11.03.2014 Views

Segmentation of Stochastic Images using ... - Jacobs University

Segmentation of Stochastic Images using ... - Jacobs University

Segmentation of Stochastic Images using ... - Jacobs University

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> <strong>using</strong><br />

<strong>Stochastic</strong> Partial Differential Equations<br />

by<br />

Torben Pätz<br />

A thesis submitted in partial fulfillment<br />

<strong>of</strong> the requirements for the degree <strong>of</strong><br />

Doctor <strong>of</strong> Philosophy in Mathematics<br />

Thesis Supervisor:<br />

Second Referee:<br />

External Referee:<br />

Thesis committee:<br />

Pr<strong>of</strong>. Dr. Tobias Preusser<br />

<strong>Jacobs</strong> <strong>University</strong> Bremen<br />

Pr<strong>of</strong>. Dr. Marcel Oliver<br />

<strong>Jacobs</strong> <strong>University</strong> Bremen<br />

Pr<strong>of</strong>. Dr. Joachim Weickert<br />

Saarland <strong>University</strong>, Saarbrücken<br />

Date <strong>of</strong> Defense: January 13, 2012<br />

School <strong>of</strong> Engineering and Science<br />

<strong>Jacobs</strong> <strong>University</strong> Bremen


Acknowledgement<br />

I would like to thank my advisor Pr<strong>of</strong>. Dr. Tobias Preusser for the guidance during my PhD studies<br />

and for giving me the opportunity to work in the rapidly growing field <strong>of</strong> stochastic modeling. Furthermore,<br />

Pr<strong>of</strong>. Preusser’s connection to Fraunh<strong>of</strong>er MEVIS enabled me to work in an inspiring<br />

environment <strong>of</strong> people working on image processing problems.<br />

Furthermore, I would like to thank Pr<strong>of</strong>. Dr. Marcel Oliver and Pr<strong>of</strong>. Dr. Joachim Weickert for<br />

being members <strong>of</strong> my dissertation committee.<br />

Special thanks go to the <strong>Jacobs</strong> <strong>University</strong> Bremen for the financial support during my PhD studies.<br />

Without the tuition waiver for the complete time <strong>of</strong> my PhD studies and the scholarship during the<br />

first two years, I would not be able to finish my studies successfully. Furthermore, special thanks go<br />

to Fraunh<strong>of</strong>er MEVIS for the student assistance contract during the first two years <strong>of</strong> my PhD studies<br />

and the possibility to use the infrastructure <strong>of</strong> the institute for my studies.<br />

I am grateful to all my colleagues at the "Modeling and Simulation" group at Fraunh<strong>of</strong>er MEVIS.<br />

Especially, I would like to thank Sabrina Haase, Hanne Tiesler, and Dr. Ole Schwen for reading and<br />

commenting on a draft <strong>of</strong> this thesis.<br />

I am also thankful to the QuocMesh collective from the work group <strong>of</strong> Pr<strong>of</strong>. Dr. Martin Rumpf at<br />

the Institute <strong>of</strong> Numerical Simulation <strong>of</strong> the Rheinische Friedrich-Wilhelms-Universität Bonn. All<br />

implementations <strong>of</strong> the methods from this thesis are done in QuocMesh and without this excellent<br />

finite element library this would be much more work. Especially, I would like to thank Dr. Ole<br />

Schwen for answering all my questions concerning QuocMesh.<br />

I am also grateful to Pr<strong>of</strong>. Dr. Robert M. Kirby from the <strong>University</strong> <strong>of</strong> Utah in Salt Lake City,<br />

USA, for the possibility to stay two weeks at the Scientific Computing and Imaging Institute in Salt<br />

Lake City. Furthermore, I would like to thank Pr<strong>of</strong>. Kirby and Pr<strong>of</strong>. Joshi for the fruitful discussions<br />

during my stay in Salt Lake City.<br />

iii


Abstract<br />

The task <strong>of</strong> segmentation, the separation <strong>of</strong> an image into foreground and background, is typically<br />

performed on noisy images, and it is a great challenge to get satisfactory segmentation results. The<br />

noise in the images depends on the acquisition modality (e.g. digital camera, MR, CT, ultrasound),<br />

the acquisition parameters (acquisition time, sound frequency, magnetic field strength) and extrinsic<br />

parameters (illumination, reflection). The acquisition step itself is a kind <strong>of</strong> physical measurement<br />

(photon density, time-<strong>of</strong>-flight <strong>of</strong> the waves, spin, absorption) and – according to good scientific<br />

practice – has to be equipped with information about the measurement error. This allows to estimate<br />

the reliability <strong>of</strong> the measurement. The last step <strong>of</strong> quantifying the measurement error is typically<br />

omitted in image processing. Neglecting the error leads to a loss <strong>of</strong> information about the influence<br />

<strong>of</strong> the input error to the result <strong>of</strong> the image processing steps. This is important in medical application,<br />

where radiologists generate decisions about the patients’ treatment based on the information<br />

extracted from the images. For example, the further treatment for cancer patients is based on the volume<br />

<strong>of</strong> the lesions segmented in the noisy images. It is important to equip the extracted information<br />

with a reliability estimate or, and this is the aim <strong>of</strong> the presented work, to be able to compute the<br />

probability density function for the extracted information depending on the estimation or modeling<br />

<strong>of</strong> the input noise.<br />

A possibility to model the image noise is to perceive a pixel inside the image as a random variable.<br />

These images are called stochastic images. Doing this, the segmentation acts on images containing<br />

random variables as pixels. This is contrary to the classical image processing task, where every<br />

pixel has a deterministic value. Applying segmentation methods based on partial differential equations<br />

(PDEs) on these stochastic images leads to stochastic PDEs (SPDEs), i.e. PDEs with stochastic<br />

coefficients or right hand side. The discretization <strong>of</strong> SPDEs is an active and fast proceeding research<br />

field and new methods for an efficient and elegant discretization are available in the literature.<br />

In this thesis, the focus is on intrusive methods for the discretization <strong>of</strong> SPDEs, because classical<br />

sampling strategies like Monte Carlo simulation or stochastic collocation are time-consuming. The<br />

approximation <strong>of</strong> random variables uses the Wiener-Askey (or generalized) polynomial chaos and<br />

the discretization <strong>of</strong> the SPDEs uses the recently developed generalized spectral decomposition and<br />

finite difference schemes for random variables.<br />

This thesis investigates the random walker segmentation, Ambrosio-Tortorelli segmentation, a regularization<br />

<strong>of</strong> the Mumford-Shah functional, and the level set based segmentation methods geodesic<br />

active contours, gradient-based segmentation, and Chan-Vese segmentation for stochastic extensions.<br />

Furthermore, a sensitivity analysis <strong>of</strong> the classical segmentation approaches uses the stochastic<br />

framework by making segmentation parameters random variables and investigating the influence <strong>of</strong><br />

the stochastic parameters on the segmentation result.<br />

The result <strong>of</strong> the presented work is a framework carrying the probability distribution <strong>of</strong> the stochastic<br />

gray values, i.e. the random variables, through all steps <strong>of</strong> the segmentation pipeline. This yields<br />

segmentation results containing, for each pixel, a probability <strong>of</strong> belonging to the object or to the<br />

background. Furthermore, this stochastic segmentation identifies regions where the image noise has<br />

an important impact on the segmentation result and regions, which are robust in the presence <strong>of</strong><br />

noise. In addition, the visualization <strong>of</strong> the resulting stochastic images/contours is investigated.<br />

v


Contents<br />

Acknowledgement<br />

Abstract<br />

Notation<br />

iii<br />

v<br />

ix<br />

1 Introduction 1<br />

2 Image <strong>Segmentation</strong> and Limitations 7<br />

2.1 Mathematical <strong>Images</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

2.2 Random Walker <strong>Segmentation</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

2.3 Mumford-Shah and Ambrosio-Tortorelli <strong>Segmentation</strong> . . . . . . . . . . . . . . . . 12<br />

2.4 Level Sets for Image <strong>Segmentation</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

2.5 Why is Classical Image Processing not Enough? . . . . . . . . . . . . . . . . . . . . 21<br />

2.6 Work Related to the <strong>Stochastic</strong> Framework . . . . . . . . . . . . . . . . . . . . . . . 23<br />

3 SPDEs and Polynomial Chaos Expansions 25<br />

3.1 Basics from Probability Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

3.2 <strong>Stochastic</strong> Partial Differential Equations . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

3.3 Polynomial Chaos Expansions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

3.4 Relation to Interval Arithmetic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35<br />

4 Discretization <strong>of</strong> SPDEs 37<br />

4.1 Sampling Based Discretization <strong>of</strong> SPDEs . . . . . . . . . . . . . . . . . . . . . . . 37<br />

4.2 <strong>Stochastic</strong> Finite Difference Methods . . . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

4.3 <strong>Stochastic</strong> Finite Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

4.4 Generalized Spectral Decomposition . . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

4.5 Adaptive Grids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45<br />

5 <strong>Stochastic</strong> <strong>Images</strong> 47<br />

5.1 Polynomial Chaos for <strong>Stochastic</strong> <strong>Images</strong> . . . . . . . . . . . . . . . . . . . . . . . . 47<br />

5.2 Generation <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> from Samples . . . . . . . . . . . . . . . . . . . . 48<br />

5.3 Comparison <strong>of</strong> the Space from [130] and the Space Used in this Thesis . . . . . . . . 52<br />

5.4 Visualization <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . 53<br />

6 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using Elliptic SPDEs 57<br />

6.1 Random Walker <strong>Segmentation</strong> on <strong>Stochastic</strong> <strong>Images</strong> . . . . . . . . . . . . . . . . . 57<br />

6.2 Ambrosio-Tortorelli <strong>Segmentation</strong> on <strong>Stochastic</strong> <strong>Images</strong> . . . . . . . . . . . . . . . 67<br />

7 <strong>Stochastic</strong> Level Sets 79<br />

7.1 Derivation <strong>of</strong> a <strong>Stochastic</strong> Level Set Equation . . . . . . . . . . . . . . . . . . . . . 79<br />

7.2 Discretization <strong>of</strong> the <strong>Stochastic</strong> Level Set Equation . . . . . . . . . . . . . . . . . . 83<br />

