Top Banner
RICE UNIVERSITY Building a 3D Atlas of the Mouse Brain by Tao Ju A Thesis Submitted in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy Approved, Thesis Committee: Joe Warren, Professor Computer Science Ron Goldman, Professor Computer Science Richard G. Baraniuk, Victor E. Cameron Professor Electrical and Computer Engineering Lydia Kavraki, Noah Harding Professor Computer Science Houston, Texas April, 2005
112

Building a 3D Atlas of the Mouse Brain - Department of Computer

Feb 09, 2022

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Building a 3D Atlas of the Mouse Brain - Department of Computer

RICE UNIVERSITY

Building a 3D Atlas of the Mouse Brain

by

Tao Ju

A Thesis Submitted

in Partial Fulfillment of theRequirements for the Degree

Doctor of Philosophy

Approved, Thesis Committee:

Joe Warren, ProfessorComputer Science

Ron Goldman, ProfessorComputer Science

Richard G. Baraniuk, Victor E. CameronProfessorElectrical and Computer Engineering

Lydia Kavraki, Noah Harding ProfessorComputer Science

Houston, Texas

April, 2005

Page 2: Building a 3D Atlas of the Mouse Brain - Department of Computer

Building a 3D Atlas of the Mouse Brain

Tao Ju

Abstract

Building and studying 3D representations of anatomical structures, such as the brain,

plays an important role in modern biology and medical science. While 3D imaging

methods such as MRI and CT have been widely applied, 2D imaging methods such

as optical microscopy typically generate images with much higher resolution. In this

thesis I describe how to construct a high-resolution 3D atlas of the mouse brain from

2D microscopic images. In particular, I focus on using computer graphics techniques,

such as image registration to correcting tissue distortions and polygonal modeling to

build surfaces representing the partitioning of anatomical regions. The methods are

being applied to construct a high-quality polygonal atlas of an adult mouse brain

from 350 histological tissue sections. While the resulting brain atlas will contribute

to a larger project of building a spatial database of gene expressions over the mouse

brain, the methods described in this thesis are well suited for the general purpose of

building polygonal atlases of arbitrary anatomical structures from tissue sections.

Page 3: Building a 3D Atlas of the Mouse Brain - Department of Computer

Acknowledgements

I would like to express my greatest thanks to my advisor, Dr. Joe Warren, whose

broad knowledge and dynamic thinking guided me throughout the process of research

and thesis writing. I would like to thank Dr. Ron Goldman for his valuable sugges-

tions on both the contents of the thesis and my presentation skills. I would also like to

thank Dr. Richard Baraniuk and Dr. Lydia Kavraki, who have spent their valuable

time reading this work and attending the defense.

I would like to thank my collaborators at Baylor College of Medicine and University

of Houston: Dr. Carson, Dr. Chiu, Dr. Gregor Eichele Dr. Thaller, Dr. Kakadiaris

and Musodiq Bello. Without their help I could not have finished this work.

I would like to give my special thanks to my wife Ming whose patient love enabled

me to complete this work.

ii

Page 4: Building a 3D Atlas of the Mouse Brain - Department of Computer

Contents

Abstract i

List of Illustrations vi

List of Tables xii

1 Introduction 1

1.1 Brain Atlases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Brain Atlases in Gene Expression . . . . . . . . . . . . . . . . . . . . 2

1.2.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.2.2 2D brain atlas . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

1.2.3 3D brain atlas . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

1.3 Atlas Creation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2 Data Collection 12

2.1 Cryo-sectioning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.2 Tissue Distortions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

3 Smooth Volume Reconstruction 16

3.1 Related work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

3.2 Method Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3.3 Warp Filtering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

Page 5: Building a 3D Atlas of the Mouse Brain - Department of Computer

3.3.1 The algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

3.3.2 Comparison . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

3.3.3 Warp representation . . . . . . . . . . . . . . . . . . . . . . . 26

3.4 Computing image warps . . . . . . . . . . . . . . . . . . . . . . . . . 26

3.4.1 Decomposing 2D warps . . . . . . . . . . . . . . . . . . . . . . 27

3.4.2 Piecewise linear representation . . . . . . . . . . . . . . . . . . 29

3.4.3 Elastic deformation . . . . . . . . . . . . . . . . . . . . . . . . 29

3.4.4 Dynamic programming . . . . . . . . . . . . . . . . . . . . . . 31

3.4.5 Warp operations . . . . . . . . . . . . . . . . . . . . . . . . . 38

3.5 Reconstruction framework . . . . . . . . . . . . . . . . . . . . . . . . 40

3.5.1 Rigid-body alignment . . . . . . . . . . . . . . . . . . . . . . . 40

3.5.2 Elastic alignment . . . . . . . . . . . . . . . . . . . . . . . . . 43

3.6 Experimental Validations . . . . . . . . . . . . . . . . . . . . . . . . . 45

3.6.1 Using a synthetic volume . . . . . . . . . . . . . . . . . . . . . 46

3.6.2 Using serial sections . . . . . . . . . . . . . . . . . . . . . . . 48

3.7 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

3.7.1 Smoothness of the volume . . . . . . . . . . . . . . . . . . . . 56

3.7.2 Limitation of 2D warps . . . . . . . . . . . . . . . . . . . . . . 58

4 Surface Construction 60

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

4.1.1 Surface from curves . . . . . . . . . . . . . . . . . . . . . . . . 61

4.1.2 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

iv

Page 6: Building a 3D Atlas of the Mouse Brain - Department of Computer

4.1.3 Proposed method . . . . . . . . . . . . . . . . . . . . . . . . . 66

4.2 Surface generation for multiple materials . . . . . . . . . . . . . . . . 66

4.2.1 Algorithm overview . . . . . . . . . . . . . . . . . . . . . . . . 66

4.2.2 Projection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

4.2.3 Topology graph . . . . . . . . . . . . . . . . . . . . . . . . . . 69

4.2.4 Polygonalization . . . . . . . . . . . . . . . . . . . . . . . . . 73

4.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

4.3.1 Automatic construction . . . . . . . . . . . . . . . . . . . . . 78

4.3.2 Manual adjustments . . . . . . . . . . . . . . . . . . . . . . . 81

4.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

5 Future Work 85

6 Summary 87

Bibliography 88

v

Page 7: Building a 3D Atlas of the Mouse Brain - Department of Computer

Illustrations

1.1 Expression images from 4 genes taken at the same sagittal

cross-section of the mouse brain. . . . . . . . . . . . . . . . . . . . . 4

1.2 Atlas at the coarsest level of subdivision (a), subdivided three times

(b), and overlaid on its defining image (c). . . . . . . . . . . . . . . . 5

1.3 Example query using a 2D brain atlas for the genes with similar

expression patterns to the gene Slc6a3 in the lower midbrain (right),

and the results computed by the database (left). . . . . . . . . . . . 7

2.1 Facilities for cryo-sectioning (a) and microscopy imaging (b). . . . . . 13

2.2 Four coronal (i.e. vertically cut from front to back) sections from a

stack of histological sections of a mouse brain acquired by

cryo-sectioning. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.3 Synthetic cross-cut through the middle of 350 coronal brain sections

in the sagittal (a) and horizontal (b) direction. . . . . . . . . . . . . . 15

3.1 Reconstruction of a smooth 2D image from deformed 1D columns

using warp filtering. (a) The original grayscale image. (b) The same

image with distorted columns. (c) The reconstructed image. . . . . . 24

Page 8: Building a 3D Atlas of the Mouse Brain - Department of Computer

3.2 (a) A subset of lines connecting corresponding points in successive

distorted columns of Figure 3.1 (b). (b) The smoothed lines of

corresponding points after reconstruction. . . . . . . . . . . . . . . . 25

3.3 Warping from a source image (a) to a target image (b) by computing

only the vertical deformation (c) and by computing deformations in

both the horizontal and vertical directions (d). . . . . . . . . . . . . . 28

3.4 Comparison of a Manhattan path (a) and a restricted path (b) whose

slope is bounded between 1/2 and 2 and whose possible nodes

(colored dark gray) lie in a skewed square along the diagonal. . . . . . 32

3.5 Weights assigned to the entries in the error table ε for computing

functions Left(i, j), Down(i, j) and Diag(i, j). . . . . . . . . . . . . . 33

3.6 A test pattern warped by applying a non-coherent warp (a) and a

coherent warp (b) from Section 14 to Section 15 in Figure 2.2. . . . . 35

3.7 Two examples of warping from a source image s to a target image t

using the global polynomial transformation in the AIR package

(AIR(s)), the deformable registration implementation in ITK

(ITK(s)), and our dynamic programming method(DP (s)). . . . . . . 37

3.8 (a) Re-parameterized functions σ∗ and ψ∗ of the restricted path

(colored in gray) along the diagonal from (0, 0) to (n, n). (b)

Symmetric addition of two diagonalized warps (colored in gray). The

resulting warp is plotted in solid black. . . . . . . . . . . . . . . . . . 39

3.9 (a) A coronal section with the line of symmetry detected. (b) The

line of symmetry is aligned to the center of the image, while the

vertical displacement is yet to be determined. . . . . . . . . . . . . . 42

vii

Page 9: Building a 3D Atlas of the Mouse Brain - Department of Computer

3.10 Reference sagittal sections from Paxino’s atlas intersected with the

plane of a coronal section at parallel line segments (thick bars) and

the resulting reference CM on that plane. . . . . . . . . . . . . . . . . 43

3.11 A tissue section before (a) and after (b) applying the bilateral filter. . 44

3.12 (a) A dark pixel and its corresponding pixels on neighboring sections.

(b) Using a majority filter, the dark pixel is replaced with the average

pixel intensity of its neighbors. (c) A reconstructed coronal section.

(d) The same coronal section after the majority filter is applied to the

stack. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

3.13 (a) A sagittal cross-section of an MRI volume of the mouse brain. (b)

A coronal section from the MRI volume. (c) The same section in (b)

after synthetic distortion. (d) A sagittal cross-section of the MRI

volume after each coronal section has been randomly distorted. (e) A

sagittal cross-section of the reconstructed volume from distorted

sections using warp filtering. . . . . . . . . . . . . . . . . . . . . . . . 47

3.14 The l2-norm from each of the 256 coronal sections of an MRI volume

to the same section after synthetic distortion (top curve) and to the

reconstructed section after warp filtering (bottom curve). . . . . . . . 48

3.15 Rigid-body alignment of coronal sections, viewed as synthetic

cross-section cuts in the sagittal direction through the stack of all

sections. (a): the original sections. (b): sections aligned using their

lines of symmetry, whose centers of mass are initially positioned in

the middle of each image. (c): sections aligned vertically using

reference centers of mass. (d): a real histological sagittal section. . . . 50

viii

Page 10: Building a 3D Atlas of the Mouse Brain - Department of Computer

3.16 Comparison of sagittal cross-sections: the original stack of 350 brain

sections after rigid-body alignment (a1), the reconstructed volume

(a2), after applying the majority image filter (a3), and a real tissue

section from Paxino’s atlas [68] at sagittal plate No. 110 (a4). . . . . 53

3.17 Comparison of horizontal cross-sections: the original stack of 350

brain sections after rigid-body alignment (b1), the reconstructed

volume (b2), after applying the majority image filter (b3), and a real

tissue section from Paxino’s atlas [68] at horizontal plate No. 157 (b4). 54

3.18 The l2-norm between successive images in the stack of 350 coronal

sections before (top curve) and after (middle curve) rigid-body

alignment. The elastic warping error between successive sections

(bottom curve) is computed as the l2-norm between the warped

image gk ◦ Φk and image gk+1 for k = 1, ..., 349. . . . . . . . . . . . . 55

3.19 Closeup looks at the hippocampus region in the synthetic sagittal

cross-section of the rigid-body aligned stack (a) and the elastically

aligned stack with filter width d = 5 (b), d = 10 (c) and d = 20 (d).

The region is indicated by red boxes in Figure 3.16 (a1) and (a2). . . 57

4.1 Constructing a surface model from planar tissue sections: two

neighboring tissue sections annotated with anatomical regions

partitioned by boundary curves (a,b); a boundary surface between the

two planes that connect the boundary curves on the two sections (c). 61

4.2 Curve networks on two neighboring sections (a,b), an invalid surface

network with self-intersecting geometry (c), and three geometrically

correct surface networks with various topology (d,e,f). . . . . . . . . . 63

ix

Page 11: Building a 3D Atlas of the Mouse Brain - Department of Computer

4.3 Surface generation: boundary curves on two neighboring sections

(a,b), wedges partitioned by the orthogonal projections of the

boundary curves on the mid-plane (c), the default material graph (d)

and the triangulated surface (e), a modified material graph (f) and

the resulting surface (g). . . . . . . . . . . . . . . . . . . . . . . . . 68

4.4 The projection of curve networks from the two sections in Figure 4.2

(a,b) (a), the default material graph (b) and the two variations (c,d)

that generate the surface networks in Figure 4.2 (c,d,e). . . . . . . . . 71

4.5 Topology modification through wedge splitting: two neighboring

sections containing non-overlapping regions of the same material type

(a,b), their orthogonal projections (c) and the default material graph

(e) that results in two separate surfaces (f), the splitting of wedges on

the mid-plane (d) and the refined topology graph (g), the modified

topology graph (h) and the resulting surface topology (i) that forms a

tube connecting regions on the two sections. . . . . . . . . . . . . . . 73

4.6 (a) A triangulation of the projected regions on the mid-plane from

Figure 4.3 (c). (b) The default topology graph. (c) The modified

topology graph. (d) A portion of the topology grid corresponding to

the gray triangles in (a) generated from the default topology graph in

(b). (e) Contoured surface by creating one polygon for each dotted

edge in (d). (f) The topology grid generated from the modified

topology graph in (c). (g) Contoured surface by creating one polygon

for each dotted edge in (f). . . . . . . . . . . . . . . . . . . . . . . . . 75

4.7 Surface reconstruction of the mouse brain before (left column) and

after (right column) smoothing: side view (a,b), bottom view (c,d),

and cross-section view (e,f). The color legend is shown on the far

right (g). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

x

Page 12: Building a 3D Atlas of the Mouse Brain - Department of Computer

4.8 Sections from the first 20 cross-sections of the mouse brain (a), the

reconstructed surface network (b), the surface bounding Cortex (c)

and Olfactory bulb (d). . . . . . . . . . . . . . . . . . . . . . . . . . . 80

4.9 Surface of the Fiber tracks (a) and colored version using neighboring

anatomical structures (b). . . . . . . . . . . . . . . . . . . . . . . . . 81

4.10 Broken pieces of Ventricle resulting from automatic construction (a)

and after manual adjustment (b). . . . . . . . . . . . . . . . . . . . . 83

4.11 A hole in Ventricles resulting from automatic construction (a) and

after manual adjustment (b). . . . . . . . . . . . . . . . . . . . . . . . 83

xi

Page 13: Building a 3D Atlas of the Mouse Brain - Department of Computer

Tables

3.1 Performance comparison of three warping methods applied to the two

examples in Figure 3.7. The l2-norm is computed between the source

image (or warped source image) and the target image. . . . . . . . . . 38

3.2 The coordinates (in mm) of the landmarks in Paxino’s atlas, and the

indices of the coronal sections in the stack that contain the landmarks. 51

3.3 Parameters in the plane equation computed for each coronal section

and the resulting fitting error. . . . . . . . . . . . . . . . . . . . . . . 51

3.4 Comparison of performance for different filtering width d and the

corresponding mean and maximum smoothness measure Sk of the

reconstructed volume. . . . . . . . . . . . . . . . . . . . . . . . . . . 52

Page 14: Building a 3D Atlas of the Mouse Brain - Department of Computer

Chapter 1

Introduction

1.1 Brain Atlases

Recent years have witnessed the fast growth of medical imaging methods, such as light

microscopy, CT, MRI and PET. These imaging methods have been applied in biology

and medical science, and the vast amount of resulting data imaged from experimental

organisms play a critical role in the study of animal functions and diseases. Due to

the anatomical heterogeneity between different species and individuals, comparative

studies of data imaged from different animals often rely on a common template, called

an atlas. An atlas represents the average shape of an anatomical structure, and serves

as a road-map for analyzing imaged data from diverse species and individuals.

By far the most intriguing yet the least understood anatomical structure is the brain.

Understanding how the brain functions and falters will not only shed light on the fight

against fatal neurogenetic diseases that have caused millions of deaths, but will also

help to answer ultimate questions about ourselves. For this reason, building an atlas

of the brain has become particularly important, as the analysis of brain images often

hold the keys to brain science. In the past few decades, biological researchers around

the world have created brain atlases of various animals such as flies [4], monkeys [17],

rats [85], and mice [56, 72, 79, 68].

In our work, we focus on building atlases of the mouse brain. The mouse is a close

relative to human beings, and the mouse genome shares a large similarity with our

own. Studying the mouse brain will not only further our understanding of the human

brain, but will also help in the study of human genetic diseases. As a motivating

1

Page 15: Building a 3D Atlas of the Mouse Brain - Department of Computer

2

application, we shall see in the next section how 2D and 3D atlases of the mouse

brain are used for the study of gene expression patterns.

1.2 Brain Atlases in Gene Expression

1.2.1 Background

One of the major recent successes in biology has been sequencing the genome of vari-

ous organisms such as fruit flies, mice and human beings. Gene sequencing represents

the first step towards the larger goal of understanding the organization and function

of biological organisms at the molecular level. In a recent project [18] that represents

the next logical step towards this larger goal, biologists are determining where each

gene in the genome is being expressed; that is, which cells are producing transcripts

for specific proteins. Using a method known as in situ hybridization, Dr. Eichele at

the Baylor College of Medicine is collecting gene expression data for all 30K genes in

the mouse genome.

This data (consisting of 2D images of the mouse brain taken at distinct cross-sections)

will play a key role in understanding the functional relationship between distinct

genes in the mouse genome. To aid in the organization and analysis of this data, we

have developed a geometric database that allows biologists to pose queries comparing

expression data for different genes. At the core of this database lies a set of 2D atlases

of the mouse brain that partition the brain into disjoint anatomical regions.

Gene expression images

Genes play a basic role in biology because genes serve as blueprints for creating pro-

teins, the building blocks of biological organisms. However, not all genes are actively

involved in protein synthesis. In a particular tissue, only a subset of the genes are

expressing (i.e. synthesizing) proteins. To retrieve gene expression patterns, biolo-

Page 16: Building a 3D Atlas of the Mouse Brain - Department of Computer

3

gists developed a technique called in situ hybridization that identifies cells expressing

a particular gene by means of antibody-antigen interactions. Given a section of a

mouse brain, in situ hybridization highlights the expression of a particular gene, and

the stained cross-section can be imaged using light-microscopy at high resolutions.

