. MIT problem

Figure 2.1 shows a typical boundary value problem for the MIT application. The

domain of the problem, , consists of a conducting region C, ( C) > 0, and a non-

conducting region N, ( N) = 0, so that = C N.

Page 33: Magnetic Induction Tomography for Imaging Cerebral Stroke

12 Principles of MIT

Figure 2.1. Schematic view of boundary value problem in MIT; C: conductive domain, N: non-

conducting region (free space).

In reality, N should be replaced by N which is of infinite extent. In this case,

would become unbounded and the boundary value problem described by (2.8) would

need to be approximated by a radiation condition [18]:

(2.9)

However, to ensure the boundedness of , N is truncated at a finite distance from the

conducting object, and the artificial domain N, and consequently the artificial

boundary , are created. By relaxing the constraint that D N = , and providing

the exact values of both n E and n H on all of , no approximation is made. The

choice of appropriate approximate boundary conditions for is discussed in chapter 5.

The conducting region will, in general, consist of a number of materials, and the

appropriate interface conditions described in (2.8) should be applied.

The exciter/sensor coils lie in N: E represents an exciter coil on which the current

source Js is applied, and S is a sensor coil on which the voltage is computed as:

(2.10)

where dl is the length element of the coil.

2.2.2. Eddy-current problem

For certain low-frequency application (e.g. non-destructive testing [19]), the problem

described by (2.8) may be simplified to the so-called eddy-current model. The eddy-

E

C

σ, ,

N : σ =0, = 0, = 0

S

Page 34: Magnetic Induction Tomography for Imaging Cerebral Stroke

13 Principles of MIT

current model is a magnetostatic approximation of the full Maxwell system, and

involves neglecting the displacement currents.

A number of justifications for this exist in the literature. These include those based on

the heuristic arguments that the eddy-current approximation is meaningful if [20]:

the dimension of the conducting object is small compared to the wavelength

the conductivities are high ( << )

More recently, these arguments have been found to be insufficient for certain

geometries, and have been refined to include the topology of the conducting object

[18,21]. Neglecting the displacement current leads to a problem which reduces to

solving a magnetostatic problem in the non-conducting region N (and consequently

N), and also allows the radiation condition to be replaced by the simpler requirement

that [18]:

(2.11)

For the medical MIT applications, however, the material properties are such that the

ratio / can be as large as 0.6 at 10 MHz (electrical properties of head tissues are

provided in chapter 4). Therefore, on the grounds of the earlier heuristics, the eddy-

current approximation must be rejected and the displacement current should be retained,

despite the fact that the frequency of operation is relatively low ( 10 MHz). Retaining

the displacement current means that wave propagation effects are included in the

problem. As the frequency increases, these effects become more important.

It is remarked that MIT systems used for imaging metallic conductors do employ the

eddy-current approximation of (2.8); e.g. [22]. Nevertheless, eddy currents (Je = E)

will form an important part of the discussion in later chapters, as will the skin depth.

The skin depth is defined as the distance measured from the surface of the medium over

which the magnitudes of the fields are reduced to 1/e, or approximately 37%, of those at

the surface of the medium. For the case of a semi-finite conductor with uniform

conductivity and permeability , it can be calculated as = (2/ 0 r )1/2

[23].

Page 35: Magnetic Induction Tomography for Imaging Cerebral Stroke

14 Principles of MIT

2.2.3. Solving the (forward) problem

Analytical solutions for the forward problem are only available for a small number of

relatively simple geometries, e.g. a conducting sphere [24,25,26,27,28] or disk [29], and

when the materials are isotropic. For complex geometries such as a realistic human head

model, or in case of dealing with anisotropic materials, numerical techniques must be

employed where the solution to the problem is approximated by discretizing the model.

The discrete forward and inverse problems are discussed in chapters 5 and 6,

respectively.

2.3. Working principles

The main elements of a complete MIT system are shown in figure 2.2. According to

Peyton et al [30], the system typically comprises three main sub-systems: a sensor

array, the electronics components, and a host computer. The sensor array consists of:

Exciter coils array

Sensor coils array

An electromagnetic confinement shield

Electronic circuitry for exciter and sensor coils

Figure 2.2. Block diagram of a typical MIT system (after Peyton et al [30]).

The outer electromagnetic shield (screen) performs several functions. In particular, it

confines the interrogating field within the imaging volume, thereby removing the

potential for interference by external conductive or magnetic objects in the vicinity. The

Screen

Page 36: Magnetic Induction Tomography for Imaging Cerebral Stroke

15 Principles of MIT

screen reduces the susceptibility to external magnetic field interference by providing a

shunt path around the object space [31]. It may also influence the sensitivity of the

system. In addition, the screen can act as a ground plane that helps to reduce undesirable

electric-field (capacitive) coupling between the coils [32]. Details of the electronic

components and circuitry are outside the scope of this study; useful information for MIT

electronic developments can be found in [33,34,35,36]. The image reconstruction,

which requires the solution of the inverse problem, is performed on the host computer.

By solving the MIT inverse problem, the temporal-spectral conductivity differences or

the static (absolute) conductivity values may be reconstructed from the measurements.

The MIT imaging can, therefore, be categorized into two groups; difference and

absolute imaging:

Difference imaging: differential or difference imaging is to estimate a change in

conductivity between two states from the MIT signals measured either at

different times (time-difference imaging), or with different frequencies

(frequency-difference imaging). The differential technique has been used for

imaging temporal-spectral phenomena in medical applications, such as lung

imaging reported by Gursoy and Scharfetter [37].

Absolute imaging: a special case of difference imaging is absolute imaging,

when one of the states is free space; an image is produced using a single set

(frame) of MIT data. Absolute imaging, thus, attempts to calculate the actual

values of the conductivity in S m

-1. The use of absolute imaging is more difficult

as the accuracy of both the measured and predicted voltages is crucial (further

discussed in chapters 6 and 9).

Depending on the application, the appropriate imaging type must be adopted. The

overall MIT procedure, excluding the hardware preparation, can be summarized as

follows:

I. Input: measure the MIT signals induced by the object to be imaged.

II. Define problem: define the MIT setup including the coil arrays and screen,

operating frequency, geometry (outer boundary) of the object. In some cases, the

geometry data is not available.

III. Forward problem: develop the physical model that represents the boundary

value problem.

Page 37: Magnetic Induction Tomography for Imaging Cerebral Stroke

16 Principles of MIT

IV. Sensitivity analysis: analyse the sensitivity of the output (MIT signals) to the

input parameters (conductivity distribution).

V. Inverse problem: apply the inverse algorithm using the sensitivity data and the

measured signals.

VI. Output: extract the conductivity distribution.

In some cases, the inverse problem is solved iteratively, in which case the steps (III) to

(VI) are repeated until convergence is obtained.

2.3.1. The MIT signal

As shown in figure 2.3, the change in the voltage due to the presence of an object

(target) has real and imaginary components. The total measured signal, i.e. primary plus

secondary, lags the primary signal by an angle .

Figure 2.3. Phasor diagram representing primary (Vp) and secondary ( V) voltages.

Provided that the skin depth of the electromagnetic field in the target is large compared

with the dimension of the target, which will often be so for a segment of the human

body [33], Griffiths et al [8] have shown that the induced voltage change, V,

normalized to the primary voltage, Vp, can be expressed as a phase change:

(2.12)

where C1 and C2 are the geometrical constants, and V << Vp.

The conduction currents induced in the target give rise to a component of V which is

proportional to frequency and conductivity. It is also imaginary and negative (it lags the

primary signal by 90 ). Displacement currents cause a real (in-phase) component

Vp Re( V)

Im( V)

Vp + V

Page 38: Magnetic Induction Tomography for Imaging Cerebral Stroke

17 Principles of MIT

proportional to the square of the frequency and permittivity. This component is

independent of the conductivity. A non-unity relative permeability also gives rise to a

real component, but with a value independent of frequency. However, since r is very

close to unity for biological tissues (| r - 1| < 10-5

), the second term on right hand side of

(2.12) can be dropped for most purposes.

In general, the real and imaginary parts of the phase change = tan-1

( V/V) may be

used for imaging the permeability-permittivity and conductivity of the target,

respectively.

For larger conductivities (e.g. in metals), the skin depth may become comparable with

the target's dimension. In this case, Re( V/V) will be affected by the conductivity and

equation (2.12) fails to predict the correct values for this component (further discussed

in chapter 5). As described by Griffiths et al [8], in such cases the target acts as a screen

between the exciter and sensor coils, and cause a negative Re( V/V).

In a practical system, an additional (undesired) coupling mechanism exists via the

capacitance between the exciter and sensor coils, and the target. This can cause large

errors in the MIT signal, and efforts by the system designers are made to reduce the

capacitive coupling. A detailed discussion of the capacitive coupling is outside the

scope of this study. Further details can be found in Goss et al [38].

2.3.2. SNR and operational frequency

Since the MIT signals are measured by a practical data-collection system, they are

inevitably contaminated by the noise in the hardware. The noise in the MIT data is

defined by the signal-to-noise ratio (SNR), which is the ratio of the signal (meaningful

information) to the noise corrupting the signal:

(2.13)

where P is the average power and A is the amplitude of the voltage. Employing higher

frequencies for MIT operation can be beneficial. High-frequency (3–30 MHz) operation

of MIT systems offers advantages in terms of the larger induced signal amplitudes

compared to systems operating in the low- or medium-frequency ranges [39].

Page 39: Magnetic Induction Tomography for Imaging Cerebral Stroke

18 Principles of MIT

While the noise amplitude in the hardware is independent of the frequency, large

induced signal amplitudes results in larger SNR. However, a higher frequency range not

only can make the simulations more complicated, it can also present difficulties in terms

of the phase stability of the distributed signals, and of achieving the required isolation

between channels in a multi-channel system [39]. Furthermore, as will be explained in

chapter 4, frequency-dependence of the dielectric properties of biological tissues may

suggest particular frequency ranges for biomedical MIT applications. Therefore,

multiple factors need to be taken into account in designing the MIT system for a

particular application.

If the MIT signals are normalized to the primary signal as in (2.12), the noise is defined

in m rather than dB, and is often referred to as phase noise.

2.4. MIT studies for brain imaging

Griffiths provided two comprehensive reviews in 2001 [40] and 2005 [1], looking at

different MIT systems for industrial and medical applications. The main focus of this

section, however, is on the numerical/experimental MIT studies for brain imaging

applications.

Tarjan and McFee in 1968 [41] measured an average value for brain conductivity

inductively. Later in 1993, Netz et al [42] suggested that brain oedema might be

detected more rapidly from conductivity changes than is possible from CT or MRI

imaging. A single-channel MIT system for in-vivo imaging was reported by Al-Zeibak

et al in 1993 [43]. However, only imaging of metallic and ferromagnetic samples at 10

MHz were reported, which were not a suitable representation of the head tissues.

Later in 1999, construction of a single-channel MIT system operating at 10 MHz was

reported by Griffiths et al [29] from the South-Wales MIT group. The MIT signals were

measured for saline solutions with conductivities ranging from 0.001 to 6 S

m

-1

encompassing the range of conductivities of biological tissues. Since then, there have

been a number of numerical and experimental studies on the feasibility of using MIT for

imaging brain lesions.

Page 40: Magnetic Induction Tomography for Imaging Cerebral Stroke

19 Principles of MIT

Merwa and Schaefetter [44] from the Graz MIT group, reported on the sensitivity study

of a single-channel MIT system to an oedematous region using a semi-realistic model of

the human brain. Their model comprised the cerebrospinal fluid (CSF) around the brain,

grey matter, white matter, ventricles, and the oedema (a spherical inclusion). The finite

element method (FEM) was used to study the effect of different sizes and positions of

the oedema on the MIT signals. It was shown that a centrally located oedema with a

conductivity contrast of 2 with respect to the background and radius of 20 mm could be

detected at 100 kHz. The same authors later reported the reconstruction of time-

differential images from the simulated oedema using a multi-channel MIT system

comprising 16 exciter and 32 sensor coils [45].

The spectroscopic distribution of the MIT signal as a function of the relative volume of

oedema and haematoma in the brain was investigated at the Berkeley MIT group by

Gonzalez and Rubinsky [46]. They considered a simple circular disk to represent the

brain and used the analytical solution to compute the signals. The total volume of the

disk was considered as the oedema/haematoma region. It was shown that the MIT signal

was sensitive to the relative volume of the oedema/haematoma at frequencies higher

than 10 MHz, and increasing the volume of tissue had the effect of lowering the

required frequency at which the signal became sensitive to the volume of the

oedema/haematoma. However, considerable simplifications were used in their model.

Later they reported the experimental studies for their simulated model [47,48]. In other

reports [49,50], they numerically compared the MIT signals due to the haematoma in a

circular sensor coil array with those in a magnetron sensor coil. They suggested that a

system with a magnetron sensor is a preferential configuration with greater sensitivity

for the centrally located lesions.

The same group recently reported on the detection of brain ischaemia in rats [51]

(ischaemia is one of the stroke types and is introduced in chapter 3). Their system

consisted of two electromagnetic coils placed around the head. Ischaemia was induced

in the rat's brain by occlusion of the right cerebral and carotid arteries. Experimental

subjects were monitored for 24 hours and measurements were made at five frequencies

in the range of 0.1-50 MHz. A mathematical head model comprising the brain, bone and

skin was also employed to simulate the MIT signals, for comparison with the

experimental data. They concluded that the experimental results were roughly in

qualitative agreement with those of the numerical model. Both theoretical and

Page 41: Magnetic Induction Tomography for Imaging Cerebral Stroke

20 Principles of MIT

experimental results suggested that the inductive signals increase as a function of

frequency and with the level of the simulated ischaemic damage, i.e. volume of the

lesion.

Researchers from China used the finite difference time domain (FDTD) method to

investigate the effect of oedema in a single channel MIT system operating at frequency

of 10 MHz [52]. Simplified spherical models were used to represent a three-layered

head geometry comprising skull, CSF and brain, and also the inclusion (oedema). By

assessing the MIT signals, it was concluded that the MIT imaging of oedema is feasible.

However, no images were reconstructed. Later they reported on the sensitivity of the

system to the site, size and conductivity of the oedema [53].

Construction of a 15-channel helmet shaped MIT device for brain imaging has been

recently reported by another group from China [54]. A hemispherical glass bowl filled

with the salt solution was used as the brain, and agar blocks of different conductivities

were placed at different peripheral locations in the solution to represent the

haemorrhage. Measurements were carried out at different frequencies raging 30 to 120

kHz and it was concluded that the MIT signal values increased linearly with the

frequency. They used interpolation algorithms to obtain 2D plots; this method suffered

from poor spatial resolution for brain imaging applications. Reconstruction of 3D

images using this MIT system has not been reported as yet.

Korjenevsky [55] from the Moscow MIT group, has obtained the first in-vivo image

using a 16-coil MIT system operating at frequency of 20 MHz. An absolute image,

reconstructed using neural network, appears to show a cross-section of the human head,

with two bright features interpreted as the lateral ventricles of the brain filled with CSF.

Vauhkonen et al [56] from the Philips MIT research group in Germany (one of the

partners in LCOMIT project), reported a numerical investigation of imaging stroke

using a 16-channel MIT system. They reconstructed images of geometrically simplified

stroke lesions with different volumes and locations in the head. Cylindrical tanks and

objects were used to simulate the head, neck, shoulder, and the stroke. It was concluded

that centrally located small inclusions (i.e. strokes) are extremely difficult to detect

since they affect the measurements only slightly. In a recent report [57], they compared

the cylindrical and the hemispherical coil setups of two MIT systems using sensitivity

analysis by numerical simulations. Different parameters for the size and the number of

Page 42: Magnetic Induction Tomography for Imaging Cerebral Stroke

21 Principles of MIT

measurement and excitation coils were tested. Their findings indicated that the

hemispherical MIT system with a smaller distance between the layer of coils and the

object shows a clearly higher sensitivity compared to the cylindrical MIT system. They

concluded that coils should be placed as close as possible to the object which, for the

human head, suggests a hemispherical coil arrangement.

The South-Wales MIT group used a saline bath to represent the brain, with an immersed

block of agar simulating stroke [58]. The conductivity contrast between the agar and the

saline was a factor of 3.3, being similar to the contrast between blood and brain at

frequency of 10 MHz. The images were reconstructed based on the measured signals

using a 16-channel MIT system, and it was concluded that stroke could be identified in

the images.

Dekdouk et al [24] from the Manchester MIT group (one of the partners in LCOMIT

project), reported a feasibility study on the delectability of oedema using a single MIT

channel. They used an analytical multilayer spherical model to represent the head. The

delectability of the oedema with regard to noise limitations of the MIT systems was

analyzed, but no images were reconstructed. They recently used the anatomically-

realistic head geometry, also used in this thesis, to solve the forward problem for strokes

occurring at different locations within the head. The simulated MIT configuration was a

16-channel cylindrical system, operating at frequency of 10 MHz. They concluded that

the MIT signal decays when the stroke approaches the centre as opposed to the

peripheral region. No images were reconstructed.

As mentioned in chapter 1, the results of this thesis have been published in literature.

Zolgharni et al [59] from the South-Wales MIT group, reported on time-differential

imaging of stroke using a simulated 16-channel MIT system, operating at frequency of

10 MHz. A realistic model was employed for solving the forward and inverse problems,

and 3D images were reconstructed [60]. In another report [61], frequency-differential

imaging (1-10 MHz) using the same simulated MIT configuration was examined.

Frequency-differential imaging of stroke using hemispherical coil arrays was recently

reported [62]. Two types of systematic errors were added to the simulated MIT signals,

and 3D images were reconstructed. Griffiths et al [63] have reported on the construction

of a helmet shaped MIT array, for which the geometrical design was based on the

findings of this thesis. To the author's knowledge, at the time of writing, the use of an

Page 43: Magnetic Induction Tomography for Imaging Cerebral Stroke

22 Principles of MIT

anatomically realistic model to estimate the MIT signals due to a stroke in the head,

reconstruction of differential stroke images, and addition of systematic errors is the first

application of its kind.

2.5. MK1 and MK2a systems

Most of the studies carried out in this thesis, have simulated one of the two existing

annular (cylindrical) measurement systems in the South-Wales MIT group (MK1 and

MK2a) as the MIT configuration. A brief description of these systems is provided in

Appendix A. From here on in this thesis, for the discussion of results and when

explaining the simulation methods, the simulated MIT configuration is simply cited as

the MK1 or MK2a system.

End of chapter 2.

Page 44: Magnetic Induction Tomography for Imaging Cerebral Stroke

Chapter 3.

Anatomy of Head and Stroke Pathology

3.1. Introduction

In order to be able to predict the MIT signals produced by the head, a realistic model is

required for the simulations. The model should consist of: (i) an accurate geometry

which represents the complex shape of the human head, and (ii) the most accurate

dielectric properties, available in literature, assigned to each head tissue. The former

needs a knowledge of gross anatomy of the head, and the latter requires a knowledge of

tissue electrical properties. The purpose of this chapter is to review the internal structure

of a healthy human head. It also describes the pathologies of interest (i.e. stroke). The

electrical properties of the tissues are discussed in the following chapter, while details of

the simulations with a realistic model are provided in chapter 7.

3.2. Head anatomy

The human head comprises various organs with particular functionalities and a detailed

review is required to explain all different structures in the head. However in this study,

regardless of the physiology, the main tissues, which are expected to mostly contribute

to the MIT signals, are discussed. Their contribution to the MIT signals will be

influenced by: (i) their volume, (ii) their location relative to the surface of the head and

consequently exciter-sensor coils, and (iii) their electrical properties. These tissues

include bones, muscles, cerebrospinal fluid, brain with its sub-tissues, and the air-filled

Page 45: Magnetic Induction Tomography for Imaging Cerebral Stroke

24 Anatomy of Head and Stroke Pathology

cavities. In the following, the tissues are introduced from outside in, and the cavities in

the head are described at the end.

3.2.1. Scalp

The scalp is the anatomical area bordered by the face anteriorly and the neck to the sides

and posteriorly [64]. The scalp electrical impedance plays a major role in head EIT

imaging, as it is engaged with currents of the highest density coming from the

electrodes and, therefore, is the first stage in which distribution of current into the rest

of the layers is determined. However, it is of less importance in MIT imaging, which is

a contact-less technique, and yet the scalp is the closest tissue to the coils.

3.2.2. Muscles

Skeletal muscles are contractile tissues which are made up of long and cylindrical

individual components known as muscle fibres (figure 3.1a). According to Mader [65],

the term muscle refers to multiple bundles of muscle fibres held together by connective

tissue. Muscles show anisotropic properties; electrical conductivity along the length of

the fibre is higher than that in the transverse direction. However, as will be discussed in

chapter 4, the anisotropy becomes less important as the frequency increases. Figure 3.1b

illustrates the head, neck, and facial muscles. The muscles of facial expressions are

located on the scalp and the face. Long strap-like neck muscles extend from base of the

human skull to the neck.

Figure 3.1. Skeletal muscles: (a) structure of a skeletal muscle fibre (after Saladin [66]), (b) muscles

of head and neck (after Mader [65]).

(a) (b)

temporalis

frontalis

masseter

sternocleidomastoid

trapezius

zygomaticus

buccinator

myofibril:

bundle of

fibres

sarcolemma

Page 46: Magnetic Induction Tomography for Imaging Cerebral Stroke

25 Anatomy of Head and Stroke Pathology

Figure 3.2. (a) simplified side view of human skull (after Villarreal [67]). (b) illustration of a typical

long bone showing the location of cortical and cancellous bones (picture adapted from [68]); skull

bones have a similar internal structure.

3.2.3. Skull

The human skull consists of 22 bones, 8 cranial and 14 facial, and contains several

cavities that house the brain and sensory organs. McCann et al [69] describe the skull as

a geometrically complex structure which can be divided into two major parts: the

cranium and the mandible. The mandible (jawbone) is the only moveable portion of the

skull. A skull that is missing a mandible is a cranium. The largest cavity in the skull,

cranial cavity, accommodates the brain and has a capacity of 1,300-1,350 cc [70].

Figure 3.2a shows a side view of the simplified human skull. The rigid part of the bone

is called osseous tissue or bone tissue which is not homogeneous. There are two types

of osseous tissue; compact and spongy (figure 3.2b). Compact (also cortical or dense)

bone forms the extremely hard exterior. Spongy (also cancellous or trabecular) bone

fills the hollow interior and has a higher surface area but is less dense, softer, weaker,

and less stiff. The tissues are biologically identical; the difference is in how the

microstructure is arranged. A realistic model of the head may consider heterogeneous

bones with their different electrical properties.

3.2.4. Meninges

The brain is protected by three membranous connective tissue coverings called the

meninges. According to Seely [71], the meninges have three layers: dura mater,

(a) (b)

Page 47: Magnetic Induction Tomography for Imaging Cerebral Stroke

26 Anatomy of Head and Stroke Pathology

arachnoid mater, and pia mater (figure 3.3a). The most superficial and thickest

membrane is the dura mater, which is in contact with the skull. The next meningeal

membrane, arachnoid, is the middle of the three meninges. The thin pia mater is tightly

bound to the convolutions of the brain and the irregular contours of the spinal cord.

The space between the arachnoid membrane and the dura mater is the subdural space

and contains only a very small amount of serous fluid. Between the arachnoid mater and

the pia mater is the subarachnoid space (figure 3.3b). The meninges are one of the

regions at which stroke may occur.

Figure 3.3. Meninges; (a) meningeal membranes surrounding brain, (b) frontal section of head to

show the meninges (after Seely [71]).

Page 48: Magnetic Induction Tomography for Imaging Cerebral Stroke

27 Anatomy of Head and Stroke Pathology

Figure 3.4. Human brain; cerebrum, seen here in longitudinal section, is the largest part of brain in

humans (after Mader [65]).

3.2.5. Brain

As Mader [65] describes, the brain is that part of the central nervous system (CNS)

contained within the cranial cavity. It is the control centre for many of the body‟s

functions and consists of the brainstem, the cerebellum, the diencephalon, and the

cerebrum (figure 3.4).

Cerebrum: the cerebrum is the largest and most obvious portion of the brain.

According to Shier et al [72], it accounts for about 80% of the mass of brain and

consists of two halves, the right and left hemispheres, which are incompletely separated

by a deep groove called the longitudinal fissure. The falx cerebri, which is a strong and

arched fold of dura mater, descends vertically in the longitudinal fissure between the

cerebral hemispheres (see figure 3.3b). Many ridges called convolutions, or gyri,

separated by shallow grooves, mark the cerebrum‟s surface. Each cerebral hemisphere

is subdivided into five lobes which are named according to the overlying cranial bones;

frontal lobe, parietal lobe, temporal lobe, occipital lobe, and insula (figure 3.5a).

Page 49: Magnetic Induction Tomography for Imaging Cerebral Stroke

28 Anatomy of Head and Stroke Pathology

Figure 3.5. Cerebrum: (a) lateral view of right hemisphere with insula exposed; colours in this

figure distinguish the lobes of cerebral hemispheres (after Shier et al [72]), (b) coronal section of

brain showing cortex, white matter, and corpus callosum (after Scanlon and Sanders [73]).

A thin layer of grey matter, 2 to 5 mm thick, called the cerebral cortex constitutes the

outermost portion of the cerebrum (figure 3.5b). Just beneath the cerebral cortex is a

mass of white matter that makes up the bulk of the cerebrum. Portions of the two

hemispheres are connected internally by the corpus callosum, a large tract of white

matter (figure 3.5b). While the bulk of the cerebrum is composed of white matter tracts,

there are masses of grey matter (so-called basal nuclei or basal ganglia) located

bilaterally deep within the white matter [73]. Since the white and grey matters have

different dielectric properties, the distinction between these tissue types is important in

modelling.

(a)

(b)

Page 50: Magnetic Induction Tomography for Imaging Cerebral Stroke

29 Anatomy of Head and Stroke Pathology

Brainstem: the brainstem connects the brain to the spinal cord. It consists of the

midbrain, pons, and medulla (see figure 3.4). These structures include many tracts of

nerve fibres and masses of grey matter called nuclei. The midbrain, the smallest region

of the brainstem, is a short section of the brain stem between the diencephalon and the

pons. According to Shier et al [72], the pons appears as a rounded bulge on the

underside of the brainstem where it separates the midbrain from the medulla (figure

3.4). The medulla is the most inferior part of the brainstem.

Diencephalon: the hypothalamus and thalamus are in the diencephalon which is the

part of brain between the brainstem and cerebrum, and is almost completely surrounded

by the cerebral hemispheres (figure 3.6). Saladin [66] describes the thalamus as a large

oval mass of grey matter which constitutes nearly four-fifths of the diencephalon. The

hypothalamus, named for its position below the thalamus, contains several masses of

nuclei that are interconnected with other parts of the nervous system (see figure 3.5b).

Cerebellum: the cerebellum is a large mass of tissue located inferior to the occipital

lobes of the cerebrum and attached to the brainstem posterior to the pons and medulla

(figure 3.4 and figure 3.6). It consists of two lateral hemispheres partially separated by a

layer of dura mater. Like the cerebrum, the cerebellum is primarily composed of white

matter with an outer thin layer of grey matter [65].

Figure 3.6. Illustration of the brainstem, diencephalon and cerebellum (after Shier et al [72]).

Page 51: Magnetic Induction Tomography for Imaging Cerebral Stroke

30 Anatomy of Head and Stroke Pathology

3.2.6. Ventricles and cerebrospinal fluid

Each cerebral hemisphere contains a relatively large cavity, the lateral ventricle. The

first lateral ventricle in the left cerebral hemisphere, and the second one in the right

cerebral hemisphere. The lateral ventricles are separated from each other by a thin

channel which lies in the midline just inferior to the corpus callosum, and according to

Graaff [70], are usually fused with each other (see figure 3.4 and figure 3.5b). A smaller

cavity, the third ventricle, is located in the centre of the diencephalon between the two

halves of the thalamus. The fourth ventricle is in the superior region of the medulla at

the base of the cerebellum. It is continuous with the central canal of the spinal cord, and

the subarachnoid space of the meninges.

Cerebrospinal fluid (CSF) is a fluid which bathes the brain and the spinal cord and

provides a protective cushion around the CNS from mechanical injury. The brain

weighs about 1,500 grams but, suspended in CSF, its buoyed weight is about 50 grams.

CSF fills the ventricles, the subarachnoid space of the brain and spinal cord, and the

central canal of the spinal cord. According to Seely [71], approximately 23 mL of fluid

fills the ventricles, and 117 mL fills the subarachnoid space.

The CSF flowing in the subarachnoid space is of some importance in head MIT

imaging. Due to its vicinity to the surface of the head, and thus to the exciter/sensor

coils, and its relatively higher conductivity (the highest of all tissues as explained in the

next chapter), CSF can have a dominant contribution to the induced MIT signals. This

can make it more difficult to reconstruct an image of a less conductive feature located

deep in the brain (an example is presented in chapter 6).

3.2.7. Spinal cord

The spinal cord is a cylinder of nervous tissue that begins at the base of the brain and

extends through a large opening in the skull (figure 3.7a). It consists of centrally located

grey matter, and peripherally located white matter. The relative size and shape of the

grey and white matter varies throughout the spinal cord [71]. The core of grey matter

roughly resembles the letter H (figure 3.7b). Within the grey commissure is the central

canal. It is continuous with the ventricles of the brain and is filled with CSF. In this

study, only the neck portion of the spinal cord, which is expected to contribute to the

MIT signals due to its vicinity to the coils, is of interest.

Page 52: Magnetic Induction Tomography for Imaging Cerebral Stroke

31 Anatomy of Head and Stroke Pathology

Figure 3.7. Spinal cord: (a) illustration of entire length, (b) drawing of a segment (after Seely [71]).

Figure 3.8. Illustration of eye balls and optic nerves (after Mader [65]).

3.2.8. Eyes and optic nerves

The visual system includes the eye balls and the optic nerves. The eye ball is made up

of three layers, enclosing three transparent structures. The outermost layer is composed

of the cornea and sclera. The middle layer consists of the choroid, ciliary body, and iris.

The innermost is the retina. Within these layers are the aqueous humor, the vitreous

humor, and the flexible lens [65]. The optic nerves transmit visual information from the

retina to the brain and are considered to be part of the CNS (figure 3.8).

Despite their relatively small dimensions, optic nerves are particularly important in head

EIT imaging. They can be used as a pathway for current injection or as electric potential

(a) (b)

optic nerve

Page 53: Magnetic Induction Tomography for Imaging Cerebral Stroke

32 Anatomy of Head and Stroke Pathology

measurements site for EIT data acquisition [74]. The reason why eyes could be used for

this purpose arises from the topology of the brain and skull. The optic nerve connects

the eye with the related occipital cortex without being interrupted by the skull and the

CSF shunting effect. It is therefore a relatively conductive pathway for the current to

flow through. In MIT imaging, however, the magnetic field can easily penetrate the

electrically-resistive skull and, therefore, the conductive pathway provided by the optic

nerves will not be as important as in EIT.

3.2.9. Cavities

As described by Mader [65], there are three large air-filled cavities in the head; nasal

cavity and two ear canals. The nasal cavity or nasal fossa is a large air-filled space

above and behind the nose in the middle of the face (see figure 3.9a and figure 3.2a).

The nose contains two nasal cavities, separated from one another by a septum composed

of bone and cartilage. The ear canal (also external auditory meatus), is a tube running

from the outer ear to the middle ear (figure 3.9b).

3.2.10. Other tissues

There are other tissues and organs such as lips, teeth, skin, and hair which are assumed

to have negligible contribution to the induced MIT signals; due to their low conductivity

and/or their small volumes. Furthermore, the brain comprises some small sub-tissues

which, in terms of the structure, are similar to white and grey matter tissues. In this

study, for simplicity in the head model, they were not accounted for as separate tissues,

but were represented as white or grey matter. More details are provided in table 4.1 in

chapter 4.

Figure 3.9. Air-filled cavities in head; (a) nasal cavity, (b) ear canal (after Mader [65]).

(a) (b)

auditory canal

olfactory

bulb

nasal

cavity

Page 54: Magnetic Induction Tomography for Imaging Cerebral Stroke

33 Anatomy of Head and Stroke Pathology

3.3. Stroke

A cerebro-vascular accident (CVA) or stroke is the rapidly developing loss of brain

function due to disturbance in the blood supply to the brain which deprives the brain

tissue of oxygen. This can be due to ischaemia (lack of blood supply) or due to a

haemorrhage (bleeding). As a result, the affected area of the brain is unable to function.

A stroke is a medical emergency and can cause permanent neurological damage,

complications, and death. It affects between 174 to 216 people per 100,000 population

in the UK each year [75], and accounts for 11% of all deaths in England and Wales

[76]. It is the major cause of serious disability and the third leading cause of death in the

United States [77,78].

3.3.1. Classification of stroke

Strokes can be classified into two major categories: ischaemic and haemorrhagic.

Ischaemia is due to the interruption of the blood supply. Haemorrhage is due to the

rupture of a blood vessel or an abnormal vascular structure. Therefore, haemorrhage and

ischaemia are opposites; haemorrhage is characterized by too much blood inside the

skull, while in ischaemia there is not enough blood supply to allow continued normal

functioning of the effected brain tissue. Brain ischaemia is much more common than

haemorrhage. Approximately 69-85% of the strokes are due to ischaemia, and the

remainder are due to haemorrhage [76,77,79].

In ischaemia, the interruption of the blood supply may be caused by embolism or

thrombosis. An embolism occurs when an object (the embolus) migrates from one part

of the body (through circulation) and causes a blockage (occlusion) of a blood vessel in

another part of the body (figure 3.10a). Thrombosis is the formation of a blood clot

(thrombus) inside a blood vessel, obstructing the flow of blood through the circulatory

system (figure 3.10b).

Bleeding (figure 3.10c) can have many different causes such as head injury or

hypertension. Caplan [79] provides a comprehensive list of common causes of

haemorrhage. There are several different subtypes of haemorrhage, named for their sites

inside of the skull. An intra-cranial haemorrhage is a bleeding within the skull which

includes two major types; intra-axial and extra-axial (see figure 3.11).

Page 55: Magnetic Induction Tomography for Imaging Cerebral Stroke

34 Anatomy of Head and Stroke Pathology

Figure 3.10. Stroke; ischaemia due to embolism (a) and thrombosis (b), haemorrhage due to

rupture of a blood vessel (c) (after Broderick and Santilli [77]).

Figure 3.11. Diagram showing different types of haemorrhage based on its location within head.

Intra-axial or cerebral

haemorrhage

(bleeding within brain)

Extra-axial

(bleeding within skull but

outside brain)

Intra-parenchymal

(bleeding within brain

tissue)

Intra-ventricular

(bleeding within

ventricles)

Epidural

Subdural

Subarachnoid

Intra-cranial haemorrhage

Page 56: Magnetic Induction Tomography for Imaging Cerebral Stroke

35 Anatomy of Head and Stroke Pathology

Figure 3.12. Examples of intra- and extra-axial haemorrhages (after Caplan [80]).

Intra-axial haemorrhage is bleeding within the brain itself, or cerebral haemorrhage.

This category includes intra-parenchymal haemorrhage, or bleeding within the brain

tissue forming a gradually enlarging haematoma, and intra-ventricular haemorrhage,

bleeding within any of the brain's ventricles (lateral, third or fourth ventricles).

Extra-axial haemorrhage, bleeding that occurs within the skull but outside of the brain

tissue, falls into three subtypes (see figure 3.12). Epidural haemorrhage, which occurs

between the dura mater (the outermost meninges) and the skull, is caused by trauma. It

is the least common type of meningeal bleeding and is seen in 1% to 3% cases of head

injury. Subdural haemorrhage results from tearing of the bridging veins in the subdural

space between the dura and arachnoid mater. Subarachnoid haemorrhage occurs

between the arachnoid and pia meningeal layers.

3.3.2. Diagnosis and treatment

If stroke is suspected, a prompt and accurate diagnosis and treatment is necessary to

minimise the brain tissue damage. An ischaemic stroke may be dealt with using

thrombolytic drugs. Thrombolytic drugs are used in medicine to dissolve blood clots in

a procedure termed thrombolysis and can limit the damage caused by the blockage of

the blood vessel. However, according to the guidelines [76,81], thrombolytic treatment

can be helpful provided that:

Page 57: Magnetic Induction Tomography for Imaging Cerebral Stroke

36 Anatomy of Head and Stroke Pathology

it is administered within three hours of onset of stroke symptoms

possibility of a haemorrhage has been definitively excluded

It should be noted that thrombolytic therapy in haemorrhagic strokes is contraindicated,

as its use in that situation would prolong bleeding into the intracranial space and cause

further damage. Therefore, a quick classification of the stroke, i.e. distinguishing

between ischaemia and haemorrhage, is of crucial importance.

Apart from the physical examination, stroke can be diagnosed through several imaging

techniques [82]: computed tomography (CT), magnetic resonance imaging (MRI),

doppler ultrasound, and arteriography. Among these, CT and MRI scans are the most

conventional and reliable techniques, and their sensitivity and specificity have been the

subject of various studies.

For instance, Chalela et al [83] compared the performance of CT and MRI in a

consecutive series of patients referred for emergency assessment of suspected stroke.

They concluded that MRI is more effective than CT for detection of ischaemia; MRI

detected stroke (ischaemic or haemorrhagic) more frequently than did CT (see table

3.1). On the other hand, CT is less affected by patient motion than conventional MRI

and also allows full access to a patient during scanning, making it of use in restless or

critically ill patients [84].

CT and MRI images allow doctors to separate brain haemorrhages from infarctions1.

They can also show whether bleeding has occurred within the skull, but outside of the

brain in the subarachnoid, subdural, or epidural spaces. Haemorrhages and infarcts

appear with different CT numbers2 on CT scans, making it quite easy to separate the

two stroke types.

Table 3.1. Performance of CT and MRI for diagnosing stroke; values in percentage [83].

MRI CT

Sensitivity Specificity Sensitivity Specificity

Haemorrhage 81 100 89 100

Ischaemia 83 98 16 96

1 Tissue death due to ischaemia is called infarction. 2 A normalised value of the calculated X-ray absorption coefficient of a pixel (picture element) in a

computed tomogram, expressed in Hounsfield units, where the CT number of air is -1000 and that of

water is zero.

Page 58: Magnetic Induction Tomography for Imaging Cerebral Stroke

37 Anatomy of Head and Stroke Pathology

Figure 3.13a is a CT scan showing a large brain infarct involving the right cerebral

hemisphere. The yellow arrow points to the infarct, which appears darker than the

surrounding brain tissue. Figure 3.13b is an MRI scan showing a small brain infarct

within the left cerebral hemisphere. The infarct appears much whiter than the

surrounding normal brain.

In contrast, the CT scan in figure 3.13c shows a brain haemorrhage. The region of

bleeding appears whiter than the surrounding tissue. The arrow points to the area of

bleeding. Finally, an MRI image from the same haemorrhage is shown in figure 3.13d.

Figure 3.13. Stroke images: (a) CT scan showing a large cerebral infarct, (b) MRI scan showing a

left cerebral infarct, (c) CT scan showing a haemorrhage in deep portion of brain, (d) MRI scan showing a brain haemorrhage similar to that in (c) (after Caplan [79]).

(a) (b)

(c) (d)

CT MRI

Hae

mo

rrhag

e

Infr

act

(Isc

hae

mia

)

Page 59: Magnetic Induction Tomography for Imaging Cerebral Stroke

38 Anatomy of Head and Stroke Pathology

Figure 3.14. CT scan of a haemorrhage worsened by thrombolytic therapy (after Harold and

Adams [85]).

Figure 3.14 shows a CT scan of a brain haemorrhage in a patient who had received

inappropriate treatment with thrombolytic therapy, resulting in a gross worsening of the

haemorrhage in the left hemisphere of the brain. This illustrates the importance of

accurate early diagnosis.

As can be seen from figure 3.13 and figure 3.14, MRI and CT scans provide excellent

spatial resolution and tissue contrast. However, both modalities involve large,

immobile, and expensive scanners, which are sometimes inaccessible within the 3-hour

window1.

Although MIT will probably never achieve the spatial resolution of MRI and CT, our

research goal is to develop it into a low-cost and portable system, that could possibly

even be carried in ambulances to allow a rapid initial diagnosis to be made.

End of chapter 3.

1 Note that the 3-hour time window is the time when drug must be administrated, and the imaging must be

conducted at a shorter time followed by the diagnosis and assessment.

Page 60: Magnetic Induction Tomography for Imaging Cerebral Stroke

Chapter 4.

Dielectric Properties of Head Tissues and

Pathophysiology

4.1. Introduction

The purpose of this chapter is to review the dielectric properties of the head tissues

introduced in the previous chapter. Each biological tissue has a distinct frequency

dependence in its passive dielectric properties, which results in a distinct spectroscopic

signature. Therefore, it should be possible to determine the tissue type by this signature.

Furthermore, various pathophysiological mechanisms can alter the biochemical

composition and structure of the cells and their environment and, therefore, result in a

change in the spectroscopic behaviour. As a result of these changes, differentiation

between various tissues‟ conditions by means of bioimpedance measurements may be

possible.

A knowledge of these mechanisms and their impact on the dielectric properties of the

head tissues is important for development of the head MIT imaging technique. For

modelling purposes, this information enables assignment of appropriate dielectric

properties to the tissues involved in the lesion, thereby estimating the expected change

in MIT signals due to the local changes introduced by the pathology. These model-

based predictions can assist in studying the feasibility of MIT applications, such as

distinguishing between ischaemic and haemorrhagic strokes.

Page 61: Magnetic Induction Tomography for Imaging Cerebral Stroke

40 Dielectric Properties of Head Tissues and Pathophysiology

A brief outline of the basic concepts of dielectric theory and terminology, and various

relaxation mechanisms (dispersions) is provided in Appendix B. In this chapter, a

literature review of the measured or estimated dielectric values of different head tissues

is carried out. The associated dielectric property changes resulting from the abnormal

conditions due to occurrence of the stroke are explained. Finally, the corresponding

values adopted in this study are presented in the end.

4.2. Dielectric properties of head tissues

Research into the dielectric properties of biological materials and their variation with

frequency has been ongoing for most of the past century. In the last few decades, the

research has been driven by the need to establish a credible database of dielectric

properties of all body tissues for use in electromagnetic dosimetry studies, where the

object is to quantify the exposure of people to external electromagnetic fields from the

knowledge of effective internal fields and currents induced in them. In these studies,

tissues are characterized by their measured dielectric properties.

One of the main problems one encounters using bioimpedance measurements is the

reliability of the results. The scatter of the data for the electrical parameters of tissues

indicates the problem of measurement reproducibility. Some of the scatter may result

from problems in measurement techniques, but some may simply be caused by the

individual variability among samples. At present, the relative importance of these two

factors in determining the overall scatter is not clear [86]. This scatter makes it difficult

to establish criteria of normality or reference value for particular measurement results.

However, in spite of these differences between reported data on dielectric properties,

some estimation of the dielectric values at frequencies of interest can still be reached.

Large differences exist in electric properties between different biological tissues. These

differences are determined, to a large extent, by the fluid content of the tissue. For

example, blood and brain conduct electric current relatively well. Skin, fat, and bone are

relatively poor conductors. Moreover, some tissues, such as bone and skeletal muscle,

are distinctly anisotropic. Therefore, when referring to published conductivity and

permittivity values, care must be taken in the orientation of the measuring electrodes

Page 62: Magnetic Induction Tomography for Imaging Cerebral Stroke

41 Dielectric Properties of Head Tissues and Pathophysiology

relative to the major axis of the tissue (e.g., longitudinal, transverse, or a combination of

both).

In the last decade, most dosimetric studies drew on data published in the scientific

literature in 1996 and made widely available on the Internet thereafter. In this chapter,

the published data is reviewed in three subdivisions; literature before 1996, the 1996

database, and literature after 1996. When available, details of tissue sample,

measurement techniques, and measuring environment are mentioned. This is because

the measurements are known to be affected by factors such as temperature, mammalian

species, subject's age, the particular electrode arrangement used, and whether the

measurement was performed in-vivo, in-situ after death, or in-vitro.

4.2.1. Literature before 1996

In 1996, Gabriel et al [87] (hereinafter as „Gabriel-I‟) conducted a comprehensive

review, extracting the dielectric properties of tissues from the literature of the past five

decades (from 1950 to 1996), and presented the results in a graphical format. The data

reported were those that corresponded most closely to living human tissues. Therefore,

human tissue and in-vivo measurements were selected in preference to animal tissue and

in-vitro. For in-vitro measurements, data obtained at temperatures closest to that of the

body and nearest to the time after death were used. Later, they made the data available

in tabular format [88].

In this study, regardless of source of the data and measurement details, the tabulated

values from Gabriel-I are plotted as „Pre-1996‟ data in figure 4.1 to figure 4.7. Each

figure, corresponding to one tissue type, will be dealt with separately in section 4.2.3.

The 'Pre-1996' data is re-visited in this study, for comparison with those published

thereafter.

4.2.2. The 1996 database

The backbone of the 1996 database is a large experimental study, conducted by Gabriel

at al [89] themselves (hereinafter as „Gabriel-II‟), providing data pertaining to human

and animal tissues at 37 C. A comparison was made between the dielectric properties of

human tissue (autopsy material) and its equivalent in one or more animal species

Page 63: Magnetic Induction Tomography for Imaging Cerebral Stroke

42 Dielectric Properties of Head Tissues and Pathophysiology

(mostly ovine). All animal tissues used were as fresh as possible, mostly within 2 hours

of the animal‟s death. Human material was obtained 24 to 48 hours after death.

Comparisons were also made between measurements carried out in-vivo on accessible

parts of the body and in-vitro on freshly excised tissue. For most tissues, the

characterization was over a wide frequency range, 10 Hz to 20

GHz, using three

previously established experimental setups with overlapping frequency ranges (10 Hz to

10 MHz, 300 kHz to 3 GHz, 130 MHz to 20 GHz). Each frequency range was covered

by a different measurement setup and it was observed that the agreement among

measurements on the three machines was particularly good at the overlapping frequency

ranges. The data obtained in Gabriel-II were compared to those presented in the context

of a review in Gabriel-I. By and large, the experimental data were well within the

confines of corresponding values from the literature.

In a follow up study by Gabriel et al [90] (hereinafter as „Gabriel-III‟), the dielectric

spectra were parameterised using a multi-dispersion model consisting of four Cole–Cole

terms and one ionic conductivity term (equation (B.8) in Appendix B, n =

4). This was

done based on the data from their two previous studies. For each tissue, the parameters

of the model enable the reconstruction of its spectrum; a procedure that can be

incorporated in the numerical studies to provide dielectric data that are broadly in line

with the vast body of literature on the subject.

While useful, the 1996 database has several limitations, as pointed out by its authors, of

which two are particularly relevant at frequencies below 1 MHz:

Most measurements were carried out on excised tissue, while data pertaining to

live tissue would have been more relevant in bio-electromagnetics studies.

For most tissues, the predictions of the model can be used with confidence for

frequencies 1 MHz because of the availability of supporting data in the

literature.

At lower frequencies, where the literature values are scarce and have larger than

average uncertainties, the model should be used with caution in the knowledge

that it provides a „„best estimate‟‟ based on the available knowledge. This is

particularly important for tissues where there are no data to support its

predictions.

Page 64: Magnetic Induction Tomography for Imaging Cerebral Stroke

43 Dielectric Properties of Head Tissues and Pathophysiology

Electrode polarization, an inevitable source of error at low-frequencies, was not

totally accounted for. It affects the data at frequencies below 100 Hz.

Because of the geometry of the sampling probe, the electric field was not

oriented in appropriate directions, making it difficult to demonstrate the

anisotropy of the dielectric properties.

The values of measured (Gabriel-II) and modelled (Gabriel-III) dielectric properties are

plotted in figure 4.1 to figure 4.7. Each figure, corresponding to one tissue type, will be

dealt with separately in section 4.2.3.

4.2.3. Literature after 1996

Although comparisons were made between the parameterised data and those from

practical measurements in 1996 and before, in order to include the most up to date

values available, a review covering the reported data since 1996 is now provided. This

serves to extend the very comprehensive work of Gabriel-I and Gabriel-II. The review is

carried out for five main tissues in the human head; i.e. muscle, brain, skull, CSF, and

blood.

Most of the literature data were in graphical rather than tabular form, and in a

logarithmic rather than linear format; such data were retrieved from the graphs. The data

were translated from the various authors‟ preferred set of parameters and units to

relative permittivity and conductivity expressed in S m

−1. In some cases, data from

recent studies are compared with the model in the 1996 database. The data are presented

in a graphical format, in order to highlight the information with respect to the frequency

coverage, and the scatter in the data. The data of the reports giving only one or two

values are not displayed in the figures, but mentioned in the review. In each figure,

details of the tissue, measurement temperature, and the reference are included in the

legend, identified by the first author's name and date of report.

In this thesis, MIT imaging in a particular range of frequency (1-10 MHz) is of interest,

being the operational frequency range of the MK1 and Mk2a systems. However,

literature pertaining to a wider frequency range is covered. The reason for that is

twofold: (i) for some tissues, the reported data in recent studies is scarce in the

mentioned frequency range and, therefore, looking at other frequencies can help to

scrutinise the 1996 database as a whole, (ii) MIT researchers are designing systems

Page 65: Magnetic Induction Tomography for Imaging Cerebral Stroke

44 Dielectric Properties of Head Tissues and Pathophysiology

operating at lower frequencies (down to 100 kHz) and, consequently, the dielectric

properties for a wider range of frequencies can be employed.

Muscle: the following 'Post-1996' data for muscle, presented in this section, is

displayed in figure 4.1 along with the 'Pre-1996', Gabriel-II, and Gabriel-III data.

Muscle tissue can be skeletal, myocardial, lingual, or other. However, we only consider

skeletal muscles which encompass the skull. Muscles are composed of fibres that are

very large individual cells and are aligned in the direction of muscle contraction (see

figure 3.1a). Electrical conduction along the length of the fibre is thus considerably

easier than conduction between the fibres in the extra-cellular matrix because the extra-

cellular matrix is less conductive than the cell. The -dispersion is more prominent and

the -dispersion is less defined in the longitudinal direction. This is in accordance with

the predictions of effective permittivity modelling of elongated structures [91].

Therefore, muscle tissue manifests typical anisotropic electric properties.

The longitudinal conductivity is higher than the transverse conductivity, even when path

differences in the charge transport are taken into account, especially in the low

frequency range. Moreover, tissue anisotropy is frequency-dependent [86]. If the

frequency of the current is high enough, the anisotropic properties disappear. For

muscle tissue this happens in the MHz frequency range. At higher frequencies, charge

movement takes place over shorter distances thus large-scale structures become less

important and capacitive coupling across membranes becomes more important.

Among different muscle types, data on skeletal muscle is the most abundant in the

literature. As a result of the anisotropy of this tissue, the data are usually presented

separately for the transverse and longitudinal directions, although some results with

random orientation have been reported. In particular, low-frequency data are very

scarce, and there are differences in what is available in the literature, partly due to the

fact that many authors do not specify the measurement orientation. The situation is not

helped by the fact that the apparent anisotropy depends on the measurement procedure,

in particular the inter-electrode distance in relation to the size of the muscle fibre [91].

Hart et al in 1999 [92] undertook 2-terminal resistivity in-vivo measurements of the

bullfrog muscle in the frequency range of 1 kHz to 1 MHz. In addition to the analytical

Page 66: Magnetic Induction Tomography for Imaging Cerebral Stroke

45 Dielectric Properties of Head Tissues and Pathophysiology

anisotropic models, they also approximately calculated the contribution of each

direction. Based on a finite difference model, they calculated the electrical resistance

between two electrodes, inserted longitudinally or transversely into the muscle, in terms

of longitudinal and transverse muscle conductivities. With their method, they confirmed

that the longitudinal conductivity is consistently higher than that of the transverse

direction. However, the difference in conductivity between these directions appeared

smaller than in previous publications, especially for the lower frequencies considered.

Gabriel et al in 2009 [93] carried out measurements on pig tissue, in-vivo, at eight

frequencies in the range between 10 Hz to 1 MHz. Conductivity data was obtained for

skeletal muscle from average values from measurement on samples from at least three

pigs. They did not observe the expected anisotropy in electrical conductivity and values

were lower than the 1996 database. They suggested that one possible explanation was

the lack of rigorous correspondence between the geometry of the measurement probe

and the structure of the tissue.

In a study by Peyman et al in 2001 [94], the dielectric properties of ten rat tissues at six

different ages (0-70 days) were measured at 37 C in the frequency range of 130 MHz

to 10 GHz. All measurements were performed within 2-4 hours of the animal's death.

Their results showed a general decrease of the dielectric properties with age which was

attributed to the changes in the water content and the organic composition of tissues.