7.3 Reinitialization <strong>of</strong> <strong>Stochastic</strong> Level Sets . . . . . . . . . . . . . . . . . . . . . . . . 84<br />

7.4 Numerical Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

vii


7.5 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using <strong>Stochastic</strong> Level Sets . . . . . . . . . . . 88<br />

8 <strong>Segmentation</strong> <strong>of</strong> Classical <strong>Images</strong> Using <strong>Stochastic</strong> Parameters 97<br />

8.1 Random Walker <strong>Segmentation</strong> with <strong>Stochastic</strong> Parameter . . . . . . . . . . . . . . . 98<br />

8.2 Ambrosio-Tortorelli <strong>Segmentation</strong> with <strong>Stochastic</strong> Parameters . . . . . . . . . . . . 101<br />

8.3 Gradient-Based <strong>Segmentation</strong> with <strong>Stochastic</strong> Parameter . . . . . . . . . . . . . . . 104<br />

8.4 Geodesic Active Contours with <strong>Stochastic</strong> Parameters . . . . . . . . . . . . . . . . . 105<br />

9 Summary, Discussion, and Conclusion 107<br />

9.1 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107<br />

9.2 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107<br />

9.3 Outlook and Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108<br />

List <strong>of</strong> Figures 111<br />

List <strong>of</strong> Tables 117<br />

A Publications Written During the Course <strong>of</strong> the Thesis 119<br />

A.1 Publications Related to <strong>Stochastic</strong> <strong>Images</strong> . . . . . . . . . . . . . . . . . . . . . . . 119<br />

A.2 Publications Related to Radi<strong>of</strong>requency Ablation . . . . . . . . . . . . . . . . . . . 119<br />

Bibliography 121<br />

viii


Notation<br />

u image function ⊗ tensor product<br />

D image domain ∂D boundary <strong>of</strong> the domain D<br />

R real numbers S ρ,k,m Kondratiev space<br />

P i finite element hat function H univariate polynomial<br />

I node set <strong>of</strong> a finite element grid Ψ multivariate polynomial<br />

D c<br />

(<br />

Cantor measure<br />

a<br />

b)<br />

binomial coefficient<br />

BV functions with bounded variation ξ basic random variable<br />

SBV special BV space (D c = 0) δ i j Kronecker delta<br />

GSBV<br />

generalized SBV space<br />

∂ f<br />

∂x<br />

partial derivative<br />

sign sign function ∂ t partial temporal derivative<br />

H d d-dim. Hausdorff measure τ time step size<br />

K<br />

edge set (discontinuities) <strong>of</strong> an<br />

image<br />

h<br />

spatial grid spacing<br />

φ phase field or level set V finite element space<br />

H 1 (D) Sobolev space H 1 over D S stochastic space, ⊂ L 2 (Ω)<br />

| · | absolute value <strong>of</strong> real numbers ‖ · ‖ x x-norm<br />

∆ Laplace operator F cumulative distribution function<br />

tanh hyperbolic tangent N normal vector<br />

κ<br />

curvature <strong>of</strong> level sets or phase<br />

fields<br />

T<br />

tangential vector (<strong>of</strong> level sets)<br />

∗ convolution operator (·) ′ derivative <strong>of</strong> univariate function<br />

E expected value W width <strong>of</strong> the tangential pr<strong>of</strong>ile <strong>of</strong><br />

a phase field<br />

Ω probability (event) space L p Lebesgue spaces<br />

H m<br />

Sobolev spaces<br />

ix


Chapter 1<br />

Introduction<br />

The development <strong>of</strong> mathematical methods for image processing became a rapidly growing research<br />

field during the last decades. The fast progress in the speed <strong>of</strong> widely available computer systems<br />

allowed the numerical implementation <strong>of</strong> complex models. A specialty is the development <strong>of</strong> segmentation<br />

algorithms based on partial differential equations (PDEs). The aim <strong>of</strong> a segmentation<br />

algorithm is the decomposition <strong>of</strong> an image into the object and the background. Typically, detecting<br />

edges inside an image or meeting a homogenization criterion for the object and the background lead<br />

to a segmentation. Widely used segmentation approaches are the random walker segmentation [59],<br />

the Mumford-Shah segmentation [107] and the related Ambrosio-Tortorelli regularization [14], and<br />

active contour methods based on level set formulations [30,31,82,138]. Besided these segmentation<br />

methods, which will be investigated in this thesis, there are other segmentation methods like region<br />

growing [127], watersheds [136], snakes [76], and graph cuts [25].<br />

Many applications use segmentation methods, e.g. quality control, machine vision, and medical<br />

image processing. For example, the further treatment for cancer patients bases on the segmented<br />

volume <strong>of</strong> the lesions from images. Fig. 1.1 shows a computed tomography (CT) image <strong>of</strong> a lung<br />

lesion and the corresponding segmentation mask.<br />

Typically, the segmentation methods act on noisy images (see Figs. 1.1 and 1.2). The image noise<br />

depends on the image acquisition modality (e.g. digital camera, MR, CT, ultrasound), the acquisition<br />

parameters (acquisition time, sound frequency, magnetic field strength), and extrinsic parameters<br />

(illumination, reflection). The acquisition itself is a physical measurement (photon density, time-<strong>of</strong>flight<br />

<strong>of</strong> the waves, spin, absorption), and it is good scientific practice to equip this measurement<br />

with information about the measurement error. This last step <strong>of</strong> quantifying the measurement error<br />

is typically omitted in image processing, leading to a loss <strong>of</strong> information about the influence <strong>of</strong> the<br />

input error to the result <strong>of</strong> the image processing steps. Furthermore, image processing operators,<br />

especially segmentation operators, do not have the ability to propagate this error information to the<br />

result. This is e.g. important in medical application, where physicians decide about the patients’<br />

treatment based on the information extracted from the images.<br />

Figure 1.1: Left: CT image <strong>of</strong> a lung lesion (the small roundish structure in the middle <strong>of</strong> the image).<br />

Right: The segmentation mask computed via region growing [127].<br />

1


Chapter 1 Introduction<br />

Figure 1.2: Noisy images from an ultrasound device (left) showing a structure in the forearm and a<br />

computed tomography (right) <strong>of</strong> a vertebra in a human spine.<br />

The aim <strong>of</strong> this thesis is to provide a representation for images containing error information and<br />

to provide a framework for the error propagation <strong>of</strong> image processing operators.<br />

The representation <strong>of</strong> images containing error information is based on a concept presented by<br />

Preusser et al. [130]. This thesis identifies pixels by random variables. We call images containing<br />

random variables as pixels stochastic images. The discretization <strong>of</strong> stochastic images uses the<br />

generalized polynomial chaos developed by Xiu and Karniadakis [160] to approximate the random<br />

variables at the pixels in a numerically meaningful way. This way <strong>of</strong> image representation is possible<br />

when information about the distribution <strong>of</strong> the gray value for a pixel is available. Repeated acquisitions<br />

<strong>of</strong> the same scene with the same imaging device or the usage <strong>of</strong> noise models can generate<br />

this information. The repeated acquisitions are only possible in rare situations, where a still scene is<br />

available and the repeated acquisition is ethically maintainable. Typically, the generation <strong>of</strong> medical<br />

images violates these conditions, because the human under investigation is alive and acquisition devices<br />

like computed tomography use high-energy radiation. Thus, for medical applications there is<br />

only a limited area for the application <strong>of</strong> these methods. For other areas like quality control, it is easy<br />

to generate samples <strong>of</strong> the same typically still scene. A possibility to overcome the need for multiple<br />

samples is the application <strong>of</strong> noise models in combination with a single image. However, the<br />

available image sample has to be as close as possible to the expected value <strong>of</strong> multiple acquisitions<br />

to get meaningful results. This is hard to achieve. Nevertheless, we present a possibility to generate<br />

a stochastic image from a sinogram <strong>of</strong> a computed tomography.<br />

Replacing the classical images in the PDE based operators by their stochastic counterparts<br />

achieves error propagation for image processing operators, but leads to stochastic partial differential<br />

equations (SPDEs). The numerical solution <strong>of</strong> SPDEs is a rapidly growing field, because these equations<br />

arise in the modeling <strong>of</strong> physical processes with uncertain parameters like heat propagation [24]<br />

or fluid dynamics [84, 93, 109]. Uncertain parameters are e.g. the thermal conductivity or the speed,<br />

because it is impossible to estimate these parameters exactly, but sometimes information about the<br />

probability density function (PDF) is available for these parameters. In the classical modeling with<br />

PDEs, one uses the expected value <strong>of</strong> the parameters for the calculation, yielding results that seem<br />

accurate, but lose the information about the distribution <strong>of</strong> the input parameters. It is a great advantage<br />

to have this information also in the output <strong>of</strong> such a calculation. The simplest method to<br />

get information about this distribution is to perform a Monte Carlo simulation [101], i.e. to perform<br />

deterministic calculations with a parameter sampled from the known input distribution. This is timeconsuming,<br />

due to the high number <strong>of</strong> runs needed to achieve a sufficient precision. To overcome this<br />

problem, methods have been developed ranging from stochastic collocation [158], a technique to use<br />

2


Figure 1.3: This thesis combines findings from image processing with findings about SPDEs to yield<br />

segmentation algorithms acting on stochastic images.<br />

special sampling points in the random space, to the stochastic finite element method (SFEM) [54],<br />

a discretization with finite elements in the random and spatial dimensions. Furthermore, the generalized<br />

spectral decomposition (GSD) [113] allows breaking down the huge equation system <strong>of</strong> the<br />

SFEM into a series <strong>of</strong> smaller systems.<br />

The main goal <strong>of</strong> this thesis is to investigate, whether it is possible to combine the results from the<br />

SPDE context with the results from image processing (see Fig. 1.3), especially for the task <strong>of</strong> image<br />

segmentation. In other words: Which methods from image processing benefit from a stochastic<br />

modeling <strong>of</strong> the input or model parameters, and how can we interpret the results from stochastic<br />

modeling? The combination <strong>of</strong> stochastic images and SPDE, in the way presented in this thesis, was<br />

never done before in the literature, although this approach yields new insights for the PDE based<br />

segmentation <strong>of</strong> images:<br />

• It determines regions where the segmentation is reliable also in the presence <strong>of</strong> image noise<br />

and regions where the image noise has a great impact on the result <strong>of</strong> a segmentation.<br />