Several examples of gene expression images at a similar sagittal (i.e. vertically cut

from front to back) cross-section of different mouse brains are shown in Figure 1.1.

The dark blue regions in each image are where the corresponding gene is expressed.

Note that each image differs slightly due to rotations and translations induced dur-

ing data collection, as well as anatomical deviations between individual mice. One

feature of these images is that they all exhibit common anatomical regions such as

the cerebellum (the dark folded region in the upper right portion of the images).

Given gene expression images, biologists want to compare expression patterns between

different genes within a region of interest. For example, they would like to pose queries

of the form: “What genes have high expressions in the cortex region?” or “What genes

have similar expression patterns with gene X in the midbrain?” Answers to these

queries often hold the keys to discovering gene functions and gene networks. With the

plan to generate expression image for over 2000 genes in the next two years, automated

methods that accurately and efficiently compares expression patterns between a large

number of genes are in great demand.

To organize gene expression data into a searchable form, we have constructed a 2D

database of gene expressions using a set of 2D atlases of the mouse brain. The atlases

are constructed one on each of eleven standard cross-sections, each corresponding to a

sagittal section of the mouse brain with particular anatomical interest. Queries to this

database are of the form: “For a given cross-section, which genes have a particular

expression pattern over a specific anatomical region?” The atlases are represented

using subdivision, a geometric technique in computer graphics that allows smooth

modelling of anatomical boundaries as well as flexible deformation of the atlas onto

Page 17: Building a 3D Atlas of the Mouse Brain - Department of Computer

4

Gene: CRY1 Gene: GLUR1B

Gene: CHAT Gene: TLE3

Figure 1.1 : Expression images from 4 genes taken at the same sagittal cross-sectionof the mouse brain.

expression images.

1.2.2 2D brain atlas

Subdivision is a fractal-like process that takes a coarse mesh and generates a sequence

of increasingly fine meshes which converge to a smooth mesh at the limit. A brief de-

scription of the subdivision methods used in studying gene expressions is presented in

[46]. We refer interested readers to an excellent introduction to subdivisions methods

in general by Warren [91].

Page 18: Building a 3D Atlas of the Mouse Brain - Department of Computer

5

We model each standard cross-section of the mouse brain as a Catmull-Clark sub-

division mesh [19] partitioned by a network of crease curves. The left portion of

Figure 1.2 shows such a coarse mesh for one sagittal cross-section of the mouse brain.

(Crease edges are thickened; crease vertices are large dots.) The middle part of this

figure shows the mesh generated by three rounds of subdivision. Note that the crease

curves partition the mesh into 15 disjoint sub-meshes, each corresponding to an im-

portant anatomical region of the mouse brain. The right portion of Figure 1.2 shows

the crease curve network and the sample image used in laying out the coarse mesh.

By tagging each quad in the coarse mesh with its corresponding anatomical region,

the partitioned smooth mesh serves as an atlas for this standard cross-section.

(a) (b) (c)

Figure 1.2 : Atlas at the coarsest level of subdivision (a), subdivided three times (b),and overlaid on its defining image (c).

Representing the brain atlas as a subdivision mesh offers several desirable features

for comparing gene expression patterns:

1. The mesh explicitly models the partitioning of the brain into anatomical regions

by smooth boundary curves, which allows queries to the associated database of

gene expression data to be restricted to anatomical regions without effort.

2. Each sub-mesh corresponding to an anatomical region contains a smooth param-

eterization, which can be used for mapping gene expressions on tissue images.

Page 19: Building a 3D Atlas of the Mouse Brain - Department of Computer

6

3. The inherent multi-resolution structure of the subdivision mesh not only allows

easy deformation by manipulating the control vertices at the coarsest level, but

also enables fast multi-level comparison of expression data mapped onto the

atlas.

Geneatlas: A 2D geometric database for gene expressions

Using subdivision meshes as 2D brain atlases, we constructed Geneatlas.org, a web-

based database of gene expressions over the mouse brain [46]. For a particular stan-

dard cross-section, the gene expression database consists of the refined subdivision

mesh for the standard atlas annotated with gene expression information for each gene

in the database. For each particular gene, we first deform the subdivision mesh onto

the gene’s corresponding expression image using a combination of affine transforma-

tions and local least-square deformations (details are described in [46]). Next, for

each quad in the deformed mesh, we estimate the number of cells covered by that

quad and their corresponding levels of expression. Note that the use of the deformed

mesh corrects for anatomical variations in individual mouse brains.

To facilitate interaction with the database, Geneatlas.org features a graphical inter-

face to allow users to pose queries of the following form: “For a given region of the

brain, which genes have a particular expression pattern?” Users may specify the tar-

get region by name or by interactively painting the desired region onto the atlas.

Target expression patterns can either be uniform patterns such as high, medium, low

or none as well as expression patterns for a given gene. For example, on the left of

Figure 1.3, the user has painted a query onto the lower midbrain of the atlas at a

particular cross-section, and he wants to find those genes whose expression pattern is

similar to the gene Slc6a3, which exhibits a check-mark expression shape.

To compute the answer to a particular query, the database compares the gene ex-

pression data for various pairs of genes. For example, if the user desires those genes

Page 20: Building a 3D Atlas of the Mouse Brain - Department of Computer

7

whose expression patterns are similar to the ith gene in the database, the query com-

pares expression data of all genes in the database to the ith gene. By computing a

norm (such as L1) that measures the similarity between the two expression patterns,

the database then reports those genes whose norm with respect to the target gene is

smallest. For example, the right portion of Figure 1.3 shows eight genes computed by

the database that answer the query shown on the left. Note that each gene exhibits a

similar expression pattern to the target gene in the region selected by the user. Using

the multi-resolution structure of the subdivision atlas, the comparison computation

can be greatly accelerated by generalizing the multi-level search technique proposed

by Chen et al.[21] for rectangular images (details are described in [46]).

The ability to search for genes with desired expression patterns in regions of interest

has the potential to lead to significant medical discoveries. For example, the target

gene Slc6a3 in the sample query in Figure 1.3 is a dopamine transporter related to

Parkinson’s Disease. Among the eight genes computed by the database, as shown

on the right of Figure 1.3, seven are known to be dopamine-related. The gene Hbn1,

however, is a newly discovered gene with unknown functions. This query suggests

further investigation into the function of the gene Hbn1 and its possible relations

with dopamine transport or even with Parkinson’s Disease. In the face of the large

Figure 1.3 : Example query using a 2D brain atlas for the genes with similar expressionpatterns to the gene Slc6a3 in the lower midbrain (right), and the results computedby the database (left).

Page 21: Building a 3D Atlas of the Mouse Brain - Department of Computer

8

number of genes that are active in the brain, the ability to answer such queries

efficiently and accurately will play an important role in identifying potential genes of

interest in genetics research and drug design.

1.2.3 3D brain atlas

The 2D brain atlases on Geneatlas.org are fundamentally limited because these atlases

are only defined on a fixed set of cross-sections of the mouse brain. Consequently,

queries into the database are limited to 2D planes of the brain on which there is an

atlas defined. Ultimately, we would like to answer queries of the form: “Which genes

have a particular expression pattern within a specific anatomical structure in 3D?” To

answer such queries, we need to construct a 3D atlas of the mouse brain that stores 3D

gene expression patterns over the entire brain volume. The 3D atlas would serve as

a 3D database of gene expressions that support fully spatial queries. The availability

of this atlas-based database with gene expressions from a significant portion of the

mouse genome would contribute immensely to the bio-informatics community.

To be able to store and compare spatial biological data, such as gene expression

patterns, we seek a 3D atlas representation with the following features:

• The atlas should model accurately the boundaries of the 3D anatomical regions

of the brain. In particular, different anatomical regions modeled in the atlas

should not overlap or leave gaps; each point in space should belong to a unique

anatomical region (and consequently should have exactly one associated level

of gene expression).

• The atlas should provide a smooth parameterization interior to each anatomical

region. The parameterization provides a common, smooth coordinate frame in

which gene expressions can be accurately stored and compared.

• In face of the large volume of gene expression data in the database, the atlas

Page 22: Building a 3D Atlas of the Mouse Brain - Department of Computer

9

should support a multi-resolution structure on which the comparison of gene

expressions can be accelerated in order to answer online queries in real-time.

Recall that our 2D brain atlases represented as quad subdivision meshes have these

desirable properties. Each 2D atlas consists of quadrilaterals associated with different

anatomical regions, and two quadrilaterals from different regions are separated by a

crease edge. After subdivision, the refined atlas consists of smoothly parameterized

quadrilaterals within each anatomical region, and the boundaries between different

regions are modelled by a network of smooth crease curves.

Similarly, we shall represent the 3D brain atlas as a tetrahedral subdivision mesh.

The 3D atlas consists of tetrahedral elements associated with different anatomical

structures, and two tetrahedra from different anatomical structures share a crease

triangular face. Using tetrahedral subdivision techniques developed by Schaefer et al.

[77], we can subdivide a coarse tetrahedral mesh and generate a refined tetrahedral

mesh. The refined 3D atlas consists of smoothly parameterized tetrahedra within each

anatomical structure, and the boundaries between different structures are modelled by

a network of smooth triangular faces. The subdivided atlas is well suited for mapping

gene expression patterns and for answering queries comparing 3D expression patterns

of different genes.

1.3 Atlas Creation

While our ultimate goal in the motivating project on mouse gene expressions is to

build a tetrahedral subdivision atlas, we focus in this thesis on building a high-

resolution polygonal atlas, which will serve as the basis for the construction of the

final tetrahedral atlas. The polygonal atlas consists of a surface network that models

the partitioning of the brain into anatomical volumes. Each anatomical volume in

the polygonal atlas can be further tetrahedralized to yield a subdivision atlas.

Page 23: Building a 3D Atlas of the Mouse Brain - Department of Computer

10

Here we briefly outline the steps involved in creating a high-resolution polygonal atlas

of the mouse brain:

1. Collect tissue sections of the mouse brain with Nissl staining imaged using light

microscopy.

2. Correct distortions on each tissue section, such as stretching and compacting,

which are induced by cryo-sectioning.

3. Annotate each tissue section with anatomical regions, and connect 2D anatom-

ical boundaries from each section to form a 3D surface network modelling the

boundaries of various anatomical divisions.

Note that our atlas is constructed from cross-section images of the brain, which

offer a much higher resolution than conventional 3D imaging data such as MRI. The

increased resolution allows our atlas to represent fine anatomical features, such as the

tubular structures of Fiber tracks and Ventricles. However, the physical sectioning of

brain tissue induces distortions that need to be corrected in order to create a smooth

atlas representing the undistorted mouse brain.

In this thesis, I present robust, efficient and accurate methods for correcting distor-

tions in the tissue sections as well as for building a surface model from corrected brain

sections. Our method for correcting distortion makes use of the sectioning direction

to automatically reproduce a smooth brain volume from consecutive tissue images.

After manual annotation, we are able to construct a geometrically and topologically

correct surface network from boundary curves on annotated sections in a robust and

flexible manner.

As an on-going collaborative project with the University of Houston and the Baylor

College of Medicine, I will be actively pursuing further research in the construction

and utilization of atlases for the study of mouse gene expressions. As part of my

future research, I plan to investigate the construction of a volumetric atlas from the

Page 24: Building a 3D Atlas of the Mouse Brain - Department of Computer

11

polygonal brain atlas generated using the techniques in this thesis. The eventual goal

is to construct a geometric database of gene expression patterns using the volumetric

atlas. The techniques described in this thesis are well suited for building high-quality

surface models of complex anatomical structures from serial tissue sections, a topic

which has become increasingly important in modern biological research.

Page 25: Building a 3D Atlas of the Mouse Brain - Department of Computer

Chapter 2

Data Collection

In medical imaging, volumetric data generated by 3D imaging methods such as MRI

and CT have wide ranging applications in the visualization and analysis of organs. 2D

imaging methods, such as optical microscopy, typically generate serial sections with

much higher resolution than MRI or CT scans. Reconstructing these 2D sections in

3D has become an important tool for understanding anatomical structures in 3D, and

in particular, for building high-resolution atlases.

2.1 Cryo-sectioning

Brain images are collected on P7 C57BL/6 Mus Musculus Brain using cyro-sectioning

[18]. The brain is first extracted from the skull and put into solution to freeze.

The frozen brain is then cut coronally into serial cryo-sections each 25µm thick.

Tissue sections are placed on slides and then a Nissl stain is applied. After staining,

a coverslip is applied to each slide. The tissue section is then imaged using light

microscopy at the resolution of 3.3µm per pixel. Given that the mouse brain is

roughly 0.6 × 0.8 × 1.2cm, imaging a single mouse brain yields around 500 images

whose dimensions are 2000 by 3500 pixels. Pictures in Figure 2.1 are taken from Dr.

Gregor Eichele’s laboratory at Baylor College of Medicine for high throughput in situ

hybridization.

12

Page 26: Building a 3D Atlas of the Mouse Brain - Department of Computer

13

(a) (b)

Figure 2.1 : Facilities for cryo-sectioning (a) and microscopy imaging (b).

2.2 Tissue Distortions

Unfortunately, raw cryo-sections cannot be used directly for 3D reconstruction. The

tissue preparation steps for light microscopy may introduce undesirable tissue distor-

tions. Similar distortions that are specific to different preparation procedures have

been closely examined by several authors [10],[28],[78],[16]. Figure 2.2 shows four

coronal (i.e. vertically cut from side to side) sections of a single mouse brain acquired

by cryo-sectioning. Each slice may exhibit tissue distortions of the following forms:

• Global rotations and translations introduced when the tissue sections are placed

on the slides.

• Regional tissue deformations in the form of vertical stretching and compacting

induced both during cutting and when the section is placed onto the slide.

Extreme deformations may lead to tearing and folding.

• Image artifacts, such as dust (dark artifact) and air bubbles (black ring-shaped

artifact), introduced during coverslipping.

Page 27: Building a 3D Atlas of the Mouse Brain - Department of Computer

14

Section 14 Section 15

Section 140 Section 141

Figure 2.2 : Four coronal (i.e. vertically cut from front to back) sections from a stackof histological sections of a mouse brain acquired by cryo-sectioning.

Observe from Figure 2.2 that tissue distortions may vary significantly even between

neighboring sections. Due to random global and regional tissue distortions, stacking

successive images will not produce a coherent 3D volume. For example, Figure 2.3

shows two synthetical cross-sections through the middle of 350 coronal (i.e. vertically

cut from side to side) sections of a single mouse brain. Note that the images look

jagged and the boundaries of anatomical structures are far from smooth.

Page 28: Building a 3D Atlas of the Mouse Brain - Department of Computer

15

(a)

(b)

Figure 2.3 : Synthetic cross-cut through the middle of 350 coronal brain sections inthe sagittal (a) and horizontal (b) direction.

Page 29: Building a 3D Atlas of the Mouse Brain - Department of Computer

Chapter 3

Smooth Volume Reconstruction

To construct a 3D atlas representing smooth anatomical structures in the mouse brain,

we need to correct tissue distortions in order to generate a smooth 3D image from

tissue sections. In this chapter we present a robust and efficient method for producing

a smooth 3D volume from distorted 2D sections. The method automatically computes

deformations of tissue sections that result in a smooth volume when stacked. The

technique works in the absence of any undistorted references. Our approach is based

on computing image warps between adjacent section pairs and then smoothing the

pairwise warps.

3.1 Related work

There has been a tremendous amount of work on 3D reconstruction from distorted

serial sections. Typically, reconstruction is achieved by deforming individual 2D sec-

tions using image registration (warping) techniques. Here we briefly review related

work on 2D image warping and 3D reconstruction.

Image warping

Image warping is the task of finding the best deformation (warp) of a source im-

age that matches a target image under a specific distance measure. Image warping

methods have been intensively studied in the field of medical imaging, and we re-

fer interested readers to survey articles by Maintz and Viergever [58], by Glasbey

[35] and by Lester and Arridge [54], as well as books by Toga [86], by Hajnal et al.

16

Page 30: Building a 3D Atlas of the Mouse Brain - Department of Computer

17

[39], and by Modersitzki [63] for excellent reviews. Software packages implementing

state-of-the-art warping methods are also available for medical applications, such as

the Automated Image Registration (AIR) package from UCLA [94] and the Insight

Segmentation and Registration Toolkit (ITK) from NLM [97].

For the purpose of 3D reconstruction, we are interested in classifying image warping

methods by the nature of the resulting image deformations. A number of methods

compute global deformations of the source image, such as rigid-body transformations

[96, 57], linear affine transformations [70, 83], and higher-degree polynomial defor-

mations [95]. Since these deformations are global in nature, they do not perform

well in the case of local variations between images [35]. To overcome this problem,

other warping methods consider local image deformations. Bookstein [13] considers

thin-plate splines in image registration, which was later applied in conjunction with

maximization of Mutual Information by Meyer et al. [61] or minimization of points

correlation by Guest et al. [37]. Alternatively, free-form, B-spline based deformations

defined on a regular grid of control points have been studied by numerous authors

including Rueckert et al. [74] and Studholme et al. [82]. Recently, a hybrid model of

global and local deformations was also considered by Pitiot et al. [69] in a piecewise

affine setting. Local warping methods have more flexible mapping functions and are

often regularized by some form of elastic energy to prevent excessive deformations.

Common optimization techniques for computing a regularized solution include solv-

ing partial differential equations (an excellent review of such methods can be found

in the book by Modersitzki [63]), finite-element methods [37], graph methods such as

max-flow [73] and min-cut [14], stochastic methods [1], and dynamic programming.

Since its first successful application in speech recognition by Sakoe and Chiba [76],

dynamic programming has been known in 1D as Dynamic Time Warping (DTW) and

has been extended in various ways for warping 2D images [55, 26, 90, 71]. Compared

to other optimization methods in the continuous domain, dynamic programming finds

discrete minimizers in a robust manner. However, due to the exponential time com-

Page 31: Building a 3D Atlas of the Mouse Brain - Department of Computer

18

plexity of a fully 2D dynamic programming task, existing DTW-based image warping

methods can only handle extremely low-resolution images.

3D reconstruction

Previous methods on warp-based 3D reconstruction from serial sections can be ap-

preciated from two perspectives: the types of image deformations performed, and the

subjects from which the deformations are computed.