Values shown in figure 4.1 are the average values taken from all ages.

Brain tissues: the following 'Post-1996' data for grey and white matter, presented in

this section, is displayed in figure 4.2 and figure 4.3 along with the 'Pre-1996', Gabriel-

II, and Gabriel-III data.

Many studies have modelled the entire brain as a single homogeneous domain, for the

sake of geometric simplicity (e.g. [95]), and several others attempted to provide an

estimation for the impedance of the whole brain, as the ratio between the impedance of

scalp, skull and brain (e.g. [96]). However, the fidelity of such approximations is

doubtful, particularly for the inner regions of the head. Considering the whole brain as

one homogeneous compartment does not account for the fact that white matter is more

Page 67: Magnetic Induction Tomography for Imaging Cerebral Stroke

46 Dielectric Properties of Head Tissues and Pathophysiology

resistive than the grey matter. Furthermore, grey and white matter may be affected by

brain pathologies in different ways1.

Latikka et al in 2001 [97] measured the resistivity of brain tissues in-vivo in 9 patients

with brain tumours during surgery with a monopolar needle at a single frequency of 50

kHz. White and grey matter had mean values of 3.91 and 3.51 m (conductivities of

0.25 and 0.28 S m

-1), respectively. Their measurements were subject to several sources

of error; the bleeding, the saline used for washing, and the leakage of CSF decreased the

resistivity of the measured brain tissues. On the other hand, the increase in the

intracranial pressure caused by the tumour increased the resistivity values. The authors

noted that the measured values differ from those given by Gabriel-I (figure 4.2 and

figure 4.3). They ascribed the differences to the fact that their measurements were done

in-vivo, thus more reliable.

Data by Peters et al in 2001 [98] were obtained by modelling the brain cortex as a

suspension of randomly distributed elongated cells suspended in CSF. Assuming the

extra-cellular volume fraction to be between 0.18 (for a rat in-vivo) and 0.28 (for a pig

in-vitro) and the conductivity of the CSF to range between 1.5 and 1.8 S m

−1, the

effective conductivity of the mixture fell to within the range 0.1-0.27 S m

−1.

In a human study by Schmid et al in 2003 [99], the dielectric properties of grey matter

in the frequency range of 800-2450 MHz were measured on 20 human brains

immediately after excision, less than 10 hours after death. The tissue temperature at the

measurement sites ranged between 18 and 25 C, and 160 different measurements were

performed on each brain. The measurements yielded a mean value (standard deviation)

of grey matter equivalent conductivity of 1.13 and 2.09 S m

-1 at 800 and 2450 MHz,

respectively. The mean value of measured relative permittivity was 58.2 and 54.7 at 800

and 2450 MHz, respectively.

Peyman et al [94] reported variation in the dielectric properties of rat brain tissue as a

function of age, at microwave frequencies. Their data pertained to the whole brain. The

observed variation was ascribed, at least partially, to the change in the ratio of grey to

white matter, which is known to occur throughout the developmental stage.

1 Blood flow in white matter is lower than in grey matter, and white matter ischaemia is typically severe,

because there is little collateral blood supply in deep white matter. Moreover, cells within the grey and

white matter have different susceptibilities to ischaemic injury [231].

Page 68: Magnetic Induction Tomography for Imaging Cerebral Stroke

47 Dielectric Properties of Head Tissues and Pathophysiology

In another study on porcine tissues [100], they were able to investigate grey and white

matter separately. In this case, no variations were observed in the dielectric spectrum of

grey matter, while statistically significant variations were observed in the dielectric

spectrum of white matter. They suggested that the observed variations were probably

related to the process of myelination1, which begins at birth and lasts to maturation. In

another study in 2007 [101], the dielectric properties of pig cerebrospinal tissues were

measured in-vivo, in the frequency range of 50 MHz - 20 GHz. Comparisons were made

for the measured values of white and grey matter with the 1996 database, and both

conductivity and permittivity values were found to be in good agreement.

In a recent study in 2009 [102], dielectric properties of ageing porcine tissues were

measured, in-vitro, in the frequency range of 50 MHz - 20 GHz. Their results showed

considerable reduction with age in both permittivity and conductivity of white matter,

where dielectric properties of grey matter remained unchanged. In figure 4.2 and figure

4.3, measured values are shown for 50 kg pigs.

Bao et al in 1997 [103] presented in-vitro complex dielectric measurements of grey and

white matter of rat brains in the frequency range between 45 MHz and 26.5 GHz at

body and room temperatures (24 and 37 C). The measurement data exhibited two

separated dispersions and they suggested an empirical model containing two Cole-Cole

functions to describe the experimental data. In figure 4.2, the measured data at 37 C is

shown.

Bone: the following 'Post-1996' data for bone, presented in this section, is displayed in

figure 4.4 and figure 4.5 along with the 'Pre-1996', Gabriel-II, and Gabriel-III data.

Bone, cortical or cancellous, has an anisotropic structure and, in consequence,

anisotropic dielectric properties. For cortical bone, the ratio of the conductivity in the

axial and radial dimensions is about 3.2 [93]. Cancellous bone is less anisotropic; the

conductivity is higher in the longitudinal direction compared to the lateral and anterior–

posterior directions. While the skull impedance is not affected by brain pathology or

functional activity, the resistivity of the skull affects the EIT measurements as it acts as

1 The change or maturation of certain nerve cells whereby a layer of myelin forms around the axons

which allows the nerve impulses to travel faster.

Page 69: Magnetic Induction Tomography for Imaging Cerebral Stroke

48 Dielectric Properties of Head Tissues and Pathophysiology

a current barrier. However in MIT imaging, conductivity of the skull is of less

importance as the magnetic field can penetrate the electrically-resistive bones.

Oostendorp et al [104] performed both in-vivo and in-vitro 4-terminal skull

measurements in the frequency range 100 Hz to 10 kHz. The in-vitro measurement was

performed on a sample of fresh skull placed within a saline environment in which the

temperature was maintained at 37 C. For the in-vivo measurement, a small current was

passed through the head by means of two electrodes placed on the scalp. The potential

distribution thus generated on the scalp was measured in two subjects for two locations

of the current injecting electrodes. Both methods revealed a skull conductivity of about

0.015 0.003 S m

-1 for the mentioned frequency range.

The dependence of the conductivity on the position and bone type was investigated by

Akhtari et al in 2002 [105], during in-vivo measurements using 4-terminal technique at

10-90 Hz on skull flaps, excised from 4 patients during surgery, with a 15-65 minutes

delay between excision and measurement. Holes were drilled in the samples to allow the

resistivity to be measured at the different layers in the bone sample. The results

indicated that the conductivities of the spongiform and compact layers of the skull have

significantly different and inhomogeneous conductivities.

A year later in 2003, the frequency dependence of the conductivity of different layers

and the bulk was measured and modelled in the same frequency range [106]. The

samples were cut from temporal, temporal-parietal and inferior occipital regions of the

skull (these regions are shown in figure 3.2).

Further conductivity measurements were made in 5 temporal bone samples temporarily

removed during epilepsy surgery by Hoekema and Wieneke in 2003 [107]. The

conductivity values were about 10 times higher than any previously published data. This

difference could be due to the fact that the measurements were made a very short time

after excision, while other samples have been dried and then re-soaked. The high

conductivity could also be due to a thin layer of saline on the surface of skull.

Sierpowska et al in 2003 [108] demonstrated the dependency of volumetric trabecular

bone mineral density and its relative permittivity. The dielectric and mechanical

properties of femoral bovine trabecular bone samples from various regions were

measured. The bones were frozen and defrosted prior to the measurement procedure.

Page 70: Magnetic Induction Tomography for Imaging Cerebral Stroke

49 Dielectric Properties of Head Tissues and Pathophysiology

The frequencies were between 50 Hz and 5 MHz, and the conductivity values ranged

from 0.02 to 0.06 S m−1

depending on the site. This compares with 0.07 S m−1

for ovine

cancellous bone in the 1996 database.

Later in 2005 [109], they measured the impedance of cadaver human trabecular knee

bone in the same frequency range. Samples of tibia and femoral trabecular bone were

defrosted to room temperature (22 C) after being kept frozen for 12 months. While

there were no significant differences between the tibia and the femoral bone throughout

the entire frequency range, human trabecular bone was found to be twice as conductive

as bovine. The authors have recently investigated the relationships between electrical

properties and microstructure of human trabecular bone [110,111].

Peyman et al in 2001 [94] observed variation in the dielectric properties of rat skull

bone as a function of developmental stage from neonate to 70 days old; no reference

was given to the type of the bone. In figure 4.4 and figure 4.5, the results are shown for

two frequencies (900 MHz and 1.8 GHz) as the average values of those with different

ages. Similar results were recently reported for porcine skull [100].

Gabriel et al in 2009 [93] carried out in-vivo measurements on the mid-sectional part of

the skull of 50 kg pigs: this part of the skull has flat areas which allow good contact

with the electrodes. The bone exposed after stripping the periostium membrane

appeared pitted and filled with red bone marrow. The measured conductivities, at eight

frequencies in the range of 10 Hz to 1 MHz, were higher than previously reported data

for bone. However, a recent in-vitro study, conducted by Peyman et al in 2009 [102],

showed a significant reduction with age in both permittivity and conductivity at 37 C.

The skull bone of 250 kg pigs (mature animals) had much lower conductivities. At the

frequency of 450 MHz, the skull conductivity values for 10, 50, and 250 kg pigs were

0.62, 0.52, and 0.19 S m

-1, respectively.

CSF: the following 'Post-1996' data for CSF, presented in this section, is displayed in

figure 4.6 along with Gabriel-II and Gabriel-III data (there was no 'Pre-1996' data

reported for CSF).

As an ionic fluid containing no cell membranes, CSF is expected to have a high

conductivity. Baumann et al in 1997 [112] measured the properties of human CSF using

Page 71: Magnetic Induction Tomography for Imaging Cerebral Stroke

50 Dielectric Properties of Head Tissues and Pathophysiology

the 4-terminal technique in a Perspex measuring cell at 10 Hz - 10 kHz. The samples

were frozen and then defrosted. They compared the measured values at room

temperature (25

C) with those obtained at body temperature (37

C). The measured

conductivity values at body temperature (1.45 S m

-1) were in better agreement with

previous data than those at room temperature (1.8 S m

-1).

Latikka et al in 2001 [97] measured the CSF resistivity in-vivo in 2 patients with brain

tumours during surgery with a monopolar needle at a single frequency of 50 kHz; it had

a mean value of 0.80 m (conductivity of 1.25 S m

-1).

Peyman et al in 2007 [101] measured the dielectric properties of pig CSF, in-vitro, in

the frequency range of 50 MHz - 6 GHz. Gabriel et al in 2009 [93] carried out

conductivity measurements on pig tissue, in-vitro, at different frequencies in the range

of 10 Hz to 1 MHz. This study was designed to obtain data at frequencies below 1 MHz

to supplement the 1996 database. The results of both studies were in good agreement

with Gabriel-III, albeit the conductivity values were slightly smaller (see figure 4.6).

Blood: the following 'Post-1996' data for blood, presented in this section, is displayed in

figure 4.7 along with the 'Pre-1996', Gabriel-II, and Gabriel-III data.

Blood is a circulating tissue composed of fluid plasma and cells. The plasma, which

comprises 55% of blood fluid, is mostly water (90% by volume). Apart from the

plasma, the blood consists of several kinds of corpuscles; about 96% of them are red

blood cells (erythrocytes), 3% are white blood cells (leukocytes), and the remainder are

platelets (thrombocytes). In haemolysed blood, erythrocytes are disrupted and their

intra-cellular material is discharged into the liquid. Therefore, the electrical properties

of whole blood and haemolysed blood are naturally very different.

Whole blood exhibits - and -dispersion, but no -dispersion [113]. The -dispersion

has a dielectric increment of about 2000, and an accompanying rise in conductivity by

0.3 S m-1

, centred around 3 MHz (haematocrit1 40%). Erythrocytes in suspension have

a frequency independent membrane capacitance with very low losses. The impedance of

lysed erythrocytes in suspension shows two clearly separated single relaxation

1 Haematocrit is the volume proportion of blood occupied by red blood cells.

Page 72: Magnetic Induction Tomography for Imaging Cerebral Stroke

51 Dielectric Properties of Head Tissues and Pathophysiology

frequencies (Debye dispersions as described in Appendix B). As will be used in later

chapters, the mentioned increment in dielectric properties of blood, centred around 3

MHz, can be exploited in frequency-differential MIT.

The -dispersion is in the lower kHz range, and the -dispersion is in the lower MHz

range. Erythrocytes are poor conductors of electric current at low frequency compared

with the extra-cellular medium or plasma which surrounds them. The conductivity of

the blood has almost linear dependence on haematocrit, so the concentration of

erythrocytes critically determines the blood‟s resistivity.

Gabriel et al in 2009 [93] carried out blood conductivity measurements on pigs

weighing around 50 kg, at different frequencies in the range of 10 Hz to 1 MHz, and

temperature of 37 C. Blood was sampled in-vitro immediately after extraction by

immersing the probe in the blood. The measured conductivities were in good agreement

with previously reported data.

Jaspard et al in 2003 [114] studied the influence of haematocrit on the dielectric

properties of animal blood, cow and sheep, in the frequency range of 1 MHz to 1 GHz

and temperature of 37 C. They found that an increase in haematocrit level induced an

increase in the number of cell membranes with a decrease in the volume of plasma.

Similarly, when haematocrit increased, a drop of conductivity and an increase of

relative permittivity were observed. Figure 4.7 shows the measured dielectric values for

41 and 42% haematocrit in cow and sheep blood, respectively.

In another study [115], they investigated the temperature dependence of the electrical

parameters of human and animal blood in the same frequency range. The results

indicated a weak sensibility and a change of sign of the temperature coefficient of the

relative permittivity (about 0.3% C−1

at 1 MHz and -0.3% C−1

at 1 GHz). The

conductivity presented a more significant variation (of the order of 1% C−1

) over the

whole operating frequency range.

Page 73: Magnetic Induction Tomography for Imaging Cerebral Stroke

52 Dielectric Properties of Head Tissues and Pathophysiology

100

102

104

106

108

1010

1012

100

102

104

106

108

Perm

ittivity

Muscle

100

102

104

106

108

1010

1012

10-2

10-1

100

101

102

Frequency (Hz)

Conductivity

Figure 4.1. Dielectric properties

of muscle against frequency;

longitudinal and transverse

directions.

Conduct

ivit

y (

S m

-1)

Rel

ativ

e p

erm

itti

vit

y

Page 74: Magnetic Induction Tomography for Imaging Cerebral Stroke

53 Dielectric Properties of Head Tissues and Pathophysiology

100

102

104

106

108

1010

1012

100

102

104

106

108

Perm

ittivity

Grey Matter

100

102

104

106

108

1010

1012

10-2

10-1

100

101

102

Frequency (Hz)

Conductivity

Figure 4.2. Dielectric properties of

grey matter against frequency.

Conduct

ivit

y (

S m

-1)

Rel

ativ

e p

erm

itti

vit

y

Page 75: Magnetic Induction Tomography for Imaging Cerebral Stroke

54 Dielectric Properties of Head Tissues and Pathophysiology

100

102

104

106

108

1010

1012

100

102

104

106

108

Perm

ittivity

White Matter

100

102

104

106

108

1010

1012

10-2

10-1

100

101

102

Frequency (Hz)

Conductivity

Figure 4.3. Dielectric properties of

white matter against frequency.

Conduct

ivit

y (

S m

-1)

Rel

ativ

e p

erm

itti

vit

y

Page 76: Magnetic Induction Tomography for Imaging Cerebral Stroke

55 Dielectric Properties of Head Tissues and Pathophysiology

100

102

104

106

108

1010

1012

100

102

104

106

108

Perm

ittivity

Cancellous bone

100

102

104

106

108

1010

1012

10-2

10-1

100

101

102

Frequency (Hz)

Conductivity

Figure 4.4. Dielectric properties of

cancellous bone against frequency.

Conduct

ivit

y (

S m

-1)

Rel

ativ

e p

erm

itti

vit

y

Page 77: Magnetic Induction Tomography for Imaging Cerebral Stroke

56 Dielectric Properties of Head Tissues and Pathophysiology

100

102

104

106

108

1010

1012

100

101

102

103

104

105

Perm

ittivity

Cortical bone

100

102

104

106

108

1010

1012

10-3

10-2

10-1

100

101

Frequency (Hz)

Conductivity

Figure 4.5. Dielectric properties of

cortical bone against frequency.

Conduct

ivit

y (

S m

-1)

Rel

ativ

e p

erm

itti

vit

y

Page 78: Magnetic Induction Tomography for Imaging Cerebral Stroke

57 Dielectric Properties of Head Tissues and Pathophysiology

100

102

104

106

108

1010

1012

100

101

102

103

Perm

ittivity

CSF

100

102

104

106

108

1010

1012

100

101

102

Frequency (Hz)

Conductivity

Figure 4.6. Dielectric properties of

CSF against frequency.

Conduct

ivit

y (

S m

-1)

Rel

ativ

e p

erm

itti

vit

y

Page 79: Magnetic Induction Tomography for Imaging Cerebral Stroke

58 Dielectric Properties of Head Tissues and Pathophysiology

100

102

104

106

108

1010

1012

100

101

102

103

104

Blood

Perm

ittivity

100

102

104

106

108

1010

1012

10-1

100

101

102

Frequency (Hz)

Conductivity

Figure 4.7. Dielectric properties of

blood against frequency.

Conduct

ivit

y (

S m

-1)

Rel

ativ

e p

erm

itti

vit

y

Page 80: Magnetic Induction Tomography for Imaging Cerebral Stroke

59 Dielectric Properties of Head Tissues and Pathophysiology

4.3. Pathophysiology and conductivity change

The brain cells can be modelled as a group of three compartments: extra-cellular space,

intra-cellular space, and the cell membrane. Both the extra- and intra-cellular spaces are

highly conductive, because they contain salt ions [116]. The lipid membrane of cells is

an insulator, which prevents currents at low frequencies from entering the cells; see

figure 4.8 for normal condition at low frequencies. Therefore, almost all the current

flows through the extra-cellular space only. Due to the cell membrane permeability at

high frequencies, the current can cross the capacitance of the cell membrane, and so

enter the intra-cellular space as well.

Ischaemia: as discussed in chapter 3, ischaemic stroke occurs due to inadequate blood

supply to part of the brain, initiating the ischaemic cascade. Brain tissue ceases to

function if deprived of oxygen for more than 60 to 90 seconds and after a few hours will

suffer irreversible injury possibly leading to death of the tissue, i.e. infarction. The

ischaemic cascade is a series of biochemical reactions that take place in the brain and is

typically secondary to stroke. It results in cerebral oedema, i.e. swelling of the brain

cells as shown in figure 4.8. This is due to the leakage of large molecules from blood

vessels [117]. These large molecules pull water into the brain tissue after them by

osmosis. Oedema causes compression and damage to the brain tissue.

Figure 4.8. Current flow through cells at both low and high frequencies. Due to ischaemia and

subsequent oedema, brain cells swell resulting in impedance increase at low frequencies. ECS:

extra-cellular space (low-frequency and high-frequency diagrams are redrawn from Holder [116]

and Holder [118], respectively).

Normal (20% ECS) Ischaemia ( 15% ECS)

Current

Cells

normal & swollen

low

frequencies

high

frequencies

Page 81: Magnetic Induction Tomography for Imaging Cerebral Stroke

60 Dielectric Properties of Head Tissues and Pathophysiology

The extra-cellular space of grey matter is about 20%, and even a small change in its

volume results in a considerable impedance increase; average reduction of extra-cellular

space from the normal can be 4-10% [119]. This effect is most pronounced at low

frequencies as the current flows cannot penetrate the cell membrane.

In summary, the dielectric behaviour of ischaemic tissue can be explained as follows:

During ischaemia there is a shift in fluid volume from the extra-cellular space to

intra-cellular space. This leads to increases in impedance at the low frequencies

due to the capacitive nature of the cell membranes. Thus, there is an increase in

impedance in the - and -dispersion regions due to ischaemia.

The impedance at high frequencies (in or around the -dispersion region) is

relatively unaffected by ischaemia, tissue excision, and tissue death. This is

largely due to the fact that the existence of cellular membranes does not affect

the high frequency measurements.

The -dispersion breaks down much more quickly than the -dispersion

(typically within a few hours after the onset of ischaemia).

Based on the above evaluations of the three different dispersion regions, it may be

concluded that the - and/or -region would be the most suitable for investigating the

electrical properties of ischaemic tissue. A number of studies demonstrated the increase

in tissue resistivity after the onset of ischaemia. For instance, Van-Harreveld and Ochs

[120] measured cerebral impedance of rabbit brain using a 2-terminal method, during

circulatory arrest. The measurements were performed at frequency of 1 kHz. Five

minutes after circulatory arrest, there was a sudden large drop in conductivity, assigned

to the development of a cortical negativity. After this initial rapid decrease, the

conductivity continued to drop at a decreasing rate. After 4-5 hours, the conductivity

became only 20-30% of the original value. The reactance change was smaller than that

of the resistance in amplitude though similar in terms of percentage change.

Holder [121] measured in-vivo, in-situ cerebral impedance changes in rats during

ischaemia. Cortical and scalp electrode arrays were positioned in the frontal occipital

plane. Four-terminal measurements were performed at frequency of 51 kHz. Cortical

measurement showed an impedance increase of 50-200%, whilst the scalp

measurements resulted in 10% of that increase (i.e. 5-20%).

Page 82: Magnetic Induction Tomography for Imaging Cerebral Stroke

61 Dielectric Properties of Head Tissues and Pathophysiology

Figure 4.9. Current flow through cells. Due to haemorrhage, increase in conductive blood volume

results in impedance decrease. BV: blood volume (diagram redrawn from Holder [118]).

Haemorrhage: haemorrhagic strokes result in tissue injury by causing compression of

tissue from an expanding haematoma. Two pathophysiological mechanisms may occur

during haemorrhagic stroke:

a ruptured vessel causes a focal or diffused blood volume expansion, which

results in a small local impedance decrease due to the high conductivity of the

blood

a rapid ischaemia of the surrounding brain region due to a loss of cerebral blood

flow and increased intracranial pressure

Therefore, a haemorrhage may appear as an inner bleed, surrounded by the ischaemic

tissue. No measurement studies appear to have been published to directly quantify the

effect of haemorrhage on the dielectric properties of the brain tissues. The physiology

suggests that the inner part of the lesion will have a mixed impedance value of the

original brain tissue and blood according to their volume proportions [74]. An increase

in the conductive blood volume will allow more current to flow, resulting in a decrease

in the bulk impedance of the region affected by haemorrhage (see figure 4.9).

The extent of blood saturation depends on the severity of stroke. For moderate cases,

the whole extra-cellular fluid will be displaced by blood, which can account for 20%

volume in grey matter. In severe cases, the original tissue could be deformed out by the

blood and, therefore, the affected region will contain blood only; a number of MIT

modelling studies have replaced the whole brain tissue with blood (e.g. [46]). While the

inner part of the region will be saturated with blood to some extent, the surrounding

layer will comprise ischaemic matter. Further studies are required in order to accurately

quantify the dielectric changes expected in this pathology.

Normal (10% BV) Haemorrhage (>10% BV)

Blood Cells

Current

Page 83: Magnetic Induction Tomography for Imaging Cerebral Stroke

62 Dielectric Properties of Head Tissues and Pathophysiology

4.4. Can MIT be applied for stroke imaging?

For haemorrhagic stroke, two types of differential imaging in clinical MIT may be of

interest: time-difference and frequency-difference. The rationale for using time-

difference MIT is based on the fact that the conductivity of blood is higher than that of

normal brain tissues, except CSF. For instance, according to the data from section 4.2.3,

blood is between 3.6 and 6.9 times more conductive than grey and white matter at a

frequency of 10 MHz. Therefore, the larger induced eddy currents in the haemorrhagic

region and the resulting MIT signals shall be distinguishable from those of a normal

brain. In other words, time-difference imaging should be able to image the conductivity

rise in the brain from before to after the onset of a haemorrhage. This could in principle

be useful for long-term monitoring at the bedside for patients at risk of recurrence or

progression of a haemorrhage; see the example of haemorrhage growth within about 4.5

hours in figure 4.10.

Figure 4.10. CT scan showing haemorrhage growth in a patient with basal ganglion intra-cerebral

haemorrhage (after Adeoye and Broderick [122]).

According to Joseph [117], multiple factors contribute to the importance of monitoring

for such patients:

Many strokes occur while patients are sleeping (also known as "wake-up"

stroke) and are not discovered until the patient wakes.

Stroke can leave some patients too incapacitated to call for help.

Occasionally, a stroke goes unrecognized by the patient or their carers.

Page 84: Magnetic Induction Tomography for Imaging Cerebral Stroke

63 Dielectric Properties of Head Tissues and Pathophysiology

An MIT device employed for stroke monitoring could set off an alarm when some

abnormal signals are detected.

For an initial diagnosis, however, frequency-difference imaging will be necessary since

a „before-stroke‟ data frame will not be available. Multi-frequency imaging relies on

being able to distinguish the „frequency signature‟ of the blood from that of other

tissues, all of which will have frequency-dependent conductivities.

The conductivity increase in white and grey matter between 1 to 10 MHz is

approximately 0.06 S m

-1 (figure 4.3) and 0.13 S

m

-1 (figure 4.2), respectively, while

blood has a conductivity change of 0.3 S m

-1 (figure 4.7). By obtaining MIT

measurements at two frequencies, it should be possible to reconstruct an image of the

conductivity difference between the two frequencies.

It would be desirable for a clinical MIT system also to be able to detect ischaemic

stroke. This will be more difficult as the existing MIT systems are operated in the

frequency range of 1-10 MHz to increase the signal to noise ratio (see section 2.3.2). In

this frequency range, -dispersion remains unaffected by the ischaemia and, therefore,

no significant impedance change occurs. Consequently, normal and ischaemic brain

tissues are likely to produce similar MIT signals.

Extension of the range of biomedical MIT systems down to 100 kHz is a realistic goal

(e.g. system reported by Scharfetter et al [123], operating at 50 kHz - 1.5 MHz). This

may allow the impedance change in -dispersion region of the brain tissue to be

exploited for MIT imaging of an ischaemic region. Even so, an MIT system capable of

detecting just haemorrhagic stroke could still be of clinical value for excluding

haemorrhage prior to administration of drugs. In this study, therefore, the intention was

to investigate MIT imaging of haemorrhagic stroke.

Absolute imaging at a single frequency may also be feasible. However, as will be

shown in chapters 6 and 9, it will be more difficult due to the fact that the errors in the

data have a larger effect on the images, and the computational time for the solution of

the inverse problem can be considerably longer.

Page 85: Magnetic Induction Tomography for Imaging Cerebral Stroke

64 Dielectric Properties of Head Tissues and Pathophysiology

4.5. Adopted values for simulations and discussion

Considering the reported dielectric properties for three main tissues in the head (brain,

CSF, and blood), there appears to be a good agreement between the suggested values by

Gabriel-III and those from the measurement studies (pre-1996, Gabriel-II, and post-

1996); evident in figure 4.1 to figure 4.7. However, there is considerable scatter of the

data for other tissues which show degrees of anisotropy, namely bones and muscle. A

critical review which investigates the reasons of this scatter is outside the scope of this

study; this can be due to the measuring techniques, sample tissues, sample age,

temperature, etc. However, by looking at different studies, the fidelity of the 1996

database can be assessed.

For skeletal muscle, there are considerable differences between the longitudinal and

transverse values from each study. Furthermore, various studies reported different

values for dielectric properties, in particular, conductivity (figure 4.1). However, both

discrepancies diminished as the frequency increased. If the frequency of current is high

enough (MHz range as in the present MIT systems), the anisotropic properties and

scatter of the data are much less important.

The orientation of the measuring electrodes relative to the axis of the bones (i.e.

longitudinal and transversal) was not specified in the reported studies and, therefore, the

level of anisotropy cannot be estimated for the bones. Noticeable differences exist

between the conductivities of the cancellous and cortical bones, with cancellous being

relatively more conductive, and the conductivities reported by various studies for each

type of bone. However, similar to muscle, the discrepancies appear to become

significantly smaller at higher frequencies (figure 4.4 and figure 4.5).

Therefore, based on the data for five important tissues in the head, it was decided that

the values of the dielectric properties suggested by Gabriel-III can be established as the

main reference for the frequency range of interest in this study, i.e. 1-10 MHz. For

inhomogeneous tissues such as bone, there are no anatomical maps to accurately model

the inhomogeneity, and here a homogeneous domain was considered to represent the

tissue. The values of the properties were calculated as weighted average of those of the

sub-tissues; in this case 50% cortical and 50% cancellous bone.

Page 86: Magnetic Induction Tomography for Imaging Cerebral Stroke

65 Dielectric Properties of Head Tissues and Pathophysiology

Similarly for those organs, e.g. eye balls, for which there are no reported properties as a

whole organ, an estimation was made based on the weighted average values of sub-

tissues' properties. Table 4.1 provides a list of such tissues/organs together with the

details of sub-tissues, and estimated values at a frequency of 10 MHz. The dielectric

properties of the domains existing in the head model are plotted in figure 4.11 for the

frequency range of 100 kHz-100 MHz. The assignment of appropriate dielectric

properties to the tissues in the head model can allow a more realistic estimation of the

expected MIT signals in the simulations. Further description of the head model is

provided in chapter 7.

Table 4.1. Weighted average dielectric properties assigned to heterogeneous or minor organs in the

head (frequency of 10 MHz).

Organ/tissue Constituent Conductivity

(S/m)

Relative

permittivity

Skull 50% cortical, 50% cancellous 0.0828 53.8

Spinal cord 33.3% white matter, 33.3% grey

matter, 33.3% nerve cells 0.225 248

Eye ball 60% vitreous humour, 10% retina,

10% sclera, 10% cornea, 10% lens 1.20 127

Ventricles 100% CSF 2.00 109

Thalamus &

hypothalamus 100% white matter 0.159 176

Basal ganglia 100% white matter 0.159 176

Cavities 100% air 0 1

Olfactory organ 100% air 0 1

Cerebral haemorrhagic lesions were also assumed to be isotropic and homogeneous

regions with blood and brain tissues as its constituents. Different percentages of blood

volume in the haemorrhage were considered in order to calculate the weighted average

of the lesion properties; this ranged 25-100 % representing moderate to severe strokes.

As previously discussed, a haemorrhage can appear as a bleeding region surrounded by

the ischaemic tissues. Any possible ischaemia surrounding the lesion was ignored as the

tissue properties are usually considered to undergo negligible conductivity changes at

MHz frequencies.

Page 87: Magnetic Induction Tomography for Imaging Cerebral Stroke

66 Dielectric Properties of Head Tissues and Pathophysiology

End of chapter 4.

105

106

107

108

101

102

103

104

105

106

107

108

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

2.2

Figure 4.11. Dielectric properties of head

tissues used for simulations.

Conduct

ivit

y (

S m

-1)

Rel

ativ

e p

erm

itti

vit

y

Frequency (Hz)

Page 88: Magnetic Induction Tomography for Imaging Cerebral Stroke

Part II

Discretized Forward and Inverse

Problems

Page 89: Magnetic Induction Tomography for Imaging Cerebral Stroke

Chapter 5.

Discrete Forward Problem

5.1. Introduction

As stated in section 2.2, the purpose of solving the MIT forward problem is to predict

the measured signals for a given geometry, distribution of dielectric properties,

operating frequency, and coil excitation current. There are two different ways of solving

such problems: analytical and numerical methods. Although theoretically exact,

analytical methods have the major drawback of the lack of generality. Analytical

solutions may handle two-dimensional problems, and in certain cases three-

dimensional, as long as the geometry and the boundary conditions are simple, and the

material properties are isotropic.

When exact analytical solutions are not available, approximate methods are sought.

Numerical solutions may handle any geometrical configurations and both isotropic and

anisotropic materials. When the problem is solved numerically, the model of the

physical system must be discretized, requiring an approximate formulation of the

problem. There are several numerical techniques that can be employed. However, any

numerical field computation involves the following steps:

Definition of the boundary value problem (BVP) that includes describing the

region over which the governing equations are to be solved and the boundary

conditions which should be applied.

Page 90: Magnetic Induction Tomography for Imaging Cerebral Stroke

69 Discrete Forward Problem

Formulation of the discrete problem by sub-dividing the domain into a series of

points or elements, and introducing an approximate variation of the field over

the elements.

Derivation of a system of equations whose approximate solution gives the

numerically computed approximate field.

The purpose of this chapter is to introduce two of the most established numerical

methods, finite element (FE) and finite difference (FD), which are employed for the

simulations in this study. The basis of each method is briefly explained and its

implementation for the MIT simulation is discussed.

In order to assess the performance of these methods in solving the forward problem, a

benchmark test is carried out, and the numerical results are compared against each other

and with that from the analytical solution. The starting point for these numerical

methods is the BVP in (2.8), introduced in section 2.2.

5.1.1. BVP simplification

Before embarking on the numerical approximation of (2.8), it is transformed into an

alternative form and a number of approximations are made. There are a number of

alternative approaches to solving (2.8), which express the problem either in terms of

primal physical field variations (e.g. E and H), or in terms of non-physical potentials.

One common approach to solve the problem is by introducing the vector potential A as

B = ×A. Considering that E = -j B, the electric field then becomes:

(5.1)

where is the electric scalar potential. Now realising that H = -1

A and by

substituting it into (2.4) we have:

(5.2)

Also, taking the divergence of (5.2) gives:

(5.3)

The equations (5.2) and (5.3) are known as the retarded potential formulation.

Page 91: Magnetic Induction Tomography for Imaging Cerebral Stroke

70 Discrete Forward Problem

Any A and can be chosen such that they give rise to the physical field B. This is

known as the gauge. To completely define the potentials and obtain a unique solution,

an appropriate gauge is required. Possibilities for gauging include the Coulomb gauge,

.A = 0, or the temporal gauge, = 0 (for further details see [124]).

By Adopting the temporal gauge, the resulting vector wave equation, known as the A-

formulation, can be derived from (5.2) as:

(5.4)

Ledger and Zaglmayr [124] have used this formulation to discuss the eddy-current

problems on multiply connected domains.

For alternative approaches, which express the forward problem in terms of a vector

wave equation for electric field E, see e.g. [34,125,126]. It must be noted that, at this

point, solving either (5.4) or a vector wave equation is equivalent to solving the first

order Maxwell's equations.

Explicitly, the forward BVP for MIT, expressed in terms of A, can be written as:

Find A such that

(5.5)

where [...] denotes the jump, and vector n is a unique outward normal at each boundary

or interface. The domain of the problem, , consists of a conducting region C, and a

non-conducting region N, so that = C N (see figure 2.1 in section 2.2.1).

Assuming that the wave propagation is negligible in N, by choosing N to be large

enough so that n A = 0 can be applied on , the approximate BVP for the forward

problem then becomes:

Page 92: Magnetic Induction Tomography for Imaging Cerebral Stroke

71 Discrete Forward Problem

Find A such that

(5.6)

By solving this problem, the A-field can be estimated everywhere in , and the induced

signals in the sensor coils can be computed by:

(5.7)

where dl is the length element of the coil.

5.2. Finite element method

The finite element method (FEM) is a numerical technique that is used to solve a BVP

governed by a differential equation and a set of boundary conditions

[127,128,129,130,131,20]. The FEM originated from the need for solving complex

elasticity and structural analysis problems in civil and aeronautical engineering.

However, it has now become widely adopted as a means of the numerical

approximation in many problems in science and engineering.

The main idea behind the method is the representation of the domain with smaller

subdomains called the finite elements (FE). The distribution of the primary unknown

quantity inside an element is interpolated based on the values at the nodes, provided that

nodal elements are used, or the values at the edges, in case that vector elements are

used; vector elements are a special kind of FE and are introduced later in this section.

The interpolation or shape functions must be a complete set of polynomials. The

accuracy of the solution depends not only on the size and distribution of the elements,

but also on the order of these polynomials, which may be linear (first-order), quadratic

(second-order), or higher order.

Page 93: Magnetic Induction Tomography for Imaging Cerebral Stroke

72 Discrete Forward Problem

The numerical solution corresponds to the values of the primary unknown quantity at

the nodes or the edges of the discretized domain. The solution is obtained after solving a

system of linear equations. To form such a linear system of equations, the governing

differential equation and associated boundary conditions must first be converted to an

integro-differential formulation by, for example, minimizing a weighted-residual

method such as the Galerkin approach [132,133]. This integro-differential formulation

is applied to a single element and with the use of proper weight and interpolation

functions the respective element equations are obtained. The assembly of all elements

results in a global matrix system which represents the entire domain of the BVP [130].

5.2.1. FE discretization

The weak form of the BVP in (5.6) can be shown to be:

Find A H0(curl) such that

for all W H0(curl), where

(5.8)

where W is a test function [17]. One may attempt to approximate each vector

component of A separately using the common nodal shape functions. However, such an

approach leads to nonphysical solutions, referred to as spurious modes [129,131]. This

can be avoided by using a special kind of FE called edge elements, also known as vector

or Nédélec elements [129,134,135]. These elements are very well suited for

approximating electromagnetic fields. The edge elements associate the degrees of

freedom to the edges of the mesh rather than the nodes.

The shape functions for edge elements are constructed such that their tangential

components are continuous across the elements' borders, whereas their normal

components are allowed to be discontinuous. In the case of the lowest order edge

elements, the vector potential A can be expressed as:

Page 94: Magnetic Induction Tomography for Imaging Cerebral Stroke

73 Discrete Forward Problem

(5.9)

where M is the number of the edges in the mesh. Note that the shape functions Ni are

vectors, and coefficients Ai are scalars that associated with the tangential components of

A on the edges of elements. Constructing the approximation (5.9) in a piecewise manner

over the domain, and using the method of Galerkin weighted residuals where Wi = Ni,

the weak form in (5.8) leads to a linear system of equations as:

(5.10)

where E is the number of elements in the mesh, e is the volume of each element, and

à = (A1,A2,...AM)T are the unknown coefficients associated with the edge element vector

shape functions.

5.2.2. Comsol Multiphysics

A commercial FE package, Comsol Multiphysics [136], was adopted to solve the MIT

forward problem. Comsol has been previously used for simulations in MIT (e.g. [50]).

The „AC/DC‟ module of Comsol is suitable for solving (5.8) and employs the FE

discretization described in 5.2.1. The AC/DC module of Comsol offers the possibility to

use elements of different order applied uniformly throughout the mesh. In the case of

higher order edge elements, the elements have additional degrees of freedom associated

with the edges and faces of the elements.

Unstructured tetrahedral FE elements, which are available in Comsol, are particularly

suitable for representing the complex anatomy of the head as automatic mesh generation

procedures are available for generating a series of elements. Hexahedral and or prism

elements are also available in Comsol, but for these elements, the construction of the

mesh is more difficult.

Page 95: Magnetic Induction Tomography for Imaging Cerebral Stroke

74 Discrete Forward Problem

Adopting Comsol allowed a faster development of the model to estimate the signals for

feasibility tests. However, the use of commercial packages such as Comsol has several

drawbacks which are discussed later in this chapter.

5.2.3. Model construction

The applied current density Js is introduced at the exciter coils. Rather than modelling

the coils as toroids, they were modelled as collection of edges forming a closed loop.

This approach, which approximates coils as 1D entities embedded in 3D space, does not

require the meshing of a separate region for the coils. The source currents were applied

as boundary conditions; the exciter coils are assumed to carry an impressed current with

a predetermined tangential component. Therefore, Js in (5.8) may be set to zero. When

employed as a sensor, the induced e.m.f. around the coil was computed using (5.7).

There are a number of ways in which the screen could be approximately modelled. Let

us first estimate the skin depth in the aluminium screen. Given = 3.77 107 S

m

-1 and

r = 1, the skin depth for the screen at the frequency of 10 MHz is 27 m which is very

small compared to the thickness of the screen (2 mm).

One possibility is to discretize the screen as an additional material, with its individual

3D domain, using tetrahedral elements. However, given the skin depth of the screen,

this would require extremely small elements to be generated in this region to capture the

variation of the electromagnetic field. Beyond the screen, the region of free space would

then extend to the truncating boundary and, in a similar manner to the other material

interfaces, the interface conditions would be naturally handled by the edge element

formulation.

In order to avoid the generation of such small elements within the screen, an alternative

approach can be employed. This consists of truncating the overall domain at the inner

surface of the screen so that the outer boundary becomes = S , and imposing a

suitable boundary condition on S. If the screen is modelled as a perfect electrical

conductor, which corresponds to setting n A = 0 on S, the skin effects in the screen

will not be taken into account. A more physically realistic approach, adopted in this

study, is to model a short distance penetration of the fields inside the screen. This can be

realised by applying the impedance boundary conditions (IBC) as:

Page 96: Magnetic Induction Tomography for Imaging Cerebral Stroke

75 Discrete Forward Problem

(5.11)

on S [137,138,139,23]. In (5.11), the dielectric properties are those for the screen. This

boundary condition is implemented by adding a boundary integral of the form:

(5.12)

to the left-hand side of (5.8), and substituting (5.11) into the integral. Since the

boundary condition involves unknown values of the electric field, this boundary integral

form will contribute to the FE system matrix. Comsol offers the possibility of using the

IBC in the model.

At the open ends of the screen (upper and lower as, for example, in the MK1 system),

the truncated infinite region N extends to the artificial boundary . The boundary

conditions that the AC/DC module of Comsol offers for imposing on are suitable for

modelling the decay of static fields in infinite regions. These boundary conditions are

appropriate if the displacement current is neglected as in the eddy-current problem

which is a magnetostatic problem in the non-conducting region. In such cases, theory

predicts that far away from the source, A should be zero and decay at the rate of r-2

(as

described by (2.11)), with r being the distance from the source. Therefore, placing the

artificial truncation boundary sufficiently far from the target and coils, and imposing

either n A = 0 or n A = 0 allows a reasonable approximation to the decay of the

static fields to be obtained.

Following this approach often leads to the requirement for excessively large

computational domains, which can make the simulations prohibitively expensive.

However, an accurate description of the decay of a static field on a much smaller

computational domain may be achieved if infinite elements are instead employed on the

upper and lower open ends [140,141]. The reason for this is that the mapped infinite

elements lead to the correct decay rate for A as r . This consequently results in a

greater accuracy compared to using either above mentioned boundary conditions, when

is placed close to the target. For the MK1 and MK2a simulations, infinite elements

were introduced at the open ends of the screen in order to capture the far-field.

Page 97: Magnetic Induction Tomography for Imaging Cerebral Stroke

76 Discrete Forward Problem

On a continuous level, the correct treatment of wave propagation effects would require

replacing the boundary condition n A = 0 on by a radiation condition (see (2.9)).

In a computational context, this would require employing some absorbing boundary

conditions (ABC) which can serve to absorb outgoing waves without reflection. A

popular set of such ABCs is the perfectly matched layer (PML) [142,143]. The PML is

a layer of artificial material surrounding the computational region, and designed to

damp waves propagating in the normal direction. The region may then be terminated by

a perfect electric conductor. The PML is not available in the AC/DC module of Comsol.

5.2.4. Solution technique

For a mesh with a small to moderate number of first-order tetrahedral elements (e.g.

150,000 elements corresponding to 180,000 degrees of freedom), the dimension of

the linear system makes it amenable to direct solution techniques. However, for a mesh

with a large number of elements, and in particular for discretizations employing second-

order elements, the increased size of the linear system makes it too expensive to be

solved by direct solvers, either in terms of the memory space requirements or in CPU

time. Therefore, an iterative strategy must be employed.

Comsol offers a variety of different iterative solution techniques. In this study, the

flexible generalised minimum residual (FGMRES) iterative solver, which is suitable for

the complex symmetric matrix obtained in this model, was employed [144,145]. The

speed of convergence was accelerated by using the symmetric successive over-

relaxation (SSOR) vector preconditioner [144].

Only first- and second-order edge elements were considered since it is well known that

when the shape functions for higher-order edge elements are not properly designed, they

lead to a rapid growth in condition number for the FE system matrices and consequently

require a large number of iterations for the iterative solver. The shape functions

employed for the higher-order edge elements in Comsol do not lend themselves to fast

iterative solution as they lead to a rapid increase in condition number and, even after the

application of the available preconditioners, their convergence remains slow. There are

alternative higher-order basis functions which lead to only a moderate growth in

condition number; e.g. see Ledger and Morgan [146,147] and references therein.

Discussion of such shape functions, however, is outside the scope of this thesis.

Page 98: Magnetic Induction Tomography for Imaging Cerebral Stroke

77 Discrete Forward Problem

5.3. Finite difference method

One of the earliest attempts to solve the partial differential equations is the finite

difference method (FDM) [148,129,149]. In the FDM, the partial derivatives are

replaced by difference equations. Space is divided into FD cells, and the equations are

written for the unknown potential of the field values at nodes. The partial differential

equations are replaced by a set of simultaneous equations. The unknown potentials are

computed on a finite set of points or nodes. The values at other locations may be

obtained by interpolation.

5.3.1. BVP

In the conducting region C, Js = 0 and the right hand side in (5.3) becomes zero. Using

the Coulomb gauge .A = 0, then (5.3) may be re-written as:

(5.13)

In arriving at this equation, the Coulomb gauge has been used, but not imposed. The

BVP before the FD discretization is therefore:

(5.14)

where n is the unit outward normal to . The boundary condition of the BVP is

obtained by equating the normal components of the total current on C to zero [150].

The BVP in (5.14) is the coupled partial differential equation which has to handle the

solution of and A together. However, by making certain assumptions, the problem

may be simplified:

wave propagation delays are negligible

the effect of the dielectric volume on A is negligible (contribution of the induced

currents to the magnetic vector potential is insignificant, and A is entirely

produced by the current source outside the dielectric volume)

skin depth is large compared with the dimensions of the dielectric volume

Page 99: Magnetic Induction Tomography for Imaging Cerebral Stroke

78 Discrete Forward Problem

Therefore, A in the BVP (5.14) is replaced by the approximation of the magnetic vector

potential created by the current source, Â, computed by:

(5.15)

where Js is the current source, and R is the distance between the source and the field

points [151,150]. In this case, the filamentary current source is applied along the exciter

coils. Therefore, Â in (5.15) can be expressed in cylindrical coordinates (r, , z) by:

(5.16)

where I is the current in the coil, a is the radius of the coil, K(k) and E(k) are elliptic

integrals of the first and second kinds [152]. This results in decoupling of the potential

equations, and the magnetic vector potential in the absence of the dielectric volume is

used directly as a source term for the computation of the electric potential. The BVP

reduces to solving (5.14) for the scalar electric potential within the dielectric volume.

This way, the displacement currents flowing outside the dielectric volume are ignored,

and the problem reduces to a magnetostatic problem in the non-conductive region N.

After solving the BVP, is known, and the total induced current density, JI, is the sum

of conductive and displacement components as:

(5.17)

This current creates a secondary magnetic vector potential, AI, which can be computed

by the Biot-Savart law:

(5.18)

where rI is the distance between the induced current and the field points. The voltage

due to the induced currents at the sensor coils may then by computed using (5.7), where

A is replaced by AI.

Page 100: Magnetic Induction Tomography for Imaging Cerebral Stroke

79 Discrete Forward Problem

Figure 5.1. A 2D representation of the grid showing the electrical connections between a node and

its 4 neighbours in the grid (6 neighbours in 3D). The admittances Y and vector voltage generator components e are known [153].

5.3.2. FD discretization

The dielectric volume is discretized using regular cubic voxels (size x), with the voxel

node located at the centre of the voxel. A constant conductivity and permittivity value is

assigned to each voxel. The complex conductivity within each voxel is considered as

= + j . The local admittance between the nodes (i,j,k) and (i-1,j,k) in figure 5.1 is

therefore computed by [154]:

(5.19)

The magnetically induced electric field is modelled as voltage generator e as:

(5.20)

where  is evaluated at midway between the two nodes of interest [152]. The current

between the nodes (i,j,k) and (i-1,j,k) is therefore given by:

(5.21)

(i,j,k) (i-1,j,k) (i+1,j,k)

(i,j+1,k)

(i,j-1,k)

Y(i+1,j,k) e(i+1,j,k)

Y(i,j-1,k)

e(i,j-1,k)

e(i-1,j,k) Y(i-1,j,k)

e(i,j+1,k)

Y(i,j+1,k)

x

y

Page 101: Magnetic Induction Tomography for Imaging Cerebral Stroke

80 Discrete Forward Problem

Similarly the current through the other five boundaries of node (i,j,k) can be computed,

and based on Kirchhoff's current law we have:

(5.22)

In this way, the scalar potential at a node can be calculated using the six nearest

neighbour values.

5.3.3. Solution technique

The BVP is solved by the successive-over-relaxation (SOR) method, which effectively

applies Kirchhoff's current law repeatedly throughout the grid as:

(5.23)

where m

(i,j,k) and m-1

(i,j,k) are the updated and existing values of the scalar potential at

node (i,j,k), and m

(i,j,k) is the change in the potential between two iterations obtained

by:

(5.24)

Page 102: Magnetic Induction Tomography for Imaging Cerebral Stroke

81 Discrete Forward Problem

In equation (5.23), f is the acceleration or relaxation factor. A value of 1.9 was deemed

to be a good choice for f and is used throughout this study. For further details on the

SOR method and its convergence properties, see e.g. Quarteroni [155] or Ledger [131].

The residual is calculated as the sum of squares of the increments in scalar potential for

each iteration over the whole grid:

(5.25)

where N1, N2, and N3 are the maximum voxel numbers is three directions. The residual

reduces continuously with each iteration, and is used as a measure of the convergence

rate and as a stopping criterion. The iteration is stopped when the residual is fallen by a

factor of 1/

m 10

20. Once is known, the induced currents within each voxel in the

dielectric volume are computed by (5.17). The resulting secondary magnetic vector

potential can be obtained by approximating the Biot-Savart law at the sensor coils as:

(5.26)

where JIn is the induced current density vector within the n-th voxel, rIn is the distance

between the node of the n-th voxel and the coil, AIn is the contribution of n-th voxel to

the secondary magnetic vector potential, and N is the number of voxels in the grid.

Subsequently, the e.m.f. in the sensor coil is computed by (5.7); note that rIn in (5.26) is

the distance to one element dl of the coil.

5.3.4. FD solver

A previously developed code [153], available in the South-Wales MIT group, was

employed for the FD simulations in this study. The FD code, hereinafter as 'FD solver',

had been developed based on the FDM described in this section, and was written in

Fortran1. The existing code was further improved to implement the electromagnetic

screen into the model (explained later in this chapter).

1 Fortran is a programming language that is suited to numeric computation and scientific computing.

Page 103: Magnetic Induction Tomography for Imaging Cerebral Stroke

82 Discrete Forward Problem

Figure 5.2. Exciter coil and conductive sphere.

5.4. Benchmark test

In order to validate the numerical techniques employed to solve the MIT forward

problem, a benchmark test was first attempted before solving the realistic head model

for haemorrhage simulations. The setup comprised a coil carrying an alternating current

as the exciter, placed coaxially with a spherical conductive object with material

properties , , and (see figure 5.2).

An analytical solution for the electric field inside a multilayered conducting sphere in

spherical coordinates (r, , ), has been derived by the LCOMIT partners based on

modified Bessel functions, In+1/2 and In-1/2, and associated Legendre polynomials, Pn

[24]. For simplicity, a homogeneous sphere is considered here for which the general

analytical solution reduces to:

(5.27)

where J0 is the current in exciter coil, 0 is shown in figure 5.2, and (r, ) are coordinates

of the point of evaluation.

Page 104: Magnetic Induction Tomography for Imaging Cerebral Stroke

83 Discrete Forward Problem

The validity of the analytical solution is restricted to cases when / << 1 , so that the

displacement current can be neglected. Following the calculation of E, the induced

eddy-current density inside the sphere can be computed from J = E. Values of r0, rC

and a were 143.7, 25 and 60 mm, respectively (see figure 5.2). The properties of the

sphere were = 0 and = 1 S m-1

, the frequency was 10 MHz, and the current in the

coil was J0 = 1 A. The ratio / was then 5.6 10-4

in the conducting region.

The benchmark problem was modelled using Comsol in which an artificial boundary

was introduced at a radius of 20 cm. The region beyond the boundary was modelled

using infinite elements. A series of meshes comprising 2,342 (figure 5.3a) to 1,135,307

(figure 5.3b) tetrahedral elements was generated in Comsol. The FE model was then

solved using first- and second-order basis functions, and the numerical results were

compared against the analytical solution.