• It quickly investigates the influence <strong>of</strong> parameters on the segmentation results, when <strong>using</strong><br />

SFEM or similar techniques for the computation.<br />

The development <strong>of</strong> SPDE methods for image segmentation is based on existing PDE segmentation<br />

methods. However, there are various methods proposed in the literature (see [144] for a review),<br />

and we limit the analysis for a stochastic extension to a few <strong>of</strong> them. Namely, these are random<br />

walker segmentation [59], Mumford-Shah segmentation [107] with the related Ambrosio-Tortorelli<br />

approximation [14], gradient-based segmentation [29], geodesic active contours [30, 82], and Chan-<br />

Vese segmentation [31]. The latter three are based on a level set formulation.<br />

Random walker segmentation [59] in contrast to other segmentation methods based on PDEs is<br />

a supervised segmentation method, meaning that the user influences the segmentation result by interactive<br />

input. For random walker segmentation, the user input consists <strong>of</strong> defining seed regions,<br />

i.e. regions where the user specifies whether they belong to the object or not. The idea <strong>of</strong> the random<br />

walker segmentation is to start random walks from the unseeded pixels and to give every pixel<br />

a probability to belong to the object dependent on the fraction <strong>of</strong> random walks reaching the seed<br />

region <strong>of</strong> the object. The direction the random walker chooses is dependent on the image gradient<br />

between neighboring pixels, i.e. the probability to walk from one pixel to another is higher, when the<br />

image gradient between the pixels is low. An implementation <strong>of</strong> the random walker algorithm uses a<br />

different strategy, because it is unnecessary to compute random walks for every pixel to compute the<br />

probabilities. Doyle and Snell [45] showed the equivalence to a Dirichlet problem. This reduces the<br />

complexity to the solution <strong>of</strong> an elliptic PDE with an unknown for every unseeded pixel.<br />

3


Chapter 1 Introduction<br />

Ambrosio-Tortorelli segmentation [14] is a regularization <strong>of</strong> the segmentation approach proposed<br />

by Mumford and Shah [107]. The idea is to compute a smooth representation <strong>of</strong> the image and<br />

the corresponding edges, respectively a phase field approximation <strong>of</strong> the edges. For the Ambrosio-<br />

Tortorelli model, the author developed a stochastic extension [1, 3], allowing to propagate information<br />

about the measurement error to the result, the smooth image and the phase field.<br />

Level set based segmentation is based on the evolution <strong>of</strong> a contour, represented by a level set<br />

function, i.e. the contour is given as the zero level set <strong>of</strong> a higher-dimensional function. A speed<br />

function controls the evolution <strong>of</strong> the contour. A typical choice for the speed function is to make it<br />

dependent on the image gradient [29, 96]. Caselles et al. [30] and simultaneously Kichenassamy et<br />

al. [82] developed improvements by adding a term that forces the contour to stay at edges. Furthermore,<br />

Chan and Vese [31] developed a segmentation method that is able to segment objects without<br />

sharp edges to the background. Instead <strong>of</strong> <strong>using</strong> gradient information, they proposed a functional<br />

that segments homogeneous regions in the image.<br />

Besides the development <strong>of</strong> stochastic segmentation algorithms, the investigation <strong>of</strong> pre- and postprocessing<br />

steps is essential to end up with a complete framework for error propagation in image processing.<br />

For example, it is necessary to develop a technique to acquire stochastic images, i.e. images<br />

whose pixels are random variables when image samples are available. This step benefits from techniques<br />

available in the literature [41, 130, 141] or from the modeling <strong>of</strong> the noise distribution. In<br />

addition, this thesis investigates the visualization <strong>of</strong> the stochastic segmentation results.<br />

Furthermore, it is possible to change the perspective and use the segmentation methods developed<br />

for stochastic images for a sensitivity analysis <strong>of</strong> the segmentation methods with respect to<br />

the segmentation parameters. The sensitivity analysis uses segmentation parameters that are random<br />

variables. The segmentation result is a stochastic image that contains information about the influence<br />

<strong>of</strong> the segmentation parameters. Thus, the stochasticity comes from the parameters and not from the<br />

input image, but the equations are nearly the same.<br />

Structure <strong>of</strong> the Thesis<br />

The thesis has the following structure: Chapter 2 presents segmentation methods for images based<br />

on PDEs. In particular, these are random walker segmentation, Ambrosio-Tortorelli segmentation,<br />

and methods based on level sets. Besides the presentation <strong>of</strong> these classical methods, this chapter<br />

discusses the drawbacks, especially for the propagation <strong>of</strong> errors. Furthermore, we review related<br />

work and highlight the differences between the related work and the methods proposed here.<br />

Chapter 3 contains an introduction into SPDEs and provides a theoretical background for the treatment<br />

<strong>of</strong> SPDEs. Furthermore, it presents the polynomial chaos expansion, a widely used tool for the<br />

approximation <strong>of</strong> random variables. The polynomial chaos expansion is the key for the numerical<br />

treatment <strong>of</strong> SPDEs and random variables, because this expansion converts the abstract idea <strong>of</strong><br />

random variables into a series expansion with deterministic coefficients. A computer can work with<br />

these coefficients, which enables the development <strong>of</strong> numerical methods for random variables. At the<br />

end <strong>of</strong> this chapter we highlight the advantages <strong>of</strong> the polynomial chaos over interval arithmetic [64].<br />

Chapter 4 investigates the discretization <strong>of</strong> SPDEs based on the polynomial chaos. The presented<br />

methods range from sampling based methods like Monte Carlo simulation and stochastic collocation<br />

to methods based on the polynomial chaos approximation for random variables. For the polynomial<br />

chaos, this chapter presents a finite difference method as well as SFEM and the GSD method.<br />

After the presentation <strong>of</strong> discretization methods for random variables and SPDEs in the previous<br />

chapters, Chapter 5 presents stochastic images. The concept <strong>of</strong> stochastic images is crucial for this<br />

thesis, because all methods developed in this thesis act on stochastic images. The main idea is to<br />

replace a pixel from a classical image by a random variable. Using the notion from stochastics, a<br />

stochastic image is a random field indexed by the position <strong>of</strong> the pixels inside the image. Besides<br />

4


the presentation <strong>of</strong> the stochastic images, Section 5.2 describes a possibility to generate stochastic<br />

images from image samples. This stochastic image generation is based on the method presented by<br />

Desceliers et al. [41], who applied an empirical Karhunen-Loève expansion on the centered covariance<br />

matrix. Section 5.4 investigates the visualization <strong>of</strong> 2D and 3D stochastic images.<br />

Chapter 6 generalizes two segmentation algorithms based on elliptic PDEs to stochastic segmentation<br />

methods acting on stochastic images. Section 6.1 deals with the extension <strong>of</strong> the random walker<br />

segmentation to stochastic images. The idea <strong>of</strong> the random walker segmentation is to prescribe a<br />

set <strong>of</strong> seed points for the objects and the background. Then a random walk starts at every unseeded<br />

point and the probability that the random walker goes from one point to another is dependent on<br />

the image intensity difference. In a stochastic image, this difference is a stochastic quantity and the<br />

probabilities for the walk <strong>of</strong> the random walker are stochastic quantities. The discretization <strong>of</strong> this<br />

method is based on the solution <strong>of</strong> a diffusion equation, because diffusion is the limit process <strong>of</strong> an<br />

infinite number <strong>of</strong> random walks.<br />

Section 6.2 investigates a stochastic extension <strong>of</strong> the Ambrosio-Tortorelli model for segmentation.<br />

The author presented this work at the European Conference on Computer Vision (ECCV) 2010 [3]<br />

and received the “ECCV 2010 Best Student Paper Award”. The idea is to replace all quantities in<br />

the Ambrosio-Tortorelli approach by their stochastic counterparts, yielding to two coupled SPDEs<br />

as the stochastic Euler-Lagrange equations for the computation <strong>of</strong> an energy minimizer. Using the<br />

GSD for the solution <strong>of</strong> the discretized SPDEs, the Ambrosio-Tortorelli method segments stochastic<br />

images computed from samples acquired via devices like digital camera or ultrasound imaging.<br />

Chapter 7 presents the last segmentation method for stochastic images investigated in this thesis,<br />

the segmentation <strong>of</strong> stochastic images with stochastic level sets. First, this chapter presents the<br />

extension <strong>of</strong> level sets to stochastic level sets, i.e. level sets evolving under an uncertain velocity.<br />

This extension is based on a parabolic approximation <strong>of</strong> the original level set equation. Having the<br />

stochastic level set extension at hand, it is possible to develop methods for the segmentation based<br />

on stochastic level sets. A method where the speed for the stochastic level set evolution is based<br />

on the image gradient and stochastic extensions <strong>of</strong> the geodesic active contour approach developed<br />

simultaneously by Caselles et al. [30] and Kichenassamy et al. [82] and the Chan-Vese approach [31].<br />

Chapter 8 deals with a sensitivity analysis <strong>of</strong> segmentation methods with respect to parameter<br />

changes. The sensitivity analysis uses the stochastic framework developed in the previous chapters,<br />

but applies it on a single deterministic input image. The stochasticity comes from the segmentation<br />

parameters that are random variables. With this modeling, we investigate the influence <strong>of</strong> the<br />

parameters on the result with the same segmentation framework developed for stochastic images.<br />

Chapter 9 contains a summary <strong>of</strong> the thesis along with a discussion. Furthermore, the chapter<br />

draws conclusions and gives directions for future work.<br />

5


Chapter 2<br />

Image <strong>Segmentation</strong> and Limitations<br />

In this chapter, we give a short review <strong>of</strong> the research in mathematical image processing and segmentation<br />

related to the work in this thesis. We focus on PDE based methods for image processing,<br />

because these methods have advantages over other image processing methods:<br />

• They are based on a continuous formulation <strong>of</strong> images, but the discretization based on finite<br />

differences or finite elements naturally leads to regular grids, characteristic for digital images.<br />