Over the past two decades, many authors have reported how global linear trans-

formations can be applied automatically to bring successive sections into alignment

[87],[53, 75, 78, 81, 67, 16, 7]. The simplicity of linear transformations not only reduces

the computational complexity but also allows convenient manual adjustment through

software interfaces [72, 22, 47]. However, as shown in several studies [28, 44, 15], the

section distortions induced by tissue preparation are local in nature. With the ad-

vance of image registration techniques, reconstruction methods have been proposed to

correct localized section distortions by using local 2D deformations [29, 36, 50, 85, 3]

as well as elastic 3D surface deformations [84, 60, 31, 33].

A large number of the above reconstruction methods compute the warp of each sec-

tion so that the warped section matches a neighboring section. For example, Durr et

al. [29] computes elastic deformation between pairs of distorted images, while Karen

et al. [47] uses software tools for user-assisted alignment between every two con-

secutive sections by rigid-body transformations. However, unlike image registration

where exact matching is preferred, the goal of reconstruction is to form a smooth vol-

ume allowing natural progression of features through successive images [36]. Recent

methods, and in particular elastic 3D surface deformation methods, focus on warping

distorted sections onto an undistorted reference, such as block-face photos [60, 50, 33]

or tissue markers [81, 7], 2D sections of a 3D in vivo image [84, 78, 67], or sections

from an existing template [40, 3, 85]. The problem with this approach is that in many

Page 32: Building a 3D Atlas of the Mouse Brain - Department of Computer

19

applications an un-sectioned reference is not always available.

There have been only a few works so far [64, 36, 93] that address the problem of

smooth 3D reconstruction using local, elastic 2D image deformations in the absence of

a reference volume (note that existing methods using elastic 3D surface deformations

such as [84, 60, 33] can not be applied due to the lack of a reference). Without

the knowledge of the original object before sectioning, these researchers based their

reconstruction on the following assumption: the shape of an anatomical structure

varies slowly with respect to section thickness. In other words, corresponding points

on adjacent sections are likely to be located close together. In a global approach,

both Guest and Baldock [36] and Wirtz et al. [93] aim at minimizing an energy

functional consisting of a distance measure between consecutive sections (e.g. spring

forces between corresponding points [36] or pixel-wise squared differences [93]) and an

elastic deformation potential on each section. Although numerical solutions can be

computed by using finite element methods [36] or by approximating the differential

equations using finite differences [93], solving the global minimization problem is

non-trivial due to the massive size of the system.

A completely different local approach was taken by Montgomery and Ross [64]. Their

idea is to reposition a point on each section by applying a local Laplacian smoothing

operator on the position of that point and the positions of corresponding points on

the two adjacent sections. However, their method requires manual delineation of

contours on each section to establish the correspondence. Moreover, the Laplacian

operator is only applied to points on the contours; hence the deformation is restricted

to the overall shape of a contoured region and has limited power in matching interior

features with neighboring sections.

Page 33: Building a 3D Atlas of the Mouse Brain - Department of Computer

20

3.2 Method Overview

Here we introduce a new local approach for reconstructing a smooth representative

volume by elastically deforming serial sections, which requires neither human inter-

vention nor the use of a reference volume. Since it is impossible to undo the tissue

distortions without the knowledge of the original object before sectioning, we base

our reconstruction on the same assumption as previous work [36] - that is, the sec-

tion thickness (25 µm) is small compared to variations in the shape of anatomical

structures. Hence the reconstruction goal is a smooth volume where a point on a

section lies close to its corresponding points on the neighboring sections. To achieve

this goal, we present:

• An automatic 3D reconstruction method, called warp filtering, based on 2D

image warps between each pair of adjacent sections. During the reconstruction,

each section is deformed using an average warp computed from the pairwise

warps within a group of neighboring sections. The average warp results in a

smooth volume by effectively repositioning each point on one section to the

weighted-average location of the corresponding points on neighboring sections.

The algorithm works with any 2D image registration techniques for computing

pairwise warps, and can be easily parallelized for speed. We performed quan-

titative and qualitative validation of the reconstruction method on both real

and synthetic data, and the results revealed the effectiveness of our method in

building a smooth volume from distorted sections.

• A new image warping algorithm based on dynamic programming for computing

regularized warps between adjacent serial sections. Due to the nature of sec-

tioning distortions, we consider a class of 2D warps that can be decomposed into

1D piecewise linear deformations with elastic constraints. The representation

of such a decomposable 2D warp greatly facilitates warp filtering. Moreover,

the decomposition allows us to extend a well-known 1D discrete minimization

Page 34: Building a 3D Atlas of the Mouse Brain - Department of Computer

21

method called Dynamic Time Warping [76] to compute 2D elastic image warps

even between images of high resolutions. Experimental results have shown

that the proposed method achieves an efficiency comparable to state-of-the-art

warping methods, while the resulting deformations often match successive tissue

sections with improved accuracy.

• A novel process for recovering the shape of the original object from deformed

sections using rigid-body transformations. The process requires minimal human

interaction, and precedes the above automatic reconstruction so that a naturally

shaped, coherent volume can be created.

We present the reconstruction method by warp filtering in Section 3.3 and describe our

image warping algorithm in Section 3.4. Section 3.5 presents the complete framework

for producing a coherent and naturally shaped volume from deformed tissue sections,

using semi-automatic rigid-body alignment and automatic warp filtering. In Section

3.6 we present our validation framework involving quantitative measures developed

from the evaluation criteria suggested by Guest and Baldock [36], and we report the

experimental results on a MRI test volume and a stack of 350 histology sections.

Finally, we discuss limitations of the method and future research in Section 3.7.

3.3 Warp Filtering

To reconstruct a volume in which corresponding points on successive sections are

located close to each other, warp filtering computes image warps that match each

section to a group of neighboring sections based on pairwise warps between successive

sections.

Page 35: Building a 3D Atlas of the Mouse Brain - Department of Computer

22

3.3.1 The algorithm

The input to the algorithm is a stack of N serial sections represented as 2D images

gk : R2 → R (k = 1, . . . , N) and pairwise warps represented as bivariate, bivalued

functions φk,k+1 : R2 → R2 (k = 1, . . . , N − 1) between every two successive images.

The warps φk,k+1 can be computed using any user-specified 2D image warping method

so that the warped image gk ◦ φk,k+11 best matches the neighboring image gk+1.

The algorithm computes new warps Φk : R2 → R2 (k = 1, . . . , N) that match each im-

age gk to a group of its neighboring images {gk−d, ..., gk+d} (d > 0). Φk is represented

as the following weighted average,

Φk =k+d∑

i=k−d

γiφk,i, (3.1)

where γi(i = k − d, ..., k + d) are binomially-distributed weights that approximate a

Gaussian filter,

γi = 2−2d

2d

i− k + d

,

and φk,i represent the warps from image gk to each image gi in the neighborhood

i ∈ [k − d, k + d]. Given the pairwise warps {φk−d,k−d+1, . . ., φk+d−1,k+d}, φk,i can be

constructed inductively as 2

φk,i =

φk,i+1 ◦ φ−1i,i+1, k − d ≤ i < k

φk,k, i = k

φk,i−1 ◦ φi−1,i, k < i ≤ k + d

(3.2)

where φk,k denotes the identity.

In effect, under the warp Φk, each point on gk is repositioned not to match exactly

the corresponding points on a single neighboring image, but to match the location of

1The operator ◦ denotes function composition: (f1 ◦ f2)(x) = f1(f2(x)).2If pairwise warps φk,k+1 are not invertible, reverse warps φk+1,k also have to be computed using

the chosen image warping method between successive image pairs.

Page 36: Building a 3D Atlas of the Mouse Brain - Department of Computer

23

the weighted-average of corresponding points on a group of images {gk−d, ..., gk+d}.Consequently, in the warped stack of images gk ◦ Φk (k = 1, . . . , N), high-frequency

noise along lines of corresponding points through successive images, which are often

induced by random sectioning distortions, are removed, and the distance between

corresponding points on adjacent images is reduced.

We illustrate warp filtering in a 2D example in Figure 3.1. Here we consider the

simplified problem of reconstructing a coherent 2D image from a sequence of deformed

1D columns of pixels gk : R → R. Figure 3.1 (a) shows a grayscale image with

letters “TMI”, and Figure 3.1 (b) shows the same image but with each column gk

synthetically distorted using a random local function. To reconstruct a smooth 2D

image, pairwise warps φk,k+1 : R → R that establish correspondences between points

on column gk to points on column gk+1 are computed between successive columns. In

Figure 3.2 (a), a subset of lines connecting corresponding points on successive columns

is plotted based on these pairwise warps. Observe that the lines are jagged due to

the local distortions that vary from column to column. Our algorithm computes the

filtered warps Φk from pairwise warps φk,k+1 and produces a stack of warped columns

gk ◦ Φk that reconstruct a smooth 2D image shown in Figure 3.1 (c). Observe in

Figure 3.2 (b) that, after warping, corresponding points on successive columns lie

close to each other and form smooth curves in space.

3.3.2 Comparison

The proposed algorithm is most closely related to the local approach taken by Mont-

gomery and Ross [64], but we present two major improvements. First, the correspon-

dence between successive sections, which was established manually using contour

lines in [64], is now computed automatically as image warps. Second, the simple

two-neighbor Laplacian operator on contour lines is replaced by a more general de-

noising Gaussian filter that operates on general warps between image pairs in a larger

Page 37: Building a 3D Atlas of the Mouse Brain - Department of Computer

24

(a)

(b)

(c)

Figure 3.1 : Reconstruction of a smooth 2D image from deformed 1D columns usingwarp filtering. (a) The original grayscale image. (b) The same image with distortedcolumns. (c) The reconstructed image.

neighborhood.

The improved local approach offers several unique advantages over the global mini-

mization methods used in [36] and [93]:

Simplicity: There is no need to set up and solve a 3D minimization problem. Our

local approach involves only 2D image warping and simple weighted-averaging.

In general, the algorithm can be used in conjunction with any 2D image regis-

tration technique for smooth 3D reconstruction.

Page 38: Building a 3D Atlas of the Mouse Brain - Department of Computer

25

(a)

(b)

Figure 3.2 : (a) A subset of lines connecting corresponding points in successive dis-torted columns of Figure 3.1 (b). (b) The smoothed lines of corresponding pointsafter reconstruction.

Stability: In contrast to global minimization over all sections, computing pairwise

warps between two individual sections is a minimization problem at a small

scale and therefore less prone to errors due to variations in the data. Such

stability is highly desirable due to the systematic and random nature of section

distortions.

Efficiency: Decomposing a global problem into local warping and filtering tasks

makes substantial performance increase possible through parallelization. In

particular, since the pairwise warp between each image pair as well as the filtered

warp for each image are computed independently, computation time can easily

be reduced linearly in a distributed computing environment. Such dramatic

performance improvements are not possible with global minimizers.

Page 39: Building a 3D Atlas of the Mouse Brain - Department of Computer

26

3.3.3 Warp representation

Although the warp filtering algorithm can be used with any image warping techniques,

efficient computation using formulas (3.1) and (3.2) requires representating pairwise

image warps φk,k+1 in an appropriate manner.

Regardless of how pairwise warps φk,k+1 are represented, the filtered warp Φk can

always be constructed implicitly by evaluating the righthand side of formulas (3.1)

and (3.2) at every point (x, y). This approach, however, needs to evaluate {φk−d,k−d+1,

. . ., φk+d−1,k+d} for every point in the image. For efficiency, using formula (3.1) and

(3.2), we can construct Φk as an explicit function, which can then be used to evaluate

Φk(x, y) directly at every point (x, y). The second approach requires the pairwise

warps φk,k+1 to be represented for the following operations on functions:

• Inversion: φ−1i,j , representing the warp from image gj to image gi.

• Composition: φi,j ◦ φj,k, representing the warp from image gi to image gk.

• Convex combination: a φi,j + b φi,k, representing a weighted average of the two

warps.

Although linear functions are feasible, warps represented as higher-degree global poly-

nomials or displacement fields in many local warping methods can not be easily

adapted to take advantage of these function operations. In the next section, we de-

scribe a new image warping method for computing a local, piecewise linear warping

function that is readily represented for efficient warp filtering.

3.4 Computing image warps

The problem of computing the optimal warp is ill-posed [63] and NP-complete [49];

hence regularization is required. For warping between successive serial cryo-sectioned

images, we consider specific types of regularized image warps such that

Page 40: Building a 3D Atlas of the Mouse Brain - Department of Computer

27

1. Each 2D warp can be decomposed into a single 1D deformation in the horizontal

direction and independent 1D deformations in the vertical direction for each

column, and

2. Each 1D deformation is represented as a monotonic piecewise linear function

with elasticity constraints.

As we shall see in the following discussion, such regularized warps are capable of

representing local deformations that characterize the differences between adjacent

serial sections. On the other hand, the regularization allows fast warp filtering by

supporting the three operations on warp functions (i.e., inversion, composition and

convex combination) and efficient warp computation using dynamic programming.

3.4.1 Decomposing 2D warps

Even if the sections have been rigidly aligned to correct the rotational and trans-

lational differences, differences between successive tissue sections still exist in the

form of anatomical variance, localized tissue distortion, and image artifacts. While

anatomical variances and image artifacts typically do not have specific orientation,

regional distortions in the form of stretching and compacting take place mostly in

the vertical (i.e., slicing) direction. To better represent the deformations caused by

sectioning distortions while accommodating other possible local variances, we shall

compute a restricted warp φ : R2 → R2 of the following form:

φ(x, y) = (φX(x), φY (x, y))

where φX , φY are single-valued functions. In other words, φX models an overall 1D

deformation in the horizontal direction (i.e., shifting of image columns) and φY models

independent 1D deformations in the vertical direction for different values of x (i.e.,

shifting of pixels in each column). A similar decomposition has been considered

previously by Agazzi et al. [1] for text recognition. Such decompositions greatly

Page 41: Building a 3D Atlas of the Mouse Brain - Department of Computer

28

simplify the task of warp computation, since 1D (vector) warps can be computed

much more efficiently and robustly than 2D (image) warps.

Previous authors, such as Cox et al.[26], have considered more restricted forms of 2D

warps. In their methods, each 2D warp consists solely of independent 1D warps on

each column (i.e., φX is the identity function). To illustrate the difference, Figure 3.3

shows a simple example of warping between two images containing rings of different

radii, a typical type of deformation between successive tissue sections. Figure 3.3

(c) shows the source image deformed using only 1D warps in the vertical direction,

and Figure 3.3 (d) shows the same source image deformed using the combination

of 1D warps in both vertical and horizontal directions. Observe that the horizontal

warp φX is necessary to model the non-vertical variances between source and target

images, which is required to handle anatomical differences and image artifacts in

tissue sections.

(a) (b)

(c) (d)

Figure 3.3 : Warping from a source image (a) to a target image (b) by computing onlythe vertical deformation (c) and by computing deformations in both the horizontaland vertical directions (d).

Page 42: Building a 3D Atlas of the Mouse Brain - Department of Computer

29

3.4.2 Piecewise linear representation

For convenient representations that will facilitate the function operations in warp fil-

tering, we consider 1D warps {φX , φY } that are monotonic, piecewise linear functions.

In particular, given two images s and t with n + 1 columns and m + 1 rows,

1. The horizontal deformation φX is represented by a pair of piecewise linear func-

tions σ, ψ : [0, K ∈ Z+] → [0, n] that match column σ(k) in image s to column

ψ(k) in image t for k = 0, . . . , K.

2. The vertical deformation φY is represented by a sequence of piecewise linear

function pairs σk, ψk : [0, L ∈ Z+] → [0,m] for k = 0, . . . , K that match the

point (σ(k), σk(l)) in image s to point (ψ(k), ψk(l)) in image t for k = 0, . . . , K

and l = 0, . . . , L.

We require the functions σ, ψ and σΣ, ψΣ3 to be invertible, so that the warp φs,t from

image s to image t can be represented as:

φs,t(x, y) = (σ(x), σx(ψ−1x

(y)))

where x = ψ−1(x) for x ∈ [0, n] and y ∈ [0,m] (linear interpolation is used when

subscript x assumes a non-integer value). The symmetry of σ and ψ allows the

inverse warp φt,s from image t to image s to be represented in the same way with

symbols σ and ψ exchanged. Such convenience will become crucial for performing

function operations during warp filtering.

3.4.3 Elastic deformation

To prevent excessive image deformation, an image is typically modelled as an elastic

material and image warps are constrained by some form of deformation energy. Due

3σΣ is short hand for the sequence σ0, σ1, . . . , σK .

Page 43: Building a 3D Atlas of the Mouse Brain - Department of Computer

30

to the piecewise-linear nature of our warp representation, we consider a discrete form

of deformation energy that consists of three terms:

1. 1st and 2nd order deformation energy in the X direction:

EX(σ, ψ) = αX

K∑

k=0

δ(k)2 + βX

K−1∑

k=0

(δ(k + 1)− δ(k))2

where δ(k) = σ(k)− ψ(k) and αX , βX are constant weights.

2. Independent 1st and 2nd order deformation energy in the Y direction for each

vertical warp:

EY (σk, ψk) = αY

L∑

l=0

δk(l)2 + βY

L−1∑

l=0

(δk(l + 1)− δk(l))2

where δk(l) = σk(l)− ψk(l), and αY , βY are constant weights.

3. Coherence between neighboring vertical warp functions:

EC(σΣ, ψΣ) = γK−1∑

k=0

L∑

l=0

((σk+1(l)− σk(l))2 + (ψk+1(l)− ψk(l))

2).

where γ is a constant weight.

Note that the inclusion of the coherence term in 3) is necessary to ensure a smooth

deformation field using a 2D warp that is decomposed into independent 1D warps.

Putting these terms together, the warping problem that we consider becomes the

minimization of the following error:

M(σ, ψ, σΣ, ψΣ) =K∑

k=0

MY (σ(k), ψ(k), σk, ψk) + EX(σ, ψ) + EC(σΣ, ψΣ) (3.3)

where MY is the warping error for each independent vertical warp under the l2-norm

(e.g., sum of pixel-wise squared differences):

MY (σ(k), ψ(k), σk, ψk) =L∑

l=0

(s(σ(k), σk(l))− t(ψ(k), ψk(l)))2 + EY (σk, ψk) (3.4)

Page 44: Building a 3D Atlas of the Mouse Brain - Department of Computer

31

3.4.4 Dynamic programming

To compute a minimizer of (3.3), we first need an efficient method for minimizing the

1D warping error in (3.4). Dynamic Time Warping (DTW) [76] is such a method that

robustly computes a discrete minimizer in 1D matching problems. Here we describe

how 1D DTW can be extended to warp 2D images by computing a discrete minimizer

of (3.3).

