Close
About
FAQ
Home
Collections
Login
USC Login
Register
0
Selected
Invert selection
Deselect all
Deselect all
Click here to refresh results
Click here to refresh results
USC
/
Digital Library
/
University of Southern California Dissertations and Theses
/
Geometrical modeling and analysis of cortical surfaces: An approach to finding flat maps of the human brain
(USC Thesis Other)
Geometrical modeling and analysis of cortical surfaces: An approach to finding flat maps of the human brain
PDF
Download
Share
Open document
Flip pages
Contact Us
Contact Us
Copy asset link
Request this asset
Transcript (if available)
Content
Geometrical Modeling and Analysis of Cortical Surfaces:
An Approach to Finding Flat Maps of The Human Brain
by
Bijan Timsari
A Dissertation Presented to the Faculty o f the Graduate School
University o f Southern California
In Partial Fulfillment o f the Requirements fo r the Degree
D octor o f Philosophy in Electrical Engineering
December 1999
Copyright 1999 - Bijan Timsari
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
UMI Number: 3110962
INFORMATION TO USERS
The quality of this reproduction is dependent upon the quality of the copy
submitted. Broken or indistinct print, colored or poor quality illustrations and
photographs, print bleed-through, substandard margins, and improper
alignment can adversely affect reproduction.
In the unlikely event that the author did not send a complete manuscript
and there are missing pages, these will be noted. Also, if unauthorized
copyright material had to be removed, a note will indicate the deletion.
®
UMI
UMI Microform 3110962
Copyright 2004 by ProQuest Information and Learning Company.
All rights reserved. This microform edition is protected against
unauthorized copying under Title 17, United States Code.
ProQuest Information and Learning Company
300 North Zeeb Road
P.O. Box 1346
Ann Arbor, Ml 48106-1346
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
UNIVERSITY OF SOUTHERN CALIFORNIA
THE GRADUATE SCHOOL
UNIVERSITY PARK
LOS ANGELES, CALIFORNIA 90007
This dissertation, written by
...............................
under the direction of h.i& Dissertation
Committee, and approved by all its members,
has been presented to and accepted by The
Graduate School, in partial fulfillment of re
quirements for the degree of
DOCTOR OF PHILOSOPHY
Dean o f G raduate S tu d ie s
Date ..M.y.s m b l .99.9,
DISSERTATION COMMITTEE
Chairperson
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Dedication
To my parents.
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Acknowledgments
First and foremost, I wish to express my sincerest gratitude to my advisor, Professor
Richard Leahy, for introducing this topic to me and for providing me the resources to
pursue my research. I am grateful for the countless hours that he spent discussing with me
and for his unbelievable patience in guiding me through the research. I feel extremely
fortunate to have been the student of such an expert and understanding mentor.
I would also like to thank the other members of my guidance committee: Professors Ulrich
Neumann, Antonio Ortega, Michael Arbib, and Alexander Sawchuk for their time and
effort in evaluating and giving valuable input to my research. I also wish to thank
Professors Michel Baudry and Michael Arbib who supported my research and gave me the
opportunity to contribute to the USC Brain Project.
I wish to express my special thanks to the faculty and staff of SIPI for providing such an
invigorating environment and for their commitment to help and support the students. They
have helped me in many different ways on numerous occasions. In particular I want to
thank Dr. Allen Weber for helping me with my computer related problems and Gloria
Halfacre, Linda Varilla and Regina Morton for their assistance in solving my
administrative problems.
During my stay at SIPI I have enjoyed the friendship of a wonderful group of people,
among them I am most grateful to Stepahnie Sandor, who got me started on my research
and helped me through my settlement at SIPI. I am also thankful to my other friends
Chighan Hsu, James Philips, Jinyi Qi, David Shattuck, John Ermer, Sylvain Baillet and
Paola Bonetto whose company made my years of graduate study an unforgettable part of
my life.
Finally I want to thank my parents for the love and support they have given me in my
entire life. They have made everything possible for me and I can never thank them enough.
iii
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Table of Contents
Dedication ii
Acknowledgement iii
List of Figures vi
Abstract xiii
CHAPTER 1 Introduction 1
1.1 Complexity of The B rain....................................................................... 1
1.2 Development of the Cortical F olding................................................. 1
1.3 Geometry of the Cerebral Cortex ........................................................2
1.4 Modeling the Cortical Surface..............................................................4
1.5 Mapping the Human Brain ................................................................... 5
1.6 Dissertation Goals .................................................................................. 6
1.6.1 Summary of Main Contributions........................ 6
1.6.2 Overview of the Dissertation.....................................................8
CHAPTER 2 Tesselation of the Human Cortical Surface 9
2.1 Survey of Current Techniques............................................................10
2.2 Triangulation Algorithm Based on Coordinate C urves................12
2.2.1 Contour T racing..........................................................................13
2.2.2 Tiling ............................................................................................ 14
2.3 Results and Discussion ........................................................................ 17
CHAPTER 3 Parametric Modeling of the Human Cortical Surface 21
3.1 Survey of Current Techniques........................................................... 21
3.2 Local Fitting of Quadratic Patches to a Triangular Mesh ...........22
CHAPTER 4 The Curvature of the Human Cortical Surface 27
4.1 Introduction ............................................................................................27
4.2 Some Basic Concepts of Differential Geometry of Surfaces .....28
4.3 Quantitative Analysis of Surface Curvature................................... 30
4.4 Curvature Maps .....................................................................................32
4.5 Experimental R esults........................................................................... 34
CHAPTER 5 Finding the Flat Map of the Human Cortical Surface 41
5.1 Mappings ................................................................................................ 41
iv
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
5.2 Survey of Current Techniques.............................................................44
5.2.1 Mechanical Model Based Flattening Algorithms .............44
5.2.2 Geometrical Model Based Flattening A lgorithm s............46
5.3 Flattening As An Optimization Problem .......................................... 48
5.3.1 Objective Function: Minimizing “Convexity” ..................50
5.3.2 Constraint Function: Preserving the Area ..........................53
5.3.3 Constraint Function: Preserving the Angle ........................54
5.3.4 Perturbation Function: Smoothing the Surface ................ 56
5.3.5 Merit Function: The complete Energy Function .............. 57
5.4 Solving the Minimization Problem: Conjugate Gradient Method 58
5.5 Results and D iscussion ..........................................................................60
5.5.1 Phantom Studies .......................................................................60
5.5.5.1 Equiareal and Conformal Maps of C ylinder 62
5.5.5.2 Equiareal and Conformal Maps of Hemisphere ..66
5.5.5.3 Isometric Maps of Half-Cylinder & Hemisphere 70
5.5.2 Isometric map of the Human Cortical Surface ..................73
5.6 Application: Multiresolution Representation of the Cortical Surface
................... 85
CHAPTER 6 Summary and Future Research 89
6.1 Conclusion.................................................................................................89
6.2 Future W ork............................................................................................. 91
Bibliography 92
V
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
List of Figures
Figure 2.1 8-Connectivity vs. 4-Connectivity, (a) Sample shape, (b) Boundary
points extracted based on 4-neighborhood system are 8-connected. (c)
Boundary points extracted based on 8-neighborhood system are 4-
connected..............................................................................................................13
Figure 2.2 18 closest neighbors of a point......................................................................... 14
Figure 2.3 Points with only two pairs of neighbors are tiled with four equal right
triangles. ............................................................................................................14
Figure 2.4 (a) A point with three pairs of neighbors forms 12 triangles, (b) A point
with overlapping neighbors forms less than 12 triangles, neighbor points
Y ’ and Z’ are overlapped and only 8 triangles are formed.......................15
Figure 2.5 Two types of unit tiling patch...........................................................................15
Figure 2.6 Tiling the surface with a unit patch at a typical point.................................16
Figure 2.7 Tessellation of a cube using coordinate curves............................................16
Figure 2.8 Adjacent triangles with outward pointing normals......................................17
Figure 2.9 (a) Phantom surface constructed from coordinate curves (2894 vertices,
5784 triangles), (b) Phantom surface tessellated using Marching Cubes
algorithm (3782 vertices,7560 triangles), (c) Resampled phantom
constructed from coordinate curves (4130 vertices, 8256 triangles). ...18
Figure 2.10 Reconstructed surface of the rotated phantom using the coordinate
curves method..................................................................................................... 19
vi
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 2.11
Figure 2.12
Figure 2.13
Figure 3.1
Figure 3.2
Figure 3.3
Figure 4.1
Figure 4.2
Figure 4.3
Figure 4.4
Reconstructed surface of the rotated phantom using the Marching Cubes
algorithm.............................................................................................................. 19
Human brain surface reconstructed using coordinate curves. No. of
vertices: 73814 No. of triangles: 147645..................................................... 20
Human brain surface reconstructed using Marching Cubes. No. of
vertices: 128984 No. of triangles: 258004...................................................20
Surface normal at a typical point of a triangulated mesh......................... 23
The local rectangular coordinate system in the neighborhood of a point
X,............................................................................................................................ 25
Fitting a local quadratic patch to a sphere, (a) Analytically defined
spherical phantom, (b) Magnified display of a small region of sphere, (c)
The same region approximated with a local quadratic patch, (d)
Superimposition of the original and approximated patches.....................26
Normal Curvature at a point............................................................................30
Sphere phantom, (a) Binary volume, (b) Tessellated Surface. (3) The
nodes in the triangulated sphere are repositioned to new locations that
conform with the analytical description of an exact sphere.....................34
Curvature maps for a sphere with a radius of 25 units, (a) Minimum
curvature map. (b) Maximum curvature map. (c) Mean curvature map.
(d) Gaussian curvature map.............................................................................35
Relative variation of the minimum curvature in the neighborhood of a
dent and a bump, (a) In the neighborhood of a dent, the minimum
curvature takes negative values, (b) In the neighborhood of a bump, the
minimum curvature takes positive values. Minimum curvature is a good
feature for locating dents..................................................................................37
Vll
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 4.5
Figure 4.6
Figure 4.7
Figure 4.8
Figure 4.9
Figure 5.1
Figure 5.2
Figure 5.3
Figure 5.4
Figure 5.5
Relative variation of the maximum curvature in the neighborhood of a
dent and a bump, (a) In the neighborhood of a dent, the maximum
curvature takes negative values, (b) In the neighborhood of a bump, the
minimum curvature takes positive values. Maximum curvature is a good
feature for locating bumps............................................................................... 37
Minimum curvature of human cortical surface mapped onto: (a) Original
brain, (b) Inflated brain.....................................................................................39
Maximum curvature of human cortical surface mapped onto: (a) Original
brain, (b) Inflated brain.....................................................................................39
Mean curvature of human cortical surface mapped onto: (a) Original
brain, (b) Inflated brain.....................................................................................40
Gaussian curvature of human cortical surface mapped onto: (a) Original
brain, (b) Inflated brain.....................................................................................40
Convexity vs. mean curvature, (a) Mean curvature takes relatively equal
values for both small and large folds, (b) Convexity takes high values for
large folds and is almost insensitive to small folds....................................50
Xf, a typical node of a tessellated surface, ni represents surface normal
and Xt j represent the edges...............................................................................51
(a) At a locally convex point all the neighbors are above the tangent
plane, (b) At a locally concave point all the neighbors are below the
tangent plane....................................................................................................... 51
Normal to the plane of a triangle....................................................................52
Cos of the angle between two vectors...........................................................54
viii
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 5.6 Tessellated half-cylinder.................................................................................62
Figure 5.7 Distribution of areal and angular distortions for the normal projection
map of the half-cylinder, (a) Areal distortion map. (b) Histogram of areal
distortion, (c) Angular distortion map. (d) Histogram of angular
distortion.............................................................................................................. 63
Figure 5.8 Variation of areal distortion in the process of generating an equiareal
map for half-cylinder, (a) 10 iterations, (b) 50 iterations, (c) 100
iterations, (d) 10000 iterations........................................................................64
Figure 5.9 Variation of angular distortion in the process of generating a conformal
map for half-cylinder, (a) 10 iterations, (b) 500 iterations, (c) 2500
iterations, (d) 10000 iterations........................................................................65
Figure 5.10 Tessellated hemisphere....................................................................................66
Figure 5.11 Distribution of areal and angular distortions for the normal projection
map of the hemisphere, (a) Areal distortion map. (b) Histogram of areal
distortion, (c) Angular distortion map. (d) Histogram of angular
distortion.............................................................................................................. 67
Figure 5.12 Variation of areal distortion in the process of generating an equiareal
map for hemisphere, (a) 10 iterations, (b) 50 iterations, (c) 100 iterations,
(d) 1000 iterations..............................................................................................68
Figure 5.13 Variation of angular distortion in the process of generating an equiareal
map for hemisphere, (a) 25 iterations, (b) 150 iterations, (c) 300
iterations, (d) 500 iterations............................................................................ 69
Figure 5.14 Isometric flat map of half-cylinder, (a) Areal distortion map. (b)
Histogram of areal error, (c) Angular distortion map. (d) Histogram of
angular error........................................................................................................71
ix
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 5.15 Isometric flat map of hemisphere, (a) Areal distortion map. (b)
Histogram of areal error, (c) Angular distortion map. (d) Histogram of
angular error........................................................................................................ 72
Figure 5.16 (a) Tessellated cortical surface, (b) Same surface after applying
smoothing relaxation....................................................................................... 74
Figure 5.17 (a) Inflated human cortical surface, (b) The front view of a cortical
hemisphere after applying a few iterations of relaxation operation to the
cortex, (c) The back view of the same cortical hemisphere.....................75
Figure 5.18 Distribution of convexity, areal and angular distortions of a typical right
hemispherical human cortex after 250 iterations with p r = io5, pv = i and
p f l = i . (a) Convexity superimposed over partially unfolded cortex, (b)
Histogram of convexity, (c) Areal distortion superimposed over the flat
map of cortex, (d) Histogram of areal distortion, (e) Angular distortion
superimposed over the flat map of cortex, (f) Histogram of angular
distortion.............................................................................................................. 77
Figure 5.19 Distribution of convexity, areal and angular distortions of a typical right
hemispherical human cortex after 500 iterations with pc = io4 , pv = i and
P a = i . (a) Convexity superimposed over partially unfolded cortex, (b)
Histogram of convexity, (c) Areal distortion superimposed over the flat
map of cortex, (d) Histogram of areal distortion, (e) Angular distortion
superimposed over the flat map of cortex, (f) Histogram of angular
distortion.............................................................................................................. 78
Figure 5.20 Distribution of convexity, areal and angular distortions of a typical right
hemispherical human cortex after 1000 iterations with pf = io 3 , ps = i and
p f l = i . (a) Convexity superimposed over partially unfolded cortex, (b)
Histogram of convexity, (c) Areal distortion superimposed over the flat
map of cortex, (d) Histogram of areal distortion, (e) Angular distortion
superimposed over the flat map of cortex, (f) Histogram of angular
distortion.............................................................................................................. 79
Figure 5.21 Distribution of convexity, areal and angular distortions of a typical right
hemispherical human cortex after 2500 iterations with pf = i o o , ps = i and
x
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
p f l = i . (a) Convexity superimposed over partially unfolded cortex, (b)
Histogram of convexity, (c) Areal distortion superimposed over the flat
map of cortex, (d) Histogram of areal distortion, (e) Angular distortion
superimposed over the flat map of cortex, (f) Histogram of angular
distortion...............................................................................................................80
Figure 5.22 Distribution of convexity, areal and angular distortions of a typical right
hemispherical human cortex after 3000 iterations with p c = io, pv = io and
p f l = io. (a) Convexity superimposed over partially unfolded cortex, (b)
Histogram of convexity, (c) Areal distortion superimposed over the flat
map of cortex, (d) Histogram of areal distortion, (e) Angular distortion
superimposed over the flat map of cortex, (f) Histogram of angular
distortion...............................................................................................................81
Figure 5.23 Distribution of convexity, areal and angular distortions of a typical right
hemispherical human cortex after 3500 iterations with p c = io, pt = ioo and
p n = ioo. (a) Convexity superimposed over partially unfolded cortex, (b)
Histogram of convexity, (c) Areal distortion superimposed over the flat
map of cortex, (d) Histogram of areal distortion, (e) Angular distortion
superimposed over the flat map of cortex, (f) Histogram of angular
distortion...............................................................................................................82
Figure 5.24 Distribution of convexity, areal and angular distortions of a typical right
hemispherical human cortex after 5000 iterations with pc = i , p s = ioo and
p a = ioo. (a) Convexity superimposed over partially unfolded cortex, (b)
Histogram of convexity, (c) Areal distortion superimposed over the flat
map of cortex, (d) Histogram of areal distortion, (e) Angular distortion
superimposed over the flat map of cortex, (f) Histogram of angular
distortion...............................................................................................................83
Figure 5.25 Distribution of convexity, areal and angular distortions of a typical right
hemispherical human cortex after 10000 iterations with p c = o , p( = 1000
and p a = i o o o . (a) Convexity superimposed over partially unfolded cortex,
(b) H istogram o f convexity, (c) Areal distortion superim posed over the
flat map of cortex, (d) Histogram of areal distortion, (e) Angular
distortion superimposed over the flat map of cortex, (f) Histogram of
angular distortion................................................................................................84
XI
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 5.26 Geometric relationship used to interpolate position of nodes during the
resampling operation......................................................................................... 86
Figure 5.27 Multiresolution representation of a hemisphere created from its isometric
flat map...................................... 87
Figure 5.28 Multiresolution representation of a cortical hemisphere created from its
isometric flat map...............................................................................................88
xii
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
A bstract
In this work we have developed computational methods for creating three-dimensional
models and unfolded two-dimensional isometric maps of the cerebral cortex. The three
dimensional model can be used for visualization of the cortical anatomy and structural
organization of the brain as well as for quantitative analysis and measurement of the
geometric features of the cortical surface. The isometric flat map provides a means for
effective inter-subject studies allowing visualization of the distribution of functional data
over unfolded maps of the brain. The flat map is also considered as a parametric model for
the cortical surface; a model appropriate for applying geometrical transformations and apt
for surface based registration techniques. Although in this research the focus has been on
human cortical surface, the algorithms and methods are general and equally well
applicable to other objects.