• It is possible to show existence and uniqueness <strong>of</strong> solutions <strong>of</strong> PDE based methods <strong>using</strong><br />

well-known results from functional analysis.<br />

• Later, we will see that PDE based methods extend naturally to stochastic images, the object<br />

under investigation in this thesis.<br />

The application <strong>of</strong> PDE models in image processing is a rapidly growing field <strong>of</strong> research. Many<br />

authors (see [17,130] for an overview) presented methods based on PDEs to solve problems arising in<br />

image processing like denoising, restoration, segmentation, registration, flow extraction, etc. Since<br />

we are interested in segmentation, the presentation focuses on results important for segmentation.<br />

Image segmentation, the separation <strong>of</strong> an image into object and background, is a repeatedly investigated<br />

problem in image processing. The literature divides the proposed methods into three<br />

categories, based on the user interaction necessary to perform the segmentation:<br />

Automatic segmentation: The user defines segmentation parameters at the beginning only, but<br />

has no possibility to refine the segmentation result.<br />

Semi-automatic segmentation: The user defines initial contours and parameters to optimize the<br />

segmentation result, but again has no chance to refine the result.<br />

Interactive segmentation: The user interactively refines the segmentation result. Thus, this<br />

method computes a segmentation result based on the user input and allows user interaction<br />

afterwards to get new input for the next iteration step.<br />

PDE based image segmentation methods are in all <strong>of</strong> these segmentation categories. The random<br />

walker segmentation [59] is an interactive segmentation approach, where the user interactively refines<br />

the segmentation result. The level set based segmentation methods [29, 96, 138] are semiautomatic,<br />

because the user has to define an initial contour as the starting point for the algorithm, but<br />

has no chance to influence the segmentation result during the run <strong>of</strong> the algorithm. The Mumford-<br />

Shah approach [107] is fully automatic. The user defines parameters only, but has no possibility to<br />

define initial contours or to modify the result locally afterwards.<br />

We organized this chapter as follows: First, we present basic definitions needed for the presentation<br />

<strong>of</strong> PDE based segmentation algorithms. Afterwards, we present five segmentation algorithms:<br />

random walker segmentation, Ambrosio-Tortorelli segmentation, and the level set based segmentation<br />

methods gradient-based segmentation, geodesic active contours and Chan-Vese segmentation.<br />

At the end, we present limitations <strong>of</strong> classical segmentation algorithms to motivate further investigations<br />

to extend these classical methods and draw conclusions.<br />

7


Chapter 2 Image <strong>Segmentation</strong> and Limitations<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

✘ ✘✘ ✘ ✘✘✘ u(x j ) = u j<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

✘ ✘ ✘ ✘✘ ✘ ✘ x i<br />

<br />

<br />

<br />

<br />

<br />

<br />

❳ ❳❳<br />

❳ ❳ ❳<br />

supp Pi (x)<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Figure 2.1: Sketch <strong>of</strong> the ingredients <strong>of</strong> a digital image. At every intersection <strong>of</strong> the regular grid lines<br />

a pixel is located and for every pixel the corresponding FE basis function has its support<br />

in the elements around this pixel.<br />

2.1 Mathematical <strong>Images</strong><br />

Before we start with the presentation <strong>of</strong> segmentation methods, we give a short overview over the<br />

notation and basic definitions for mathematical image processing. The primary object is the image:<br />

Definition An image is a function u from the image domain D ⊂ IR d ,d ∈ {2,3}, into the real numbers,<br />

i.e. u : D → IR. In what follows, the image domain D is a rectangular domain.<br />

Mathematical images are defined on a continuous space, i.e. they have an infinite number <strong>of</strong> values.<br />

An image acquired by a digital imaging device, e.g. a digital camera or advanced devices like CT [66]<br />

or MR [91], is called a digital image and the image intensities are known on a finite point set only:<br />

Definition A digital image (see Fig. 2.1) is a set <strong>of</strong> image intensities at the intersections <strong>of</strong> regular<br />

grid lines, called pixels. We denote the pixel value <strong>of</strong> the ith pixel <strong>of</strong> the digital image u by u i . The<br />

set <strong>of</strong> all pixels <strong>of</strong> a digital image is denoted by I and called the image grid.<br />

The link between this continuous definition and the pixel representation <strong>of</strong> digital images is the usage<br />

<strong>of</strong> an interpolation rule. Let us denote by P i the bilinear (2D) or trilinear (3D), basis function <strong>of</strong> the<br />

i-th pixel belonging to the multi-linear finite element space <strong>of</strong> the grid I . Then a digital image is<br />

interpolated at every point x in the image domain D by <strong>using</strong> the interpolation<br />

u(x) = ∑ u i P i (x) . (2.1)<br />

i∈I<br />

Remark 1. In what follows, we deal with gray value images only. This is not a strong restriction,<br />

because color images are typically composed <strong>of</strong> three color channels and it is possible to apply the<br />

methods presented in the following on these color channels separately when there is no coupling<br />

between the channels.<br />

Until now, we have no regularity assumptions on the image u, but to show existence and uniqueness<br />

<strong>of</strong> solutions <strong>of</strong> image processing methods, we have to restrict the analysis to images with a prescribed<br />

regularity. For the methods used in this thesis, the space <strong>of</strong> functions <strong>of</strong> bounded variation and<br />

generalizations <strong>of</strong> this space are important.<br />

8


2.2 Random Walker <strong>Segmentation</strong><br />

Definition The space <strong>of</strong> functions <strong>of</strong> bounded variation is<br />

{<br />

∫<br />

}<br />

BV(D) = u ∈ L 1 (D) : |Du|dx < ∞<br />

D<br />

. (2.2)<br />

Following [17] and <strong>using</strong> the Lebesgue decomposition theorem (see [32]) the derivative <strong>of</strong> a BVfunction<br />

decomposes into three parts, the absolutely continuous part ∇udx, the jump part D j u and<br />

the Cantor part D c u. This leads us to the definition <strong>of</strong> the class <strong>of</strong> special BV-functions:<br />

Definition The class <strong>of</strong> BV-function for which the Cantor part vanishes, i.e. D c u = 0, is called the<br />

space <strong>of</strong> special functions <strong>of</strong> bounded variation (SBV).<br />

Based on the space SBV we define the space <strong>of</strong> generalized functions <strong>of</strong> bounded variation (GSBV):<br />

Definition The space GSBV consists <strong>of</strong> all functions u ∈ L 1 (D) satisfying<br />

i.e. the truncated function belongs to SBV for all T .<br />

∀T > 0 : u T = sign(u)max(T,|u|) ∈ SBV , (2.3)<br />

In particular, the spaces SBV and GSBV are useful, when we introduce Mumford-Shah segmentation<br />

and the related Ambrosio-Tortorelli approximation.<br />

2.2 Random Walker <strong>Segmentation</strong><br />

Random walker segmentation performs on a single image u : D → IR defined on the image domain D.<br />

Since random walker segmentation is based on pixel values only, we need no additional assumptions<br />

about the smoothness <strong>of</strong> the images. The main idea <strong>of</strong> random walker segmentation is that the user<br />

prescribes a set <strong>of</strong> seed points for the object and the background. From the remaining unseeded<br />

points, random walks start and the percentage <strong>of</strong> random walks reaching the object seeds is the<br />

probability <strong>of</strong> the pixel to belong to the object.<br />

Before we begin to introduce random walker segmentation, we have to define notation for the<br />

graph representation <strong>of</strong> the image. A graph G is a pair G = (V,E) containing vertices or nodes<br />

v ∈ V and edges e ∈ E ⊂ V ×V. We denote an edge connecting the vertices v i and v j by e i j and<br />

identify e i j with e ji , because we are interested in nondirectional graphs only. Every edge has a<br />

weight w(e i j ) =: w i j that describes the costs for <strong>using</strong> the edge. A graph containing edge weights is<br />

a weighted graph. The summation <strong>of</strong> all edge weights for a node i,<br />

d i =<br />

∑<br />

{ j∈V :e i j ∈E}<br />

w(e i j ) , (2.4)<br />

is the degree <strong>of</strong> the node i.<br />

For random walker segmentation, we identify the image u with a graph G. The pixels <strong>of</strong> the digital<br />

image are the graph nodes and every pixel (respectively node) is connected to the neighboring nodes<br />

by a weighted edge. Fig. 2.2 shows a graph corresponding to a 3 × 3 image. For random walker<br />

segmentation the graph weights are<br />

w(e i j ) = exp ( −β(g i − g j ) 2) , (2.5)<br />

where (g i − g j ) 2 is the normalized difference between the image intensities at position i and j:<br />

(g i − g j ) 2 =<br />

The parameter β is the only free parameter that the user chooses.<br />

(u i − u j ) 2<br />

max {k,l∈V :ekl ∈E}(u k − u l ) 2 . (2.6)<br />

9


Chapter 2 Image <strong>Segmentation</strong> and Limitations<br />

✈<br />

✈<br />

✈<br />

✈<br />

✈<br />

✈<br />

✈✘ ✘ ✘ v k<br />

✎☞<br />

w 7<br />

✍✌ jk<br />

✘✘✘<br />

✈ v j<br />

✘ ✘ ✘ e i j<br />

✘✈<br />

✘ ✘ v i<br />

Figure 2.2: The graph generated from a 3 × 3 image contains 9 nodes and 12 edges. The edges e mn<br />

connect the nodes (the black dots) v l . Every edge e mn has a weight w mn describing the<br />

costs for traveling along this edge.<br />

2.2.1 Relation to the Dirichlet Problem<br />

The simulation <strong>of</strong> an infinite number <strong>of</strong> random walks is equivalent to solving a combinatorial Dirichlet<br />

problem [45,59]. Therefore, we review the Dirichlet problem in this section and show its relation<br />

to random walks. The Dirichlet integral is<br />

R[u] = 1 ∫<br />

|∇u| 2 dx . (2.7)<br />

2<br />

D<br />

A minimizer <strong>of</strong> the Dirichlet integral is a harmonic function, i.e. a function satisfying ∆u = 0 and the<br />

prescribed boundary conditions. The Dirichlet integral presented above is only useful for graphs with<br />

equal weights, for different weights we have to use a Dirichlet integral that respects these weights:<br />