The relative L2 error measure for the numerically calculated induced eddy current in the

sphere was approximated using an 11-point Gauss Quadrature rule:

error = (5.28)

where EL is the total number of FEs, and wp is the Gauss coefficient in a reference

tetrahedron multiplied by the determinant of the Jacobian. The term ep = |Jnu - Jan|2 is

the magnitude of the error at each integration point inside the tetrahedron, where Jnu and

Jan are the numerical and analytical current densities, respectively.

Figure 5.3. Typical FE meshes generated for benchmark test; a coarse (a), and a fine (b) mesh. Note

that the presence of coil, sphere, and layer of infinite elements (separate subdomains) may result in

generation of non-uniform meshes, e.g. in (a).

(a) (b)

Page 105: Magnetic Induction Tomography for Imaging Cerebral Stroke

84 Discrete Forward Problem

Figure 5.4. Error in the numerically computed eddy current inside the conducting sphere versus

the number of degrees of freedom; results for first- and second-order elements plotted in broken

red and solid blue lines, respectively.

As expected, the error when using second-order elements was smaller than that with

first-order elements for the same number of degrees of freedom (figure 5.4). As the

mesh was refined, the numerical results tended to the true (analytical) solution.

Examining the gradient of the error curves revealed that the second-order elements

converged approximately twice as fast as the first-order elements. The error for second-

order elements ranged from 3.76 to 0.25% which indicates the numerical results are in

good agreement with the analytical solution.

In order to attain a global error of 3%, approximately 1,135,000 and 3,800 elements of

first- and second-order were required, respectively, which stresses the superiority of

second-order edge elements in terms of the accuracy.

Comsol uses Nédélec‟s edge elements of the first type, so that a first-order element has

6 degrees of freedom, whereas a second-order element has 20 [17,134]. It is expected

that with further refinement of the mesh (i.e. increasing the number of elements), the

error would decease further, but the time taken to solve the model would become

prohibitively long.

103

104

105

106

107

10-1

100

101

102

Degrees of freedom

err

or

%

Page 106: Magnetic Induction Tomography for Imaging Cerebral Stroke

85 Discrete Forward Problem

5.4.1. FE vs. FD vs. Analytical

In order to assess the performance of the FDM, the sphere was modelled using a cubic

grid of 1 mm voxel size. The model was solved for two conductivities:

= 1 S m

-1 where the skin depth was = 160 mm

= 100 S m

-1 where the skin depth was = 16 mm

The FE model was also solved for = 100 S m

-1 using the finest mesh generated in the

previous section; for = 1 S m

-1 the FE solution was already available. The values of

the radial current density along r (see figure 5.5) were plotted for the analytical, FE, and

FD solutions.

Figure 5.5. Cross section of sphere showing r.

For = 1 S m

-1, the skin depth is larger than the sphere diameter. In this case, there is a

good agreement between the three models (figure 5.6a). Close to the surface of the

sphere (r 60 mm), the FD model slightly departs from the analytical and FE solution.

This is due to the discretization error on the boundary, whereas a tetrahedral mesh in the

FE model can properly represent the curved geometry.

When the skin depth is smaller than the sphere diameter, i.e. as in the case when =

100 S m

-1, the FDM fails to predict the correct values of the current density (figure

5.6b). However, there is a good agreement between the analytical and FE solutions.

Note that the FD solution for either of the conductivities can be obtained simply by

multiplying the values in the other one by the corresponding ratio of the conductivities.

exciter coil

Page 107: Magnetic Induction Tomography for Imaging Cerebral Stroke

86 Discrete Forward Problem

Figure 5.6. Plots of current density |J| along r for = 1 S m-1

(b) and = 100 S m

-1 (c).

Figure 5.7 shows the real and imaginary components of the current density together

with its norm along r, from the FE results computed for two conductivities. As can be

seen, for the large skin depth at = 1 S m

-1, the induced current is predominately

imaginary. However, for the smaller skin depth at = 100 S m

-1, the real component of

the current density is considerably larger.

0 0.01 0.02 0.03 0.04 0.05 0.060

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

0.18

0.2

r (m)

|J|

Analytical

FE

FD

0 0.01 0.02 0.03 0.04 0.05 0.060

2

4

6

8

10

12

14

16

18

20

r (m)

|J|

Analytical

FE

FD

(a)

(b)

Page 108: Magnetic Induction Tomography for Imaging Cerebral Stroke

87 Discrete Forward Problem

Figure 5.7. FE results: real, imaginary, and norm of current density along r for = 1 S m-1

(left)

and = 100 S m

-1 (right).

Figure 5.8 illustrates these quantities as the current maps on a cross-sectional plane

cutting through the sphere. The high magnitude current densities on the surface are

evident for = 100 S m

-1.

Figure 5.8. Maps of induced current densities in A m

-2 on a cross-section of sphere for FE results

when = 1 and 100 S m-1

.

Real Imaginary Absolute value

r (m) r (m)

Curr

ent

den

sity

(A

m-2

) |J|

|Im(J)|

|Re(J)|

|J| |Im(J)|

|Re(J)|

= 1 = 100

= 1

= 100

Page 109: Magnetic Induction Tomography for Imaging Cerebral Stroke

88 Discrete Forward Problem

5.5. Screen implementation in FD model

As stated in section 2.3, the MIT sensor array usually includes an electromagnetic

screen which can have a significant effect on the MIT signals. Therefore, implementing

the screen into the simulations is an essential aspect of the MIT models. The method by

which the screen is modelled in Comsol was explained earlier in section 5.2.3.

However, in the FD solver the region outside the conductive domain is not discretized,

so the screen could not be included in the FD models in a straightforward way. As

described by Ktistis et al [156], in order to add the screen to the FD solver, the forward

problem can be solved using the exciter-sensor mutual impedance changes, Z, for a

given target. Based on the Lorentz reciprocity theorem, Z for a pair of exciter-sensor

coils (a and b) can be derived in terms of the real inner product of the magnetic and

electric fields as:

(5.29)

where

Za is the mutual impedance between the coils when the domain properties are μa,

εa, σa and an alternating current I, of angular frequency ω, is applied to coil a to

produce the magnetic field Ha and the electric field Ea.

Zb is the mutual impedance for domain properties μb, εb, σb and Hb, Eb fields

produced by coil b when excited by the same current and frequency.

The volume of integration, , should include the regions where domain properties are

changed. Equation (5.29) is not exact as it ignores the error term ((σb-σa , εb-εa , μb-μa)2)

which is second-order with respect to the dielectric properties changes (for further

details see [157]). Assuming the permeability change to be negligible, which is usually

valid for biological tissues, (5.29) can be simplified as:

(5.30)

If the impedance changes are measured with respect to empty space, the effect of the

permittivity εa is negligible and σa = 0, and therefore:

Page 110: Magnetic Induction Tomography for Imaging Cerebral Stroke

89 Discrete Forward Problem

(5.31)

In empty space (air), the electric field is solely produced by the magnetic vector

potential A as:

(5.32)

The current density inside the target may can be computed by Jb = ( b + j b) Eb. Using

(5.31) to obtain the voltage changes in the sensor coils as V = I Z, then gives:

(5.33)

which can be used to compute the voltage change in the sensor coil (coil a) with respect

to the empty space. The vectors Aa and Jb are computed for the volume of integration,

i.e. conductive volume C. Note that Aa (hereinafter as 'A-field') is calculated for empty

space and is not affected by the target; it is related only to the geometry of the sensor

array and its operating parameters.

Therefore, if the A-field is available for an MIT sensor array, computed in presence of

the screen, it can be fed to the FD solver for computing V for any target of interest.

The A-field can be computed by the FE model in Comsol which includes the screen:

the A-field calculated for the exciter coils replaces (5.16) for solving the eddy-

current problem and computing Jb within the conductive volume.

the A-field calculated for the sensor coils is used for computing the voltage in

the sensors using (5.33), instead of (5.7) and (5.26).

5.5.1. Screen for MK1 and MK2a systems

In order to the assess validity of (5.33), FE models were constructed in Comsol for the

MK1 and MK2a systems (figure 5.9). Instead of truncating the domain at the internal

surface of the screen and adding the infinite elements at the above-bottom open ends,

the overall domain was considered to be a sphere of 30 cm radius. The screen domain

was added to the model as a vacuum, with IBC on its boundaries; material of the screen

was not discretized. A layer of infinite elements was then added to capture the far-field.

Page 111: Magnetic Induction Tomography for Imaging Cerebral Stroke

90 Discrete Forward Problem

Figure 5.9. FE models for computing A-field for MK1 and MK2a systems.

Although this approach requires the generation of more elements in the non-conductive

region (extra elements beyond the screen are required), the construction time of a one-

piece spherical layer of infinite elements is considerably shorter in Comsol than that for

the two-piece upper-lower positioned layers. However, the solution time will be longer.

A series of increasingly fine meshes of tetrahedral elements was generated, and the FE

model was solved using the second-order elements:

Mesh-1: 16,968 elements, 101,816 degrees of freedom

Mesh-2: 30,708 elements, 188,860 degrees of freedom

Mesh-3: 53,923 elements, 333,636 degrees of freedom

Mesh-4: 108,811 elements, 683,740 degrees of freedom

In order to assure the accuracy of the results, the magnitude of the vector potential, |A|,

was monitored along an arbitrary test line passing through the imaging domain within

the sensor array; line specified by (-0.1,-0.1,-0.1)-(0.1,0.1,0.1) in a Cartesian

coordinates whose origin was the centre of the spherical domain, and z-axis was

directed along the axis of the cylinder. Figure 5.10 shows the convergence results for

the MK1 system, with Ex1 (exciter coil 1) activated (for coils numbering in the MK1

system, see Appendix A). Mesh-4 deemed to have converged and further refinement of

the mesh did not significantly alter the results. For other activated exciter coils, similar

behaviour was observed. Therefore, the computed A-field from this mesh was employed

for the FD simulations.

MK1 MK2a

screen layer of infinite

elements

imaging region

Page 112: Magnetic Induction Tomography for Imaging Cerebral Stroke

91 Discrete Forward Problem

Figure 5.10. Plots of |A| for 4 meshes along a test line.

The A-field was extracted for a centrally located cylinder, 12 cm radius and 50 cm

height; this is the imaging region where the target is placed (see figure 5.9). The A-field

was extracted on regular cubic grids of 5 and 10 mm. The FD solver was amended to

implement the extracted A-field as the source. The discrete form of (5.33) can be

expressed by:

(5.34)

where x is the grid size, N is the number of voxels in the grid, Aan and Jbn are the fields

within the n-th voxel, and Jbn is computed by (5.17). Equation (5.34) gives the voltage

for one exciter-sensor coil combination, and it must be applied to each pair of coils.

In the following step, a conductive cylinder of 0.1 S m

-1, 10 cm radius and 20 cm height,

was centrally located within the sensor array. For this cylinder, the skin depth ( = 50.3

cm when r = 1 and f = 10 MHz) was larger than its dimensions. The FE model was

solved using the same mesh selected for the A-field computations. The solution time for

one activated exciter was 15 min1.

1 All simulations in this thesis were conducted on a Dell Precision 390 workstation, which has a CoreTM2

6700 CPU, with an internal clock frequency of 2.66 GHz and 8 GB of memory.

Page 113: Magnetic Induction Tomography for Imaging Cerebral Stroke

92 Discrete Forward Problem

The FD model, with the screen incorporated, was also solved using the grid sizes of 5

and 10 mm. The solution time for one activated exciter was approximately 6 and 0.4s,

for 5 and 10 mm gird sizes, respectively. The induced secondary voltages obtained from

FE and FD simulations are plotted in figure 5.11, for Ex = 1 and Se = 1,2,..,16 (as

described in Appendix A, throughout this thesis, exciter and sensor coils are labelled as

Ex and Se, respectively).

As it can be observed, there was a good agreement between the FE and FD results. The

FD results with 10 mm grid size slightly departed from the FE results at Se1 and Se16,

where the magnitude of the voltage was at its largest (see magnified view for Se16). The

5 mm grid size can represent the curved surface of the conductive cylinder more

accurately. The FD model was also solved without the screen. Large discrepancies

between the results from the model without the screen, and those with the screen

included are evident in the figure.

Figure 5.11. Induced secondary voltage at 16 sensor coils for Ex1 activated (FE mesh comprised

108,811 tetrahedral elements).

A similar procedure was carried out for computing the A-field for the MK2a system.

The converged FE mesh comprised 103,375 tetrahedral elements corresponding to

643,900 degrees of freedom.

0 2 4 6 8 10 12 14 16

-20

-10

0

x 10-5

Se

Secondary

voltage (

V)

FE

FD - 10 mm

FD - 5 mm

FD - 5 mm (without screen)

14.5 15 15.5 16 16.5

-5

-4

-3

-2

-1

0

1x 10

-5

Se

Secondary

voltage (

V)

10-5

-3

-5

Page 114: Magnetic Induction Tomography for Imaging Cerebral Stroke

93 Discrete Forward Problem

5.5.2. Simulated vs. measured data

With the screen included in the FD model, the FD and FE results can now both be

compared against the practical measured data. In this section, the results from the

experimental study carried out by Griffiths et al [8] are reproduced through simulations.

As the target, they used a cylindrical glass vessel (diameter 120 mm, height 200 mm)

filled with saline solution with the following nominal conductivities: 0.5, 1.0, 1.5, 2.0

and 3.0 S m

−1. Tap water (0.012 S

m

−1) was used for an additional measurement to

provide a sample with a very low conductivity.

In their study, the measurements were undertaken using Ex1-Se9 combination (opposite

coils) of the MK1 system operating at the frequency of 10 MHz (see figure 5.12). First,

the primary voltage Vp was measured (sample absent). Then sample was placed at the

centre of array, and the secondary voltage Vs was acquired.

Here, the same setup was simulated by the FE and FD models. Recalling from section

2.3.1, for a fixed frequency, when the skin depth is larger than the target dimensions,

then we can expect Re( V/Vp) and Im( V/Vp) .

Simulated and measured results for Im( V/Vp) are shown in figure 5.13a and figure

5.13b, respectively. As can be seen, the imaginary component closely obeyed the

expected linear relationship with the conductivity in all three data sets (measured, FE,

and FD). A fitted linear line for each data set is shown in the figure, where the

coefficients of the line equation are also provided.

Figure 5.12. Plan view of the MIT array showing the position of the sample and coils within the

electromagnetic screen. Red (outer) and black (inner) coils are exciters and sensors, respectively.

9

1

Screen

Sample

Page 115: Magnetic Induction Tomography for Imaging Cerebral Stroke

94 Discrete Forward Problem

Figure 5.13. Values of Im( V/Vp) in percentage against conductivity: (a) FE and FD results, (b) measured data where straight fitted line is also shown (after Griffiths et al [8]).

The real component Re( V/Vp), plotted in figure 5.14, was much smaller in magnitude

than the imaginary and was positive, but only for the two lowest values of conductivity

(0.012 and 0.5 S m

−1). As the conductivity increased, the curve in the measured and FE

results became negative and increased in magnitude. It is likely that the small, positive

value at low conductivities arose from the permittivity of the saline solution, when the

eddy currents were predominantly displacement currents.

As the conductivity increased, and the skin depth became comparable with the diameter

of the sample, the assumption of a weak interaction, on which the FD model is based,

became invalid. Therefore, the FD model failed to predict the correct values for the real

component and, as can be seen in figure 5.14a, the values remain unchanged for a fixed

0 0.5 1 1.5 2 2.5 3-10

-8

-6

-4

-2

0

Conductivity (Sm-1)

Im (

V/V

p)

(%)

(a)

(b)

FE

-3.28 - 0.03

FD

-3.28 - 1.6 10-10

Page 116: Magnetic Induction Tomography for Imaging Cerebral Stroke

95 Discrete Forward Problem

value of permittivity ( r = 80), regardless of the conductivity. The skin depth would be

equal to the diameter of the sample, 120 mm, at a conductivity of 1.76

S m

−1.

Figure 5.14. Values of Re( V/Vp) in percentage against conductivity: (a) FE and FD results, (b) measured data (after Griffiths et al [8]).

As an extra step and in order to investigate the frequency dependence of the conduction

and displacement currents, the FE model was solved for two fixed conductivity values

of 0.012 and 3 S

m

-1 ( r = 80), where frequency was varied between 10 kHz to 10 MHz.

The measures of the conduction and displacement current densities induced throughout

the cylindrical vessel were computed by integrating the norm of current density over the

cylinder using an 11-point Gauss Quadrature. These values are plotted in figure 5.15

against the frequency.

0 0.5 1 1.5 2 2.5 3-1.4

-1.2

-1

-0.8

-0.6

-0.4

-0.2

0

0.2

Conductivity (Sm-1)

Re (

V/V

p)

(%)

(a)

(b)

FD

FE

Page 117: Magnetic Induction Tomography for Imaging Cerebral Stroke

96 Discrete Forward Problem

The norm of displacement current was linearly proportional to the frequency squared,

and remained the same for a fixed value of relative permittivity and frequency,

irrespective of the conductivity. The norm of conduction current, however, was almost

linearly proportional to the frequency. It was also proportional to the conductivity. The

average ratio between the values for 3 and 0.012 S m-1

was 249.2 where 3/0.012 = 250.

Note that at the frequency of 10 MHz and for = 0.012 S m-1

, the displacement currents

are larger than the conduction currents, resulting in a positive value for Re( V/Vp)

shown in figure 5.14.

Figure 5.15. Total values of conduction and displacement current densities induced throughout the

cylindrical vessel.

5.6. Discussion

Comparison of the FE results with the analytical solution (section 5.4) and practical

measurements (section 5.5.2) ensured the validity of the employed FE techniques. A

very strong advantage of the FEM is its ability to deal with complex geometries by

using unstructured grids. This can be of importance for modelling a realistic human

head geometry, with its complex features. The employment of Comsol for the solution

of the forward problem in this study, allowed a rapid development of the model to

104

105

106

107

10-12

10-10

10-8

10-6

10-4

10-2

Frequency (Hz)

Induced c

urr

ent density thro

ughout cylin

der

(mA

)

Displacement

Conduction

= 3 S m-1

Conduction

= 0.012 S m-1

Page 118: Magnetic Induction Tomography for Imaging Cerebral Stroke

97 Discrete Forward Problem

estimate the signals for feasibility tests. However, the use of commercial packages, such

as Comsol, has two major drawbacks:

For large 3D models comprised of more than 100,000 elements, most of the

computational effort associated with the FE simulation is devoted to the solution

of a large sparse set of linear equations. For meshes of this size, direct solvers

are inappropriate due to the excessive computational resources that they require.

Therefore, iterative forward solution techniques must be adopted. However,

without suitable pre-conditioners, these can be slow to converge. Although

commercial FE packages offer a range of solution techniques, their performance

for MIT was not found to be satisfactory, and led to excessive run times. This

could jeopardise one major advantage of using MIT which is a rapid diagnosis.

For an optimised implementation of iterative inverse algorithms, it is necessary

to solve multiple forward problems in order to update the Jacobian (sensitivity

distribution matrix); the sensitivity matrix is introduced in the next chapter. To

this end, it is essential to have unrestricted access to the source code in order to

introduce the necessary modifications. Besides the excessive computational time

for solving multiple forward problems, another major drawback is the fact that

access to the source codes is normally impossible in commercial packages.

In contrast, the FD solver offers full flexibility and accessibility, allowing the necessary

amendments to be done. Furthermore, the FD solution time is considerably shorter; for

the particular example presented in section 5.5.1, the FD model performed 150 times

faster than the FE model, with comparable results.

Comparison of the FD results against those from the analytical (section 5.4.1), measured

(section 5.5.2), and FE (section 5.5.1) data indicated that the FD model can be used for

the MIT simulations. However, strictly its applicability is limited to the cases when the

target's dimension are larger than the skin depth. Although by examining the imaginary

part of the signal computed by the FD model, there is surprisingly good agreement with

the FE model and the experimental results when the target is of comparable dimension

to the skin depth (figure 5.13).

Therefore, when appropriate, the FDM has been employed for the simulations in the

remainder of this thesis. Furthermore, when conducting preliminary studies for the

Page 119: Magnetic Induction Tomography for Imaging Cerebral Stroke

98 Discrete Forward Problem

introduction of various image reconstruction algorithms (chapter 6), and system

geometrical optimisation (chapter 8), FDM was the method of choice.

When using the measured data, it is crucial to simulate the MIT system as accurately as

possible. This includes implementing the screen into the model. Figure 5.11 indicated

that the effect of screen can be significant on the induced voltages.

End of chapter 5.

Page 120: Magnetic Induction Tomography for Imaging Cerebral Stroke

Chapter 6.

Discrete Inverse Problem

6.1. Introduction

As stated in section 2.2, the aim of solving the MIT inverse problem is to recover the

electrical properties from the measured signals at the sensor coils. The purpose of this

chapter is to explain the general framework of the discretized MIT inverse problem, and

to review the image reconstruction algorithms with details of their implementations.

The discussion starts with time-difference imaging, and describes the difficulties

associated with the MIT inverse problem which are common to all types of MIT

imaging. This is followed by frequency-difference and absolute imaging.

Although time- and frequency-difference imaging techniques are discussed in different

sections, they are essentially similar, and all the inverse algorithms applicable to time-

difference imaging can also be used for frequency-difference imaging. Absolute

imaging requires iterative image reconstruction techniques, and is dealt with in a

different section. When necessary, images are reconstructed after introducing each

individual algorithm and the results are discussed.

In practice, the distribution of dielectric properties cannot assumed to be continuous. In

order to approximate the inverse solution, similar to the forward problem, the inverse

problem is discretized which yields a solution in matrix representation:

The imaging volume is discretized into n elements (also voxels or pixels).

The conductivity value throughout each element is assumed to be constant.

Page 121: Magnetic Induction Tomography for Imaging Cerebral Stroke

100 Discrete Inverse Problem

n unknown conductivity values are arranged in a vector.

After solving the inverse problem, the obtained conductivity values are then

assigned to the corresponding elements in order to reconstruct the image.

Figure 6.1. One geometrical example of the model (to scale): (a) diagram of the 3D model, (b) x-y

and y-z cross-sectional planes for image representations.

Throughout this chapter, geometrically simplified 3D models are used for the

simulations. The model comprises a homogeneous cylindrical tank (background) of 10

cm radius and 16 cm height (figure 6.1a). One or more cylindrical inclusions, 2 cm

radius and 4 cm height, are placed within the background. The inclusions have various

conductivities different from that of the background. The forward problem (and inverse

problem) is solved using the FD solver and for the MK1 system, with the screen

included. The discretized problem has a voxel size of 10 mm resulting in n = 5056

conductive voxels in 3D. Throughout, the parameter to be estimated is the conductivity.

For displaying the reconstructed images, two cross-sectional planes, passing through the

tank and inclusion, are extracted from the reconstructed volume image in order to show

the conductivity values (figure 6.1b). In some cases, 1D images are also plotted in order

to closely investigate the computed conductivity values along a line passing through the

imaging volume.

It is remarked that by using the same set of voxels for the forward and inverse problems,

and in some cases using the noise-free simulated data, a so-called inverse crime is

committed (inverse crime is discussed at the end of this chapter). In this regard, it must

be noted that the purpose of this chapter is to introduce the concept of the inverse

problem and the difficulties associated with solving it, briefly review some of the

y-z

x-y

inclusionbackground x-y y-z

(a) (b)

Page 122: Magnetic Induction Tomography for Imaging Cerebral Stroke

101 Discrete Inverse Problem

existing image reconstruction algorithms, and provide a basis for dealing with more

realistic cases later in this thesis. The inverse algorithms and techniques, explained in

the following sections, are well-established methods and have been presented before for

EIT simulations. Here, they are explained in the context of MIT, and MIT related

examples are provided.

6.2. Time-difference imaging

As stated in chapter 2, time-difference imaging is to calculate the conductivity change

between two states, separated in time. Depicted in figure 6.2, is an example of the two

states with different conductivity distributions (hereinafter as „State-I‟ and „State-II‟).

The goal of the inverse problem in this case is to estimate from V, where V is the

voltage change measured in the sensor coils between the two states. There are several

difficulties associated with solving this problem including its being non-linear,

underdetermined, and ill-posed.

Figure 6.2. Diagram illustrating the conductivity distributions for two states in x-y planes.

6.2.1. Non-linearity

In MIT, the field strength, computed by (5.6) or (5.14), varies as a function of position

and distribution of dielectric properties throughout the domain. Therefore, eddy currents

(J = E) and the signals induced by them, are a non-linear function of the distribution

of dielectric parameters of the entire domain. In order to obtain the conductivity

State-I State-II

Page 123: Magnetic Induction Tomography for Imaging Cerebral Stroke

102 Discrete Inverse Problem

distribution one must solve a system of equations which relate every voxel in the

domain to every measurement.

One approach for solving the discretized inverse problem can be obtained by linearizing

the model. First, the forward problem, when approximated by either (5.6) or (5.14), may

be considered in the form:

(6.1)

where n is a column vector of the (real) conductivity values in each voxel, V m

is

the column vector of voltages induced in the sensor coils resulting from , and F is the

corresponding non-linear function which represents the approximate solution of the

forward problem [158]. If F and are known, the voltage can be calculated. This is the

forward problem that was discussed in chapter 5, and for the presumed boundary

conditions, has a unique solution.

Assuming small conductivity changes between the two states, one can expand (6.1)

using a truncated Taylor series. With the second and higher order terms in the series

sufficiently small, the non-linear problem may be approximated by:

(6.2)

Here, F/ is the Jacobian of the (discrete) forward operator F( ), computed at some

initial conductivity estimate 0, which relates interior conductivity perturbations to

perturbations in the measured signals. In this context, the Jacobian is often called the

sensitivity matrix.

Let us assume that 0 is the conductivity distribution for the homogeneous State-I

(figure 6.2) and is known. Using the linearized problem in (6.2):

(6.3)

The Ax = b is essentially a system of linear equations in which A is the sensitivity

matrix, b is measured data (voltage changes), and x is the vector of conductivity

changes to be determined. Now let us consider the case when State-I is not

homogeneous and its conductivity distribution ( 1) is not known (see figure 6.3).

Page 124: Magnetic Induction Tomography for Imaging Cerebral Stroke

103 Discrete Inverse Problem

Figure 6.3. Diagram illustrating the conductivity distribution for two inhomogeneous states.

Similarly, the problem can be linearized for the two states around an initial guess 0:

(6.4)

which again yields a system of linear equations. Note that the linearization is valid as

long as there are no large differences between 0 and the conductivity distribution for

State-I. For instance, if State-I has a conductivity distribution with large contrasts, a

homogeneous 0 may result in a sensitivity matrix that does not accurately represent the

variations in the voltage obtained in the forward model for changes in the conductivity

distribution. This is further discussed in sensitivity analysis in section 6.3.1.

6.2.2. Underdetermined

The MIT inverse problem will usually suffer from a shortage of information. The reason

for this is twofold. Firstly, the number of available measurements is typically fairly

small. Secondly, it is a common practice to discretize the domain by a large number of

elements (voxels), inside of which the conductivity is assumed to be constant, in

particular in 3D imaging.

For instance in the MK1 system, with 16 exciter and 16 sensor coils, a total of m =

16 16 = 256 coil combinations is obtained. However, since the exciter and sensor coils

State-I State-II

Page 125: Magnetic Induction Tomography for Imaging Cerebral Stroke

104 Discrete Inverse Problem

are placed closely, only 136 of these are independent due to the reciprocity theorem

which states that reversing the sensor and exciter coils would give an identical mutual

inductance between the coils [158,159].

One may attempt to overcome this problem by increasing the number of channels in the

system as the MIT is a contact-less technique. In general, a C-channel system, with each

channel acting as both the exciter and sensor, will deliver C (C+1)/2 independent coil

combinations [160]. More information could also be obtained by conducting the data

collection for several planes of measurement; e.g. positioning the MKI sensor array at

different heights with respect to the target, which results in more coil combinations.

However, the number of coils that can be implemented, and the number of planes of

measurement that can be undertaken are limited by the hardware and processing time

restrictions.

Furthermore, one could choose to reduce the number of voxels (unknowns), although

this may result in blocky images. Finally, another possibility would be to use an

alternative basis for representing the unknown conductivity (e.g. a Fourier basis

[161,162,163,164]). In this way, the number of unknowns can be reduced, but still

maintaining the smoothness of the solution. However, in this study, the conductivity is

discretized by constant values inside each voxel, and the number of unknowns will be

larger than the number of measurements. Therefore, the problem will be

underdetermined. A more precise definition of over- and underdetermined problems is

provided in the following.

In the overdetermined case (A m×n

, m >

n), there is at least one solution to Ax = b, if

the number of linearly dependent rows of A (i.e. r = rank(A)) is less or equal to the

number of columns of A. The least squares (LS) solutions are the ones that satisfy:

(6.5)

where ||.|| is the Euclidean norm. In the case of full column rank (r = n), the LS solution

is unique and may be obtained by:

(6.6)

where A+ = (A

TA)

-1A

T is the Moore-Penrose pseudo-inverse [165].

Page 126: Magnetic Induction Tomography for Imaging Cerebral Stroke

105 Discrete Inverse Problem

In the underdetermined case (A m×n

, m <

n), there are infinitely many solutions or no

solution to Ax = b. In the case of the former and full rank (r = m), the solution with the

smallest Euclidean norm ||x|| can be computed by:

(6.7)

where in this case, A+ = A

T (AA

T)

-1 is the Moore-Penrose pseudo-inverse [165,166]. In

general, the pseudo-inverse for under- and overdetermined problems (with or without

full rank) can be obtained using the singular value decomposition which is explained

later in this chapter. However, even in situations with full row or column rank,

computing x is not trivial as AAT and A

TA are ill-conditioned with respect to inversion,

i.e. the inverse problem is ill-posed. Further numerical difficulties can arise when the

right hand side is not consistent due to the errors in the measurements (i.e. b does not lie

in the range of A).

6.2.3. Ill-posedness

Hadamard [167,168] attempted to describe the generic properties of the inverse

problems which arise from physical and natural phenomena. A mathematical model of a

physical problem is well-posed if the solution of the mathematical model:

is consistent with the observations (in this case, the MIT signals).

is unique.

depends continuously on the data (observations).

The MIT inverse problem is severely ill-posed. Although by imposing some constraints

on the solution, one may ensure the validity of the first two criteria, it certainly fails the

third one. This means even small perturbations in the measurements, b, can cause large

changes in the solution, x. In practice this means for any given measurement precision,

there may be arbitrarily large changes in the obtained conductivity (since b is a

measured quantity, it is subject to various observational errors). The ill-posedness of the

linearized inverse problem is due to the fact that the matrices AAT or A

TA are ill-

conditioned with respect to inversion.

The ill-posedness of the problem does not mean that a meaningful approximate solution

cannot be computed. Rather, it implies that the standard methods in numerical linear

Page 127: Magnetic Induction Tomography for Imaging Cerebral Stroke

106 Discrete Inverse Problem

algebra for solving (6.3), e.g. pseudo-inverse, cannot be used in a straightforward

manner to compute a solution. Therefore, solving the inverse problem requires special

treatment. A common approach includes the regularization or truncation of a singular

value expansion. But first, computation of the sensitivity matrix and its reliability to

represent the model is discussed.

6.3. Sensitivity analysis

6.3.1. Sensitivity computation

The Jacobian or sensitivity distribution matrix may be computed numerically by altering

the electrical properties of each degree of freedom (i.e. each voxel or element) and

solving the forward problem. This way, the sensitivity of each point in the domain to the

conductivity change can be assessed. For instance, figure 6.4 illustrates a homogeneous

initial guess 0 = 1 S m-1

, where the conductivity of the perturbed element is raised to

twice of that in the domain (2 S m

-1).

This „perturbation method‟ requires the solution of multiple forward problems (once for

every perturbed element), and can be cumbersome, in particular, for 3D image

reconstruction. The sensitivity matrix may be pre-computed for single-step differential

image reconstruction algorithms. However, for the non-linear absolute imaging

algorithms, which requires the sensitivity to be updated after each iteration (section 6.8),

the use of perturbation method for the computation of sensitivity matrix would be

prohibitively long.

Figure 6.4. Perturbation method for sensitivity calculation; highlighted voxels represent the

perturbed elements (pixels).

1st perturbed pixel 2nd perturbed pixel

Page 128: Magnetic Induction Tomography for Imaging Cerebral Stroke

107 Discrete Inverse Problem

An alternative approach is to use the mutual impedance changes, Z, based on the

Lorentz reciprocity theorem and derive the generalized representation for each pair of

exciter-sensor coils in terms of the inner product of electric fields [169,156]. To this

end, similar to section 5.5 and neglecting the permittivity in (5.30), V for one pair of

coils can be expressed by:

(6.8)

where Ea and Eb are the electric fields produced when either of the coils is activated.

For a discretized domain comprising n homogeneous elements, then we have:

(6.9)

where k is the volume of k-th element. Assuming a current of I =

1

A, the sensitivity

entry Ak of the k-th element which measures the rate of voltage change Vk due to the

small change in the properties of that element is then defined as:

(6.10)

Note that k represents the change in the conductivity of the k-th element between a

and b. Therefore, using the same conductivity for a and b results in Ak being the

gradient of the non-linear model, F, at k.

Strictly speaking, what has been computed in (6.10) is a Fréchet derivative which is

obtained by neglecting higher order terms [125] and, hereinafter, is referred to as the

E.E sensitivity. The employment of this method to compute the sensitivity matrix is,

therefore, significantly faster than the perturbation method, which is a finite difference

approximation of the Fréchet derivative [125]. Compared to the perturbation method,

which requires the solution of n forward problems for each pair of coils (and

consequently each row of the sensitivity matrix), here the same entries using the E.E

sensitivity approach can be derived from the solution of two forward problems; solution

for the activated exciter, and solution for the sensor when employed as the exciters.

Page 129: Magnetic Induction Tomography for Imaging Cerebral Stroke

108 Discrete Inverse Problem

The entry Ak for one element not only depends on the electrical properties of the

material within that element, but also on the distribution in the whole sensing area since

the surrounding area affects the current flow in that element. Therefore, the initial guess

of conductivity distribution for computing the sensitivity matrix is of some importance.

The initial conductivity distribution can be assumed using some prior knowledge. For

instance, it can be taken as the typical distribution that is expected in the target. For a

low contrast target the nearest guess could be a uniform distribution.

On a regular grid of size x, the entries can be simply computed by:

(6.11)

where (Ea)k and (Eb)k are the electric fields within the k-th voxel, and Ne and Ns are the

total number of the exciter and sensor coils, respectively. The sensitivity matrix of the

entire system, A, has therefore as many columns as there are elements in the discretized

volume (k = 1,2,...,n), and as many rows as there are coil combinations or coil pairs (i =

1,2,...,m) where m = Ne Ns. Each row of A corresponds to the sensitivity of one coil

combination to all elements. Each column corresponds to the sensitivity of all coil

combinations to one element. In order to compute the sensitivity matrix based on the

reciprocity method, the FD solver was amended to implement the required sub-routines.

In order to compare the perturbation and E.E sensitivity matrices, a homogeneous

conductivity distribution of 1 S

m

-1 was considered for the cylinder shown in figure

6.1a. The E.E sensitivity matrix was computed simulating the MK1 system which

provides 256 coil combinations. The entries of an arbitrary column in the sensitivity

matrix, corresponding to the sensitivity of all coil combinations to one voxel in the

discretized cylinder, are plotted in figure 6.5 (thick black line). Note that the plot is

shown only for a limited number of the coil combinations for the convenience of

display.

The perturbation sensitivity was computed by raising the conductivity of the same

voxel, while the conductivity of the remaining voxels remained at 1 S

m

-1. Three levels

of conductivity change were considered for the perturbed voxel: 0.1, 0.5, and 1 S

m

-1

corresponding to 10, 50, and 100% conductivity change, respectively. The computed

sensitivity entries are also shown in figure 6.5.

Page 130: Magnetic Induction Tomography for Imaging Cerebral Stroke

109 Discrete Inverse Problem

Figure 6.5. Comparison of sensitivity matrices computed based on perturbation and E.E methods.

Percentage values correspond to conductivity change in the perturbed voxel with respect to the

background.

As can be seen, when the conductivity change in the perturbed voxel with respect to the

background is large ( 50%), the entries in the perturbation matrix deviate from those in

the E.E one. This is to be expected as the E.E method assumes small conductivity

changes, i.e. ignores the second-order terms with respect to the conductivity change (see

section 5.5). However, for a small conductivity change ( 10%), there is a good

agreement between the entries of the two sensitivity matrices.

If the MIT signals are normalized to the primary signal to obtain the phase changes, the

entries of the sensitivity matrix in (6.11) should also be divided by the primary signal

for the corresponding coil combination:

(6.12)

Therefore, the equation (6.3) will be in the form:

(6.13)

100 105 110 115 120 125 130 135 140

-6

-4

-2

0

2

4

x 10-8

Coil combination

A

E.E

10%

50%

100%

Page 131: Magnetic Induction Tomography for Imaging Cerebral Stroke

110 Discrete Inverse Problem

In order to examine the reliability of linearization in equations (6.3) and (6.4), an

investigation was carried out using the setup in figure 6.1 as follows:

i. Ex1 (exciter 1) and Se9 (sensor 9) in the MK1 system were considered to

represent a single channel system with opposite coils (for coil numberings see

Appendix A).

ii. State-I denotes the case of a homogeneous background conductivity distribution

of 0 = 1 S m-1

. The forward problem was solved for this distribution.

iii. The sensitivity matrix and voltage in the sensor coil, V0 = F( 0), were recorded.

Note that for a single channel system, there is only one sensor coil and the

voltage can be expressed as a scalar.

iv. An inclusion was added to the homogeneous background conductivity

distribution (see the diagram in figure 6.6). State-II denotes the conductivity

distribution 1 which includes the inclusion.

v. The conductivity of the inclusion was varied from 0.1 to 1.9 S m-1

in 0.1 steps;

equivalent to 90% perturbation with respect to the background.

vi. For each case, the forward problem was solved and the voltage V1 = F( 1) was

recorded.

vii. The difference between the voltages ( V = V1 – V0) is plotted against the

conductivity change (solid blue curve in figure 6.6). This is the voltage change

computed by the non-linear model, i.e. forward model.

viii. In the next step, using the linearized model V = b = Ax = A , the sensitivity

matrix was used to predict the voltage changes; values of V are shown as red

circles.

As can be observed, there is good agreement between the non-linear and linear models

when the conductivity change is small. However, as the conductivity changes become

larger, the predicted voltage changes by the linear model depart from those computed by

the non-linear model. At 90% conductivity change (inclusion of 0.1 S m-1

), the error in

the linear model was 35%. For large values of | | > 4 S m-1

(i.e. 400%, not shown in

the figure), the change in the voltage tends to saturate the change in the conductivity.

Steps (i) to (viii) were repeated for an inhomogeneous background as the State-I, where

the inclusion of 1.2 S m-1

was placed in the background of 1 S m-1

(see diagram in

figure 6.7). The sensitivity matrix, however, was computed for the homogeneous

Page 132: Magnetic Induction Tomography for Imaging Cerebral Stroke

111 Discrete Inverse Problem

background as before. Similarly, State-II was simulated for an added inclusion with up

to 90% conductivity change with respect to the background.

The results are plotted in figure 6.7. As expected, a similar behaviour was observed,

indicating that for small enough conductivity changes, the inverse problem can be

treated using the linear model. Examining other exciter-sensor coil combinations in the

MK1 system, revealed the same conclusion.

Figure 6.6. Comparison of non-linear and linear models; is contrast between added inclusion

and background.

Figure 6.7. Comparison of non-linear and linear models; is contrast between added inclusion

and background.

-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1-6

-4

-2

0

2

4

6

8

x 10-7

x =

V

F(1) - F(

0)

A0 (

1 -

0)

-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1-6

-4

-2

0

2

4

6

8

x 10-7

x =

V

F(2) - F(

1)

A0 (

2 -

1)

0

1

1

2

Page 133: Magnetic Induction Tomography for Imaging Cerebral Stroke

112 Discrete Inverse Problem

6.3.2. Ill-conditioning

One suitable numerical tool that can be used for the analysis of discrete ill-posed

problems is the singular value decomposition (SVD) of the sensitivity matrix A. The

SVD reveals the difficulties associated with the ill-conditioning of A. An idea of how

severely ill-conditioned the sensitivity matrix is, can be obtained by evaluating its

singular values. To this end, the SVD of A is computed as a decomposition of the form:

(6.14)

where Um m = (u1,u2,...,um) and Wn n = (w1,w2,...,wn) are unitary matrices with

orthogonal columns, UTU = Im and W

TW = In. The matrix is the m by n matrix

diag(s1 , s2 , .... ) padded with zeros and s1 s2 .... > 0 [170,171,172]. The numbers si

are the singular values of A while the vectors ui and wi are the left (output) and right

(input) singular vectors of A, respectively. The matrix A is said to be ill-conditioned if

both of the following criteria are satisfied:

Singular values of A decay to zero.

Ratio between the largest and smallest non-zero singular values (s1/sm) is large.

The singular values of the sensitivity matrix, which was computed for the example in

the previous section, are plotted in figure 6.8. The singular values fell abruptly by

several orders of magnitude, indicating that the sensitivity is severely ill-conditioned

and the inverse problem is, therefore, ill-posed (s1/s256 = 9.86×106).

Using (6.14), matrix A can be expanded into the summation of a number of matrices:

(6.15)

As a result, the application of a linear transformation in the form Ax is equivalent to the

application of a sequence of linear transformations which have a contribution of

decreasing magnitude. Notice that coefficients si are ordered in terms of magnitude. The

SVD can be used as a tool for solving over- and underdetermined problems. The LS

solution of Ax = b corresponds to:

Page 134: Magnetic Induction Tomography for Imaging Cerebral Stroke

113 Discrete Inverse Problem

(6.16)

The solution of this problem is equivalent to [168]:

(6.17)

where A+ is the pseudo-inverse given by:

(6.18)

Since W and U matrices are unitary, their inverse is simply their transpose, and + is the

transpose of with the non-zero singular values si replaced by their reciprocals 1/si.

Equation (6.18) generalises the earlier expressions given in (6.6) and (6.7) for the over-

and underdetermined problems with full row/column rank and, in the case of exact

arithmetic, the results are the same. Note that in the overdetermined case with r = n,

there is only one solution to , and this solution has minimal norm. In

the overdetermined case without full column rank and in the underdetermined case, r <

n and there are an infinite number of solutions to , and the solution

with minimal norm is sought.

Equation (6.17) is referred to as the singular value expansion of the solution from

which, multiple results can be observed:

The solution is now a summation of r components instead of simply A+b.

For each component of the series, the singular value si is in the denominator and

a very small value can potentially have a large impact on the solution.

Since each term uiTb/si is a scalar coefficient, it leaves the solution x as

essentially a linear combination of wi vectors. Notice that x represents the

reconstructed MIT image. Therefore, the right singular vectors wi (i = 1,...,r) are

the orthogonal bases comprising the image.

Among the singular values of the sensitivity, the smallest ones are the most vulnerable

to noise. These are often extremely small, and when contaminated by the numerical

error and then inverted, they grow into drastically magnified errors, causing instability

Page 135: Magnetic Induction Tomography for Imaging Cerebral Stroke

114 Discrete Inverse Problem

in the solution. They also amplify the measurement noise in data vector b. This can lead

to 'blowing-up' of the solution (artificially large values in the solution), and introducing

artefacts in the reconstructed image. This explains why computing the solution using the

equation (6.7) may lead to poor quality results.

Similar to the work by Lionheart et al for EIT [173], some arbitrary singular vector

maps are shown in figure 6.8, adjacent to their corresponding singular values. Each

singular vector plot was constructed by placing the elements in column vector wi in the

appropriate voxel positions in the grid to reconstruct the 3D volume, and then summing

the absolute values in voxels along the height of cylinder (z-direction) to obtain the 2D

(x-y) maps.

It can be appreciated from the plot that the singular vectors corresponding to the large

singular values contain information mainly related to the periphery and locations close

to the surface of the domain. As the magnitude of the singular values decrease, the

corresponding singular vectors tend to contain information about the centre of the

domain and/or details of spatial resolution. Note that by using absolute values to obtain

2D maps, the changes in sign of the singular vector components are not observed.

Figure 6.8. Singular spectrum plotted on a logarithmic scale, normalized to the largest singular

value. Also shown some arbitrary singular vector maps; index = 1, 3, 60, 136, 223.

0 50 100 150 200 25010

-7

10-6

10-5

10-4

10-3

10-2

10-1

100

Singular value index

Norm

alis

ed s

ingula

r valu

e

Page 136: Magnetic Induction Tomography for Imaging Cerebral Stroke

115 Discrete Inverse Problem

In order to eliminate the cause of instability in equation (6.17), filtering out the small

singular values is essential; a process which also improves the conditioning of the

sensitivity matrix. Removing singular values from the sensitivity matrix, however,

causes the sensitivity to drop. Since some of these small singular values may represent

important details and desirable parts of the images, by filtering out the singular values

the information within their singular vector is lost, and the reconstructed image may

suffer from a low spatial resolution. The filtering process is often performed with

numerical tools such as TSVD and Tikhonov regularization.

6.4. Truncated singular value decomposition

As previously discussed, complications arise due to the terms with small singular values

in the series in (6.17). Therefore, the most straightforward way to avoid this situation is

to eliminate these components entirely from the summation in (6.17). This can be

achieved by setting a truncation level for the minimum magnitude of singular values

that are to be retained in the summation. Let 0 < k < r be the number of terms retained

after truncation. The solution then becomes:

(6.19)

which is the truncated singular value decomposition (TSVD) solution corresponding to

subject to [168]. Here, k is the truncation level, and

(6.20)

By truncating the small singular values, all the noise components parallel to uk+1, ...,ur

which would be strongly amplified, are rejected. The equation (6.19) can be rewritten

using a filtering function, f, in the form:

(6.21)

Page 137: Magnetic Induction Tomography for Imaging Cerebral Stroke

116 Discrete Inverse Problem

In order to examine the effect of truncation, consider the case where State-I corresponds

to a uniform background of 1 S m-1

, and State-II consists of the same background, but

with the addition of two inclusions (0.5 and 1.5 S m-1

). The sensitivity matrix is

computed according to State-I, and the inverse problem is solved with three different

levels of truncation in (6.19): k = 100, 180, 256 corresponding to sk/s1 = 1.4 10-3

,

4.0 10-5

, 1.0 10-7

(see figure 6.9). Note that k = 256 indicates no truncation is done, and

all the singular values are included in the solution.

Figure 6.9. Levels of truncation of singular values.

The first column in figure 6.10a shows the target to be imaged, and the second to fourth

columns are the reconstructed images with different levels of truncation. As can be

observed, with a high level of truncation (k = 100), more than half of the singular values

are dropped (also see figure 6.9); this results in a blurred but relatively smooth image.

Incorporating 180 singular values, details of the inclusions are clearly better visualised.

Both the shape and magnitude of the reconstructed conductivity change are improved;

see figure 6.10b which shows a 1D image of the conductivity values along a line in x-y

cross-sectional plane.

A further improvement is observed when all the available information is used (k = 256).

For this case, the recovered inclusions are slightly better defined, are less displaced

towards the boundaries, and the maximum magnitude of 0.3 and 0.42 S m-1

is obtained

0 50 100 150 200 25010

-7

10-6

10-5

10-4

10-3

10-2

10-1

100

Singular value index

Norm

alis

ed s

ingula

r valu

e

k = 100

k = 180

k = 256

Page 138: Magnetic Induction Tomography for Imaging Cerebral Stroke

117 Discrete Inverse Problem

for positive and negative conductivity changes, respectively. The fact that even with

incorporating all 256 singular values, and using the noise-less data, the reconstructed

image is still different from the target clearly indicates the shortage of information (the

inverse problem being underdetermined).

Figure 6.10. Reconstructed images using TSVD; (a) columns from left: target and images with

different levels of truncation, rows from top: x-y and y-z cross-sectional planes, colour-map shows

conductivity change in S m-1

, (b) 1D representation of target and reconstructed conductivity

changes along a line in x-y plane.

Next, the simulated data is contaminated with a relatively small level of error of 1 10-3

0.1%. A randomly generated noise of 60 dB (standard deviation) is added to the MIT

signals (for further details see Appendix C.6).

0 -8 -6 -4 -2 0 2 4 6 8 10

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

Length (cm)

Co

nd

uctivity c

ha

ng

e (

S/m

)

True values

k = 100

k = 180

k = 256

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

Target k = 100 k = 180 k = 256

y-z

x

-y

(a)

(b)

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

Page 139: Magnetic Induction Tomography for Imaging Cerebral Stroke

118 Discrete Inverse Problem

In this case, the introduction of noise leads to the situation where the right hand side

vector b is no longer consistent. Therefore, as the problem is underdetermined and

nearly rank deficient, no meaningful solution can be found using either the pseudo-

inverse A+ = A

T (AA

T)

-1. In the case of SVD without truncation, we have:

(6.22)

which is large due to the appearance of small singular values and an inconsistent right

hand side. However for the TSVD, the resulting solution has a smaller norm [168]:

(6.23)

and, consequently, the use of TSVD does still allow meaningful solution to Ax = b to be

obtained.

The results for the noisy data are shown in figure 6.11. For the high level of truncation

(k = 100), effect of noise is insignificant as the incorporated singular values are above

the noise level (see figure 6.9). For the intermediate level of truncation, although a

portion of the singular values fall below the noise level, the reconstructed image is

almost identical to its noise-free counterpart. However, a closer scrutiny of different 1D

images reveals small differences between the two images, with the noise-free image

being more accurate.

With no truncation (k = 256), the solution is unstable and large artefacts, both in terms

of magnitude and volume, can easily be observed in 2D representation of the image.

The visualization of the inclusions is unfeasible as the artefacts are dominant. Similarly,

large oscillations can be seen in the 1D image.

While TSVD cuts the data, the generalized Tikhonov regularization can do both

filtering and implementing a-priori information. In the next section, the Tikhonov

regularization method is first explained in the standard form, followed by the

generalized form.

Page 140: Magnetic Induction Tomography for Imaging Cerebral Stroke

119 Discrete Inverse Problem

Figure 6.11. As figure 6.10, but with 60 dB noise added to simulated data.

6.5. Tikhonov regularization

6.5.1. Standard Tikhonov

The Tikhonov regularization [173], is often presented as a technique for obtaining

approximate LS solutions to overdetermined problems that are ill-posed. It has

advantages over the TSVD approach as the SVD of a matrix does not have to be

computed. The disadvantage is that it requires a regularization parameter that must be

tuned to individual problems and applications. An approximate LS solution to the

0 -8 -6 -4 -2 0 2 4 6 8 10

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

Length (cm)

Co

nd

uctivity c

ha

ng

e (

S/m

)

True values

k = 100

k = 180

k = 256

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

Target k = 100 k = 180 k = 256

y-z

x

-y

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

(a)

(b)

Page 141: Magnetic Induction Tomography for Imaging Cerebral Stroke

120 Discrete Inverse Problem

overdetermined problem Ax = b with the standard Tikhonov regularization is

characterized by minimizing the objective function:

(6.24)

where is the scalar regularization parameter and ||x||2 is the penalty term, constraining

the solution. Regularization can be viewed as a tool to trade-off the influence of the

measurements on the solution against its compliance with the prior assumptions, i.e. not

letting ||x||2 get too big. The value of controls this trade-off; as 0, x tends to the LS

solution A+b. The solution of the minimization (6.24) is given by:

(6.25)

Here, In and Im are identity matrices of the appropriate sizes. The positive definite

matrices (AAT+ In)

-1 and (A

TA+ Im)

-1 can be interpreted as filters in the image and data

spaces respectively: the former is back-projection followed by spatial filtering, and the

latter is filtering of the data followed by back-projection [174].

In order to see whether the Tikhonov regularization can also be applied to the solution

of underdetermined systems note that the Tikhonov solution (6.25) is equivalent to the

solution in the SVD form in (6.21) when the following filtering is employed [168]:

(6.26)

Note that for si >> then f (si) 1, and if si << then f

(si) 0. Therefore, the Tikhonov

regularization has a similar effect as the TSVD. We know that the application of the

TSVD leads to a LS solution satisfying subject to .

Therefore, it can be concluded that the Tikhonov regularization can also be applied to

the solution of ill-posed underdetermined systems [175]. Computationally, the second

of the rearrangements shown in (6.25) will be advantageous for underdetermined

problems as it leads to the creation of a smaller matrix. In the following, only the so-

called left inverse form of the Tikhonov based schemes will be presented, although

advantageous right inverse form are also available [168].

By replacing the definition of f (si) given by (6.26) together with (6.21), the modified

singular values of the sensitivty matrix may be computed. These are shown in figure

Page 142: Magnetic Induction Tomography for Imaging Cerebral Stroke

121 Discrete Inverse Problem

6.12 which allow comparison with the results for the TSVD in figure 6.9. As the

original singular values decrease, the modified singular values grow larger and,

therefore, their effect in the solution (6.19) is inceasingly diminished.