1D Warping

First we consider the 1D problem of computing the vertical warp pair (σk, ψk) between

column σ(k) in image s and column ψ(k) in image t, assuming the existence of σ and

ψ. DTW discretizes (σk(l), ψk(l)) (l = 0, ..., L) as a path on the integer grid from

(0, 0) to (m,m). The original DTW algorithm computes a Manhattan path from

(0, 0) to (m, m), as shown in Figure 3.4 (a). However, the resulting piecewise linear

functions (σk, ψk) are not invertible because these functions are piecewise constant.

To enforce invertibility, we adopt a technique similar to [52] that restricts the slope

of every segment of the path to be 1/2, 1 or 2. A sample path is shown in Figure 3.4

(b). Using DTW, the path with bounded slope that minimizes the 1D warping error

MY (σ(k), ψ(k), σk, ψk) defined in (3.4) can be found in three steps:

1. Construct a m + 1 by m + 1 error table ε, so that

εi,j = (s(σ(k), i)− t(ψ(k), j)2 + αY (i− j)2,∀ i, j ∈ [0,m]

2. Construct a m + 1 by m + 1 auxiliary table e, in which ei,j is the error of the

minimal-error path with bounded slope from entry (0, 0) to entry (i, j) in the

Page 45: Building a 3D Atlas of the Mouse Brain - Department of Computer

32

error table ε. We can compute ei,j inductively as follows:

ei,j =

Min

ei−2,j−1 + Left(i, j) + βY

ei−1,j−2 + Down(i, j) + βY

ei−3,j−3 + Diag(i, j)

, i > 0 or j > 0

εi,j, i = 0 and j = 0

∞, i < 0 or j < 0

(3.5)

where each function Left, Down, and Diag computes a weighted sum of entries

in ε according to the direction and length of the previous path segment, as shown

in Figure 3.5. Note that the sum of the weights is proportional to the length of

the path segment in each direction.

3. The minimal-error path that leads to em,m encodes the piecewise linear functions

(σk, ψk) : [0, L] → [0,m], where L is the number of path segments and eσk(l),ψk(l)

are entries along the path for l = 0, . . . , L.

(0,0)

(m,m)

(0,0)

(m,m)

(a) (b)

Figure 3.4 : Comparison of a Manhattan path (a) and a restricted path (b) whoseslope is bounded between 1/2 and 2 and whose possible nodes (colored dark gray) liein a skewed square along the diagonal.

The time and space complexity of both of the above algorithms are O(m2). However,

we can dramatically improve both the speed and the memory usage. First, εi,j and

ei,j need to be computed only for grid points i, j that lie within a skewed square

Page 46: Building a 3D Atlas of the Mouse Brain - Department of Computer

33

1

2/3

2/3

1/3

1/3 1

2/3

1/3

2/3

1/3

ji,ε

1-j2,-iε

ji,ε

2-j1,-iε

1

1

1

1/2

1/21/2

1/21/2

1/2

ji,ε

3-j3,-iε

Figure 3.5 : Weights assigned to the entries in the error table ε for computing functionsLeft(i, j), Down(i, j) and Diag(i, j).

of size (m + 2)/3 along the diagonal, as shown in Figure 3.4 (b). The restriction

effectively reduces the computation time by a factor of 9. Furthermore, we can limit

our computation to grid points {i, j} within a narrow band along the diagonal, whose

half-width is the maximum deformation ‖σk(l) − ψk(l)‖. When warping between

adjacent tissue sections, we typically observe a maximum deformation of less than

10% in both the horizontal and vertical directions; hence banding further reduces the

computation time by approximately a factor of 5.

2D Warping

The 2D warping task M(σ, ψ, σΣ, ψΣ) is computed in two stages. First, observing a

similar form on the right-hand side of the 2D minimization goal in (3.3) (ignoring

the coherence term EC) and the 1D minimization goal in (3.4), we consider functions

(σ, ψ) such that (σ(k), ψ(k)) (k = 0, ..., K) form a path on the integer grid from (0, 0)

to (n, n) with bounded slope. Similarly, using DTW, such a path that minimizes

the sum of the vertical warping error∑K

k=0 MY (σ(k), ψ(k), σk, ψk) and the horizontal

elastic energy EX(σ, ψ) can be found in three steps:

1. Construct an n + 1 by n + 1 error table ε, so that

εi,j = MY (i, j, σi, ψj) + αX(i− j)2,∀ i, j ∈ [0, n]

where (σi, ψj) is the minimal-error vertical warp between column i in s and

Page 47: Building a 3D Atlas of the Mouse Brain - Department of Computer

34

column j in t, computed using the previous 1D algorithm.

2. Construct an n+1 by n+1 auxiliary table e in the same way as in (3.5) except

the constant βY is replaced by βX .

3. The minimal-error path that leads to en,n encodes the piecewise linear functions

(σ, ψ) : [0, K] → [0, n], where K is the number of path segments and eσ(k),ψ(k)

are entries along the path for k = 0, . . . , K.

Next, given (σ, ψ), the vertical warps (σΣ, ψΣ) can be computed using the previous

1D algorithm. However, these minimal-error vertical warps yield 2D warps that may

vary significantly from column to column. For example, Figure 3.6 (a) shows the

result of applying a 2D warp computed from Section 14 to Section 15 in Figure

2.2, without considering the coherence among vertical warps, to a test pattern (a

sequence of uniformly spaced horizontal bars). Notice the high-frequency noise in the

warped test pattern due to the large disparity between vertical warps in neighboring

columns. Typically, the true 2D deformations induced during the sectioning process

are much smoother. To incorporate the coherence energy EC(σΣ, ψΣ), given (σ, ψ),

instead of computing a single best vertical warp (σk, ψk) for every k = 0, . . . , K, we

compute a subset of the low-error warps (σkΣ, ψkΣ

) in step 3 of the 1D algorithm.

Finally, we add an extra dynamic programming pass to choose the best warp (σk, ψk)

from each group (σkΣ, ψkΣ

) that minimizes the sum of the vertical warping error∑K

k=0 MY (σ(k), ψ(k), σk, ψk) and the coherence energy EC(σΣ, ψΣ) (each vertical warp

function must first be re-parameterized to have the same domain [0, L], see Section

3.4.5). Figure 3.6 (b) shows that the coherent warp applied to the same test pattern

generates exhibits a much smoother deformation.

The dominant operation in 2D warping is the construction of the error table ε in step 1,

which has time complexity O(n2m2). However, the computation can be implemented

efficiently by restricting the calculation to a subset of entries and using appropriate

banding, as described in 1D warping. Experimental results have shown that the speed

Page 48: Building a 3D Atlas of the Mouse Brain - Department of Computer

35

(a) (b)

Figure 3.6 : A test pattern warped by applying a non-coherent warp (a) and a coherentwarp (b) from Section 14 to Section 15 in Figure 2.2.

of our algorithm is comparable to other state-of-the-art methods while producing

higher quality warps.

Examples and comparisons

Figure 3.7 illustrates two warping the adjacent Nissl-stained cryo-sections shown in

Figure 2.2. In each example, the goal is to find the warp that deforms the source

section s to match the target section t. Sections are represented as grayscale images

of dimensions n = 850 and m = 670. We compare the result of our method 4 to the

results of two other popular warping methods: Automated Image Registration (AIR)

[94], and NLM Insight Segmentation and Registration Toolkit (ITK) [97]. For AIR, we

used the 182-parameter 12-degree global polynomial non-rigid transformation model.5

For ITK, we used the FEM-based local deformable registration method.6 The l2-norm

between the warped source image and the target image in each method as well as the

computation time are compared in Table 3.1. Note that our dynamic programming

4With elastic energy weights αX = αY = 0.00001, βX = βY = 0.02, γ = 0.005.5In our test, align warp program was used with model menu number (m) 32, threshold (t1,t2) 2,

and convergence threshold (c) 0.00001.6In our test, FEMRegistrationFilter was used with mean square metric, single-resolution, 8 pixels

per element, 40 iterations with elasticity (E) and density (RhoC) set to 105.

Page 49: Building a 3D Atlas of the Mouse Brain - Department of Computer

36

method achieves efficiency comparable to a global warping method, while the quality

of the resulting 2D warp is often better than a conventional local warping method

when deforming between successive tissue sections.

Observe from Figure 3.7 that the proposed warping method also handles extreme

deformations (e.g. sever folding in the second example) and image artifacts (e.g.

the big air bubble in the first example) in a reasonable manner. These random

artifacts and distortions are examples of incompatible features between the source

and target images, which imply that there does not exist a perfect warp that exactly

matches one image to the other. Such incompatible features may also include tissue

tears, or appearing(disappearing) anatomical features (e.g. the dense circular feature

in upper middle of the target image in the first example) that are typical through

successive tissue sections. The proposed warping method matches compatible features

between the two images well, while producing moderate modifications in places where

incompatibility arises without triggering excessive deformations.

Page 50: Building a 3D Atlas of the Mouse Brain - Department of Computer

37

t t

s ‖s− t‖ s ‖s− t‖

AIR(s) ‖AIR(s)− t‖ AIR(s) ‖AIR(s)− t‖

ITK(s) ‖ITK(s)− t‖ ITK(s) ‖ITK(s)− t‖

DP (s) ‖DP (s)− t‖ DP (s) ‖DP (s)− t‖

Figure 3.7 : Two examples of warping from a source image s to a target image tusing the global polynomial transformation in the AIR package (AIR(s)), the de-formable registration implementation in ITK (ITK(s)), and our dynamic program-ming method(DP (s)).

Page 51: Building a 3D Atlas of the Mouse Brain - Department of Computer

38

Example 1 Example 2

Time l2-norm Time l2-norm

Source Image (s) – 11717.1 – 17069.7

Polynomial transformation (AIR(s)) 36 s 5907.5 252 s 12419.5

Deformable registration (ITK(s)) 1325 s 5005.9 1190 s 14314.6

Dynamic programming (DP (s)) 165 s 3951.9 173 s 9207.4

Table 3.1 : Performance comparison of three warping methods applied to the twoexamples in Figure 3.7. The l2-norm is computed between the source image (orwarped source image) and the target image.

3.4.5 Warp operations

As we mentioned before, the piecewise linear 2D warp representation facilitates the

three operations involved in warp filtering (i.e. inversion, composition and convex

combination). First, however, we need to resolve a subtle problem that arises in our

dynamic programming approach. Horizontal warps (σ, ψ) between different images

may have different domains K, and even vertical warps (σΣ, ψΣ) in a same 2D warp

may have different domains L. For convenience, we re-parameterize every horizontal

warp σ, ψ : [0, K] → [0, n] onto a fixed domain [0, 2n] as follows:

σ∗(x) = σ(ω−1(x)), ψ∗(x) = ψ(ω−1(x)) (3.6)

where ω(x) : [0, K] → [0, 2n] is a monotonic, invertible mapping defined as ω(x) =

σ(x) + ψ(x). Intuitively, (3.6) re-parameterizes the two functions σ and ψ along the

diagonal from (0, 0) to (n, n), as shown in Figure 3.8 (a). Every vertical warp (σk, ψk)

can be diagonalized onto the domain [0, 2m] in a similar fashion.

Page 52: Building a 3D Atlas of the Mouse Brain - Department of Computer

39

)x(*σ

)x(*ψ

(0,0)

(n,n)

x

(0,0)

(n,n)

(a) (b)

Figure 3.8 : (a) Re-parameterized functions σ∗ and ψ∗ of the restricted path (coloredin gray) along the diagonal from (0, 0) to (n, n). (b) Symmetric addition of twodiagonalized warps (colored in gray). The resulting warp is plotted in solid black.

1D warp operations

• Inversion: Given a pair of 1D warps {σ, ψ}, the inverted warp {σ, ψ}−1 is easily

generated by exchanging the two functions as {ψ, σ}.

• Composition: The composition of two pairs of warps {σ, ψ} and {σ, ψ} with

domain [0, 2n], represented as {σ, ψ} ◦ {σ, ψ}, is computed as a new symmetric

warp {σ(ψ−1), ψ(σ−1)}. Note that the new warping functions are defined on the

domain [0, n], so diagonalization (3.6) must be performed to re-parameterize

σ(ψ−1) and ψ(σ−1) onto the domain [0, 2n].

• Convex combination: Given two pairs of warps {σ, ψ} and {σ, ψ} with do-

main [0, 2n], the sum a{σ, ψ} + b{σ, ψ} is easily computed as the symmetric

warp {aσ + bσ, aψ + bψ}. Note that the new warping functions preserves the

monoticity, invertibility and the same domain. Figure 3.8 (b) illustrates this

operation for a = b = 1/2.

Page 53: Building a 3D Atlas of the Mouse Brain - Department of Computer

40

2D warp operations

Let {σ, ψ, σΣ, ψΣ} be a symmetric 2D warp between two images, the inverted warp

is easily generated by exchanging the symbols σ and ψ.

Given two 2D warps {σ, ψ, σΣ, ψΣ} and {σ, ψ, σΣ, ψΣ}, the convex combination can

be computed simply as {aσ + bσ, aψ + bψ, aσΣ + bσΣ, aψΣ + bψΣ}. Composition of the

two warps can be performed as follows. First, the horizontal warp in the composite

2D warp, denoted by {σ, ψ}, is computed as the composite of the horizontal warps

{σ, ψ} ◦ {σ, ψ}. Then, the vertical warps in the composite 2D warp, denoted by

{σΣ, ψΣ}, are computed so that

{σk, ψk} = {σk1 , ψk1} ◦ {σk2 , ψk2}

where k1 = σ−1(σ(k)) and k2 = ψ−1(ψ(k)).

3.5 Reconstruction framework

Having described the algorithms for computing and filtering the image warps, we

are now able to present the complete framework for reconstructing a coherent vol-

ume from a stack of parallel coronally-sliced histological sections of a mouse brain.

Reconstruction proceeds in two steps: rigid-body alignment and elastic alignment.

3.5.1 Rigid-body alignment

We first perform an initial alignment of the sections using rigid-body transformations.

Sectioning the mouse brain not only induces local tissue distortions, but also causes

each tissue section to be randomly rotated and shifted. These global deformations

need to be undone before the elastic alignment step for the following two reasons:

1. Rotational differences between successive images can not be captured well by

our 2D warping algorithm, which is tailored to deformations characterized by

Page 54: Building a 3D Atlas of the Mouse Brain - Department of Computer

41

vertical variations within each column and horizontal shifting of columns.

2. Random translations and rotations of tissue sections disrupt the spatial con-

tinuity between successive sections in the un-sectioned brain, which makes it

hard to recover the correct and natural 3D shape of the brain.

To undo the global deformations, one can apply PCA (principal component analysis)

[43] to align all the sections by their centers of mass (CM) and their principal direc-

tions of variations. However, PCA-based methods typically cannot reliably determine

the orientation of an image with a near-circular shape, such as the coronal sections

of the brain. More importantly, since the brain is naturally curved, the CM of suc-

cessive sections of the un-sectioned brain form a curve in space instead of a straight

line. Aligning images by their CM will have the effect of turning a curved brain into

a straight tube. Therefore, we develop a new semi-automatic method for reliably

aligning coronal brain sections while preserving the original shape of an un-sectioned

brain.

Symmetry detection

We first use Marola symmetry measures [59] to detect the bilateral symmetry in each

coronal slice and align each image by identifying the lines of symmetry of the tissue

with the vertical mid-line of the image (see Figure 3.9). The bilateral symmetry of the

brain not only allows us to determine the exact orientation of each coronal slice, but

also restricts the translational variation between successive sections to the vertical

direction on each image. Symmetry detection and alignment is implemented as an

automatic algorithm followed by user inspection and correction.

Page 55: Building a 3D Atlas of the Mouse Brain - Department of Computer

42

(a) (b)

Figure 3.9 : (a) A coronal section with the line of symmetry detected. (b) The lineof symmetry is aligned to the center of the image, while the vertical displacement isyet to be determined.

Vertical alignment

The purpose of vertical alignment is to recover the shape of the un-sectioned object,

especially in the stacking direction (i.e. the direction orthogonal to the sectioning

plane). Since an un-sectioned volume is not available, we use the 31 sagittal histolog-

ical sections in the Paxino’s atlas [68] as the reference to align our coronal sections.

The idea is to compute a reference CM by cutting the reference sagittal sections with

the plane corresponding to an experiment coronal section (see Figure 3.10), and to

translate the coronal section in the vertical direction so that its CM agrees vertically

with the reference CM.

The main question is how to determine the plane corresponding to each experimental

coronal section in the coordinate system of Paxino’s atlas 7. Note that the orientation

of slicing during the sectioning process usually deviates from the standard coronal

plane in Paxino’s atlas due to the difficulty of positioning the brain during sectioning.

However, we do know that the coronal sections are parallel and uniformly spaced, and

thus we can represent the kth section by the plane αx+βy + γz + δ− k = 0 in which

7In Paxino’s atlas, we let the X axis be orthogonal to the sagittal planes, the Y axis be orthogonal

to the horizontal planes, and the Z axis be orthogonal to the coronal planes

Page 56: Building a 3D Atlas of the Mouse Brain - Department of Computer

43

Reference CM

Figure 3.10 : Reference sagittal sections from Paxino’s atlas intersected with the planeof a coronal section at parallel line segments (thick bars) and the resulting referenceCM on that plane.

the parameters α, β, γ, δ are to be yet determined.

Here we present a simple method for determining the plane equations for each coronal

section with minimal human assistance. First, a group of n landmarks specified by the

anatomists are located in Paxino’s atlas with coordinates (xi, yi, zi) for i = 1, . . . , n.

Next, we let the user determine the index si of the experimental coronal section on

which the ith landmark appears (note that the exact location of the landmark on the

section is not required). Finally, the parameters for the plane equations are computed

by minimizing the quadratic error:

E =n∑

i=1

(αxi + βyi + γzi + δ − si)2 (3.7)

3.5.2 Elastic alignment

After rigid-body transformations, we next need to elastically warp each coronal sec-

tion so that a coherent volume is formed. We first apply the warping algorithm

presented in Section 3.4 to compute the warps between successive images, and then

re-parameterize each image using the Gaussian-filtered warps presented in Section

3.3.

Image filters can be applied to further improve the quality of the volume reconstructed

Page 57: Building a 3D Atlas of the Mouse Brain - Department of Computer

44

by warp filtering. Since the distance between adjacent sections (25µm) far exceeds

the size of a cell (∼ 10µm), matching individual cells on two adjacent sections is not

practical. To avoid aligning cellular details while matching macro features (e.g. the

dark folds of the cerebellum), we can apply a smoothing filter on the tissue sections

before computing the pairwise warps. Although a Gaussian filter could be used, an

edge-preserving filter is more appropriate because boundaries between anatomical

structures are retained. Figure 3.11 shows the result of applying the bilateral filter

[88] that we used in our experiments (other filters may also be used). Note that the

filtered image exhibits clearer boundaries between different anatomical features. After

performing warp filtering on the bilaterally filtered images, the final warps would then

be applied to the original images for accurate reconstruction.