R w [u] = 1 ∫<br />

w(x)|∇u| 2 dx . (2.8)<br />

2<br />

D<br />

A minimizer <strong>of</strong> (2.8) is a function that satisfies ∇ · (w∇u) = 0. To compute the minimizer <strong>of</strong> the<br />

discrete Dirichlet problem, i.e. to find a minimizer <strong>of</strong> the discrete version <strong>of</strong> (2.8), we introduce the<br />

combinatorial Laplacian matrix:<br />

⎧<br />

⎨ d i if i = j<br />

L i j = −w i j if v i and v j are adjacent nodes<br />

(2.9)<br />

⎩<br />

0 otherwise .<br />

Using the combinatorial Laplacian matrix, the graph discrete version <strong>of</strong> (2.8) is<br />

R[x] = 1 2 xT Lx , (2.10)<br />

where x is a vector containing all nodes/pixels <strong>of</strong> the graph resp. the image, i.e. x = (v 1 ,...,v n ).<br />

The user prescribes seed points V M = V O ∪V B for the object V O and background V B , see Fig. 2.3.<br />

These points act as boundary conditions for the Dirichlet problem, because the probability that a<br />

random walk starting at a seed point reaches it is one. The unseeded points V U are the degrees <strong>of</strong><br />

freedom. Reordering the nodes according to the set they belong to, (2.10) is written in block form<br />

R[x U ] = 1 [<br />

x<br />

T<br />

2 M xU] [ ][ ]<br />

T L M B xM<br />

B T L U x U<br />

= 1 (2.11)<br />

(<br />

x<br />

T<br />

2 M L M x M + 2xUB T T x M + xUL T )<br />

U x U .<br />

In (2.11) L M is the part <strong>of</strong> the matrix L describing the dependencies between the seed points V M .<br />

L U is the part with the dependencies between the unseeded pixels. Finally, B, respectively B T , is the<br />

10


2.2 Random Walker <strong>Segmentation</strong><br />

Figure 2.3: Left: Definition <strong>of</strong> the seed regions for the object (yellow) and the background (red).<br />

Middle: The probability that a random walker reaches an object seed. Black denotes<br />

probability zero, white probability one. Right: Random walker segmentation result <strong>of</strong> the<br />

ultrasound image. As input we used the seed regions from the left image and β = 200.<br />

part <strong>of</strong> the matrix describing the coupling between the seeded and unseeded pixels. Differentiation<br />

<strong>of</strong> (2.11) yields a minimizer <strong>of</strong> (2.11) given by the solution <strong>of</strong><br />

L U x U = −B T x M . (2.12)<br />

Remark 2. For 2D-images, the matrices L U and B are band matrices with five bands used only. The<br />

numerical solution <strong>of</strong> the system benefits from the use <strong>of</strong> numerical methods that make use <strong>of</strong> this<br />

special matrix structure, e.g. , it is necessary to store five bands as single vectors only. Furthermore,<br />

arithmetic operations for matrices with band structure can be implemented efficiently [57].<br />

As already mentioned, the random walker segmentation is an interactive segmentation method. Due<br />

to the fast calculation <strong>of</strong> the random walker result, the user interactively defines new seed regions or<br />

eliminates unwanted seed regions to get an optimal segmentation result. Fig. 2.4 shows three steps<br />

<strong>of</strong> the computer/user interaction for the refinement <strong>of</strong> a segmentation result.<br />

Another, more mathematical, motivation for the derivation <strong>of</strong> (2.12) is that we have to solve<br />

−∇ · (w∇u) = 0 in D<br />

u = 1 on V O<br />

u = 0 on V B .<br />

(2.13)<br />

Transforming this PDE to homogeneous boundary conditions [124], applying the reordering <strong>of</strong> the<br />

nodes, and <strong>using</strong> the combinatorial Laplacian, we end up with (2.12). Eq. (2.12) is a system <strong>of</strong> linear<br />

equations solvable by <strong>using</strong> iterative methods, e.g. the method <strong>of</strong> conjugate gradients.<br />

Remark 3. The presented segmentation method, the random walker segmentation, sounds like a<br />

stochastic method for segmentation that includes randomness and uncertainty, but this is false. Using<br />

the equivalence to the Dirichlet problem presented above, the method computes deterministic weights<br />

for an elliptic PDE on a graph. Thus, all “randomness” is lost. The “randomness” comes from the<br />

interpretation <strong>of</strong> the result: Every pixel gets a value between zero and one, and we interpret these<br />

values as probabilities for reaching the seed region from this specific pixel.<br />

The random walker segmentation result is a probability for every pixel for belonging to the object (see<br />

middle <strong>of</strong> Fig. 2.3). Typically, the threshold 0.5 distinguishes between the object and the background.<br />

A pixel having a probability above 50% is assigned to the object and a pixel with a probability below<br />

50% to the background. Fig. 2.3 shows a random walker segmentation result.<br />

11


Chapter 2 Image <strong>Segmentation</strong> and Limitations<br />

Figure 2.4: From left to right: Three steps <strong>of</strong> the interactive random walker segmentation. We show<br />

the seeds and the image to segment in the upper row and the segmentation corresponding<br />

to this particular choice <strong>of</strong> the seeds in the lower row. The addition <strong>of</strong> seed regions for<br />

the object and the background yield an iterative refinement <strong>of</strong> the segmentation.<br />

2.3 Mumford-Shah and Ambrosio-Tortorelli <strong>Segmentation</strong><br />

The minimization <strong>of</strong> a functional, as seen in the random walker segmentation, is a common technique<br />

for segmentation problems. The next method, the Mumford-Shah segmentation, bases on<br />

the minimization <strong>of</strong> a functional, too. The Mumford-Shah functional is not as easy as the random<br />

walker functional, because the Mumford-Shah functional involves two unknowns, the image and an<br />

additional edge set. This leads to a couple <strong>of</strong> mathematical problems for the theoretical pro<strong>of</strong> <strong>of</strong><br />

existence and uniqueness <strong>of</strong> minimizers and it is hard to discretize the Mumford-Shah functional<br />

directly. We avoid the numerical problems by introducing the Ambrosio-Tortorelli approximation<br />

that Γ-converges to the Mumford-Shah functional [14].<br />

Mumford and Shah [107] proposed to minimize the functional<br />

∫<br />

∫<br />

E MS (u,K) := (u − u 0 ) 2 dx + µ |∇u| 2 dx + νH d−1 (K) , (2.14)<br />

D\K<br />

where u 0 : D → IR is the initial image, u : D → IR is an image that is smooth and differentiable in D\K,<br />

K ⊂ D a set <strong>of</strong> discontinuities, µ,ν are nonnegative constants, and H d−1 (K) is the d −1-dimensional<br />

Hausdorff measure <strong>of</strong> the edge set K. The aim is to find an image u and a set K such that the functional<br />

is minimal. Roughly speaking, the minimizer u must be an image, which is close to the initial u 0 away<br />

from the edges (then ∫ D\K (u − u 0) 2 dx is small) and smooth away from the edges (then ∫ D\K |∇u|2 dx<br />

is small). Moreover, the length <strong>of</strong> the edge set K must be small (then H d−1 (K), measuring the length<br />

<strong>of</strong> the edge set, is small). The direct minimization <strong>of</strong> the Mumford-Shah energy is difficult due to<br />

the different nature <strong>of</strong> u and K: u is a function and K is a set. In addition, the pro<strong>of</strong> <strong>of</strong> existence <strong>of</strong> a<br />

D\K<br />

12


2.3 Mumford-Shah and Ambrosio-Tortorelli <strong>Segmentation</strong><br />

minimizer is a challenging problem, cf. [35]. Since the functional is not differentiable, the estimation<br />

<strong>of</strong> minimizers based on the Euler-Lagrange equations is impossible. Instead, researchers proposed<br />

regularized approximations (see [17]). The following paragraph summarizes one <strong>of</strong> these methods,<br />

proposed by Ambrosio and Tortorelli [14].<br />

Remark 4. All components <strong>of</strong> the Mumford-Shah functional are essential to get a segmentation <strong>of</strong><br />

the image u, i.e. it is impossible to omit one <strong>of</strong> the components to end up with a mathematically and<br />

numerically easier problem. If we omit the first component, we have no control over the difference<br />

between the image and the smooth approximation and u = 0, K = /0 minimize the remaining parts. We<br />

obtain another trivial solution if we omit the second component: Now u = u 0 and K = /0 minimize the<br />

functional. When omitting the last component, K = D minimizes the functional. Thus, the Mumford-<br />

Shah functional contains the minimal number <strong>of</strong> components necessary for segmentation, and it is<br />

essential to discretize them well to get meaningful numerical solutions.<br />

2.3.1 Ambrosio-Tortorelli <strong>Segmentation</strong><br />

As already mentioned, the Ambrosio-Tortorelli segmentation [14] is a kind <strong>of</strong> regularization <strong>of</strong> the<br />

Mumford-Shah functional. Ambrosio-Tortorelli segmentation uses a function φ : D → IR, the phase<br />

field, instead <strong>of</strong> the edge set K. The phase field is a smooth indicator function <strong>of</strong> the edge set K. It<br />

is zero on the edge set K and goes smoothly to one away from the edge set. An additional variable ε<br />

controls the width <strong>of</strong> the transition zone. When ε goes to zero, the phase field goes to the characteristic<br />

function <strong>of</strong> the edge set. In a following section, we will recap that the Ambrosio-Tortorelli<br />

energy converges in the Γ-sense to the Mumford-Shah energy [14].<br />

The idea <strong>of</strong> the Ambrosio-Tortorelli segmentation for a given initial image u 0 is to find a phase<br />

field φ and a smooth image u minimizing the energy<br />

where<br />

E AT [u,φ] := Efid,u ε [u] + Eε reg,u[u,φ] + Ephase ε [φ] , (2.15)<br />

E ε fid,u [u] = ∫<br />

D<br />

∫<br />

Ereg,u[u,φ] ε =<br />

∫<br />

Ephase ε [φ] =<br />

D<br />

D<br />

1<br />

2 (u − u 0) 2 dx<br />

µ ( φ 2 )<br />

+ k ε |∇u| 2 dx<br />

(<br />

νε|∇φ| 2 + ν 4ε (1 − φ)2) dx .<br />

(2.16)<br />

The first energy, the fidelity energy, ensures closeness <strong>of</strong> the smoothed image to the original u 0 . The<br />

second energy, the regularization energy, measures smoothness <strong>of</strong> u apart from areas where φ is<br />

small (the edges), and enforces φ to be small close to edges. The parameter k ε ensures coerciveness<br />