Three different values of are adopted in order to examine how the singular values are

affected by different levels of regularization. Note that = 0 denotes no adjustment,

similar to no truncation for k = 256.

Figure 6.12. Modified singular values (si + /si) for 3 different levels of regularization.

A similar method, called the damped SVD method [176], replaces si2 by si in the filter

factors defined by (6.26). This method has a similar effect, but with a more gradual

transition between including large and rejecting small singular values. The use of the

dynamic (variable) in (6.26) has also been suggested for seismic tomography [177].

Reconstructed images using the standard Tikhonov method are shown in figure 6.13. As

expected, for the high and intermediate levels of regularization ( = 1 10-15

to 1 10-18

),

the inverse problem is stable, but the images are blurred. With no regularization, the

noise added to the data is amplified in the solution which, as before, is dominated by the

noise. The Tikhonov images are almost identical to their counterparts reconstructed by

the TSVD method.

0 50 100 150 200 25010

-12

10-10

10-8

10-6

10-4

10-2

Singular value index

Sin

gula

r valu

e

= 1 10-15

= 1 10-18

= 0

Page 143: Magnetic Induction Tomography for Imaging Cerebral Stroke

122 Discrete Inverse Problem

Figure 6.13. Reconstructed images using standard Tikhonov regularization with 60 dB noise added

to simulated data. Columns from left: target and images with different levels of regularization

indicated by value of . Rows from top: x-y and y-z cross-sectional planes. Colour-map shows

conductivity change in S m-1

.

6.5.2. Generalized Tikhonov

The penalty term ||x||2 in the standard Tikhonov form prevents extreme values of the

reconstructed conductivity change, but does not enforce any smoothness on the values

[173]. As an alternative, ||Rx||2 can be used for the penalty term where R is called the

regularization matrix, which can introduce additional constraints on the computed

conductivity distribution.

The choice of R depends on the particular application being considered. However,

common choices involve approximate first- or second-order (finite difference)

differential operators. Further refinement of the Tikhonov regularization may be

achieved by penalizing the differences between the solution, x, and some reference

background, xref, which can include some known non-smooth behaviour. Finally, there

is the possibility that we may not wish to fit all measurements to the same accuracy, in

particular as some may have larger errors than others. This can be achieved by

considering some diagonal weighting matrix [173]. The generalized Tikhonov

procedure is then of the form:

(6.27)

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

Target = 10-15 = 10-18 = 0

y-z

x

-y

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

Page 144: Magnetic Induction Tomography for Imaging Cerebral Stroke

123 Discrete Inverse Problem

which reduces to the standard Tikhonov form for the choice of R = I, W = I, and xref =

0. The solution for (6.27) will be [34,178]:

) (6.28)

Throughout this study, the weighting matrix = I has been employed. Notice that xref is

different from the initial conductivity 0 corresponding to State-I. Recalling from

equations (6.2) and (6.3), we have x = = - 0. By applying ||x - xref||2, the

difference between the reconstructed conductivity change, x, and the expected

conductivity change, xref, is penalized. For instance, consider a case when the a-priori

information includes some anatomical features such as the approximate location of

different tissue types in the domain. Then in time-difference imaging of haemorrhage,

the conductivity change for the elements representing any tissue type in which the

bleeding is unlikely to occur (e.g. bones) can be set to zero in the vector xref.

Similar to the standard form, the generalized Tikhonov method can be explained in

terms of the singular values. However, in this case, the generalized singular value

decomposition (GSVD) needs to be employed. For further details on the GSVD see e.g.

[176,168].

6.5.3. Neighbouring matrix

A common choice for the regularization matrix is to use a discrete approximation of the

difference operators on the discretized volume. This corresponds statistically to

assuming a correlation between neighbouring voxel values. Elements sharing a face as

in figure 6.14a (e.g. [125]) or at least a node as in figure 6.14b (e.g. [179]) may be

considered neighbours. Hereinafter, the former is called R-face, and the latter R-node.

Figure 6.14. Correlation between neighbouring elements with common face (a), common node (b),

and a combination of faces and edges (c) in a 2D image. Values shown in voxels correspond to a

central finite difference approximation of Laplacian operator.

-1 -1 -1

-1 8 -1

-1 -1 -1

-1

-1 4 -1

-1

-1 -2 -1

-2 12 -2

-1 -2 -1

(a) (b) (c)

Page 145: Magnetic Induction Tomography for Imaging Cerebral Stroke

124 Discrete Inverse Problem

For a discrete Laplacian operator in a regular cubic mesh comprising n voxels, each

element in the regularization matrix is defined as:

(6.29)

Here, N is the number of elements neighbouring the central element. Another type of

operator can be realised by taking into account the degree of neighbourhood or

correlation between the elements (figure 6.14c).

Applying the penalty term ||Rx||2 enforces some degree of smoothness on the image by

preventing the neighbouring elements having sharp conductivity contrasts. Those

regularization matrices that can enforce the smoothness while preserving the desired

edges and genuine sharp contrasts, are introduced later for absolute imaging algorithms.

Detailed description of neighbouring regularization matrices can be found in [166].

Images were reconstructed using the generalized Tikhonov method, where R is the

neighbouring matrix (R-face) and xref =

0. For a high level of regularization (

= 10

-15),

the smoothness is significantly enforced on the image which makes the internal

inclusions nearly invisible (see figure 6.15).

Figure 6.15. As in figure 6.13, but reconstructed using (6.28) with R-face and xref = 0.

The magnitude of the obtained conductivity changes for the inclusions are comparable

to that for the image reconstructed using the standard Tikhonov in figure 6.13. For =

10-18

, the inclusions are clearly visible, while the background is smooth. The small

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

Target = 10-15 = 10-18 = 0

y-z

x

-y

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

Page 146: Magnetic Induction Tomography for Imaging Cerebral Stroke

125 Discrete Inverse Problem

artefacts, which appreaed in the image reconstructed using the standard Tikhonov, are

removed here due to the effect of R. Finally, for no regularization ( = 0), the

reconstructed image is entirely degraded by the presence of artefacts.

In order to compare the results obtained from the standard and generalized Tikhonov

methods, 1D plots of the conductivity change are considered. Figure 6.16a shows the

values along a cross-sectional line cutting through the inclusions. As can be seen, while

the magnitude of the reconstructed values for the inclusions is comparable in both

methods, the transition from negative to positive conductivity changes is smooth in the

results obtained from the generalized Tikhonov with R-face. Also, the positions of the

features, where the maximum magnitude is obtained, are more accurately recovered

using the generalized Tikhonov form.

Figure 6.16. Comparison of standard and generalized Tikhonov: 1D plots showing reconstructed

conductivity changes ( = 1 10-18

) along cross-sectional lines in y-direction (a) and x-direction (b).

-10 -8 -6 -4 -2 0 2 4 6 8 10-0.06

-0.04

-0.02

0

0.02

0.04

0.06

0.08

Length (cm)

Co

nd

uctivity c

ha

ng

e (

S/m

)

-10 -8 -6 -4 -2 0 2 4 6 8 10

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

Length (cm)

Co

nd

uctivity c

ha

ng

e (

S/m

)

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

(a)

(b)

True values

Standard Tikhonov

(Identity)

Generalized Tikhonov

(R-face)

Page 147: Magnetic Induction Tomography for Imaging Cerebral Stroke

126 Discrete Inverse Problem

Figure 6.16b shows the 1D plots corresponding to a different crossing line. This plot

shows the smoothness of the reconstructed conductivity changes in the background. For

the standard Tikhonov with Identity matrix, some artefacts with magnitudes as large as

0.08 S m

-1 can be observed. The plot for generalized Tikhonov method shows an

smoother conductivity change in the background.

6.6. Error estimation and regularization parameter

Figure 6.11b illustrated the importance of selecting the regularization parameter

properly. For very low levels of regularization, reconstructions will have oscillatory

artefacts due to the noise amplification. In contrast, with too much regularization, the

reconstructed images tend to be smooth or damped. The study of optimal regularization

parameters is an active area in ill-posed problems. The majority of the suggested

methods for this purpose exploit certain types of the error in solution of the inverse

problem, and attempt to minimize it. In order to derive the mathematical representation

of the error, let us express the linearized model in the form:

(6.30)

where xtrue is the sought true distribution in the target, and represents the discrete

noise vector with zero mean (E( ) = 0) and2 variance (standard deviation = ).

Following (6.25), the error in the solution computed for regularization parameter can

be expressed by:

(6.31)

where e is the solution error [180]. Note that in practice, this error is not available as

xtrue is unknown. Therefore, it cannot be used for optimizing the regularization

parameter. However, when comparing the sensitivity of spherical and cylindrical MIT

arrays for stroke imaging in chapter 8, e is employed for selecting ; this helped to

obtain the best feasible images for each system.

It is worth mentioning in passing that while a mathematical method attempts to

minimize the errors in the solution, it may not necessarily result in the best potentially

Page 148: Magnetic Induction Tomography for Imaging Cerebral Stroke

127 Discrete Inverse Problem

useful images. For instance, if the image artefacts consistently appear in certain regions

of the head, which can be excluded as the bleeding sites by a radiologist, then a noisy

image with a clear stroke may be preferable from a clinical point of view, even if with

larger mathematical errors. Examples of such images are provided in chapter 8.

The predictive error, P , estimates the discrepancy between the predicted voltage by the

linearized model and that obtained from the true solution as:

(6.32)

where C = AB . Again, this error cannot be employed for error minimization as it

involves the usually unknown true solution. For the purpose of optimising , the

regularized residual may be defined as:

(6.33)

Several methods have been suggested for selecting the regularization parameter. Two of

the most commonly used methods are:

Generalized cross validation (GCV) method proposed by Wahba [181].

L-curve method proposed by Hansen [168].

Each method is briefly explained in this section. Detailed studies and comparisons of

these methods can be found in the work of Hansen [168] or Vogel [180].

6.6.1. GCV

The GCV method is based on statistical considerations [181]. It attempts to minimize

the predictive error, P . Since xtrue is not usually available, the GCV method

alternatively aims at minimizing the function:

(6.34)

which is computable (for further details see Appendix C.1).

To evaluate the GCV function, the regularized inverse C = AB = AW++

UT must be

computed for different values of , where ++

is the inverse of if the singular values

Page 149: Magnetic Induction Tomography for Imaging Cerebral Stroke

128 Discrete Inverse Problem

are replaced by (si + /si). Also, the residual r can be computed using (6.33). For

differential imaging, this can be easily obtained by using the SVD of A.

However, since the sensitivity matrix is updated for absolute imaging (explained later in

this chapter), this can be a cumbersome process when computing the SVD of A must be

repeated. In addition, the GCV function can have a very flat minimum, making it

difficult to isolate a clear minimum [182].

To examine the efficiency of the GCV method, three levels of noise were added to the

simulated data in previous sections. For the standard Tikhonov method and each data

set, a plot of the GCV function for different values of is shown in figure 6.171.

For a relatively low level of 60 dB noise, the optimized is 3.6 10-18

. As the noise level

increases, a higher level of regularization is suggested by the GCV method. For 40 dB

noise, is increased by approximately a factor of 100.

Figure 6.17. Regularization parameter selection using GCV method for three noise levels.

In order to illustrate the importance of using the correct values of for a specified level

of noise, the images presented in figure 6.18 were reconstructed using the computed

1 For the GCV and L-curve methods, a regularization toolbox, which is available as freeware, was

employed. The toolbox is developed in Matlab [233,232]. The command 'fminbnd', which finds the

minimum of a function in a given domain, is used in the toolbox to find the optimum value of .

10-25

10-20

10-15

10-10

10-21

10-20

10-19

10-18

10-17

10-16

10-15

G(

)

3.6050e-018

6.4012e-016

3.4753e-017

40 dB

50 dB

60 dB

Page 150: Magnetic Induction Tomography for Imaging Cerebral Stroke

129 Discrete Inverse Problem

values of suggested by the GCV method. The fourth column from left, shows the

image with 40 dB added noise, but reconstructed using the optimized for the 60 dB

case, i.e. 3.6 10-18

. Clearly, this level of regularization is low for 40 dB noise as the

image is dominated by the artefacts. To avoid these artefacts, a larger is required, as

suggested by the GCV method (6.4 10-16

). However, the use of a larger value of has

the disadvantage that the image becomes more blurred.

Figure 6.18. Reconstructed images by standard Tikhonov method and using suggested value of by

GCV method. Colour-bar in S m

-1.

6.6.2. L-curve

The L-curve method is a more recent technique for choosing [168]. The general idea

is to display the trade-off between the two quantities of the objective function in (6.24);

i.e. the penalty term versus the corresponding residual error norm for each of a set of

regularisation parameter values. This results in a plot which typically has an L-shaped

curve; thus the L-curve name. Intuitively, the best regularisation parameter should lie on

the corner of the L-curve, since:

for larger values of , the residual increases rapidly without a significant

reduction in the norm of the solution.

for smaller values of , the norm of the solution increases rapidly without much

decrease in the residual.

The sharpness of this corner varies from problem to problem; not all curves have a

pronounced enough corner to allow unambiguous selection of [182].

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

60 dB 40 dB 40 dB

y-z

x

-y

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

5 10 15 20

2

4

6

8

10

12

14

16

18

20

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5Target = 3.6 10-18 = 6.4 10-16 = 3.6 10-18

Page 151: Magnetic Induction Tomography for Imaging Cerebral Stroke

130 Discrete Inverse Problem

Figure 6.19. Plots of L-curve for selecting the regularization parameter for three noise levels.

Similar to the previous section, the optimized value of was computed for three levels

of noise using the L-curve method. The suggested values are close to those obtained by

the GCV method (see figure 6.19).

When attempting the generalized Tikhonov regularization with R-face, however, both

methods failed in predicting an optimized value for . Figure 6.20a indicates that care

must be taken when examining a range of regularization parameters to minimize the

GCV function. The GCV function has points of inflection, making the automatic

detection of the desired minimum more difficult. For 50 dB noise, the smallest value in

the range ( 9 10-37

) is suggested as the optimized , whereas the desired local

minimum occurs at a larger value ( 10-20

). The optimized values for all three noise

levels were too small resulting in under-regularized solutions. This is in line with the

observations of others [182,183,184].

The L-curve method resulted in the same value of ( 1.6 10-15

), regardless of the noise

level; the corners for all curves were coincident (figure 6.20b). However, the solution

obtained for this choice of would be over-regularized. Notice that other L-corners

occur for each curve which depend on the noise level (shown by green lines), but the

algorithm has found the sharpest. In general, the L-curve criterion does not work well

when the solution has a higher degree of smoothness [185], which is the case when

applying the neighbouring matrices.

10-10

10-9

10-8

10-7

10-6

10-5

10-2

10-1

100

101

102

103

residual norm || A x - b ||2

so

lutio

n n

orm

|| x || 2

2.2183e-018

5.6477e-017

40 dB 50 dB 60 dB

3.9390e-016

Page 152: Magnetic Induction Tomography for Imaging Cerebral Stroke

131 Discrete Inverse Problem

Figure 6.20. Regularization parameter selected by GCV (a) and L-curve (b) methods for three noise

levels (40, 50, and 60 dB).

6.7. Frequency-difference imaging

Although time-difference MIT imaging, at a fixed frequency, could potentially be useful

for many applications, it cannot be applied for imaging when:

No change in the conductivity distribution of the target occurs.

The data set for one of the two states occurring in time, State-I or State-II, is

unavailable.

The latter, for instance, applies to an initial diagnosis of haemorrhage when no previous

data-set for the patient is available. Frequency-difference imaging may alternatively be

attempted. This relies on being able to distinguish the frequency signature of the

10-40

10-35

10-30

10-25

10-20

10-15

10-21

10-20

10-19

10-18

10-17

10-16

G(

)

4.4128e-022

8.2933e-037

8.5527e-020

40 dB

60 dB

50 dBlocal

minima

10-10

10-9

10-8

10-7

10-6

10-5

10-6

10-4

10-2

100

102

104

106

108

residual norm || A x - b || 2

so

lutio

n s

em

i-n

orm

|| L

x || 2

5.5e-22 1.1e-21 4e-19

1.663e-15

60 dB 50 dB 40 dB

(a)

(b)

local

minimum

Page 153: Magnetic Induction Tomography for Imaging Cerebral Stroke

132 Discrete Inverse Problem

constituents of the target from each other. Note that in order to be able to apply

frequency-difference imaging, at least one of the constituents must have a frequency-

dependent conductivity.

In order to form the reference data set for State-I, the measurement/simulation is carried

out for one frequency (f1). Subsequently, data for State-II is acquired at a different,

usually higher, frequency (f2). Throughout this thesis, the phase changes are employed

as the MIT data sets for frequency-difference imaging. Of interest is the change in tissue

conductivity, , as the frequency changes from f1 to f2. However, as the equation

(2.12) shows, the change in the signal's phase is proportional to both and f.

A corrected phase change, , is computed so as to represent only the change in

conductivity:

(6.35)

where V1 and Vp1 are the State-I voltages measured at frequency of f1, and V2 and Vp2

are for State-II measured at frequency of f2. The sensitivity matrix must be computed

accordingly at frequency f2 (Scharfetter et al [186] applied a similar correction

procedure to the voltages).

Once the corrected phase difference has been obtained for all coil combinations ( ),

the linearized inverse problem is then treated in the same way as discussed in the

foregoing part of this chapter (see equations (6.12) and (6.13)). As before, the linearized

model assumes that there are no large conductivity changes between the two states. In

other words, the sensitivity matrix computed for one frequency can be used for solving

the inverse problem if the morphology of the sensitivity map does not depend on the

frequency [186].

If f2 > f1 in (6.35), and given the fact that the conductivities of all known biological

materials either increase or remain unchanged as the frequency increases, the estimated

conductivity changes by the inverse problem are expected to be non-negative. This

physical constraint may be applied as a-priori information by truncating the negative

values from the estimated conductivity changes before plotting the image.

Page 154: Magnetic Induction Tomography for Imaging Cerebral Stroke

133 Discrete Inverse Problem

Table 6.1. Conductivities (in S m

-1) of white matter and blood at 1 and 10 MHz, and conductivity

change between two frequencies.

1 MHz 10 MHz 10 MHz - 1 MHz

Background (white matter) 0.1 0.16 0.06

Inclusion (blood) 0.8 1.1 0.3

The constituents of the cylinder with inclusion (figure 6.21a) were considered to be

white matter (representing brain) and blood (representing haemorrhage). Two

frequencies of 1 and 10 MHz were considered. The conductivity values are given in

table 6.1. The acquired data for State-I and State-II corresponded to 1 and 10 MHz,

respectively.

Frequency-difference images were reconstructed from the simulated phase changes with

1 m added phase noise, and using the generalized Tikhonov method (R-face, xref = 0)

and (6.35). Again, neither of the GCV and L-curve methods worked properly. The

solution was slightly under-regularized for both cases ( GCV = 1.3 10-10

and L-curve =

2.9 10-11

). A better image in terms of the smoothness was obtained for an empirically

adjusted value of = 1 10-9

.

Figure 6.21. Frequency-difference imaging; (a): target, (b) and (c): reconstructed images using

generalized Tikhonov method (1 m added phase noise, R-face, xref = 0, = 1 10-9

), but plotted with

different colour-bars in S m

-1.

5 10 15 20

2

4

6

8

10

12

14

160

0.05

0.1

0.15

0.2

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.05

0.1

0.15

0.2

0.25

0.3

5 10 15 20

2

4

6

8

10

12

14

160

0.05

0.1

0.15

0.2

0.25

0.3

5 10 15 20

2

4

6

8

10

12

14

160

0.05

0.1

0.15

0.2

0.25

0.3

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

5 10 15 20

2

4

6

8

10

12

14

160

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

5 10 15 20

2

4

6

8

10

12

14

160

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.05

0.1

0.15

0.2

0.25

5 10 15 20

2

4

6

8

10

12

14

160

0.05

0.1

0.15

0.2

0.25

5 10 15 20

2

4

6

8

10

12

14

160

0.05

0.1

0.15

0.2

0.25

0.3

y-z

x

-y

(a) (b) (c)

Page 155: Magnetic Induction Tomography for Imaging Cerebral Stroke

134 Discrete Inverse Problem

As can be seen in figure 6.21b, although the inclusion is reconstructed, it can be barely

seen, and the obtained conductivity change values are under-estimated ( 0.1 S m

-1).

The under-estimation also happened in time-difference images reconstructed in the

previous sections. However, it is more evident here. The inclusion may be better

visualized if the maximum limit in the colour-bar is reduced as in figure 6.21c.

If State-II measurements are carried out for multiple frequencies (multi-frequency

imaging), more information will be available, which may allow the spectroscopy of the

target; e.g. biological tissue characterization [187]. This was not pursued in this thesis.

6.8. Absolute imaging

Absolute imaging is an attempt to quantify the actual conductivity distribution of the

target relative to empty space (0 S m

-1), where the conductivity of the target is supposed

not to change during the data acquisition. In absolute imaging, there is only one MIT

data set available. While single-step linear methods are useful for differential imaging,

absolute imaging is likely to require a non-linear algorithm to be successful. This is due

to the fact that the use of linear approximation is only valid for small deviations from

the reference conductivity (see figure 6.6).

In medical applications, the conductivity contrasts can be considerably large; e.g.

contrast between CSF (2 S

m

-1) and skull (0.08 S

m

-1) at frequency of 10 MHz. A

linearized model may be able to calculate a small change in conductivity between two

states, but reconstructing the absolute conductivity values will require an iterative non-

linear solution.

6.8.1. Iterative non-linear solution

Let us assume the voltage Vm is measured for an unknown conductivity distribution .

Revisiting equation (6.2), this voltage can be expressed by a Taylor expansion around

some initial conductivity 0. In equation (6.3), the voltage V0, which corresponded to

the measured voltage for State-I, is not available in absolute imaging. Instead, it can be

estimated by solving the forward problem for the known initial guess 0 (usually a

homogeneous distribution in lack of any a-priori information).

Page 156: Magnetic Induction Tomography for Imaging Cerebral Stroke

135 Discrete Inverse Problem

The sensitivity matrix is also computed for the same conductivity distribution. By

setting x = = - 0 and b = V = Vm - V0, the absolute values, , can then be

calculated by, for example, the standard Tikhonov method as in (6.25):

(6.36)

In presence of a large contrast, the estimated absolute values may not be accurate as the

linearized approximation fails to properly model the problem. However, the new

estimated conductivity distribution, , can be employed as a closer guess to solve the

linearized model with an updated sensitivity matrix. This results in an iterative

algorithm as:

(6.37)

where k+1 is the conductivity distribution computed at (k+1)-th iteration, and k is the

conductivity estimated by the previous iteration, which is used as the initial guess for

the current iteration. The terms Vk and Ak are the voltage vector and sensitivity matrix

calculated by solving the forward problem for the conductivity distribution k.

Therefore, as stated by Lionheart et al [173], the essence of a non-linear algorithm is to

repeat the process of calculating the sensitivity and solving a regularized linear model. It

starts by taking advantage of the known voltages due to an initial guess for the

conductivity distribution, and approaches the optimum solution in an iterative manner.

Obviously, the iteration procedure must be conducted in a way to improve the

convergence of the solution. This may be done by adopting the proper choice of

regularization level, and applied constraints (further discussed later). However, as will

be shown, in presence of a high level of noise in the data, the solution may not

converge.

In absolute imaging, similar to frequency-difference imaging, negative values can be

truncated from the solution as a-priori information. This may be applied after each

iteration prior to using the most recent conductivity distribution for the computation of

the sensitivity matrix. A flowchart for the iterative image reconstruction algorithms is

shown in figure 6.22.

Page 157: Magnetic Induction Tomography for Imaging Cerebral Stroke

136 Discrete Inverse Problem

Figure 6.22. Flowchart for iterative non-linear MIT image reconstruction.

There is no single correct choice for the stopping criterion. One can monitor the

difference between the voltages ||rk||2 = ||Vm – Vk||

2 and the iterations may be stopped

when the error reaches a pre-defined tolerance value which indicates that the assumed

conductivity, k, and the absolute values, k+1, are close enough. For instance, the

Morozov stopping criterion suggests that the tolerance value can be estimated from the

known level of noise in the measured data [168]. Various types of stopping criteria can

be found in [168,180,188,189]. As an alternative approach, || k+1 k||2 may be

monitored.

In order to assess the performance of an algorithm by the numerical experiments, the

error in the conductivity itself can also be monitored; ||ek||2 = || k+1 true||

2 where true

is the known conductivity distribution of the simulated target. Obviously, this method of

monitoring will not be feasible for practical MIT imaging where true is unknown.

Solve forward

problem

Stopping

criteria

Measured

voltages

Start with initial

guess

Compute

sensitivity

Compute

voltages

Solve inverse

problem

stop

Continue

Update

conductivity

Solution verified

Page 158: Magnetic Induction Tomography for Imaging Cerebral Stroke

137 Discrete Inverse Problem

6.8.2. Newton-type methods

The iterative algorithm presented in the previous section, falls in the category of

Newton-type methods for solving the inverse problem. The key idea of any Newton-

type method consists in repeatedly linearizing the forward operator, V = F( ), around

some approximate solution , and then solving the linearized problem. There exist

various formulas in the literature for the Newton-type iterative algorithms, which are

explained by the LS minimization and using Tikhonov regularization1. The only

difference is the penalty term added to the LS solution.

For instance, the penalty term || ||2 which attempts to constrain the conductivity

change between two iterations results in equation (6.37), referred to as Levenburg-

Marquardt [190]. One may consider constraining || k+1||2 (absolute conductivity values)

or || k+1 - ref||2 (difference between the expected and estimated absolute conductivity

values). The latter results in a formula which is referred to as Gauss-Newton [173].

In general, one may allow the regularization parameter to vary between the iterations

( k), or consider it to be constant ( ). The automatic update of between the iterations

is possible if a reliable method for selecting is available.

With the aim of increasing the computational speed of the iterative process, the

sensitivity matrix can be calculated only once, although the forward solution is

calculated repeatedly as the conductivity is updated. Since the sensitivity matrix

remains the same in all iterations (same as that in the first iteration), it can be pre-

calculated off-line. This semi-linear algorithm is also called the Newton-Kantorovich

method [2,191]. Furthermore, if rather than solving the forward problem, the MIT

voltage is simply calculated by multiplying the sensitivity matrix with the most recent

conductivity distribution (i.e. Vk =

A k), then the method is referred to as the iterative

Tikhonov regularization [190]:

(6.38)

However as figure 6.6 indicated, the sensitivity matrix can only predict the voltages if

the conductivity change is small. Therefore, the applicability of semi-linear methods

may be limited to low-contrast targets.

1 Regularization methods other than Tikhonov regularization for Newton-type algorithms can be found in

Kaltenbacher et al [230].

Page 159: Magnetic Induction Tomography for Imaging Cerebral Stroke

138 Discrete Inverse Problem

6.8.3. Absolute imaging using phase changes

As discussed in the forgoing sections for differential imaging, it is also often useful to

normalize the MIT signals to the primary voltage for absolute imaging. This may help

to reduce some of the errors in the data. For instance, if there is an error in the gain of

one of the measurement channels in the hardware, it will affect the secondary and

primary voltages in the same way, and the error will be cancelled when forming =

tan-1

( V/Vp). This can be of great importance when attempting absolute imaging for

which the difference between the measured and simulated data is used.

The equivalent to formulation (6.37) for the phase changes in the generalized Tikhonov

form is:

(6.39)

where sensitivity entries are computed by (6.12). As a case study, absolute imaging of

the cylinder with inclusion, was performed. The simulations were conducted at the

frequency of 10 MHz. The conductivity values of the target were, therefore, 0.16 and

1.1 S m

-1 for the background (white matter) and inclusion (blood), respectively (see

figure 6.23). In order to gain more information, the MK1 array was placed at 3 different

levels as shown in the figure, resulting in 256 3 = 768 MIT signals.

The images were reconstructed from the phase changes, i.e. equation (6.39). It was

found that normalization slightly reduces the ill-posedness of the inverse problem in this

case. The condition number of a normalized sensitivity matrix, computed for a uniform

conductivity, was some 5 times less than that of an un-normalized one.

Figure 6.23. Simulated target; coil array is placed at three levels along height of cylindrical tank.

Colour-bar in S m

-1.

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.2

0.4

0.6

0.8

1

5 10 15 20

2

4

6

8

10

12

14

16 0

0.2

0.4

0.6

0.8

1

5 10 15 20

2

4

6

8

10

12

14

16 0

0.2

0.4

0.6

0.8

1x-y y-z

4 cm

0

-4 cm

Page 160: Magnetic Induction Tomography for Imaging Cerebral Stroke

139 Discrete Inverse Problem

As stated previously, both the GCV and L-curve methods require the solution of the

inverse problem for different levels of regularization in order to select . This can be

done by computing the SVD of A, which must be repeated for each iteration when the

sensitivity matrix is updated. Therefore, the automatic update of between iterations

can be prohibitively long. Furthermore, figure 6.20 indicated that these methods, when

employed with the generalized Tikhonov solution, may result in under- or over-

regularized solutions.

Here, an alternative approach is suggested for selecting k. The iterations can be started

with a relatively large value of 1. This assures the stability and robustness of the

algorithm for the first few iterations. The initial guess can, therefore, be chosen with

less caution. In order to accelerate the convergence and achieve a better spatial

resolution, the regularization level is reduced after each iteration as:

(6.40)

where l 1 is a predefined ratio. Note that k is used to compute k+1 which is the

solution computed at (k+1)-th iteration. The residual error ||rk||2 = || m – k||

2 (radians)

and the solution error ||ek||2 = || k true||

2 (S

m

-1) are monitored between the iterations.

The iteration process can be stopped once an increase in the residual error is observed

(||rk+1||2 > ||rk||

2) and the last updated conductivity may be considered as the converged

solution. This is the solution for which the minimum residual is achieved. Here,

however, the process was continued for a few more iterations after the increase in the

residual. This allowed examining the solutions obtained thereafter.

For the mentioned case study (figure 6.23), the iterations were conducted using R-face,

1 = 1

S

m

-1, and 1

=

1 10

-8. As the first attempt, a 50% drop in the regularization level

between the iterations were considered (l = 0.5). Solving the forward problem for each

iteration, including all 3 planes of measurement, took 75 s, while 6

s was the solution

time for the inverse problem. Plots of the monitored quantities are shown in figure 6.24.

The residual error ||rk||2 fell abruptly during the first few iterations and, after a slow

decreasing trend, started increasing from iteration 53; this occurred at 52 = 4.4 10-24

.

The second row in figure 6.24 shows the magnified plots at iterations 51-53. The

solution error ||ek||2 showed a similar trend to that in the residual.

Page 161: Magnetic Induction Tomography for Imaging Cerebral Stroke

140 Discrete Inverse Problem

Figure 6.24. Absolute imaging using (6.39) with R-face, 0 = 1 S m-1

, l = 0.5 and 1 = 1 10-8

. First

row: plots of ||rk||2 and ||ek||

2 for iteration procedure. Second row: same plots as in first row but

magnified at iteration 52. Also shown reconstructed images at iterations 52, 53, and 54. Upper limit

of colour-bars (S m

-1) is set to maximum conductivity value for each image.

Based on the residual plot, the solution obtained at iteration 52 can be considered as the

converged solution. The reconstructed images obtained from the 52th, 53th, and 54th

0 10 20 30 40 50

-0.05

0

0.05

0.1

0.15

0.2

Iteration No.

0 10 20 30 40 504.5

5

5.5

6

6.5

7

7.5

8

Iteration No.

50 51 52 53 54

0

0.5

1

1.5

2

2.5

3

3.5

x 10-3

Iteration No.

50 51 52 53 54

5.25

5.3

5.35

5.4

5.45

Iteration No.

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

5 10 15 20

2

4

6

8

10

12

14

16 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

5 10 15 20

2

4

6

8

10

12

14

16 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

5 10 15 20

2

4

6

8

10

12

14

16

18

200

1

2

3

4

5

6

7

8

9

10

5 10 15 20

2

4

6

8

10

12

14

160

1

2

3

4

5

6

7

8

9

10

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

5 10 15 20

2

4

6

8

10

12

14

16

18

200

1

2

3

4

5

6

7

8

9

10

52th 53th 54th

y

-z x

-y

||rk||2 ||ek||

2

||ek||2 ||rk||

2

Page 162: Magnetic Induction Tomography for Imaging Cerebral Stroke

141 Discrete Inverse Problem

iterations are also illustrated in the figure where the colour-bar limits are set to the

maximum conductivity value in each image. In order to enhance the visualization, for

each image, the colour-bar limit is set to the maximum value in that particular image.

The conductivity values of the inclusion (blood) are clearly under-estimated in the

converged solution. A conductivity of 0.47 S

m

-1 is obtained for blood in the 52th

iteration (target is 1.1 S m

-1).

If the iterations are continued after the increase in the residual, the process will be

unstable due to the solution being under-regularized. A large ||ek||2 and high amplitude

artefacts can be observed in the plot and reconstructed image at the 54th iteration.

One may start the iteration process with a different level of regularization (in previous

example 1 = 1 10-8

). Obviously, starting with a larger will require the algorithm to be

carried out for a few more iterations to arrive at the converged solution. However, a

small value of can have a significant effect on the solution. In order to examine this,

let us consider 1 = 1.8 10-20

to start the iterations with. This is equivalent to 40 =

1.8 10-20

in the previous example for which the algorithm was stable (see the plots at

iteration 40 in figure 6.24).

Absolute imaging was performed where all the parameters (except 1) were similar to

the previous example. Results are shown in figure 6.25. The algorithm was unstable and

the residual started increasing from iteration 4. The smallest solution error ||ek||2 was

obtained at the first iteration which corresponds to the initial guess 1. Also the solution

errors were considerably larger than those in figure 6.24. The reconstructed images are

dominated by the artefacts which clearly indicates the inverse solution being under-

regularized.

The reason why the algorithm was unstable in this case can be realised from the

algorithm itself. Revisiting the equation (6.39), we can express the inverse problem as x

= B b where x = k+1 - k, b = m - k, and B = (AkTAk + kRk

T Rk)

-1Ak

T. For the first

few iterations, there are large values in vector b. This is due to the difference between

the initial guess ( 0) and target conductivity distribution.

Therefore, for a small value of k, B can amplify such large values in b, resulting in

high amplitude artefacts in the solution. These artefacts produce more artefacts in the

next iterations; evident in images in figure 6.25.

Page 163: Magnetic Induction Tomography for Imaging Cerebral Stroke

142 Discrete Inverse Problem

Figure 6.25. Absolute imaging using (6.39) with R-face, 0 = 1 S m-1

, l = 0.5 and 1 = 1.8 10-20

.

Errors ||rk||2 and ||ek||

2 are plotted against iterations. Reconstructed images correspond to iterations

1, 2, and 8. Colour-bar (S m

-1) limits are set to maximum conductivity for each image.

For the example in figure 6.24, however, the solution at iteration 40 is much closer to

the target than the initial guess (evident in the ||ek||2 plot), particularly for the

background which has a large volume. Therefore, the values in b (contributions from

both the background and inclusion) are smaller (evident in the ||rk||2 plot). This prevents

the large oscillations in the solution until grows extremely small, making the

algorithm unstable at iteration 53.

Therefore, unless a close estimate for 1 is available, it is suggested to initially over-

regularize the inverse problem. The only disadvantage would be the need for a few

more iterations to arrive at the solution.

0 2 4 6 8 1050

100

150

200

250

300

350

400

450

Iteration No.

0 2 4 6 8 100

0.2

0.4

0.6

0.8

1

1.2

1.4

Iteration No.

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

5 10 15 20

2

4

6

8

10

12

14

16 0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

5 10 15 20

2

4

6

8

10

12

14

16

18

200

1

2

3

4

5

6

5 10 15 20

2

4

6

8

10

12

14

160

1

2

3

4

5

6

5 10 15 20

2

4

6

8

10

12

14

160

1

2

3

4

5

6

5 10 15 20

2

4

6

8

10

12

14

16

18

200

1

2

3

4

5

6

5 10 15 20

2

4

6

8

10

12

14

160

1

2

3

4

5

6

7

8

9

10

5 10 15 20

2

4

6

8

10

12

14

160

1

2

3

4

5

6

7

8

9

10

1st 2nd 8th

y

-z x

-y

||ek||2 ||rk||

2

Page 164: Magnetic Induction Tomography for Imaging Cerebral Stroke

143 Discrete Inverse Problem

Figure 6.26. Absolute imaging using (6.39) with R-face, 0 = 1 S m-1

, l = 0.1 and 1 = 1 10-8

. Errors

||rk||2 and ||ek||

2 are plotted against iterations. Reconstructed images correspond to iterations 10, 18,

and 19. Colour-bar (S m

-1) limits are set to maximum conductivity value in each image.

Another parameter which influences the algorithm is the ratio used to reduce the

regularization level between the iterations; l = 0.5 was previously employed. In order to

examine the effect of this ratio on the solution, the results in figure 6.26 were obtained

using l = 0.1, where all other parameters were similar to the example in figure 6.24.

As can be seen, the residual increased from iteration 19, corresponding to 18 = 1 10-25

,

which, not surprisingly, is at an earlier stage of the iteration process compared to the

53th iteration in figure 6.24. However, unlike the diverged solution at iteration 53 which

provides a still useful image (see figure 6.24), here the diverged solution at the 19th

iteration is useless. This is due to a quicker drop in the regularization level when l = 0.1

is adopted.

0 5 10 15 20 25

-0.05

0

0.05

0.1

0.15

0.2

Iteration No.

0 5 10 15 20 254.5

5

5.5

6

6.5

7

7.5

8

Iteration No.

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

5 10 15 20

2

4

6

8

10

12

14

160

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

5 10 15 20

2

4

6

8

10

12

14

16 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

5 10 15 20

2

4

6

8

10

12

14

16 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

5 10 15 20

2

4

6

8

10

12

14

16 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

5 10 15 20

2

4

6

8

10

12

14

16

18

200

1

2

3

4

5

6

7

8

9

10

10th 18th 19th

y

-z x

-y

||rk||2 ||ek||

2

Page 165: Magnetic Induction Tomography for Imaging Cerebral Stroke

144 Discrete Inverse Problem

Figure 6.27. Plots of ||rk||2 and ||ek||

2 for absolute imaging using (6.39) with R-face, 0 = 1 S m

-1, l =

0.9 and 1 = 1 10-8

.

If l = 0.9 is used, the iterations reach the maximum allowable number defined in the

algorithm (100 iterations) before arriving at the converged solution (see figure 6.27).

Based on the presented examples, l = 0.5 deemed to be a good compromise. This allows

the algorithm to reach the solution in a more gradual and stable trend than that for l =

0.1. On the other hand, fewer iterations are required than that for l = 0.9.

One may reasonably argue that 53 iterations for l = 0.5 is still a large number making

the process long. It must be noted that in all the mentioned examples, noise-free data

have been used. When using a noisy data, it is likely that the divergence in the residual

will occur at the earlier stages of iteration. In order to demonstrate this, two phase noise

levels of 1 and 10 m were added to the simulated phase changes. Absolute imaging was

performed in a similar way ( 1 = 1 10-8

and l = 0.5).

For 1 m noise, the first increase in the residual ||rk||

2 occurred at iteration 8,

corresponding to 7 = 1.5 10-10

(see figure 6.28). However, since the regularization

level is still high, the divergence of the residual is gradual and the following iterations

of the algorithm do not produce a substantial increase in the residual. As can be seen,

the image reconstructed at the 10th iteration is still useful. Thereafter, as grew

smaller, large increases occurred in the residual. The solution error ||ek||2 showed a

similar trend.

Clearly for this example, it is more difficult to decide where to stop the iterations.

According to the Morozov stopping criterion [168], the residual error cannot be

0 20 40 60 80 100

-0.05

0

0.05

0.1

0.15

0.2

Iteration No.

0 20 40 60 80 1004.5

5

5.5

6

6.5

7

7.5

8

Iteration No.

||rk||2 ||ek||

2

Page 166: Magnetic Induction Tomography for Imaging Cerebral Stroke

145 Discrete Inverse Problem

expected to be smaller than the noise in data, thus stopping the iterations when arriving

at the noise level. Here, the phase noise is 1 m = 1.7 10

-5 radians, and the minimum

residual error at iteration 7 is ||rk||2 = 4 10

-4 radians, which is larger than the noise level.

Therefore, the Morozov stopping criterion cannot be applied.

One possible approach is to continue the iterations for a few more steps after the first

increase in the residual, and then examine the obtained images. In this particular

example, the minimum solution error occurs at the 7th iteration where also the residual

is minimum. However, the image obtained from the 10th iteration shows a closer

conductivity value for the inclusion ( 0.35 S

m

-1), despite being less smooth in the

background (white matter). The solution at the 14th iteration is under-regularized.

Figure 6.28. Absolute imaging using (6.39) with R-face, 0 = 1 S m-1

, l = 0.5 and 1 = 1 10-8

. with 1

m additive phase noise. Errors ||rk||2 and ||ek||

2 are plotted against iterations. Reconstructed images

correspond to iterations 7, 10, and 14. Colour-bar (S m

-1) limits are set to maximum conductivity

value in each image.

0 5 10 15

-0.05

0

0.05

0.1

0.15

0.2

Iteration No.

0 5 10 154.5

5

5.5

6

6.5

7

7.5

8

Iteration No.

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.05

0.1

0.15

0.2

0.25

0.3

5 10 15 20

2

4

6

8

10

12

14

160

0.05

0.1

0.15

0.2

0.25

0.3

5 10 15 20

2

4

6

8

10

12

14

160

0.05

0.1

0.15

0.2

0.25

0.3

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.05

0.1

0.15

0.2

0.25

0.3

0.35

5 10 15 20

2

4

6

8

10

12

14

16 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

5 10 15 20

2

4

6

8

10

12

14

16 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

5 10 15 20

2

4

6

8

10

12

14

16 0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

5 10 15 20

2

4

6

8

10

12

14

16 0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

7th 10th 14th

y

-z x

-y

||rk||2 ||ek||

2

Page 167: Magnetic Induction Tomography for Imaging Cerebral Stroke

146 Discrete Inverse Problem

Figure 6.29. Absolute imaging using (6.39) with R-face, 0 = 1 S m-1

, l = 0.5 and 1 = 1 10-8

. with 10

m additive noise. Errors ||rk||2 and ||ek||

2 are plotted against iterations. Reconstructed images

correspond to iterations 3, 6, and 7. Colour-bar (S m

-1) limits are set to maximum conductivity

value in each image.

For the phase noise level of 10 m , the increase in the residual ||rk||

2 occurred at even an

earlier stage at iteration 5, corresponding to 4 = 1.2 10-9

(see figure 6.29). None of the

reconstructed images showed a clear inclusion within the background, and the artefacts

are dominant in all images. Notice the solution error ||ek||2 is larger than that obtained

from the noise-free data in figure 6.24. This level of noise was simply too high for the

image reconstruction algorithm to succeed.

As stated in Appendix A, a phase noise level of 1 m is achievable in the recently

developed MK2a system in the South-Wales MIT group. Absolute imaging using the

measured data by the MK2a system is presented in chapter 9.

0 5 10 15

-0.05

0

0.05

0.1

0.15

0.2

Iteration No.

0 5 10 154.5

5

5.5

6

6.5

7

7.5

8

Iteration No.

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.05

0.1

0.15

0.2

0.25

0.3

0.35

5 10 15 20

2

4

6

8

10

12

14

16 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

5 10 15 20

2

4

6

8

10

12

14

16 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

5 10 15 20

2

4

6

8

10

12

14

16 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

5 10 15 20

2

4

6

8

10

12

14

16 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

5 10 15 20

2

4

6

8

10

12

14

16 0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

5 10 15 20

2

4

6

8

10

12

14

16 0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

3rd 6th 7th

y

-z x

-y

||ek||2 ||rk||

2

Page 168: Magnetic Induction Tomography for Imaging Cerebral Stroke

147 Discrete Inverse Problem

6.8.4. Three-layer geometry

In order to make the geometry more realistic, a layer of peripheral CSF, 1 cm thick, was

added to the geometry. As the a-priori information, conductivity values below zero and

above 2 S m

-1 were not admissible; CSF is the most conductive constituent in the head.

Absolute images were reconstructed using the iteration procedure explained in the

previous section, and the noise-free simulated data.

As can be observed in figure 6.30, although the CSF layer is reconstructed ( 1.6 S m

-1),

the blood (inclusion) is entirely invisible. The likely reason for this is the fact that the

presence of a highly conductive layer on the surface and close to the coils, produces

very large signals. This, in turn, makes it more difficult to visualize internal features

with lower conductivities. The solution started diverging after 3 iterations, despite

initially over-regularizing the inverse problem. Varying different parameters in the

algorithm (e.g. 1 and l) did not improve the images.

Although the meningeal CSF has a relatively smaller thickness in the human head, the

CSF filled ventricles, located within each hemisphere, may also have the same

dominating effect on the signals. Therefore, absolute imaging of haemorrhage may be

difficult using the existing inverse algorithms.

Figure 6.30. Absolute imaging of a 3-layer model including CSF, white matter, and blood. Colour-

bar in S m

-1.

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

5 10 15 20

2

4

6

8

10

12

14

160

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

5 10 15 20

2

4

6

8

10

12

14

160

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

5 10 15 20

2

4

6

8

10

12

14

160

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

y-z

x-y

Target Image

Page 169: Magnetic Induction Tomography for Imaging Cerebral Stroke

148 Discrete Inverse Problem

6.8.5. Edge-preserving

The quadratic penalty term ||Rx||2 in Tikhonov regularization, when R is a neighbouring

matrix, penalizes the discontinuities resulting in smooth edges and blurred images. In

some applications, the distribution of the electrical properties is approximately

piecewise constant; e.g. in medical imaging where each tissue type has its own electrical

properties. Therefore, the use of algorithms which can preserve the edges in the target is

desirable.

One approach to reconstruct the edges is to substitute the neighbouring matrix in

Tikhonov regularization by a modified difference operator which allows the presence of

the discontinuities:

(6.41)

where (R )i evaluates the conductivity gradient between the element of interest and its

neighbours, n is the total number of elements, and is an auxiliary function that is used

to mark the edges, and then encourage their reconstruction [192].

The process starts with using one of the difference operator types for the first iteration.

Once the new conductivity distribution is computed, the gradient between each two

neighbouring elements is examined. Based on this gradient, the regularization matrix for

the next iteration is adjusted in such a way that its smoothing effect is reduced for the

large gradients; the larger the gradient, the less the penalization. This is based on the

assumption that sharp gradients are genuine edges in the target and, therefore, should

not be smoothed out. After a few iterations, the edges are gradually introduced. The

edge-preserving regularization has been employed for MIT imaging by Casanova et al

[193]. Details on modifying the regularization matrix, and various types of the auxiliary

functions can be found in [193,192,194,195].

The edge-preserving technique was examined for the model in figure 6.23, and the

results were compared with those obtained by the R-face. For noise-free simulated data,

there was a visible improvement in the reconstructed image, both in terms of the

sharpness of the inclusion edges and the conductivity magnitude. This is evident in the

surface plots of the x-y cross-sectional plane in figure 6.31, upper row.

Page 170: Magnetic Induction Tomography for Imaging Cerebral Stroke

149 Discrete Inverse Problem

Figure 6.31. Surface plots of x-y plane in images for noise free (upper row) and 5 m noise (lower

row). Left column is for images reconstructed using R-face, and right column is for edge-preserving

regularization matrix.

However, when a noise level of 5 m was added to the simulated phase changes, the

edge-preserving matrix did not show any superiority over the ordinary neighbouring

regularization matrix (see figure 6.31, lower row). This was due to the fact that the

presence of noise in the data can produce artefacts and, consequently, large gradients in

the image. Such gradients are considered as desired edges, and are amplified in

subsequent iterations.

The iteration procedure was stopped after only 4 iterations where the residual error

started diverging. Continuation of the iterations would result in clearer edges for the

inclusion, however, with highly magnified artefacts. Therefore, this method of edge-

preserving can be very sensitive to the noise in the data.

An alternative approach to preserve the edges is the total variation (TV) regularization

function which is also appropriate for problems where discontinuous jumps in the

model are expected. The particular advantage of the TV regularization is that the

regularization term does not penalize discontinuous transitions in the model any more

than smooth transitions.

R-face, noise-free Redge, noise-free

R-face, 5 m noise Redge, 5 m noise

Page 171: Magnetic Induction Tomography for Imaging Cerebral Stroke

150 Discrete Inverse Problem

As described by Aster et al [176], the discretized TV regularised inverse problem can be

formulated as:

(6.42)

where L is a discretization of the gradient operator. The employment of ||L || as the

regularisation penalty term allows the reconstruction of discontinuous profiles. The TV

minimisation belongs to the important classes of problems known as „Minimisation of

sum of norms‟ or „Linear l1 problems‟ [173]. Two good reviews on the TV

regularization are provided by Vogel [180] and Borsic et al [196]. They suggest various

methods for solving (6.42), including the 'lagged diffusivity fixed point' and 'Primal-

Dual' methods, which are outside the scope of this thesis. However, as part of the

LCOMIT project, the TV method was applied for producing 2D images of oil/water in

industrial pipelines.

6.9. Imaging centrally located features

MIT imaging of the centrally located inclusions is examined in this section. This is of

direct relevance for the case of bleeding deep within the brain.

6.9.1. Differential imaging

Time-difference imaging: the model comprised the white matter as the background

(0.16 S m

-1) with and without blood (1.1 S

m

-1) for State-II and State-I, respectively (see

target in figure 6.32). In order to gain more information, similar to absolute imaging, 3

planes of measurement were used (see figure 6.23). A phase noise level of 1 m was

added to the simulated phase changes. The images were reconstructed using the

generalized Tikhonov method (R-face, xref = 0) and (6.35), and using the regularization

levels suggested by the GCV and L-curve methods (see figure 6.32).

Clearly, the L-curve method resulted in an over-regularized solution (middle column).

A better visualization of the inclusion was achieved if the GCV method was used to

select (right column). In both cases, the obtained conductivity change values for the

inclusion were highly under-estimated ( 10-3

S m

-1) where the target value is 0.94 S

m

-1.

Page 172: Magnetic Induction Tomography for Imaging Cerebral Stroke

151 Discrete Inverse Problem

Figure 6.32. Time-difference imaging. Left: target, Middle and Right: reconstructed images using

generalized Tikhonov method (R-face, xref

=

0) and suggested by L-curve and GCV methods,

respectively. Colour-bars in S m

-1.

Figure 6.33. Frequency-difference imaging. Left: target, Middle and Right: reconstructed images

using generalized Tikhonov method (R-face, xref

=

0) and different values. Colour-bars in S

m

-1.

Frequency-difference imaging: the model comprised the white matter and blood

where the conductivity values for 1 and 10 MHz were adopted from table 6.1 (see target

in figure 6.33). Similarly, 3 planes of measurement were employed and 1 m phase

noise was added to the simulated phase changes. The images were reconstructed using

the generalized Tikhonov method (R-face, xref = 0) and (6.35). Both the GCV and L-

curve methods suggested under-regularized solutions ( GCV= 5 10-11

, L-curve= 9 10-11

).

5 10 15 20

2

4

6

8

10

12

14

16

18

20-10

-8

-6

-4

-2

0

2

4

6

x 10-4

5 10 15 20

2

4

6

8

10

12

14

16-10

-8

-6

-4

-2

0

2

4

6

x 10-4

5 10 15 20

2

4

6

8

10

12

14

16-10

-8

-6

-4

-2

0

2

4

6

x 10-4

5 10 15 20

2

4

6

8

10

12

14

16

18

20-1

-0.5

0

0.5

1

1.5

2

2.5

x 10-3

5 10 15 20

2

4

6

8

10

12

14

16-1

-0.5

0

0.5

1

1.5

2

2.5

x 10-3

5 10 15 20

2

4

6

8

10

12

14

16-1

-0.5

0

0.5

1

1.5

2

2.5

x 10-3

5 10 15 20

2

4

6

8

10

12

14

16

18

20-0.1

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

5 10 15 20

2

4

6

8

10

12

14

16-0.1

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

5 10 15 20

2

4

6

8

10

12

14

16-0.1

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.05

0.1

0.15

0.2

0.25

0.3

5 10 15 20

2

4

6

8

10

12

14

160

0.05

0.1

0.15

0.2

0.25

0.3

5 10 15 20

2

4

6

8

10

12

14

160

0.05

0.1

0.15

0.2

0.25

0.3

5 10 15 20

2

4

6

8

10

12

14

16

18

20

0.01

0.02

0.03

0.04

0.05

0.06

5 10 15 20

2

4

6

8

10

12

14

16

0.01

0.02

0.03

0.04

0.05

0.06

5 10 15 20

2

4

6

8

10

12

14

16

18

20

0.01

0.02

0.03

0.04

0.05

0.06

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

5 10 15 20

2

4

6

8

10

12

14

160

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

5 10 15 20

2

4

6

8

10

12

14

160

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

y-z

x-y

Target L-curve = 9 10-11 = 5 10-9

Target L-curve = 9 10-5 GCV = 2 10-6

y-z

x

-y

Page 173: Magnetic Induction Tomography for Imaging Cerebral Stroke

152 Discrete Inverse Problem

The middle column in figure 6.33 shows the image reconstructed using L-curve. Since

GCV was similar to L-curve, the image in the right column was instead reconstructed

using a larger value of , in an attempt to reduce the artefacts in the image. This resulted

in a spread inclusion of under-estimated amplitude ( 0.07 S

m

-1), where the target is 0.3

S m

-1. As can be seen, however, the background conductivity change was reconstructed

more accurately ( 0.05 S m

-1), where the target value is 0.06 S

m

-1.