Based on the fact that every conformal and equiareal mapping is isometric, we have
formulated the calculation of isometric mapping between surfaces as a constrained
optimization problem. We have designed an energy function whose minima occur when
the surface points are positioned in an unfolded configuration. Two constraint functions
imposing the requirements of preservation of angles and areas guarantee that the surface
will continuously deform, subject to a simultaneous conformal and equiareal mapping. A
conjugate gradient method minimizes the energy function, allowing the surface to
automatically unfold and converge to a flat plane.
To perform numerical processing on the surface we have defined appropriate
mathematical models to simplify the calculations as much as possible. For executing the
computationally intensive iterative process of conjugate gradient we have used a simple
first order approximation of the surface described as a triangular mesh. To create this
model we have used our tessellation algorithm that generates meshes with fewer triangles
xiii
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
than other commonly used methods such as the Marching Cubes algorithm. For
calculating intrinsic and extrinsic curvatures of the surface, where we need higher order
approximations, we have used local parametric approximation obtained by fitting
quadratic patches to the surface. We have applied these methods to computer generated
phantom data and realistic physical data and obtained satisfactory results.
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
c h a p t e r 1 Introduction
1.1 Complexity of The Brain
The brain is probably the most complex and the least known object in this universe.
Physically, it is a sophisticated entity, an approximately 1500 gram mass, containing on the
order of 101 1 neurons (the fundamental functional unit of nervous tissue), interconnected
by on the order of 1014 synapses (functional contacts between neurons). Structurally, it is
an intricate system whose circuitry and internal organization is not yet well understood.
Functionally, it is an elaborate machine, the central commander of all the activities of the
body with the mysterious capability of conducting the most complicated tasks. And in
2
appearance, it is a convoluted object, covered with a highly folded 2500 cm wide, 3 mm
thick layer of gray matter.
1.2 Development of the Cortical Folding
The mammalian cerebral cortex is formed from six profoundly convoluted neocortical
layers of gray matter surrounding a volume of fibrous white matter. The cortical foldings
are thought to be the result of the gradual evolution of two initially smooth hemispheres
trying to fit in a restricted volume. It is established that during the folding period the cortex
continuously grows. This growth is either a uniform growth which will increase the size of
the cortical surface but has no effect on its intrinsic geometry and shape, or is a differential
growth that may affect both [60]. One of the earliest models proposed to explain the
folding process of the cortex suggests that a cortex develops under the combined restraints
1
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
of uniform growth and minimal radial distortion and tends to fold along lines of minimal
curvature [57].
In spite of the existing controversy among neuroanatomists on the relationship between
the pattern of folding of the mammalian cerebral cortex and the functional and anatomical
structure of the brain, there is agreement on the fact that the deviation of the surface from
sphericity has a significant effect on several aspects of brain function[30]. Also, it is
generally accepted that the relative expanse of cortex in more functionally complex
mammals is associated with an increase in the number of functional units as opposed to an
increase in the complexity of units.
There is much evidence to support the theory that the cortex is subdivided into areas, each
with a specialized role. Increased specialization of cortical areas may account for the
necessity for a larger brain in human beings [42]. However, it should not be assumed that
all areas will be distinct and non-overlapping. Functions such as recognition, perception,
motor skills and planning are divided between several areas of the cortex.
1.3 Geometry of the Cerebral Cortex
The most interesting morphological features of the cerebral cortex in mammals are the
diverse and complex arrangement of their sulci and gyri. Sulci are the prominent fissures
visible on the cortical surface of a mammalian brain and gyri is the name used for
collective reference to the ridges that appear on the cortex. In the larger and more
complicated brains, the gyri and sulci exhibit elaborate, contorted and irregular three
dimensional convolution patterns. These convolutions increase the functional area of the
brain without increasing the size of cerebrum beyond a biologically feasible limit.
Despite their anatomic and functional similarities, even the gyri and sulci that consistently
appear in all normal anatomies exhibit pronounced variability in size and configuration.
Sulci and gyri are classified as either primary or secondary. Primary sulci are those which
are most consistent in position across individuals; they tend to be deeper and appear earlier
2
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
in cortical development. Secondary sulci exhibit more variation in position and
appearance [30].
The geometry of cortex can be studied from two perspectives: intrinsic and extrinsic. The
extrinsic geometry of the cortex determines its anatomical appearance and the shape of
white matter. The intrinsic geometry affects the relative position of functional areas and
the spread of activity within the surface itself. Intrinsic geometry is more fundamental
because it constrains the range of possible extrinsic properties. For the cortex, intrinsic
geometry has received less attention probably because it is more difficult to measure
practically or to calculate mathematically.
Previous studies of the intrinsic curvature of the cortex date back to the early 1980’s when
Van Essen and Maunsell reported the total amount of intrinsic curvature of cortex [60].
According to the Gauss-Bonnet theorem [22], the integral of Gaussian curvature over a
surface topologically equivalent to sphere is equal to 4 n . Mathematically, the two cortical
hemispheres are not topologically equivalent to a sphere due to the injection of the
callosum into the white matter. However, the perimeter of the hole is relatively small and
the theorem applies approximately. Thus the integral of Gaussian curvature is not an
appropriate measure for capturing the amount of intrinsic curvature of the cortex.
To measure the intrinsic curvature of the cortex, Griffin [30] used an alternative property
called compactness defined as the ratio between the mean geodesic of the surface and the
mean geodesic of a sphere of equal area. The mean geodesic of a surface is equal to the
average geodesic distance between pairs of points on the surface. He showed that the
cortex is compact (the compactness is less than one) and does not have the form of a large
crumpled bag but in fact has an intrinsic geometry considerably different from a sphere.
His experiment was a confirmation of the observations reported by Bok [5] on the human
brain and Todd [58] on the ferret brain about the significant variation of the intrinsic
curvature of the cortical surface. As we will see later in this thesis the Gaussian or intrinsic
3
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
curvature of the human cortex is a sharply varying quantity with large variance and mean
close to zero. This implies the possibility of developing a semi-isometric mapping
between the cortical surface and the plane.
1.4 Modeling the Cortical Surface
The surface of the human cortex is an incredibly complicated surface compared to the
objects that are normally subject to computer modelling. Computer based surface
modelling deals with objects that are either represented by a number of flat polygons or
made up of smooth patches. These smooth patches are often formed by combinations of
bounded parametric functions and make elegant and efficient descriptions for the surface
but can be difficult to design and manipulate. Alternatively, the polygonal representation,
while usually inefficient in terms of capturing the smoothness of the object, is
considerably easier to deal with numerically and much easier to create in an automated
way from sampled data.
Creating computer based representations of the human brain is an active area of research.
In general, two class of approaches to this problem exist: contour based and voxel based,
both of which use the data provided by scanning techniques such as Magnetic Resonance
Imaging (MRI) and Computed Tomography (CT). Scanning devices generate a stack of
slice images through an object. Each pixel in the image corresponds to some characteristic
at that point in space, for example proton density in MRI or X-ray absorption in CT scans.
These images together form a solid representation of the head. The contour approach
attempts to trace the outline of the characteristic of interest and use these contours to form
the surface. Surface reconstruction from voxel data sets typically use a variant of the
Marching Cubes algorithm [37],
1.5 Mapping the Human Brain
Recent developments of computer aided medical imaging, have made it possible to record,
display and manipulate 3-D models of functional activity in the living human brain, with
4
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
no apparent harm to the subject. However, the complexity and variability of cortical
convolutions make it difficult to accurately and systematically analyze these functional
neuroimaging data across individuals. The localization of high resolution functional
activation onto highly convoluted human cortex is most easily visualized using two
dimensional brain maps. By definition, 2-D maps are simplified representations of a 3-D
object and their usefulness is determined by their accuracy, clarity and ability to display a
particular type of information. Currently, the most accurate maps of the brain are those
generated section by section from the information abstracted from 2-D samples. These are
cryosection imaging based maps with very high resolution and rich information content.
The major limitation of this type of map is its inability to represent 3-D information,
which is a direct consequence of its two dimensional nature [4][31].
To overcome this, numerous 3-D visualization tools have been developed over the years to
facilitate viewing of the brain from different perspectives and manipulating it interactively
[59] [43]. Three dimensional methods are necessary when we want to reconstruct a model
of the brain with the size, shape and distances all preserved. While this approach can solve
the problem of displaying 3-D data, it can not present the information as accurately as the
cryosection maps.
A third, more abstract approach involves developing 2-D maps of the brain similar to
those of the earth used for summarizing and comparing different information such as
political boundaries, transportation routes and sociological data. To understand the
compromises needed to generate a flat map of the brain it is useful to compare it with the
simpler case of the earth. In cartography for creating a globe map, only one of the three
variables, area, angle or distance, can be maintained, leading to equiareal, conformal or
isometric maps. The basic approach involves placing a graticule (the network of lines of
latitude and longitude) on the surface of the original object and transforming it in a
systematic way onto a flat surface [54], Over 200 different projections of the earth have
5
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
been developed, none of them ideal, and all with certain advantages and limitations
depending on their intended use [55].
The surface of the brain is considerably more irregular than that of the earth, although in
principal it can also be transformed to a flat plane. This dissertation is about developing a
mechanism for performing such transformation.
1.6 Dissertation Goals
One of the major obstacles in performing quantitative analysis on the human brain is the
difficulty of defining a simple, accurate and robust mathematical model that is suitable for
the intended study. To study the functional organization of the brain it is essential to have a
correct representation of the cortex. At the same time, efficiently dealing with a complex
object requires the model to remain as simple as possible. The goal of this work is to
develop authentic 3-D and 2-D representations of the cortical surface. To achieve this goal
in particular, we have:
1. Developed an algorithm for creating a 3-D triangular approximation of the brain cortex.
2. Presented a method for calculating local differential properties of the cortical surface.
3. Defined an optimization problem for generating minimum distortion flat maps of brain.
4. Introduced a multi-resolution framework for representation of the cerebral cortex.
1.6.1 Summary of Main Contributions
A new Tessellation Algorithm: We develop an algorithm for building a triangular mesh
from a set of planar contours representing coordinate-constant slices of an object. We
show the advantage of this method over the Marching Cubes algorithm, an alternative
method commonly used for extracting boundary surfaces from volumetric data, by
comparing the results of both methods on phantom and realistic data.
Calculating the Surface Curvature: We present a method for quantifying the intrinsic
and extrinsic curvatures of the surface. We calculate the minimum, maximum, mean and
6
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Gaussian curvatures for phantom and real brain data and show that the primary folding of
the cortex appear along the lines of principal curvatures.
Map M aking as an Optimization Problem: We formulate the problem of making flat
maps as a problem in non-convex optimization theory in which the goal is to reach a flat
configuration of points with minimum distortion in area and angle and consequently in
distance. We define three cost functions. The first indicates the total convexity of a surface,
minimization of which results in an unfolded surface. The other two functionals measure
the total difference between areas and angles of the constructing triangles of two surfaces.
We evaluate the efficiency of this approach by generating flat maps for several phantoms
and real data sets.
Equiareal-Conformal M ap Instead o f Isom etric M ap: We take advantage of a well
known theorem in differential geometry to replace the quantitatively expensive and
practically challenging process of creating isometric maps of complex surfaces with the
equivalent task of generating a simultaneously equiareal and conformal map of the same
surface.
Uniformly Sampled Multi-resolution Surfaces: We use the 2-D isometric flat maps to
perform a uniform resampling on the original 3-D surface. Using the uniformly resampled
surface we have generated a multi-resolution representation useful for multi-scale
processing of the surface.
1.6.2 Overview of the Dissertation
In chapter 2, we describe our new tesselation algorithm and demonstrate its advantages
over the Marching Cubes algorithm. Chapter 3 is dedicated to parametric modeling of
surfaces. Here we present a method for local description of surfaces by quadratic
polynomials. In chapter 4, first we provide a brief overview of some basic concepts of the
7
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
differential geometry of surfaces. Then we calculate the minimum, maximum, mean and
Gaussian curvature maps of several surfaces including human cortex. Chapter 5 discusses
our strategy for creating flat maps. Mathematical definitions for three of the most
important types of map (equiareal, conformal and isometric) are presented followed by a
complete description of the optimization framework that we use for creating these maps.
Sample maps for developable as well as non-developable phantom surfaces are displayed
as evidence for efficiency of our method. To prove that the objective of this research has
been fulfilled, the isometric flat map of a typical cortical hemisphere is presented in this
chapter. As an example of the application of isometric maps, this chapter ends with images
of the uniformly resampled cortical surface in multiple resolutions. Finally, chapter 6
describes possible extensions of this work.
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
c h a p t e r 2 Tesselation o f Human Cortical Surface
Development of a mathematical representation of the brain surface suitable for automatic
feature identification is often a prerequisite for brain visualization, quantitative analysis of
brain properties, multimodal registration, and mapping and unfolding the cortical surface.
With the current state of brain imaging technology, MR imaging is the only noninvasive
technique that generates the anatomical details of the brain with a resolution high enough
for reasonable quantitative measurements of geometric properties of the brain. Extracting
the human cortical surface from MR images has been a topic of interest for many
researchers over the past two decades. A range of computerized techniques has been
proposed for this purpose with varying level of automation. Difficulties such as image
intensity inhomogeneities, extremely convoluted cortical structures, and topology pose
difficult challenges. Many of these techniques use an intensity based approach, and
attempt to determine appropriate thresholds for an initial segmentation. Some methods are
semi-automated and their role is only to provide an initial segmentation to assist a trained
expert to do the final labeling of brain structures [64]. Fully automated methods have also
been proposed. Some of them employ statistical image analysis techniques [62] [34],
others use morphological image processing [8][6][45]. There are still other methods that
take an indirect approach and register the brain to a reference atlas and then perform the
labeling through tissue matching with respect to the atlas [14].
9
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
2.1 Survey of Current Techniques
The immediate output of a brain surface extraction algorithm is an unorganized set of
points defining either the outer or inner boundary of cortical gray matter. The problem of
reconstructing a three-dimensional surface from a set of points is a major practical
obstacle in such diverse fields as medical imaging and computer aided geometric design.
The problems are not limited to finding the best interpolants, but also concern the
definition of the shape of the object, estimates of geometric properties such as area,
volume, curvature, and extraction of elementary shapes. The difficulty of the problem
increases with the complexity of the shape of the object. As with any other practical
application, computational resources are always a limiting factor in computational
complexity.
Solutions to this problem fall into one of the two broad categories of volume based and
surface based methods. Volume based approaches assume that the data are available as a
set of discrete values defined on a 3-D grid. Several methods have been proposed for
surface reconstruction based on such an arrangement of data, the most popular of which is
the Marching Cubes algorithm [37] and its extensions. The Marching Cubes technique
compares the intensity values at eight contiguous voxels from two adjacent slices forming
a cube with a predefined isovalue to determine how the surface intersects this cube.
Trilinear interpolation along each of the voxel edges is used to find the vertices of the
triangular patch that forms part of the isosurface passing through a given voxel. The
Marching Cubes algorithm can be applied directly to the original brain MRI; however, due
to the limited efficiency of the threshold based schema it uses for determining the location
of surface points, it is often preferable to generate the brain surface data by using a custom
designed segmentation method. Marching Cubes is then applied to the resulting binary
level image defining the desired brain volume.
10
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
One obvious problem with Marching Cubes is the large number of vertices, and
consequently triangles, needed to represent the surface. As each boundary cube can
generate up to 4 sub-pixel facets, the resulting set of triangles can be quite large. The
original algorithm proposed in [37] has several ambiguous cases which can lead to holes
in the final image [24]. These ambiguities occur either when the wrong decision is made
about whether a vertex is inside or outside a surface, or if the triangles chosen do not fit
together properly. A number of extensions have been proposed to the Marching Cubes
algorithm to avoid the ambiguous cases. These include marching tetrahedra [49] and
dividing cubes [13].
The input data to a surface based approach are a series of planar contours corresponding to
the intersections of the object surface with a set of parallel planes. A polyhedral
approximation of the surface is obtained by linking each set of contours from two
consecutive slices with triangular patches. The desired surface is the best approximating
polyhedron. The difference between the methods in this category lie in the definition of
the best approximation. Different metric functions have been used to distinguish good
surfaces from bad [50][51]. Some methods optimize a cost function based on a graph
describing the set of all possible arrangements of triangular patches to determine the
desired surface. In this formalism a given path in the graph defines a possible solution. The
optimal path may be found by maximizing the volume [35] or minimizing the area [28] of
the polyhedral representation of the surface, or by optimizing some global figure of merit
assigned to individual triangles through a relaxation process [61]. A variety of other
metrics for the graph cost function have been used, each claiming to improve some “good”
feature of the surface [40][51]. These optimal methods are generally slow and time
consuming. Other methods take advantage of less computationally expensive heuristic
methods to create visually acceptable surfaces. They define triangular patches one by one
using a local decision criteria [12][15][16][29]. Heuristic methods can generate incorrect
11
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
results if the contours are very different in shape, orientation and alignment. Furthermore,
most of these algorithms fail to handle ambiguous cases that result from having multiple
contours in a single section (correspondence ambiguity) or different numbers of multiple
contours in adjacent sections (branching ambiguity).
To solve the correspondence problem some methods make the assumption that the data is
sampled densely enough that by examining the overlap between contours on adjacent
sections, unambiguous solutions are found. More general methods incorporate high level
information about the shape of the objects in the reconstruction process [52][7]. Proposed
solutions for the branching problem include algorithms that require user interaction
[12][65] as well as methods that compute Delaunay triangulation of the set of vertices
belonging to the contours in adjacent sections [3]. A problem related to the branching
problem is tiling a pair of contours with very different shapes. The proposed solution to
this problem consists of creating an interpolated contour between the two adjacent slices,
and thus reducing the dissimilarity between contours to be tiled [44] [25].
2.2 Triangulation Algorithm Based on Coordinate Curves
The common strategy in surface based triangulation algorithms is to reconstruct the
external surface of the object using a set of contour points extracted from one series of its
parallel planar sections. The polyhedral approximation of the surface is obtained by
linking contour points in consecutive sections with triangular patches. These algorithms
were originally developed to conform with the structure of the data acquired by CT and
MRI techniques in which a three-dimensional image of a body’s structure is constructed
by computer from a series of plane cross-sectional images made along an axis.