<strong>of</strong> the differential operator and existence <strong>of</strong> solutions, because φ 2 may vanish. The third energy,<br />

the phase energy, drives the phase field towards one and ensures small edge sets via the term |∇φ| 2 .<br />

The parameter ε controls the scale <strong>of</strong> the detected edges, µ the amount <strong>of</strong> detected edges, and ν the<br />

behavior <strong>of</strong> the phase field. k ε is a small regularization parameter.<br />

The relation between the first two components <strong>of</strong> the Ambrosio-Tortorelli and the Mumford-Shah<br />

energy are obvious. The third component, the phase energy, is a combination <strong>of</strong> a term forcing φ to<br />

be one and the term ∫ D ε|∇φ|2 . In the limit ε → 0, it can be shown to be equal to H d−1 (K) by <strong>using</strong><br />

the co-area formula [105].<br />

A minimizer <strong>of</strong> this energy is an image that is flat away from edges and a phase field, which<br />

is close to zero at edges only. To obtain a minimizer <strong>of</strong> an energy, a widely used technique is to<br />

solve the Euler-Lagrange equations resulting from this energy. For the computation <strong>of</strong> the Euler-<br />

Lagrange equations, we have to compute the first variation <strong>of</strong> the above energies <strong>using</strong> the Gâteaux<br />

13


Chapter 2 Image <strong>Segmentation</strong> and Limitations<br />

Figure 2.5: Left: The initial (noisy) US image treated as input for the Ambrosio-Tortorelli approach.<br />

Middle: The smooth Ambrosio-Tortorelli approximation <strong>of</strong> the initial image. Right: The<br />

corresponding phase field, i.e. the approximation <strong>of</strong> the edge set <strong>of</strong> the smoothed image.<br />

derivatives [17]. In the following θ : D → IR is a test function. For the fidelity energy, we get:<br />

d<br />

dε E ∫<br />

f id[u + εθ]<br />

∣ = 2(u + u 0 )θ dx . (2.17)<br />

ε=0<br />

For the other energies, we get similar results. The Euler-Lagrange equations <strong>of</strong> (2.15) are<br />

D<br />

−∇ · (µ(φ 2 + k ε )∇u ) + u = u 0<br />

( 1<br />

−ε∆φ +<br />

4ε + µ )<br />

2ν |∇u|2 φ = 1<br />

4ε . (2.18)<br />

This is a system <strong>of</strong> two coupled elliptic PDEs. We seek u,φ ∈ H 1 (D) as the weak solutions <strong>of</strong> these<br />

Euler-Lagrange equations. An implementation solves both equations alternately, letting either u or φ<br />

vary alternatingly until they reach a fixed point as the joint solution <strong>of</strong> both equations. Fig. 2.5 shows<br />

an exemplary result <strong>of</strong> the Ambrosio-Tortorelli segmentation approach.<br />

2.3.2 Γ-Convergence<br />

As already stated, the Ambrosio-Tortorelli functional approximates the Mumford-Shah functional.<br />

We show that the Ambrosio-Tortorelli functional converges in a variational sense towards the<br />

Mumford-Shah functional. This variational convergence is called Γ-convergence [14]:<br />

Definition The sequence <strong>of</strong> functionals F n : X → IR Γ-converges to the functional F if<br />

1. For every x ∈ X and for every sequence x n converging to x ∈ X,<br />

F(x) ≤ liminf<br />

n→∞ F n(x n ) . (2.19)<br />

2. For every x ∈ X there exists a sequence x n converging to x ∈ X such that<br />

F(x) ≥ limsupF n (x n ) . (2.20)<br />

n→∞<br />

The pro<strong>of</strong> <strong>of</strong> the Γ-convergence <strong>of</strong> a function sequence consists <strong>of</strong> two steps: First, we have to prove<br />

(2.19) for all sequences and then we have to construct a sequence fulfilling (2.20). This last step is<br />

the challenging task when proving Γ-convergence [14]. Using the definition <strong>of</strong> Γ-convergence and<br />

the space GSBV introduced in Section 2.1, the following theorem from [14, 17] holds.<br />

14


2.3 Mumford-Shah and Ambrosio-Tortorelli <strong>Segmentation</strong><br />

Theorem 2.1. Define Ẽ AT : L 1 (D) × L 1 (D) → IR + by<br />

{<br />

EAT (u,φ) if (u,φ) ∈ H<br />

Ẽ AT (u,φ) :=<br />

1 (D) × H 1 (D),0 ≤ φ ≤ 1<br />

+∞ otherwise<br />

(2.21)<br />

and G : L 1 (D) × L 1 (D) → IR + by<br />

{<br />

EMS (u) if u ∈ GSBV(D) and φ = 1 almost everywhere<br />

G(u,φ) =<br />

+∞ otherwise.<br />

(2.22)<br />

If k ε = o(ε), then Ẽ AT Γ-converges to G(u,φ) for ε → 0.<br />

The convergence <strong>of</strong> the Ambrosio-Tortorelli energy towards the Mumford-Shah energy enables us<br />

to use the coupled pair <strong>of</strong> PDEs obtained as Euler-Lagrange equations <strong>of</strong> the Ambrosio-Tortorelli<br />

energy and to solve this with very small ε. The result is a phase field that is close to the characteristic<br />

function <strong>of</strong> the edge set <strong>of</strong> the Mumford-Shah functional.<br />

2.3.3 Edge Continuity and Edge Consistency<br />

The classical Mumford-Shah model and the Ambrosio-Tortorelli approximation lack a step linking<br />

edges. This step is necessary to enforce the detection <strong>of</strong> closed contours in the images. Otherwise,<br />

the appearance <strong>of</strong> partially detected, breaking up contours is possible, see Fig. 2.5. For example,<br />

Erdem et al. [49] introduced such a step for the Ambrosio-Tortorelli model. The idea is to use a<br />

modified diffusion coefficient in the image equation. This modified coefficient does not depend on<br />

the phase field exclusively, but contains information about the continuity and directional consistency<br />

<strong>of</strong> the detected edges. To be more precise, Erdem et al. [49] proposed to use the equation<br />

−∇ · (µ((cφ) 2 + k ε )∇u ) + u = u 0 , (2.23)<br />

instead <strong>of</strong> the first equation <strong>of</strong> (2.18). The additional factor c is the product <strong>of</strong> the two factors from<br />

the directional consistency c dc and the edge continuity c h , i.e.<br />

c = c dc · c h . (2.24)<br />

If c < 1, the diffusivity decreases, allowing to form new edges in the image, whereas c > 1 leads to<br />

an increased diffusivity, allowing to smooth away unwanted edges.<br />

Directional Consistency<br />

The directional consistency tries to judge the quality <strong>of</strong> the detected edges based on information<br />

from surrounding pixels. The idea is that an edge is reliable if the gradients <strong>of</strong> the image for pixels<br />

in directions perpendicular to the edge are in parallel. For inaccurately detected edges, e.g. due to<br />

noise, these gradients are typically not aligned. To do so, Erdem et al. [49] introduced<br />

(c dc ) i = ζ dc<br />

i<br />

+ 1 − ζ i<br />

dc<br />

(2.25)<br />

φ i<br />

for all pixels i ∈ I , where ζi<br />

dc measures the alignment <strong>of</strong> the gradients. This factor increases the<br />

diffusion, if the image gradients around the detected edge are not aligned. As the feedback measure<br />

for the alignment <strong>of</strong> the gradients they proposed to use<br />

( ( ))<br />

1<br />

ζi<br />

dc = exp ε dc |η s | ∑ ∇v<br />

j∈η s i · ∇v j − 1 , (2.26)<br />

15


Chapter 2 Image <strong>Segmentation</strong> and Limitations<br />

where ∇v i and ∇v j are the normalized gradients at position i respectively j, i.e. ∇v k = ∇u k /|∇u k |.<br />

Eq. (2.26) is close to one if the gradients are aligned (than the scalar product ∇v i · ∇v j is close to<br />

one) and close to zero if the gradients are not aligned. The set η s contains s pixels in the direction<br />

perpendicular to the image gradient and the parameter ε dc controls the influence <strong>of</strong> the gradient.<br />

For the numerical experiments we used ε dc = 0.25 and four pixels in directions perpendicular to the<br />

image gradient, i.e. |η s | = 4.<br />

Edge Continuity<br />

To avoid the breaking up <strong>of</strong> edges, Erdem et al. [49] proposed to use an additional feedback measure,<br />

which lowers the diffusivity around detected edges, to allow a growth <strong>of</strong> the detected edges. Using a<br />

simplified version <strong>of</strong> the original model [49], the feedback measure is<br />

(c h ) i =<br />

1<br />

1 + φ i − φ 2<br />

i<br />

, (2.27)<br />

where φ is the phase field. We use a slight modification <strong>of</strong> the above feedback measure by adding an<br />

additional scale factor α, allowing us to weight the deviation <strong>of</strong> the phase field from 0 and 1:<br />

1<br />

(c h ) i =<br />

1 + α ( )<br />

φ i − φi<br />

2 . (2.28)<br />

In the numerical experiments, we set α = 10. Thus, the diffusivity decreases in regions, where the<br />

phase field is away from zero and one, i.e. regions, where the edge detection is in an intermediate state<br />

between smoothing the structure away and building up a sharp edge. Fig. 2.6 shows a comparison<br />

between Ambrosio-Tortorelli segmentation with and without the additional factor c.<br />

2.4 Level Sets for Image <strong>Segmentation</strong><br />

The Ambrosio-Tortorelli segmentation approach introduces a new quantity besides the image and a<br />

smooth representation: the phase field. This phase field approximates the edge set in the Ambrosio-<br />