(a) (b)

Figure 3.11 : A tissue section before (a) and after (b) applying the bilateral filter.

Even after reconstruction, image artifacts such as tissue folds and air bubbles still

remain. To clean up the image artifacts, we found that an effective solution is to

apply a majority filter to corresponding points through successive images based on

the pairwise image warps. For each pixel on an image, if the intensity of the pixel

is the highest or the lowest among its corresponding pixels in a group of nearby

images, we replace the pixel’s intensity by the average intensity of its neighbors (see

Figure 3.12 top). The filter removes outliers along matched pixels through different

images, which are often introduced by bubbles or tissue folding on a single image.

Page 58: Building a 3D Atlas of the Mouse Brain - Department of Computer

45

The bottom of Figure 3.12 shows a coronal section in the reconstructed volume before

and after the majority filter is applied. Observe that the dark folds in the original

image are much less apparent in the filtered image. The purpose of majority filtering

is to ensure a smooth appearance to the reconstructed volume, which is ideal for

overall visualization but may not be suitable for analysis of fine anatomical details.

For efficient implementation, majority filtering of pixel intensities can be performed

at the same time as the Gaussian filter is applied to the pairwise image warps.

(a) (b)

(c) (d)

Figure 3.12 : (a) A dark pixel and its corresponding pixels on neighboring sections.(b) Using a majority filter, the dark pixel is replaced with the average pixel intensityof its neighbors. (c) A reconstructed coronal section. (d) The same coronal sectionafter the majority filter is applied to the stack.

3.6 Experimental Validations

Here we report on the validation methods and the corresponding results for evaluating

the effectiveness of the proposed reconstruction method both qualitatively (i.e. by

visual examination) and quantitatively (i.e. by computing distance and smoothness

measures). These methods form a general framework for validating any warp-based

3D reconstruction algorithm. The evaluation is carried out on two sets of data:

a synthetic volume with known distortions, and real serial sections with unknown

Page 59: Building a 3D Atlas of the Mouse Brain - Department of Computer

46

distortions. All computations are performed on a commodity PC with 1.5GHz AMD

Athlon processor and 2.5GB memory.

3.6.1 Using a synthetic volume

We first apply our reconstruction method to a stack of synthetically distorted cross-

sections from an existing 3D volume. In our experiment, we use a MRI volume of

the C57BL/6J mouse brain that has been generously provided by the LONI group

at UCLA. The volume has dimension 2563 with a uniform spacing of 56µm. Figure

3.13 (a) shows a sagittal cross-section of the MRI volume. We take each of the

256 coronal cross-sections of the volume and apply a B-spline based local image

distortion that varies randomly from section to section. Figure 3.13 (b) and (c)

show one of the coronal sections from the MRI volume before and after the synthetic

distortion. Notice that deformation takes place mainly in the vertical direction in

order to simulate the actual cutting distortions in real cryo-sections. The sagittal

cross-sectional view of the distorted volume is shown in Figure 3.13 (d).

The reconstruction proceeds by first computing the pairwise warps between successive

sections followed by warp filtering with width d = 5. The complete computation took

742 seconds (12 minutes and 22 seconds), with 661 seconds spent on warp computation

and 81 seconds on warp filtering and final image deformation.

Visual validation

Figure 3.13 (e) shows the sagittal cross-section at the same location through the re-

constructed stack. Our reconstruction method removes the majority of high frequency

noises caused by synthetic distortions (see Figure 3.13 (d)) and is able to form smooth

and easily-recognizable anatomical structures.

Page 60: Building a 3D Atlas of the Mouse Brain - Department of Computer

47

(a)

(b) (c)

(d)

(e)

Figure 3.13 : (a) A sagittal cross-section of an MRI volume of the mouse brain. (b)A coronal section from the MRI volume. (c) The same section in (b) after syntheticdistortion. (d) A sagittal cross-section of the MRI volume after each coronal sectionhas been randomly distorted. (e) A sagittal cross-section of the reconstructed volumefrom distorted sections using warp filtering.

Page 61: Building a 3D Atlas of the Mouse Brain - Department of Computer

48

Quantitative validation

We further compute the l2-norm from each coronal cross-section in the original MRI

volume to the same section after synthetic distortion and after reconstruction; the

results are plotted in Figure 3.14. The l2-norm measures quantitatively how close

the reconstructed volume is to the real volume. Observe that the sum of all the

section-wise l2-norms is reduced by 46% after automatic reconstruction.

0

20

40

60

80

100

120

1 51 101 151 201 251

Coronal MRI Sections

L2

Dis

tan

ces

Figure 3.14 : The l2-norm from each of the 256 coronal sections of an MRI volumeto the same section after synthetic distortion (top curve) and to the reconstructedsection after warp filtering (bottom curve).

3.6.2 Using serial sections

We next test our method in reconstructing a 3D volumetric representation of the

cell-density for the mouse brain from serial sections. The input is a stack of 350

Nissl-stained images acquired by cyro-sectioning coronally a single frozen adult mouse

brain (data preparation is detailed in Chapter 2). Each image is 850× 670 pixels at

a resolution of 25µm per pixel. The reconstruction consists of rigid-body alignment

and elastic alignment using warp filtering.

Page 62: Building a 3D Atlas of the Mouse Brain - Department of Computer

49

Figure 3.15 (a) shows a synthetic cross-section cut in the sagittal direction through

the original stack of 350 coronal sections, and Figure 3.15 (b) shows a cross-section

cut at the same position after rigid-body alignment using the line of symmetry in

each coronal section. In Figure 3.15 (b), the CM of each coronal section is initially

positioned at the middle of the image. To determine the correct vertical alignment,

we let the user specify the indices of the coronal sections containing the 9 landmarks

whose coordinates (expressed in the Paxino’s atlas) are shown in Table 3.2. The plane

equations αx + βy + γz + δ − k = 0 are computed for the coronal sections that best

fit the landmarks at the specified planes. Table 3.3 compares the parameters and the

resulting fitting error E (as defined in equation 3.7) computed in two scenarios: using

standard coronal planes (i.e., set α = β = 0) and using arbitrary plane orientations.

Note that allowing deviation of the plane from the standard coronal direction reduces

the fitting error by 90.7%. In particular, a large non-zero value of β indicates a signif-

icant tilting of the slicing plane around the X axis (about 8.7 degrees) during physical

sectioning. Figure 3.15 (c) shows the result after aligning the sections vertically with

the reference CM computed by intersecting the computed planes with the sagittal

sections in Paxino’s atlas. Note that the CM of the aligned sections now form a curve

in space, and the sagittal profile of the stack resembles that of a real sagittal section

shown in Figure 3.15 (d).

Computing 349 pairwise warps between successive sections consumes 27MB and 976

minutes (16 hours and 16 minutes), averaging 168 seconds for each single warp. The

performance of the subsequent warp filtering stage at different filtering width d (de-

fined in Section 3.3.1) is summarized in Table 3.4. Since every pairwise warp φk,k+1 is

computed independently (and so is every filtered warp Φk), both warping and filtering

can be greatly sped up by distributing the computation of different warps to differ-

ent processors. Using this simple parallel scheme on a cluster of 16 processors, for

example, all pairwise warps would be computed within an hour, while warp filtering

would finish in approximately ten minutes at filtering width d = 20.

Page 63: Building a 3D Atlas of the Mouse Brain - Department of Computer

50

(a) (b)

(c) (d)

Figure 3.15 : Rigid-body alignment of coronal sections, viewed as synthetic cross-section cuts in the sagittal direction through the stack of all sections. (a): the originalsections. (b): sections aligned using their lines of symmetry, whose centers of massare initially positioned in the middle of each image. (c): sections aligned verticallyusing reference centers of mass. (d): a real histological sagittal section.

Page 64: Building a 3D Atlas of the Mouse Brain - Department of Computer

51

Landmarks X (Lateral) Y (Interaural) Z (Bregma) Section Index

1 −2.25 1.5 −4 121

2 2.25 1.5 −4 123

3 0 4.3 −2.5 152

4 0.6 4 −0.9 209

5 −0.6 4 −0.9 210

6 0 1.5 0.25 255

7 0 3 1.15 280

8 0.3 2 2.5 325

9 −0.3 2 2.5 327

Table 3.2 : The coordinates (in mm) of the landmarks in Paxino’s atlas, and theindices of the coronal sections in the stack that contain the landmarks.

Standard Coronal Plane Arbitrary Plane

α 0 0.29932

β 0 −4.90077

γ 31.9736 31.9222

δ 243.405 256.331

E 300.081 27.7732

Table 3.3 : Parameters in the plane equation computed for each coronal section andthe resulting fitting error.

Page 65: Building a 3D Atlas of the Mouse Brain - Department of Computer

52

Filter width d Time (min) Memory (MB) Mean Sk Max Sk

0 (Original stack) – – 85.43 1165.68

1 4.4 37 7.71 74.39

2 17.2 91 3.15 24.83

5 46.6 178 1.33 7.04

10 83.4 269 1.09 6.22

20 172.1 449 1.04 6.15

Table 3.4 : Comparison of performance for different filtering width d and the corre-sponding mean and maximum smoothness measure Sk of the reconstructed volume.

Visual validation

Figures 3.16 (a2) and 3.17 (b2) show the result of reconstruction at filtering width d =

20. Note that anatomical features become much more coherent in the reconstructed

volume. The reconstruction also recovers the shape of some key structures, such as

the folds in the cerebellum to the left and the hippocampus region in the middle;

these profiles are barely recognizable or completely illegible in the original stack. For

qualitative validation, we compare the reconstructed volume to real histology sections

from Paxino’s Atlas [68] at similar sagittal (Figure 3.16 (a4)) and horizontal (Figure

3.17 (b4)) locations. We find that the shapes of the anatomical structures recovered in

the reconstructed volume closely resemble the shapes of the corresponding structures

in real tissue sections.

Quantitative validation

Objective validation with serial sections is difficult due to the lack of knowledge of

the original object before sectioning. One commonly used technique is to measure

the correlation between adjacent sections before and after reconstruction, as done

Page 66: Building a 3D Atlas of the Mouse Brain - Department of Computer

53

Figure 3.16 : Comparison of sagittal cross-sections: the original stack of 350 brainsections after rigid-body alignment (a1), the reconstructed volume (a2), after applyingthe majority image filter (a3), and a real tissue section from Paxino’s atlas [68] atsagittal plate No. 110 (a4).

by Wirtz et al. [93]. However, as pointed out by Guest and Baldock [36], such

criteria is inappropriate since the purpose of reconstruction is not to exactly match

one section to the next section. Inspired by the smoothness measure proposed by

Guest and Baldock [36], we propose the following two-step procedure to measure the

smoothness of our reconstruction:

Correspondence evaluation

We first evaluate the quality of the pairwise warps φk,k+1 that establish the corre-

spondences between adjacent images gk, gk+1 for k = 1, . . . , N . This evaluation is

accomplished by computing the l2-norm between each warped image gk ◦ φk,k+1 and

target image gk+1, as plotted in Figure 3.18; the l2-norm reveals the quality of the

Page 67: Building a 3D Atlas of the Mouse Brain - Department of Computer

54

Figure 3.17 : Comparison of horizontal cross-sections: the original stack of 350 brainsections after rigid-body alignment (b1), the reconstructed volume (b2), after apply-ing the majority image filter (b3), and a real tissue section from Paxino’s atlas [68]at horizontal plate No. 157 (b4).

match. For comparison, we plot the l2-norm between successive images in the origi-

nal stack as well as in the rigidly registered stack. Observe that the pairwise warps

reduce the pixel-wise differences between neighboring sections substantially, by 60%

to 90% in comparison to the original stack. Correspondence evaluation is a neces-

sary step because the smoothness of corresponding features will make no sense if the

correspondences are computed incorrectly.

Page 68: Building a 3D Atlas of the Mouse Brain - Department of Computer

55

0

5000

10000

15000

20000

25000

30000

35000

1 51 101 151 201 251 301

Coronal Sections

L2

Dis

tan

ce

Figure 3.18 : The l2-norm between successive images in the stack of 350 coronalsections before (top curve) and after (middle curve) rigid-body alignment. The elasticwarping error between successive sections (bottom curve) is computed as the l2-normbetween the warped image gk ◦ Φk and image gk+1 for k = 1, ..., 349.

Smoothness evaluation

We compute a smoothness measure Sk on each section k as follows:

Sk =

∑ni=0

∑mj=0(B(i, j) + C(i, j)− 2A(i, j))2

nm(3.8)

where A(i, j) = Φ−1k (i, j) denotes the location of (i, j) in the warped image gk ◦ Φk,

B(i, j) = Φ−1k−1(φk−1,k(i, j)) and C(i, j) = Φ−1

k+1(φ−1k,k+1(i, j)) denote the locations of

the corresponding points of (i, j) on the two neighboring sections gk−1 ◦ Φk−1 and

gk+1 ◦Φk+1 after warping. As a reminder, φ refers to pairwise warps between adjacent

sections, and Φ refers to the filtered warp. Sk is similar to the CAM measure proposed

by [36], which effectively measures how close a point lies to its corresponding points on

neighboring sections. For comparison, the smoothness measure of each section before

reconstruction is computed by setting Φk−1, Φk and Φk+1 to the identity function.

Table 3.4 reports the mean and maximum smoothness measure Sk for all 350 sections

in the reconstructed volume for different filtering widths d. Observe that Sk decreases

Page 69: Building a 3D Atlas of the Mouse Brain - Department of Computer

56

by an order of magnitude just by warp filtering using a single neighbor on each side

of a section (i.e., d = 1). A reduction by two orders of magnitude is observed in both

the mean and maximum Sk for d ≥ 5. To interpret the numbers, we notice from (3.8)

that Sk reflects the average distance from every point A(i, j) to the mid-point of its

two corresponding points (B(i, j) + C(i, j))/2. For example, at d = 20, Sk = 1.04,

hence the average distance ‖(B(i, j)+C(i, j))/2−A(i, j)‖ = 0.51. In other words, in

the reconstructed volume, each point deviates on average from the middle of its two

corresponding points on the neighboring sections by only 0.51 pixels. These results

appear to improve on the experimental results of a global reconstruction method

reported in [36] that uses a similar measure of smoothness.

Figures 3.16 (a3) and 3.17 (b3) show the cross-section cuts of the reconstructed vol-

ume after incorporating the two image filters (the majority filter is applied to 3 neigh-

bors on each side of a given pixel). In comparison with Figures 3.16 (a2) and 3.17

(b2), computing warps on edge-enhanced images better reconstructs the anatomi-

cal boundaries, and majority filtering along lines of corresponding points through

successive sections improves the smoothness of the appearance.

3.7 Discussion

3.7.1 Smoothness of the volume

The smoothness of the reconstructed volume using our method is controlled by the

filtering width d. Inappropriate choices of d, however, may cause the method to

either flatten the highly-curved or sharp features (when d is too big) or fail to remove

the distortion errors (when d is too small). This problem of producing the desired

amount of smoothness in the reconstructed volume arises in all smoothness-based

reconstruction methods such as [36] and [93].

A good choice of d depends on the input data. In most cases, the scale of random tissue

Page 70: Building a 3D Atlas of the Mouse Brain - Department of Computer

57

distortions is small in comparison to the size of anatomical features, and the value

of d can be determined via experiments and visual evaluation. For example, Figure

3.19 shows the hippocampus region in a sagittal cross-section cut through the original

stack of 350 images and the reconstructed stack with filter width d = 5, 10, 20. Note

that although a choice of d = 5 already achieves a good local smoothness measure in

Table 3.4, the dark stripe at the top of hippocampus still exhibits low-frequency noise

inherent in the input data, which spans a larger neighborhood and requires a wider

filtering width. At d = 20, the waviness is gone and the region becomes smooth. Still,

the reconstructed volume needs to be examined to ensure that other key anatomical

features have not been overly smoothed.

(a) (b) (c) (d)

Figure 3.19 : Closeup looks at the hippocampus region in the synthetic sagittal cross-section of the rigid-body aligned stack (a) and the elastically aligned stack with filterwidth d = 5 (b), d = 10 (c) and d = 20 (d). The region is indicated by red boxes inFigure 3.16 (a1) and (a2).

Sometimes a reasonable choice of d fails to exist, for example, when the scale of the

largest distortion noise exceeds that of the smallest anatomical feature. To remove

large distortions in one place while preserving small meaningful features in other lo-

cations, we use a more flexible method by finding independent filter widths dk for

each section gk or even independent widths dk(i, j) for each individual pixel gk(i, j).

These local filter widths adapt to the regional noise-signal ratio based on the corre-

spondence information between adjacent sections. Such local choices of filter widths

can easily be incorporated into the warp filtering framework in (3.1) and (3.2). More-

Page 71: Building a 3D Atlas of the Mouse Brain - Department of Computer

58

over, edge-preserving filters, such as a bilateral filter [42], can be used in place of the

Gaussian filter during warp filtering to preserve sharp features in the reconstructed

volume that are often flattened out using the Gaussian filter.

3.7.2 Limitation of 2D warps

Any restricted image deformation, such as the regularized 2D warp described in Sec-

tion 3.4, has its limitations. Our warp representation models only those 2D deforma-

tions that can be characterized by non-uniform deformations in the vertical direction

and a simple uniform deformation in the horizontal direction. Eligible deformations

include translation, scaling, uniform growth/shrinkage of anatomical structures in the

horizontal direction, and sectioning distortions in the vertical direction. Deforma-

tions that are not well modelled include rotation and non-uniform growth/shrinkage

of anatomical structures in the horizontal direction. Although rotational differences

between adjacent sections can be eliminated using rigid-body registration, we do ob-

serve, in a few cases, non-uniform variance of anatomical structures in the horizontal

direction that are not modelled accurately by our method. Even for these rare cases,

our method behaves in a reasonable manner, as observed in the uniform appearance

of the graph of warping error in Figure 3.18. To better handle a wider range of

2D deformations, we are currently investigating more flexible ways of decomposing

2D warps into 1D warps, such as allowing columns to bend and break during the

horizontal deformation as discussed in [55, 90, 71].