Representing a three dimensional object by a series of two dimensional sections hides the
information about the connectivity of sections along the imaging axis; however, it
provides sufficient data for finding points located on the surface of the object and their
neighborhood relations in the plane of imaging. Since the volume image data acquisition
12
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
is based on a rectangular grid, one can always rearrange the data set to effectively change
the imaging axis and find the neighborhood relations in all three sets of mutually
orthogonal planes defining the imaging grid.
Mathematically, a point located on a surface with a parametric description of S(u,v) is
uniquely specified by two constant parameter (coordinate) curves passing through it. In a
rectangular coordinate system, the contours extracted from the cross sections of an object
with planes normal to any of the three axes correspond to coordinate curves. Thus each
surface point is defined as the point of intersection between contours belonging to two
perpendicular planes. We have developed a two step algorithm for triangulating the
surface that is based on this property.
2.2.1 Contour Tracing
In a cross-sectional image of a three dimensional object, boundaries of regions correspond
to intersections of the object’s surface with the plane of sectioning. The connectivity
structure of the surface is directly related to how the boundary points are defined. If an 8-
neighborhood system is used to define boundary points, the resulting contours will be 4-
connected, and correspondingly, boundary points defined based on a 4-neighborhood
system result in 8-connected contours (Figure (2.1)).
Figure 2.1. 8-Connectivity vs. 4-Connectivity, (a) Sample shape, (b) Boundary points extracted based on
4-neighborhood system are 8-connected. (c) Boundary points extracted based on 8-neighborhood system
are 4-connected.
13
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
The union of the three sets of 8-connected contour points
extracted from sections perpendicular to coordinate axes
constitutes an 18-connected surface. Figure (2.2) displays the
eighteen closest neighbors of a point located in the center of the
3 x 3 x 3 cube.
By tracing the contours in a consistent manner (either in clockwise or counter clockwise
direction) the two closest neighbors of each surface point in the plane of the contours are
determined. Where a point is located in the intersection of multiple contours in a single
plane, more than one pair of closest neighbors are found for that point. This is the situation
where two constituting elements of the object are tangent to each other. By separating the
neighbors into two pairs we guarantee that in the tiling step the correct surface
representation of each element will be constructed.
A typical surface point has three pairs of closest neighbors lying on three mutually
orthogonal planes that uniquely localize the point in space. A triangle formed by
connecting the point to two of these six neighbors is the smallest possible patch, within the
limits of the image resolution, that one can use to tile the surface. The surface constructed
from such triangles will have a resolution corresponding to that of the original image.
2.2.2 Tiling
Surface points can be divided into two groups depending
on the number of contours passing through that point.
Those that are in the intersection of contours lying on two
perpendicular planes and those that are in the intersection
of contours belonging to three mutually orthogonal
planes.
Points in the first group are located on planar regions to which one of the coordinate axes
is locally orthogonal. This group of points have only two pairs of closest neighbors to
14
Figure 2.3. Points with only
two pairs of neighbors are tiled
with four equal right triangles.
□■
■
Figure 2.2. 18 closest
neighbors of a point.
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
which they can be linked with four equal right triangles. As an example, Figure (2.3)
shows the point (X0,YQ ZQ ) that is shared between a constant-X and a constant-Y curve and
is located on a patch that is contained in the (Z = ZQ ) plane. The four nearest neighbors of
this point are (X0 ± 1,F0, ZQ ) and (X0, Y0 ± 1, ZQ ) .
The second group consists of points that have
at least three pairs of closest neighbors on the
surface. A point of this type can form up to
twelve triangles when linked to all its
neighbors (Figure (2.4.a)). In some locations
one of the neighbors in one pair may coincide
with a neighbor in another pair. Such an arrangement of points will result in less than
twelve triangles (Figure (2.4.b)).
One can not simply join all the points to their neighbors with triangular patches to tile the
surface since this would result in many redundant triangles and doubly covered regions.
Instead, by orderly tiling different regions of the surface with triangles selected based on
some constraints, a relatively uniform triangulation of the surface can be obtained.
The unit tiling patch is composed of two triangles with a
common side (Figure (2.5)).When a point is to be connected to
two of its neighbors, this unit patch is fit to the surface by
superimposing the common side of the triangles on the edge
linking the neighbors of that point. To avoid double coverage of the surface, after fitting a
unit patch to a set of four points no other patch will be allowed to cover that set.
Type (1) Type (2)
Figure 2.5. Two types
of unit tiling patch.
Figure 2.4. (a) A point with three pairs of
neighbors forms 12 triangles, (b) A point with
overlapping neighbors forms less than 12
triangles, neighbor points Y' and Z' are
overlapped and only 8 triangles are formed.
15
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
In Figure (2.6) we have shown a small region of a
surface to explain how the tiling algorithm works at a
typical point A. First we pick two neighbors of A,
namely B and D. We then find their other common
neighbor, C. We cover the patch limited to these four
point with triangles ABD and CBD and label this patch as processed. When we move to
point B, neighbors A and C will be ignored since they have already been processed. The
same rule applies to point C with neighbors B and D and point D with neighbors A and C.
Depending on the order in which the points are being processed, the shape of the
triangulation may change. For example if point B was processed before point A then
ABCD would be tiled with a unit patch of type(2) instead of type(l) (dotted triangles in
Figure (2.6)).
However, this does not have any effect on the
shape of the constructed surface in planar areas
where all the four points are coplanar. In other
areas the resulting change in the shape of the
surface is on the same scale as the voxel size and
either result could be considered “correct”.
As we described in Section (2.2.1), surface points
located on planar patches parallel to coordinate
planes (normal to coordinate axes) have only two
pairs of equidistant neighbors. These points, when tiled with unit patches, create a surface
made of regularly arranged equal right triangles. Figure (2.7) shows a regularly
triangulated cubic surface constructed with this algorithm.
Figure 2.7. Tessellation of a cube
using coordinate curves.
Figure 2.6. Tiling the surface with
a unit patch at a typical point.
16
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Once the entire surface is tiled all of the triangles need to
be checked for orientation consistency. The vector normal
to the plane of a triangle represents the orientation of the
surface at that location and it must point outward with
respect to the object. Since our tiling operation is a local process and the outward direction
is determined using global information, for each point on the object we have to calculate
and store the outward pointing vector before beginning the tessellation. Alternatively, we
can select the simpler option of reorienting wrongly oriented triangles. The normal vector
to the plane of a triangle is calculated by taking the cross product of the vectors formed
from two sides of the triangle, therefore the normals to two adjacent triangles sharing a
side will both point in the same direction if vectors representing the common side are in
opposite directions (Figure (2.8)). Starting from a triangle whose normal is pointing
outward, we find all adjacent triangles and if necessary change their orientations so that
every pair of adjacent triangles have consistent orientations as described.
2.3 Results and Discussion
To evaluate this method we compared its performance with the Marching Cubes algorithm
[37]. To compare the ability of the two methods to reconstruct surfaces with sharp comers
vs. surfaces with smooth curves we used a phantom made of a hemisphere attached to a
cube. Figure (2.9.a) shows the surface generated by our method consisting of 2894
vertices and 5784 triangles. Figure (2.9.b) shows the output of the Marching Cubes
algorithm. This surface is made up of 3782 vertices and 7560 triangles. In Figure (2.9.c)
we have resampled the object to make the number of surface points used in our method
close to the number of vertices generated by the Marching Cubes algorithm. The
resampled object resulted in 4130 vertices and 8256 triangles. The Marching Cubes
method rounds right-angled edges whereas our method preserves them. However, when
constructing smooth surfaces our method compared to Marching Cubes creates a rougher
17
i i
J
y
Figure 2.8. Adjacent triangles
with outward pointing normals.
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
surface. This is due to the smaller number of vertices and triangles. By increasing the
number of vertices we observe that with a comparable number of triangles, our methods
creates a smoother surface.
(a) (b) (c)
Figure 2.9. (a) Phantom surface constructed from coordinate curves (2894 vertices, 5784 triangles), (b)
Phantom surface tessellated using Marching Cubes algorithm (3782 vertices,7560 triangles), (c)
Resampled phantom constructed from coordinate curves (4130 vertices, 8256 triangles).
The fact that we use coordinate curves to determine the location of surface points implies
that we should expect to see a staircase shape for the reconstructed surface of objects with
planar faces that are not positioned parallel to any of the coordinate axes. To examine if
the same effect will be observed when the object’s surface is reconstructed by the
Marching Cubes method we applied both methods to a rotated version of the phantom
used in the previous study. The results are displayed in Figure (2.10) and Figure (2.11). It
can be seen that both methods suffer from the staircase effect.
18
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 2.10. Reconstructed surface of the rotated Figure 2.11. Reconstructed surface of the rotated
phantom using the coordinate curves method. phantom using the Marching Cubes algorithm.
To compare the outputs of these methods for real medical images we used a data set
representing the cortical surface of a human brain extracted from 150 axial magnetic
resonance images of the head of a normal subject, acquired with a total of 256 x 256
voxels per slice. After tracing the contours in transaxial, coronal and sagittal sections, a
total of 73814 points were found to represent the outer surface of the brain. These surface
points were then linked with 147645 triangles to create the surface displayed in Figure
19
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
(2.12.a). For the same object, the Marching Cubes algorithm generated 128984 vertices
and 258004 triangles (Figure (2.13.b)). Using the same parameter settings for rendering
the two surfaces one can not observe any major degradation in the visual quality of the
surface generated by our method. Given the significant decrease in the number of vertices
and triangles without substantially compromising the quality of the resulting surface this
method is a potentially attractive alternative to the Marching Cubes algorithm.
i V A ’ iv r
, ’ . € ■
...... . w n# -Vw ;
^ ^ “ 5«;^saP 9 • 1 ‘ -J f-
s W i T i T , 7 ! < „ in * *
i
. * ? 4 1 '
' J« iaM § iafi8 3 5 ‘ -
i r " % 5 0 h . k « •
tu * u
, '« 'O b , «
M ! 4 1
'%S jV ’ ^p
Figure 2.12. Human brain surface reconstructed
using coordinate curves. No. of vertices: 73814
No. of triangles: 147645.
:5
H
Figure 2.13. Human brain surface reconstructed
using Marching Cubes. No. of vertices: 128984
No. of triangles: 258004.
20
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
c h a p t e r 3 Parametric Modeling o f Human Cortical Surface
Differential geometry is the mathematics of using calculus to study geometric properties
of curves and surfaces. Although biological surfaces (and in particular the cortex) have a
non-zero thickness and so are not true surfaces, considering the relatively constant small
thickness (3-4 mm) of human cerebral cortex compared to its large surface area (2000-
2500 cm2), application of the techniques of differential geometry in the analysis of human
cortical surface can be shown to be useful [30][57]. In order to appreciate geometric
properties of the brain and its effects on cortical architecture and function we need to
develop a precise mathematical representation adaptable for differential analysis of the
cortex. This representation allows us to make statements about the geometry of the surface
and study the variability of the cortical surface in an ensemble of brain data.
3.1 Survey of Current Techniques
The problem is to fit a mathematical model to a surface given as a piecewise linear
function representing a collection of triangular faces pasted together along their edges. For
a detailed description of the available methods for solving this problem refer to [41]. Here
we present only a brief overview of the general approaches to this problem.
Two general types of surface fitting algorithms exist, interpolation and approximation. In
interpolation the constructed surface satisfies the given data precisely and the number of
surface pieces is approximately the same as the number of data points. A triangular mesh
21
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
is the lowest order interpolating model for a surface. As the number of data points
increases the space and computational requirements of interpolated shape representations
become a major disadvantage. In approximation, we construct surfaces that do not
necessarily satisfy the given data precisely, but only approximately. In this case it is often
desirable to specify a maximum bound on the deviation of the surface from the given data,
and to specify some constraints, i.e. data that is to be satisfied precisely.
Most fitting algorithms also fall into one of two categories: global or local. In a global
algorithm, a system of equations or an optimization problem defined collectively in terms
of all the data points needs to be solved. Theoretically, a perturbation of any one input data
item can change the shape of the entire surface; however, the magnitude of the change
decreases with increasing distance from the affected data item. Local algorithms are more
geometric in nature, constructing the surface segment-wise, using only local data for each
segment. A perturbation in a data item only changes the surface locally. These algorithms
are usually computationally less expensive than global methods.
3.2 Local Fitting of Quadratic Patches to a Triangular Mesh
To study the differential geometry of human cerebral cortex we need to measure its local
properties. That means those properties that depend only on the behavior of the surface in
the neighborhood of a point. The mathematical representation resulting from a local
surface fitting method is adequate for studying these properties. To use methods of
differential calculus, the function describing the surface must satisfy continuity and
differentiability requirements appropriate to the property under study. Our goal is to
measure different curvatures of the human cortical surface including minimum, maximum,
mean and Gaussian curvature. A model that can present meaningful values for the
curvatures of human cortex must at least belong to the class of smooth regular surfaces of
? 3
order two (C ), i.e the surfaces defined by the mapping X(u, v):U — > R from an open set
2 3
U c R into R such that all the partial derivatives of order two or less are continuous in
22
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
U , and Xu x l v ^ [0 ,0 ,0]? for all (u, v) e U , where Xu and X v are partial derivatives with
respect to u and v and x is the cross product [22].
To keep our model as simple as possible we have chosen a model satisfying the minimum
requirements based on the schema described by Hamann [32]. At each vertex in the
triangular mesh, a quadratic function is fit by locally approximating the parametric surface
by using terms up to order two in its Taylor series.
We assume that a triangulated mesh is constructed from the contours representing the
3
surface using the algorithm described in Chapter 2. Let X i e R , i = 0 ,1 ,..., N - 1
represent the vertices of the mesh and 9 ^ (X; -) the set of all Kth -order neighbors of vertex
X-. We define the Kth -order neighbors of a point Xf - as all the 1st -order neighbors of its
(k - 1 )th -order neighbors that are not common with its (k - 2)th -order neighbors, with the
1st-order neighbors defined as the immediate neighbors of X( - and each point defined as
its own 0th -order neighbor.
At each vertex, X; -, we calculate the surface normal, » ., by averaging the normals to the
planes of triangles connected to that vertex, i.e.
' N ^ j = o \\ajXbj\
(Eq. 3.1)
where tij and 6. are the vectors representing the sides
of the triangles connected to vertex X( . and N[ is the
total number of such triangles. The order of the vectors
is selected such that the cross product always points
outward with respect to the surface (Figure (3.1)).
The tangent plane Tp at a point X. of the surface is
defined by nj ■ (X-X{) = 0 where ni is the unit
surface normal at Xt. We define an orthonormal coordinate system in Tp as follows:
Figure 3.1. Surface normal at a
typical point of a triangulated mesh.
23
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
First, based on the outward unit normal vector n( -, we define a vector mi perpendicular to
n i according to
m i =
J T l - ( n iy + n iz ) ’ n ix’ n ix t
ix
^ nir - ( nix + niz^ n iy^
iy
^ - l n iz’ n iz’ - ( n ix + n iy r f ’
iz
i f n. ± 0
i f n iy * 0
i f i*. * 0
(E q . 3 .2 )
and then form the coordinate system using three orthonormal vectors (figure(3.2)):
ni p i =
t a
rn i
Q i = n i x P i
(E q. 3 .3 )
/
r ]
2 n
C f = a r g m in ^ X j-\x i + U j P + VjQl + ([Uj, Vj]C[Uj, Vj] )n.
^ C Z e ^ ( Z )
1 J
/
The quadratic patch approximating the parametric surface passing through the point Z ( .
can be written as:
X ( u , v ) = X i + u P i + v Q i + ([u, v ] C ( [ m, v]‘) n i (Eq.3.4)
where C; - is a 2 x 2 symmetric matrix. The unknown matrix C( - is determined by solving a
minimization problem defined as:
f 1 2"
(Eq. 3.5)
where e 9 ^ (X-) represents all the Kth -order neighbor points of vertex X - , with
J I I
P t
K > 1. Let X j = [uj, Vj] be the projection of point X- into the tangent plane Tp . The
p
local coordinates of the points X- in the tangent plane are determined by:
[Uj, V j f = [ t f - AT,.) • P t, (Xj - X t) • Q t\ (Eq. 3.6)
We can restate the minimization problem in terms of distances of the points in the
neighborhood 9 ^ ( X t) from the tangent plane. We denote such distances with h j , i.e.
P P
hi = ( X j - X - ) • n i . Since X. = X- + + v-Q ■ , we can write
J J J 1 J 1 J 1 J 1
24
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
c , . = arg
By introducing matrices:
min X \hj ~ ^ U P vj \ C ' [ uP vj ^
v C i
(Eq. 3.7)
u0 2m0v0 v0
C,i
h0
.C f =
C1 2
, and Hi =
“At,., 2M lV i_1 V Ni_, vNi_i
C22
kNi_l
we can use matrix algebra and rewrite the problem as
(
Ci = arg (Eq. 3.8)
The minimum mean squared estimate of C( - can be easily found by solving the following
normal equations [53]:
(Eq. 3.9)
Figure 3.2. The local rectangular coordinate system in the neighborhood of a point X± .
25
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
To study the accuracy of the surface description obtained based on this method we used
the spherical phantom displayed in Figure (3.3). We selected a small region of the
phantom (Figure (3.3.b)) and fit a quadratic patch to it. The center of patch is marked with
a black dot in Figure (3.3.b). As it can be seen in Figure (3.3.c) and (d) the accuracy of the
quadratic patch in describing the surface decreases with distance from the center of the
patch.
(c) (3)
Figure 3.3. Fitting a local quadratic patch to a sphere, (a) Analytically defined spherical phantom,
(b) Magnified display of a small region of sphere, (c) The same region approximated with a local quadratic
patch, (d) Superimposition of the original and approximated patches.
26
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
c h a p t e r 4 The Curvature o f Human Cortical Surface
4.1 Introduction
In this chapter we want to study local geometric properties of the cortex. In other words,
our focus will be on the properties that do not pertain to the geometric configuration of the
surface as a whole but depend only on the form of the configuration in an arbitrary small
neighborhood of a point under consideration. Among many local geometric quantities
defined for the surface, curvature information is the most valuable since it quantifies the
structure of the sulci and gyri, thus providing the basis for comparative studies.