Tortorelli approach, but even with the above modifications there is no guarantee to obtain a connected<br />

phase field in the end.<br />

Level set based segmentation methods use another viewpoint. They place a closed curve somewhere<br />

in the image and try to adjust this initial curve to the edges, respectively the object boundaries<br />

in the image. This approach assures that the final segmentation result is a closed contour. For the<br />

representation <strong>of</strong> the contour, there are methods for an explicit [76] or an implicit [29,31,40,96,138]<br />

representation available in the literature. A famous method for the explicit representation is the snake<br />

model [76], but explicit representations have drawbacks: parametrization <strong>of</strong> the curve, distribution<br />

<strong>of</strong> the nodes describing the curve, dependence <strong>of</strong> the result on the parametrization, etc. To avoid<br />

these shortcomings we focus on the implicit representation based on level sets in the following.<br />

Dervieux and Thomasset [40] and Osher and Sethian [121, 138] developed the level set method.<br />

The main idea is the implicit representation <strong>of</strong> a curve by embedding a curve C 0 ⊂ IR n into a higherdimensional<br />

function φ : IR n+1 → IR and to identify the zero level set <strong>of</strong> φ with the curve, i.e.<br />

C 0 = {x ∈ IR n : φ(0,x) = 0} . (2.29)<br />

It is possible to describe the motion <strong>of</strong> the curve by the motion <strong>of</strong> the level sets <strong>of</strong> φ. At any time t > 0,<br />

we get the curve C(t) back from the level set representation via C(t) = {x ∈ IR n : φ(t,x) = 0}. Using<br />

this concept, we describe the motion <strong>of</strong> the curve <strong>using</strong> the level set equation [138]<br />

φ t + F|∇φ| = 0 , (2.30)<br />

16


2.4 Level Sets for Image <strong>Segmentation</strong><br />

input image without edge linking edge linking<br />

image<br />

phase field<br />

n/a<br />

Figure 2.6: Comparison <strong>of</strong> the Ambrosio-Tortorelli model (left) and the extended model <strong>using</strong> the<br />

edge linking procedure (right). Data set provided by PD Dr. Christoph S. Garbe.<br />

where F : IR n+1 → IR is the speed in the normal direction. For the discretization <strong>of</strong> (2.30) Osher<br />

and Sethian [121] developed numerical methods based on Hamilton-Jacobi equations. The use<br />

<strong>of</strong> simple finite difference approximations, like central differences, fails due to the hyperbolic nature<br />

<strong>of</strong> (2.30) [138]. In addition, Sethian [138] developed efficient methods, like the Narrow Band<br />

Method, where the equation is solved in the surrounding <strong>of</strong> the zero level set only.<br />

Due to numerical reasons, level set methods use signed distance functions, i.e. functions that satisfy<br />

|∇φ| = 1, as level set function. Since the function loses this attribute during the evolution <strong>of</strong><br />

the curve, we reinitialize the signed distance function from time to time. For this purpose, methods<br />

are available ranging from iterative methods, e.g. solving φ t = sign(φ)(1 − |∇φ|) to steady state<br />

(see [138]), to efficient – every grid point is only visited once – Fast Marching methods [138].<br />

Besides the application for image segmentation, other research fields like computer-aided design,<br />

flow simulations, or optimal path planning [138] use level sets. Furthermore, it is possible to use a<br />

level set approach for the simulation <strong>of</strong> phase change problems. In this context, the author investigated<br />

the simulation <strong>of</strong> the phase change during radio-frequency ablation [10, 12, 13].<br />

2.4.1 Phase Field Models<br />

Phase fields, like the phase field in the Ambrosio-Tortorelli approach, have a close relation to level<br />

sets. In fact, the literature [55,137] refers to level set methods as “sharp interface approach”, because<br />

level sets know the position <strong>of</strong> the interface precisely due to the implicit tracking. On the other hand,<br />

one refers to phase fields as “diffusive interface approach”, because phase fields are constant away<br />

from the interface and vary smoothly near the interface. Phase field methods treat the transition<br />

zone around the interface as a zone with mixed content <strong>of</strong> the regions separated by the interface.<br />

17


Chapter 2 Image <strong>Segmentation</strong> and Limitations<br />

Phase fields are frequently used for interface tracking (see [23, 143] and the references therein), also<br />

the image processing community uses phase fields for segmentation purposes [15, 123]. In contrast<br />

to the phase field used in the Ambrosio-Tortorelli approach, the phase fields needed in this context<br />

differentiate between object and background, i.e. they describe closed contours. They have a value <strong>of</strong><br />

−1 in the object, 0 at the interface, +1 on the background and vary smoothly between these values.<br />

Typically, the phase field is ±1 away from the interface and changes smoothly inside a small layer<br />

with thickness ε around the interface in a tangential pr<strong>of</strong>ile. A phase field equation [143] like<br />

(<br />

φ t + F|∇φ| + u e · ∇φ = b ∆φ + φ(1 − φ 2 )<br />

)<br />

ε 2 (2.31)<br />

controls the evolution <strong>of</strong> the diffusive phase field interface. In this equation, φ is the phase field, u e an<br />

external advection, b the interface speed depending on the curvature, ε the thickness <strong>of</strong> the diffusive<br />

interface, and F again the speed in the normal direction. In contrast to the level set approach (2.30),<br />

this is a parabolic equation, which avoids the numerical difficulties arising in the discretization <strong>of</strong><br />

the level set equation. When the curvature-depending interface speed vanishes, i.e. when b = 0, this<br />

equation is kept parabolic by adding a counter term introduced by Folch [51]. Following [143] this<br />

counter term leads to the parabolic equation<br />

(<br />

φ t + F|∇φ| + u e · ∇φ = b ∆φ + φ(1 − φ 2 ( ))<br />

)<br />

∇φ<br />

ε 2 − |∇φ|∇ ·<br />

, (2.32)<br />

|∇φ|<br />

where b is a purely numerical parameter, because the curvature at the end cancels out the Laplacian<br />

and the term φ(1−φ 2 )<br />

. It is easier to discretize (2.32) than (2.30) <strong>using</strong> central differences. Sun et<br />

ε 2<br />

al. [143] developed a phase field equation based on nonlinear preconditioning [56]<br />

(<br />

φ t + a|∇φ| + u e · ∇φ = b ∆φ + 1 ε (1 − |∇φ|2 ) √ ( ) ( ))<br />

φ<br />

∇φ<br />

2tanh √ − |∇φ|∇ ·<br />

, (2.33)<br />

2ε |∇φ|<br />

which is an integrated reinitialization scheme for the phase field. The phase field φ in this equation<br />

becomes a signed distance function. Thus, this equation is a parabolic level set equation with integrated<br />

reinitialization. Again, it is possible to discretize (2.33) <strong>using</strong> simple difference schemes. This<br />

connection between phase fields and level sets allows us to use known segmentation algorithms from<br />

the level set context and embed them into this nonlinear preconditioned phase field equation. The<br />

discretization <strong>of</strong> the parabolic phase field equations is easier in the stochastic context, cf. Chapter 7.<br />

2.4.2 Gradient-Based <strong>Segmentation</strong><br />

The idea <strong>of</strong> the level set propagation is useful to segment objects inside an image. The simplest<br />

approach for segmentation based on level sets is to use a speed F in the level set equation that<br />

depends on characteristics <strong>of</strong> the image. Popular is F = F(|∇u|), i.e. to stop the evolution on edges<br />

inside the image (see [29, 96, 138] and the references therein). Caselles et al. [29] proposed to use<br />

with<br />

g u =<br />

φ t + g u |∇φ| = 0 (2.34)<br />

1<br />

(1 − εκ) , (2.35)<br />

1 + |∇G σ ∗ u|<br />

where u is the image, G σ a Gaussian smoothing filter with width σ, κ the curvature <strong>of</strong> the level set<br />

function and ε a small scale parameter that controls the influence <strong>of</strong> the curvature smoothing term.<br />

Although this idea sounds simple, the method achieves good results when a high gradient separates<br />

the objects from the background (see Fig. 2.7). One major drawback <strong>of</strong> this method is the need for<br />

18


2.4 Level Sets for Image <strong>Segmentation</strong><br />

Figure 2.7: <strong>Segmentation</strong> <strong>of</strong> a medical image based on a level set propagation with gradient-based<br />

speed function. The time increases from left to right and the zero level set (red line)<br />

approximates the boundary <strong>of</strong> the object (a liver mask) at the end.<br />

finding a stopping criterion. The evolution speed g u is always positive, even close to edges. Thus, it<br />

is possible that the zero level set passes the edge. A typical stopping criterion is to stop the evolution<br />

when the difference between the level sets <strong>of</strong> subsequent time steps is small. This occurs when the<br />

level set reached the boundary <strong>of</strong> the object and the speed dropped down. Using methods that are<br />

more sophisticated, it is possible to stop the zero level set at the edge. Thus, these methods have a<br />

convergent solution. The next section presents one <strong>of</strong> these methods, geodesic active contours.<br />

Remark 5. It is also possible to formulate the gradient-based segmentation based on the phase field<br />

model presented in the last section. This yields the equation<br />

(<br />

φ t + g u |∇φ| = ε ∆φ + φ(1 − φ 2 )<br />

)<br />

ε 2 . (2.36)<br />

2.4.3 Geodesic Active Contours<br />

Caselles et al. [30] and simultaneously Kichenassamy et al. [82] developed geodesic, or minimal<br />

distance, active contours. They minimize an energy B that depends on the curve C and on the<br />

parametrization <strong>of</strong> the curve C(q) : [0,1] → IR 2 :<br />

∫ 1<br />

B(C) = α |C ′ (q)| 2 dq + β<br />

0<br />

∫ 1<br />

0<br />

g u (|∇u(C(q))|) 2 dq , (2.37)<br />

where g u is the edge indicator from the last section. They computed a minimizer <strong>of</strong> this energy<br />

by <strong>using</strong> a level set representation <strong>of</strong> the curve and computing the Euler-Lagrange equations <strong>of</strong> the<br />

resulting energy. This leads to a level set equation with an additional advection term that forces the<br />

zero level set to stay in regions with high gradient:<br />

φ t = −α∇g u · ∇φ − βg u |∇φ| + εκ|∇φ| . (2.38)<br />

The user chooses the parameters α,β and ε. For given parameters and an initial level set we solve<br />

to steady state. Fig. 2.8 shows a typical geodesic active contours segmentation result.<br />