While we will continue our research on more accurate and efficient warping meth-

ods, notice that comparing the quality of 2D warps generated by various methods

is difficult due to the lack of standard benchmark data. In an effort to promote

the development of benchmarks for 2D image warping methods as well as for 3D

reconstruction methods, we have made the 350 histological sections in our exper-

iment as well as our preliminary reconstruction results available for download at

Page 72: Building a 3D Atlas of the Mouse Brain - Department of Computer

59

http://www.geneatlas.org/gene/data/histology.zip. We encourage those interested

in this problem to apply their methods to this data.

Page 73: Building a 3D Atlas of the Mouse Brain - Department of Computer

Chapter 4

Surface Construction

Once the distortions on the tissue sections have been corrected and a smooth image

volume has been generated, we can go on to build a surface model representing the

anatomy of the mouse brain. Given a stack of parallel 2D sections of an anatomical

structure, a 3D surface can be constructed in three steps:

1. Annotate each section with anatomical regions that share common boundary

curves. Figure 4.1 (a,b) show two neighboring annotated sections of the mouse

brain.

2. Connect 2D boundary curves on neighboring sections to form a 3D boundary

surface between the two sections. Figure 4.1 (c) shows an example of a boundary

surface between the sections in (a,b).

3. Concatenate the surfaces constructed in Step 2 between successive sections to

form the model of the entire anatomical structure.

In our experiment, annotation of the mouse brain sections is provided by anatomists

at the Baylor College of Medicine. Since surface concatenation is straightforward,

the key step is generating a surface between the curves on two neighboring sections.

Surface generation is a challenging task because the annotated sections may contain

boundary curves of complex topology and geometry, which may vary significantly

from section to section. For example, the two sections shown in Figure 4.1 (a,b)

are only 25 µm apart in the mouse brain, yet the complexity of annotation and the

60

Page 74: Building a 3D Atlas of the Mouse Brain - Department of Computer

61

(a) (b) (c)

Figure 4.1 : Constructing a surface model from planar tissue sections: two neighboringtissue sections annotated with anatomical regions partitioned by boundary curves(a,b); a boundary surface between the two planes that connect the boundary curveson the two sections (c).

difference in the annotated regions makes it hard even for a human to create the

intermediate surface.

Here we present a new surface construction technique that is capable of building

geometrically and topologically correct surface networks between annotated tissue

sections with arbitrarily complex boundary curves. Using this technique, we can

construct a high-resolution model of the mouse brain automatically from the 350

planar tissue sections that have undergone distortion-correction as in the previous

chapter. The new technique also allows experts to conveniently adjust the topology

of the reconstructed surface to comply with the topology of the actual brain.

4.1 Introduction

4.1.1 Surface from curves

A typical way of building a 3D surface model is from 2D curves on planar sections.

Here we consider a stack of uniformly spaced parallel sections. Each section contains

a non self-intersecting curve network that partitions the section into closed regions,

Page 75: Building a 3D Atlas of the Mouse Brain - Department of Computer

62

as shown in Figure 4.1 (a,b). In particular, each region is associated with a material

type. In our application, each annotated region is associated with its anatomical type

(e.g. cortex, cerebellum, etc.), which are represented by different colors in Figure 4.1.

The goal is to connect the curve networks on each section to form a surface network

in 3D that partitions the entire space into volumes associated with corresponding

materials. The surface network should possess the following properties:

1. Interpolation: the surface network should exactly interpolate the curve networks

on each section, and the material types of the volumes should agree with those

of the regions on each section.

2. Geometric correctness: the surface network should be a closed mesh partitioning

the space into disjoint volumes. In particular, the mesh should not contain self-

intersections or gaps.

3. Topological correctness: the topology of the surface network should agree with

the topology of the original object (e.g. the un-sectioned mouse brain in our

study) from which the curve networks are derived.

Note that the second property is particularly important for building a 3D atlas of the

mouse brain. To map experimental brain data on the annotated brain atlas, we must

associate any point in the atlas with a unique anatomical structure. This requirement

will be violated if the surfaces bounding different anatomical structures intersect or

leave gaps. A geometrically correct surface network will generate anatomical volumes

that are exact partitions of the brain.

Given curve networks on two parallel sections, there usually exist multiple, geomet-

rically correct, yet topologically distinct surface networks. For example, Figure 4.2

(d,e,f) show three topologically different surface networks connecting the curve net-

works in Figure 4.2 (a,b). Since it is not possible to infer the topology of the original

Page 76: Building a 3D Atlas of the Mouse Brain - Department of Computer

63

Top Bottom

(a) (b) (c)

(d) (e) (f)

Figure 4.2 : Curve networks on two neighboring sections (a,b), an invalid surfacenetwork with self-intersecting geometry (c), and three geometrically correct surfacenetworks with various topology (d,e,f).

object solely from the curve networks, manual adjustment is often necessary to pro-

duce topologically correct surfaces.

4.1.2 Background

Surface reconstruction from curve networks has been studied extensively in the past

three decades, and numerous methods have been proposed. In many fields the problem

is also known as contour interpolation or contour tiling, and the curve networks on

each section are often referred to as contour lines [48]. Here we attempt only to

provide a brief review of some of these methods, and we refer interested readers to

excellent surveys by Hagen [38] and by Sloan and Painter [80].

With few exceptions [92], previous methods are designed to build a surface between

sections that are partitioned by the curve network into regions of only two material

Page 77: Building a 3D Atlas of the Mouse Brain - Department of Computer

64

types: outside and inside (e.g. air and tissue or empty and solid). Early approaches

attempt to find a triangular tiling that connects the contour lines on the neighbor-

ing sections while optimizing a quality measure, such as the minimum surface area

criteria in the work of Fuchs et al. [30]. While Fuch’s and later approaches [24, 62]

produce fairly good looking surfaces connecting simple contour lines, the techniques

may generate surfaces with either self-intersections or gaps between complex sections

containing multiple and nested contours.

Recent work on surface reconstruction from two-material sections focuses on handling

curve networks of arbitrary topology and geometry. Boissonnat presents a method

based on Delaunay triangulations [12], which is refined by Geiger [34]. Cheng et

al. [23] improve Boissonnat’s method further to generate surfaces without having to

compute 3D Delaunay triangulations. Several researchers have used implicit functions

to interpolate between 2D contours, such as the method of Herman et al. [41] and

Csebfalvi et al. [27]. The variational approach of Turk et al. [89] combines the two

steps of building and interpolating implicit functions. Generating smooth surfaces

by solving Partial Differential Equations (PDF) has also been considered by Bloor

et al.[11] and later by Chai et al. [20], who further studied smooth connections of

neighboring surfaces sharing a common contour line. Finally, a group of researchers

([9, 6, 66, 51, 8]) have developed methods based on computing orthogonal projections

of two neighboring sections in order to construct surfaces with correct geometry and

reasonable topology. In particular, the methods of Oliva et al. [66] and Barequet

et al. [8] compute the areas of differences on the projected sections and triangulate

these areas using Voronoi diagrams [66] or Straight Skeletons [8].

The above methods all have severe difficulty in extending to sections containing re-

gions associated with multiple materials. A common approach [34] is to treat each

material type separately and build boundary surfaces for one material at a time.

This approach, however, will result in invalid geometry such as self-intersections. For

example, each of the two sections shown in Figure 4.2 (a,b) contains three materi-

Page 78: Building a 3D Atlas of the Mouse Brain - Department of Computer

65

als colored as white, green and purple 1. Building surfaces separately for the green

material and the purple material will result in self-intersecting geometry as shown in

Figure 4.2 (c).

To the best of our knowledge, the only method for surface reconstruction between

sections with multiple material types is the approach proposed recently by Weinstein

[92]. Weinstein’s method builds surfaces simultaneously for all materials using a

volumetric approach. Each section is first voxelized onto a uniform grid, where each

grid point is associated with a material type; then the surfaces between volumes

of different materials are generated using contouring on the voxel grid. Weinstein’s

approach produces geometrically correct surfaces for multiple-material sections that

may contain curve networks of arbitrary topology. However, due to the use of discrete

voxels, the surface produced by Weinstein’s approach only approximates rather than

interpolates the curve network on each section.

Most existing algorithms, and particularly recent ones for handling complex curve

networks, are designed to completely automate surface reconstruction, leaving little

room for the user to adjust the topology of the resulting surface network. The few

exceptions include the method of Christiansen and Sederberg [24], where user inter-

action is required to guide the triangulation in cases of excessive ambiguity. Software

packages, such as SURFdriver [65], allow a limited degree of user interaction dur-

ing surface reconstruction, such as capping and connecting regions. However, these

mechanisms for user interaction are restricted to sections containing two material

types. The topology of the surface network becomes much more complicated with

the increase in the number of materials due to the interaction between volumes of

different material types. In particular, topology modification methods that treat each

material type separately cannot guarantee to create a geometrically correct surface.

1In grayscale printing, green is dark gray while purple is light gray.

Page 79: Building a 3D Atlas of the Mouse Brain - Department of Computer

66

4.1.3 Proposed method

In the next section, we introduce a new technique for surface reconstruction in the

presence of multiple materials. Our method is able automatically to create a geo-

metrically correct surface with reasonable topology from serial sections containing

arbitrarily complex curve networks and any number of material types. Our technique

is inspired by the projection-based approach taken by Oliva et al. [66] and Barequet

et al. [8], but our method can handle multiple material types. Unlike Weinstein’s

method [92], our approach produces 3D surface networks that exactly interpolate the

2D curve networks on the input sections. For example, Figure 4.2 (d) shows the result

produced automatically by our method given the two sections in (a,b). Notice that

the purple volume extends into the green volume without causing self-intersecting

geometry.

Another novelty of our method lies in its ability to allow users to modify the topology

of the surface network in a convenient and creative manner. Our method is composed

of two phases: topology specification and geometry construction. While a default

surface topology is assumed based on orthogonal projections of the curve networks,

the user is free to modify this basic topology and to create a wide range of variations.

The geometry construction phase guarantees to create an intersection-free and gap-

free surface network consistent with the specified topology. For example, Figure 4.2

(e,f) show two variations to the surface network in (d), where the connectivity of the

purple and green volumes have been modified manually.

4.2 Surface generation for multiple materials

4.2.1 Algorithm overview

Like most other methods for building surfaces from parallel sections, our method first

computes a layer of surface network between every two neighboring sections and then

Page 80: Building a 3D Atlas of the Mouse Brain - Department of Computer

67

concatenates successive layers to form a complete model. Given two neighboring

sections (an example is shown in Figure 4.3 (a,b)), our method proceeds in three

steps:

1. Project the curve networks from the two sections orthogonally onto a common

plane, referred to as the mid-plane (Figure 4.3 (c)). The projected curves

partition the mid-plane into closed regions.

2. Each closed region on the mid-plane is the orthogonal projection of a cylindrical

volume (or a wedge) between the two sections. Create a topology graph, shown

in Figure 4.3 (d), that describes how each wedge is partitioned vertically into

slabs, and how slabs between neighboring wedges are connected.

3. Polygonalize the boundary between two connected slabs of different materials.

The polygonal surface network partitions the space into volumes corresponding

to connected slabs of the same material in the topology graph (Figure 4.3 (e)).

The key difference between our method and most other methods for two-material

surface construction is that instead of considering how 2D curves are connected to

form 3D surfaces, we consider how 2D regions on each section are connected to form

3D volumes. In our method, each volume consists of basic volumetric units called

slabs. In the topology graph, connected slabs of the same material represent a con-

tinuous volume, while connected slabs of different materials represent the interface

between neighboring volumes. The volumetric approach provides an intuitive way of

representing and modifying the material transition between sections, and naturally

results in a closed surface network without self-intersections or gaps.

The automatic surface generation produces a surface network that relates regions on

different sections with overlapping orthogonal projections. However, one can revisit

and modify the graph structure to alter how volumes are formed between the two

sections. For example, Figure 4.3 (f,g) show a modified topology graph from the

Page 81: Building a 3D Atlas of the Mouse Brain - Department of Computer

68

Top

(a)

Bottom

(b)

A

B C D

(c)

B C D

A

(d) (e)

B C D

A

(f) (g)

Figure 4.3 : Surface generation: boundary curves on two neighboring sections (a,b),wedges partitioned by the orthogonal projections of the boundary curves on the mid-plane (c), the default material graph (d) and the triangulated surface (e), a modifiedmaterial graph (f) and the resulting surface (g).

default graph in (d) and the resulting surface. The polygonalization step is guaranteed

to produce a geometrically correct surface network given any valid graph, while at

the same time ensuring exact interpolation of the curve networks on the two sections.

4.2.2 Projection

Given two neighboring sections, orthogonal projection of the curve networks onto the

mid-plane is performed in two steps. First, we compute intersection points of the

Page 82: Building a 3D Atlas of the Mouse Brain - Department of Computer

69

curves from different sections projected on the mid-plane. These intersections are

added to the existing set of vertices in the curve network on each section 2. Second,

closed regions are identified on the mid-plane by tracing closed boundary cycles in the

projected curve network. Note that a region may be bounded by multiple boundary

cycles, which can be detected using a scan-line based algorithm.

4.2.3 Topology graph

Each closed region partitioned by the projected curve networks on the mid-plane

becomes the cross-section of a wedge extending from the top to the bottom section.

Since the projection of a wedge on each section is a subset of some region on that

section, each wedge is associated with a distinct material at its top and bottom. This

observation suggests the vertical transition of materials within each wedge.

Definition

The topology graph is built to determine how the space between two sections is

partitioned into volumes of various materials. A node in the topology graph represents

a slab of volume within a wedge associated with a particular material. In the example

of Figure 4.3 (d), the nodes belonging to the same wedge are organized along a vertical

line to illustrate the vertical composition of slabs within a wedge. In particular, the

top and bottom slabs in each wedge are associated with the corresponding materials

on the top and bottom section, and successive slabs in the same wedge consist of

different materials.

The edges in the topology graph connect successive nodes within a wedge or nodes

between neighboring wedges. The edges determine how volumes are formed from

2To maintain topological connectivity between surface networks generated for successive sections,

the curves on each section are augmented by intersection points with curves on its two neighboring

sections

Page 83: Building a 3D Atlas of the Mouse Brain - Department of Computer

70

slabs and how volumes of different materials interact. Specifically,

• An edge connecting nodes of the same material (shown as thickened lines in

Figure 4.3 (c)) represents a continuous volume formed by the corresponding

slabs.

• An edge connecting nodes of different materials (shown as dotted lines in

Figure 4.3 (c)) represents a boundary between the volumes to which the corre-

sponding slabs belong.

By default, the topology graph contains edges connecting successive nodes in a same

wedge as well as the top and bottom nodes between neighboring wedges, as shown

in Figure 4.3 (c). The default graph directly relates two regions on different sec-

tions with overlapping projections: if two regions consist of the same material, they

are connected into a volume; otherwise, the regions generate two volumes sharing a

common boundary, as shown in Figure 4.3 (d).

Graph variation

The default topology graph can be modified by inserting nodes between the top

and bottom nodes of each wedge, and by adding edges connecting inserted nodes to

existing ones. The addition of nodes and edges effectively builds intermediate layers of

structures between the two sections, making it possible to construct complex surface.

For example, the default graph in Figure 4.3 (d) can be modified by inserting a node

associated with the white material to wedge C and connecting the new node to the

existing white nodes in neighboring wedges B and D, as shown in Figure 4.3 (f).

While the purple and green volume shares a common boundary in the default graph,

they are separated by a continuous volume of white material in the modified graph,

as reflected in the resulting surface network in Figure 4.3 (g).

Our topology representation allows the user to design complicated surface topology

Page 84: Building a 3D Atlas of the Mouse Brain - Department of Computer

71

A

B

CD

E

A

B

C

D

E

A

B

C

D

E

A

B

C

D

E

(a) (b) (c) (d)

Figure 4.4 : The projection of curve networks from the two sections in Figure 4.2(a,b) (a), the default material graph (b) and the two variations (c,d) that generatethe surface networks in Figure 4.2 (c,d,e).

by modifying a simple graph structure. Although an arbitrary number of nodes and

edges can be added to the graph, the addition should adhere to the following rules to

ensure a valid topology:

1. To prevent self-intersecting volumes, intersecting edges are prohibited in the

topology graph. Assume we number the successive nodes in each wedge from

bottom to top, if two edges connect the ith and jth node of one wedge to the

kth and lth node of a neighboring wedge, we require (i− j) ∗ (k − l) ≥ 0.

2. To prevent isolated volumes between two sections, each newly added node in

the topology graph must be connected to a top or a bottom node of some wedge

through a chain of edges between nodes with the same material type.

As a more complex example, Figure 4.4 shows the default topology graph and two

variations that generate the surface networks in Figure 4.2 (d,e,f). Note that the mod-

ification of the topology graph satisfies the above two rules to ensure non-intersecting

and well-connected volumes.

Page 85: Building a 3D Atlas of the Mouse Brain - Department of Computer

72

Wedge splitting

Modifying the topology graph by adding nodes and edges is confined to existing

wedges. To represent a larger class of surface topology (e.g., connecting regions with

non-overlapping projections), we allow an existing wedge to be split into smaller

wedges. The refined topology graph after wedge splitting allows for finer control over

the resulting surface topology. Moreover, as we shall see in the next section, splitting

is also used for generating the surface geometry.

We explain wedge splitting using the example of Figure 4.5. The two sections in (a,b)

contain regions whose orthogonal projections on the mid-plane, shown in (c), do not

overlap. As a result, the default topology graph in (e) generates two separate pieces of

surface, as shown in (f). Wedge splitting proceeds in two steps. First, new boundary

curves are added to the projected curve networks in (c) that divides wedge A into

two smaller wedges, D and E, as shown in (d). Second, the default topology graph

in (e) is refined in (g) to reflect the division of the wedges. Each newly created wedge

(e.g., D) inherits the nodes and edges from the original wedge A as well as edges

to A’s neighboring wedges (e.g., B). Meanwhile, corresponding nodes in neighboring

new wedges, D and E, are connected by edges.

Wedge splitting offers more freedom in designing variations of surface topologies be-

tween sections. For example, the refined topology graph in Figure 4.5 (g) after wedge

splitting can be further modified into (h), where nodes and edges are added to form

a continuous volume between the originally disconnected green nodes from wedges

B and C. The resulting surface shown in (i) forms a tubular structure between the

green regions on the two sections.