Curvature is a local property indicating a measure for how rapidly a surface pulls away
from the tangent plane in a neighborhood of a point. In what follows we will introduce the
relevant concepts (principal curvatures, principal directions, Gaussian and mean
curvatures) without using local coordinates to emphasize the geometric content of the
definitions. However, for computational purposes, it is important to express all concepts in
local coordinates.
Throughout our discussion we assume that X(u, v) = (x(u, v), y(u, v), z(u, v ) ) : U - > R is
the parametric representation of the surface defined on U, an open subset of R with
coordinates u and v and it satisfies all the required conditions. We therefore make the
following assumptions:
27
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
( 1). X is differentiable. This means that the surface X(u, v) belongs to the class of smooth
surfaces of order r, Cr (r>l), i.e. the functions x(u,v), y(u,v), z(u, v)are r times
continuously differentiable in U .
(2). X is a homeomorphism. Since X is continuous this means that X has an inverse X~l
which is continuous.
2 3
(3). X is regular. This means that for each P e U , the differential dXp :R — > R is one-
to-one or equivalently the Jacobian matrix
J =
is of rank 2 in U . Still another way of expressing this condition is that ^ x ^ * 0 , which
means there is no singular point on the surface i.e. a unique unit surface normal always
exists at each point of the surface.
4.2 Some Basic Concepts of Differential Geometry of Surfaces
There are metrics for a surface that can be calculated without leaving the surface.
Geometrically, these metrics describe quantities such as lengths of curves, angles of
tangent vectors and areas of regions that can be measured on the surface without referring
to the ambient space R where the surface lies. Such properties are based only on the first
fundamental form of the surface.
DEFINITION 1. The infinitesimal distance element between two points on a curved
2 2 2
surface is given by ds = X u ■ X ydu + 2Xu • X ydudv + X u • X ydv and is called the first
fundamental form of the surface. Furthermore, the following quantities are called the first
fundamental or metric coefficients of the surface.
28
dx dx
du dv
dy dy
du dv
dz dz
du dv
(Eq. 4.1)
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Ex (u, v) = X u ■ X u
Fx (u, v) = X u • X v
Gx (u, v) = X y ■ X v
(Eq. 4.2)
(Eq. 4.3)
(Eq. 4.4)
The first fundamental form gives the distance ds between neighboring points (u, v)and
(u + du,v + dv) on a surface, to first order in du and d v . The distance element ds lies in
the tangent plane of the surface at (u, v)and therefore yields no information on how the
surface curves away from the tangent plane at that point. To investigate surface curvature,
we must examine the displacement between neighboring points (u, v) and (u + du,v + dv)
to second order in du and dv. This is what can be determined from the second
fundamental form of a surface.
DEFINITION 2. The projection of the second order infinitesimal displacement between
two points of a curved surface on the unit surface normal is equal to one half of the second
2 2 2
fundamental form of the surface given by d h = n • X uudu + 2n ■ X uvdudv + n ■ X vvdv ,
in which n represents the unit surface normal. The following quantities are called the
second fundamental coefficients of the surface and form the basis for defining and
analyzing the curvature of a surface.
We are going to study the geometric shape of a surface X in the neighborhood of a typical
point P. We consider all surface curves C:(u = u(t), v = v(t)) passing through P and
denote their curvatures by k = k ( t ) . We use symbol y = y (t) to represent the angle
between the unit principal normal vector N(t) to the curve C and the corresponding unit
normal vector n(u(t), v(t)) to the surface X . The family of planes containing the normal n
at the given point P cut the surface in a family of normal section curves for that point.
Lx (u, v) = n ■ X uu
Mx (u, v) = n ■ X uv
Nx (u, v) = n ■ X yv
(Eq. 4.5)
(Eq. 4.6)
(Eq. 4.7)
29
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
DEFINITION 3. The curve of intersection of the surface X and a plane passing through
both the tangent to C at P and the normal to the X at P is called the normal section curve
of X at P .
The curvature of all curves C with a common
tangent direction at P , depends only on the
angle y between their unit principal normal
vector N at P and the corresponding unit
normal vector n to X . The length of the
projection of the vector kN over n is called the
normal curvature of C (Figure (4.1)).
kN
Figure 4.1. Normal Curvature at a point.
DEFINITION 4. Let C be a regular curve in X passing through P . The number
Kn - Kcosy is called the normal curvature of C at P where k is the curvature of C at P
and y is the angle between the unit principal normal vector N of the curve C at P and the
corresponding unit normal vector n to the surface X at P .
DEFINITION 5. The maximum normal curvature and the minimum normal curvature
k2 are called the principal curvatures at P .
DEFINITION 6. The product of the principal curvatures is called Gaussian curvature and
their average is called mean curvature.
4.3 Quantitative Analysis of Surface Curvature
At a given point for a specified direction (u, v)the normal curvature can be expressed in
terms of the first and second fundamental coefficients of the surface as
k = -
• 2 • • -2
Lu + 2Muv + Nv
72 7 7 7i
Eu + 2 Fuv + Gv
(Eq. 4.8)
30
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
The radius of curvature (the radius of the circle that best approximates the normal section
curve) is R - |k|_1 . The sign convention in the above equation gives positive curvature if
the center of the curvature lies on the opposite side of the surface with respect to the
surface normal (convex) and negative curvature when both lie on the same side (concave).
At each point Equation (4.8) associates a curvature k with each direction (u, v)on the
3k
surface. The direction in which k attains extremum values occur when — = Oand
du
die
— = 0 . These conditions can be written as
dv
(L + KE)u + (M + KF)v = 0 (Eq. 4.9)
(L + K.E)u + (M + KF)v = 0 (Eq. 4.10)
For these equations to possess a consistent solution k must satisfy
K2 - 2 H K + K = 0 (Eq. 4.11)
where the Gaussian curvature K and mean curvature H are defined by
K = L N - M (E q 4 12)
E G - F
„ 2FM - (EN + GL)
H = ------------i ( E q . 4.13)
2 ( E G - F )
The two solutions to the quadratic Equation (4.11) are the extremum values of the
curvature as a function of direction which are called the principal curvatures. They
represent the upper and lower bounds on the curvature at a given point:
Kmax = H + J h 2 - K (E q.4.14)
Kmin = H - j H 2 - K (Eq. 4.15)
The Gaussian and mean curvatures introduced above are readily seen to be the product and
average of the principal curvatures, respectively:
K = KmaxKmin < Etl-4-16>
H = l / 2 ( Kmax + Kmin) ™ 7)
31
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
When H2 = K , Kmax = Kmin and k is independent of direction. Such a point is called an
umbilical or spherical point since the surface locally approximates a sphere at that point.
When the principal curvatures have the same sign (as in a hill or pit) the Gaussian
curvature is positive and the point is called elliptic. Hyperbolic points are those points that
have negative Gaussian curvature or principal curvatures with opposite signs (as in saddle
points). When either one of the principal curvatures is zero the Gaussian curvature is zero
(as in ruled or cylindrical surfaces). Surfaces that have zero Gaussian curvature
everywhere are developable, i.e. they can be laid flat on a plane without stretching or
tearing. If both principal curvatures are zero the point is called planar but if only one of
them is zero it is called parabolic.
4.4 Curvature Maps
A convenient method for displaying the variation of a scalar quantity over a surface is by
means of a color-coded map, an even gradation of color corresponding to a range of values
for that quantity.
The three primary color components of each pixel are determined by applying some linear
mapping to the value of the scalar quantity associated to that pixel. The linearity of the
mapping is desired because it makes computations such as color interpolation easier.
Suppose we wish to interpolate from color A to color B , we might write
C = (1 - a)A + aB and sweep a from 0 to 1. This is the typical way that Gouraud and
Phong shading are implemented for rendering of triangulated surfaces. We would like
equal increments of a to result in steps of C that were of perceptually equal size. This
simple interpolation scheme works only if the mappings between the value of the scalar
quantity and the red, green and blue color components of each pixel are linear.
32
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
By using the same mapping function for the three color components the resulting map will
be a grey scale image with different shades of grey corresponding to different values of the
scalar quantity.
We have generated different curvature maps for several objects including synthetic
phantoms and real human brain. The surface that we start with is reconstructed from the
slice images of the objects using the tessellation algorithm introduced in Chapter 2. To
calculate the curvatures we fit a local quadratic model to the surface patch in the
neighborhood of each node of the tessellated surface as was described in Chapter 3.
Following the notation used in Chapter 3, let X(u, v) be a parametrization at a point
X t e S of a surface S given by
X(u, v) = X i + uPi + vQ i + {[u, v]Ci[u ,v f)n i (Eq.4.18)
The first and second fundamental coefficients of the surface can be calculated from this
representation using the Equations (4.2)-(4.7). According to the equations o f Weingarten
[22] the principal curvatures K j and k2 of the surface are defined as the eigen values of
the matrix
- i - i
-A =
L M E F
M N _ _ F G
and it can be shown [33] that this matrix is equal to the coefficient matrix of the second
order terms in Equation (4.18), i.e.
-A = 2 C- (Eq. 4.20)
So, in order to calculate the minimum and maximum curvatures at each point we find the
eigenvalues of the corresponding matrix C{ at that point, and then we use Equations (4.16)
and (4.17) to calculate the mean and Gaussian curvatures.
33
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
4.5 Experimental Results
To examine the accuracy of this method in calculating different curvatures of a surface we
have tested the algorithm on several phantoms. All phantoms were designed to provide the
same type of data set that we expected from a real cortical surface. To generate a phantom
first, a binary volume composed of a stack of cross-sectional images is created. This
volume represents a discretized approximation of an object with a well defined analytical
description that can be used to verify the results of our calculations. The phantoms are
then tessellated using the algorithm described in Chapter 2. The tessellated object is then
slightly deformed by adjusting the location of its vertices to reduce the difference between
(a) (b) (c)
Figure 4.2. Sphere phantom, (a) Binary volume, (b) Tessellated Surface. (3) The nodes in the triangulated
sphere are repositioned to new locations that conform with the analytical description of an exact sphere.
the discretized shape and the continuous form of the object. In contrast to a discretized
volume representation of objects in a rectangular grid, in a mesh representation the
location of nodes does not necessarily have to be on the grid points. Figure (4.2.a) shows a
tessellated discretized spherical phantom in rectangular grid. Figure (4.2.b) shows the
same phantom, and the modified tessellated phantom that conforms with an analytically
modeled sphere is displayed in Figure (4.2.c).
Figure (4.3) shows minimum, maximum, mean and Gaussian curvature maps of a sphere.
A grey scale colormap is used to display the variation of the curvatures. As can be easily
34
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
verified, the difference between calculated and mathematically derived values of the
curvatures of the sphere are small. Given that the values of all curvatures should be
constant (minimum curvature: l/R, maximum curvature: 1 /R, mean curvature: l/R,
Gaussian curvature: l/R , where R is the radius of sphere) a close investigation of the
variation pattern of these curvatures indicates that the error is primarily due to the
deviation of the shape of the phantom from a real sphere.
M aximum C u rv a tu re . S p h e r e (r-25) M inimum C u rv a tu re . S p h e re (r-25)
00407
0 0 * 0 6
0.0408
0.0404
0.0407
jo 0406
0.0406
0.O4O2
0.0406
0.0405 [0.0401
(a) (b)
M ean C u rv a tu re . S p h e r e p - 2 5 )
1 0 ■ 1 0
0.0406
i 0.0407
0.0406
i 0.0406
! 0.0406
0.0405
1 0.0404
1 0.0404
1 0.0404
10.0403
G au ssian C u rv a tu re , S p h e r e (r-25)
-
%
(c)
Figure 4.3. Curvature maps for a sphere with a radius of 25 units, (a) Minimum curvature map. (b) Maximum
curvature map. (c) Mean curvature map. (d) Gaussian curvature map.
The second set of experiments were performed on phantoms specifically designed to
simulate cortical folds. To study the variation of different curvatures in the neighborhood
of a depression and a protuberance in a surface we created two new phantoms by making
35
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
several dents and bumps in our spherical phantom. In Figures (4.4) and (4.5) we have
displayed the minimum and maximum curvature maps of these two phantoms,
respectively. Comparing the curvature maps of the indented sphere with those of the
bumpy one, we can observe that curvature values in the neighborhood of a dent are always
negative whereas those corresponding to a bump are positive. One can justify this fact by
noting that the normal curvature is proportional to Cosy where y is the angle between the
unit principal normal vector of a curve passing through a point and the corresponding unit
normal vector to the surface at that point [22], Since a dent and a bump of equal size are
mirror images of each other with respect to the surface of the sphere, the principal normal
vector of a pair of corresponding curves passing through a pair of corresponding points -
one on a bump the other on a dent - will point in opposite directions. However, since the
normal vector to the surface is always pointing outward there will be a 180 degree
difference between the values of the angle y for a dent and a bump which accounts for the
negative sign.
We can also verify that the absolute value of the curvature depends on the size of the bump
or dent. The mathematical explanation of this fact is based on the notion of radius of
curvature. At each point of the surface one may fit circles to the point and two of its
neighbors such that these circles lie in the normal sections of the surface. The maximum
and minimum curvature of the surface at a given point is equal to the inverse of the
maximum and minimum radii of such circles passing through that point. As the size of the
bump or dent becomes smaller the radius of the circles that can be fit to it decreases,
causing an increase in the maximum and minimum curvature and consequently the mean
and Gaussian curvatures.
36
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Minimum Curvature (sphere: R=25, Bump: r=10) Minimum Curvature (Sphere: Rs25, Dent: rs10)
r
i #
(a) (b)
Figure 4.4. Relative variation of the minimum curvature in the neighborhood of a dent and a bump, (a) In
the neighborhood of a dent, the minimum curvature takes negative values, (b) In the neighborhood of a
bump, the minimum curvature takes positive values. Minimum curvature is a good feature for locating dents.
Maximum Curvature (Sphere: R=25, Bump: rs10) Maximum Curvature (Sphere: Rs25, Dent: r»10)
(a) (b)
Figure 4.5. Relative variation of the maximum curvature in the neighborhood of a dent and a bump, (a) In
the neighborhood of a dent, the maximum curvature takes negative values, (b) In the neighborhood of a
bump, the minimum curvature takes positive values. Maximum curvature is a good feature for locating
bumps.
37
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Comparing the minimum and maximum curvatures in the neighborhood of a bump and a
dent as demonstrated in Figures (4.4) and (4.5), one can observe that maximum curvature
is a perfect feature for locating “hills” in a surface, and similarly minimum curvature for
identifying “valleys”. In the neighborhood of a hill, all the points have positive normal
curvatures. Using the fact that maximum curvature is the maximum normal curvature of
all curves passing through a point, we can conclude that maximum curvature takes the
largest positive value on top of a hill. By similar reasoning, the minimum curvature takes
the most negative value at the bottom of a valley. We can use this property of maximum
and minimum curvature to locate sulci and gyri on the cortical surface.
We have calculated the minimum, maximum, mean and Gaussian curvature of a typical
human cortical surface. The curvatures have been plotted both on the original extracted
cortical surface of our sample data and its inflated version. In order to generate a
reasonable curvature map for the cortex we have to ignore a few out of range values
resulting from computational errors. The curvature of these points are replaced by the
average value of the curvature of the points in their neighborhood. To determine the limit
of acceptable values, we use the sorted histogram of the quantity and pick a threshold
value such that more than 99% of points have values less than the threshold.
From the Figures (4.6) and (4.7), we can see that minimum and maximum curvatures can
distinctly identify the location of sulci and gyri on the cortex. Large negative values of the
minimum curvature (dark grey areas) correspond to sulci and large positive values of the
maximum curvature (light grey areas) match with the gyri.
38
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
M inim um C u rv a tu re . S a m p le C ortical S u rfac e M inim um C u rv a tu re , Brain
Figure 4.6. Minimum curvature of human cortical surface mapped onto: (a) Original brain, (b) Inflated brain.
M axim um C u rv a tu re , S a m p le C ortical S urface M axim um C u rv a tu re , Brain
Figure 4.7. Maximum curvature of human cortical surface mapped onto: (a) Original brain, (b) Inflated brain.
The mean curvature (Figure (4.8)) by itself is particularly informative for ascertaining both
types of major cortical folds. It takes positive values where the cortex is folded inward
along the crown of a gyrus, and negative values where the cortex is folded outward along
the fundus of a sulcus.
Finally, Figure (4.9) displays the Gaussian curvature of the human cortical surface.
Clearly, the Gaussian curvature of brain cortex is not a constant value equal to zero; hence
it is not a developable surface. However, its mean is close to zero and its variance is very
small.
39
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
M ean C u rv a tu re . S am ple C ortical S u rfac e M ean C u rv a tu re , Brain
Figure 4.8. Mean curvature of human cortical surface mapped onto: (a) Original brain, (b) Inflated brain.
G a u ss ia n C u rv a tu re , S a m p le C ortical S u rfac e G a u ss ia n C u rv a tu re , Brain
Figure 4.9. Gaussian curvature of human cortical surface mapped onto: (a) Original brain, (b) Inflated brain.
40
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
c h a p t e r 5 Finding the Flat Map of Human Cortical Surface
In the last two decades, a growing interest has developed among neurologists and
engineers in studying the problem of unfolding the cortical surface and generating flat
maps of the human brain. A central problem in the study of cortical organization and
function is the difficulty in analyzing regions buried within the deep and irregularly
convoluted sulci of the cortex. The difficulty is not simply in visualization, because buried
regions can be exposed by taking two dimensional slices through the cortex using various
histological or noninvasive imaging techniques. Rather, the main challenges are studying
the functional architecture of cortex and measuring detailed maps of neural activity on the
complex curved surface of the brain.
This chapter is intended to provide a review of existing cortex unfolding and flattening
methods and to establish the motivation for the design of our geometric distortion
minimization techniques for generating flat maps of the human brain. We start with a brief
overview of the mathematical definitions of three important mappings.
5.1 Mappings
The first systematic investigation of the problem of surface mapping is attributed to the
construction of the globe map for nautical or other purposes. Of particular theoretical and
practical interest are mappings that preserve certain geometric properties. The most
important types are isometric or length-preserving mappings, conformal or angle-
preserving mappings and equiareal or area-preserving mappings. All these mappings
41
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
belong to a general class of mappings called diffeomorphisms. A differentiable map c p
from a portion X of a regular surface onto a portion Y of a regular surface is called
diffeomorphism if its inverse c p -1 exists and is differentiable.