6.9.2. Absolute imaging

Absolute imaging was performed using the model in figure 6.23, but when the inclusion

was placed in the centre of the cylinder (see target in figure 6.34). The same iterative

algorithm explained in section 6.8.3 was employed for image reconstruction ( 1= 1 10

-8

and l = 0.5). The images were obtained from the noise-free simulated phase changes and

those with 1 m additive noise, and are presented in figure 6.34.

An under-estimated ( 0.24 S

m

-1) and spread inclusion was reconstructed from noise-

free data at the 22nd iteration (again, this is the point at which the residual error started

increasing). The obtained background conductivity ( 0.15 S m

-1) was more accurate,

where the target values is 0.16 S m

-1. Continuing the iterations resulted in an oscillatory

behaviour of the residual error, but a better reconstruction of a the inclusion.

Figure 6.34. Absolute imaging using (6.39) with R-face, 0 = 1 S m-1

, l = 0.5 and 1 = 1 10-8

, using

noise-free data and with 1 m additive noise. Colour-bar (S m

-1) limits are set to maximum

conductivity value in each image.

5 10 15 20

2

4

6

8

10

12

14

16 0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

5 10 15 20

2

4

6

8

10

12

14

16

18

20

0

0.05

0.1

0.15

0.2

5 10 15 20

2

4

6

8

10

12

14

160

0.05

0.1

0.15

0.2

5 10 15 20

2

4

6

8

10

12

14

160

0.05

0.1

0.15

0.2

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.05

0.1

0.15

0.2

0.25

0.3

5 10 15 20

2

4

6

8

10

12

14

160

0.05

0.1

0.15

0.2

0.25

0.3

5 10 15 20

2

4

6

8

10

12

14

160

0.05

0.1

0.15

0.2

0.25

0.3

5 10 15 20

2

4

6

8

10

12

14

16

18

20 0

0.2

0.4

0.6

0.8

1

5 10 15 20

2

4

6

8

10

12

14

16 0

0.2

0.4

0.6

0.8

1

5 10 15 20

2

4

6

8

10

12

14

16 0

0.2

0.4

0.6

0.8

1

Target noise-free , 22nd 1 m noise , 9th

y

-z x

-y

Page 174: Magnetic Induction Tomography for Imaging Cerebral Stroke

153 Discrete Inverse Problem

The presence of even a small level of noise (1 m ) made the inclusion invisible (see the

right column in figure 6.34). Despite the fact that the background conductivity (0.16 S

m-1

) is much smaller than the inclusion (1.1 S m

-1), and the lack of a highly conductive

layer such as CSF, the contribution of the inclusion to the induced MIT signals is

relatively small. This is due to its location at the centre and further away from the

exciter-sensor coils, and the fact that the eddy current is zero at the centre, which results

in a zero sensitivity at the very centre. Therefore, such small signals produced by the

inclusion are highly affected by the noise, making it more difficult to reconstruct the

feature. Compare the image obtained at the 9th iteration here with that at the 10th

iteration in figure 6.28 (both are affected by the same noise level and are reconstructed

using the same level of regularization). The peripheral inclusion was clearly visible.

6.10. Other iterative methods

A drawback of Newton-type methods is the need for the application of the discretized

inverse of F( ), in order to build the search direction for minimizing the objective

functions. In general, when dealing with large 3D problems, the SVD-based pseudo-

inverse and Tikhonov regularization solutions can become cumbersome and, in some

cases, impractical.

Alternatively, some iterative methods, which have been introduced to solve linear

algebraic systems, can also be used to regularize the solution of linear ill-posed

problems. The basic feature is that the number of iterations plays the role of a

regularization parameter [197]. These iterative methods do not require the inversion of

sensitivity matrix to find the search direction. Although more iterations may be

required, the cost of each iteration is usually cheaper.

From these iterative algorithms, a few have been employed for electrical tomography

techniques (MIT, EIT, ECT) and for both absolute and differential imaging. These

methods include Steepest descent, Landweber, Conjugate gradient, Algebraic

reconstruction technique (ART), and Simultaneous iterative reconstruction technique

(SIRT). These algorithms are outlined in Appendix C. Among these, the Landweber

method was briefly examined for time-frequency imaging of haemorrhage [60], where a

detailed investigation was not pursued due to the time restrictions.

Page 175: Magnetic Induction Tomography for Imaging Cerebral Stroke

154 Discrete Inverse Problem

6.11. Inverse crime

As introduced in this chapter, the concept of the inverse problem is to find the

constituents of a material from a set of observations. As was done in this chapter, a

relatively common practice to examine the inverse solvers‟ efficiency is to predict the

observations (e.g. voltages in MIT) through numerical simulations, and then solve the

inverse problem. Since the practical observations are subject to measurement errors,

there will be differences between the predicted and measured data. The errors in

measured data can have several sources:

SNR: noise in the electronics is always present in the MIT signals. This

type of error is measured in terms of SNR, introduced in section 2.3.2.

Geometrical error in the sensor array: any inaccuracies in the geometry of

the MIT sensor array can influence the signals. These include the location

and orientation of the coils, and also the shape and dimensions of the

electromagnetic screen. An accurate model must simulate the exact

geometrical design.

Boundary errors: a common strategy in electrical tomography techniques is

to assume the boundary of the eddy current region as the a-priori, and solve

the forward problem for this specified region. In the case of head imaging,

this would be the skin surface. In practice, however, this a-priori needs to

be defined by some other measuring means, such as optical techniques.

Therefore, error is inevitable. Furthermore, once the boundaries of the

eddy-current region is known, it needs to be registered to the MIT sensor

array; i.e. defining its location within the array. This, in turn, may introduce

additional errors.

The employment of the same solver for predicting the observations and estimating the

material distributions, with no noise added to the predicted data, is known as the inverse

crime. Lionheart [2] and Lionheart et al [173] have suggested several guidelines in

order to avoid the inverse crime. These guidelines and the strategies employed in this

study are as follows:

In the absence of measured data, different models should be employed as

the predictor and estimator. If two models are not available, two different

Page 176: Magnetic Induction Tomography for Imaging Cerebral Stroke

155 Discrete Inverse Problem

discretizations must be used. This way, when solving the inverse problem,

the boundaries of the inclusions will not be forced to be of that in the

forward problem. The discretization for the estimator should not be just a

subdivision of the mesh in the predictor; e.g. making each cubic voxel into

eight smaller ones. In the next chapter, when using the realistic head model,

the FE and FD models are used as the predictor and estimator, respectively.

In chapter 8, where the imaging performance of different MIT arrays is

examined, the FD solver is employed for all simulations due to its faster

performance. However, different discretizations are used.

Noise in the practical data-collection systems must be added to the

predicted data. For the simulations of our practical MIT data-collection

systems, realistic levels of noise for the MK1 and MK2a are used. When

using random noise generators, multiple runs are examined.

Simulated errors in the boundary of the eddy-current region are added to

the predicted data in chapter 8. These include displacement and scaling

errors.

It is also suggested to use objective regularization parameters, rather than

subjective ones, i.e. using an automatically selected . In this thesis,

automatic and empirically selected parameters are both examined.

6.12. Discussion

Differential and absolute MIT imaging were introduced in this chapter. First, the general

framework of the discretized MIT inverse problem and the difficulties associated with

solving it was reviewed. Next, one of the most commonly used image reconstruction

techniques, the Tikhonov regularized solution, was applied to imaging geometrically

simplified models. Time- and frequency-differential images of peripherally located

inclusions were reconstructed, where the effects of noise and the regularization matrix

on the inverse solution were examined. The concept of error estimation and the

automatic selection of the regularization level was introduced. The application of two

automatic methods (GCV and L-curve) for selecting revealed that while these

methods can suggest a proper level of the regularization in some cases, they may result

in under- or over-regularized inverse solutions for other cases.

Page 177: Magnetic Induction Tomography for Imaging Cerebral Stroke

156 Discrete Inverse Problem

In theoretical studies, when the simulated target is known, one can empirically adjust

the regularization level to obtain the images that are most similar to the target; as was

done for some examples in this chapter. However, in practice the target is unknown, and

no single optimised can be selected empirically.

As stated previously, the automatic methods require the solution of the inverse problem

for various levels of regularization to provide one value for . Therefore, it is suggested

rather than the automatic dismissal of different levels of the regularization and using

only one value of , a visual inspection of the reconstructed images for the examined

range of values may be useful. Note that since the inverse problem solutions are

already available for different values of , the only extra step would be the graphical

representation of the images.

This approach might be helpful for imaging cerebral haemorrhage, where one may be

able to tell the difference between a genuine haemorrhage and an artefact produced due

to an under-regularized solution, even if the target is unknown. Examples of such cases

are provided in the following chapters. This visual inspection of the images can be used

until a reliable automatic method, applicable to the various cases, is available. Note that

the development of such an automatic method was not the purpose of this study.

For absolute imaging, even if reliable, the automatic methods for selecting can be

computationally expensive when the sensitivity matrix is updated between the

iterations. On the other hand, a visual inspection of the images after each iteration is

simply impractical. Therefore, an alternative approach was suggested in section 6.8.3,

where the inverse problem is initially over-regularized and is gradually decreased as

the iterations proceed. This approach will be applied to absolute imaging using the

measured data in chapter 9.

When the inclusion was centrally located, both differential and absolute imaging proved

to be more difficult. This is due to the fact that the MIT technique is inherently less

sensitive to the central regions, where the field strength and, consequently, the eddy

currents are smaller. Therefore, imaging a haemorrhage deep within the brain may be

much more difficult. This will be examined in the haemorrhagic simulations for a

realistic head model later in this thesis.

End of chapter 6.

Page 178: Magnetic Induction Tomography for Imaging Cerebral Stroke

Part III

Stroke Imaging

Page 179: Magnetic Induction Tomography for Imaging Cerebral Stroke

Chapter 7.

Simulated Stroke Imaging Using the

MK1 System

7.1. Introduction

In this chapter, realistic simulations for stroke imaging with the MK1 system are

presented. First, the head model is described. Introduction of the stroke domains into the

model is explained, followed by the finite element (FE) and finite difference (FD) mesh

construction. Next, the results of the forward problem are presented. This includes mesh

convergence tests and discussing the MIT signals.

After extracting the predicted MIT signals due to the stroke, time- and frequency-

difference imaging is performed. Images are reconstructed from the simulated data after

adding random noise of the same level existing in the practical data-collection systems.

Finally, the power absorbed in the head due to its exposure to the electromagnetic fields

is investigated.

7.2. Head model

7.2.1. UCL head mesh

A tetrahedral head mesh, originally developed by UCL for EIT, was used as a basis for

the description of the head geometry in this study. It is an anatomically-realistic, multi-

Page 180: Magnetic Induction Tomography for Imaging Cerebral Stroke

159 Simulated Stroke Imaging Using the MK1 System

layer mesh which was developed from segmented MRI data and using I-Deas1

[198,199]. This mesh has been employed for a number of EIT imaging studies (e.g.

[199,200,201,202,203,204]). The UCL mesh comprises 12 tissues types: muscle, skull,

CSF, grey matter, white matter, ventricles, spinal cord, optic nerves, eye balls, nasal

cavity, ear canals, and the olfactory organ.

Figure 7.1 shows the triangular surface elements of the UCL FE mesh, and figure 7.2

illustrates 6 transverse cross-sectional planes cutting the head at different levels along

the z-axis. Each tissue type is defined by a separate domain, shown in different colours.

Since this mesh was developed for EIT studies, which involves much lower frequencies

than MIT, the air-filled cavities within the head (e.g. olfactory organ) were not

discretized. The total volume of the head, excluding the cavities, is 3,684 cm3. There

are 53,336 tetrahedral elements in the head.

Figure 7.1. UCL head mesh, triangular surface elements of skin and internal compartments.

Figure 7.2. Transverse planes showing internal tissues separated by different colours.

1 Integrated design engineering analysis software.

x

y

z

Page 181: Magnetic Induction Tomography for Imaging Cerebral Stroke

160 Simulated Stroke Imaging Using the MK1 System

Figure 7.3. Stroke regions in the right hemisphere of brain; side view represents the vertical level of

the LP which was similar for all three strokes.

7.2.2. Stroke domains

Intra-cerebral haemorrhages were introduced into the head within the right hemisphere

of the brain. This was done by separating some of the existing tetrahedral elements

within the head as the stroke domains. Three lesions of varying sizes and sites were

considered: a large peripheral (LP), a small peripheral (SP), and a small deep (SD) with

volumes of 49.4, 8.2, and 7.7 cm3, respectively (see figure 7.3). These volumes ranged

from 4.9 to 0.8% of the total volume of the brain. These particular examples were

considered in order to assess the sensitivity of the MIT signals to the location and

volume of the lesion.

7.2.3. FE mesh construction

In order to prepare the mesh for the MIT simulations, the missing components such as

the electromagnetic screen and the coils were required to be added to the model.

Furthermore, the air-filled cavities within the head needed to be included in the

discretized domain.

As the first attempt, the geometry of the head was extracted from the UCL mesh. Note

that only a tetrahedral mesh was received from UCL, but not the geometry description.

To this end, the UCL mesh was imported into Comsol and the geometry was extracted

from it. However, the complexity of the geometry prevented its extension to add the

missing components. Comsol could not perform any changes in the geometry, but it

allowed generation of new meshes for the existing geometry.

Page 182: Magnetic Induction Tomography for Imaging Cerebral Stroke

161 Simulated Stroke Imaging Using the MK1 System

Therefore to overcome this problem, an alternative approach was undertaken. It

included addition of the missing components in the form of tetrahedral elements to the

UCL mesh, and then extracting the geometry from the completed mesh.

The FLITE1 mesh generation software was employed to extend the mesh in a five-step

process as shown in figure 7.4:

I. First, the missing components, external to the head, were inserted into a

cylindrical domain with a centrally located ellipsoidal vacuum which was

slightly larger than the head.

II. A tetrahedral mesh was then generated for this domain using Comsol (II).

III. Next, based on the triangular boundary elements on the surfaces of the

ellipsoid and head, a tetrahedral mesh was grown to fill the space between

them using FLITE (III).

IV. In a similar way, the internal cavities were filled with a tetrahedral mesh; as

an example, nasal cavity shown in (IV).

V. Finally, all separate meshes were merged to construct the complete MIT

mesh (V).

The total volume of the head, including its air-filled cavities, was 3,709 cm3.

The completed mesh was then imported into Comsol, and the geometry was extracted

from this mesh. A series of tetrahedral meshes were generated using the extracted

geometry in Comsol for subsequent simulations. Further details on these meshes will be

provided later in this chapter.

Figure 7.5 shows the location of the coils relative to the head and LP stroke within the

brain, where red and blue coils represent the exciter and sensor coils, respectively.

1 FLITE is a fully automatic mesh generation software that is developed by O. Hassan and K. Morgan of

the Civil and Computational Engineering Research Centre, School of Engineering, Swansea University.

For further details see e.g. [235,234].

Page 183: Magnetic Induction Tomography for Imaging Cerebral Stroke

162 Simulated Stroke Imaging Using the MK1 System

Figure 7.4. FE mesh construction; details given in text.

Figure 7.5. Level of coil array relative to head and LP stroke (red = exciter, blue = sensor).

y

z

(II)

(III)

(IV)

(V)

Page 184: Magnetic Induction Tomography for Imaging Cerebral Stroke

163 Simulated Stroke Imaging Using the MK1 System

Figure 7.6. FD mesh with two cuts illustrating the internal structure, including the LP stroke.

7.2.4. FD mesh construction

Since in the FD solver, only the conductive volume (head) is discretized, the UCL head

mesh, with its internal cavities first filled with tetrahedra (figure 7.4IV), was discretized

into cubic voxels for the FD model. Uniform cubic grids of different sizes, ranging from

2 to 7.5 mm were prepared for subsequent simulations. In a grid of 2.5 mm size, for

instance, 243083 cubic voxels are used to represent the head. Figure 7.6 shows the 2.5

mm cubic grid, where the red domain in the right hemisphere is the LP stroke.

7.2.5. Dielectric properties

All dielectric properties were assumed to be isotropic. As explained in section 4.5, the

stroke lesions were assumed to be homogeneous regions with brain tissues and blood as

constituents. The lesion conductivity was set to a weighted average such as:

(7.1)

where 0.25 p 1 shows the proportion of the volume occupied by blood, and can be

thought of as the severity of bleeding. The permittivity was computed in a similar way.

Table 7.1 gives the adopted values for different tissues at frequencies of 1 and 10 MHz,

based on the survey carried out in chapter 4. As an example, for a lesion in white matter

saturated with 75% blood, and at frequency of 10 MHz: lesion = 0.25 0.159 +

0.75 1.097 = 0.863 S m

-1.

x y

z

Page 185: Magnetic Induction Tomography for Imaging Cerebral Stroke

164 Simulated Stroke Imaging Using the MK1 System

Table 7.1. Conductivity (S m

-1) and relative permittivity values assigned to each tissue type at the

frequencies of 1 and 10 MHz. The mass density of tissues (kg m

-3), adopted from [205], are to be

used for the power absorption calculations later in section 7.6. a filled with air.

1 MHz 10 MHz

Mass density Organ/tissue Conductivity

Relative

permittivity Conductivity

Relative

permittivity

Muscle 0.503 1836 0.617 171 1040

Skull 0.057 197 0.0828 54 1850

CSF 2.000 109 2.002 109 1010

Grey matter 0.163 860 0.292 320 1030

White matter 0.102 480 0.159 176 1030

Ventricles as CSF as CSF as CSF as CSF as CSF

Spinal cord 0.133 670 0.225 248 1040

Optic nerves 0.130 926 0.223 155 1040

Eyeballs 1.127 897 1.20 127 1037

Nasal cavitya 0 1 0 1 -

Ear canalsa 0 1 0 1 -

Olfactory organa 0 1 0 1 -

Blood 0.822 3026 1.097 280 -

7.3. Forward problem

The forward problem was solved using both the FE and FD methods. In this section,

first the FE simulations are discussed, followed by the FD results.

7.3.1. FE mesh convergence

The FE model was solved using the techniques explained in chapter 5. In order to

ensure that the simulations with the head were accurate, a convergence test was first

performed. As described in section 7.2.3, a series of increasingly fine meshes of

tetrahedra, comprising 43538 to 776852 elements, were generated using Comsol (see

table 7.2). The models were solved for the normal head (no stroke present) and a

frequency of 10 MHz. In order to represent the normal head, the dielectric values of the

haemorrhage domain was set to those of the brain tissues, i.e. white or grey matter.

Page 186: Magnetic Induction Tomography for Imaging Cerebral Stroke

165 Simulated Stroke Imaging Using the MK1 System

Table 7.2. Series of meshes for the convergence tests; NS: not solved due to memory restrictions.

Mesh Number of elements Degrees of freedom

First-order elements

Degrees of freedom

Second-order elements

1 43,538 51,294 277,470

2 74,812 87,940 476,024

3 82,852 98,114 529,318

4 97,144 114,025 617,530

5 155,253 182,305 987,062

6 183,369 216,540 1,169,898

7 229,379 270,249 1,461,356

8 276,787 326,283 1,764,010

9 306,808 361,534 NS

10 33,488 393,344 NS

11 383,093 453,105 NS

12 471,113 556,270 NS

13 587,941 694,427 NS

14 776,852 918,028 NS

The computations were performed for both first- and second-order elements, and for the

activated Ex1 (for coil numbering see figure A.2 in Appendix A). For models with

degrees of freedom larger than 2000000 (cells marked 'NS' in table 7.2), the memory

required to solve the model exceeded the maximum memory available (8 GB).

For the solved models, the induced voltage in the sensor coils was monitored. As

expected, the second-order elements performed better than the first-order in terms of the

accuracy, and achieved convergence of the voltage at a smaller number of elements (see

figure 7.7). From Mesh-2 onwards, second-order elements showed a steady predicted

sensor voltage. However, for a sufficiently fine mesh (Mesh-7), first-order elements

returned comparable results albeit at greater computational cost. Figure 7.7 shows the

typical convergence results for the computed voltage for three arbitrary sensors.

Monitoring the results at all sensors revealed similar behaviour.

Based on the convergence results, all subsequent calculations employed Mesh-3 (3rd

point from left on each curve for second-order elements), with 82,852 second-order

tetrahedral elements of which 54,019 elements represented the head. This mesh was

deemed to be a good comprise between the accuracy of solution and computational run

Page 187: Magnetic Induction Tomography for Imaging Cerebral Stroke

166 Simulated Stroke Imaging Using the MK1 System

time. Solving the model for one activated exciter for this mesh took approximately 4

hours, and to compute a complete set of data (i.e. 16 exciters) took 64 hours.

A typical result of the magnetic field produced by one exciter coil (Ex13), at frequency

of 10 MHz, is illustrated in figure 7.8a. The corresponding map of the induced eddy-

current density in a transverse slice, for the normal head, reveals a high current density

in the CSF due to its larger conductivity relative to the skull and brain tissue (figure

7.8b). With the LP stroke lesion present, it too contains a higher current density than the

surrounding normal brain tissue because of its blood content and, consequently, higher

conductivity (figure 7.8c).

Figure 7.7. Convergence of the induced voltage, for activated Ex1, in sensors Se6, Se9, and Se13 for

first- and second-order elements plotted in broken red and solid blue lines, respectively.

Figure 7.8. Typical results. (a) Projection on the transverse plane of magnetic field lines with Ex13

activated with a current of 1 A. The colour of a field line indicates the magnitude of the magnetic

field at that point; the colour-bar indicates the field strength in A m

−1. (b) Map of induced current

density (A m

−2) in a transverse slice through the normal brain, 25 mm above the centre of the coils.

(c) as (b) but with the LP stroke present.

105

106

-0.5

0

0.5

1

1.5

2

2.5x 10

-5

Degrees of freedom

Vo

lta

ge

(V

)

13

9

6

Shield

Exciter

Sensor

(a) (b) (c)

Page 188: Magnetic Induction Tomography for Imaging Cerebral Stroke

167 Simulated Stroke Imaging Using the MK1 System

7.3.2. MIT signals (FE results)

Simulations were first carried out for the 256 coil combinations with the conductivity of

all head tissues set to zero, and their relative permittivities set to unity, so that the

simulated problem corresponded to that obtained if no head were present. The primary

voltage Vp was recorded. Simulations were then carried out for the normal head, and the

in-quadrature voltage, V1, was recorded. Similarly, with a lesion present, the in-

quadrature voltage, V2, was recorded. The phase difference, , was calculated for each

coil combination as:

(7.2)

Profiles of for each of the three strokes saturated with 75% blood, and for Ex1

activated at 10 MHz, are plotted in figure 7.9. Note that the values of due to the

presence of three haemorrhages are in the range of milli-degrees which indicates that

sensitive hardware is required to measure them accurately.

The LP stroke produced the largest signals ( max = 65 m in Se11). This is not

surprising as a larger volume of blood with higher conductivity replaced the brain

tissues.

Figure 7.9. Profiles of for three strokes, simulated at frequency of 10 MHz and with Ex1

activated. Also shown is the phase angle for normal head relative to empty space (values are

divided by 20 for convenience of display).

0 2 4 6 8 10 12 14 16-10

0

10

20

30

40

50

60

70

80

Sensor number

(m

)

LP

SP

SD

/20

Page 189: Magnetic Induction Tomography for Imaging Cerebral Stroke

168 Simulated Stroke Imaging Using the MK1 System

The values of for the two smaller strokes (SP and SD), reveal that the more centrally

located the bleeding, the smaller the MIT signals are. This is also to be expected as the

sensitivity of the MIT system decays towards the centre of brain as a result of the

reduced inductive coupling of the lesion with the coils. For comparison, the phase

angles relative to the empty detector volume when the head was first introduced, =

tan-1

(V1/Vp), are also shown. They are very much larger, with a maximum of 1560 m .

In figure 7.10, the complete sets of 256 values of are plotted in magnitude, sorted in

descending order, for each of the three stroke lesions. As expected, the largest signals

were produced by the LP stroke (| |max = 71

m ), and the smallest were from the SD

lesion (| |max = 3.4

m ). Given the phase noise level of the MK1 system (17

m ),

approximately 70 of the 256 signals (27%) would be detectable for the LP stroke, but

none for the smaller strokes1.

In order to be able to measure the same percentage of the signals produced by the SD

stroke as detectable in the LP stroke, a noise level of about 1 m would be required. This

level of the phase noise is achievable in the MK2a system [206]. The tolerable noise

level for producing acceptable images will be examined in sections 7.4 and 7.5.

Figure 7.10. The 256 values of magnitude of sorted in descending order for each of three strokes,

simulated at frequency of 10 MHz.

1 A transmission line matrix (TLM) model was also used as an independent check of the FE simulations.

Detailed discussion can be found in a report by Zolgharni et al [60].

0 50 100 150 200 250

0

10

20

30

40

50

60

70

Coil combination

| |

(m

)

LP

SP

SD

Page 190: Magnetic Induction Tomography for Imaging Cerebral Stroke

169 Simulated Stroke Imaging Using the MK1 System

Figure 7.11. The 256 values of magnitude of sorted in descending order for LP stroke with

different blood percentages.

Finally, the phase changes for the LP stroke, containing different percentages of blood

(25 to 100%), are plotted in figure 7.11. Clearly, as the blood percentage increases, the

phase changes produced grow larger. For only 25% blood in the lesion, | |max = 34

m

is obtained. This indicates that at early stages from the onset of haemorrhage, it would

be more difficult to detect the presence of bleeding. When the blood entirely replaces

the brain tissues, i.e. 100% blood content, then | |max = 94

m . However, in the

remainder of this thesis, the strokes saturated with 75% blood are considered in order to

be compatible with the EIT studies using this head model (e.g. [202]).

7.3.3. FD vs. FE

For the FD model, similar to the FE simulations, the conductivity values were adopted

from table 7.1. Since the FD solver cannot handle holes within the conductive domain,

for the cavities within the head, a small enough conductivity of 10-4

S m

-1 was

considered to represent the air.

The permittivity of all tissues, however, was assumed to be zero. The reason for this

approximation is that it allows a stationary iterative solver (such as SOR explained in

section 5.3.3), to be used for solving the linear system of equations. For the FD

0 50 100 150 200 2500

10

20

30

40

50

60

70

80

90

100

Coil combination

| |

(m

)

100%

75%

50%

25%

Page 191: Magnetic Induction Tomography for Imaging Cerebral Stroke

170 Simulated Stroke Imaging Using the MK1 System

methods, stationary iterative solvers can be implemented without the need to assemble

the large matrix. Clearly, this leads to considerable computational savings.

If the true permittivity values were to be included in the model, this would change the

system matrix from being symmetric positive definite to being complex symmetric. For

such matrices, the convergence of stationary iterative solvers can no longer be

guaranteed. For the permittivity values listed in table 7.1, convergence was not

obtained.

By adopting an alternative iterative solution technique (or a direct approach), linear

systems with complex matrices can be solved, but this would require the assembly and

storage of a large matrix. Considering the large number of cubic voxels used to

represent the head, this would have substantial negative impact on the computational

footprint of the FD algorithm. As the computational saving was one of the main reasons

for adopting the FD algorithm in this thesis, this approach was not pursued here.

In order to examine the accuracy of the FD results, and the effect of ignoring the tissue

permittivity, the FD results can be compared with those obtained from the FE model

that included the true permittivity values (available from the previous section). To this

end, the normal head (no stroke present) was simulated at frequency of 10 MHz, using

different grid sizes, generated in section 7.2.4 (see table 7.3).

Table 7.3. Series of FD meshes for the convergence tests.

Mesh Grid size (mm) Number of conductive voxels

1 7.5 9 012

2 6 17 583

3 5 30 401

4 4 59 338

5 3 150 865

6 2.5 243 083

Note that the MK1 system includes an aluminium screen. Therefore, the A-field was

extracted from Comsol for each one of these grids (screen implementation in FD solver

was explained in section 5.5), and the FD models were solved.

Page 192: Magnetic Induction Tomography for Imaging Cerebral Stroke

171 Simulated Stroke Imaging Using the MK1 System

The induced voltages in 16 sensors for Ex1 activated are plotted in figure 7.12, where

the converged FE results are also plotted for comparison (thick broken line). As can be

seen, the FD results showed a consistent trend. As the grid size reduced, the discrepancy

between the FE and FD results grew smaller. For the grid of 2.5 mm size, the FD results

were comparable to those obtained from the FE model.

The discrepancy between the FE results and FD results (2.5 mm grid size) is plotted in

figure 7.13. The largest differences are observed at sensor coils where the magnitude of

the induced voltage is relatively large (e.g. Se1 and Se9).

Figure 7.12. Secondary voltage in 16 sensors for Ex1 activated, in normal head at frequency of 10

MHz. Numbers 7.5, 6, 5, 4, 3, and 2.5 indicate grid sizes used in FD model. Converged FE results

are plotted in broken line for comparison.

Figure 7.13. Discrepancy between FE and FD (2.5 mm grid size) results in figure 7.12.

0 2 4 6 8 10 12 14 16

-4

-3

-2

-1

0

1

x 10-5

Se

Secondary

voltage (

V)

0 5 10 15

-3

-2

-1

0

1

2

3

4x 10

-7

0 5 10 150

5

10

15

20

0 2 4 6 8 10 12 14 16

-4

-3

-2

-1

0

1

x 10-5

Se

Secondary

voltage (

V)

7.5

6

5

4

3

2.5 FE

VFE - VFD (V)

Sensor number

Sensor number

Page 193: Magnetic Induction Tomography for Imaging Cerebral Stroke

172 Simulated Stroke Imaging Using the MK1 System

The error obtained for other coil combinations showed a similar behaviour, indicating

that there was a good agreement between the FE results and those obtained from the FD

model using a 2.5 mm grid size. This is despite the fact that the permittivity values in

the FD model were set to zero, therefore, neglecting the displacement currents within

the head. The error observed in figure 7.13 can have several reasons such as the

discretization errors.

In order to examine the effect of neglecting displacement currents, while excluding the

other sources of error, the FD model must be solved with and without the true

permittivity values. This would require implementing better iterative solvers in the FD

code. Alternatively, the FE model could be solved were the permittivity values are set to

zero. However, the iterative solvers, available in Comsol, did now allow zero

permittivities. The use of a direct solver was also impractical for such a large model.

The error values shown in figure 7.13 deemed to be in the acceptable range for the

purpose of this chapter which is to use different models as the predictor and estimator,

therefore avoiding the inverse crime (see section 6.11). Solving the FD model for the

grid of 2.5 mm size, activating all 16 exciters, took 65 min, which is considerably

shorter than 64 hours in the FE simulations. In the remainder of this chapter, this grid

will be used for computing the sensitivity matrix in the FD model. As the predicted MIT

measurements, signals computed by the FE model will be employed.

7.4. Time-difference imaging

From the values of (obtained from the FE model), time-difference images of the

conductivity change were reconstructed. Here, the conductivity change corresponded to

that from before (normal head) and after stroke.

First, the sensitivity matrix was computed by solving the FD model for the head with a

uniform conductivity of 1 S

m

−1, activating all 16 exciters and 16 sensors in sequence.

The entries of the sensitivity matrix were computed using (6.12), i.e. normalized to the

primary signal. In solving the inverse problem, it was assumed that the boundary of the

eddy-current region (skin surface) was known as a-priori information. The entire eddy-

current region (head), therefore, remained fixed at 243,083 cubic elements.

Page 194: Magnetic Induction Tomography for Imaging Cerebral Stroke

173 Simulated Stroke Imaging Using the MK1 System

Since the sensitivity was an excessively large matrix (256 243083), the Tikhonov

regularized solution in (6.25) was used for reconstructing the images, where the

regularization matrix was Identity matrix. The images were reconstructed from the

noise-free and noisy phase changes where two levels of Gaussian phase noise, 1 and 17

m (standard deviation), corresponding to the noise in MK1 and MK2a systems, were

added to the simulated data. Three cross-sections were extracted from the reconstructed

volume image for display. For each image, the normalized solution error was computed

as:

(7.3)

where rec and true are the reconstructed and true conductivity change distributions,

and ||.||2 is the Euclidean norm. Note that the target ( true) and image ( rec) have

different discretizations, simulated by the FE and FD models, respectively. For

computing E, the target was interpolated onto 2.5 mm voxels, similar to that in the

image, as if it had been simulated using this grid.

GCV method: first, the regularization parameter was selected by using the GCV

method (explained in section 6.6.1). Plots of the GCV function for different values of

are shown in figure 7.14, while the images presented in figure 7.15 were reconstructed

using the values of λ which minimized the GCV function. Note that, for convenience of

display, the sagittal plane (third row) is not to scale.

For the noise-free data, the solution is clearly under-regularized, where the magnitude of

the obtained conductivity changes are as large as 40 S m

-1 (2nd column from left). These

correspond to negative and positive values appearing as small volume artefacts evident,

mainly, on the boundary of the transverse plane. A large error of E = 4.4 103% was

obtained for this image.

Also the GCV plot shows that the GCV function is constantly decreasing as grows

smaller until 10-22

. Therefore, for any range of examined values of , the minimum

value within the range would always be considered as the optimised . However, an

inflection point in the curve, suggestive of the development of a minimum shown by the

arrow, occurs at = 1 10-11

. As will be shown in later images, this is where the

optimum value of lies.

Page 195: Magnetic Induction Tomography for Imaging Cerebral Stroke

174 Simulated Stroke Imaging Using the MK1 System

Figure 7.14. Regularization parameter selection using GCV method for three phase noise levels (0,

1 m , 17 m ) added to phase changes for LP stroke.

Target - LP

0

= 7.3 10-22

E = 4415%

1 m

= 3.3 10-11

E = 93.7%

17 m

= 4.0 10-8

E = 96.7%

Figure 7.15. Columns left to right: simulated LP target, and reconstructed images with 0, 1 m , 17

m additive noise. Rows from top to bottom: transverse, coronal, and sagittal cross-section planes.

Colour-bars in S m

-1. is selected by GCV method.

10-30

10-25

10-20

10-15

10-10

10-5

10-16

10-14

10-12

10-10

10-8

G(

)

4.02e-8

3.29e-11

7.28e-22

17 m

1 m

0

1 10-11

Page 196: Magnetic Induction Tomography for Imaging Cerebral Stroke

175 Simulated Stroke Imaging Using the MK1 System

For 1 m of phase noise, a clear minimum in the GCV function can be observed at =

3.3 10-11

. A visible stroke feature of under-estimated magnitude ( 0.15 S

m

-1) is

reconstructed in the image (3rd column). However, its visibility is degraded slightly by

the presence of small volume artefacts of higher amplitude ( 3 S

m

-1). The visibility of

the stroke can be improved by lowering the upper limit in the colour-bar, which in the

current image is set to the maximum value obtained in the inverse solution. An error of

E = 93.7% was obtained for this image.

With 17 m noise in the simulated measurements, the stroke is feature is entirely

displaced towards the head surface; evident in the transverse and coronal planes of the

image in 4th column. This is due a high level of regularization for = 4 10-8

that

minimizes the GCV function (see GCV plot). A smaller value of would improve the

localisation of the stroke, but artefacts would have degraded the image. The computed

error (E = 96.7%) is slightly larger than that in the image with 1 m noise.

L-curve method: next, the L-curve method (as described in section 6.6.2) was used for

selecting , and the results are shown in figure 7.16 and figure 7.17. Apart from the

noise-free case, the inverse solutions are comparable with those suggested by the GCV

method, but slightly more regularized.

The optimised regularization parameter for the noise-free data ( = 3.3 10-12

) resulted

in a stroke feature with an under-estimated magnitude ( 0.15 S

m

-1) in the image (2nd

column from left). The smallest error is obtained for the noise-free image (E = 93.2%)

which is considerably smaller than that obtained based on the GCV method (E =

4.4 103%).

For the noisy data, slightly over-regularized solutions were obtained for the values of

suggested by the L-curve method (1.6 10-9

for 1 m and 1.2 10

-7 for 17 m ). The stroke

feature is blurred and spread in the vertical direction, evident in the coronal planes of

the images (3rd and 4th columns). It is also displaced towards the surface. However, the

asymmetry in the images due to the stroke is clearly noticeable.

Page 197: Magnetic Induction Tomography for Imaging Cerebral Stroke

176 Simulated Stroke Imaging Using the MK1 System

Figure 7.16. Regularization parameter selection using L-curve method for three phase noise levels

(0, 1 m , 17 m ) added to phase changes for LP stroke.

Target - LP

0

= 3.3 10-12

E = 93.2%

1 m

= 1.6 10-9

E = 95.3%

17 m

= 1.2 10-7

E = 97.2%

Figure 7.17. As figure 7.15, but selected by L-curve method.

10-6

10-4

10-2

100

102

104

106

residual norm || A x - b ||2

solu

tion n

orm

|| x || 2

3.9659e-026

1.26e-71.62e-9

3.32e-12

17 m

1 m

0

Page 198: Magnetic Induction Tomography for Imaging Cerebral Stroke

177 Simulated Stroke Imaging Using the MK1 System

= 1.0 10-13 = 1.0 10-12 = 1.0 10-11 = 1.0 10-10

= 1.0 10-9 = 1.0 10-8 = 1.0 10-7 = 1.0 10-6

Figure 7.18. Reconstructed images of LP stroke with 1 m phase noise added to phase changes, and

for different values of . Upper limit in each colour-bar (S m

-1) is set to maximum value in the

corresponding image.

Now let us adjust the regularization level empirically, assuming that the target is

unknown. Different values of , ranging from 1 10-13

to 1 10-6

, were considered and

the images were reconstructed from the phase changes with 1 m noise (see figure 7.18).

These are the typical images that a radiologist might obtain by varying .

Page 199: Magnetic Induction Tomography for Imaging Cerebral Stroke

178 Simulated Stroke Imaging Using the MK1 System

For 1 10-12

, the reconstructed images may be dismissed given the magnitude of the

obtained conductivity change (> 2 S m

-1) and their appearance within the images. Note

that while the background (head volume) appears smooth, there are high amplitude

values in the solution which indicates the presence of small volume artefacts, spread

throughout the volume. A few high amplitude voxels can just be seen in the sagittal

plane at the back of head for = 1 10-13

. Therefore, one may decide either the data is

too noisy, or the solution is highly under-regularized.

On the other hand, if the assessment starts with values of 1 10-7

, an evident

asymmetry can be observed in the images, arousing suspicions of some abnormality in

the head. However, as will be shown in the next chapter, this asymmetry could also be

due to some errors in the assumed boundaries for the eddy-current region (head). An

increase in would result in disappearance of the asymmetry, and reconstruction of

smoother images.

Therefore, one may appreciate that the desired level of the regularization should lie in

the range 1 10-11

1 10-8

, for which the best visibility of the stroke is achieved,

while no significant noise artefacts are introduced.

The best visualization of the stroke for 1 m noise in the data is achieved at = 1 10

-10

in figure 7.18. This image is reproduced in figure 7.19 along with those for other noise

levels (0 and 17 m ), all of which were obtained by the empirical adjustment of .

Note that these images are obtained using the values of optimised based on the visual

assessment of the images. Such images may not necessarily result in the smallest values

of E for the definition of the error given by (7.3), and the minimum error may be

achieved using a different value of .

In chapter 8, while comparing the performance of different configurations for the MIT

sensor arrays by reconstructing the images for each array, the regularization level is

optimised based the computed error. In this way, the best feasible image, from an

strictly mathematical point of view, is obtained when the minimum error is achieved for

each MIT system. Obviously, this way of adjusting requires the targets to be known.

Page 200: Magnetic Induction Tomography for Imaging Cerebral Stroke

179 Simulated Stroke Imaging Using the MK1 System

Target - LP

0

= 1.0 10-11

E = 91.6%

1 m

= 1.0 10-10

E = 92.6%

17 m

= 1.0 10-8

E = 96.7%

Figure 7.19. As figure 7.15, but selected empirically.

Similarly, the images were reconstructed for the smaller strokes, and are shown in

figure 7.20 (SP stroke) and figure 7.21 (SD stroke), where the regularization parameter

was selected empirically. As can be seen, the magnitude of computed conductivity

change for the lesions in all images is considerably under-estimated ( 0.06 S m

-1).

For the SP stroke with 1 m added noise, the stroke is visible, although spread and

blurred. In case of 17 m noise, the stroke is barely visible as an asymmetric feature in

the coronal plane, displaced towards the boundaries.

For the SD stroke, apart from the unrealistic noise-free case where a spread and under-

estimated feature is reconstructed, the stroke is not visible. The likely reason is the fact

that when regularizing the solution to diminish the effect of noise, the genuine phase

changes due to the stroke are also dismissed. This is in line with the observations in

section 6.9, which indicated that MIT imaging of the centrally located features is more

difficult than the peripheral ones.

Page 201: Magnetic Induction Tomography for Imaging Cerebral Stroke

180 Simulated Stroke Imaging Using the MK1 System

Target - SP

0

= 5.0 10-12

E = 97.8%

1 m

= 1.0 10-9

E = 99.2%

17 m

= 1.0 10-7

E = 99.9%

Figure 7.20. As figure 7.19, but for SP stroke.

Target - SD

0

= 1.0 10-10

E = 99.5%

1 m

= 1.0 10-8

E = 100%

17 m

= 1.0 10-7

E = 100.5%

Figure 7.21. As figure 7.19, but for SD stroke.

Page 202: Magnetic Induction Tomography for Imaging Cerebral Stroke

181 Simulated Stroke Imaging Using the MK1 System

Figure 7.22. (a): profiles of noise-free for SD stroke (solid blue line) and those with 1 m added

phase noise (broken red line). (b): difference between two plots in (a).

As an example, figure 7.22a shows plots of the phase differences, noise-free and with 1

m noise, for the SD stoke. Notice that the noisy data deviates significantly from the

noise-free one. Figure 7.22b shows the difference between the two data sets, i.e. noise,

on a similar y-axis scale. As can be seen, the noise is comparable to the genuine phase

diferences induced by the stroke. The variance of the noise-free data and the noise was

1.094 and 1.030 m , respectively.

0 50 100 150 200 250-3

-2

-1

0

1

2

3

4

5

6

Coil combination

(m

)

0 50 100 150 200 250-3

-2

-1

0

1

2

3

4

5

6

Coil combination

no

ise

-fre

e

- n

ois

y

(m)

(a)

(b)

Page 203: Magnetic Induction Tomography for Imaging Cerebral Stroke

182 Simulated Stroke Imaging Using the MK1 System

7.5. Frequency-difference imaging

Frequency-difference imaging was performed using the approach introduced in section

6.7. To this end, all simulations mentioned in section 7.3.2 were repeated, but now at a

frequency of 1 MHz. Note that the convergence test presented in section 7.3.1 was

carried out at frequency of 10 MHz. Here, it was assumed that the convergence results

will also apply at the lower frequencies.

Frequency-difference images were reconstructed from the phase changes computed by

the FE model and corrected using (6.35). The sensitivity matrix was the same as the one

used for time-difference imaging: computed by the FD model at frequency of 10 MHz

using a grid of 2.5 mm size and uniform conductivity of 1

S

m

-1. Its entries were

normalized to the primary signal as in (6.12).

The Tikhonov regularized solution, expressed by the equation (6.25), was used to solve

the inverse problem, and to reconstruct the conductivity changes between 1 and 10

MHz. Any negative values were truncated from the solution as a-priori information

(this physical constraint was discussed in section 6.7).

Figure 7.23 shows the GCV and L-curve plots for optimising for the noise-free data

corresponding to the LP stroke, and figure 7.24 shows the reconstructed images using

these suggested levels of regularization.

Figure 7.23. Regularization parameter selection using GCV and L-curve method for noise-free

frequency-difference data for LP stroke.

10-30

10-20

10-10

10-20

10-15

10-10

G(

)

10-15

10-10

10-5

100

101.2

101.3

101.4

101.5

101.6

residual norm || A x - b ||2

solu

tion n

orm

|| x || 2

GCV L-curve

= 2.88 10-21

= 4.1 10-10

Page 204: Magnetic Induction Tomography for Imaging Cerebral Stroke

183 Simulated Stroke Imaging Using the MK1 System

Target - LP

GCV

= 2.9 10-21

E = 84.5%

L-curve

= 4.1 10-10

E = 70.9%

Figure 7.24. Columns left to right: simulated LP target, and reconstructed images from noise-free

data with selected using GCV and L-curve methods. Rows from top to bottom: transverse,

coronal, and sagittal cross-section planes. Colour-bars in S m

-1.

While, similar to time-difference imaging, GCV suggested an extremely small value of

= 2.9 10-21

for the noise-free data, the L-curve method returned a reasonable value of

= 4.1 10-10

. However, as can be seen in the reconstructed images, the stroke feature is

not visible in either case. A slight left-right asymmetry can be barely observed in the

transverse and coronal planes of the images corresponding to the L-curve method (3rd

column from left).

Even the empirical adjustment of the regularization parameter did not improve the

images. Furthermore, other attempts for imaging the smaller strokes, or reconstruction

from the noisy data, also failed.

Note that in frequency-difference imaging, the change in the conductivity of blood in

the lesion needs to be visualised against changes in all the other tissues (see target in

figure 7.24). Therefore, the non-zero values in the reconstructed images are not

necessarily artefacts, but can be genuine conductivity changes in other tissues. Such

non-zero values can be seen in the coronal and sagittal planes of the image in 3rd

column from left.

Page 205: Magnetic Induction Tomography for Imaging Cerebral Stroke

184 Simulated Stroke Imaging Using the MK1 System

However, since the sensitivity is not uniform throughout the head, the conductivity

change is not reconstructed equally well in all tissues. For instance, the central region of

the head in the sagittal plane seems to have no conductivity change ( 0 S m

-1), whereas

this region is filled by frequency-dependent tissues (see target).

In this chapter, MIT measurements were simulated with the coil array in just one

position relative to the head (see figure 7.5). As will be shown in chapter 8, by

conducting the measurements (simulations) for multiple planes and therefore acquiring

more MIT data, meaningful frequency-difference images of the stroke can be

reconstructed. Since, in this chapter, the FE model was employed for predicting the

MIT signals, simulation of more planes of measurement was prohibitively long.

7.6. Specific absorption rate

For reasons of safety, it is important that the power absorbed in the tissue from the

electromagnetic field should be within international safety limits. The absorbed power

can lead to a temperature increase in the tissues due to Joule heating. The Joule heating

is particularly important when dealing with the brain tissues. An increase in temperature

by even 1 C can have profound effects on a single neuron and neuronal network

function [207].

Specific absorption rate (SAR) is a measure of the rate at which the energy is absorbed

by the tissue when, for instance, exposed to a radio frequency (RF) electromagnetic

field. At a given point in the tissue, SAR is defined as the power absorbed per unit mass

of tissue and has units of W kg

-1 as:

(7.4)

where m(x) is the local mass density of the tissue (table 7.1) and |E(x)| is the magnitude

of the electric field at this point [208,209]. While SAR is commonly used to measure

power absorbed from mobile phones (e.g. [210]) and during MRI scans (e.g. [211]),

tissue heating is generally considered to be negligible in MIT. However, no values,

either measured or modelled, appeared to have been published prior to commencing this

study. Here, the FE model was used to compute the SAR throughout the head.

Page 206: Magnetic Induction Tomography for Imaging Cerebral Stroke

185 Simulated Stroke Imaging Using the MK1 System

Figure 7.25. SAR distribution in normal head for Ex1 activated at 10 MHz; (a) skin surface, (b)

slice at through centres of coils. Colour-bar indicates logarithm to the base 10 of value in W kg

-1.

Eyes are potentially the most critical organs (tissues) regarding the heat generation here.

The reason for this is twofold:

due to the low perfusion in the eye, the temperature may increase even with

relatively low SAR levels [212].

as an external organ, and due to its vicinity to the exciter coils, eye is located in

the areas of high current density where the fields are relatively large, resulting in

higher SAR levels (see figure 7.25).

Therefore, in terms of the heat generation in the eyes, the worst conditions may happen

when the exciter coils 1-3 and 15-16 in the MK1 systems are activated; these are the

closest coils to the eyes (for coil numbering in the MK1 system see Appendix A.1).

Average SAR in one eye was maximum with Ex1 activated, and was 4.1 10-5

W kg

-1.

The computed SAR with Ex1 activated at frequency of 10 MHz was maximum near the

eyes, with a value of 1 10-3

W kg

-1 (see figure 7.25). The overall maximum (not

shown in the figure) was slightly larger, at 3 10-3

W kg

-1, and occurred in the skin and

facial muscle inferior to the eye when Ex3 or Ex15 were activated. This computation

was for an exciter coil of 1 turn carrying a current of 1 A.

Before proceeding to an assessment of the SAR values encountered in practice, let us

consider the effect of changing the excitation current or the number of turns on the

(a) (b)

Page 207: Magnetic Induction Tomography for Imaging Cerebral Stroke

186 Simulated Stroke Imaging Using the MK1 System

excitation coil, the product of which can be taken as a single variable, the Ampere-turns.

If the Ampere-turns is changed by a factor, , the value of the source current density Js

in the boundary value problem (5.6) also changes by the factor . It follows from the

linearity of equations (5.6) and (5.7) that the vector potential, A, and the signal, V,

measured in the sensor coil will all then change by the same factor .

This will be true without the presence of the conductive domain, C, (see figure 2.1),

and in this case the measured V will be the primary signal, Vp. It will also be true for the

in-quadrature part of V (V1 or V2) measured in the presence of C. Hence, as Vp, V1 and

V2 all change by the same factor , a change in the Ampere-turns has no effect on the

values of produced by a given stroke lesion which is computed by (7.2).

The phase noise in the measurement will, however, be affected by the value of . If the

phase noise arises from a fixed level of voltage noise in the receiver amplifier, as will

normally be so, changing the magnitudes of the signals by a factor , will cause an

inversely proportional change in phase noise. Thus, changing the Ampere-turns of the

exciter coil will cause the phase noise to change by a factor 1/ . Finally, the SAR at any

point in C also depends on the Ampere-turns of the exciter coil. Being proportional to

the square of the electric field, as identified by the equation (7.4), the SAR is

proportional to 2.

For the MK1 system (coils of 2 turns, current of 0.14 A as described in Appendix A.1),

= 2 0.14 = 0.28 relative to the numerical model. Multiplying the modelled values of

SAR by 2 = 0.0784 leads to a maximum SAR of 2.4 10

-4 W

kg

-1. The IEC

requirements for MRI scanners, which cover similar frequencies to those used here,

place a limit for the head of 10 W kg

-1, averaged over 6 min [213]. The estimated SAR

for the MK1 system is thus many orders of magnitude below this limit and will indeed

produce negligible heating.

However, in order to reduce the phase noise and the integrating time in the data-

collection systems, new MIT designs are now tending to plan for much higher

excitation currents, possibly up to at least 10 A, to improve the performance in terms of

the noise and speed. For a coil of 2 turns, a current of 10 A ( = 20) would bring the

maximum SAR to 1.2 W kg

-1 which is still below, but is now approaching the IEC limit.

SAR modelling should therefore now become an integral part of MIT system design.

Page 208: Magnetic Induction Tomography for Imaging Cerebral Stroke

187 Simulated Stroke Imaging Using the MK1 System

7.7. Discussion

In this chapter, it was intended to make the simulations as realistic as possible by:

using an anatomically correct geometry

avoiding the inverse crime by employing different solvers as the predictor and

estimator

avoiding the inverse crime by adding phase noise to the simulated data; the noise

levels were comparable with those currently achievable in practical data-

collection systems

MIT imaging of the stroke lesions of different volumes and sites in the brain was

examined. Based on the obtained results, time-difference imaging of peripherally

located haemorrhages appears to be feasible, given the phase noise levels as low as 1 m

which is achievable in the MK2a system (see Appendix A.2).