Page 86: Building a 3D Atlas of the Mouse Brain - Department of Computer

73

Top

(a)

Bottom

(b)

A

B C

(c)

D

B CE

(d)

B C

A

(e) (f)

B C

D

E

B C

D

E

D

(g) (h) (i)

Figure 4.5 : Topology modification through wedge splitting: two neighboring sectionscontaining non-overlapping regions of the same material type (a,b), their orthogonalprojections (c) and the default material graph (e) that results in two separate surfaces(f), the splitting of wedges on the mid-plane (d) and the refined topology graph (g),the modified topology graph (h) and the resulting surface topology (i) that forms atube connecting regions on the two sections.

4.2.4 Polygonalization

While the topology graph abstractly describes how the 2D regions on each section

are connected to form volumes, it remains to determine the geometry of the surface

that forms the boundary between adjacent volumes. Polygonalization is performed

by first refining the abstract topology graph into an actual volumetric grid, referred

to as the topology grid. Contouring this volumetric grid produces a closed surface

Page 87: Building a 3D Atlas of the Mouse Brain - Department of Computer

74

network that partitions the space between the two sections into volumes of desired

topology. Finally, the contoured surface is smoothed to improve the fairness of the

surface while preserving geometric correctness.

Topology grid

The topology grid is an instantiation of the topology graph onto a 2D triangulation

of wedges. The regions enclosed by the projected curve networks on the mid-plane

are triangulated, as shown in Figure 4.6 (a). The wedges corresponding to the tri-

angulated regions are split into small wedges, each with the shape of a triangular

prism, while the topology graph is refined into a grid structure. Figure 4.6 (d,f) show

portions of the topology grid generated from the default and the modified topology

graph in Figure 4.6 (b,c).

In our implementation, we use the Straight Skeleton triangulation proposed by Aich-

holzer and Aurenhammer [2] and adopted by Barequet et al. [8] for their projection-

based contour interpolation. Straight Skeleton triangulation adds internal vertices

within each region that forms a skeleton of the region suitable for interpolating curve

networks. Since surfaces will only be formed between nodes of different materials on

the grid, we need to triangulate only those wedges that contain two or more nodes of

different materials. For example, wedge A in Figure 4.6(a) is not triangulated since

wedge A contains only a single node in the topology graph of (b,c).

Contouring the topology grid

Just like the topology graph, each edge in the topology grid containing a material

change represents the boundary between two neighboring volumes, and hence gives

rise to the polygons on the final surface. Here we extend the multi-material contouring

method of Ju et al. [45] from a voxel grid onto the topology grid. The special structure

of the topology grid, unlike that of a regular voxel grid, allows us to produce contoured

Page 88: Building a 3D Atlas of the Mouse Brain - Department of Computer

75

B C D

A

B C D

A

(a) (b) (c)

a b

d c

a b

c

(d) (e) (f) (g)

Figure 4.6 : (a) A triangulation of the projected regions on the mid-plane from Figure4.3 (c). (b) The default topology graph. (c) The modified topology graph. (d) Aportion of the topology grid corresponding to the gray triangles in (a) generated fromthe default topology graph in (b). (e) Contoured surface by creating one polygon foreach dotted edge in (d). (f) The topology grid generated from the modified topologygraph in (c). (g) Contoured surface by creating one polygon for each dotted edge in(f).

Page 89: Building a 3D Atlas of the Mouse Brain - Department of Computer

76

surfaces that exactly interpolate the curve networks.

To employ contouring algorithms that are typically designed for a voxelized grid,

however, we need to define two more topology elements on the topology grid: faces

and cells. A face on the topology grid is a closed cycle of edges connecting nodes

between two neighboring triangulated wedges, such as {a, b, c, d} in Figure 4.6 (d)

and {a, b, c} in Figure 4.6 (f). We refer to the portion of the grid shown in Figure

4.6 (d,f) as a column. A column consists of nodes and edges on triangulated wedges

whose projections on the mid-plane surround a single vertex. A cell is a layer on

the column that is made up of grid faces bounded between two closed cycle of edges

around the column. For example, the column in Figure 4.6 (b) contains one cell,

while the column in Figure 4.6 (d) contains two cells.

Assuming the top and bottom sections lie on planes z = 0 and z = 1 with the

mid-plane at z = 0.5, contouring proceeds in two simple steps,

1. Create one vertex for each cell on the topology grid. We assign coordinate

{x, y, (2i− 1)/2n} to the vertex of the ith (i = 1, . . . , n) cell in the triangulated

wedges that project onto triangles on the mid-plane sharing a common vertex

{x, y, 0.5}.

2. Create one polygon for each edge on the topology grid that contains a material

change. The polygon connects the vertices of the grid cells sharing that edge.

Figure 4.6 (e,g) show the contoured surfaces in the grid columns of (d,f). Note

that the contour contains polygons that are dual to grid edges connecting nodes of

different materials. The polygons project onto the triangulation on the mid-plane

and interpolate polygonal boundary segments on the two sections.

Dualing the topology grid is guaranteed to produce a closed contour that partitions

the space between the two sections into distinct volumes whose topology and materials

are specified by the topology graph. The surface is also free of self-intersections:

Page 90: Building a 3D Atlas of the Mouse Brain - Department of Computer

77

the only self-intersection that could take place is between triangles on the contour

projecting to the same triangle on the mid-plane, which is impossible since the vertices

of the grid cells sharing successive edges within the same triangulated wedge are

monotonically increasing (or decreasing).

Our contouring method lifts the triangles from the mid-plane to form polygons on the

3D contour. A similar mechanism was originally used by Barequet [8], whose method

only handles sections containing two materials. More importantly, Barequet’s method

forbids topology modification by maintaining a one-to-one map between a triangle on

the mid-plane and a triangle on the surface. Our method allows complex surface

topology to be generated by modifying the topology graph, which may result in

multiple triangles on the surface projecting onto the same triangle on the mid-plane

(see Figure 4.6 (g)).

Surface smoothing

The vertices of the surface produced by contouring lie on pre-set coordinates with

uniform spacing in the z direction, which may produce a surface with a stair-stepping

appearance. To achieve a smoother-looking surface while maintaining geometric cor-

rectness, we apply a Laplacian smoothing operator:

z∗i = αzi + (1− α)

∑j∈N(i) zj

‖N(i)‖

where zi denotes the z coordinate of vertex i, N(i) denotes the set of edge-adjacent

vertices of vertex i, and α is a scalar value between 0 and 1. In our implementation, we

choose α = 0.5. Note that smoothing the z-coordinate of vertices will not cause self-

intersections on the smoothed surface so long as the monotonicity of the z-coordinate

of vertices with the same projection on the mid-plane is explicitly preserved.

Page 91: Building a 3D Atlas of the Mouse Brain - Department of Computer

78

4.3 Results

Here we present the result of applying our surface generation method to serial sections

of the mouse brain after distortion-correction. The sections have been annotated first

to illustrate 17 major anatomical types 3. Annotation is performed using a Java-based

user-interface that allows the creation of smooth boundary curves using subdivision.

A surface network interpolating the curve networks on each section is then generated

automatically using our method. Finally, the surface topology is verified and corrected

manually, followed by automatic re-generation of the corrected surface network.

4.3.1 Automatic construction

Figure 4.7 (a,c,e) show the side, bottom and cross-section views of the surface network

automatically constructed using our method for the 350 sections of the mouse brain.

The input curve networks contain 200404 vertices, and the output surface contains

1132731 polygons and 590902 vertices. The surface network partitions space into

different anatomical structures, indicated by colors (the legend is shown in Figure 4.7

(g)), whose cross-sections coincide with the annotation on each of the 2D sections.

Note that although individual layers of surfaces between consecutive sections have

been smoothed, the model in Figure 4.7 (a,c,e) looks jagged due to the manual layout

of the curve networks. We applied a preliminary smoothing pass to the entire surface

model using a non-shrinking smoothing operator [Taubin] and the results are shown in

Figure 4.7 (b,d,f). In addition, edges on the surface network shared by three or more

polygons representing the ridges where three or more anatomical structures meet are

smoothed separately to generate a smooth curve network.

3The annotated structures are: Cortex, Cerebellum, Striatum, Basal Forebrain, Amygdala, Hip-

pocampus, Hypothalamus, Thalamus, Olfactory Bulb, Midbrain, Pons, Medulla, Ventral Striatum,

Globus Pallidus, Septum, Fibers, and Ventricles.

Page 92: Building a 3D Atlas of the Mouse Brain - Department of Computer

79

(a) (b)

(c) (d)

(e) (f)

Cortex

Cerebellum

Basal Forebrain

Amygdala

Hippocampus

Hypothalamus

Thalamus

Olfactory Bulb

Striatum

Midbrain

Pons

Medulla

Ventral Striatum

Globus Pallidus

Septum

Fibers

Ventricles

(g)

Figure 4.7 : Surface reconstruction of the mouse brain before (left column) and after(right column) smoothing: side view (a,b), bottom view (c,d), and cross-section view(e,f). The color legend is shown on the far right (g).

Page 93: Building a 3D Atlas of the Mouse Brain - Department of Computer

80

(a) (b)

(c) (d)

Figure 4.8 : Sections from the first 20 cross-sections of the mouse brain (a), thereconstructed surface network (b), the surface bounding Cortex (c) and Olfactorybulb (d).

In Figure 4.8 we take a closer look at the the surface constructed for the first 20

cross-sections of the mouse brain. We observe in particular the close interaction

between two anatomical structures: Cortex (red) and Olfactory bulb (green). While

the Cortex starts at the top surrounded by the Olfactory bulb, the Olfactory bulb

is surrounded by the Cortex at the bottom. The reconstructed surfaces for the two

anatomical structures are shown in Figure 4.8 (c,d).

An important property of our brain model, in contrast with existing polygonal mod-

Page 94: Building a 3D Atlas of the Mouse Brain - Department of Computer

81

els of the mouse brain [5], is that the various anatomical structures share common

boundary surfaces that do not leave gaps or self-intersect. Figure 4.9 (a) shows the

reconstruction of a complex structure in the mouse brain, the Fiber tracks. To view

the adjacency relation between Fiber tracks and its neighboring anatomical struc-

tures, we color the surface of Fiber tracks by its neighboring anatomical structures

in Figure 4.9 (b). Note that the coloring also provides us with information about the

3D anatomy of the brain that is difficult to obtain from 2D images, which will be

helpful for anatomists and brain researchers.

(a) (b)

Figure 4.9 : Surface of the Fiber tracks (a) and colored version using neighboringanatomical structures (b).

4.3.2 Manual adjustments

Due to the small distance between adjacent 2D sections in the mouse brain (i.e.

25µm), the automatically constructed surface network possesses the correct topology

in most places. However, the reconstructed surface fails to produce desired topology

occasionally when there is a large migration of anatomical features between successive

sections. Here we show two examples of repairing incorrect surface topology on the

reconstructed Ventricles through manual modification of the topology graph.

Page 95: Building a 3D Atlas of the Mouse Brain - Department of Computer

82

Figure 4.10 (a) shows a tubular portion of Ventricles that is broken into separate

pieces after automatic reconstruction due to a large displacement of corresponding

regions on successive sections. The surface is colored by the neighboring anatomical

structures. The surface was corrected using the wedge splitting technique discussed

in Section 4.2.3, and the regenerated surface, shown in Figure 4.10 (b), models a

continuous tube.

Figure 4.11 (a) shows a portion of the Ventricles containing a hole due to the migration

of its cross-sections. A continuous reconstruction of Ventricles is produced in Figure

4.11 (b) after modifying the topology graph between the corresponding sections in a

manner similar to the graph modification in Figure 4.3 (f).

Page 96: Building a 3D Atlas of the Mouse Brain - Department of Computer

83

(a) (b)

Figure 4.10 : Broken pieces of Ventricle resulting from automatic construction (a)and after manual adjustment (b).

(a) (b)

Figure 4.11 : A hole in Ventricles resulting from automatic construction (a) and aftermanual adjustment (b).

Page 97: Building a 3D Atlas of the Mouse Brain - Department of Computer

84

4.4 Discussion

The current surface generation method proceeds in two steps: automatic generation

and user-assisted modification. The first step produces a geometrically correct surface

with minimal topological variation; users are totally responsible for checking and

correcting invalid topologies. A major direction for future research is to develop

automatic methods for producing particular types of topologies, which would greatly

reduce the amount of human interaction. A simple example would be to always

connect a particular anatomical region (e.g. the nerve track) on successive sections.

More complicated examples includes tracking feature points (e.g. points where three

or more regions meet) on successive curve networks. These topology requirements

can be enforced by developing rules that would result in a desirable topology graph

in the automatic generation step.

Page 98: Building a 3D Atlas of the Mouse Brain - Department of Computer

Chapter 5

Future Work

The geometric techniques described in previous chapters can generate high-resolution

surface models from distorted tissue sections of a complex anatomical structure. We

are actively expanding our work to create a volumetric atlas for the larger goal of

analyzing gene expressions. I identify the following particular topics for my future

research:

Surface smoothing and simplification: although the preliminary smoothing scheme

discussed in Section 4.3.1 significantly improves the surface quality of the orig-

inal model, artifacts such as waviness can still be observed from the smoothed

surface in Figure 4.7 (d). One possible cause is that the concatenated contours

between successive sections have a fairly non-uniform triangulation, which may

reduce the effectiveness of a linear smoothing scheme. I intend to investigate

other mesh fairing methods that would deliver a smoother surface model. For

simplification, existing methods, such as QSlim [32] and simplification envelops

[25], can either be directly utilized or modified to work on surface networks.

Surface tetrahedralization: After the surface network has been simplified to a

reasonable size, we need to tetrahedralize the surface network to form a vol-

umetric model. In particular, we desire that the tetrahedrons of neighboring

anatomical structures should share common boundary faces and the sizes of the

tetrahedrons should adapt to the shape of the anatomical structure. Although

robust, adaptive tetrahedralization techniques have been developed for closed

meshes, these techniques are not yet available for surface networks that par-

85

Page 99: Building a 3D Atlas of the Mouse Brain - Department of Computer

86

tition space into multiple volumes. Hence the problem of stable and adaptive

tetrahedralization of surface networks will be an exciting area of research.

Atlas utilization: As in Geneatlas.org, our ultimate goal is to construct a 3D database

of gene expression patterns over the mouse brain. The availability of the vol-

umetric atlas will allow us to map expression data from different brains onto

the same atlas for accurate and efficient comparison. Constructing a spatial

database will involve extensive research on 3D registration of the atlas onto

experimental images, efficient comparison of expression data over the atlas, and

developing graphical tools for querying and visualizing gene expression pat-

terns. The availability of an atlas-based database containing expression data of

the twenty thousand genes in the mouse genome will be an invaluable resource

for biology and for medical researchers.

Page 100: Building a 3D Atlas of the Mouse Brain - Department of Computer

Chapter 6

Summary

In this thesis, I have describe how computer graphics techniques are applied in con-

structing a 3D geometric atlas of the mouse brain anatomy. In particular, I have

presented:

1. An automatic and robust method for elastically reconstructing a smooth 3D

brain from tissue cross-sections containing local distortions induced by physi-

cal sectioning. The method employs a novel image warping algorithm based

on dynamic programming and designed for adjacent tissue sections formed by

directional slicing.

2. A robust and flexible method for creating a geometrically and topologically

correct surface network from complex boundary curves on successive planar

sections. The method is best suited for building surface models of a complex

structure containing multiple anatomical regions and allows interactive user-

guided modification.

The combination of these two techniques presents a robust and efficient framework

for generating a high-quality polygonal atlas of an anatomical structure from cross-

sectional images. As part of a larger project, the surface model generated using these

techniques may become the basis for creating a volumetric atlas of the mouse brain

for the study of gene expression study.

87

Page 101: Building a 3D Atlas of the Mouse Brain - Department of Computer

Bibliography

[1] Agazzi, O., Kuo, S., Levin, E., and Pieraccini, R. Connected and de-

graded text recognition using planar hidden markov models. In IEEE Inter-

national Conference on Acoustics, Speech, and Signal Processing (ICASSP-93)

(April 1993), vol. 5, pp. 113–116.

[2] Aichholzer, O., and Aurenhammer, F. Straight skeletons for general

polygonal figures in the plane. In Computing and Combinatorics (1996), pp. 117–

126.

[3] Ali, W., and Cohen, F. Registering coronal histological 2-d sections of a rat

brain with coronal sections of a 3-d brain atlas using geometric curve invariants

and b-spline representation. IEEE Transactions on Medical Imaging 17, 6 (1998),

957–966.

[4] Armstrong, J., Kaiser, K., Mller, A., Fischbach, K., Merchant, N.,

and Strausfeld, N. Flybrain, an on-line atlas and database for the drosophila

nervous system. Neuron 15, 1 (1995), 17–20.

[5] at UCLA, M. A. P. http://www.loni.ucla.edu/map/index.html.

[6] Bajaj, C. L., Coyle, E. J., and Lin, K.-N. Arbitrary topology shape

reconstruction from planar cross sections. Graph. Models Image Process. 58, 6

(1996), 524–543.

[7] Bardinet, E., Ourselin, S., Dormont, D., Malandain, G., Tande, D.,

Parain, K., Ayache, N., and Yelnik, J. Co-registration of histological,

88

Page 102: Building a 3D Atlas of the Mouse Brain - Department of Computer

89

optical and mr data of the human brain. In Proceedings of MICCAI 2002(5th

International Conference, Tokyo, Japan) (2002), pp. 548–555.

[8] Barequet, G., Goodrich, M. T., Levi-Steiner, A., and Steiner, D.

Straight-skeleton based contour interpolation. In SODA ’03: Proceedings of the

fourteenth annual ACM-SIAM symposium on Discrete algorithms (2003), Society

for Industrial and Applied Mathematics, pp. 119–127.

[9] Barequet, G., and Sharir, M. Piecewise-linear interpolation between polyg-

onal slices. In SCG ’94: Proceedings of the tenth annual symposium on Compu-

tational geometry (1994), ACM Press, pp. 93–102.

[10] Berthold, C., Rydmark, M., and Corneliuson, O. Estimation of section-

ing compression and thickness of ultrathin sections through vestopal-w-embedded

cat spinal roots. Journal of Ultrastructure Research 80 (1982), 42–52.

[11] Bloor, M., and Wilson, M. Local control of surfaces generated using partial

differential equations. Computer Graphics 18, 3 (1994), 161–169.

[12] Boissonnat, J.-D. Shape reconstruction from planar cross sections. Comput.

Vision Graph. Image Process. 44, 1 (1988), 1–29.

[13] Bookstein, F. Principal warps: thin-plate splines and the decomposition of

deformations. IEEE Transactions on Pattern Analysis and Machine Intelligence

11 (1989), 567–585.

[14] Boykov, Y., Veksler, O., and Zabih, R. Fast approximate energy mini-

mization via graph cuts. In Proceedings of International Conference on Computer

Vision (September 1999), vol. 1, pp. 377–384.