Isometric m apping. An isometric mapping or isometry is a diffeomorphism cp: X — > Y ,
such that the length of any arc on Y is the same as that of its inverse image on X .
A diffeomorphism cp: X Y is isometric if and only if at corresponding points of X and
Y , when referred to the same coordinate systems, the coefficients of the first fundamental
forms on X and Y are equal [22] [36].
Conformal mapping. A conformal or angle-preserving mapping is a diffeomorphism
cp: X — » Y , such that the angle of intersection of every arbitrary pair of intersecting arcs on
Y is the same as that of the inverse images on X at the corresponding point. The angle
between two intersecting curves is, by definition, the angle between the tangents to these
curves at the point of intersection.
A diffeomorphism cp: X — > Y is conformal if and only if, when on X and Y the same
coordinate systems have been introduced, the coefficients of the first fundamental forms
are proportional [22] [36].
Equiareal mapping. An equiareal or area-preserving mapping is a diffeomorphism
cp: X -» Y , such that every part of X is mapped onto a part of Y which has the same area.
A diffeomorphism cp: X — » Y is equiareal if and only if at corresponding points of X and
Y, when referred to the same coordinate systems, the discriminants of the first
fundamental forms on X and Y are equal. The discriminant of the first fundamental form
is defined as EG - F2 , where E, F , and G are the first fundamental coefficients [22] [36].
It can be proved that every isometric mapping is also conformal and equiareal, and every
equiareal and conformal mapping is isometric. It is of basic importance that there is no
42
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
difference in the measurement of lengths, angles, and areas in isometric surfaces, although
the surfaces, when considered from the embedding space may have an entirely different
geometric shape.
The concept of isometric mapping is usually used in the same sense as bending
deformation, a surface deformation that preserves the length of every arc on the surface.
Unfolding is considered as a special type of bending deformation. Surfaces which can be
transformed into each other by bending are called applicable.
Two distinct points on a curved surface can be connected by many different paths on the
surface, and each path has in general a different length. Those paths corresponding to
minima of the path length function are analogous to straight lines in Euclidean space and
are known as geodesics. A bending maps geodesics to straight lines.
The geodesic distance between two points is defined to be equal to the length of the
geodesic connecting them. Geodesic distance depends only on the first fundamental form
of the surface and it remains unchanged under bending. Properties that remain unchanged
under bending are called bending invariants. Other examples of bending invariants are
surface area and Gaussian curvature. Counter examples are Euclidean distance, enclosed
volume and mean curvature.
The properties of a surface which depend on the first fundamental form only and are
bending invariant are called intrinsic properties. Intrinsic properties are invariant under
transformations which do not stretch or tear the surface. Correspondingly, there are
properties of the surface that depend on the particular configuration of the surface in the
space. These properties are called extrinsic properties. Thus a sheet of paper has the same
intrinsic properties whether it is flat or crumpled but not the same extrinsic properties.
43
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
5.2 Survey of Current Techniques
We have categorized published works on unfolding and flattening of the cortical surface
into two general categories of mechanical and geometrical approaches, which are based on
the underlying model they use for representing the brain surface.
5.2.1 Mechanical Model Based Flattening Algorithms
In this class of algorithms the material of the brain is modeled as an isotropic elastic
substance that is moving under a system of forces. Depending on the definition of the
system of forces and the final equilibrium state of the flattened brain, different physical
interpretations can be made for the behavior of algorithms in this class.
The simplest solution in this category, proposed by Dale and Sereno, is a shrink-wrap
method for blowing up the elastic surface of cortex by simulating the way a folded balloon
is blown up [18]. The basic idea behind this methods is to start with a three-dimensional
wire frame reconstruction with the correct topology and then gradually deform the shape
of the surface by rubber sheet transformations to conform to the cortical sheet. The
location of each vertex on the surface is updated iteratively according to elastic forces
between neighboring vertices and repulsive and attractive forces along local surface
normals which depend on the MRI data at the vertex. The criteria used in this unfolding
scheme are not optimal in terms of preserving any metric property of the resulting map.
They preserve only local properties of the surface and therefore do not rule out large-scale
distortions caused by locally correlated errors.
An improved version of this method has been recently proposed [27], in which the
resulting map is forced to have a minimum distortion in some geometric measures of the
surface. In this method the external force is an unfolding force that acts to flatten the
surface by making each node closer to being coplanar with its neighbors, and at the same
time, to resist internal forces that are attempting to maintain the intrinsic curvature of the
44
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
original surface. The internal forces model the elastic properties of the surface. They act to
restore the length between adjacent nodes to their reference values in order to preserve
surface area as well as distances within the cortex.
With a similar approach, Van Essen and his team have developed algorithms for creating
two-dimensional maps of the cortex by modeling it as an elastic sheet which is
continuously deformed to a planar configuration with minimal energy. The energy is
defined as a measure of the difference between local lengths and angles on the two
dimensional map and the corresponding lengths and angles in the three-dimensional
surface. The first version of this algorithm started from an arbitrary planar configuration
with the same topology as cortex, possibly with different geometry. The map was then
subjected to random perturbation and configurations were selected that had progressively
lower energy, until a mapping of minimum overall distortion was asymptotically
approached [9].
Later, in 1995, Carman et al. [11] enhanced the cortex model to include deformations
generated by a system of forces. In the new model, internal forces acted to preserve the
topology and geometry of the sheet while external forces acted to unfold and flatten it. The
unfolding force is designed to drive the overall surface into a planar configuration and to
act against internal forces attempting to maintain intrinsic curvature or local folding. The
internal forces are the longitudinal and torsional forces, which together model the elastic
properties of the surface. The starting configuration for the dynamic unfolding process,
which also provided the reference geometry, was a wire-frame representation of the
reconstructed brain surface.
Drury, et al. [23] have recently extended this method by incorporating the concept of
multiresolution processing. By making a global shape change on a coarse-resolution
representation of the surface and then transferring it to finer-resolution representations,
45
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
which are later subjected to more localized shape refinements, they reduce the
computational demands of the unfolding process.
This enhanced method represents some improvements over the previous ones developed
by this group in terms of accuracy of results and speed of processing. The criteria used for
creation of the flat map is to minimize the distortion in local lengths and angles with
reference to those in some wire-frame representation of reconstructed brain surface.
Because of this, although this method is not explicitly formulated to find an isometric
mapping between the brain surface and plane, the outcome is equivalent.
5.2.2 Geometrical Model Based Flattening Algorithms
In the second class of algorithms the cortical surface is considered to be a geometric entity
not a mechanical object. In this geometrical model, minimizing the distortion in some
geometric measure of interest replaces the objective of finding the equilibrium state of an
elastic sheet subject to a system of forces. There are four measures one may want to retain
in the flat map of the cortex: the neighborhood relationship between different regions of
brain, the area of these regions, the length of curves connecting a single point to the other
points on the surface, and the angles between these curves. The transformations which
preserve these measures are called topological, equiareal, isometric and conformal,
respectively. Although there have been some reports on equiareal mappings of rat brain
surface to the plane [1], isometric mappings appear to be the preferred subject of study for
many researchers.
Seeking the optimal preservation of the metric structure of cortex, Schwartz et al., [48]
formulated a variational problem to maximize the goodness of fit of the “distance matrix”
of the original 3-D model and the flattened model of the brain. The distance matrix is
defined as the matrix of all possible inter-point distances. In a polyhedral model of the
surface, distances must be defined as minimal geodesic distances along paths embedded in
the surface so that there is a match between straight lines in the plane and distances
46
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
defined on the polyhedron. The goal is to determine a set of points in the plane such that
the corresponding distance matrix in the plane provides a best fit, in the least squares
sense, to the distance matrix of the original surface. This is done by applying a Newton-
Raphson gradient descent algorithm on the difference between the three-dimensional
(geodesic) distance matrix and that of a corresponding set of initially random two-
dimensional points.
The resulting flat map is optimal in the sense that it is derived from a variational principle
that optimizes the overall fit between the curved and planar surfaces. However the major
drawback of this algorithm is the difficulty of finding the minimal distances. Their
suggested solution to the problem of finding geodesic distances on a surface has an
exponential complexity and the requirements of performing this procedure in terms of
both memory and time becomes prohibitive as the level of complexity of the surface
increases.
So far, all the methods that we have referred to, in spite of their differences in approach are
similar in the sense that they start from the polyhedral model of the cortex and base all the
calculations on that model. Alternatively, one can start from a parametric model of the
cortex as suggested by Davitzikos. In [19][20], he models the outer cortex as a thick sheet
of grey matter which is deformed toward its central layer as an equilibrium is reached
between two sets of opposing forces: external forces trying to unfold the surface and
internal forces acting as elastic resistance which preserves the connectivity of the surface.
The equilibrium configuration of the surface, which is the solution to the associated
elasticity differential equation, provides a numerical description of cortex as a function of
two parameters. Davitzikos then uses this parametric representation as the initial value for
a fixed point algorithm whose final answer will be a parameterization isometric to the
plane. This algorithm uses the fact that first fundamental coefficients for the plane are
E = G - 1 and F = 0 . Instead of trying to reparameterize the entire surface X(u, v) to
47
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
globally change these values, only the isoparametric curves (Constant-u curves, and
Constant-v curves) are reparameterized to become unit speed curves (satisfying
E - G = 1 ) and also intersect each other at right angles (satisfying F = 0).
5.3 Flattening As An Optimization Problem
In the process of developing the flat map of a particular surface the first question to be
answered is whether the surface is isometric to the plane. A curved surface that can be
brought into isometric correspondence with a plane is said to be developable. Since an
isometric mapping does not change the intrinsic geometry of a surface, the fact that the
Gaussian curvature of plane is zero everywhere implies that the Gaussian curvature of a
developable surface must also be identically zero.
Most surfaces are not developable, for example a sphere has a constant Gaussian curvature
and can not be mapped isometrically into the plane with zero error. This is the so-called
map-maker’s problem, the problem of portraying the surface of the earth on a flat plane.
Given that the earth is approximately spherical, some distortion of conformality, distance,
direction, scale, and area always results from projecting the earth onto a plane. Some
projections minimize distortions in some of these properties at the expense of maximizing
errors in others. Some projection are attempts to moderate distortion in all of these
properties.
The results of our experiments with human brain (Chapter 4), as well as the reports of
other researchers conducting experiments on macaque monkey [47][33], indicate that the
Gaussian curvature of cortical surface is not zero; however, it is relatively constant over
most of the surface. In contrast to the sphere, this result suggests that flattening of cortex
with some small distortion is possible, with larger than average errors occurring in the
neighborhood of sulci and gyri where the Gaussian radius of curvature is larger than
average [47].
48
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
The work we will describe in this section is an effort to find a practical approximate
solution to a problem for which no exact solution exists. Formally, our approach entails
the specification of the flat map of cortex as a solution to a constrained optimization
problem. We formulate the constraints as non-negative energy functions in three
dimensional space with zeros at point-configurations satisfying the constraints. We refer to
the constraint functions as energy functions not because they model the energy of actual
physical systems, but because they play a role similar to that of physical energy functions.
Intuitively, energy constraints may be viewed as forces that pull the surface into the
desired configuration and hold it there, although they are not necessarily intended to be
physically realistic forces.
We are seeking an optimal configuration of points which minimizes the integral of a
surface metric subject to some geometric constraints. An ideal surface metric for our
purpose is one that reaches its extremum value when the cortical surface deforms to an
unfolded configuration. The role of geometric constraints is to impose the requirements
for an isometric mapping. One can choose the preservation of geodesic distances as a
constraint directly characterizing isometry. However, computing geodesic distances on an
arbitrary manifold is a difficult problem and so far none of the proposed solutions are
computationally tractable for huge data sets such as the human brain [63][39][21].
Alternatively, we have chosen to use constraint functions that preserve areas and angles,
thus imposing equiareal and conformal mapping requirements. This will serve our purpose
because, as we mentioned earlier in this chapter, if we find a transformation that is both
equiareal and conformal, it can be shown to be isometric.
5.3.1 Objective Function: Minimizing “Convexity”
By defining a measure for quantifying the folding pattern of a surface one can numerically
evaluate the deviation of a surface from a flat configuration. Ideally, such a measure should
be able to discriminate between surface folds based on their size and their folding
49
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
direction. Mean curvature captures the folding pattern of a surface but it is equally
sensitive to large and small folds, yet it distinguishes between fold-in and fold-out and
presents the difference with a sign change (Figure (5.1.a)).
M ean C urvature. Sam ple C ortical S urface Convexity, Sam ple C ortical S urface
Figure 5.1. Convexity vs. mean curvature, (a) Mean curvature takes relatively equal values for both small and
large folds, (b) Convexity takes high values for large folds and is almost insensitive to small folds.
Computational cost can be considered a drawback for mean curvature if it is used in a cost
function of an optimization problem dealing with large data sets. As explained in Chapter
4, the calculation of mean curvature at each point of the surface involves finding the
coefficients of a local polynomial patch through solving an overdetermined system of
linear equations, a process requiring the computationally demanding task of calculating
the pseudo-inverse of a matrix.
Motivated by the definition of locally convex surfaces [22] we have come up with an
alternative measure for determining the convexity of a surface that is fast and easy to
calculate for tessellated surfaces.
DEFINITION. A surface X is locally convex at a point P if there exists a neighborhood N
of P (subset of X ) such that N is contained in one of the closed half-spaces determined by
the plane tangent to X at P . If in addition, N has only one common point with the tangent
plane, then X is called strictly locally convex at P .
50
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
We define the convexity of a surface S at a point Xt to
be equal to the average of the inner products of the
unit surface normal vector at point X- and the unit
edge vectors X t - linking the point X { to its neighbors
X j, or equally the average of the cosines of the angles
between the normal and the edges (Figure (5.2)).
_ l v x ijx x ij +1
' N^ x ij*x i j + i\\
N . - l
Conv-• = - V n.
I N. Zj 1
(Xj-Xi)
Figure 5.2. Xj, a typical node of a
tessellated surface, rij represents surface
normal and Xy represent the edges.
(Eq. 5.1)
j = 0
(r \
"i /
/ A X __^ J
/ \ ij / )
(a)
(b
Figure 5.3. (a) At a locally convex point all the neighbors
are above the tangent plane, (b) At a locally concave point
all the neighbors are below the tangent plane.
Assuming that the surface normal is
always pointing outward, we can
verify that if the surface is locally
convex at a point, all the neighbors of
that point are above the tangent
plane; therefore, the angles between
the normal and edges are all acute
and convexity is positive (Figure
(5.3.a)). On the other hand, if the surface is locally concave at a point, all the neighbors of
that point are below the tangent plane, hence all the angles are obtuse and convexity is
negative (Figure (5.3.b)).
o
For a locally planar portion of a surface all the angles are 90 and the convexity is zero.
Based on this convexity measure we have defined an energy functional Ic that quantifies
the deviation of a surface from a planar configuration and is zero only when the surface is
a flat plane.
N - 1 N - 1
Convi = ^ E
i = 0 / = 0
N , - l
N, E ",
} = 0
(Eq. 5.2)
51
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
In Equation (5.2) we have used the symbols X-t :i = 0,1, 1 (N is the total number
of vertices of the mesh) to refer to the vertices of the mesh representation of the surface,
and ni to refer to the surface normal at point Xi . We assume that each point X{ has a total
of Ni neighbors that we represent by X- : j = 0 , 1 , N{ - 1 . Ic is proportional to the
average squared convexity over all points on the surface. The reason for using the squared
value is to treat convex and concave points equally. The global minimum of this function is
where the surface reaches a planar configuration.
In order to minimize lc , we need to calculate the gradient of Ic with respect to the vector
of vertex positions X = (XQ , Xp ..., XN_ j /w h e r eXt = (Xix, X iy, X iz) :i = 0,1 , N - I.
For this purpose we take the derivative of lc with respect to Xt ,
0/,.
N,- 1
; = o
l
w f ° nvi
1
\ xr xi \ \
-(nr (Xj- X i))(Xj- X i)
(Eq. 5.3)
1
‘ p r x , 1
In the above equation and all future relations that we will derive for the derivatives of our
energy functionals, we ignore the dependency of the unit surface normal on vector of
dn-
vertex positions and therefore we assume that ■ = — = 0 ,for all i = 0, 1, 1 and
dxk
k = 0, 1, 1 .
5.3.2 Constraint Function: Preserving the Area
In order to define the constraint energy functional corresponding to a;xb
i L
the preservation of area, we consider the i triangle in the surface
tessellation with unit normal vector nj , and edges a- and b{
connecting the vertex Xi to two of its neighbors. The area of this
triangle is given by half the norm of the cross product of the edges
n, =
I a,-x b J I
XA i
at
Figure 5.4. Normal to
the plane of a triangle.
ai and b{.
52
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
The areal constraint function ls is defined to penalize the difference between the current
area and the original area occupied by each triangle. It is proportional to the average of the
squared difference in the areas of the triangles constructing the mesh in their current forms
and their original shapes.
T - 1 T - 1
(Eq. 5.4)
t 1 V ' r p O . ^ 1 X - ' T I L » I n \ 1 / 0 , 0 .
!s = = r S L 2 l(fli X^ -) “ 2 (ai Xbi )
i = 0 i = 0
In this relation T refers to the total number of triangles in the mesh and n indicates the
iteration number, with n = 0 being the starting value corresponding to the original
cortical surface. In order to get a closed form expression when calculating the derivative of
the area of each triangle, we use the following relation:
si = k aixbi = V aixbi)-ni
(Eq. 5.5)
2 1 1 * ' I I 2'
where « ■ is the unit vector normal to the plane of the triangle. To demonstrate the
functional dependence of ls on the position of the vertices of the surface one can express
a • and b( as functions of the vertex Xi and its neighbors as
T - 1
= f X f 5 « ( ^ ■ - 4 ,) X ( 4 , - 4 ,» • <) - 4
T ^ V 2
/ = 0
(Eq. 5.6)
where we have renamed to Xc . and used Figure (5.4) as reference to describe the sides
of the triangle in terms of its vertices.