Imaging of a centrally located small haemorrhage, apart from the unrealistic case of

noise-free data, was not successful using the adopted reconstruction algorithms. The

likely reason is the fact that the system is considerably less sensitive to the central

regions and, therefore, smaller signals are produced by the lesions located in such

regions; as was shown in figure 7.9. Regularizing the inverse solution diminishes the

effect of these small signals on the solution.

Automatic selection of the regularization parameter was found to be unsatisfactory in

some cases. The best images, in terms of stroke visibility, were obtained by visual

inspection of the images and empirically adjusting the regularization level. In this

regard, it is remarked that the purpose of this study is not to improve the inverse

solution techniques, but to employ existing methods to examine the feasibility of stroke

imaging using MIT. Therefore, by adjusting the regularization parameter, the most

potentially useful images were obtained. The inverse solution for different values of

can be obtained using the SVD of the sensitivity matrix, as explained in section 6.6.1.

Subsequently, the reconstructed images could be examined by a radiologist to extract

the most meaningful ones.

Frequency-difference imaging proved to be more difficult as nearly all the materials in

the imaging domain have frequency-dependent conductivities. It is possible that the use

Page 209: Magnetic Induction Tomography for Imaging Cerebral Stroke

188 Simulated Stroke Imaging Using the MK1 System

of geometrically customised systems, with higher sensitivity to the region of interest

(brain), might help to improve the visualisation of the desired feature, i.e. stroke. This is

investigated in chapter 8.

Although absolute imaging is increasingly used in geophysics [214], its application to

stroke imaging currently seems impractical using the available numerical techniques

and computing facilities. This would take more than one day for the FE model, and one

hour for the FD model, to perform each iteration of the non-linear solution for the head

model. This could jeopardise one of the main reasons for employing MIT for stroke

imaging, which is a quick diagnosis. Furthermore, as will be shown in chapter 9,

absolute images are very sensitive to the different types of errors, in particular

systematic errors.

Further improvements in the computational facilities and simulation methods, such as

the development of more time-efficient numerical packages together with parallel

computing1, may allow the feasibility of using iterative absolute image reconstruction

for stroke imaging. However, at the moment, differential imaging of the stroke is

preferable provided that:

'before-stroke' data set is available to perform time-difference imaging

hardware is capable of multi-frequency measurements to perform frequency-

difference imaging

SAR computations for the head revealed that, at present, MIT is probably still a

harmless and safe imaging method regarding heat generation. However, the SAR

computation should become an integrated part of MIT system design in future, as higher

coil currents are being considered by hardware designers.

End of chapter 7.

1 Parallelization approaches used in the FD solver are currently being investigated at the University of

Glamorgan as part of the LCOMIT project (e.g. [236]).

Page 210: Magnetic Induction Tomography for Imaging Cerebral Stroke

Chapter 8.

Customized Coil Arrays, Systematic

Errors, and the MK2b System

8.1. Introduction

As discussed in chapter 7, a customised MIT coil array may help to improve the

sensitivity of the system to the regions of interest in the target. The induced eddy-

current patterns in MIT and, therefore, the sensitivity distribution of the system not only

depend on the material property distribution of the target, but also on the geometrical

design of the MIT sensor array. Aspects of the geometrical design include:

the screen geometry

the number of channels and the number of coil turns in each channel

the coil dimensions (i.e. radii for circular coils)

the spatial coil configuration (position and orientation)

Peyton et al [32] investigated the effect of screen dimensions on the sensitivity maps.

They concluded that the screen geometry has little influence on the spatial sensitivity

distribution for a particular exciter-sensor pair; entries of the sensitivity matrix are

mainly only scaled by a constant factor. Watson et al [33] have shown that a coil-screen

standoff distance of 60-80% of the coil diameter gives a reasonable suppression of inter-

coil capacitive coupling without excessively damping the inductive signals. Therefore,

once the coil diameter and their spatial location in the MIT sensor array are known, an

appropriate diameter for the screen can be estimated.

Page 211: Magnetic Induction Tomography for Imaging Cerebral Stroke

190 Customized Coil Arrays, Systematic Errors, and the MK2b System

As discussed in section 7.6 for SAR computations, the coil turns (and the excitation

current as an operating parameter) have no effect on the eddy-current patterns; the

induced currents are scaled by a constant factor. The phase noise in the measurements,

however, will be affected by these factors; more coil turns induce larger signals,

resulting in a larger SNR (see equation (2.13)). The remaining factors are the number of

channels (coil combinations) and their spatial configuration. The coil radii are also

limited by the space available to accommodate the specified number of coils.

In a numerical study, Gursoy and Scharfetter [215] examined the effect of the location

of the sensors on the Tikhonov regularized time-differential images, for some fixed

exciter coils. They observed that by positioning the sensor coils at certain locations, the

resolution of the reconstructed image could be improved for a particular region, albeit at

the cost of other regions' resolution. They referred to such designs as 'regionally

focused' MIT.

In a later study [216], they investigated the effect of orientation of the sensor coils on

the same images, again assuming that the exciter coils were fixed. They concluded that

each coil orientation had its own merits and shortcomings for different imaging regions.

Eichardt et al [57] compared cylindrical (annular) and hemispherical coil arrays using a

sensitivity analysis. Their simulation results indicated that the hemispherical array

showed a higher sensitivity at the top of the brain compared to the cylindrical MIT

system. These findings, in fact, confirmed the results of the current chapter in this

thesis, which had been published prior to their report [62].

The purpose of this chapter is to compare the imaging performance of the customised

annular and hemispherical MIT arrays. To this end, frequency-difference imaging of

stroke lesions in the realistic head model is examined. Furthermore, the effect of

different values of SNR and examples of systematic errors (in terms of inaccuracies in

the boundaries) on the reconstructed images are investigated.

Later in this chapter, the geometrical design study of the latest MIT system in the

South-Wales group (MK2b) is briefly explained. The MK2b system, under construction

at the time of writing this thesis, is optimised for stroke imaging in terms of the 12-

tissue head model.

Page 212: Magnetic Induction Tomography for Imaging Cerebral Stroke

191 Customized Coil Arrays, Systematic Errors, and the MK2b System

8.2. Forward and inverse problems

In this chapter, the FD solver was used as the predictor and estimator models for solving

the forward and inverse problems, respectively. Employing the much faster FD solver

for all simulations allowed a variety of different arrays to be examined in a shorter time

(recalling from section 7.3.3, the FE and FD models took 64 hours and 65 min to

solve the forward problem for the MK1 system, respectively).

In order to avoid an inverse crime, two different discretizations were used in the FD

model throughout this chapter:

the forward problem was solved using a grid of 2 mm size, comprising 474897

voxels

the inverse problem was solved using a grid of 2.5 mm size, comprising 243083

voxels (sensitivity matrix was computed using this grid)

Based on the comparisons made between the FE and FD results in section 7.3.3, the FD

model with a grid of at least 2.5 mm size provided results of acceptable accuracy.

However, the simulations in that section were performed for the MK1 system. Here, it is

assumed that the same accuracy could be achieved if similar comparisons were to be

made for the MIT arrays under investigation in this chapter. This assumption meant that

the convergence tests were not carried out for every single MIT array, in the interests of

the computational time.

Initially, MIT arrays without an electromagnetic screen were considered, again in the

interests of the computational time. In this way, the simulations could be performed

without the process of introducing the screen into the FD model which requires the FE

model to be solved for free space (see section 5.5).

Therefore, the second assumption made in this chapter is that the effect of screen on the

reconstructed images, if any, would be similar in all the examined MIT arrays. In other

words, if a particular array performs better than the other ones, this superiority would be

the case with and without the screen. Therefore, the comparisons made between

different arrays without the screen are still meaningful.

As discussed in section 2.3, the screen helps to minimize the interference from

surrounding dielectric objects in a practical MIT system, and also the capacitive

Page 213: Magnetic Induction Tomography for Imaging Cerebral Stroke

192 Customized Coil Arrays, Systematic Errors, and the MK2b System

coupling between coils. However, these issues are irrelevant in the simulations

presented in this chapter. Later in this chapter, when the final design of the MK2b

system is defined, the screen is added to the system in order to obtain practically

realistic images. As will be shown, the images are influenced by the presence of screen.

8.3. Hemispherical and annular arrays

8.3.1. Hemispherical (Helmet) array

The first helmet shaped MIT array considered in this chapter consisted of 28 exciter and

28 sensor coils of 10 mm radius, positioned in 4 rows on a hemisphere (figure 8.1). Two

different helmet radii were considered:

100 mm, referred to as 'H100'

120 mm, referred to as 'H120'

The closest distance between a coil and the surface of the head was 5 mm for H100.

This particular configuration of the coils was adopted as it conformed better to the

region of interest for imaging, i.e. the brain. These arrays provided 28 28 = 784 coil

combinations.

Figure 8.1. Helmet MIT arrays: (a) side view of hemispheres H100 and H120 in position over head,

(b) coil positions on hemisphere H100, magnified in relation to (a). Angles of elevation of coil

centres in each ring are indicated with respect to lowest ring. Each of first three rows from bottom

includes 16 coils, and the upper row includes 8 coils. Exciters and sensors are shown as red and

black coils, respectively. Origin of coordinate system is defined as centre of helmet.

(a) (b)

z

y

67.5

45

22.5

H100 H120

Page 214: Magnetic Induction Tomography for Imaging Cerebral Stroke

193 Customized Coil Arrays, Systematic Errors, and the MK2b System

Figure 8.2. Annular MIT array. Three levels of array for each plane of measurement are indicated

as Z1, Z2, and Z3 which are level with first lower rows of H100 (see figure 8.1b). Two pairs of coils

for each plane of measurement are shown. Exciters and sensors are shown as red and black coils,

respectively. Coordinate system is that in Helmet arrays.

8.3.2. Annular array

The annular coil array modelled was similar to the MK1 system, consisting of 16 pairs

of exciter/sensor coils arranged in two concentric rings, but here the rings were smaller,

having radii of 100 and 110 mm for exciters and sensors, respectively. The radii of the

coils themselves were also smaller, at 10 mm, to match the coils of the helmet.

The array was positioned at 3 different heights with respect to the simulated head as

shown in figure 8.2. This provided a total of 768 (3×16×16) coil combinations,

comparable to that in the helmet systems (784). The middle plane passed through the

centre of the LP stroke. The spacing of the three planes of measurement are different in

order to match more closely the vertical extent of the helmet configurations; these

planes are level with three lower coil rings in H100.

8.3.3. Forward problem

In order to predict the phase changes due to the stroke, the forward problem was solved

using the grid of 2 mm size. Solving the FD model to compute each of the MIT data

sets, e.g. 784 measurements for H100 array at one frequency, took approximately 6

hours. The forward problem was solved for the frequencies of f1 = 1 MHz and f2 = 10

MHz.

Z3 = 70 mm

Z2 = 38 mm

Z1 = 0

1 cm

Page 215: Magnetic Induction Tomography for Imaging Cerebral Stroke

194 Customized Coil Arrays, Systematic Errors, and the MK2b System

Figure 8.3. Singular spectrum for three MIT arrays, normalised to the largest singular value and

plotted on a logarithmic scale.

8.3.4. Sensitivity analysis

The sensitivity matrix, A, was computed using the grid of 2.5 mm size, for a uniform

conductivity of 1 S

m

-1 and at the frequency of 10 MHz. All the sensitivity entries were

computed using (6.12), i.e. normalized to the primary signal. For each of the MIT

arrays, this took approximately 2 hours.

The SVD of A was computed as described in section 6.3.2. The singular values,

normalized to the largest singular value s1 in each system, are plotted on a logarithmic

scale in figure 8.3. The ratio between the largest and smallest singular values for the

annular, H100, and H120 sensitivity matrices was 9.1×107, 4.0×10

5, and 1.9×10

6,

respectively. This indicates that, for the particular discretization used, the sensitivity of

the H100 array is 230 times better conditioned than that of the annular system.

Total sensitivity maps for each system are plotted in transverse, coronal, and sagittal

cross-sections in figure 8.4. To this end, the sum of absolute values in each column of

the sensitivity matrix, corresponding to all coil combinations for one voxel in the grid,

were used to compute the total sensitivity map values (entries):

(8.1)

where aij are the entries of the sensitivity matrix A, m and n are the number of rows

(coil combinations) and columns (voxels) in A.

H100

H120

Annular

Page 216: Magnetic Induction Tomography for Imaging Cerebral Stroke

195 Customized Coil Arrays, Systematic Errors, and the MK2b System

Therefore, Map1 n = (map1, map2,..., mapn) is a row vector from which the 3D map is

constructed by placing each entry mapj in the corresponding voxel of the grid.

Absolutes values of aij were considered for the convenience of display.

Among the three arrays, a maximum total sensitivity value of 18 10-4

radian.m. was

obtained for the annular array in the nose region, evident in the sagittal plane. The upper

limit of the colour-bar in all maps was, therefore, set to this value.

As can be seen, the maps show a strong decay of total sensitivity values from the

periphery close to the coils towards the centre of the head. In the centre, the values are

almost zero; dark regions evident in all cross-sections.

The H100 array has a better sensitivity distribution in the desired brain region (see all

cross-sections in middle column). This is due to the coils' locations and orientations

conforming to the upper part of the head. Since the coils are positioned closer to the

scalp, it reveals higher values of total sensitivity than does the H120 array.

Annular H100 H120

Figure 8.4. Total sensitivity plots for three MIT arrays. Rows from top to bottom: transverse,

coronal, and sagittal cross-sections. Colour-bar in radian.m.S-1

.

10-4

Page 217: Magnetic Induction Tomography for Imaging Cerebral Stroke

196 Customized Coil Arrays, Systematic Errors, and the MK2b System

Target, ( )max = 0.24 ( )max = 0.27 ( )max = 0.24 ( )max = 0.18

Figure 8.5. A reconstructed image of LP stroke, shown with different colour-bars in S m

-1. The

values given above each image are upper limits in the colour-bars.

8.3.5. Image display

Prior to discussing the image reconstruction and the obtained results, let us examine the

image plots in terms of the colour representations. Figure 8.5 shows an arbitrary image,

reconstructed for the LP stroke. The maximum value of the estimated conductivity

change, occurring in one of the image voxels, is ( )max = 0.27 S m

-1. The second

column from left, shows the plotted image when the upper limit of the colour-bar is set

to 0.27. The visibility of stroke is poor. In the third column, the upper limit is set to the

maximum value in the simulated target, i.e. 0.24. Although still blurred, the stroke

visualization is slightly improved. The last column shows the same image when the

upper limit is further adjusted to 0.18. As can be seen, the visibility of stroke is now

clearly better. An extremely low upper limit would result in a totally uniform image.

Therefore, the presence of even a small volume artefact occurring in one voxel, but with

a high magnitude, may jeopardise the visibility of the desired feature. Hence, the proper

choice of the colour scales can be important for the visual assessment of the images. In

practice, once the image values are computed, the radiologist should be able to easily

change the colour scales in a visualization toolbox, in order to see whether a meaningful

image can be observed.

Page 218: Magnetic Induction Tomography for Imaging Cerebral Stroke

197 Customized Coil Arrays, Systematic Errors, and the MK2b System

8.3.6. Image reconstruction

A phase noise level of 1 m was added to all the predicted phase changes. The Tikhonov

regularized solution as in (6.25) was used to solve the inverse problem, and reconstruct

the conductivity changes between 1 and 10 MHz. Any negative values were truncated

from the solution as a-priori information (this physical constraint was explained in

section 6.7). The error in the solution obtained using a regularization parameter, , was

computed as:

(8.2)

where ||.||2 is the standard L2 norm. Note that the distribution of the conductivity change

in the target ( true) and image ( ) have different discretizations. For the purpose of

computing E, the target was interpolated onto 2.5 mm voxels, as if it had been simulated

using this grid.

The error was computed for a wide range of values (10-12

- 10-6

), and the optimised

regularization parameter was considered to be the value of that minimized the error.

In this way, the best feasible image, from a mathematical point of view, was obtained

for each array. The reconstructed images using the optimised values of were then

compared. For computing the norms, three volumes were considered for integrations:

entire head, to compute Ehead

only brain region, to compute Ebrain

the volume of the head superior to the base of brain, comprising the brain and

upper parts of the skull-CSF-muscles, to compute Eupper

These volumes are shown in blue in figure 8.6.

Figure 8.6. Integration volumes, shown as blue regions, for computing Ehead, Ebrain, and Eupper.

Ehead Ebrain Eupper

Page 219: Magnetic Induction Tomography for Imaging Cerebral Stroke

198 Customized Coil Arrays, Systematic Errors, and the MK2b System

Figure 8.7. Error Ehead in reconstructed images plotted in percentage as a function of , for three

MIT arrays.

The optimal λ, which gave the minimum value of Ehead, was 7.5 10-10

, 1.0 10-9

, and

5.0 10-10

for the annular, H100, and H120 arrays, respectively (see figure 8.7). The

corresponding minimum values of the error were 57.5%, 58.5%, and 55.5%.

The reconstructed images using the optimised values of , are presented in figure 8.8,

three middle columns. Since the reconstructed features were under-estimated in terms of

the conductivity values, an upper limit of the colour-bar (0.18), lower than the

maximum value in the simulated target (0.24), was chosen for displaying the

reconstructed images.

The maximum value for each particular image, ( )max, is provided to clarify whether

and to what extend the high magnitude values are being suppressed for the illustration.

Only some of the artefacts in the reconstructed images saturate the colour-bar, and none

of the reconstructed stroke features.

A blurred stroke feature, of amplitude 0.13 S m

-1, can be seen in the image from the

annular array (target is 0.24 S m

-1). The stroke is better visualised with H100 where the

feature has amplitude > 0.16 S m

-1. Increasing the radius of the helmet to 120

mm

(H120) has the advantage of reducing the artefacts of raised image value occurring near

the coils (see sagittal planes), but at the expense of reducing the stroke visibility.

10-12

10-11

10-10

10-9

10-8

10-7

10-6

50

60

70

80

90

100

Eh

ea

d

10-12

10-11

10-10

10-9

10-8

10-7

10-6

50

60

70

80

90

100

Eb

rain

Annular

H100

H120

Page 220: Magnetic Induction Tomography for Imaging Cerebral Stroke

199 Customized Coil Arrays, Systematic Errors, and the MK2b System

Target - LP

( )max = 0.24

Annular

( )max = 0.35

= 7.5 10-10

H100

( )max = 0.32

= 1.0 10-9

H120

( )max = 0.32

= 5.0 10-10

Annular - 1 plane

( )max = 0.35

= 5.0 10-11

Figure 8.8. Columns left to right: simulated LP target, and reconstructed images from phase

changes with 1 m added noise for annular (three planes), H100, H120, and annular (with one

plane) MIT arrays; was optimised based on Ehead. Maximum conductivity change occurred in

each images is given as ( )max. Rows from top to bottom: transverse, coronal, and sagittal planes.

Colour-bar in S m

-1.

Compared with the images obtained from the MK1 system in the previous chapter

(figure 7.24), conducting the measurements now for three planes, to acquire more data,

clearly improves the performance of the annular array. In order to further examine this,

data obtained from only one plane of measurement in the present annular array (figure

8.2) was used for image reconstruction. This was the lowest coil ring (Z = 0), which is

almost at the same level as the MK1 coils shown in figure 7.5. The reconstructed image

is displayed in the last column from left of figure 8.8, where was optimised in the

same way explained. The visibility of the stroke feature is clearly much degraded.

Figure 8.9 shows the errors Eupper and Ebrain as a function of . The minima exhibited by

the curve for Eupper were broad, indicating that there was no sharp optimisation of λ. The

likely reason for this is the fact that for high levels of regularization (the range shown

by a green arrow), the stroke feature vanishes, but so do the high amplitude peripheral

Page 221: Magnetic Induction Tomography for Imaging Cerebral Stroke

200 Customized Coil Arrays, Systematic Errors, and the MK2b System

artefacts outside the brain, thus lowering the error. The three middle columns in figure

8.10, show the reconstructed images using the optimised values of for Eupper. All

images are slightly more regularized than those for Ehead.

Figure 8.9. As figure 8.7, but for Eupper (a) and Ehead (b).

The fifth column from left in figure 8.10, shows an over-regularized image for H100 ( o

= 2.5 10-7

). According to the values of Eupper, two images shown for H100 have very

similar errors; 63.1% for optimised , and 64% for the over-regularizing o. This is a

clear example of where from a strictly mathematical point of view and using the

previously defined error measure by equation (8.2), the inverse solutions in both cases

exhibit similar accuracies. However, one is clinically more useful by visualizing the

10-11

10-10

10-9

10-8

10-7

10-6

60

65

70

75

80

85

90

95

100E

up

pe

r

10-12

10-11

10-10

10-9

10-8

10-7

10-6

50

60

70

80

90

100

Eb

rain

10-12

10-11

10-10

10-9

10-8

10-7

10-6

50

60

70

80

90

100

Eb

rain

Annular

H100

H120

10-12

10-11

10-10

10-9

10-8

10-7

10-6

50

60

70

80

90

100

Eb

rain

Annular

H100

H120

(a)

(b)

Page 222: Magnetic Induction Tomography for Imaging Cerebral Stroke

201 Customized Coil Arrays, Systematic Errors, and the MK2b System

stroke feature much more clearly, despite the presence of high-amplitude artefacts in the

scalp (see sagittal plane in middle column, figure 8.10).

It is therefore suggested that visual inspections should be conducted alongside with the

mathematical methods for assessing the images. In practice, when the target is

unknown, the suggested value of by an automatic method can be employed as an

initial estimation for deciding on the range of values to be examined (for this case

GCV = 8 10-11

, L-curve = 1 10-10

). A series of images using these values can be

reconstructed, and then visually inspected.

The images reconstructed using the optimised values of for Ebrain were slightly more

regularized than those for Ehead (not shown here as they were very similar to figure 8.8).

Because of the better visualisation of the stroke feature in the images obtained by H100

(see figure 8.8 and figure 8.10, middle column), only this array will be considered for

the remainder of this section.

Target - LP

( )max = 0.24

Annular

( )max = 0.29

= 2.5 10-9

H100

( )max = 0.27

= 2.5 10-9

H120

( )max = 0.27

= 7.5 10-10

H100

( )max = 0.22

o = 2.5 10-7

Figure 8.10. As figure 8.8, but reconstructed using values of minimising Eupper. Also fifth column

from left shows image reconstructed for H100 and using a high level of regularization.

Page 223: Magnetic Induction Tomography for Imaging Cerebral Stroke

202 Customized Coil Arrays, Systematic Errors, and the MK2b System

SP-target

( )max = 0.24

SP-image

( )max = 0.25

= 7.5 10-9

SD-target

( )max = 0.22

SD-image

( )max = 0.26

= 2.5 10-9

Normal-target

( )max = 0.13

Normal-image

( )max = 0.25

= 5.0 10-9

Figure 8.11. Columns left to right: simulated SP target, reconstructed SP image, simulated SD

target, reconstructed SD image, simulated normal brain, reconstructed normal image. Rows from

top to bottom: transverse, coronal, and sagittal cross-sections. Images reconstructed using phase

changes with 1 m added noise to H100 data, and optimised for Ebrain. Colour-bar in S m

-1.

Images reconstructed by minimizing Ebrain, are shown in figure 8.11 for the smaller

strokes and also for the normal head (no stroke present). Notice that, in this figure only,

the upper limit was further reduced to 0.15 S

m

-1. This is because the reconstructed

conductivity changes were further under-estimated compared to those obtained for the

LP stroke.

All images appeared very similar. The SP stroke can be barely recognised as a slight

left-right asymmetry in the transverse and coronal planes (2nd column from left). The

SD stroke cannot be visualised at all (4th column).

The image of the normal head (last column) is not completely symmetrical. The likely

reason is an unsymmetrical head geometry and the presence of 1 m random phase noise

in the data.

Page 224: Magnetic Induction Tomography for Imaging Cerebral Stroke

203 Customized Coil Arrays, Systematic Errors, and the MK2b System

8.4. Simulated errors

In this section, different types of error are added to the data, and their effect on the

reconstructed images of the LP stroke is investigated for the H100 array.

8.4.1. Phase noise

Although the images in the previous section were all reconstructed from the data

containing 1 m noise, here different levels of Gaussian phase noise (up to 20 m°

standard deviation) were considered in order to test the effect of the noise level on the

images.

First, using noise-free data, the regularization parameter was optimised as described in

the previous section: the value of that minimised Eupper was adopted ( = 2.5 10-9

).

Subsequently, noise of different levels were added to the data and all the noisy images

were reconstructed using the same = 2.5 10-9

. Strictly speaking, each noisy data set

will have its own optimal , which may be different from the adopted one. However, the

optimised for the noise-free data was used throughout this section.

The image error due to the added noise was evaluated as:

(8.3)

where noisy and noise-free are the images reconstructed using the data with and

without noise. The integration was carried out over the upper portion of head; Eupper in

figure 8.6.

The error Enoise thus is a measure of the spurious content in the image due to the noise,

as a percentage of the noise-free image content, and increases roughly in proportion

with the phase noise added to the simulated measurements (figure 8.12). Three runs of

the model, each with its own randomly generated noise for each noise level, were used.

Examples of images are shown in figure 8.13, including two noise levels of particular

interest: 17 and 1

m° relating to the MK1 and MK2a practical data-collection systems

(see Appendix A). Compared with the noise-free image, the image is slightly degraded

Page 225: Magnetic Induction Tomography for Imaging Cerebral Stroke

204 Customized Coil Arrays, Systematic Errors, and the MK2b System

by a noise level of 1 m°, and the stroke feature remains clearly visible. For 17 m° added

phase noise, however, artefacts dominate the image and the stroke cannot be visualized.

Figure 8.12. Value of error, Enoise, in reconstructed images using H100 (λ = 2.5x10-9

) for different

levels of added Gaussian phase noise (standard deviation). Three simulations (different colours) for

each noise level are shown along with the fitted straight line.

Target - LP

( )max = 0.24

Noise-free

( )max = 0.25

1 m

( )max = 0.26

3 m

( )max = 0.34

17 m

( )max = 1.44

Figure 8.13. Columns left to right: simulated LP target, and reconstructed images for H100 with

different levels of added phase noise ( = 2.5 10-9

). Rows from top to bottom: transverse, coronal,

and sagittal cross-sections. Colour-bar in S m

-1.

0 2 4 6 8 10 12 14 16 18 200

20

40

60

80

100

120

140

Noise level (m )

En

ois

ey = 6.43 x

Page 226: Magnetic Induction Tomography for Imaging Cerebral Stroke

205 Customized Coil Arrays, Systematic Errors, and the MK2b System

Of course, the noise artefacts will change for every run of the random number generator,

but qualitatively the images will remain fairly similar. The artefacts in the image can be

reduced by increasing λ, but at the disadvantage of blurring or losing the stroke feature.

The purpose here is to define the tolerable noise level which does not degrade the

visibility of stroke.

By inspection of the images for different runs of randomly generated noise, it appeared

that an acceptable noise level would be between 1 and 4 m , and could reasonably be set

at 3 m . This leads to an acceptable value of the error measure, Enoise, of approximately

20% (see figure 8.12).

8.4.2. Systematic errors

It is widely appreciated that MIT image reconstruction will be improved by introducing

as much a-priori information as possible. One example of this is to determine the

boundary of the eddy-current region, in this case the shape of the head, by some

independent means such as by optical shape scanning (see e.g. [217]). This boundary is

of great importance as it marks the large contrast in dielectric properties between the

tissues and air.

The sensitivity matrix will then be computed for this known region. Accurate

calibration of the optical method will be important, and so will accurate registration of

the optically-determined boundary with the known geometry of the MIT array.

Inevitably there will be errors in these processes.

In order to examine the effect of such inaccuracies, two examples of geometrical

systematic errors were simulated. The first simulated a mis-registration equivalent to a

displacement of the head relative to the helmet, separately in x-, y- and z-directions

(figure 8.14a). The second involved a simple scaling error simulated by an expansion or

contraction of the entire head (figure 8.14b).

Since the coils do not form part of the FD discretization, but instead are perfect circles

whose coordinates and orientations are specified separately, it is very easy to model

these geometrical changes. A shift of the head was introduced by altering the

coordinates of all the coils by a fixed amount, whilst leaving the FD mesh unchanged.

Page 227: Magnetic Induction Tomography for Imaging Cerebral Stroke

206 Customized Coil Arrays, Systematic Errors, and the MK2b System

An expansion or contraction of the head was introduced by altering the FD mesh size

whilst leaving the coils unchanged.

Figure 8.14. Exaggerated geometrical errors: light pink and dark blue colours show the 'true' and

'erroneous' boundaries, respectively. (a) displacement in y-direction. (b) contraction.

Note that the data computed using the changed („erroneous‟) boundary will differ from

the measured data, i.e. data computed using the original („true‟) boundary. This

difference is the systematic error due to the use of wrong geometry that may happen in

practice. For image reconstruction, the sensitivity matrix was computed using the

changed boundary of the head. The simulated MIT measurements at 1 and 10 MHz

were computed using the original head boundary. No random noise was included in any

of these simulations. The image in the 2nd column from left in figure 8.15 (true

boundary) is identical to that in figure 8.13 (noise-free), and is reproduced here for ease

of comparison.

Displacement of the head in the negative x-direction, produced artefacts which appear

as increased values on the surface that faces towards the positive x-direction, and

decreased values on the surface that faces towards the negative x-direction (figure 8.15,

3rd and 4th columns from left).

A displacement in the opposite direction produced the opposite effect (figure 8.15, 5th

and 6th columns). Obviously, larger displacements resulted in more pronounced

artefacts. For instance, compare those for x = +3 and x = +5 mm in columns 5 and 6.

The artefacts for x = +5 mm have degraded the stroke feature. Similarly,

displacements of the head in y- and z-directions produced artefacts near to the surfaces

facing the corresponding directions (see figure 8.16).

(a) (b)

Page 228: Magnetic Induction Tomography for Imaging Cerebral Stroke

207 Customized Coil Arrays, Systematic Errors, and the MK2b System

Target - LP

( )max = 0.24

True boundary

( )max = 0.25

x = -3 mm

( )max = 0.32

x = -5 mm

( )max = 0.39

x = +3 mm

( )max = 0.31

x = +5 mm

( )max = 0.38

Figure 8.15. Columns left to right: simulated LP target, and reconstructed images with different

boundary errors in x-direction. Rows from top to bottom: transverse, coronal, and sagittal planes.

Target - LP

( )max = 0.24

True boundary

( )max = 0.25

y = -5 mm

( )max = 0.36

y = +5 mm

( )max = 0.42

z = -5 mm

( )max = 0.39

z = +5 mm

( )max = 0.27

Figure 8.16. As figure 8.15, but for y- and z-direction displacements.

y

z

x

y

z

x

Page 229: Magnetic Induction Tomography for Imaging Cerebral Stroke

208 Customized Coil Arrays, Systematic Errors, and the MK2b System

Contraction of the head (scaling error) produced increased values near the surface of the

head (figure 8.17, 3rd and 4th columns from left). This is to be expected. Note that in

contracting the geometry, a layer of conductive material is in fact removed (ignored)

from the periphery. This would be the pink layer in figure 8.14b. Therefore, when the

sensitivity matrix is computed for the contracted geometry, the signals received from

the absent material is not accounted for in the sensitivity matrix. The contribution to the

signals from this material results in higher values of conductivity at the surface in the

images. Therefore, artefacts of raised conductivity will appear on the boundary.

Target - LP

( )max = 0.24

True boundary

( )max = 0.25

3% Cont.

( )max = 0.36

5% Cont.

( )max = 0.45

3% Exp.

( )max = 0.21

5% Exp.

( )max = 0.20

Figure 8.17. As figure 8.15, but for contraction and expansion scaling errors.

Expansion of the head, however, produced artefacts as reduced image values on the

periphery (figure 8.17, 5th and 6th columns). Note that since haemorrhage is always

going to be a region with positive image values, its visibility may even improve in the

case of an expansion of the geometry.

Page 230: Magnetic Induction Tomography for Imaging Cerebral Stroke

209 Customized Coil Arrays, Systematic Errors, and the MK2b System

Indeed, in the absence of a sufficiently accurate method of a-priori boundary

determination, it may be worth deliberately using an oversized boundary for the

computation of the sensitivity matrix in order to ensure that all image errors will be

negative, and not confused with a haemorrhage. This possibility deserves further

investigation.

In a similar way to the calculation of Enoise in section 8.4.1, the error due to inaccuracies

in the boundary was computed as:

(8.4)

where erroneous and errorless are the images reconstructed with and without boundary

errors ( errorless is identical to noise-free in equation (8.3)).

This time, Eboundary is a measure of the error in the image content due to the systematic

errors expressed as a percentage of the original image content (image without boundary

error). The values of Eboundary are plotted in percentage in figure 8.18, for displacement

and scaling errors.

Figure 8.18. Image error, Eboundary, for different displacements of head or percentage scaling.

Applying the same acceptance criterion as for the level of random noise, i.e. Eboundary <

20%, indicates that the mis-registration displacement error should be no more than 3-4

mm, and the size scaling error should be no more than about 3%.

-5 -4 -3 -2 -1 0 1 2 3 4 50

5

10

15

20

25

30

35

40

Displacement (mm) - Scaling (%)

Eb

ou

nd

ary

x-direction

y-direction

z-direction

Scaling

Page 231: Magnetic Induction Tomography for Imaging Cerebral Stroke

210 Customized Coil Arrays, Systematic Errors, and the MK2b System

8.5. MK2b system

As discussed previously, conducting the data acquisition for multiple planes of

measurements in annular arrays can help to improve the spatial resolution of the

reconstructed images, despite the fact that not all the added information is independent

from those obtained from the first plane of measurement. In practice, this may be done

by applying a scanning protocol such as coil or target movements.

However, any type of movement could result in the introduction of further errors in the

data; e.g. boundary errors. Therefore, it is preferable to carry out the measurements in

one simple process, avoiding the movement of the head or sensor array.

Based on the results of previous sections in this chapter, it appears that hemispherical

arrays might be the most suitable coil configurations for stroke imaging. For a simulated

phase noise level of < 3 m , acceptable images of a peripheral stroke were produced for

this configuration. The electronics developed by the South-Wales group for the MK2a

system, can achieve a low level of phase noise (1 m°), but is not compatible with 56

coils as used in the H100 array.

In order to employ the new electronics architecture, the number of exciters must be a

multiple of 16, and the number of sensors a multiple of 7 [63]. By taking into account

the physical and practical constraints to design a feasible system, this section describes

a numerical study for the design of the MK2b system.

8.5.1. Concept of MK2b array

As shown in figure 8.19, the system is envisaged as an array of separate exciter and

sensor coils positioned on a hemisphere (radius RC), inside an hemispherical

electromagnetic screen (outside radius RS). An inner, concentric, plastic hemispherical

shell (inside radius RP) is to be fixed inside the coil array so as to completely enclose

and protect the coils.

For the electromagnetic screen, hemispherical spun aluminium bowls of 3 mm thickness

are commercially available. Plastic shells (3 mm thickness) can be vacuum formed in

our own Radiotherapy Physics Mould Room.

Page 232: Magnetic Induction Tomography for Imaging Cerebral Stroke

211 Customized Coil Arrays, Systematic Errors, and the MK2b System

Figure 8.19. Conceptual diagram of MK2b system (not to scale).

Figure 8.20. Standard measurement indexes for head (after ARTUR [218]).

8.5.2. Dimensions

The plastic shell needs to be small enough to allow the coils to be placed as close as

possible to the head surface in order to maximise the MIT signals, and at the same time

large enough to accommodate most adult heads. The 1988 U.S. Army Anthropometry

Survey provides head measurements from 1774 males and 2208 females [218]. By

examining the data, the most relevant dimension is the length (anterior–posterior

distance from between the eyebrows to the back of the cranium, as shown in figure

8.20). A hemisphere based on the head length, will certainly be large enough to suit the

breadth and coronal arc for that head.

The measurement data fell within the length of 209.5 mm for 96.5% of the males and

99.95% of the females. A further margin of 10 mm was allowed for cushioning, leading

to an inside shell radius of RP =

115

mm in figure 8.19. The coils are positioned further

out by another 10 mm, on a hemisphere of radius RC = 125 mm. The coil-shell

separation is, therefore, 7 mm. A coil-screen standoff distance of 32 mm is chosen as

this allows the use of a commercially available size of spun aluminium hemisphere with

outside radius of RS = 160 mm.

coils

screen

shell

head

RS

RC

RP

Length Breadth Coronal Arc

Page 233: Magnetic Induction Tomography for Imaging Cerebral Stroke

212 Customized Coil Arrays, Systematic Errors, and the MK2b System

Figure 8.21. Top view of three coil arrays; red and black coils are exciters and sensor coils,

respectively.

8.5.3. Number of coils

The number of coils in the array determines the maximum coil diameter that will fit

onto the hemisphere. Initially, three arrays were considered as shown in figure 8.21:

30 coils of 45 mm diameter: 16 exciters + 14 sensors

46 coils of 35 mm diameter: 32 exciters + 14 sensors

60 coils of 30 mm diameter: 32 exciters + 28 sensors

The exciter and sensor coils are considered to be distributed in the array (figure 8.21). In

order to compare the performance of these three arrays, frequency-difference imaging of

the LP stroke was performed. The FD solver, with grid sizes of 2 and 2.5 mm, was used

for the forward and inverse problems (see section 8.2). The screen was excluded from

the models at this stage, so that the simulations could be conducted without the need for

time consuming step of solving the FE model (see section 8.2).

The forward problem was solved for the frequencies of 1 and 10 MHz, and 1 m noise

was added to the simulated phase changes. In order to find the optimised , the error

Ehead was minimised (see equation (8.2) and figure 8.6). The minimum computed values

of Ehead were 67, 64.4, and 64.5% for arrays comprising 30, 46, and 60 coils,

respectively.

The reconstructed images are shown in figure 8.22 (three middle columns). All three

arrays produced images which were poor compared to those obtained from the 58-coil

array (H100) in the previous section (e.g. see figure 8.13 for 1 m noise). The likely

reasons are:

30 coils 46 coils 60 coils

Page 234: Magnetic Induction Tomography for Imaging Cerebral Stroke

213 Customized Coil Arrays, Systematic Errors, and the MK2b System

Target - LP

( )max = 0.24

30-coil

( )max = 0.50

= 7.5 10-11

46-coil

( )max = 0.31

= 2.5 10-10

60-coil

( )max = 0.34

= 2.5 10-10

46-coil + screen

( )max = 0.45

= 2.5 10-10

Figure 8.22. Columns left to right: simulated LP target, and reconstructed images for 30-coil, 46-

coil, 60-coil, and 46-coil+screen arrays; reconstructed with 1 m noise, and optimised for Ehead.

Rows from top to bottom: transverse, coronal, and sagittal cross-sections. Colour-bar in Sm-1

.

larger distance between the coils and head (RC = 125 mm in figure 8.19

compared with 100 mm for H100 array).

use of fewer exciter-sensor coils (for 30-coil and 46-coil arrays)

Nevertheless, the stroke feature can still be recognised in all images.

For the 30-coil array (2nd column from left), the stroke is considerably displaced

towards the face (evident in transverse plane), and the image contains artefacts of high

amplitude (( )max = 0.5 S m

-1). The 46-coil and 60-coil arrays have a fairly similar

performance, both in terms of the error Ehead and the visualization of stroke, which is

better than the performance of the 30-coil array.

Since adding a further 14 sensors to form the 60-coil array would have increased the

cost of the electronic components, the 46-coil array was chosen for the practical

implementation.

Page 235: Magnetic Induction Tomography for Imaging Cerebral Stroke

214 Customized Coil Arrays, Systematic Errors, and the MK2b System

For the 46-coil array with 35 mm diameter coils, the screen standoff distance (32

mm) is

91% of the coil diameter, which is slightly larger than the range 60-80% used by

Watson et al [33].

Finally, following the method described in section 5.5, the screen was added to the 46-

coil array by solving the FE model. The reconstructed image is shown in the last

column in figure 8.22. Comparing to the image obtained by the same array without the

screen (middle column), adding the screen clearly improved the stroke visibility. It is

less under-estimated in amplitude ( 0.17 S m

-1). However, the minimum error remained

the same (Ehead = 65%). This is due to the presence of artefacts of high amplitude on the

periphery (( )max = 0.45 S m

-1). The remainder of this chapter employs the 46-coil

array with the screen included.

8.5.4. Other coil arrangements and quality measurement

Another investigation in the design of the sensor array was to decide which coils acted

as the exciters, and which ones as the sensors. A distributed arrangement1, with the

exciters and sensors interspersed, was previously considered (see figure 8.21).

According to the Lorentz reciprocity theorem discussed in section 5.5, the mutual

inductance between a pair of exciter-sensor coils is unchanged if they are interchanged.

However, their relationship with other coils in the array will change. Therefore,

different coil arrangements may not perform in the same way.

The number of possible ways to select 32 exciters from 46 coils is 46C32 = 46!/(32! 14!)

= 239,877,544,005 arrangements. Clearly, a number of assumptions should be made to

limit the possible arrangements, since not all of them can be numerically investigated

due to time restrictions.

For practical implementation, it is preferable to separate the exciter and sensor coils, so

to make the circuitry simpler; e.g. reducing cable lengths. Two examples of such

arrangements are shown in figure 8.23, one symmetrical about the x- and y-axes (right

array), and one asymmetric about y-axis (left array).

1 Note that throughout this thesis, the term 'coil configuration' is used to describe the spatial positioning of

the coils (location and orientation) and the number of exciters and sensors. The term 'coil arrangement'

refers to which coils are exciters and which are sensors for a given coil configuration.

Page 236: Magnetic Induction Tomography for Imaging Cerebral Stroke

215 Customized Coil Arrays, Systematic Errors, and the MK2b System

Figure 8.23. Examples of separated coil arrangements favoured from the point of view of the

electronics layout; red coils are exciters.

Target - LP

( )max = 0.24

= 1 10-9

( )max = 0.28

= 1 10-10

( )max = 0.40

= 1 10-11

( )max = 0.81

Figure 8.24. Columns left to right: simulated LP target, and images reconstructed from noise-free

data and using different values of for asymmetric arrangement shown in figure 8.23 (left array).

Rows from top to bottom: transverse, coronal, and sagittal cross-sections. Colour-bar in Sm-1

.

The use of asymmetric coil arrangements, however, results in artefacts on the

reconstructed images. In order to demonstrate this, the asymmetric arrangement in

figure 8.23 (left array) was simulated. Images of the LP stroke were reconstructed from

noise-free data and with three different levels of regularization: optimised value of =

10-10

, one smaller ( = 10-11

) and one larger ( = 10-9

) value than the optimised value

(see figure 8.24).

y

x

Page 237: Magnetic Induction Tomography for Imaging Cerebral Stroke

216 Customized Coil Arrays, Systematic Errors, and the MK2b System

As can be seen in the transverse and coronal planes in figure 8.24, artefacts of relatively

large volume are evident in the left hemisphere of brain, on the opposite side to the

stroke location, and also in the central regions. Therefore, symmetric arrangements were

intuitively favoured in this chapter. A detailed investigation is required in order to

realise the possible reasons for this behaviour in asymmetric arrangements.

Even after excluding the asymmetric arrangements, the number of possibilities is still

too large to permit assessment of every individual arrangement. Therefore, six

symmetric arrangements, separated and distributed, were considered for comparison;

labelled as D1 to D6 as shown in figure 8.25.

Figure 8.25. Six selected coil arrangements for the 46 coil configuration; red coils are exciters. D6 is

the arrangement used previously in figure 8.21 (middle array).

A quality measure, which has been employed in geophysics for tomographic problems

[219], was used for a quantitative comparison of the different coil arrangements. Gursoy

and Scharfetter [215] used this measure for comparing the performance of sensor coils

placed in different locations in a MIT array (i.e. different coil configurations). Here, the

coil configuration is fixed, but the arrangement is varied as in figure 8.25.

A measure of the orthogonality of the rows in the sensitivity matrix is given by:

D1 D2 D3

D4 D5 D6

Page 238: Magnetic Induction Tomography for Imaging Cerebral Stroke

217 Customized Coil Arrays, Systematic Errors, and the MK2b System

(8.5)

where ri and rj are the i-th and j-th rows of the sensitivity matrix. The quality measure is

defined as:

(8.6)

where m is the total number of rows (coil combinations) in the sensitivity matrix. The

quality measure reflects the linear independence of the different exciter-sensor

combinations [219]. The more orthogonal the rows are, the more independent

information they provide, and the less ill-conditioned the sensitivity matrix is. A matrix

with orthogonal rows ( ij = 0) will result in the maximum value of Qm = m (m-1)/2. On

the contrary, a matrix with totally dependent rows ( ij = 1) will have the worst quality

measure of Q = 0. For the 46-coil array, m = 32 14 = 448 and the maximum value of Q

is Qm = 100128.

The sensitivity matrix was computed for arrangements D1-D6, where its entries were

computed using (6.12), i.e. normalized to the primary signal. The quality measure was

calculated for each design using (8.6), and was normalized to Qm (=100128). The

normalized values are shown in figure 8.26 as percentages. As can be seen, these values

spanned only a small range, (79.2-80.3%), indicating that there was no clear best coil

arrangement. No quality measure values in the work by Gursoy and Scharfetter [215]

were reported to be compared against the values obtained here.

Figure 8.26. Quality measures for 6 designs, D1-D6, normalized to Qm = 100128.

D1 D2 D3 D4 D5 D679

79.2

79.4

79.6

79.8

80

80.2

80.4

80.6

Qu

alit

y (

%)

Page 239: Magnetic Induction Tomography for Imaging Cerebral Stroke

218 Customized Coil Arrays, Systematic Errors, and the MK2b System

According to the quality measurements, none of the considered coil arrangements (D1-

D6) will have a distinctly better performance than the rest; the reconstructed images

confirmed this.

Initially, for comparison, no noise was added to the simulated data so to avoid any

random error affecting the images. In order to find the optimised , the error Ehead was

minimised (see equation (8.2) and figure 8.6). In the presence of screen, the computed

conductivity changes tended to have higher amplitudes. Therefore, the upper limit of

colour-bar was set to the maximum value in the simulated LP target (( )max = 0.24 S

m-1

) for illustrations in figure 8.27.

The stroke feature can be clearly visualized in all images, with some being more under-

estimated in amplitude (D1-D2 & D5), and some being slightly more spread (D3 & D6)

evident in the transverse planes.

D1

( )max = 0.33

= 2.5 10-10

D2

( )max = 0.35

= 2.5 10-10

D3

( )max = 0.58

= 1.0 10-10

D4

( )max = 0.33

= 2.5 10-10

D5

( )max = 0.42

= 2.5 10-10

D6

( )max = 0.42

= 1.0 10-10

Figure 8.27. Columns left to right: reconstructed images for D1 to D6 designs, respectively, where

was optimised for Ehead. Rows from top to bottom: transverse, coronal, and sagittal cross-sections.

Colour-bar in S m

-1 with upper limit of 0.23.

Page 240: Magnetic Induction Tomography for Imaging Cerebral Stroke

219 Customized Coil Arrays, Systematic Errors, and the MK2b System

D1

( )max = 0.53

D2

( )max = 0.37

D3

( )max = 0.61

D4

( )max = 0.40

D5

( )max = 0.39

D6

( )max = 0.37

Figure 8.28. As figure 8.27, but with 1 m added noise.

Figure 8.28 shows the images with 1 m added noise, where was optimised using the

noisy data. A value of = 2.5 10-10

was obtained for all coil arrangements. In the image

obtained from D6 (last column from left), the stroke feature was slightly better

visualized than in the rest of the noisy images obtained from D1-D5 coil arrangements.

More investigations are being carried out in order to make the final decision on the

arrangement of the coils.

8.5.5. MK2b construction

At the time of writing, the MK2b system is under construction. Figure 8.29 presents two

photographs from the early stages of the construction at the Medical Physics & Clinical

Engineering Department, Singleton Hospital, Swansea. The array is now with the

LCOMIT partners at the University of Glamorgan for the addition of the electronics and

finalizing the sensor array.

Page 241: Magnetic Induction Tomography for Imaging Cerebral Stroke

220 Customized Coil Arrays, Systematic Errors, and the MK2b System

Figure 8.29. Photographs of MK2b system at early stages of construction. Left: screen and coil

formers before mounting. Right: array (screen, formers, plastic shell) prior to adding the

electronics and coil windings.

8.6. Discussion

Hemispherical arrays conform more closely to the shape of the head than a conventional

annular array, thus increasing the sensitivity of the MIT array to conductivity changes.

By examining the total sensitivity maps, it was shown that for the coils placed

sufficiently close to the head, a better sensitivity distribution can be achieved, in

particular, in the upper parts of the head (map for H100 in figure 8.4).

Hemispherical arrays offer the possibility of using a larger number of the coils, than in

the annular ones. This allows the measurements to be carried out without the need to

move the head or the array, and therefore avoids further systematic errors.

For the chosen discretizations and the inverse solution techniques, no clear optimised

level of regularization could be found for brain imaging. In practice, it may be useful to

reconstruct a series of images with different levels of regularization, and inspect them

visually in order to decide whether a meaningful feature can be recognised in the

images. For instance, just the detection of an asymmetry in the brain may be very

significant clinically. Furthermore, presenting the images with different colour limits

may improve the visualization of the features under-estimated in amplitude.

Frequency-difference MIT imaging of a large intra-cerebral haemorrhage ( 50 ml),

peripherally located, seems to be feasible, provided that the phase noise level in the

hardware is small enough ( 3 m ), and any error in the assumed a-priori boundaries

remains low ( 4 mm for displacement and 3% for scaling).

Screen

Former

Page 242: Magnetic Induction Tomography for Imaging Cerebral Stroke

221 Customized Coil Arrays, Systematic Errors, and the MK2b System

However, imaging of a small haemorrhage (8 ml), located deep within the brain, proves

to be more difficult, even with noise-free measured data, and no systematic errors in the

assumed boundaries of the head. This is due to the lower sensitivities of MIT to the

central regions of the imaging domain.

In this chapter, only two types of the boundary errors were examined; displacement and

scaling. Mis-registration of the measured boundaries may also occur, for example, in

terms of rotation of the head about all three axes. Further errors can also be introduced

if the patient's head moves during the measurement process, as suggested by Gursoy

and Scharfetter [220], unless the head's position is somehow fixed within to the sensor

array. Further investigations are required in order to examine the effect of such errors on

the images.

Contraction of the head geometry, as a scaling systematic error, produced artefacts of

small volume but high amplitude on the periphery (figure 8.17). Such raised

conductivities close to the scalp could be mistaken for small subarachnoid

haemorrhages. On the contrary, artefacts of negative sign were produced by an

expansion of the head (same figure). However, such artefacts cannot be confused by

haemorrhage which appears as a raised conductivity region. The stroke visibility may

even improve, as shown for the particular example presented in the images.

Therefore, in case of uncertainty about the accuracy of the measured boundaries, it is

recommended that using a head geometry, slightly larger than the available one, might

help to avoid the appearance of artefacts of high amplitude. If the only geometrical data

available is the approximate overall dimensions of the head in three directions, one

might consider computing the sensitivity matrix for a sphere or ellipsoid large enough to

accommodate the head. In the FE mesh construction in figure 7.4(II), such an ellipsoid

was shown in which the head was confined. Obviously, detailed investigations are

required in order to examine the images that one would obtain using these approximate

boundaries for the eddy-current region.

The design process of the MK2b MIT system was described. Different coil

configurations and arrangements were examined, and frequency-differential images of

the LP stroke were reconstructed. The conductivity distribution in the target was

assumed to be known, and was employed for adjusting the regularization level during

image reconstructions. This allowed achievement of the best possible images for each

Page 243: Magnetic Induction Tomography for Imaging Cerebral Stroke

222 Customized Coil Arrays, Systematic Errors, and the MK2b System

MIT array, using the particular inverse solution technique employed. The best image

was taken as that which produced the minimum error in the estimated conductivity

change values.

End of chapter 8.

Page 244: Magnetic Induction Tomography for Imaging Cerebral Stroke

Part IV

Experiments and Conclusions

Page 245: Magnetic Induction Tomography for Imaging Cerebral Stroke

Chapter 9.

Experiments

9.1. Introduction

One of the main purposes in the LCOMIT project was to design a practical MIT

imaging system for biomedical and other low-conductivity applications. This has been

pursued by the LCOMIT partners at the Universities of Glamorgan and Manchester.

The MK2a system, described in Appendix A.2, is a recently developed hardware with a

low level of phase noise (1 m ), which may allow imaging low-conductivity targets.

The purpose of this chapter is to carry out the first practical MIT imaging using the

MK2a system and the image reconstruction algorithms introduced in chapter 6. To this

end, a geometrically simplified model comprising cylindrical domains, similar to those

employed in chapter 6, are used. Absolute and differential imaging of the targets are

performed.