[15] Boyle, J., Jenkins, J., McKay, I., McPhaden, A., and Lindop, G.

An assessment of the distortion of arteries due to sectioning in endomyocardial

biopsies. Journal of Pathology 181, 2 (1997), 243–246.

Page 103: Building a 3D Atlas of the Mouse Brain - Department of Computer

90

[16] Brey, E., King, T., Johnston, C., McIntire, L., Reece, G., and

Patrick, C. A technique for quantitative three-dimensional analysis of mi-

crovascular structure. Microvascular Research 63 (2002), 279–294.

[17] Cannestra, A., Santori, E., Holmes, C., and Toga, A. A three-

dimensional multimodality brain map of the nemestrina monkey. Brain Research

Bulletin 43, 2 (1997), 141–148.

[18] Carson, J., Thaller, C., and Eichele, G. A transcriptome atlas of the

mouse brain at cellular resolution. Current Opinion in Neurobiology 12, 5 (2002),

562–565.

[19] Catmull, E., and Clark, J. Recursively generated b-spline surfaces on

arbitrary topological meshes. Computer-Aided Design 16, 6 (1978), 350–355.

[20] Chai, J., Miyoshi, T., and Nakamae, E. Contour interpolation and surface

reconstruction of smooth terrain models. In VIS ’98: Proceedings of the confer-

ence on Visualization ’98 (1998), IEEE Computer Society Press, pp. 27–33.

[21] Chen, J. Y., Bouman, C. A., and Allebach, J. P. Multiscale branch

and bound image database search. Storage and Retrieval for Image and Video

Databases V (1997), 133–144.

[22] Chen, X., Ai, Z., Rasmussen, M., Bajcsy, P., Auvil, L., Welge, M.,

Leach, L., Vangveeravong, S., Maniotis, A., and Folberg, R. Three-

dimensional reconstruction of extravascular matrix patterns and blood vessels in

human uveal melanoma tissue: techniques and preliminary findings. Investigative

Ophthalmology and Visual Science 44, 7 (2003), 2834–2840.

[23] Cheng, S.-W., and Dey, T. K. Improved constructions of delaunay based

contour surfaces. In SMA ’99: Proceedings of the fifth ACM symposium on Solid

modeling and applications (1999), ACM Press, pp. 322–323.

Page 104: Building a 3D Atlas of the Mouse Brain - Department of Computer

91

[24] Christiansen, H. N., and Sederberg, T. W. Conversion of complex con-

tour line definitions into polygonal element mosaics. In SIGGRAPH ’78: Pro-

ceedings of the 5th annual conference on Computer graphics and interactive tech-

niques (1978), ACM Press, pp. 187–192.

[25] Cohen, J., Varshney, A., Manocha, D., Turk, G., Weber, H., Agar-

wal, P., Brooks, F., and Wright, W. Simplification envelopes. In SIG-

GRAPH ’96: Proceedings of the 23rd annual conference on Computer graphics

and interactive techniques (1996), ACM Press, pp. 119–128.

[26] Cox, I., Hingorani, S., Rao, S., and Maggs, B. A maximum likelihood

stereo algorithm. Comput. Vis. Image Und. 63, 3 (1996), 542–567.

[27] Csebfalvi, B., Neumann, L., Kanitsar, A., and Groller, E. Smooth

shape-based interpolation using the conjugate gradient method. In Proceedings

of Vision, Modeling, and Visualization (2002), pp. 123–130.

[28] Deverell, M., Bailey, N., and Whimster, W. Tissue distortion in three

dimensional reconstruction of wax or plastic embedded microscopic structures.

Pathol Res Pract 185, 5 (1989), 598–601.

[29] Durr, R., Peterhans, E., and von der Heydt, R. Correction of distorted

image pairs with elastic models. European Journal of Cell Biology 48, 25 (1989),

85–88.

[30] Fuchs, H., Kedem, Z. M., and Uselton, S. P. Optimal surface reconstruc-

tion from planar contours. Commun. ACM 20, 10 (1977), 693–702.

[31] Gabrani, M. Multidimensional spline theory and surface-based alignment of

brains.

[32] Garland, M., and Heckbert, P. S. Surface simplification using quadric

error metrics. In Proceedings of ACM SIGGRAPH 1997 (1997), ACM Press /

Page 105: Building a 3D Atlas of the Mouse Brain - Department of Computer

92

ACM SIGGRAPH, New York. E. Fiume, Ed., Computer Graphics Proceedings,

Annual Conference Series, ACM, pp. 209–216.

[33] Gefen, S., Tretiak, O., and Nissanov, J. Elastic 3-d alignment of rat

brain histological images. IEEE Transactions on Medical Imaging 22, 11 (2003),

1480–1489.

[34] Geiger, B. Three-dimensional modeling of human organs and its application

to diagnosis and surgical planning. Tech. Rep. RR-2105.

[35] Glasbey, C., and Mardia, K. A review of image warping methods. Journal

of Applied Statistics 25 (1998), 155–171.

[36] Guest, E., and Baldock, R. Automatic reconstruction of serial sections

using the finite element method. BioImaging 3 (1996), 154–167.

[37] Guest, E., Berry, E., Baldock, R., Fidrich, M., and Smith, M. Robust

point correspondence applied to two- and three- dimensional image registration.

IEEE Transactions on Pattern Analysis and Machine Intelligence 23 (2001),

1–15.

[38] Hagenr, H. Focus on Scientific Visualization. Springer Verlag, 1993.

[39] Hajnal, J., Hawkes, D., and Hill, D. Medical Image Registration. CRC

Press, 2001.

[40] He, L., Sarrafizadeh, R., and Houk, J. Three-dimensional reconstruction

of the rubrocerebellar premotor network of the turtle. Neuroimage 2, 1 (1995),

21–33.

[41] Herman, G. T., Zheng, J., and Bucholtz, C. A. Shape-based interpola-

tion. IEEE Comput. Graph. Appl. 12, 3 (1992), 69–79.

Page 106: Building a 3D Atlas of the Mouse Brain - Department of Computer

93

[42] Jiang, W., Baker, M., Wu, Q., Bajaj, C., and Chiu, W. Applications of

a bilateral denoising filter in biological electron microscopy. Journal of Structural

Biology 144 (2003), 114–122.

[43] Jolliffe, I. Principal Component Analysis. Springer-Verlag, 1986.

[44] Jones, A., Milthorpe, B., and Howlett, C. Measurement of microtomy-

induced section distortion and its correction for 3-dimensional histological recon-

structions. Cytometry 15, 2 (1994), 95–105.

[45] Ju, T., Losasso, F., Schaefer, S., and Warren, J. Dual contouring of

hermite data. In SIGGRAPH ’02: Proceedings of the 29th annual conference on

Computer graphics and interactive techniques (2002), ACM Press, pp. 339–346.

[46] Ju, T., Warren, J., Eichele, G., Thaller, C., Chiu, W., and Car-

son, J. A geometric database for gene expression data. In Proceedings of the

Eurographics/ACM SIGGRAPH symposium on Geometry processing (2003), Eu-

rographics Association, pp. 166–176.

[47] Karen, P., Jirkovska, M., Tomori, Z., Demjenova, E., Janacek, J.,

and Kubinova, L. Three-dimensional computer reconstruction of large tissue

volumes based on composing series of high-resolution confocal images by gluemrc

and linkmrc software. Microscopy Research and Technique 62, 5 (2003), 415–422.

[48] Keppel, E. Approximating complex surfaces by triangulation of contour lines.

IBM JOURNAL OF RESEARCH AND DEVELOPMENT XIX (1975), 2–11.

[49] Keysers, D., and Unger, W. Elastic image matching is np-complete. Pattern

Recogn. Lett. 24, 1-3 (2003), 445–453.

[50] Kim, B., Boes, J., Frey, K., and Meyer, C. Mutual information for

automated unwarping of rat brain autoradiographs. Neuroimage 5 (1997), 31–

40.

Page 107: Building a 3D Atlas of the Mouse Brain - Department of Computer

94

[51] Klein, R., Schilling, A., and Straßer, W. Reconstruction and simplifi-

cation of surfaces from contours. Graphical models 62, 6 (2000), 429–443.

[52] Kovar, L., and Gleicher, M. Flexible automatic motion blending with

registration curves. In Proceedings of the 2003 ACM SIGGRAPH/Eurographics

Symposium on Computer Animation (2003), Eurographics Association, pp. 214–

224.

[53] Lamers, W., Laan, A., Huijsmans, D., Smith, J., Los, J., and Strac-

kee, J. Deformation-corrected computer-aided 3-d reconstruction of immuno-

histochemically stained sections of embryonic organs. European Journal of Cell

Biology 48 (1989), 103–106.

[54] Lester, H., and Arridge, S. R. A survey of hierarchical non-linear medical

image registration. Pattern Recognition 32, 1 (1999), 129–149.

[55] Levin, E., and Pieraccini, R. Dynamic planar warping for optical character

recognition. In Proc ICASSP (1992), vol. III, pp. 149–152.

[56] MacKenzie-Graham, A., Lee, E., Dinov, I., Bota, M., Shattuck, D.,

Ruffins, S., Yuan, H., Konstantinidis, F., Pitiot, A., Ding, Y., Hu,

G., Jacobs, R., and Toga, A. A multimodal, multidimensional atlas of the

c57bl/6j mouse brain. Journal of Anatomy 204, 2 (2004), 93–102.

[57] Maes, F., Collignon, A., Vandermeulen, D., Marchal, G., and

Suetens, P. Multimodality image registration by maximization of mutual in-

formation. IEEE Transactions on Medical Imaging 16 (1997), 187–198.

[58] Maintz, J. A., and Viergever, M. A. A survey of medical image registra-

tion. Medical Image Analysis 2, 1 (1998), 1–37.

[59] Marola, G. On the detection of the axes of symmetry of symmetric and almost

symmetric planar images. IEEE Trans. Pattern Anal. Mach. Intell. 11, 1 (1989),

Page 108: Building a 3D Atlas of the Mouse Brain - Department of Computer

95

104–108.

[60] Mega, M., Chen, S., Thompson, P., Woods, R., Karaca, T., Ti-

wari, A., Vinters, H., Small, G., and Toga, A. Mapping histology

to metabolism: Coregistration of stained whole-brain sections to premortem pet

in alzheimers disease. Neuroimage 5 (1997), 147–153.

[61] Meyer, C., Boes, J., Kim, B., Bland, P., Zasadny, K., Kison, P. V.,

Koral, K., Frey, K., and Wahl, R. Demonstration of accuracy and clin-

ical versatility of mutual information for automatic multimodality image fusion

using affine and thin plate spline warped geometric deformations. Medical Image

Analysis 3 (1997), 195–206.

[62] Meyers, D., Skinner, S., and Sloan, K. Surfaces from contours. ACM

Trans. Graph. 11, 3 (1992), 228–258.

[63] Modersitzki, J. Numerical Methods for Image Registration. Oxford University

Press, 2004.

[64] Montgomery, K., and Ross, M. Improvements in semiautomated serial-

section reconstruction and visualization of neural tissue from tem images. Three-

Dimensional Microscopy (Proc SPIE) 2184 (1994), 264–267.

[65] Moody, D., and Lozanoff, S. Surfdriver: A practical computer program for

generating three-dimensional models of anatomical structures.

[66] Oliva, J.-M., Perrin, M., and Coquillart, S. 3d reconstruction of com-

plex polyhedral shapes from contours using a simplified generalized voronoi dia-

gram. Computer Graphics Forum 15, 3 (1996), 397–408.

[67] Ourselin, S., Bardinet, E., Dormont, D., Malandain, G., Roche, A.,

Ayache, N., Tande, D., Parain, K., and Yelnik, J. Fusion of histological

sections and mr images: towards the construction of an atlas of the human basal

Page 109: Building a 3D Atlas of the Mouse Brain - Department of Computer

96

ganglia. In Proceedings of MICCAI 2001(4th International Conference, Utrecht,

The Netherlands) (2001), pp. 743–751.

[68] Paxinos, G., and Franklin, K. The Mouse Brain in Stereotaxic Coordinates,

Second Edition. Academic Press, 2000.

[69] Pitiot, A., Malandain, G., Bardinet, E., and Thompson, P. Piecewise

affine registration of biological images. In International Workshop on Biomedical

Image Registration (2003).

[70] Rangarajan, A., Chui, H., and Bookstein, F. L. The softassign pro-

crustes matching algorithm. In Information Proc. Med. Imaging (1997), pp. 29–

42.

[71] Ronee, M., Uchida, S., and Sakoe, H. Handwritten character recogni-

tion using piecewise linear two-dimensional warping. In Proceedings of the Sixth

International Conference on Document Analysis and Recognition (ICDAR ’01)

(September 2001), pp. 39–43.

[72] Rosen, G., Williams, A., Capra, J., Connolly, M., Cruz, B., Lu,

L., Airey, D., Kulkarni, K., and Williams, R. The mouse brain library

@ www.mbl.org. In International Mouse Genome Conference (2000), vol. 14,

p. 166.

[73] Roy, S. Stereo without epipolar lines: A maximum-flow formulation. Interna-

tional Journal of Computer Vision 34, 2 (1999), 147–161.

[74] Rueckert, D., Sonoda, L., Hayes, C., Hill, D., Leach, M., and

Hawkes, D. Non-rigid registration using free-form deformations: application to

breast mr images. IEEE Transactions on Medical Imaging 18, 1 (1999), 712–721.

[75] Rydmark, M., Jansson, T., Berthold, C., and Gustavsson, T.

Computer-assisted realignment of light micrograph images from consecutive sec-

Page 110: Building a 3D Atlas of the Mouse Brain - Department of Computer

97

tion series of cat cerebral cortex. Journal of Microscopy 165 (1992), 29–47.

[76] Sakoe, H., and Chiba, S. Dynamic programming algorithm optimization for

spoken word recognition. IEEE Trans. Acoustics, Speech and Signal Processing

26, 1 (1978), 43–49.

[77] Schaefer, S., Hakenberg, J., and Warren, J. Smooth subdivision of

tetrahedral meshes. In Proceedings of the Eurographics/ACM SIGGRAPH sym-

posium on Geometry processing (2004), Eurographics Association, pp. 151–158.

[78] Schormann, T., Dabringhaus, A., and Zilles, K. Statistics of defor-

mations in histology and application to improved alignment with mri. IEEE

Transactions on Medical Imaging 14, 1 (1995), 25–35.

[79] Sidman, R. High resolution brain atlas of c57bl/6j in 2d and 3d.

[80] Sloan, K. R., and Painter, J. From contours to surfaces: testbed and initial

results. In CHI ’87: Proceedings of the SIGCHI/GI conference on Human factors

in computing systems and graphics interface (1987), ACM Press, pp. 115–120.

[81] Streicher, J., Donat, M., Strauss, B., Sprle, R., Schughart, K., and

Mller, G. Computer-based three-dimensional visualization of developmental

gene expression. Nature Genetics 25, 2 (2000), 147–152.

[82] Studholme, C., Constable, R., and Duncan, J. Accurate alignment of

functional epi data toanatomical mri using a physics-based distortion model.

IEEE Transactions on Medical Imaging 19 (2000), 1115–1127.

[83] Thevenaz, P., Ruttimann, U., and Unser, M. A pyramid approach to

subpixel registration based on intensity. IEEE Transactions on Medical Imaging

7, 1 (1998), 27–41.

Page 111: Building a 3D Atlas of the Mouse Brain - Department of Computer

98

[84] Thompson, P., and Toga, A. A surface-based technique for warping three

dimensional images of the brain. IEEE Transactions on Medical Imaging 15, 4

(1996), 402–417.

[85] Timsari, B., Tocco, G., Bouteiller, J., Baudry, M., and Leahy, R.

Accurate registration of autoradiographic images of rat brain using a 3-d atlas. In

International Conference on Imaging Science, Systems, and Technology (1999).

[86] Toga, A. Brain Warping. Academic Press, 1999.

[87] Toga, A., and Arnicar-Sulze, T. Digital image reconstruction for the study

of brain structure and function. Journal of Neuroscience Methods 20, 1 (1987),

7–21.

[88] Tomasi, C., and Manduchi, R. Bilateral filtering for gray and color images.

In Proceedings of the IEEE International Conference on Computer Vision (1998),

pp. 59–66.

[89] Turk, G., and O’Brien, J. F. Shape transformation using variational im-

plicit functions. In SIGGRAPH ’99: Proceedings of the 26th annual conference

on Computer graphics and interactive techniques (1999), ACM Press/Addison-

Wesley Publishing Co., pp. 335–342.

[90] Uchida, S., and Sakoe, H. A monotonic and continuous two-dimensional

warping based on dynamic programming. In Proceedings of the 15th International

Conference on Pattern Recognition (1998), vol. 1, pp. 521–524.

[91] Warren, J., and Weimer, H. Subdivision Methods for Geometric Design.

Morgan-Kaufmann, 2002.

[92] Weinstein, D. Scanline surfacing: Building separating surfaces from planar

contours. In VISUALIZATION ’00: Proceedings of the 11th IEEE Visualization

2000 Conference (VIS 2000) (2000), IEEE Computer Society.

Page 112: Building a 3D Atlas of the Mouse Brain - Department of Computer

99

[93] Wirtz, S., Fischer, B., Modersitzki, J., and Schmitt, O. Super-fast

elastic registration of histologic images of a whole rat brain for three-dimensional

reconstruction. In Proceedings of the SPIE Medical Imaging 2004 (February

2004), vol. 5730.

[94] Woods, R., Grafton, S., Holmes, C., Cherry, S., and Mazziotta, J.

Automated image registration: i. general methods and intrasubject, intramodal-

ity validation. Journal of Computer Assisted Tomography 22 (1998), 139–152.

[95] Woods, R., Grafton, S., Watson, J., Sicotte, N., and Mazziotta, J.

Automated image registration: ii. intersubject validation of linear and nonlinear

models. Journal of Computer Assisted Tomography 22 (1998), 153–165.

[96] Woods, R., Mazziotta, J., and Cherry, S. Mri-pet registration with

automated algorithm. Journal of Computer Assisted Tomography 17, 4 (1993),

536–546.

[97] Yoo, T., Ackerman, M., Lorensen, W., Schroeder, W., Chalana,

V., Aylward, S., Metaxes, D., and Whitaker, R. Engineering and al-

gorithm design for an image processing api: A technical report on itk - the

insight toolkit. In Proc. of Medicine Meets Virtual Reality (2002), IOS Press

Amsterdam, pp. 586–592.