Taking the partial derivative of I with respect to klh vertex results in
dX‘
i = 0
dx:
(Eq. 5.7)
ds;
In order to calculate —- we use the following relations from vector analysis [56]
dxn k
53
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
J-(A BxC) = A - B x ^ + A- ^ - x C + ^ - B x C
du ou ou ou
A (BxC) = B (C x A ) = C ■ (Ax B)
(Eq. 5.8)
(Eq. 5.9)
to get
as”
dx:
n , rn vrn \
n i X ( Bf — Ci)
• l / - w r 'l - * r 'L \
niX(XCi - XAi)
Yn _ Yn
A k ~ Ai
Yn _ Yn
A k ~ Bi (Eq. 5.10)
cos a =
M p n u n — ■
ct
Figure 5.5. Cos of the
angle between two vectors.
x k = x a
0 else
5.3.3 Constraint Function: Preserving the Angle
The constraint of preservation of angles must be included in our
objective function in order to enforce the conformality of the flat
map with the original surface. Given that there are T triangles in
the tessellated surface, we can define 3 T equality constraints for
the angles of the triangulated mesh. The satisfaction of the
conformality constraint is accomplished by adding to the objective function a term that
prescribes a high cost for violating these equalities. This term must be a function of the
difference between the angles in the current configuration of the surface and those in its
original configuration.
The difference between any one-to-one function of the angles is also equally useful. Cosa
is a monotonically decreasing function of a for 0 < a < 180. Therefore, a cost function
defined to minimize the difference between the cosine of the angles of a triangle with
respect to some reference will indirectly minimize the difference between the angles of
that triangle and those of the reference one. The advantage of the cosine over other
monotonic functions of an angle is that it can be easily calculated as the inner product of
the unit vectors in the directions parallel to the sides of angle (Figure (5.5)).
Based on this argument we have defined the angular constraint function Ia as
54
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
37" - 1
0 2
(Eq. 5.11)
i = 0
where T is the total number of triangles in the mesh and n is the iteration number starting
from zero.
The dependency of / on the position of the vertices of the surface becomes clear when
C o s a t is replaced with its equivalent in terms of the sides of the angle a i .
3jT- 1
/ = — y
a 37 Zj
i = 0
n
Ci
i n
n
bi ci
^ 0
- C o s a {
Using b( = X Ci- X Ai and c( - = X Bi- X Ai we will have:
3 T -
i = — y
a 37 Z j
i = 0
/ Yn Yfl Yf> YH ^
Ci ~ Ai Bi ~ Ai
Yn y h
Ci ~ Ai
Yn Yn
Bi Ai
dl
C osol}
(Eq. 5.12)
(Eq. 5.13)
In order to calculate V/ we need to determine . Taking partial derivative of I with
dXk
respect to k vertex we will get:
3T - 1
a 2 . n 0. d n
- = — - > (cos a- -c o s a ,) cos a.
n 3T 1 ‘ -w,n 1
dX
where -
d d d V
dXl W k x dXn k y dXn kz;
i = 0
and
dX'l
d n
cosa- =
dX
(b + c) n( b c
I , + coscc- — - + ----
H ' t |* ||2
b n C
— ii - cosa - — —
C 1 ii ii 2
— c o sa ( -
||c||
n b
\ \ H
Yn - Yn
A k - Ai
Yn - Yn
A k - Bi
Yn _ Yn
A k ~ Ci
else
(Eq. 5.14)
(Eq. 5.15)
55
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
5.3.4 Perturbation Function: Smoothing the Surface
When formulating our optimization problem we ideally are looking for a global minimum
of function / subject to the constraints ls and Ia . However, both from the theoretical and
computational points of view, we must in many circumstances be content with a relative
minimum point. Theoretically, when deriving necessary conditions based on differential
calculus, or practically when searching for the minimum point by a convergent stepwise
procedure, comparisons of the values of nearby points is all that is possible. Global
solutions can only be found if the problem possesses certain convexity properties that
guarantee that any local minimum is a global minimum. In order to be able to take
advantage of these convexity properties the objective function and the constraint set must
both be convex [2][38], a requirement that is obviously not satisfied for our problem.
Common gradient-search techniques that are often used for solving optimization problems
quickly become trapped in local minima. Local minima are found in an iterative gradient
based minimization routine in which constraints are enforced softly using penalty
functions, in points at which the gradient of the cost plus penalty is zero.
To deal with the unavoidable problem of local minima we have considered a fourth energy
functional lp , in our total objective function that will become effective only when the
solution is trapped in a local minimum. This functional is essentially a spring force
function that drives the surface to a smoother configuration forcing the current solution
away from local minimum.
N - l N t - l
= X \x r x , f
i = 0 j = 0
In this relation X(:i = 0,1 , N - 1 represent the vertices of the surface and Xj:
j = 0 ,1 ,..., N{ - 1 represent the neighbors of X -. The derivative of Ip with respect to Xk
is:
56
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
5.3.5 Merit Function: The complete Energy Function
The complete energy function incorporating the objective function of minimizing
convexity and the constraint functions of preserving the areas and the angles is given by
where the coefficients Pc , (3; and Pa define the relative importance of unfolding versus
the preservation of areas and angles, respectively. By defining ys = and
ya = Pa/P c we obtain
Here we have explicitly expressed the dependency of the merit function on X , the position
vector of the surface, and ya and ys , respectively the areal and angular penalty
coefficients. The procedure for solving the optimization problem given by Equation (5.19)
is this: We let {ysk} , {yak} ,k - 1, 2,... be two sequences tending to infinity such that for
each k, ysk>0, ysk+l>ysk and yak>0, yak+l>Yak. Then, for each k we solve the
problem
to obtain a solution point Xk . Any limit point of the sequence {Xk \ is a solution to our
optimization problem.
The actual minimization of /(ysk, yak, X) is accomplished using a partial conjugate
gradient method with line searching based on Armijo’s rule [38].
5.4 Solving the Minimization Problem: Conjugate Gradient Method
There are many algorithms designed to solve nonlinear programming problems that
belong to the general class of iterative decent algorithms. The simplest and oldest
(Eq. 5.18)
(Eq. 5.19)
min ^ s k ’Vak’X)
(Eq. 5.20)
57
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
algorithm in this category is the steepest descent method. The method of steepest descent
is defined by the iterative algorithm
X k + 1 = X k ~ a k ^ ^ X k^
where a k is a nonnegative scalar minimizing I(Xk- a kVI(Xk) ) . In words, from the point
Xk we search along the direction of the negative gradient -Vl(Xk) to a minimum point on
this line; this point is taken to be Xk + j .
Steepest descent converges very slowly. As an alternative, conjugate direction methods are
characterized by an accelerated convergence rate with little additional computational
overhead. The most widely used conjugate direction method is the conjugate gradient
method in which the variable Xk is iteratively updated according to
x k + i = x k + a kd k < Ec<-5-21>
where ak is obtained by a line search
I ( X k + CLkd k) = m i n I ( X k + a d k) (Eq. 5.22)
and d k is generated by
rf* = - v / W + V * - i < E c > -5-23>
where
^ VI(Xk)\vi(Xk) - VI(Xk_ l))
X k = ------------------------------------------------- (Eq. 5.24)
V/(Z;t_ 1)
The conjugate gradient method is designed to give convergence in a finite number of steps
for quadratic functions. Since our problem is not quadratic we must deal with the loss of
conjugacy that results from non quadratic terms in the cost function. Also, because of
inaccurate line searches it is not unusual for the method to start generating inefficient
search directions after a few iterations. For these reasons it is important to operate the
method in cycles of conjugate direction steps, with the first step in each cycle being a
58
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
steepest descent step. To decide when we need to restart a cycle we wait until one of the
following two conditions is satisfied: either n iterations have taken place since the
preceding restart, n being the dimension of the domain space of our minimization
problem, or if
VI(Xk)tVl(Xk_ 1)\ > Y | | V / ( ^ _ 1)||2 (Eq. 5.25)
where y is a fixed scalar with 0 < y < 1. This relation is a test on loss of conjugacy, for if
the generated directions were conjugate then we would have S7I(Xk)tVI(Xk _ j ) = 0
The complete conjugate gradient algorithm using restart is [38]:
Step 1. Given X Q compute VI(Xo) and set dQ = -V/(A'0)
Step 2. For k = 0, 1, . .. ,« - 1 :
2a. Set Xk+l = Xk + a kdk where a k minimizes l(Xk + adk) .
2b. Compute VI{Xk + j ) .
2c. If k = n - 1 or XI(Xk)tVI(Xk_ l)\ >y||V/(Z/t_ j) |2 go to Step 3 else, set
d k = -VI(Xk) + X kd k _ j
where
Step 3. Replace XQ by Xn and go back to Step 1.
An important practical issue relates to the line search accuracy that is necessary for
efficient computation. On one hand, an accurate line search is needed to limit the loss of
direction conjugacy, and on the other hand, insisting on a very accurate line search can be
computationally expensive. A practical and popular line search method is Armijo’s rule
[38]. Armijo’s rule guarantees that the selected a is neither too large nor too small. To
implement Armijo’s rule we choose fix scalars s, P e (0, 1) and a e (0, 1) and we set
tn ■
a k = $P ' where m . is the first nonnegative integer m for which the inequality
59
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
I(Xk) - l(Xk + S$md k) > -G Sf> mVI(Xk)’ dk (E q . 5 .2 6 )
holds. In Equation (5.26) d k is the descent direction which for the gradient method is
equal to the negative gradient -VI(Xk) . Usually a is chosen from the interval
[10- 5,10_1] and the reduction factor P e [0.1,0.5] depending on the accuracy of our
estimate of the initial step-size .v.
5.5 Results and Discussion
The flattening algorithm introduced in Section (5.3) develops an isometry between any
arbitrary surface and a flat plane on the basis of a theorem in differential geometry which
states that every equiareal and conformal mapping is isometric [36]. To evaluate the
performance of this algorithm we have applied it to two types of data: artificially
generated phantoms with well known geometric properties (cylinder, cone, sphere) and
real physical data (human cortical surface).
5.5.1 Phantom Studies
Our main focus in the first set of phantom studies was evaluation of the correctness and
accuracy of the objective functions in generating conformal and equiareal maps. It is
essential to establish the effectiveness of this approach, i.e. using the framework of solving
an optimization problem in order to develop a transformation between surfaces, before it
can be applied to real data sets. Both developable and non-developable surfaces were
considered in this study: a cylinder as the developable surface and a sphere as the non-
developable surface.
To evaluate the ability of the energy functionals defined in Section (5.3) to generate
optimal maps with minimum areal and angular distortions, we conducted a test in which
the functionals were expected to restore the area and angles of triangles forming the
tessellated surface of some computer generated phantom that was subjected to distortion.
As the distorted surface we used the flat map of the phantom obtained by simple
60
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
projection of the surface points into a plane. We applied the error minimization procedure
as described in Section (5.4), once with the objective of minimizing the areal errors and
once for minimization of the angular errors. In the process of minimization, the
distribution of distortion was calculated and mapped over the surface of the phantom. In
these distortion maps the color assigned to each triangle was chosen to be proportional to
the percentage of the relative error corresponding to that triangle. The histogram of areal
and angular errors were also calculated as a quantitative measure for the distribution of
distortions. In the case of areal distortion for each triangle this error is simply equal to the
difference between the area of that triangle in the current configuration and its area in the
original three dimensional surface, divided by the average area of the triangles in the three
dimensional surface. In other words, the areal error for the ith triangle, efrea, is defined
as:
An 40
Area A i ~ A i
et - — fZ \ x 100 < E c l- 5,27>
1 v i . 0
f S - 4 ■ ■
T
i = 0
where An; is the area of the ith triangle after n iterations, A0/ is its area in the original
surface and T is the total number of triangles in the tessellation. According to this relation,
a negative error indicates a shrunken triangle whereas a positive error corresponds to an
expanded one. For angular distortion the error for the ith triangle is defined as:
Angle 1
3
Coso^k - Costx k
1 3
*= 1
Cosa^k
(Eq. 5.28)
where a k = 1,2, 3 are the three angles of the ith triangle after n iterations and a°k,
k - 1, 2, 3 are the same angles in the original 3-D surface. In this case, the error is always
non-negative and larger values of error represents more distorted triangles.
61
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
A useful indicator for expressing the level of success of an optimization algorithm is the
magnitude of the relative error in the quantity that the algorithm is minimizing. The
average relative error in area and angle are defined as: T_ l
Average Relative Error in Area: EArga = j, ^ g Area
i = 0
r - i
Average Relative Error in Angle: EAng[g = i ^ e Angle
i = 0
Ideally, when the objective is completely accomplished the relative error will become
zero. We suggest that a relative error in the range of 0.01% to 0.1% is probably
satisfactory for most practical purposes.
5.5.1.1 Equiareal and Conformal Maps of Cylinder
Figure (5.6) shows a tessellated half-cylinder phantom
that is generated by specifying the location of its
vertices analytically. Figure (5.7.a) shows the areal
distortion for the projectional map of the half-cylinder
and figure (5.7.c) shows the angular distortion for the
same map. Figures (5.7.b and 5.7.d) display the
histogram of areal and angular distortions for this map,
respectively. As can be seen, the distortion is minimum
in the center of the map with a symmetrically increasing trend towards the boundaries.
In Figure (5.8) we have demonstrated the change in the distribution of areal distortion over
time as the difference between the area of the flat map and the half-cylinder is minimized
by the iterative process of conjugate gradient. The figure clearly shows that the map
expands from all four comers in order to accommodate the compressed triangles in the left
Figure 5.6. Tessellated half-cylinder.
62
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Areal Distortion (Relative Error - -3623% )
Histogram of Angular Distortion
Angular Distortion (Relative Error = 34.98%)
(c) (d)
Figure 5.7. Distribution of areal and angular distortions for the normal projection map of the half-cylinder,
(a) Areal distortion map. (b) Histogram of areal distortion, (c) Angular distortion map. (d) Histogram of
angular distortion.
and right sides. From Figure (5.8) we observe that the iterative process of conjugate
gradient not only reduces the total areal error, but also distributes this error approximately
uniformly over the entire map. While in the beginning of the process the histogram of
areal error is concentrated on the negative side of the axis at the end it is almost
symmetrically centered around zero. This can be verified also by checking the
configuration of triangles in the map. When the process starts the map contains
compressed triangles in the left and right borders, but it gradually changes such
63
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Areal Distortion (Relative Error - -1.57%) Areal Distortion (Relative Error - -11.24%)
- 0 .0 5 - 0 X 4 - 0 .0 3 - o jh - 0 . 0 1
Areal Distortion (Relative Error - -0X001%) Areal Distortion (Relative Error - -0 01%)
o x is
(c) (d)
Figure 5.8. Variation of areal distortion in the process of generating an equiareal map for half-cylinder.
(a) 10 iterations, (b) 50 iterations, (c) 100 iterations, (d) 10000 iterations.
64
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Angular Distortion (Relative Error - 8.93%) Angular Distortion (Relative Error - 19J9S%)
Angular Distortion (Relative Error - 0.01%) Angular Distortion (Relative Error - 1 i(7%)
(c) (d)
Figure 5.9. Variation of angular distortion in the process of generating a conformal map for half-cylinder,
(a) 10 iterations, (b) 500 iterations, (c) 2500 iterations, (d) 10000 iterations.
65
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
that at the end, pairs of compressed and expanded triangles (black and white triangles in
Figure (5.8.d)) are uniformly spread all over the map. Figure (5.9) shows the angular
distortion map of the half-cylinder in different stages of the development of its conformal
map. This figure indicates that the reconfiguration of the points is in the direction of
aligning the sides of adjacent triangles and making them right triangles. As the sequence
of errors converges to zero we would expect that all triangles reshape to a right triangle. As
expected, the concentration of errors is on the two extreme sides of the map where the
triangles are significantly distorted due to the initial projection into the plane. Again,
based on the distortion maps and histograms we can verify that as a result of the iterative
minimization process the angular error is reduced and distributed approximately
uniformly over the map.
5.5.1.2 Equiareal and Conformal Maps of
Figure (5.10) is a picture of the tessellation of
an analytically defined hemisphere. The areal
and angular distortion maps of the normal
projection of the hemisphere to the plane of its
greatest circle, as well as the histogram of areal
and angular errors are displayed in Figure
(5.11). We can see that the distortion is
minimum on the top of the hemisphere where
the tangent plane is parallel to the plane of projection and it increases as we move to the
boundaries of the surface. We also note that the initial total areal error of -49.97%
conforms with the results one would expect from mathematical calculation. Since the area
2 2
of the greatest circle ( nR ) is half of the total area of the hemisphere (2nR ), projecting
the hemisphere onto the plane of its greatest circle will naturally result in 50% error.
Hemisphere
Figure 5.10. Tessellated hemisphere.
66
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Similarly to the half-cylinder, the distribution of areal and angular errors in the process of
developing equiareal and conformal maps of the hemisphere were calculated, and a few
representative samples of the results are displayed in Figure (5.12) and Figure (5.13),
respectively. This series of figures shows that the error is initially concentrated in the
regions around the boundaries of the surface, but after a sufficient number of iterations,
distributes evenly over the entire map.
Areal Distortion (Relative Error - -4857% )
-1 - 0 5 -0.0 -0.7 - 0 5 -OS -0.4 -0.3 -0.2 -0.1 0
Histogram of Areal Distortion
-1.1 -1 -0 5 - 0 5 -0.7 - 0 5 -0.5 -0.4 - 0 5 - 0 2 -0.1
1W
Angular Distortion (Relative Error - 3257%)
0 0.05 0.1 0.15 0.2 0 2 5 0 5 0.35 0.4 0.45 0.5
Histogram of Angular Distortion
60 -
40 •
^ j — J
0 0 • 1
- n n
0 0 ■
6 0 •
i
20 •
0.1 0 2 0 5 0.4 0 5 0 5 0.7 0 5 0 5 1 1.1
(c) (d)
Figure 5.11. Distribution of areal and angular distortions for the normal projection map of the hemisphere,
(a) Areal distortion map. (b) Histogram of areal distortion, (c) Angular distortion map. (d) Histogram of
angular distortion.