9.2. Experimental setup

All the measurements described in this chapter were carried out using the MK2a system

(see Appendix A) and at frequency of 10 MHz. The experimental setup comprised a

cylindrical tank (19.5 cm diameter) as the background, and a smaller solid cylinder (4

cm diameter, 4 cm height) as the inclusion. The inclusion was placed within the tank at

two different positions as shown in figure 9.1; two setups are referred to as 'case-A' and

'case-B'.

Page 246: Magnetic Induction Tomography for Imaging Cerebral Stroke

225 Experiments

Figure 9.1. Conceptual diagram of targets in case-A and case-B. Upper row shows x-y cross-

sections, and lower row shows vertical cross-sections along broken lines shown in upper row.

Distances show displacement of inclusion relative to centre of tank.

The tank was filled with aqueous saline solution ( = 0.205

S

m

-1) to the depth of 16 cm,

and the inclusion was a piece of Perspex ( = 0

S

m

-1). The position of the inclusion

within the tank was maintained using a length of nylon fishing line (see photograph in

figure 9.2a), and the tank was placed at the centre of the MK2a array (see figure 9.2b).

The secondary voltage, Vs, was measured for 9 equally-spaced (2 cm) planes of

measurement by moving the tank within the array in z-direction (figure 9.3). This

provided 14 14 9 = 1764 coil combinations. In order to adjust the vertical level of the

tank, wooden spacer disks of 1 cm thickness were placed beneath the tank. A plastic

guide was employed for the horizontal placement of the tank within the array (see figure

9.2b). Three runs of measurements were conducted:

uniform tank with no inclusion

inclusion placed within the tank according to case-A (figure 9.1)

inclusion placed within the tank according to case-B (same figure)

Prior to placing the target within the array for each plane of measurement, the primary

signal for empty space, Vp, was acquired. In theory, the primary signal measured at

different times should remain the same. However, multiple measurements for Vp were

performed so the Vs could be normalized to a Vp measured close in time so as to

minimise the effect of drift in the hardware.

5 cm

5 cm

5 cm

case-A case-B

Tank

Inclusion

y

x

z 16 cm

19.5 cm

Page 247: Magnetic Induction Tomography for Imaging Cerebral Stroke

226 Experiments

Figure 9.2. Photographs of experimental setup showing inclusion within tank (a) and tank within

MK2a array (b).

Figure 9.3. Planes of measurement: (a) levels of 9 planes (P1-P9) relative to tank, (b) & (c) highest

and lowers positions of tank within array, respectively.

(a)

(b)

Nylon line

Inclusion

Tank

Tank

MK2a

array

2 cm

(a) (b) (c)

Screen

Tank

Ex array

Se array

P1

P1 P9

P9

Guide

P5

Page 248: Magnetic Induction Tomography for Imaging Cerebral Stroke

227 Experiments

9.3. Simulations

The experiment was simulated using the FD model. Revisiting figure 5.11 in chapter 5,

grid sizes of 1 cm provided results of acceptable accuracy for simulating a cylindrical

domain of 20 cm diameter and height. Here, in order to achieve a better accuracy, the

smallest grid size which the FD solver could handle was considered. This corresponded

to a grid of 0.05 cm size, allowing a better approximation of the curved surfaces in the

geometry using the cubic voxels. The use of smaller grid sizes than 0.05 cm resulted in

excessive numbers of voxels for representing the cylindrical tank, thus preventing the

FD solver to run (the arrays sizes defined in the FD solver grew too large). For the

employed grid, solving the forward model for each run of measurement (9 planes of

measurement corresponding to 1764 coil combinations) took approximately 15 hours.

All three runs for the uniform tank, case-A, and case-B were simulated.

9.4. Signal analysis

Figure 9.4a is a plot of the 1764 simulated phase changes, , due to the uniform tank.

The data from each plane of measurement can be distinguished from those from the

other planes by their maximum values. The maximum magnitude of the phase changes

predicted by the model is 1.83 , which occurs for the middle plane of measurement.

Figure 9.4b shows the same plot, but for the measured data plotted on the same y-axis

scale. The maximum magnitude of the phase changes is 2.27 . As can be seen, despite

the presence of some 'spikes' in the phase changes, the data corresponding to each plane

of measurement can be distinguished from those for other planes.

The difference between the simulated and measured data, plotted in figure 9.4c, shows a

discrepancy as large as 0.44 in magnitude. This phase error is significant compared

with the genuine phase changes produced by the tank itself. The largest discrepancies

are observed for the phase changes of large magnitude, corresponding to the middle

plane of measurement. The root mean square (r.m.s.) phase error is 0.0843 .

The discrepancy between the simulated and measured data can significantly affect the

absolute imaging, which is examined later in this chapter. Also, the possible reasons for

this discrepancy are investigated in the discussion.

Page 249: Magnetic Induction Tomography for Imaging Cerebral Stroke

228 Experiments

Figure 9.4. Phase changes in degrees for uniform tank: (a) simulated data for tank of 19.2 cm

diameter, (b) measured data for tank of 19.5 cm diameter, (c) difference between (a) and (b).

0 200 400 600 800 1000 1200 1400 1600 1800-2.5

-2

-1.5

-1

-0.5

0

0.5

Coil combination

Phase c

hanges (

)

0 200 400 600 800 1000 1200 1400 1600 1800-2.5

-2

-1.5

-1

-0.5

0

0.5

Coil combination

Phase c

hanges (

)

0 200 400 600 800 1000 1200 1400 1600 1800-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

Coil combination

Phase e

rror

()

(a)

(b)

(c)

one plane of

measurement

tank (simulated)

tank (measured)

tank (measured)

- tank (simulated)

Page 250: Magnetic Induction Tomography for Imaging Cerebral Stroke

229 Experiments

Figure 9.5. Phase differences caused by addition of inclusion for case-A, relative to uniform tank:

(a) simulated, (b) measured.

The phase differences due to the addition of the inclusion, = tank + inclusion - tank, are

plotted in figure 9.5 for case-A. The simulated data predicts a maximum phase

difference of 0.064 magnitude (figure 9.5a). There are a few spikes in the measured

data with much larger magnitudes (| |max = 0.67 ). These spikes are out of y-axis limits

in figure 9.5b which is set to [-0.15,0.15] for convenience of display. By closer

scrutiny of the two plots, the genuine phase differences, which can be used for

differential imaging, can be recognised in the measured data. However, the presence of

random high magnitude spikes is also significant. These can produce artefacts in the

reconstructed images. Similar plots are shown in figure 9.6, for case-B.

200 400 600 800 1000 1200 1400 1600 1800-0.15

-0.1

-0.05

0

0.05

0.1

Coil combination

Phase d

iffe

rence (

)

200 400 600 800 1000 1200 1400 1600 1800-0.15

-0.1

-0.05

0

0.05

0.1

Coil combination

Phase d

iffe

rence (

)

(a)

(b)

tank + inclusion - tank

(simulated)

tank + inclusion - tank

(measured)

Page 251: Magnetic Induction Tomography for Imaging Cerebral Stroke

230 Experiments

Figure 9.6. As figure 9.5, but for case-B.

9.5. Image reconstruction

In order to solve the forward problem, a grid size of 0.05 cm was employed in section

9.3. However, the same grid cannot be used for solving the inverse problem as it would

result in an excessively large sensitivity matrix. Furthermore, it would be time-

consumingly expensive to perform any iterative absolute imaging which requires

solving multiple forward problems for updating the sensitivity matrix.

Therefore, similar to chapter 6, a grid size of 1 cm was used for solving the inverse

problem. By using this grid size, a tank of 20 cm diameter (0.5

cm larger than the one

200 400 600 800 1000 1200 1400 1600 1800-0.15

-0.1

-0.05

0

0.05

0.1

Coil combination

Phase d

iffe

rence (

)

200 400 600 800 1000 1200 1400 1600 1800-0.15

-0.1

-0.05

0

0.05

0.1

Coil combination

Phase d

iffe

rence (

)

(a)

(b)

tank + inclusion - tank

(simulated)

tank + inclusion - tank

(measured)

Page 252: Magnetic Induction Tomography for Imaging Cerebral Stroke

231 Experiments

used in the experiments and forward simulations) was simulated for computing the

sensitivity matrix. The tank was 16 cm in height.

9.5.1. Differential imaging

For differential imaging, the sensitivity matrix was computed for a uniform conductivity

of 1 S

m

-1, and its entries were normalized to the primary signals as in (6.12). The

generalized Tikhonov method was used together with the neighbouring regularization

matrix as in (6.28). The regularization level was chosen empirically as none of the

automatic methods suggested an appropriate value for . For image representations, 8

transverse slices cutting the tank at different heights, were selected from the

reconstructed volume image (see figure 9.7).

Figure 9.7. Transverse slices cutting through the tank used for image representations.

The first column from left in figure 9.8, shows the target in case-A, when interpolated

on a 1 cm grid size. The second column represents the reconstructed images from the

noise-free simulated data ( = 1 10-8

).

As can be seen, the inclusion is clearly visible in slices 7 and 9. Two upper and lower

slices (5 and 11) also show the feature, which does not exist in the target. The likely

cause for this is the use of neighbouring matrix which does not allow sharp gradients in

the image, thereby slightly spreading the reconstructed inclusion. Also the reconstructed

inclusion is under-estimated in amplitude where the maximum magnitude of the

obtained conductivity change is 0.15 S m-1

(target is 0.205 S m-1

).

Slice 1

Slice 15

z

Page 253: Magnetic Induction Tomography for Imaging Cerebral Stroke

232 Experiments

Target

case-A

Image

(simulated data)

Image

(measured data)

1

3

5

7

9

11

13

15

Figure 9.8. Case-A: Columns from left: level of cross-section along tank, target on 1 cm grid,

differential images from simulated, measured data.

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

Page 254: Magnetic Induction Tomography for Imaging Cerebral Stroke

233 Experiments

Target

case-B

Image

(simulated data)

Image

(measured data)

1

3

5

7

9

11

13

15

Figure 9.9. As figure 9.8, but for case-B.

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

5 10 15 20

2

4

6

8

10

12

14

16

18

20 -0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

Page 255: Magnetic Induction Tomography for Imaging Cerebral Stroke

234 Experiments

The third column shows the images reconstructed from the measured data. An inclusion

can be observed in the images in the correct location (slices 7 and 9), with an even more

under-estimated magnitude ( 0.06 S

m

-1) in comparison to the second column. This is

because a higher level of regularization level, = 1 10-7

, was used in order to reduce

the artefacts.

The presence of artefacts is evident in all slices in this column; some with lower

conductivity change (slice 1), and some with raised values (slice 15). The likely

reasons for these artefacts are discussed later in this chapter.

The differential images for case-B are shown in figure 9.9.

9.5.2. Absolute imaging

Absolute imaging was performed using the iterative method described in section 6.8.3

for the uniform tank. Initially, the images were reconstructed using noise-free simulated

data (figure 9.10, upper row). The computed conductivity values were slightly higher

( 0.22 S m

-1), with lower conductivities on the periphery. This is likely due to the fact

that the signals were computed using a tank of 19.5 cm diameter, and the sensitivity

matrix was computed for a tank of 20 cm diameter. Recalling from chapter 8 for stroke

imaging, an expansion in the assumed boundaries resulted in lower conductivity values

on the head surface.

On the day of measurements, a phase noise level of 3 m was estimated in the MK2a

system. In order to see how such a noise level may affect the images, this noise was

added to the simulated phase changes, and change in the image was insignificant (see

middle row).

Absolute imaging from the measured data was not successful, and the inverse solution

diverged (the residual started increasing) after a few iterations (lowest row). As can be

seen, the image at iteration 8 is totally degraded by the presence of artefacts which are

produced because of the noisy signals. The application of various levels of

regularization and types of regularization matrix (Identity, neighbouring) did not

improve the images.

Page 256: Magnetic Induction Tomography for Imaging Cerebral Stroke

235 Experiments

Figure 9.10. Absolute imaging of uniform tank of 0.205 S m

-1 conductivity. Transverse slides cutting

halfway up the tank. Rows from top to bottom: images reconstructed from noiseless simulated data,

simulated data with 3 m additive noise, and measured data, respectively. Colour-bar in S

m

-1.

9.6. Further discussion

The discrepancy between the simulated and measured data in, for example, figure 9.4c

can have several reasons. Apart from the possible discretization error in the simulated

data, the error in the measured data may be due to the phase noise in the MK2a system,

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.05

0.1

0.15

0.2

0.25

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.05

0.1

0.15

0.2

0.25

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.05

0.1

0.15

0.2

0.25

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.05

0.1

0.15

0.2

0.25

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.05

0.1

0.15

0.2

0.25

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.05

0.1

0.15

0.2

0.25

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.05

0.1

0.15

0.2

0.25

Iteration 2 Iteration 8

5 10 15 20

2

4

6

8

10

12

14

16

18

200

0.05

0.1

0.15

0.2

0.25

Simulated

Simulated +

3 m noise

Measured

Target

Page 257: Magnetic Induction Tomography for Imaging Cerebral Stroke

236 Experiments

systematic errors, and the conductivity values. In this section, the phase changes for the

uniform tank are examined more closely. To this end, three arbitrary coil pairs (Ex1-

Se8, Ex2-Se9, Ex12-Se6) are chosen for the purpose of comparison between the

simulated and measured phase changes. Figure 9.11 shows the measured phase changes

plotted against the simulated ones, where each colour represents one pair of coils and 9

points for each colour correspond to the 9 planes of measurement.

As can be seen, for some coil pairs, the measured data is larger than the simulated,

whereas for other coil pairs the simulated data is larger. Furthermore, if one particular

coil pair is considered (e.g. the one represented by the blue colour), the measured data

for some planes of measurement is larger than the simulated one, while for other planes,

the simulated data is larger.

It is a common practice to consider a calibration factor for the measuring devices, where

the measured data are calibrated by, for example, dividing or multiplying them by a

factor. However, as discussed, coil pairs show an inconsistent behaviour for each plane

of measurement. Therefore, a single calibration factor cannot be applied to adjust the

measured signals. In the following, the effect of each potential source of error on the

signals is investigated separately.

Figure 9.11. Comparison of measured and simulated phase changes for uniform tank of 0.205 S m

-1

conductivity. Blue, red, and green colours represent Ex1-Se8, Ex2-Se9, and Ex12-Se6 coil pairs,

respectively. For each coil pair (each colour), 9 points are shown which correspond to 9 planes of

measurement for that particular coil pair. The line of identity is also drawn.

-2.5 -2 -1.5 -1 -0.5 0 0.5-2.5

-2

-1.5

-1

-0.5

0

0.5

Measured ( )

Sim

ula

ted

()

Page 258: Magnetic Induction Tomography for Imaging Cerebral Stroke

237 Experiments

9.6.1. Discretization error

Considering the grid size of 0.05 cm used for solving the forward problem, the

simulated signals are expected to have a good accuracy. Solving the FD model for a

smaller grid size was impractical as mentioned in section 9.3.

However, in order to further assure the accuracy of the FD results, the FE model was

solved for the uniform tank. Since the solution time in the FE model is long, the model

was solved for only one plane of measurement (middle plane, P5 in figure 9.3a). For a

fine mesh, comprising 65,000 second-order elements, the solution time was 6 hours.

The phase errors, difference between the FE and FD phase changes, are plotted in figure

9.12, where the y-axis limits is similar to that in figure 9.4c for the ease of comparison.

Note that the coil combinations in x-axis refer to the middle plane of measurement.

Figure 9.12. Discrepancy between the FD and FE phase changes for the uniform tank and middle

plane of measurement.

As can be seen, the phase errors are much smaller than those in figure 9.4c. The r.m.s.

phase error is 0.006 with a maximum magnitude of 0.01 . Therefore, the large

discrepancies between the measured and simulated data cannot be due to the

discretization error in the FD model.

700 750 800 850 900 950 1000 1050-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

Coil combination

Phase e

rror

()

Page 259: Magnetic Induction Tomography for Imaging Cerebral Stroke

238 Experiments

9.6.2. Phase noise

As mentioned previously, the phase noise in the hardware was estimated to be 3 m

during the measurements. This level of noise had no significant influence on the

reconstructed images in figure 9.10, middle row. In order to assess the effect of this

noise noise on the phase changes, the difference between the noise-free and noisy

simulated phase changes is plotted in figure 9.13. The discrepancy in the data is small

compared to those in figure 9.4c. The r.m.s. phase error (noise) in figure 9.13 is 0.003

with a maximum magnitude of 0.01 . This indicates that the large discrepancies

between the measured and simulated data cannot be due to the phase noise in the

hardware.

Figure 9.13. Discrepancy between the noise-free and noisy simulated phase changes for the uniform

tank.

9.6.3. Conductivity value error

The conductivity of the saline, measured at different times during the experiment, varied

from 0.2 to 0.21 S m-1

. This variation in the observed values may be due to the error in

the sampling equipment for which the accuracy was unknown. Note that a nominal

conductivity of 0.205 S m

-1 was considered and used for the forward simulations.

According to equation (2.12), for a fixed frequency and when the skin depth is larger

than the target dimensions, the phase change is linearly proportional to the conductivity.

0 200 400 600 800 1000 1200 1400 1600 1800-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

Coil combination

Phase e

rror

()

Page 260: Magnetic Induction Tomography for Imaging Cerebral Stroke

239 Experiments

Revisiting figure 5.13, the imaginary part of the phase changes is linearly proportional

to the conductivity, regardless of the value of conductivity itself (in the range < 3 S m

-1

shown in the figure). Therefore, a slightly larger or smaller conductivity will affect the

imaginary component of all the phase changes by a constant factor.

In order to take out the possible conductivity errors from the measured data, the

conductivity was adjusted as follows:

A range of conductivity values between 0.19 and 0.25 S m

-1, with a step size of

0.001 S m

-1, was considered.

The phase changes for each conductivity were obtained by scaling the data set

computed for = 0.205 S m

-1.

The phase error or the discrepancy between the measured data and each of the

obtained data sets was computed as in figure 9.4c.

The r.m.s. phase error was calculated and is plotted in figure 9.14 as a function

of conductivity.

Figure 9.14. Values of r.m.s. for the phase error as a function of conductivity; phase error is

computed as the discrepancy between the measured and scaled simulated phase changes.

The adjusted conductivity was considered to be the value for which the minimum r.m.s.

in the phase error was obtained. This corresponded to a = 0.223 S m

-1.

0.19 0.2 0.21 0.22 0.23 0.24 0.250.05

0.06

0.07

0.08

0.09

0.1

0.11

0.12

0.13

Conductivity (S/m)

R.M

.S.

= 0.205 S m-1

r.m.s. = 0.0843 a = 0.223 S m-1

r.m.s. = 0.0613

Page 261: Magnetic Induction Tomography for Imaging Cerebral Stroke

240 Experiments

The phase errors, corresponding to a, are plotted in figure 9.15; this corresponded to

the difference between the measured and adjusted simulated data. In comparison to

those shown in figure 9.4c, the phase errors now have relatively smaller magnitudes

with a maximum of 0.28 (this was 0.44 in figure 9.4c). Nevertheless, there still exist

phase errors indicating that the remaining discrepancy between the simulated and

measured data may not be due to the phase noise in the hardware or the conductivity

error, but due to systematic errors.

Figure 9.15. Discrepancy between the measured phase changes for the uniform tank and the

simulated ones after the conductivity adjustment.

9.6.4. Systematic errors

Systematic errors can arise if the boundary (i.e. position and/or shape of the tank) was

not exactly as assumed in the model, or the tank was not exactly centred within the coil

array. Furthermore, the surface of the saline in the tank may have been in error by

perhaps a few millimetres due to inaccurate filling or movement during data acquisition.

This may contribute, for example, in producing the apparently the low conductivity in

slice 1 in figure 9.8 and figure 9.9.

The vertical alignment of the tank within the sensor array for different planes of

measurement was achieved by placing wooden spacer disks beneath the tank. A total

number of 16 disks of 1 cm thickness was used to obtain the required spacing between

0 200 400 600 800 1000 1200 1400 1600 1800-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

Coil combination

Phase e

rror

()

Page 262: Magnetic Induction Tomography for Imaging Cerebral Stroke

241 Experiments

the upper and lower planes of measurement. The height of all 16 disks together was

16.2 cm, again resulting in an error of a few millimetres. A further error of about the

same magnitude could occur during the horizontal placement of the tank within the

guide as shown in figure 9.2b.

In order to simulate an error in the tank alignment within the sensor array, the tank was

displaced in x- and y-directions for 2 mm. This corresponded to the tank being 2.8 mm

off-centre within the array. The forward problem was solved for the displaced tank and

the adjusted value of saline conductivity ( a = 0.223 S m

-1).

The phase errors, being the difference between the simulated phase changes for the

displaced tank and the tank at correct position, are plotted in figure 9.16. The r.m.s.

phase error is 0.033 with a maximum magnitude of 0.11 . As can be seen, these phase

errors are comparable with those in figure 9.15 in terms of the magnitude. Therefore,

the dis-positioning of the tank can be one of the likely contributing reasons for the large

discrepancy between the simulated and measured data.

Figure 9.16. Discrepancy between the simulated phase changes for the uniform with adjusted

conductivity when placed at the centre of sensor array and when displaced for 2.8 mm.

Other potential contributing factors may be due to the inaccurate geometry of the sensor

array itself. These may include the size, positioning, and orientation of the coils, and

also the geometry of the screen in the hardware.

0 200 400 600 800 1000 1200 1400 1600 1800-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

Coil combination

Phase e

rror

()

Page 263: Magnetic Induction Tomography for Imaging Cerebral Stroke

242 Experiments

Since the data-collection hardware design was stable, with a relatively small phase noise

level of 3 m , the discrepancy between the measured and simulated data deemed to be

mainly due to a combination of the systematic errors.

The measurements reported here were the very first conducted on the MK2a data-

collection system. There is much scope for further investigation of the errors, and this is

suggested for future work in the last chapter.

End of chapter 9.

Page 264: Magnetic Induction Tomography for Imaging Cerebral Stroke

Chapter 10.

Conclusions and Future Work

10.1. Introduction

The feasibility of MIT imaging of cerebral haemorrhage was investigated through

numerical simulations. This chapter summarises the work carried out in preparation of

this thesis, and discusses the main conclusions drawn from the study. It also outlines

potential future work in a continuation of this study.

10.2. Remarks and conclusions

10.2.1. Chapter 1

The MIT imaging technique was introduced, and potential applications of MIT were

discussed. The motivation and objectives of this research were also provided.

10.2.2. Chapter 2

The physical principles of MIT imaging were described together with the governing

equations. The concept of the forward and inverse problems was introduced. The main

sub-systems of a practical MIT measurement device were outlined, and the working

principles were explained. The history of MIT studies for brain imaging, conducted by

different research groups, was also provided.

Page 265: Magnetic Induction Tomography for Imaging Cerebral Stroke

244 Conclusions and Future Work

10.2.3. Chapter 3

The anatomy of a healthy human head was briefly reviewed. Different types of stroke

pathology (ischaemic and haemorrhagic) were introduced, and further insight into the

motivation for this study was provided.

10.2.4. Chapter 4

A literature survey of the dielectric properties of the head tissues at different frequencies

was carried out. The most reliable dielectric property values available in the literature

were extracted to be employed for the simulations. The pathophysiology of the stroke

and its influence on the dielectric properties of the tissues were described, and the

applicability of MIT for stroke imaging was discussed. It was concluded that, for the

current range of operating frequencies, MIT imaging of haemorrhage may be feasible.

10.2.5. Chapter 5

Two of the most common numerical techniques, finite element (FE) and finite

difference (FD), were used for solving the discretized forward problem. The boundary

value problem for each technique was described, and its implementation for solving the

forward problem was explained.

The validity of the employed numerical techniques was examined by conducting

benchmark tests. These included performing convergence tests and comparison of the

numerical results against the analytical solutions for a simplified geometry (conducting

sphere). The numerical results were also compared with the measured practical data.

The results revealed that both methods can be reliably used for MIT simulations.

However, the application of the FD model is limited to the cases when the skin depth is

larger than the target dimensions. In terms of the solution time, the FD model performed

better and, for the examples provided, it was up to 150 times faster than the FE model.

10.2.6. Chapter 6

The discretized MIT inverse problem was outlined, and difficulties associated with

solving it were explained. The Tikhonov regularized method was initially examined for

Page 266: Magnetic Induction Tomography for Imaging Cerebral Stroke

245 Conclusions and Future Work

solving the inverse problem for geometrically simplified 3D models. Differential and

absolute images of an inclusion placed within a tank were reconstructed. Imaging

centrally located inclusions proved to be difficult, indicating that the MIT system is less

sensitive to the central regions.

The effect of the regularization matrices on the images was examined. Furthermore, the

regularization level, controlled by the regularization parameter, was found to have

profound influence on the accuracy of the inverse solution, and therefore quality of the

reconstructed images. Two automatic methods for the selection of the regularization

parameter were tested and it was shown that, for some of the presented examples, they

may result in under- or over-regularized solutions, so can be unreliable.

For non-linear iterative absolute imaging, an alternative scheme for the adjustment of

the regularization level between the iterations was employed. This included an initial

over-regularization of the inverse problem, followed by a gradual decrease in the

regularization parameter. For differential imaging, it was suggested that a subjective

assessment along with the use of automatic methods can be helpful for selecting a

proper level of regularization.

10.2.7. Chapter 7

An anatomically realistic head geometry, employed for the simulations, was described.

Three stroke lesions, of various sizes and sites within the brain, were considered. The

MIT forward problem was solved for a 16-channel annular MIT system (MK1). The

largest signals were produced by a peripherally located large stroke (50 ml), where the

smallest ones were due to a small lesion (7 ml), located deep within the brain. It was

also shown that more severe strokes (containing a higher percentage of blood within the

lesion), result in stronger MIT signals. For all strokes, the produced signals were small

(< 100 m phase difference) which indicated that sensitive and robust hardware is

required to measure them accurately.

In order to the make the simulations more realistic, phase noise at different levels was

added to the simulated data. Time- and frequency-differential imaging of the stroke

lesions was performed. For a phase noise level of 1 m , peripheral strokes could be

visualized in time-difference images. However, apart from the unrealistic noise-free

Page 267: Magnetic Induction Tomography for Imaging Cerebral Stroke

246 Conclusions and Future Work

cases, images of a deep stroke were degraded by the noise and the lesion was not

visible.

Frequency-difference imaging using the MK1 system was less successful, and only the

large peripheral stroke could be visualized as a left-right asymmetry in the images.

The amount of absorbed energy by the head was examined by computing the specific

abortion rate. It was concluded that, at present, MIT is a harmless and safe imaging

method regarding heat generation.

10.2.8. Chapter 8

Customised MIT arrays, hemispherical and annular, were investigated for stroke

imaging. For the particular arrays examined, the sensitivity matrix of the hemispherical

arrays were found to be slightly less ill-conditioned. Plots of the total sensitivity maps

revealed that for the coils placed sufficiently close to the head, a better sensitivity

distribution can be achieved, in particular, in upper parts of the head. This is because

hemispherical arrays conform more closely to the shape of the head than a conventional

annular array.

The imaging performance of the arrays was investigated by reconstructing frequency-

differential images. The true conductivity distribution of the target was assumed to be

known, and was used for adjusting the regularization level during image reconstruction.

This allowed achievement of the best possible images for each MIT array.

It was shown that making measurements in multiple planes with the annular array can

improve the reconstructed images by providing more independent data. On the other

hand, hemispherical arrays offers the possibility of using a larger number of coils,

allowing the measurements to be carried out without the need to move the head or the

array, and therefore avoiding further systematic errors.

Two types of systematic errors in the assumed a-priori boundaries of the head for

computing the sensitivity matrix were considered: displacement and scaling. It was

shown that displacement and scaling should remain less than 4 mm and 3%,

respectively. Otherwise, the artefacts produced by the errors may degrade the images.

However, the displacement errors were examined separately in the three coordinate

Page 268: Magnetic Induction Tomography for Imaging Cerebral Stroke

247 Conclusions and Future Work

directions. A simultaneous displacement in two or three directions, may lower the 4 mm

limit.

A scaling error in terms of contraction of the head boundaries produced misleading

artefacts of positive sign on the periphery, which could be mistaken for a genuine

feature due to a haemorrhage. Conversely, artefacts of negative sign were produced by

an expansion of the head which could not be mistaken as a haemorrhage.

The design study of the MK2b data-collection system was discussed. This included

examining different coil configurations and arrangements. A hemispherical array,

comprising 32 exciter coils and 14 sensor coils, was adopted as the final design based

on which the system is currently under construction.

10.2.9. Chapter 9

Practical MIT imaging using the measured data with the MK2a system was performed

for the very first time. While acceptable dime-differential images were reconstructed,

absolute imaging was less successful. The likely reasons for the observed discrepancy

between the simulated and measured MIT data were discussed. These included

discretization error, phase noise, conductivity error, and systematic errors. It was

concluded that systematic errors could be the main contributing reason to the large

discrepancies between the two data sets.

10.3. Discussion and future work

In this section, first the suggested future work that could be carried out as an immediate

follow up to this thesis is discussed. Next, potential research areas which could be

considered as long term investigations in order to improve MIT imaging in general are

described.

10.3.1. Immediate follow up research

a) In chapters 7 and 8, a single generic geometry was employed without variation

of the geometry for different subjects. It may be useful to perform the

Page 269: Magnetic Induction Tomography for Imaging Cerebral Stroke

248 Conclusions and Future Work

simulations for various subjects with different head shapes and sizes. This can

be done by extracting the head geometry from the MRI scans.

b) It is worth investigating the imaging of various types of stroke such as

subarachnoid or epidural haemorrhages.

c) More realistic simulations may also model the small substructures in the brain,

and take into account the effects of age and temperature.

d) Following the results in chapter 9 which were the very first from the MK2a

system, the electronics must be checked carefully to ensure the stability and

robustness of the hardware components.

e) The cylindrical phantoms and the measurement protocol (correct adjustment of

the phantom within the array and conductivity of the saline solution) should also

be improved in order to minimise the discrepancies between the simulated and

measured data.

f) Measurements can be repeated several times in order to acquire more data, and a

similar examination of the measured MIT signals as in chapter 9 could be

performed following the steps (d) and (e).

g) For stroke imaging, the construction of the MK2b system, designed based on the

findings in chapter 8, should be completed as the first step.

h) Based on the findings in chapter 8, it is of importance to accurately define the

boundaries of the eddy-current region (head surface) for computing the

sensitivity matrix. Therefore, immediate follow up work will include

examination of different shape measuring systems such as optical tools, and

utilizing photogrammetry software packages to determine the boundaries of the

head.

i) An accurate method must be developed for registration of the obtained boundary

in step (h) with the MK2b coil array so as to avoid systematic displacement

errors.

j) In order to validate the numerical feasibility tests, it would be timely to carry out

practical measurements using anatomically realistic phantoms representing the

head and using the MK2b system. The phantoms should be real head-shaped and

its constituents should have dielectric properties similar to those in the head

tissues. It is appreciated that it will be impractical to construct a phantom with

all the characteristics and complexity of a human head. However, it must be

noted that by being realistic, the phantoms should be multi-layer, representing

Page 270: Magnetic Induction Tomography for Imaging Cerebral Stroke

249 Conclusions and Future Work

the most important tissue types such as brain, CSF, and muscles. They might

also include real animal blood to simulate a haemorrhage. The dielectric

properties of all components of the phantoms must first be checked using

wideband impedance analyser (available at the University of Glamorgan).

k) If images, reconstructed from experiments in step (j), provide the required

confidence, an application for funding for a clinical pilot study on stroke should

be made.

10.3.2. Forward and inverse problems

Most numerical simulations of MIT, including this study, have neglected the effects of

wave propagation. When measuring small signals in biomedical MIT, propagation

delays might be significant and would appear at the sensor coil as a phase lag which

could be confused with the phase lag caused by the eddy-current signal. An analysis of

the error due to neglecting the wave propagation delays warrants a detailed numerical

investigation.

A long simulation time could jeopardise one major advantage of using MIT which is a

rapid diagnosis. On the other hand, the application of too many assumptions to simplify

the simulations could jeopardise the accuracy of the results. For the numerical

techniques and simulation environment, it would be invaluable to develop a dedicated

MIT software package which is:

comprehensive (implementing forward and inverse problems)

flexible (adjustable for particular imaging applications)

robust (employing improved iterative solvers)

time-efficient

The use of such a package would allow conducting the feasibility tests for stroke

imaging or any other application of interest.

Future research in the solution of the inverse problem may include exploring more

efficient ways of image reconstruction and, in particular, developing automatic methods

for selection of the regularization level.

Page 271: Magnetic Induction Tomography for Imaging Cerebral Stroke

250 Conclusions and Future Work

One possibility to improve the image quality for monitoring purposes, could be the use

of an individualised conductivity distribution of the head before stroke. The internal

structure of the head can be mapped from MRI or CT scans, and then incorporated into

the reconstruction algorithm (e.g. through the reference conductivity vector in

generalized the Tikhonov technique in equation (6.28)). For those tissues outside the

likely regions for bleeding (e.g. bones and muscles), the conductivity change can be

forced to be zero. This can be done by modifying the regularization matrix to highly

penalize the conductivity change for these tissues (opposite to what is done to preserve

the edges in equation (6.41)).

Even if it is impractical to map each individual's head before the application of MIT

monitoring, the use of a generic map obtained from a subject with the same head size

may be useful.

10.3.3. Practical measurements and systematic errors

With an achievable phase noise of 1 m and a low phase drift [206], it seems that the

current generation of practical measuring systems should be accurate and stable enough

for imaging peripherally located haemorrhages. Systematic errors will occur during the

data acquisition process and may include:

inaccuracies in the assumed a-priori boundaries of the eddy-current region (e.g.

size-scaling)

inaccurate registration (misalignment) of the assumed boundaries to the sensor

array (displacement and rotation)

movement of the target during measurements (displacement and rotation)

While the effect of size-scaling and displacement errors on the images was examined in

this study, further research may include investigating the patient's movement, which has

previously been reported only for simplified geometries [220].

Another source of systematic errors is the inaccuracies in geometry of the sensor array

itself; e.g. the axes of the coils in the hemispherical or annular arrays are not exactly

directed towards the centre of array. One may consider this as a fault in the hardware, or

an error in the model which does not simulate the hardware exactly as it is. Either way,

this will result in discrepancy between the predicted and measured signals. One method

Page 272: Magnetic Induction Tomography for Imaging Cerebral Stroke

251 Conclusions and Future Work

for correcting these error may be to examine the primary signals (no target present),

which depend only on the geometry of sensor array and driving parameters (current and

frequency). This may allow estimation of correct coordinates and orientations of the

coils.

10.3.4. Computational vs. Measurement Accuracy

Ideally, we would have liked to have shown that the simulations are at least as accurate

as the measurement system and data, if not more so. However, the question of accuracy

and how it is measured poses a number of interesting points.

Firstly, the accuracy of a numerical model is not governed solely by the numerical

discretization. The accuracy of a numerical simulation is in fact governed by a number

of different factors including the approximation of the mathematical model to the

physical phenomena, uncertainty in the geometry, boundary conditions and materials,

the fidelity of the numerical discretization, and the magnitude of the tolerance used in

the iterative solvers.

Secondly, an interesting question arises as to how the accuracy should be measured. In

the simulations conducted in this study, various outputs and fields were computed that

vary with position in space. When analytical solutions were available, the accuracy was

measured by computing the norm of the error in a given quantity, as was done in the

case of the convergence study presented in figure 5.4. However, analytical solutions

were not available for the complex geometries such as the head model, making such

studies impossible. On the other hand, experimental data was available for a limited

number of realistic case studies that allowed the comparison with computed quantities

at specific locations in the model, such as the comparison of the computed and

measured voltage in the sensor coils shown in figure 5.13. The problem with such

comparisons is that although the experimental data may show a close agreement with

the computed voltage, this may not necessarily mean that the other fields elsewhere in

the domain have the same level of accuracy.

The derivation of the sensitivity matrix shown in chapter 6 required that the

conductivity changes were small. The comparisons shown in figure 6.5 indicated that

the reciprocity based sensitivity can accurately predict the signal if the conductivity

Page 273: Magnetic Induction Tomography for Imaging Cerebral Stroke

252 Conclusions and Future Work

changes are small. However, for changes in the background conductivity which

exceeded 10%, the reciprocity based sensitivity became inaccurate.

Thirdly, there is a considerable level of uncertainty in the measured data obtained from

practical measurement system. For example, one known uncertainty is the SNR. In

reality, however, the overall uncertainty could be much worse than 1 m (1.74 10

-5

radians) suggested by the SNR alone due to the existence of various systematic errors

(see chapter 9). Also, the exact positioning and geometry of the coils may in practice be

slightly different to that set out in the system design and, therefore, lead to uncertainties

in the measurement system. Thus, unless the magnitude of these uncertainties is

known, it could lead to the false conclusions that the simulated model was not as

accurate as the measurement system.

Nevertheless, there is strong evidence to support the claim that the numerical

approximations of the MIT forward and inverse problems were accurate. This was

illustrated by the closeness of the agreement between the numerical and analytical

results presented in figure 5.4, and the agreement between the numerical and practical

results illustrated in figure 5.13.

For instance, the least squares fit in figure 5.13 showed that the difference between the

FE and the measured results was 0.052%, the difference between the FD and measured

results was 0.055%, whereas the allowable SNR translates to 1.74 10-3

%. An initial

observation of these results might lead to the conclusion that the results are not as

accurate as the SNR. However, we reject this conclusion on the basis that there are a

large number of uncertainties that are associated with both the simulations and the

measurements, and that for the discretizations chosen. These uncertainties dominate

over the numerical discretization error. That is to say, that for a finer discretization

(small meshing spacing and/or the use of higher order shape functions), we would not

expect the quoted differences to reduce substantially.

The comparisons presented in chapter 5 were used as a basis for choosing the

discretization for the complicated head model in later chapters. Clearly for a new

problem, the discretization required to achieve a specified level of accuracy may be

different to that required by the examples where the analytical solutions and

experimental data were available. However, for cases where no such data is available, a-

Page 274: Magnetic Induction Tomography for Imaging Cerebral Stroke

253 Conclusions and Future Work

posteriori error estimation could be used to evaluate the error in the FE discretization

and, subsequently, adaptively refine the discretization.

The topic of a-posteriori error estimation is widely established in many fields, and a

large number of estimation techniques have been developed (see e.g. [221]). An

interesting class of error estimation is that of goal orientated error estimation and

adaptivity. In this technique, the numerical discretization is adaptively refined to reduce

the error in a chosen output of interest. Such techniques have been successfully applied

to the reduction of the error associated with the scattering width output of Maxwell‟s

equations [222]. A wide range of other error estimators exist for the Maxwell problem;

e.g. the a-posteriori error estimation technique developed by Schöberl [223]. Other error

estimators have also been applied and developed in the context of the eddy-current

problems (see e.g. [224]).

To the author's knowledge, a-posteriori error estimation and adaptivity have not yet

been applied to the problems in MIT. However, an adaptive meshing approach has been

developed by Molinari [225] for EIT applications. An interesting extension of the

current project would be to incorporate a goal orientated error estimation and adaptivity

procedure for refining the discretization so as to reduce the error in the chosen outputs

of interest. In the context of MIT problem, this could be reduction in the error

associated with the voltage in the forward problem, or reduction of the error in the

sensitivity matrix in the inverse problem. Such error estimator could then be used to

subsequently adapt the discretization of the both forward and inverse problems [175].

A-posteriori error estimation can be seen as one aspect of the uncertainty quantification.

Further techniques of the uncertainty quantification could also be explored for

quantifying the uncertainties that are not associated with the numerical discretization;

e.g. mathematical modelling and measurement errors (for further details see [226]).

10.3.5. Sensor array design

Although it is stated that the MK2b system is optimized for stroke imaging, in the

author's opinion, care must be taken when using the term optimization. An optimized

design for a system is a design that offers the best possible performance regarding all

different aspects of that system. In an MIT device, for instance, this will include

hardware components comprising the electronic circuitry and sensor array, host

Page 275: Magnetic Induction Tomography for Imaging Cerebral Stroke

254 Conclusions and Future Work

computing unit and its software components comprising the image reconstruction

algorithms, and the data acquisition protocol.

Even optimizing the sensor array alone, will require exploration of all different

possibilities for the geometrical design of its sub-systems. In this study, only annular

and hemispherical arrays were examined. A sensor array comprising coils of different

sizes, with an irregular spatial positioning may generate better eddy-current patterns

throughout the head. Obviously, the possibilities can be limited by the practical

implementations.

Therefore, as a follow up to the work undertaken in this study, it might be worth

investigating the effect of a larger variety of coil configurations (size, location, and

orientation of the coils), targeting specific regions within the imaging domain. For

example, imaging of meningeal haemorrhages (subarachnoid or epidural) may be

realised by employing small sized coils located close to the interrogating region.

10.3.6. Image resolution

MIT inherently suffers from poor spatial resolution. This is due to insufficient

independent information and, more importantly, the fact that the underlying continuum

inverse problem is exponentially ill-posed. One may identify the low resolution of the

MIT imaging technique as a weakness. In this respect, it must be noted that the

motivation for using MIT is somewhat different from that of the conventional imaging

modalities such as MRI. Despite its limited resolution, MIT is employed to provide a

reliable, portable, and low-cost imaging tool. Furthermore, MIT is measuring different

physical parameters from those measured in CT or MRI, namely conductivity and

permittivity, that may even give an earlier indication of disease in some cases. In this

context, the reconstructed MIT images can be considered as a source of information

aiming to answer certain questions; in the case of stroke imaging, the presence of the

bleeding and its approximate location and volume.

However, in order to improve the resolution, a hybrid approach combining MIT with

ultrasound has been proposed by the South-Wales MIT group. This potentially provides

similar electrical conductivity information to MIT but with the spatial resolution of

ultrasound (in principle sub-mm).

Page 276: Magnetic Induction Tomography for Imaging Cerebral Stroke

255 Conclusions and Future Work

The ultrasound wave causes the charged ions within the tissue to move in presence of a

strong static magnetic field. A moving charge is subject to the Lorentz force. Positive

and negative ions move in opposite directions and this acts as the source of current. The

electric current density generated depends on the velocity of the ions (ultrasound wave),

static magnetic field strength, and density and mobility of the ions (conductivity). If the

induced current can be measured by some means outside of the object, the internal

conductivity distribution may be reconstructed.

One approach is to sense the magnetic field produced by the currents using a sensor

coil. This should provide the advantages of ultrasound, in terms of spatial resolution,

but with the additional conductivity information overlaid on the ultrasound images. The

proposed hybrid approach could be sensitive just to the small region of tissue excited by

the ultrasound pulse and, therefore, may be a much better posed problem than MIT.

The numerical techniques employed in this study could be used as a basis for

preliminary tests. The Lorentz force current multi-poles produced by the ultrasound

pulse and also static magnetic field throughout the domain of interest may be computed

by the addition of extra terms to the boundary value problems presented in this thesis.

Therefore, one interesting follow up study could be to examine the feasibility of

employing this novel approach.

End of chapter 10.

Page 277: Magnetic Induction Tomography for Imaging Cerebral Stroke

Part V

Appendices and Bibliography

Page 278: Magnetic Induction Tomography for Imaging Cerebral Stroke

Appendix A. System Description

A.1. MK1 system

The MK1 data-collection system consists of a circular array of 16 coil modules

(channels), switchable between exciter and sensor modes (figure a.1). The object to be

imaged is placed inside the coil array. The array is enclosed within an aluminium

electromagnetic screen which is 350 mm in diameter and 250 mm in height. The coils

are wound on Perspex formers of 50 mm diameter, each of 2 turns, mounted inside the

electromagnetic screen. The standoff distance of the coils from the screen being 30 mm

for the exciters and 40 mm for the sensors (figure A.1). The coils are placed halfway up

the screen, 125 mm height from the base. Sixteen corresponding exciter and sensor

circuit modules are housed in metal enclosures attached to the outside of the screen. The

electronic components have been described in more detail elsewhere [33].

Figure A.1. An image of frond-end of the 16-channel MK1 system [33].

Page 279: Magnetic Induction Tomography for Imaging Cerebral Stroke

258 Appendix A. System Description

Figure A.2. Diagram of the MK1 configuration; dimensions in mm: (a) the coil array inside the

electromagnetic screen; labels represent the coil numbering, (b) detail of an individual coil module;

the coil windings are shown in black (sensor) and red (exciter).

Activating the exciter coils in turn and measuring the signals in the sensors, provides

256 coil combinations. Numbering of the coils is shown in figure A.2a. Throughout this

thesis, the exciter and sensor coils will be referred to as Ex and Se; e.g. exciter coil 1 is

labelled as Ex1.

The exciter coil current is 141 mA in amplitude (100 mA r.m.s.). The system operates at

a frequency of 10 MHz. The averaged phase noise across all channel combinations is 17

m . For simplicity, the simulations described in this thesis have all assumed coils of just

1 turn and an excitation current of 1 A in amplitude. This simplification has no effect on

the phase changes; the only implication is for the heat generation in the target (head),

which is discussed in chapter 7.

A.2. MK2a system

The MK2a data-collection system consists of two circular arrays of 14 coil modules,

one of exciters and the other of sensors (figure A.3). The arrays are enclosed within an

aluminium electromagnetic screen, which is 350 mm in diameter and 300 mm in height.

The centres of the exciter and sensor coils lie on planes 40 mm above and below the

centre of the screen (figure A.4a). The coils are wound on Perspex formers of 44 mm

diameter, each of 2 turns, mounted inside the electromagnetic screen. The standoff

distance of the coils from the screen is 40 mm for all coils (figure A.4b).

(a) (b)

13 5

9

1

Page 280: Magnetic Induction Tomography for Imaging Cerebral Stroke

259 Appendix A. System Description

The MK2a system is capable of operating at multiple frequencies in the range of 0.5-14

MHz. The system achieves a measurement phase noise of less than 1 m for a time

constant of 1s for frequencies in the range of 1-14 MHz [206]. Similar to MK1, the

simulation for the MK2a system considered coils of 1 turn. The MK2a system provides

196 coil combinations for each set of the measurements, i.e. one plane of measurement.

Figure A.3. An image of frond-end of the 14-channel MK2a system.

Figure A.4. Geometrical specifications of the system; dimensions in mm: (a) exciter and sensor coils

sit on red and blue broken lines, respectively, (b) details of an individual module.

End of Appendix A.

350

300

190

110

40

44

(b) (a)

Page 281: Magnetic Induction Tomography for Imaging Cerebral Stroke

Appendix B. Dielectric Theory and

Terminology

In a polarisable material such as biological tissue, the polarization does not respond

instantaneously to an applied electric field as the rates at which polarizations can occur

are limited. Therefore, as the frequency of the applied electric field is increased, some

polarizations will no longer be able to attain their d.c. or low frequency values.

There are a number of different dielectric mechanisms, connected to the way a medium

reacts to the applied field. Each dielectric mechanism is centred around its characteristic

frequency, which is the reciprocal of the characteristic time1 of the process. According

to Pethig [227], the slowest polarization mechanism is often the dipolar orientation of

macromolecules and this is usually the first polarization term to disappear as the

frequency is increased. The fall of the polarizability, with consequent reduction of

permittivity and the occurrence of energy absorption, is referred to dielectric relaxation

or dispersion [227].

The frequency at which this fall in dipole polarization occurs can vary from very low

frequencies of the order 1 Hz for large macromolecules to frequencies up to 1012

Hz for

small molecules. Atomic and electronic polarizations occur at frequency comparable

with the natural frequencies of vibration of the atoms in a molecule ( 1014

Hz) and

corresponding to electronic transitions between different energy levels in the atom

( 1015

Hz), respectively.

1 Characteristic or relaxation time refers to the time needed for a system to relax under external stimuli.

Page 282: Magnetic Induction Tomography for Imaging Cerebral Stroke

261 Appendix B. Dielectric Theory and Terminology

In order to derive a mathematical model for the relaxation process, the constitutive

relationship (2.6) can be re-written as:

(B.1)

where the polarization P is defined as the induced dipole moment per unit volume. The

electric field and polarization are related by the dielectric susceptibility, , as P = 0 E.

The polarization P is composed of two parts P1 and P2, where P1 arises from atomic and

electronic polarizations and P2 arises from the much slower process of dipolar

reorientation. For the frequency range of interest in MIT applications (< 10

8 Hz), it can

be assumed that P1 responds instantly to the applied field E, and has the constant value

P1 = 1E. The part P2 lags behind E in such a way that at any moment, P2 approaches its

final value 2E at a rate of:

(B.2)

where is the characteristic relaxation time. This shows that the polarization approaches

its final value, in response or relaxation mode, exponentially with the time constant .

By integration of (B.2), the total polarization for a harmonic field can be expressed in

the frequency domain as [227]:

(B.3)

From (B.1) and (B.3), the permittivity can be treated as a complex function of the

frequency of the applied field:

(B.4)

where is the permittivity measured at a sufficiently high frequency for the dipolar

polarization to have been disappeared and s is the limiting low frequency permittivity.

The real and imaginary parts of the permittivity are given by:

Page 283: Magnetic Induction Tomography for Imaging Cerebral Stroke

262 Appendix B. Dielectric Theory and Terminology

(B.5)

Here, ' is related to the stored energy (a measure of the induced polarization per unit

field) and '' is related to the dissipation of energy within the medium (out-of-phase loss

factor). These equations are commonly known as the Debye dispersion formulas which

describe the relaxation for ideal systems in terms of permittivity as a function of

frequency; it is assumed that there is only a single orientational process and the

equilibrium is attained exponentially with time. It is often useful to define the complex

conductivity as in-phase or lossy ( ') and capacitive ( '') components:

(B.6)

where dc is the static ionic conductivity and ac is the conductivity increase due to the

dispersion. The total conductivity is thus made of two terms corresponding to the

residual static conductivity and polarization losses. Figure B.1 illustrates the idealised

variation in the permittivity, loss factor, and conductivity with frequency for a Debye

single dispersion relaxation, plotted based on the equations (B.5) and (B.6). Note that

the relaxation time corresponds to a relaxation frequency fr = 1/2 . At the relaxation

frequency, the permittivity is halfway between its limiting values and the loss factor at

its highest.

Figure B.1. Normalized permittivity ( ' - )/( s - ), loss factor ''/( s - ), and conductivity

0 ''/( s - ) for a single time constant relaxation plotted against f/fr (after Barnes and Greenebaum [91]).

Page 284: Magnetic Induction Tomography for Imaging Cerebral Stroke

263 Appendix B. Dielectric Theory and Terminology

Biological tissue is a heterogeneous material containing water, dissolved organic

molecules, macromolecules, ions, and insoluble matter. The presence of ions plays an

important role in the interaction with an electric field, providing means for ionic

conduction and polarization effects. Ionic charge drift creates conduction currents and

also initiates polarization mechanisms through charge accumulation at structural

interfaces, which occur at various organizational levels. Their dielectric properties will

thus reflect contributions to the polarization from both structure and composition.

The complexity of both the structure and the composition of biological material is such

that each dispersion region may be broadened by multiple contributions to it, and

biological materials exhibit a certain asymmetrical distribution of relaxation times. The

broadening of the dispersion can be compensated empirically by introducing a

distribution parameter. This results in an alternative to the Debye equation known as

Cole-Cole equation [91]:

(B.7)

where the distribution parameter, , is a measure of the broadening of the dispersion.

Furthermore, biological materials rarely show a single time constant Debye response.

Therefore, the complex permittivity is usually a complicated function of frequency; a

superimposed description of dispersion phenomena occurring at multiple frequencies.

Dispersion data in a biological material is based on the electrical examination of the

material as a function of frequency, i.e. dielectric spectroscopy. The relaxation

mechanisms can be divided in three groups of -, -, and -dispersions; each relating to

cell membranes, organelles inside cells, double layer counter-ion relaxation,

electrokinetic effects, etc [113]:

The - or low-frequency dispersion (typically below 1 kHz) is characterized by

very high permittivity values and can be ascribed, at least partially, to counter-

ion diffusion effects. Other mechanisms, e.g. interactions in the vicinity of the

cell membrane, may contribute to -dispersion.

The -dispersion occurs at intermediate frequencies (in the hundreds of kilohertz

region) and originates mostly from the polarisation of the cellular membranes

and those of membrane-bound intra-cellular bodies. Cellular membranes act as

barriers to the flow of ions between the intra- and extra-cellular media. Any

Page 285: Magnetic Induction Tomography for Imaging Cerebral Stroke

264 Appendix B. Dielectric Theory and Terminology

damage to the cell membrane changes the features of the -dispersion. Other

contributions to the -dispersion come from the polarization of protein and other

organic macromolecules.

The -dispersion is due to the dipolar polarization of tissue water at frequencies

in excess of a few hundred MHz, where the response of tissue water is

dominant.