67
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Areal Distortion (Relative Error - -0.001%) Areal Distortion (Relative Error - -1.95%)
-0X11 -0.008 -0.006 -0XI04 -0X102 0 0.002 0.004 0.006 0.006 0.01
Figure 5.12. Variation of areal distortion in the process of generating an equiareal map for hemisphere,
(a) 10 iterations, (b) 50 iterations, (c) 100 iterations, (d) 1000 iterations.
68
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Angular Distortion (Relative Error = 4.34%) Angular Distortion (Relative Error s 15.47%)
(a) (b)
Angular Distortion (Relative E rror» 0.93%)
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05
120
1 0 0
4 0
20
O
O 0. 0.02
Angular Distortion (Relative Error = 0.01%)
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05
so
so
40
2 0
oo
eo
so
40
0 . 0 1 o.os
( C ) ( d )
Figure 5.13. Variation of angular distortion in the process of generating an equiareal map for hemisphere,
(a) 25 iterations, (b) 150 iterations, (c) 300 iterations, (d) 500 iterations.
69
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
5.5.1.3 Isometric Maps of Half-Cylinder and Hemisphere
In the next phase of our phantom studies we wanted to evaluate the accuracy of our
method in generating isometric maps of surfaces. By using both constraint functions
simultaneously preserving area and angles, the projection map of the phantoms were
forced to converge to a simultaneous equiareal and conformal map, or equivalently an
isometric map. The combined objective function in this case is
' = PA + PA + P/p ( E < > - 5 - 2 9 >
where Is and Ia represent areal and angular constraint functions respectively, with (3 5 and
Pa determining their relative weights. Since we desire to have both constraints enforced at
the same time we set = Pa • As mentioned earlier in Section (5.3.4), this objective
function is nonconvex and it is possible that the sequence of solutions for this optimization
problem traps in local minima. The last term 1 in Equation (5.29) is the perturbation
function which helps the sequence escape from a local minimum. However, since this term
effectively smooths the surface through local averaging, it counteracts the other two terms.
To reach a compromise between these two opposing effects we use an adaptive
mechanism for selecting $p . Typically we start with Pp = 0 . As soon as the sequence is
trapped in a local minimum (this happens when the values of both objective functions Is
and 1 do not change in two consecutive iterations of the conjugate gradient algorithm
while the magnitude of the gradient vector is less than a small threshold), we assign a fixed
value to Pp (usually Pp = 0 1P5) for the next iteration of the algorithm. Next time the
algorithm gets stuck a fraction (usually 0 .1) of the previous Pp will be used and this
continues until becomes less than 10- 5 .
Following the described procedure we applied our algorithm to both phantoms and
calculated the final angular and areal distortion maps (Figure (5.14) and Figure (5.15)).
70
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Areal Distortion (Relative Error = I).58%)
Histogram of Angular Distortion
Angular Distortion (Relative Error s 4.93%)
( C ) ( d )
Figure 5.14. Isometric flat map of half-cylinder, (a) Areal distortion map. (b) Histogram of areal error,
(c) Angular distortion map. (d) Histogram of angular error.
71
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Histogram of Areal Distortion
Areal Distortion (Relative Error = 0.98%)
Histogram of Angular Distortion
Angular Distortion (Relative E rror»1.06%)
(c) (d)
Figure 5.15. Isometric flat map of hemisphere, (a) Areal distortion map. (b) Histogram of areal error,
(c) Angular distortion map. (d) Histogram of angular error.
72
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
5.5.2 Isometric map of Human Cortical Surface
In this section we present the results of applying our algorithm to a human cortical surface.
This cortical surface was extracted from the MR images of the head of a normal subject by
using a morphological based segmentation method [45] [46] and was post processed
through manual editing to correct the topology. The corrected volume was then tessellated
using the algorithm described in Chapter 2. Since the triangles created by the tesselation
process are formed by linking pairs of neighboring points in 3-D, and the neighborhood
system used for this purpose only allows a limited number of configurations for the
neighboring points, at the voxel scale the tesselation tends to be jagged with a uniform
shape pattern. To alleviate this effect, we smooth the initial tessellation using a relaxation
operator [23]. Vertices are iteratively repositioned according to
x \ + 1 = ( 1 - X)X] + XXi (E q . 5 .3 0 )
where Xt is the position of the vertex i , t is the iteration number and A . is a smoothing
parameter in the [0,1] interval (usually X - 0 .7 5 ). Xj is the weighted average of the
centers of all triangles that include vertex X{ and is calculated as
(C q. 5 .3 1 )
' j 6 %
where a \ is the total area of all triangles that contain vertex X t , is the set of all such
triangles, a- is the area of triangle j and C-is the center of triangle j . The degree of
smoothing applied should be sufficient to eliminate all the obvious irregularities, but not
so much as to seriously distort the cortical folds. Figure (5.16) shows a sample cortical
surface before and after applying the smoothing relaxation operator.
73
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 5.16. (a) Tessellated cortical surface, (b) Same surface after applying smoothing relaxation
Inflating the topologically corrected cortical surface through minimization of its convexity
eventually results in a closed convex surface. In this experiment we were interested in
obtaining a planar flat map of the cortex, so rather than the closed surface of the entire
brain we used the open surface of a half brain. Being a folded surface, cutting the cortex
with a blade parallel to one of the coordinate planes, even if it passes through the inter-
hemispheric fissure, does not necessarily generate two open cortical hemispheres. To find
the proper cutting angle we used the semi-inflated brain (Figure (5.17.a)) and cut it with a
plane with variable orientation passing through a point located on the interhemispheric
fissure. The orientation that resulted in two open halves of the inflated brain was used to
cut the original cortex. The front and back views of the right cortical hemisphere obtained
in this way are displayed in Figure (5.17).(b-c).
74
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
„ _ . L*=_-
' V ,
Figure 5.17. (a) Inflated human cortical surface, (b) The front view of a cortical hemisphere after applying
a few iterations of relaxation operation to the cortex, (c) The back view of the same cortical hemisphere.
The merit function used for developing the isometric flat map of the cortical hemisphere
consists of four components.
1 = Mc + M, + P a 7 « +Vp ( E q -5 - 3 2 )
The convexity cost function / needs to be included in this case because unlike the
cylinder or sphere, the cortex is not a convex surface. The minimization routine starts with
a relatively large value of (3 C (Pc = 105P5 and P4 = Pa ) and with every decade of
decrease in the convexity Pc . is decreased by a factor of 10. As the surface gradually
unfolds, its total convexity continuously decreases. Once the total convexity falls below a
pre-defined threshold, we assume the surface has been unfolded sufficiently. After that, Pc
is kept constant. At this stage, due to the small relative weights of area preserving and
angle preserving constraints during the unfolding process, a significant amount of
distortion has been introduced to the surface. This distortion can be reduced by trying to
minimize the difference between the angles and area of the triangles in their current
configuration and their configuration in the original surface through increasing the relative
values of P5 and Pa with respect to Pc . Increasing P4 and Pa has to be done gradually and
in steps. An abrupt increase in the weights of any of the two constraints will move the
surface in the solution space so rapidly that it may exhibit unstable behavior. As with the
75
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
half-cylinder and the hemisphere we use an adaptively changing as described in
Section (5.5.1.3), to apply a small perturbation to the surface when the algorithm traps in
local minima. Figure (5.18) through Figure (5.24) displays samples of the output
generated by the algorithm during the course of development of the flat map of the cortical
hemisphere. In this series of pictures, in each figure, the left column in the top row shows
the convexity map projected on the current configuration of the surface. The left column in
the second and third rows show the areal and angular distortion maps projected on the final
flat map of the surface. The right column shows the histogram of convexity, areal error and
angular error, respectively from top to the bottom.
76
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Convexity Map
•0.25 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 0.25
Histogram of Convexity Distortion
0.2 0.4 0.6 0.8
7FT 7aT
Areal Distortion (Relative Error ■ 35.32%)
•0.25 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 0.25
T cf
Histogram of Areal Distortion
0.1 0.2 0.3 0.4
W
Angular Distortion (Relative Error s 32.76%)
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Histogram of Angular Distortion
0.2 0.4 0.6 0.8
w (e)
Figure 5.18. Distribution of convexity, areal and angular distortions of a typical right hemispherical human
cortex after 250 iterations with p r = io5 , P ( = l and P „ = i • (a) Convexity superimposed over partially
unfolded cortex, (b) Histogram of convexity, (c) Areal distortion superimposed over the flat map of cortex,
(d) Histogram of areal distortion, (e) Angular distortion superimposed over the flat map of cortex, (f)
Histogram of angular distortion.
77
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Histogram of Convexity Distortion
Convexity Map
Histogram of Areal Distortion Areal Distortion (Relative Error = 51.85%)
Histogram of Angular Distortion Angular Distortion (Relative Error s 28.61%)
(e) (f)
Figure 5.19. Distribution of convexity, areal and angular distortions of a typical right hemispherical human
cortex after 500 iterations with p c = to4 , P v = 1 and P „ = 1 ■ (a) Convexity superimposed over partially
unfolded cortex, (b) Histogram of convexity, (c) Areal distortion superimposed over the flat map of cortex,
(d) Histogram of areal distortion, (e) Angular distortion superimposed over the flat map of cortex, (f)
Histogram of angular distortion.
78
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Histogram of Convexity Distortion
Convexity Map
Histogram of Areal Distortion
Histogram of Angular Distortion Angular Distortion (Relative Error s 39%)
( e ) (f )
Figure 5.20. Distribution of convexity, areal and angular distortions of a typical right hemispherical human
cortex after 1000 iterations with p c = to3 , p s = l and P „ = i • (a) Convexity superimposed over partially
unfolded cortex, (b) Histogram of convexity, (c) Areal distortion superimposed over the flat map of cortex,
(d) Histogram of areal distortion, (e) Angular distortion superimposed over the flat map of cortex, (f)
Histogram of angular distortion.
79
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Histogram of Convexity Distortion
Convexity Map
Histogram or Areal Distortion Areal Distortion (Relative Error s 52.22%)
Histogram of Angular Distortion Angular Distortion (Relative Error s 48.12%)
(e) (f)
Figure 5.21. Distribution of convexity, areal and angular distortions of a typical right hemispherical human
cortex after 2500 iterations with [ic = 100, p s = l and |? n = l . (a) Convexity superimposed over partially
unfolded cortex, (b) Histogram of convexity, (c) Areal distortion superimposed over the flat map of cortex,
(d) Histogram of areal distortion, (e) Angular distortion superimposed over the flat map of cortex, (f)
Histogram of angular distortion.
80
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Histogram of Convexity Distortion
Convexity Map
Histogram of Areal Distortion
Angular Distortion (Relative Error s 33.89%) Histogram of Angular Distortion
(e) (f)
Figure 5.22. Distribution of convexity, areal and angular distortions of a typical right hemispherical human
cortex after 3000 iterations with p c = to, p s = to and p a = 1 0. (a) Convexity superimposed over partially
unfolded cortex, (b) Histogram of convexity, (c) Areal distortion superimposed over the flat map of cortex,
(d) Histogram of areal distortion, (e) Angular distortion superimposed over the flat map of cortex, (f)
Histogram of angular distortion.
81
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Convexity Map
••.25 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 0.25
Histogram of Convexity Distortion
-0.0 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1
Histogram of Areal Distortion Areal Distortion Relative Error s 21.95%)
Angular Distortion (Relative Error s 21.40%) Histogram of Angular Distortion
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Figure 5.23. Distribution of convexity, areal and angular distortions of a typical right hemispherical human
cortex after 3500 iterations with p (. = 1 0, P v = 100 and = 100. (a) Convexity superimposed over partially
unfolded cortex, (b) Histogram of convexity, (c) Areal distortion superimposed over the flat map of cortex,
(d) Histogram o f areal distortion, (e) Angular distortion superimposed over the flat map of cortex, (f)
Histogram of angular distortion.
82
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Histogram of Convexity Distortion
Histogram of Areal Distortion
Histogram of Angular Distortion
Angular Distortion (Relative Error = 17.65%)
(e) (f)
Figure 5.24. Distribution of convexity, areal and angular distortions of a typical right hemispherical human
cortex after 5000 iterations with p r = 1 , P s . = too and p n = too. (a) Convexity superimposed over partially
unfolded cortex, (b) Histogram of convexity, (c) Areal distortion superimposed over the flat map of cortex,
(d) Histogram of areal distortion, (e) Angular distortion superimposed over the flat map of cortex, (f)
Histogram of angular distortion.
83
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Histogram of Convexity Distortion
Histogram of Areal Distortion Areal Distortion (Relative Error a 6.26%)
Histogram of Angular Distortion
(e ) (f )
Figure 5.25. Distribution of convexity, areal and angular distortions of a typical right hemispherical human
cortex after 10000 iterations with P c = o , p s = 1000 and p o = 1000. (a) Convexity superimposed over partially
unfolded cortex, (b) Histogram of convexity, (c) Areal distortion superimposed over the flat map of cortex,
(d) Histogram of areal distortion, (e) Angular distortion superimposed over the flat map of cortex, (f)
Histogram of angular distortion.
84
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
5.6 Application: Multiresolution Representation of Cortical Surface
In many applications, using a multiresolution strategy can save computational time. In a
multiresolution framework, global shape changes are first made on a coarse resolution
representation of the surface, and then transferred to progressively finer-resolution
representations. The goal is to break down the processing of a complex surface into
multiple levels with manageable complexities.
To create a multiresolution representation, the surface of interest must be represented by a
regularly connected lattice in which all interior nodes have equal number of neighbors.
Currently, no general tessellation technique exists that is capable of creating such a lattice.
Potentially, one can resample the polyhedral mesh generated by available techniques and
thereby identify a set of nodes belonging to the surface tessellation that forms an
appropriately regular lattice. Unfortunately, there is not any reliable way to achieve this
objective in a single step, i.e., by direct resampling of a highly convoluted 3-D surface.
However, such a resampling is straightforward to achieve on a 2-D surface. Using the
isometric flat map of the surface one can resample the 2-D surface into a regular grid and
then by interpolation determine the corresponding 3-D coordinates for each node in the
resampled grid. The advantage of using isometric maps for this purpose, as opposed to
other types of flat maps, is that the side lengths and angles of the triangles will be
maintained when mapping back the coordinate data from the 2-D domain to the 3-D space.
The first step is to generate a regular hexagonal array of nodes that completely covers and
extends slightly outside the 2-D map of the surface. To perform the resampling we have to
determine the geometric relationship between every single node of the resampling grid
and the triangle in the original map that contains this node. Let us assume that the node gi
falls inside the triangle A a b c . The triangle A abc defines a local barycentric coordinate
system [26] in which the coordinates of gt reflect the fractional areas of the three triangles
85
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
formed by this point and any two of the three vertices of Aabc (Figure (5.26)). These
coordinates are given by
Affine
Transformation
Triangle in 3-D
S p ace
Triangle in 2-D
Domain
A
C z a t
A a = [(-£*) *(&-*,■)] ' n i (Eq. 5.33)
B
Figure 5.26. Geometric relationship used to interpolate
position of nodes during the resampling operation.
where ni represents the surface normal at point gi and Ar is the area of Aa b c . It can be
easily verified that if the areas Aa, Ab and Ac are all positive, then g. is contained within
the triangle Aa b c . We use this fact to determine which one of the points on the hexagonal
grid are inside the 2-D map. To determine whether each grid point g ; is interior or exterior
to the original 2-D surface, the closest surface point Sj to the grid point g- is identified
and Equations (5.33) to (5.35) are used to decide whether g t is contained within any of the
triangles sharing the vertex S -. For each resampled grid point lying within the original
surface, the corresponding location G{ in the 3-D space is determined by a local affine
transformation from its location in the 2-D domain. Assuming that the resampled grid
point g t is contained within the triangle Aabc the position of the corresponding node G;
is uniquely specified by the linear combination of the position vectors of the vertices of the
corresponding triangle AABC with coefficients determined by the barycentric coordinates
Aa, Ab and Ac . In other words [26],
The uniformly resampled reconstruction of the surface can then be downsampled by
deleting alternate rows of nodes and alternate nodes within each remaining row to create a
lower resolution representation of the surface.
G i = A c A + A bB + A cC
(Eq. 5.36)
86
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Based on this method we have created a multiresolution representation for the hemispheric
phantom and the cortical surface data set whose isometric flat maps are shown in Figure
(5.15) and Figure (5.25), respectively. Figure (5.27) displays the original phantom along
with the same surface when uniformly resampled in three different resolutions. Figure
(5.28) shows the cortical hemisphere at three different resolutions. A magnified picture of
a potion of the surface demonstrates the uniformity of the tessellation and the level of
detail of surface description.
H i
(a) (b)
•
4^
(c) (d)
Figure 5.27. Multiresolution representation of a hemisphere created from its isometric flat map.
87
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
'-"v. : ■ W -r‘ ' + .
ft.-* ; /J • •
, * it/i-
-‘5 r ~ t - r ' 4 w - m ~ -
I ' - ' s. ffi?L {
* “ isf
Figure 5.28. Multiresolution representation of a cortical hemisphere created from its isometric flat map.
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
c h a p t e r 6 Summary and Future Research
6.1 Conclusion
In this work we have developed a set of tools and methods for geometric modeling and
quantitative analysis of complex surfaces. Although in the process of this research we
focused on human cortical surface, the developed algorithms and methods are general and
equally well applicable to other objects. We presented a method for triangulation of the
surface of an object from a set of planar contours representing its constant coordinate
slices. Using examples, we showed that this method creates meshes that are comparatively
simpler than those generated by the Marching Cubes algorithm. This simplicity, which is
practically reflected as a reduction in the number of triangles constituting the mesh, results
in savings in processing time for numerical calculations on the surface.
We demonstrated how the polyhedral mesh created by this tool can be used for
approximating the surface with local quadratic patches that are appropriate for calculating
differential geometric properties of the surface. In particular we showed how minimum,
maximum, mean and Gaussian curvatures can be easily calculated from this local
description. We used this model to calculate the curvature of a typical human cortical
surface and showed that minimum and maximum curvatures can distinctly identify the
locations of sulci and gyri on the cortex. We also showed that the mean curvature can
provide a true image of the general pattern of cortical foldings. In addition, we calculated
the Gaussian curvature of the brain surface and observed that it is a sharply varying
89
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
quantity with large variance and mean close to zero. From this we concluded that the
human brain surface is not a developable surface but it can be semi-isometrically mapped
into a plane.