Figure B.2. Dispersion regions , , and (after Grimnes and Martinsen [113]).

Figure B.2 shows that the permittivity in biological materials typically diminishes with

increasing frequency. The dispersion regions are here shown as originating from clearly

separated Cole–Cole dispersions. The dielectric spectrum of a biological tissue may be

expressed as the summation of multiple (n) Cole–Cole dispersions with the addition of

the static conductivity term [91]:

(B.8)

Where adequate experimental data is available, and with a choice of parameters

appropriate to each tissue, equation (B.8) can be used to predict the dielectric behaviour

over the desired frequency range.

End of Appendix B.

Page 286: Magnetic Induction Tomography for Imaging Cerebral Stroke

Appendix C. Supplements to chapter 6

C.1. GCV method

The GCV method works in the basis of 'leave-one-out' cross validation [176]. It is based

on the philosophy that if an arbitrary element bi of the vector b is left out, then the

corresponding regularized solution should predict this observation well, and the choice

of regularization parameter should be independent of an orthogonal transformation of b.

In the context of MIT, the models are obtained by leaving one of the m data points,

corresponding to one coil combination, from the fitting process in the LS solution.

Consider the modified Tikhonov regularization problem at which a data point bk is

ignored from the measurement vector b:

(C.1)

The solution to this problem is called xk, which would accurately predict the missing

data bk. A regularization parameter, , is sought to minimize the predictive errors for all

k:

(C.2)

Computing G( ) involves solving m problem of this form. In order to simplify this

computation, consider the definition:

Page 287: Magnetic Induction Tomography for Imaging Cerebral Stroke

266 Appendix C. Supplements to chapter 6

(C.3)

The solution xk

, therefore, solves:

(C.4)

This is known as the 'leave-one-out lemma', which gives:

(C.5)

where is B defined in equation (6.31). It is straightforward to derive:

(C.6)

where C = AB . Subtracting both sides of the equation from 1 gives:

(C.7)

Considering:

(C.8)

the equation (C.7) can then be simplified as:

(C.9)

By substituting (C.9) into (C.2) , the function can be expressed by:

(C.10)

The equation (C.10) can be further simplified by replacing (C )k,k in the sum with the

average value 1/m trace(C ). Therefore, the GCV methods aims at minimizing:

(C.11)

which will be minimised at the same as that in (6.34). For further details see [176].

Page 288: Magnetic Induction Tomography for Imaging Cerebral Stroke

267 Appendix C. Supplements to chapter 6

C.2. Steepest descent

The steepest descent method is generally used to find the local minimum of a nonlinear

function using its gradient. The search direction used in this method is the negative of

the gradient vector at every iteration point. It requires an initial estimated solution 0. In

the searching process, at the k-th iteration, k is replaced by k+1, which is a better

estimate of the solution.

The name of steepest descent method is earned because it uses the steepest descent

direction, - f( ), as the search direction, when the goal is to minimize f( ) = ||A -

Vm||2. If A is available, then we have:

(C.12)

Therefore, moving in the direction opposite to the gradient of f( ), the iteration

procedure can be written as:

(C.13)

where the positive scalar αk is the optimal step-length [190]. At each iteration, the

search direction and the optimal step-length in the search direction are updated. The

detailed procedure for finding the optimum step-length can be found in [166,170,228].

C.3. Landweber method

If the step-length in the steepest descent method is chosen to be fixed, then one obtains

the Landweber method expressed by [180]:

(C.14)

The Landweber method may require many iteration steps before reaching a minimum

point. It is an easy to implement algorithm, and has been employed for MIT imaging

(e.g. [229]). The modified types of the Landweber method can be found in [230].

Page 289: Magnetic Induction Tomography for Imaging Cerebral Stroke

268 Appendix C. Supplements to chapter 6

C.4. Conjugate gradient method

A reduction of the iteration steps may be achieved for the steepest descent method, if

instead of (C.13), one uses an iteration procedure, where the search direction is different

from the gradient direction. In the conjugate gradient method, one can obtain a set of

linearly independent conjugate gradient directions at each iteration from an

orthogonalization of the successive gradients [166]. For quadratic functions, the

conjugate gradient method has been proved to converge to the unique global minimum,

by moving along such successive non-interfering directions [228]. Implementation of

the conjugate gradient method has been suggested for EIT technique [173]. However, its

convergence, when applied to non-linear ill-posed problems has not been proved [230].

C.5. ART and SIRT

The algebraic reconstruction technique (ART) and simultaneous iterative reconstruction

technique (SIRT) are the variants of the Kaczmarz‟s algorithm [176]. These algorithms

were originally developed for tomographic applications, and have been applied to MIT

technique (e.g. [35]). ART is a version of Kaczmarz‟s algorithm that has been modified

especially for the tomographic reconstruction problem. It uses only one set of projection

data in each iteration step. The ART iterative formula for solving the system A = Vm

is:

(C.15)

where ak+1 is the (k+1)-th row vector of the sensitivity matrix A, and Vk+1 is the (k+1)-th

element of the vector voltage Vm. Because only one set of measurement data is used in

each iteration step, ART may not converge if the measurement contains a significant

error [190]. SIRT is a variation on ART which overcomes this disadvantage. Assuming

that the sensitivity matrix is composed of m rows, then m vectors can be obtained

according to equation (C.15) using ART. The conductivity is updated using the average

of those m vectors and the iterative formula of the SIRT algorithm is:

(C.16)

Page 290: Magnetic Induction Tomography for Imaging Cerebral Stroke

269 Appendix C. Supplements to chapter 6

where is the relaxation factor, and diag(AAT) is a vector composed of diagonal

components of AAT where the division means that each numerator is divided by the

corresponding denominator. If necessary, can be replaced by a vector; this means that

the m vectors based on ART are not treated equally when trying to obtain the average

vector [190]. SIRT can give slightly better images than ART, at the expense of being a

slightly slower algorithm [176]. Note that SIRT formula in (C.16) is similar to the

Landweber method in (C.14) with a re-scaled matrix.

C.6. Noise

Throughout this study, Gaussian pseudo-random noise is used for the simulations.

When examining the tolerable noise level for the image reconstruction, several draws of

randomly generated noise, but with the same standard deviation, were employed. The

following is a MATLAB function for generating noise based on equation (2.13):

function Vnoisy = addNoise(V,SNRdB0) % SNRdB0 is the noise level in dB % V is the column vector noise-free voltage % randn(n,1) : returns n values with Standard deviation=1

n = length(V); Vpower = V'*V/n; % or Vpower = mean(V.^2); SNR0 = 10^(SNRdB0/10); Npower0 = Vpower/SNR0; Noise = sqrt(Npower0)*randn(n,1); Vnoisy = V + Noise;

If the data set is in terms of phase changes (i.e. voltage normalized to the primary

signal), the noise is generated with a specified standard deviation in milli-degree:

function PHInoisy = addNoise(PHI,SNR0) % SNR0 is the noise level in milli-degree % PHI is the column vector noise-free phase change % randn(n,1) : returns n values with Standard deviation=1

n = length(PHI);

PHInoisy = PHI + SNR0*(pi/180000)*randn(n,1);

End of Appendix C.

Page 291: Magnetic Induction Tomography for Imaging Cerebral Stroke

References

[1] H Griffiths, "Magnetic induction tomography," in Electrical Impedance Tomography - Methods,

History and Applications., D S Holder, Ed.: IOP Publishing, 2005, ch. 8, pp. 213-38.

[2] W R Lionheart, "EIT reconstruction algorithms: pitfalls, challenges and recent developments,"

Physiological Measurement, vol. 25, pp. 125-142, 2004.

[3] W Daily, A Ramirez, A Binley, and D LeBrecque, "Electrical resistance tomography," The Leading

Edge, vol. 23, pp. 438-442, 2004.

[4] Q Marashdeh, L S Fan, and W Warsito, "Electrical Capacitance Tomography - A Perspective," Industrial & Engineering Chemistry Research, vol. 47, pp. 3708-3719, 2008.

[5] A Binley and A Kemna, "DC Resistivity and Induced Polarization Methods," in Hydrogeophysics, Y Rubin and S S Hubbard, Eds.: Springer, 2005, pp. 129-56.

[6] M M Rezos, J J Schultz, and S A Smith, "Controlled research utilizing a basic all-metal detector in the search for buried firearms and miscellaneous weapons," Forensic Science International, vol.

195, pp. 121-7, 2010.

[7] E P Lopes, "Electromagnetic geophysical tomography," IEE Proceedings F, Radar and Signal

Processing, vol. 139, pp. 27-35, 1992.

[8] H Griffiths, W Gough, S Watson, and R J Williams, "Residual capacitive coupling and the

measurement of permittivity in magnetic induction tomography," Physiological Measurement, vol.

28, pp. 301-311, 2007.

[9] M Soleimani, "Simultaneous Reconstruction of Permeability and Conductivity in Magnetic

Induction Tomography," Waves and Applications, vol. 23, pp. 785-98, 2009.

[10] M Soleimani, Image and Shape Reconstruction Methods in Magnetic Induction and Electrical

Impedance Tomography, PhD Thesis.: The University of Manchester, 2005.

[11] R Binns, A R Lyons, A J Peyton, and W D Pritchard, "Imaging molten steel flow profiles,"

Measurement Science and Technology, vol. 12, pp. 1132-1138, 2001.

[12] T Takagi, H Huang, H Fukutomi, and J Tani, "Numerical evaluation of correlation between crack

size and eddycurrent testing signal by a very fast simulator," IEEE Transactions on Magnetics, vol.

34, pp. 2581-2584, 1998.

Page 292: Magnetic Induction Tomography for Imaging Cerebral Stroke

271 <References

[13] EP/E009697/1 EPSRC. Imaging low-conductivity materials in Magnetic Induction Tomography -

LCOMIT. [Online]. http://gow.epsrc.ac.uk/ViewGrant.aspx?GrantRef=EP/E009697/1 (year accessed: 2010)

[14] EP/E009158/1 EPSRC. Imaging low-conductivity materials in Magnetic Induction Tomography - LCOMIT. [Online]. http://gow.epsrc.ac.uk/ViewGrant.aspx?GrantRef=EP/E009158/1 (year

accessed: 2010)

[15] EP/E009832/1 EPSRC. Imaging low-conductivity materials in Magnetic Induction Tomography -

LCOMIT. [Online]. http://gow.epsrc.ac.uk/ViewGrant.aspx?GrantRef=EP/E009832/1 (year

accessed: 2010)

[16] Philips. Philips Research Aachen. [Online].

http://www.research.philips.com/profile/locations/route-de.html (year accessed: 2010)

[17] P Monk, Finite Element Methods For Maxwell's Equations.: Oxford University Press, 2003.

[18] H Ammari, A Buffa, and J.-C Nédélec, "A justification of eddy currents model for the Maxwell

equations," SIAM Journal on Applied Mathematics, vol. 60, pp. 1805-23, 2000.

[19] C J Hellier, Handbook of Nondestructive Evaluation.: McGRAW-HILL, 2003.

[20] A Bossavit, Computational Electromagnetism: Variational Formulations, Complementarity, Edge Elements.: Academic Press, 1997.

[21] K Schmidt, O Sterz, and R Hiptmair, "Estimating the eddy-current modelling error," IEEE Transactions on Magnetics, vol. 44, pp. 686-9.

[22] M Soleimani, W R Lionheart, A J Peyton, X D Ma, and S R Higson, "A three-dimensional inverse finite-element method applied to experimental eddy-current imaging data," IEEE Transactions on

Magnetics, vol. 42, pp. 1560-7, 2006.

[23] J D Jackson, Classical Electrodynamics, 3rd ed.: John Wiley & Sons, 1999.

[24] B Dekdouk, D W Armitage, C Ktistis, M Zolgharni, and A Peyton, "A feasibility study on the

delectability of edema using magnetic induction tomography using an analytical model," in Proc.

IFMBE, vol. 22, 2008, pp. 736-739.

[25] T P Theodoulidis, N V Kantartzis, and T D Tsiboukis, "Analytical and numerical solution of the eddy-current problem inspherical coordinates based on the second-order vector

potentialformulation," IEEE Transactions on Magnetics, vol. 33, pp. 2461-2472, 1997.

[26] G Mrozynski and E Baum, "Analytical determination of eddy currents in a hollow sphereexcited by

an arbitrary dipole," IEEE Transactions on Magnetics, vol. 34, pp. 3822-3829, 1998.

[27] A A Kolyshkin and R Vaillancourt, "Series solution of an eddy-current problem for a sphere

withvarying conductivity and permeability profiles," IEEE Transactions on Magnetics, vol. 35, pp.

4445-4451, 1999.

[28] A Nakazawa, W Yang, and A J Peyton, "An analytical approach to obtaining 3D sensitivity maps

for electro-magnetic tomography," International Journal of Systems Science, vol. 2, pp. 607-619,

2006.

[29] H Griffiths, W R Stewart, and W Gough, "Magnetic induction tomography: a measuring system for

biological tissues," in Annals of the New York Academy of Sciences, vol. 873, 1999, pp. 335-45.

[30] A J Peyton, M S Beck, A R Borges, and J E Oliveira, "Development of Electromagnetic

Tomography (EMT) for Industrial Applications. Part 1: Sensor Design and Instrumentation," in 1st

Page 293: Magnetic Induction Tomography for Imaging Cerebral Stroke

272 <References

World Congress on Industrial Process Tomography, Buxton, 1999, pp. 306-312.

[31] Z Z Yu, P F Worthington, S Stone, and A J Peyton, "Electromagnetic screening of inductive

tomography sensors," in European Concerted Action on Process Tomography, Bergen, 1995, pp.

300-310.

[32] A J Peyton, S Watson, R J Williams, H Griffiths, and W Gough, "Characterising the Effects of the

External Electromagnetic Shield on a Magnetic Induction Tomography Sensor," in 3rd world congress on industrial process tomography, Banff, Canada, 2003, pp. 352-357.

[33] S Watson, R J Williams, W Gough, and H Griffiths, "A magnetic induction tomography system for samples with conductivities below 10 S/m," Measurement Science and Technology, vol. 19, p.

045501 (11pp), 2008.

[34] M Vauhkonen, M Hamsch, and C H Igney, "A measurement system and image reconstruction in

magnetic induction tomography," Physiological Measurement, vol. 29, pp. 445-454, 2008.

[35] X Ma, A J Peyton, S R Higson, A Lyons, and S J Dickinson, "Hardware and software design for an

electromagnetic induction tomography (EMT) system for high contrast metal process applications,"

Measurement Science and Technology, vol. 17, pp. 111–118, 2006.

[36] J Rosell-Ferrer, R Merwa, P Brunner, and H Scharfetter, "A multifrequency magnetic induction

tomography system using planar gradiometers: data collection and calibration," Physiological

Measurement, vol. 27, pp. 271-280, 2006.

[37] D Gursoy and H Scharfetter, "Feasibility of lung imaging using magnetic induction tomography,"

SFB-Report No. 2009–013, 2009.

[38] D Goss, R O Mackin, E Crescenzo, H S Tapp, and A J Peyton, "Understanding the Coupling

Mechanisms in High Frequency EMT," in 3rd world congress on industrial process tomography,

Banff, Canada, 2003, pp. 364-369.

[39] S Watson, R J Williams, and H Griffiths, "Frequency downconversion and phase noise in MIT,"

Physiological Measurement, vol. 23, pp. 189-194, 2002.

[40] H Griffiths, "Magnetic induction tomography," Measurement Science and Technology, vol. 12, pp.

1126-31, 2001.

[41] P P Tarjan and R McFee, "Electrodeless measurements of the effective resistivity of the human

torso and head by magnetic induction," IEEE Transactions on Biomedical Engineering, vol. 15, pp.

266-278, 1968.

[42] J Netz, E Forner, and S Haagemann, "Contactless impedance measurement by magnetic induction -

a possible method for investigation of brain impedance," Physiological Measurement, vol. 14, pp.

463-471, 1993.

[43] S Al-Zeibak and N H Saunders, "A fesibility study of in-vivo electromagnetic imaging," Physics in

Medicine and Biology, vol. 38, pp. 151-60, 1993.

[44] R Merwa, K Hollaus, O Biro, and H Scharfetter, "Detection of brain oedema using magnetic

induction tomography: a feasibility study of the likely sensitivity and detectability," Physiological

Measurement, vol. 25, pp. 347-54, 2004.

[45] R Merwa and H Scharfetter, "Magnetic Induction Tomography: A feasibility study of brain oedema

detection using a finite lement human head model," in IFMBE, vol. 17, 2007, pp. 480-3.

[46] C A Gonzalez and B Rubinsky, "A theoretical study on magnetic induction frequency dependence

Page 294: Magnetic Induction Tomography for Imaging Cerebral Stroke

273 <References

of phase shift in oedema and haematoma," Physiological Measurement, vol. 27, pp. 829-38, 2006.

[47] C A Gonzalez and B Rubinsky, "The detection of brain oedema with frequency-dependent phase

shift electromagnetic induction," Physiological Measurement, vol. 27, pp. 539-52, 2006.

[48] C A Gonzalez, R Rojas, C Villanueva, and B Rubinsky, "Inductive Phase Shift Spectroscopy for

Volumetric Brain Edema Detection: An Experimental Simulation," in Proceedings of the 29th

Annual International Conference of the IEEE EMBS, Lyon, 2007, pp. 2346-9.

[49] C A Gonzalez, R Rojas, and B Rubinsky, "Circular and Magnetron Inductor/Sensor Coils to Detect

Volumetric Brain Edema by Inductive Phase Shift Spectroscopy: A Sensitivity Simulation Study," in IFMBE, vol. 17, Graz, 2007, pp. 315-9.

[50] R Rojas, B Rubinsky, and C A Gonzalez, "The effect of brain hematoma location on volumetric inductive phase shift spectroscopy of the brain with circular and magnetron sensor coils: a

numerical simulation study," Physiological Measurement, vol. 29, pp. 255-66, 2008.

[51] C A Gonzalez, C Villanueva, C Vera, O Flores, and R D Reyes, "The detection of brain ischaemia

in rats by inductive phase shift spectroscopy," Physiological Measurement, vol. 30, pp. 809-19,

2009.

[52] Q Ming-Xin, J Li-Cheng, W Hai-Bing, and L Hua, "Simulation study of detecting conductivity of

brain tissues with magnetic induction based on FDTD," in 27th Annual IEEE Conference

Engineering in Medicine and Biology, Shanghai, 2005, pp. 4369-72.

[53] Q Ming-Xin, Y Wang, X Hu, M Jiao, and W Hai-Bing, "Study of MIT Phase Sensitivity for

Detecting a Brain Edema Based on FDTD Method," in ICBBE, 2007, pp. 660-3.

[54] Z Xu, H Luo, W He, C He, and X Song, "A multi-channel magnetic induction tomography

measurement system for human brain model imaging," Physiological Measurement, vol. 30, pp.

175-86, 2009.

[55] A V Korjenevsky. [Online]. http://www.cplire.ru/html/tomo/mitimage.html (year accessed: 2010)

[56] M Vauhkonen, M Yan, D Chen, M Hamsch, and C H Igney, "Phase changes in magnetic induction

tomography caused by different types of inhomogeneities," in EIT conference, Dartmouth, 2008.

[57] R Eichardt, C H Igney, J Kahlert, M Hamsch, and M Vauhkonen, "Sensitivity Comparisons of

Cylindrical and Hemi-spherical Coil Setups for Magnetic Induction Tomography," in IFMBE,

2009, pp. 269-72.

[58] S Watson, A Morris, R J Williams, W Gough, and H Griffiths, "The Cardiff magnetic induction

tomography system," , 2002, p. p116.

[59] M Zolgharni, P D Ledger, and H Griffiths, "Forward modelling of magnetic induction tomography:

a sensitivity study for detecting haemorrhagic cerebral stroke," Medical & Biological Engineering

& Computing, vol. 47, pp. 1301-1313, 2009.

[60] M Zolgharni, P D Ledger, D W Armitage, D S Holder, and H Griffiths, "Imaging cerebral

haemorrhage with magnetic induction tomography: numerical modelling," Physiological

Measurement, vol. 30, pp. S187-S200, 2009.

[61] M Zolgharni, H Griffiths, and D S Holder, "Imaging haemorrhagic cerebral stroke by frequency-

difference magnetic induction tomography: numerical modelling," in Proc. IFMBE, vol. 22, 2008,

pp. 2464-2467.

[62] M Zolgharni, P D Ledger, and H Griffiths, "Frequency-difference MIT imaging of cerebral

haemorrhage with a hemispherical coil array: numerical modelling," Physiological Measurement,

Page 295: Magnetic Induction Tomography for Imaging Cerebral Stroke

274 <References

pp. in-press, 2010.

[63] H Griffiths, M Zolgharni, S Watson, and P D Ledger, "The Cardiff Mk2b MIT head array:

optimising the coil configuration," in EIT2010, Gainesville, FL, 2010, p. submitted.

[64] O Faiz and D Moffat, Anatomy at a Glance.: Blackwell Science Ltd, 2002.

[65] S S Mader, Understanding Human Anatomy and Physiology, 5th ed.: McGraw-Hill, 2004.

[66] K S Saladin, Anatomy & Physiology: The Unity of Form and Function, 3rd ed.: McGraw-Hill,

2003.

[67] M R Villarreal. (2007) Wikimedia Commons. [Online].

http://commons.wikimedia.org/wiki/File:Human_skull_side_simplified_(bones).svg (year

accessed: 2010)

[68] SEER Training Modules. [Online]. http://training.seer.cancer.gov/anatomy/skeletal/tissue.html

(year accessed: 2010)

[69] S McCann and E Wise, Kaplan Anatomy Coloring Book.: Kaplan Publishing, 2005.

[70] V D Graaff, Human Anatomy.: McGraw−Hill, 2001.

[71] R Seeley, Anatomy and Physiology, 6th ed.: McGraw-Hill, 2003.

[72] D Shier, J Butler, and R Lewis, Hole's Human Anatomy and Physiology, 9th ed.: McGraw−Hill,

2001.

[73] V C. Scanlon and T Sanders, Essentials of Anatomy and Physiology, 5th ed.: F. A. Davis Company,

2007.

[74] L Horesh, "Some Novel Approaches in Modelling and Image Reconstruction for Multi-Frequency

Electrical Impedance Tomography of the Human Brain," 2006.

[75] J Mant, D Wade, and S Winner, "Wikipedia," in Health care needs and assessment: stroke in

Health Care Needs Assessment: The Epidemiology Based Needs Assessment Reviews.: Radcliffe

Medical Press, Oxford, 2004.

[76] National clinical guidelines for stroke.: Royal College of Physicians of London, 2004.

[77] P J Broderick and D J Santilli, Prevention and treatment of ischemic stroke.: American Academy of

Family Physicians, 1998.

[78] C H Sherry and S E Eric, Intravenous thrombolysis, in Acute ischemic stroke: An Evidence-based

Approach, D M Greer, Ed.: John Wiley & Sons, 2007.

[79] L R Caplan, Stroke.: American Academy of Neurology, 2006.

[80] L R Caplan, Caplan’s stroke : a clinical approach, 4th ed.: Saunders Elsevier, 2009.

[81] I E Silverman and M M Rymer, Ischemic Stroke.: Clinical Publishing, 2009.

[82] Neurologychannel (2009). [Online]. http://www.neurologychannel.com/stroke/diagnosis.shtml

(year accessed: 2010)

[83] J A Chalela, C S Kidwell, and L M Nentwich, "Magnetic resonance imaging and computed

tomography in emergency assessment of patients with suspected acute stroke: a prospective

Page 296: Magnetic Induction Tomography for Imaging Cerebral Stroke

275 <References

comparison," Lancet, vol. 369, pp. 293-8, 2007.

[84] A P Barber and S M Davis, Limitations of current brain imaging modalities in stroke, in Magnetic

Resonance Imaging in Stroke, edited by Stephen Davis, Marc Fisher, and Steven Warach.:

Cambridge University Press, 2003.

[85] P Harold and J Adams, Treatment of Patients with Acute Ischemic Stroke, in A Primer on Stroke

Prevention Treatment: An Overview Based on AHA/ASA Guidelines, edited by Larry B. Goldstein.: John Wiley & Sons, 2009.

[86] D Miklavcic, N Pavselj, and F F Hart, "Electrical properties of tissues," Wiley Encyclopedia of Biomedical Engineering, vol. 6, pp. 3578-89, 2006.

[87] C Gabriel, S Gabriel, and E Corthout, "The dielectric properties of biological tissues: I. Literature survey," Physics in Medicine and Biology, vol. 41, pp. 2231-2249, 1996.

[88] Dielectric Properties of Body Tissues. [Online]. http://niremf.ifac.cnr.it/tissprop/ (year accessed: 2010)

[89] S Gabriel, R W Lau, and C Gabriel, "The dielectric properties of biological tissues: II. Measurements in the frequency range 10 Hz to 20 GHz," Physics in Medicine and Biology, vol. 41,

pp. 2251-2269, 1996.

[90] S Gabriel, R W Lau, and C Gabriel, "The dielectric properties of biological tissues: III. Parametric

models for the dielectric spectrum of tissues," Physics in Medicine and Biology, vol. 41, pp. 2271-

2293, 1996.

[91] F S Barnes and B Greenebaum, Bioengineering and Biophysical Aspects of Electromagnetic Fields,

3rd ed.: CR C Press, 2006.

[92] F X Hart, N J Berner, and R L McMillen, "Modelling the anisotropic electrical properties of

skeletal muscle," Physics in Medicine and Biology, vol. 44, pp. 413-421, 1999.

[93] C Gabriel, A Peyman, and E H Grant, "Electrical conductivity of tissue at frequencies below 1

MHz," Physics in Medicine and Biology, vol. 54, pp. 4863-4878, 2009.

[94] A Peyman, A A Rezazadeh, and C Gabriel, "Changes in the dielectric properties of rat tissue as a

function of age at microwave frequencies," Physics in Medicine and Biology, vol. 46, pp. 1617-

1629, 2001.

[95] J O Ollikainen, M Vauhkonen, P A Karjalainen, and J P Kaipio, "Effects of local skull

inhomogeneities on EEG source estimation," Medical Engineering & Physics, vol. 21, pp. 143-154,

1999.

[96] D Gutierrez, A Nehorai, and C H Muravchik, "Estimating brain conductivities and dipole source

signals with EEG arrays," IEEE Transactions on Biomedical Engineering, vol. 51, pp. 2113-2122,

2004.

[97] J Latikka, T Kuurne, and H Eskola, "Conductivity of living intracranial tissues," Physics in

Medicine and Biology, vol. 46, pp. 1611-1616, 2001.

[98] M J Peters, J G Stinstra, and M Hendriks, "Estimation of the Electrical Conductivity of Human

Tissue," Electromagnetics, vol. 21, pp. 545-557, 2001.

[99] G Schmid, G Neubauer, and P R Mazal, "Dielectric Properties of Human Brain Tissue Measured

Less Than 10 h Postmortemat Frequencies From 800 to 2450 MHz," Bioelectromagnetics, vol. 24,

pp. 423-430, 2003.

Page 297: Magnetic Induction Tomography for Imaging Cerebral Stroke

276 <References

[100] A Peyman, S Holden, and C Gabriel, "Dielectric Properties of Biological Tissues at Microwave

Frequencies, Final technical report," MTHR Department of Health, U.K., 2005.

[101] A Peyman, S J Holden, S Watts, R Perrott, and C Gabriel, "Dielectric properties of porcine

cerebrospinal tissues at microwave frequencies: in vivo, in vitro and systematic variation with age," Physics in Medicine and Biology, vol. 52, pp. 2229-2245, 2007.

[102] A Peyman, C Gabriel, E H Grant, G Vermeeren, and L Martens, "Variation of the dielectric properties of tissues with age: the effect on the values of SAR in children when exposed to walkie-

talkie devices," Physics in Medicine and Biology, vol. 54, pp. 227-241, 2009.

[103] J Bao, S Lu, and W D Hurt, "Complex dielectric measurements and analysis of brain tissues in the

radio and microwave frequencies," IEEE Transactions on Microwave Theory and Techniques, vol.

45, pp. 1730-1741, 1997.

[104] T F Oostendorp, J Delbeke, and D F Stegeman, "The conductivity of the human skull: results of in

vivo and in vitro measurements," IEEE Transactions on Biomedical Engineering, vol. 47, pp. 1487-

1492, 2000.

[105] M Akhtari, H C Bryant, and A N Mamelak, "Conductivities of Three-Layer Live Human Skull,"

Brain Topography, vol. 14, pp. 151-167, 2002.

[106] M Akhtari, H C Bryant, and D Emin, "A Model for Frequency Dependence of Conductivities of the

Live Human Skull," Brain Topography, vol. 16, pp. 39-55, 2003.

[107] R Hoekema and G H Wieneke, "Measurement of the Conductivity of Skull, Temporarily Removed

During Epilepsy Surgery," Brain Topography, vol. 16, pp. 29-38, 2003.

[108] J Sierpowska, J Toyras, and M A Hakulinen, "Electrical and dielectric properties of bovine

trabecular bone - relationships with mechanical properties and mineral density," Physics in

Medicine and Biology, vol. 48, pp. 775-786, 2003.

[109] J Sierpowska, M A Hakulinen, J Toyras, and J S Day, "Prediction of mechanical properties of

human trabecular bone by electrical measurements," Physiological Measurement, vol. 26, pp.

S119-S131, 2005.

[110] J Sierpowska, M A Hakulinen, and J Toyras, "Interrelationships between electrical properties and

microstructure of human trabecular bone," Physics in Medicine and Biology, vol. 51, pp. 5289–

5303, 2006.

[111] J Sierpowska and M J Lammi, "Effect of human trabecular bone composition on its electrical

properties," Medical Engineering & Physics, vol. 29, pp. 845-852, 2007.

[112] S B Baumann, D R Wozny, S K Kelly, and F M Meno, "The Electrical Conductivity of Human

Cerebrospinal Fluid at Body Temperature," IEEE Transactions on Biomedical Engineering, vol. 44,

pp. 220-223, 1997.

[113] S Grimnes and O G Martinsen, Bioimpedance and Bioelectricity Basics, 2nd ed.: Elsevier Ltd.,

2008.

[114] F Jaspard, M Nadi, and A Rouane, "Dielectric properties of blood: an investigation of haematocrit

dependence," Physiological Measurement, vol. 24, pp. 137-147, 2003.

[115] F Jaspard and M Nadi, "Dielectric properties of blood: an investigation of temperature

dependence," Physiological Measurement, vol. 23, pp. 547-554, 2002.

[116] D S Holder, "Brief introduction to bioimpedance," in Electrical Impedance Tomography - Methods,

Page 298: Magnetic Induction Tomography for Imaging Cerebral Stroke

277 <References

History and Applications., D S Holder, Ed.: IOP Publishing, 2005, ch. Appendix A.

[117] J U Becker. emedicine. [Online]. http://emedicine.medscape.com/article/793904-overview (year

accessed: 2010)

[118] D S Holder, "Electrical impedance tomography of brain function," in Electrical Impedance

Tomography - Methods, History and Applications., D S Holder, Ed.: IOP Publishing, January 2005,

ch. 4.

[119] D Liang, S Bhatta, V Gerzanich, and M Simard, "Cytotoxic edema: mechanisms of pathological

cell swelling," Neurosurg Focus, vol. 22, p. E2, 2007.

[120] A Van H and S Ochs, "Cerebral Impedance Changes After Circulatory Arrest," The American

journal of physiology, vol. 187, pp. 180-192, 1956.

[121] D S Holder, "Electrical impedance tomography with cortical or scalp electrodes during global

cerebral ischaemia in the anaesthetised rat," Clinical Physics and Physiological Measurement, vol. 13, pp. 87-98, 1992.

[122] O Adeoye and J P Broderick, Intracerebral Hemorrhage, in A Primer on Stroke Prevention Treatment: An Overview Based on AHA/ASA Guidelines, edited by Larry B. Goldstein.: John Wiley

& Sons, 2009.

[123] H Scharfetter, A Kostinger, and S Issa, "Hardware for quasi-single-shot multifrequency magnetic

induction tomography (MIT): the Graz Mk2 system," Physiological Measurement, vol. 29, pp. 431-

443, 2008.

[124] P D Ledger and S Zaglmayr, "hp-Finite element simulation of three dimensional eddy current

problems on multiply connected domains," 2010.

[125] M Soleimani and W Lionheart, "Absolute Conductivity Reconstruction in Magnetic Induction

Tomography Using a Nonlinear Method," IEEE Transactions on Medical Imaging, vol. 25, pp.

1521-1530, 2006.

[126] R Merwa, K Hollaus, and H Scharfetter, "Numerical solution of the general 3D eddy current

problem for magnetic induction tomography (spectroscopy)," Physiological Measurement, vol. 24,

pp. 545-554, 2003.

[127] O C Zienkiewicz, R L Taylor, and J Z Zhu, The Finite Element Method: Its Basis and

Fundamentals, 6th ed.: Elsevier Butterworth-Heinemann, 2005.

[128] W H A Schilders and E J W Maten, Handbook of Numerical Analysis, Special Volume: Numerical

Methods in Electromagnetics.: Elsevier B.V., 2005.

[129] A Bondeson, T Rylander, and P Ingelstrom, Computational Electromagnetics.: Springer, 2005.

[130] A C Polycarpou, Introduction to the Finite Element Method in Electromagnetics.: Morgan &

Claypool, 2006.

[131] P D Ledger, Computational Electromagnetics.: Swansea University, 2009.

[132] V Thomée, Galerkin Finite Element Methods for Parabolic Problems, 2nd ed.: Springer, 2006.

[133] B Rivière, Discontinuous Galerkin methods for solving elliptic and parabolic equations : theory

and implementation.: Siam, 2008.

[134] J C Nedelec, "Mixed finite elements in R3," Numerische Mathematik, vol. 35, pp. 314-341, 1980.

Page 299: Magnetic Induction Tomography for Imaging Cerebral Stroke

278 <References

[135] O Biro, "Edge element formulations of eddy current problems," Computer Methods in Applied

Mechanics and Engineering, vol. 169, pp. 391-405, 1999.

[136] (2009) Comsol Multiphysics. [Online]. http://www.comsol.com/ (year accessed: 2010)

[137] R Garg, Analytical and Computational Methods in Electromagnetics.: Artech House, 2008.

[138] H Ammari and J C Nedelec, "Generalized impedance boundary conditions for the maxwell

equations as singular perturbations problems," Communications in Partial Differential Equations,

vol. 24, pp. 24-38, 1999.

[139] C Buccella, M Feliziani, and F Maradei, "Magnetic Field Computation in a Physically Large

Domain With Thin Metallic Shields," IEEE Transactions on Magnetics, vol. 41, pp. 1708-1711,

2005.

[140] P Bettess, "Infinite elements," International Journal for Numerical Methods in Engineering, vol.

11, pp. 53-64, 2005.

[141] O C Zienkiewicz, C Emson, and P Bettess, "A Novel Boundary Infinite Element," International

Journal for Numerical Methods in Engineering, vol. 19, pp. 393-404, 1983.

[142] J P Berenger, "A perfectly matched layer for the absorption of electromagnetic waves," Journal of

Computational Physics, vol. 114, pp. 185-200, 1994.

[143] J P Berenger, Perfectly Matched Layer (PML) for Computational Electromagnetics.: Morgan &

Claypool, 2007.

[144] COMSOL AB. (2008) Comsol Multiphysics User‟s Guide v3.5.

[145] Y Saad, "A flexible inner-outer preconditioned GMRES algorithm," SIAM Journal on Scientific

Computing, vol. 14, pp. 461-469, 1993.

[146] P D Ledger and K Morgan, "The application of the hp finite element method to electromagnetic

problems," Archives of Computational Methods in Engineering, vol. 12, pp. 235-302, 2005.

[147] J Schoberl and S Zaglmayr, "High order Nedelec elements with local complete sequence

properties," COMPEL: The International Journal for Computation and Mathematics in Electrical and Electronic Engineering, vol. 24, pp. 374-84, 2005.

[148] M V K Chari and S J Salon, Numerical Methods in Electromagnetism.: Academic Press, 2000.

[149] M N O Sadiku, Numerical techniques in electromagnetics, 2nd ed.: CRC Press, 2001.

[150] N G Gencer, M Kuzuoglu, and Y Z Ider, "Electrical impedance tomography using induced

currents," IEEE Transactions on Medical Imaging, vol. 13, pp. 338-350, 1994.

[151] N G Gencer and M N Tek, "Forward problem solution for electrical conductivity imaging via

contactless measurements," Physics in Medicine and Biology, vol. 44, pp. 927-940, 1999.

[152] D W Armitage, H H LeVeen, and R Pethig, "Radiofrequency-induced hyperthermia: computer

simulation of specific absorption rate distributions using realistic anatomical models," Physics in

Medicine and Biology, vol. 28, pp. 31-42, 1983.

[153] A Morris, H Griffiths, and W Gough, "A numerical model for magnetic induction tomographic

measurements in biological tissues," Physiological Measurement, vol. 22, pp. 113-119, 2001.

[154] K Asami, "Dielectric dispersion in biological cells of complex geometry simulated by the three-

dimensional finite difference method," Journal of Physics D: Applied Physics, vol. 39, pp. 492-499,

Page 300: Magnetic Induction Tomography for Imaging Cerebral Stroke

279 <References

2006.

[155] A Quarteroni, R Sacco, and F Saleri, Numerical Mathematics, 2nd ed.: Springer, 2007.

[156] C Ktistis, D W Armitage, and A J Peyton, "Calculation of the forward problem for absolute image

reconstruction in MIT," Physiological Measurement, vol. 29, pp. 455-464, 2008.

[157] E Somersalo, D Isaacson, and M Cheney, "A linearized inverse boundary value problem for

Maxwell‟s equations," Journal of Computational and Applied Mathematics, vol. 42, pp. 123-36,

1992.

[158] M Vauhkonen, Electrical impedance tomography and prior information, PhD Thesis.: Kuopio

University, 1997.

[159] L D Landau and E M Lifshitz, Electrodynamics of Continuous Media, Volume 8 of Course of

Theoretical Physics.: Pergamon Press, 1984.

[160] A J Peyton. Electrical Tomography for Industrial Applications. [Online]. (EIS/Tomography

Systems)" http://www.intranet.eee.manchester.ac.uk/intranet/pg/coursematerial/ (year accessed:

2010)

[161] F Hettlich and W Rundell, "The determination of a discontinuity in a conductivity from a single

boundary measurement," Inverse Problems, vol. 14, pp. 67-82, 1998.

[162] H Ki and D Shen, "Numerical inversion of discontinuous conductivities," Inverse Problems, vol.

16, pp. 33-47, 2000.

[163] K Kolehmainen, S R Arridge, W R Lionheart, M Vauhkonen, and J P Kaipio, "Recovery of region

boundaries of piecewise constant coeffcients of elliptic PDE from boundary data," Inverse Problems, vol. 15, pp. 1375-91, 1999.

[164] V Kolehmainen, M Vauhkonen, J P Kaipio, and S R Arridge, "Recovery of piecewise constant

coeffcients in optical diffusion tomography," Optics Express, vol. 7, pp. 468-80, 2000.

[165] Å Björck, Numerical Methods for Least Squares Problems.: SIAM, 1996.

[166] G R Liu and X Han, Computational inverse techniques in nondestructive evaluation.: CRC Press LLC, 2003.

[167] J Hadamard, Lectures on Cauchy's Problem in Linear Partial Differential Equations.: Yale University Press, New Haven, 1923.

[168] C Hansen, Rank-deficient and discrete ill-posed problems : numerical aspects of linear inversion.: SIAM, 1998.

[169] K Hollaus, C Magele, R Merwa, and H Scharfetter, "Fast calculation of the sensitivity matrix in magnetic induction tomography by tetrahedral edge finite elements and the reciprocity theorem,"

Physiological Measurement, vol. 25, pp. 159-168, 2004.

[170] G Allaire and S M Kaber, Numerical Linear Algebra.: Springer, 2008.

[171] L Hogben, Handbook of Linear Algebra.: Taylor & Francis Group, 2007.

[172] L N Trefethen and D Bau, Numerical Linear Algebra.: SIAM, 1997.

[173] W R Lionheart, N Polydorides, and A Borsic, "The reconstruction problem," in Electrical

Impedance Tomography - Methods, History and Applications., D S Holder, Ed.: IOP Publishing,

Page 301: Magnetic Induction Tomography for Imaging Cerebral Stroke

280 <References

2005, ch. 1.

[174] W R Lionheart, "Reconstruction Algorithms for Permittivity and Conductivity Imaging," in 2nd

World Congress on Industrial Process Tomography, Hannover, Germany, 2001.

[175] P D Ledger, Personal communications, Swansea University, 2010.

[176] R C Aster, B Borchers, and C H Thurber, Parameter Estimation and Inverse Problems.: Elsevier

Academic Press, 2005.

[177] Lin-Ping Song and Shu-Yi Zhang, "Singular value decomposition-based reconstruction algorithm

for seismic traveltime tomography," IEEE Transactions on Image Processing, vol. 8, pp. 1152 -

1154, 1999.

[178] J C Santamarina and D Fratta, Discrete Signals and Inverse Problems: An Introduction for

Engineers and Scientists.: John Wiley & Sons, 2005.

[179] R Merwa, K Hollaus, P Brunner, and H Scharfetter, "Solution of the inverse problem of magnetic

induction tomography (MIT)," Physiological Measurement, vol. 26, pp. 241-250, 2005.

[180] C R Vogel, Computational methods for inverse problems.: SIAM, 2002.

[181] G Wahba, Spline models for observational data.: SIAM, 1990.

[182] B M Graham and A Adler, "Objective selection of hyperparameter for EIT," Physiological

Measurement, vol. 27, pp. S65-S79, 2006.

[183] M A Lukas, "Strong robust generalized cross-validation for choosing the regularization parameter,"

Inverse Problems, vol. 24, p. 034006 (16pp), 2008.

[184] M A Lukas, "Robust generalized cross-validation for choosing the regularization parameter,"

Inverse Problems, vol. 22, pp. 1883–1902, 2006.

[185] P C Hansen, T Jensen, and G Rodriguez, "An adaptive pruning algorithm for the discrete L-curve

criterion," Journal of Computational and Applied Mathematics, vol. 198, pp. 483-492, 2007.

[186] P Brunner, R Merwa, and H Scharfetter, "Reconstruction of the shape of conductivity spectra using

differential multi-frequency magnetic induction tomography," Physiological Measurement, vol. 27,

pp. S237-S248, 2006.

[187] H Scharfetter, R Casañas, and J Rosell, "Biological tissue characterization by magnetic induction

spectroscopy (MIS): requirements and limitations," IEEE Transactions on Biomedical Engineering,

vol. 50, pp. 870-880, 2003.

[188] A G Ramm, Inverse Problems: Mathematical And Analytical Techniques With Applications To

Engineering.: Springer, 2005.

[189] A B Bakushinsky and M Y Kokurin, Iterative Methods for Approximate Solution of Inverse

Problems.: Springer, 2004.

[190] W Q Yang and L Peng, "Image reconstruction algorithms for electrical capacitance tomography,"

Measurement Science and Technology, vol. 14, pp. 1-13, 2003.

[191] M Soleimani, "Computational aspects of low frequency electrical and electromagnetic tomography:

A review study," International Journal of Numerical Analysis and Modeling, vol. 5, pp. 407-440, 2008.

Page 302: Magnetic Induction Tomography for Imaging Cerebral Stroke

281 <References

[192] P Charbonnier, L Blanc-Feraud, and G Aubert, "Deterministic Edge-Preserving Regularization in

Computed Imaging," IEEE Transactions on Image Processing, vol. 6, pp. 298-311, 1997.

[193] R Casanova, A Silva, and A R Borges, "MIT image reconstruction based on edge-preserving

regularization," Physiological Measurement, vol. 25, pp. 195-207, 2004.

[194] D F Yu and J A Fessler, "Edge-Preserving Tomographic Reconstruction with Nonlocal

Regularization," IEEE Transactions on Medical Imaging, vol. 21, pp. 159-173, 2002.

[195] C V Alvino and A J Yezzi, "Tomographic reconstruction of piecewise smooth images," in

Proceedings of the 2004 IEEE Computer Society Conference on Computer Vision and Pattern Recognition (CVPR’04), 2004.

[196] A Borsic, B M Graham, A Adler, and W B Lionheart, "Total Variation Regularization in Electrical Impedance Tomography," ISSN 1749-9097, 2007.

[197] M Bertero, Introduction to Inverse Problems in Imaging.: IOP, 1998.

[198] R H Bayford, A Gibson, and A Tizzard, "Solving the forward problem in electrical impedance

tomography for the human head using IDEAS (integrated design engineering analysis software), a finite element modelling tool," Physiological Measurement, vol. 22, pp. 55-64, 2001.

[199] A Tizzard, L Horesh, and R J Yerworth, "Generating accurate finite element meshes for the forward model of the human head in EIT," Physiological Measurement, vol. 26, pp. S251-S261,

2005.

[200] A A Bagshaw, A D Liston, and R H Bayford, "Electrical impedance tomography of human brain

function using reconstruction algorithms based on the finite element method," NeuroImage, vol. 20,

pp. 752-764, 2003.

[201] A Romsauerova, A McEwan, L Horesh, R H Bayford, and D S Holder, "Multi-frequency electrical

impedance tomography (EIT) of the adult human head: initial findings in brain tumours,

arteriovenous malformations and chronic stroke, development of an analysis method and

calibration," Physiological Measurement, vol. 27, pp. S147-S161, 2006.

[202] L Horesh, O Gilad, A Romsauerova, A McEwan, and D S Holder, "Stroke type differentiation by

multifrequency electrical impedance tomography-a feasibility study," in Proc IFMBE 11, Paper

1741, 2005.

[203] L Fabrizi, A McEwan, T Oh, E J Woo, and D S Holder, "A comparison of two EIT systems suitable

for imaging impedance changes in epilepsy," Physiological Measurement, vol. 30, pp. S103-S120,

2009.

[204] O Gilad, L Horesh, and D S Holder, "A modelling study to inform specification and optimal

electrode placement for imaging of neuronal depolarization during visual evoked responses by

electrical and magnetic detection impedance tomography," Physiological Measurement, vol. 30, pp.

S201-S224, 2009.

[205] N Makris, L Angelone, and S Tulloch, "MRI-based anatomical model of the human head for

specific absorption rate mapping," Medical & Biological Engineering & Computing, vol. 46, pp.

1239-1251, 2008.

[206] H C Wee, S Watson, R Patz, H Griffiths, and R J Williams, "A Magnetic Induction Tomography

system with sub-millidegree phase noise and high long-term phase stability," in IFMBE

Proceedings, vol. 22, 2008, pp. 744-747.

[207] M M Elwassif, Q Kong, and M Vazquez, "Bio-heat transfer model of deep brain stimulation-

Page 303: Magnetic Induction Tomography for Imaging Cerebral Stroke

282 <References

induced temperature changes," Journal of Neural Engineering, vol. 3, pp. 306-315, 2006.

[208] C M Collins and M B Smith, "Signal-to-Noise Ratio and Absorbed Power as Functions of Main

Magnetic Field Strength, and Definition of “90°” RF Pulse for the Head in the Birdcage Coil,"

Magnetic Resonance in Medicine, vol. 45, pp. 684-691, 2001.

[209] J W Hand and R W Lau, "Electromagnetic and Thermal Modeling of SAR and Temperature Fields

in Tissue Due to an RF Decoupling Coil," Magnetic Resonance in Medicine, vol. 42, pp. 183-192, 1999.

[210] A Hadjem and D Lautru, "Study of Specific Absorption Rate (SAR) Induced in Two Child Head Models and in Adult Heads Using Mobile Phones," IEEE Transactions on Microwave Theory and

Techniques, vol. 53, pp. 4-11, 2005.

[211] A Trakic, S Crozier, and F Liu, "Numerical modelling of thermal effects in rats due to high-field

magnetic resonance imaging (0.5–1 GHz)," Physics in Medicine and Biology, vol. 49, pp. 5547-

5558, 2004.

[212] C M Collins, W Liu, and J Wang, "Temperature and SAR calculations for a human head within

volume and surface coils at 64 and 300 MHz," Journal of Magnetic Resonance Imaging, vol. 19,

pp. 650-6, 2004.

[213] "Medical electrical equipment-particular requirements for the safety of magnetic resonance

equipment for medical diagnosis," IEC 60601-2-33, 1995, Revised 2002.

[214] W R Lionheart, Personal communications, Swansea University, 2010.

[215] D Gursoy and H Scharfetter, "Optimum Receiver Array Design for Magnetic Induction

Tomography," IEEE Transactions on Biomedical Engineering, vol. 56, pp. 1435-1441, 2009.

[216] D Gursoy and H Scharfetter, "The effect of receiver coil orientations on the imaging performance

of magnetic induction tomography," Measurement Science and Technology, vol. 20, p. 105505

(9pp), 2009.

[217] W A Wan-Daud, A J Peyton, and H S Tapp, "Towards an integrated system using multiple cameras

and structured lighting for the measurement of human body shape and volume," Biometrie

Humaine et Anthropologie, vol. 21, pp. 121-127.

[218] C C Gordon, T Churchill, and C Clauser, Anthropometric Survey of U.S. Army Personnel: Methods

and Summary Statistics 1988.: U.S. Army Natick RD&E Center, 1998. [Online].

http://www.dtic.mil/dticasd/docs-a/anthro_military.html (year accessed: 2010)

[219] A Curtis, A Michelini, and D Leslie, "A deterministic algorithm for experimental design applied to

tomographic and microseismic monitoring surveys," Geophysical Journal International, vol. 157,

pp. 595-606, 2004.

[220] D Gursoy and H Scharfetter, "Reconstruction artefacts in magnetic induction tomography due to

patient‟s movement during data acquisition," Physiological Measurement, vol. 30, pp. S165-S174,

2009.

[221] M Ainsworth and J T Oden, A Posteriori Error Estimation in Finite Element Analysis.: John Wiley

& Sons, 2000.

[222] P D Ledger, J Peraire, K Morgan, O Hassan, and N P Weatherill, "Adaptive hp finite element

computations of the scattering width output of Maxwell‟s equations," International Journal for

Numerical Methods in Fluids, vol. 43, pp. 953-78, 2003.

Page 304: Magnetic Induction Tomography for Imaging Cerebral Stroke

283 <References

[223] J Schöberl, "A posteriori error estimates for Maxwell equations," Mathematics Computation, vol.

77, pp. 633-649, 2008.

[224] R Beck, R Hiptmair, and B Wohlmuth, "Hierarchical error estimator for eddy current computation,"

in 2nd European Conference on Numerical Methods and Advanced Applications, Singapore, 1999, pp. 110-20.

[225] M Molinari, High fidelity imaging in electrical impedance tomography, PhD Thesis.: University of Southampton, 2003.

[226] D E Post and L G Votta, "Computational Science Demands a New Paradigm," Physics Today, vol. 58, pp. 35-41, 2005.

[227] R Pethig, Dielectric and Electronic Properties of Biological Materials.: John Wiley & Sons, 1979.

[228] J R Shewchuk, "An Introduction to the Conjugate Gradient Method Without the Agonizing Pain,"

1994.

[229] S Watson, C Ktistis, and B Dekdouk, "Developments in Magnetic Induction Tomography for Oil-

Industry Applications," in 5th World Congress on Industrial Process Tomography, Bergen, Norway, 2007.

[230] B Kaltenbacher, A Neubauer, and O Scherzer, Iterative Regularization Methods for Nonlinear Ill-Posed Problems.: Hubert & Co. GmbH & Co. KG, Göttingen, 2008.

[231] R G Gonzalez, J A Hirsch, and W J Koroshetz, Acute Ischemic Stroke, Imaging and Intervention.: Springer, 2006.

[232] P C Hansen, "Regularization Tools version 4.0 for Matlab 7.3," Numerical Algorithms, vol. 46, pp. 189-194, 2007.

[233] P C Hansen. Regularization Tools Version 4.1 (for Matlab Version 7.3). [Online].

http://www2.imm.dtu.dk/~pch/Regutools/index.html (year accessed: 2010)

[234] U Tremel, F Deister, and O Hassan, "Parallel generation of unstructured surface grid," Engineering

with Computers, vol. 21, pp. 36-46, 2005.

[235] F Desiter, U Tremel, O Hassan, and N P Weaterill, "Fully automatic and fast mesh size

specification for unstructured grid generation," Engineering with Computers, vol. 20, pp. 237-248,

2004.

[236] Y Maimaitijiang, M A Roula, and S Watson, "Parallelization methods for implementation of a

magnetic induction tomography forward model in symmetric multiprocessor systems," Parallel

Computing, vol. 34, pp. 497-507, 2008.