We also introduced a strategy for developing isometric flat maps of surfaces. Based on the
fact that every simultaneous conformal and equiareal mapping is an isometric mapping we
formulated a constrained optimization problem consisting of an objective function with a
minimum occurring at the unfolded configuration of the surface and two constraint
functions guaranteeing the preservation of all areas and angles in the surface. Because of
the non-convexity of this model and the possibility of the solution vector becoming stuck
in local minima, we have employed a heuristic scheme using a simple smoothing operator
that allows us to add a perturbation to the solution vector which facilitates escape from
local minima. A conjugate gradient method minimizes the combined energy function and
unfolds the surface gradually, deforming it to a flat form corresponding to the global
minimum of the total energy function.
We applied this method to data sets representing surfaces of both developable and non-
developable types. The result of our experiments confirmed that this method is capable of
generating isometric flat maps for developable surfaces and semi-isometric flat maps for
non-developable surfaces. We then used this algorithm on some physical data representing
the real human brain surface to obtain the semi-isometric flat map of the cortex.
Finally, as a sample application of the developed techniques, we showed how the isometric
flat map of a surface can be used for constructing a multiresolution representation for that
surface.
All the described methods were tested, first with computer generated phantom data for
verifying their correctness, and then with real physical data to demonstrate their
practicality and efficiency. All methods provided satisfactory results and hence we assert
that they are practically plausible.
90
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
6.2 Future Work
There are possibilities for improving the techniques that we have developed and
continuing research in areas that we have explored.
Surface tessellation based on coordinate curves seems to be a promising alternative to the
Marching Cubes algorithm with the advantage of generating simpler meshes with fewer
number of triangles. However, more comprehensive studies are required to evaluate the
performance of this method. Issues such as accuracy of the method in reconstructing the
detailed shape features of the surface and its complexity compared to other available
methods need to be investigated.
An available tool for calculating the curvature of cerebral cortex opens a number of
avenues for research. Exploring the correlation between the differential geometry of the
cortex and the development of cortical folding is an active area of research where this tool
can be fully utilized. Automatic identification of sulcal and gyral organization of the brain
is another area that can be explored for possible applications in image registration and
matching for inter-subject studies.
The isometric flat map provides many possibilities for effective inter-subject studies. It
allows visualization of the distribution of functional data over unfolded maps of the brain.
Also it can be considered as a parametric model for the cortical surface; a model
appropriate for applying geometrical transformations and apt for surface based registration
techniques.
91
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Bibliography
[1] G. Alvarez-Bolado, G. Rosenfeld, and L. W. Swanson (1995), “On Mapping Patterns
in The Embryonic Forebrain.” J. Comp. Neurol. 355, 237-295.
[2] D. P. Bertsekas (1995), Nonlinear programming, Athena Scientific, Belmont,
Massachusetts.
[3] J. D. Boissonnat (1988), “Shape Reconstruction from Planar Cross-Sections.”
Comput. Vision, Graph., Image Proc. 44, 1, 1- 29.
[4] C. Bohm, T. Greitz, D. Kingsley, B. M. Berggren and L. Olsson (1983), “Adjustable
Computerized Brain Atlas for Transmission and Emission Tomography.” Am. J.
Neuroradiol. 4:731:733
[5] S. T. Bok, The Histonomy o f the Cerebral Cortex, Amesterdam: Elsevier.
[6] M. Bomans, K-H Hohne, U. Tiede and M. Riemer (1990), “3-D Segmentation of MR
Images of the Head for 3-D Display.” IEEE Trans. Medical Imaging, vol. MI-9, pp. 177-
183.
[7] Y. Bresler, J. A. Fessler and A. Macovski (1989), “A Bayesian Approach to
Reconstruction from Incomplete Projections of Multiple Objects in 3D Domain.” IEEE
Trans. Pattern Anal. Mach. Intell. 11, 8, 840-858.
[8] M. E. Brummer, R. M. Mersereau, R. L. Eisner and R. R. J. Lewine (1993),
“Automatic Detection of Brain Contours in MRI Data Sets.” IEEE Trans. Med. Imaging,
vol. 12, 153-166.
[9] G. J. Carman, D. C. Van Essen (1985), “Computational Algorithms for the Production
of Two-Dimensional Maps of Cerebral Cortex.” Soc. Neurosci, Abstr. 11: 1243.
[10] G. J. Carman (1990), Mappings o f the Cerebral Cortex, Ph.D. dissertation, California
Institute of Technology.
[11] G. J. Carman, H. A. Drury and D. C. Van Essen (1995), “Computational Methods
for Reconstructing and Unfolding the Cerebral Cortex.” Cerbral Cortex.
[12] H. N. Christiansen and T. W. Sederberg (1978), “Conversion of Complex Contour
Line Definition into Polygonal Element Mosaics.” Comput. Graph. 12 3, 187-192
[13] H. Cline, W. Lorenson, S. Ludke, C. Crawford and B. Teeter (1988), “Two
Algorithms for Three-Dimensional Reconstruction of Tomograms.” M e d ic a l P h y sic s,
15(3):320-327.
[14] L. Collins, C. Holmes, T. M. Peters, A. C. Evans (1996), “Automatic 3-D
Segmentation of Neuroanatomical Structures from MRI.” Human Brain Mapping 3(3),
190-208.
92
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
[15] P. N. Cook and S. Batnitsky (1981), “Three-Dimensional Reconstruction from Serial
Sections for Medical Applications.” Proc. o f the 14th Hawaii Int. Conf. on System
Sciences, 2, pp. 358-389.
[16] P. N. Cook (1981), “Three-Dimensional Reconstruction from Surface Contour for
Head C.T. Examinations.” J. Comput. Assist Tomography 5,1.
[17] L. T. Cook, P. N. Cook, K. R. Lee, S. Batnitzky, B. Y. S. Wong, S. L. Fritz, J. Ophir,
S. J. Dwyer, L. R. Bigongiari and A. W. Templeton (1980), “An Algorithm for Volume
Estimation Based on Polyhedral Approximation.” IEEE Trans. Biomed. Eng. 27, 9, 493-
500
[18] A. M. Dale and M. I. Sereno (1993), “Improved Localization of Cortical Activity by
Combining F.EG and MEG with MRI Cortical Surface Reconstruction: A Linear
Approach.” J. Cognitive Neurosci. 5: 2, pp. 162-176.
[19] C. A. Davatzikos and J, L. Prince (1995), “An Active Contour Model for Mapping
the Cortex.” IEEE Trans. Medical Imaging, Vol. 14, No. 1.
[20] C. A. Davatzikos and R. N. Bryan (1996), “Using a Deformable Surface Model to
Obtain a Shape Representation of the Cortex.” IEEE Trans, on M edical Imaging 15: 785-
795.
[21] E. W. Dijkstra (1959), A Note on Two Problems in Connexion with Graphs.”
Numerische Mathematik 1: 269-271.
[22] M. P. Do Carmo (1976), Differential Geometry o f Curves and Surfaces, New Jersey,
Prentice-Hall Inc..
[23] H. A. Drury, D. C. Van Essen, C. H. Anderson, C. W. Lee, T. A. Coogan and J. W.
Lewis (1996), “Computerized Mappings of the Cerebral Cortex: A Multiresolution
Flattening Method and A Surface-Based Coordinate System.” J. Cognitive Neurosci. 8: 1.
pp. 1-28.
[24] M. J. Durst Letters (April 1998): “Additional Reference to Marching Cubes.”
Comput. Graph., 22(2):72-73.
[25] A. B. Ekoule, F. C. Peynn And C. L. Odet (1991), “A Triangulation Algorithm from
Arbitrary Shaped Multiple Planar Contours.” ACM Trans. Graph. 10, 2, 182-199.
[26] G. Farin (1993), Curves and Surfaces fo r Computer Aided Geometric Design, 3rd
ed., San Diego, CA Academic Press.
[27] B. Fischl, M. I. Sereno and A. M. Dale (1999), “Cortical Surface-Based Analysis II:
Inflation. Flattening, and a Surface-Based Coordinate System.”. Neuroimage 9(2): 195-
207.
[28] H. Fuchs, Z. M. Kedem and S. P. Uselton (1997), “Optimal Surface Reconstruction
from Planar Contours.” Commun. ACM 20, 693-702
[29] S. Ganapathy and T. G. Dennehy (1982), “A New General Triangulation Method for
Planar Contours.” Comput. Graph. 16, 3, 69-75.
93
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
[30] L. D. Griffin (1994), “The Intrinsic Geometry of the Cerebral Cortex,” J. ofTheor.
Biol. 166, 261-273.
[31] T. Greitz, C. Bohm, S. Holte and L. Eriksson (1991), “A Computerized Brain Atlas:
Construction. Anatomical Content and Applications.” J. Comp. Assist. Tomogr. 15(1):26-
38.
[32] B. Hamann (1993), “Curvature Approximation for Triangulated Surfaces.”
Computing, Pages 139-153. Springer-Verlag.
[33] S. C. Joshi, J. Wang, M. I. Miller, D. C. Van Essen and U. Grenander (1995), “On the
Differential Geometry of the Cortical Surface.” Proc. o f the SPIE Int. Sym. on Optical
Science, Engineering, and Instrumentation San Diego, CA, pp. 304-311.
[34] T. Kapur, E. Grimson, W. M. Wells and R. Kikinis (1996), “Segmentation of Brain
Tissue from Magnetic Resonance Images.” Medical Image Analysis, vol. 1, pp. 109-127.
[35] E. Keppel (1975), “Approximating Complex Surfaces by Triangulation of Contour
Lines.” IBM. J. Res. Dev. 19, 2-11
[36] E. Kreyszig (1959), Differential Geometry, Totonto: University of Toronto Press,
(Mathematical expositions; no. 11).
[37] W. E. Lorensen and H. E. Cline (1987), “Marching Cubes: A High Resolution 3D
Surface Construction Algorithm.” ACM SIGGRAPH Vol 21 No. 4.
[38] D. Luenberger (1989), “Linear and Nonlinear Programming,” 2nd Ed., Addison-
Wesley.
[39] S. D. Ma and A. Gagalowicz (1985), “Determination of Local Coordinate Systems
for Texture Synthesis on 3-D Surfaces.” Eurographics’85, C. E. Vandoni ed. Elsevier
Science Publishers.
[40] D. Meyers and S. Skinner (1992), “Surface from Contours.” ACM Trans. Graph. 11,
3, 228-258.
[41] L. Piegl and W. Tiller (1997), The NURBS Book, 2nd Ed., New York, Springer-
Verlag.
[42] J. L. Ringo (1991), “Neuronal Interconnection as A Function of Brain Size.” Brain
Behav. Evol. 38, pp. 1-6.
[43] P. E. Roland and K. Zilles (1994), Trends Neurosci. 17, pp. 458-467.
[44] J. O’Rourke and V. Subramanian (1991), On Reconstructing Polyhedra from Parallel
Slices, Tech. Rep. TR 008, Dept, of Computer Science, Smith College, Northampton,
Mass..
[45] S. Sandor, R. Leahy and B. Timsari (1996), “Generating Cortical Constraints for
MEG Inverse Procedures Using MR Volume Data”. In Aine, C.J., F lynn, E.R., Okada, Y.,
Stroink, G., Swithenby, S.J., and Wood, C.C. (Eds.) Biomag96: Advances in
Biomagnetism Research, Springer-Verlag, New York.
94
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
[46] S. Sandor and R. Leahy (1997), “Surface Based Labelling of Cortical Anatomy
Using a Deformable Atlas.” IEEE Trans, on Medical Imaging, Vol 16, pp 41-54.
[47] E. Schwartz and B. Merker (1986), “Computer-Aided Neuroanatomv: Differential
Geometry of Cortical Surfaces and an Optimal Flattening Algorithm.” IEEE Computer
Graphics and Applications, 36-44.
[48] E. L. Schwartz, A. Shaw, and E. Wolfson (1989), “A Numerical Solution to the
Generalized Mapmaker’s Problem: Flattening Nonconvex Polyhedral Surfaces.” IEEE
Trans. Pattern Anal. Machine Intell. Vol. 11, No. 9. pp. 1005-1008.
[49] R Shirley and A. Tuckman (1990), “A Polygonal Approximation to Direct Scalar
Volume Rendering.” Comput. Graph., 24(5):63-70.
[50] K. R. Jr. Sloan and J. Painter (1987), “From Contours to Surfaces: Testbed and Initial
Results.” Proc. o f CHI + G I’ 87 (Toronto, Canada), pp. 115-120
[51] K. R. Jr. Sloan and J. Painter (1988), “Pessimal Guesses May Be Optimal: A
Counterintuitive Search Result.” IEEE Trans. Pattern Anal. Mach. Intell. 10, 6, 949-955.
[52] B. I. Soroka (1981), “Generalized Cones from Serial Sections.” Comput. Graph.
Image Proc. 15, 154-166.
[53] G. Strang (1980), Linear Algebra and its Applications, Orlando, Florida 32887:
Academic Press Inc., 2nd ed..
[54] J. P. Synder (1987), Map Projections: A Working Manual, US Geological Survey
Professional Paper 1395.
[55] J. P. Synder (1993), Flattening the Earth: Two Thousand Years o f Map Projections,
University of Chicago Press.
[56] M. R. Spiegel (1959), Theory and Problems o f Vector Analysis and An Introduction
to Tensor Analysis, Schaum Publishing Co. New York.
[57] P. H. Todd (1982), “A Geometric Model for the Cortical Folding Pattern of Simple
Folded Brains.” J. Theor. Biol. 97, pp. 529-538.
[58] P. H. Todd (1986), “The Intrinsic Geometry of Biological Surface Growth,” Lecture
Notes in Biomathematics, Vol. 69 (Levin, S. ed.) Berlin Springer-Verlag.
[59] A. W. Toga, ed. (1990) Three Dimensional Neuroimaging, Raven.
[60] D. C. VanEssen and J. H. R. Maunsell (1980), “Two Dimensional Maps of the
Cerebral Cortex.” J. Comp. Neurol., 191, pp. 255-281.
[61] Y. F. Wang and J. K. Aggrawal (1985), “Construction of Surface Representation from
3D Volumetric Scene Description.” Proc. o f Computer Vision and Pattern Recognition,
San Francisco, California.
[62] W. Wells, R. Kikinis, E. Grimson, F. Jolesz (1996), “Adaptive Segmentation of MRI
Data.” IEEE Trans. Med. Imaging, vol. 12, pp. 59-69.
95
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
[63] E. Wolfson and E. L. Schwartz (1989), “Computing Minimal Distances on Arbitrary
Two-dimensional Polyhedral Surfaces.” IEEE Pattern Anal. Machine Intell. Vol. 11, No.
9,pp. 1001-1005.
[64] A. Worth, N. Makris, J. Meyer, V. Caviness and D. Kennedy (1997), “Automated
Segmentation of Brain Exterior in MR Images Driven bv Empirical Procedures and
Anatomical Knowledge,” Proceedings o f the 15th International Conference on
Information Processing in Medical Imaging, pp. 99-112. Springer-Verlag.
[65] M. J. M. Zyda and R. J. Allan (1987), “Surface Construction from Planar Contours.”
Comput. Graph. 11, 4, 393-408.
96
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Linked assets
University of Southern California Dissertations and Theses
Conceptually similar
PDF
Complexity -distortion tradeoffs in image and video compression
PDF
Computer-aided lesion detection in positron emission tomography: A signal subspace fitting approach
PDF
Automated segmentation and analysis of human cerebral cortex imagery
PDF
Intelligent systems for video analysis and access over the Internet
PDF
Design and analysis of server scheduling for video -on -demand systems
PDF
Contributions to efficient vector quantization and frequency assignment design and implementation
PDF
Design and performance analysis of low complexity encoding algorithm for H.264 /AVC
PDF
Design and applications of MPEG video markup language (MPML)
PDF
Color processing and rate control for storage and transmission of digital image and video
PDF
Contribution to transform coding system implementation
PDF
Contributions to image and video coding for reliable and secure communications
PDF
High-frequency mixed -signal silicon on insulator circuit designs for optical interconnections and communications
PDF
Adaptive video transmission over wireless fading channel
PDF
Error resilient techniques for robust video transmission
PDF
Energy and time efficient designs for digital signal processing kernels on FPGAs
PDF
Contributions to coding techniques for wireless multimedia communication
PDF
Advanced video coding techniques for Internet streaming and DVB applications
PDF
Compression, correlation and detection for energy efficient wireless sensor networks
PDF
A modular approach to hardware -accelerated deformable modeling and animation
PDF
Data compression and detection
Asset Metadata
Creator
Timsari, Bijan (author)
Core Title
Geometrical modeling and analysis of cortical surfaces: An approach to finding flat maps of the human brain
School
Graduate School
Degree
Doctor of Philosophy
Degree Program
Electrical Engineering
Publisher
University of Southern California
(original),
University of Southern California. Libraries
(digital)
Tag
engineering, electronics and electrical,OAI-PMH Harvest
Language
English
Contributor
Digitized by ProQuest
(provenance)
Advisor
Leahy, Richard (
committee chair
), Neumann, Ulrich (
committee member
), Ortega, Antonio (
committee member
)
Permanent Link (DOI)
https://doi.org/10.25549/usctheses-c16-624930
Unique identifier
UC11334861
Identifier
3110962.pdf (filename),usctheses-c16-624930 (legacy record id)
Legacy Identifier
3110962.pdf
Dmrecord
624930
Document Type
Dissertation
Rights
Timsari, Bijan
Type
texts
Source
University of Southern California
(contributing entity),
University of Southern California Dissertations and Theses
(collection)
Access Conditions
The author retains rights to his/her dissertation, thesis or other graduate work according to U.S. copyright law. Electronic access is being provided by the USC Libraries in agreement with the au...
Repository Name
University of Southern California Digital Library
Repository Location
USC Digital Library, University of Southern California, University Park Campus, Los Angeles, California 90089, USA
Tags
engineering, electronics and electrical