2.4.4 Chan-Vese <strong>Segmentation</strong><br />

The segmentation methods presented so far are based on a high gradient that separates the object<br />

from the background. When such a gradient is not present, the methods fail to segment the object.<br />

19


Chapter 2 Image <strong>Segmentation</strong> and Limitations<br />

Figure 2.8: <strong>Segmentation</strong> <strong>using</strong> geodesic active contours. Left: The initial image. Right: Solution <strong>of</strong><br />

the geodesic active contour method initialized with small circles inside the object.<br />

Chan and Vese [31] proposed a method that is independent <strong>of</strong> gradient information. Instead, they<br />

proposed to segment homogeneous regions inside the image. To be more precise, Chan and Vese [31]<br />

proposed to minimize the functional<br />

∫<br />

∫<br />

F(c 1 ,c 2 ,C) = µ · Length(C) + ν · Area(inside(C)) + λ 1 |u 0 − c 1 | 2 dx + λ 2 |u 0 − c 2 | 2 dx .<br />

inside(C)<br />

The corresponding Euler-Lagrange equation is<br />

( ( ) )<br />

∇φ<br />

φ t = δ(φ) µ∇ · − ν − λ 1 (u 0 − c 1 ) 2 + λ 2 (u 0 − c 2 ) 2<br />

|∇φ|<br />

outside(C)<br />

(2.39)<br />

, (2.40)<br />

where δ is the Dirac δ-function [42]. This equation is a parabolic PDE that contains a curvature<br />

smoothing term, a term penalizing the segmented area, and two terms penalizing variations from<br />

the mean value <strong>of</strong> the segmented object and the background. Instead <strong>of</strong> δ, we use a regularized<br />

δ-function δ ε for the discretization given by the derivative <strong>of</strong> the Heaviside approximation<br />

H ε = 1 (<br />

1 + 2 ( z<br />

) )<br />

2 π arctan . (2.41)<br />

ε<br />

By <strong>using</strong> H ε from above, δ ε is<br />

1<br />

δ ε (x) =<br />

πε + π . (2.42)<br />

ε<br />

x2<br />

The mean value <strong>of</strong> the object and the background can be computed <strong>using</strong> the Heaviside function:<br />

∫<br />

D<br />

c 1 (φ) =<br />

u ∫<br />

0(x)H ε (φ(x))dx<br />

∫<br />

D<br />

D H , resp. c 2 (φ) =<br />

u 0(x)(1 − H ε (φ(x)))dx<br />

∫<br />

ε(φ(x))dx<br />

D (1 − H . (2.43)<br />

ε(φ(x)))dx<br />

The user chooses λ 1 ,λ 2 , µ,ν. The advantage <strong>of</strong> the Chan-Vese model is that it does not need edges<br />

in the image to segment objects. In fact, the model is independent <strong>of</strong> gradient information. Instead,<br />

it tries to separate homogeneous regions in the image. Fig. 2.9 shows a typical result <strong>of</strong> Chan-Vese<br />

segmentation on an image without edges.<br />

This concludes the presentation <strong>of</strong> classical segmentation algorithms based on PDEs. The presented<br />

segmentation algorithms range from interactive, nearly parameter free algorithms, like random<br />

walker segmentation, over semi-automatic with a moderate number <strong>of</strong> variables, like the level<br />

set based algorithms, to automatic methods like Mumford-Shah segmentation, where no user interaction<br />

is necessary. All these segmentations are able to produce accurate results on a wide range<br />

20


2.5 Why is Classical Image Processing not Enough?<br />

Figure 2.9: <strong>Segmentation</strong> <strong>of</strong> an object without sharp edges <strong>using</strong> the Chan-Vese approach. In red, we<br />

show the steady-state solution <strong>of</strong> the Chan-Vese segmentation method initialized with a<br />

small circle inside the object.<br />

<strong>of</strong> images and from the perspective <strong>of</strong> segmentation <strong>of</strong> single images, there is no need for new concepts.<br />

Nevertheless, the approaches presented in this chapter have drawbacks regarding robustness<br />

with respect to noise, reproducibility, and error propagation. The next section investigates this.<br />

2.5 Why is Classical Image Processing not Enough?<br />

In the last sections, we introduced five segmentation methods and showed that all these segmentation<br />

methods perform well on some selected medical images. Besides the segmentation <strong>of</strong> single images,<br />

a segmentation method has to fulfill other features not presented so far:<br />

• It is unclear how robust the methods are with respect to image noise.<br />

• The robustness <strong>of</strong> the methods for parameter changes and different initializations is unclear.<br />

• Propagating error information through these algorithms is hard, i.e. if information about measurement<br />

errors at the image acquisition is available, it is impossible to propagate this information<br />

through the segmentation to get segmentation results containing the error information.<br />

We organized this section as follows: First, we give an introduction to image noise, show how the<br />

image noise influences the image quality for different acquisition modalities and how image noise<br />

is modeled mathematically. Then, we investigate the noise robustness <strong>of</strong> the presented segmentation<br />

methods, and finally, we discuss error propagation in classical image segmentation methods.<br />

2.5.1 Image Noise<br />

Image noise is a serious problem when dealing with medical images and images from digital cameras.<br />

Different noise sources degrade the images. A principal problem <strong>of</strong> image acquisition devices is the<br />

noise due to the random arrival <strong>of</strong> the photons. Light or X-ray emission is a stochastic process [44].<br />

In addition, the instrumentation noise due to thermal effects in the acquisition device degrades the<br />

image quality. Further sources <strong>of</strong> image noise are the quantization noise due to the conversion<br />

from analog to digital signals and the compression process for the images, if any. Physical effects<br />

influencing the path <strong>of</strong> the photons, like blurring, diffraction, and scattering, cause image noise, too.<br />

21


Chapter 2 Image <strong>Segmentation</strong> and Limitations<br />

Figure 2.10: A test pattern corrupted by uniform (left), Gaussian (middle), and speckle noise (right).<br />

It is possible to reduce some <strong>of</strong> the noise sources by averaging the image values over a period.<br />

When a signal is available for a period L, the expected value <strong>of</strong> a pixel is<br />

1<br />

E(a) = lim<br />

L→∞ L<br />

∫ L<br />

0<br />

a(x)dx . (2.44)<br />

When the probability density function (PDF) <strong>of</strong> the process a is known the integral reduces to<br />

E(a) =<br />

∫ ∞<br />

where ρ is the PDF. The variance <strong>of</strong> the stochastic process is (cf. [44])<br />

σ 2 =<br />

∫ ∞<br />

With these quantities, the signal-to-noise ratio (SNR) [44] is<br />

−∞<br />

−∞<br />

aρ(a)da , (2.45)<br />

(a − E(a)) 2 ρ(a)da . (2.46)<br />

SNR = |E(a)|2<br />

σ 2 . (2.47)<br />

One divides the noise sources into additive and multiplicative noise sources. Fig. 2.10 shows three<br />

noise models. Additive noise is modeled via<br />

g(x) = f (x) + n(x) , (2.48)<br />

where g is the measured signal, f the true signal and n the noise. Multiplicative noise is modeled via<br />

g(x) = f (x) + n(x) f (x) . (2.49)<br />

The multiplicative noise depends on the image value. In what follows, we use the additive noise<br />

model, because we are not directly interested in the noise modeling, but need a noise model as input<br />

for the stochastic image processing framework. Once the noise is characterized, the noise is no<br />

longer a free parameter and (2.49) can be expressed as<br />

g(x) = f (x) + ñ(x) , (2.50)<br />

and it is possible to use the additive model.<br />

All these sources <strong>of</strong> noise influence the image quality and it is not well understood how the noise<br />

influences the segmentation result, e.g. how the image noise influences the segmented object volume.<br />

This is due to the construction <strong>of</strong> typical segmentation algorithms. They have no knowledge about the<br />

noise that corrupted the image to segment and it is impossible to apply the segmentation algorithm<br />

on noise realizations, apart from artificial test data corrupted with a known noise model. In the<br />

next two sections, we present two problems related to image noise that cannot be investigated with<br />

deterministic segmentation models, apart from <strong>using</strong> a sampling based approach.<br />

22


2.6 Work Related to the <strong>Stochastic</strong> Framework<br />

2.5.2 Robustness<br />

Robustness <strong>of</strong> segmentation methods is desirable in two ways: The methods should be robust with<br />

respect to the noise and with respect to the segmentation parameters.<br />

Robustness with respect to the segmentation parameters, e.g. the β for random walker segmentation<br />

or µ,ν, and ε for Ambrosio-Tortorelli segmentation, is necessary to get stable results. When the<br />

segmentation result changes significantly for small parameter changes, the results are arbitrary, and<br />

it is not recommendable to base e.g. medical diagnoses on such a segmentation result. It is possible<br />

to investigate this kind <strong>of</strong> robustness by comparing the results <strong>of</strong> segmentations with slightly modified<br />

parameters or by treating the segmentation parameters as random variables and investigate the<br />

variance <strong>of</strong> the segmentation result. Chapter 8 <strong>of</strong> this thesis is about this.<br />

Robustness with respect to noise is an essential property <strong>of</strong> a segmentation method for medical<br />

images. The real, noise-free, image is not available, and it is a random choice which noise realization<br />

the image at hand shows. It is desirable for segmentation methods to be robust with respect to the<br />

noise realization, i.e. the segmentation result should not vary much for noise realizations. To investigate<br />

this, it is possible to run the segmentation on noise realizations or to make image pixels random<br />

variables, describing the process leading to the image noise. The first way is time-consuming. We<br />

will see this later in the thesis, e.g. in Section 6.1. The second way is the fundamental idea <strong>of</strong> this<br />

thesis. It needs a theoretical foundation, which the following chapters will present.<br />

2.5.3 Error Propagation<br />

Image processing widely neglects error propagation. Nearly all methods in image processing consider<br />

the available data as the “truth”, but as we saw, the real data is not available. Instead, we have to<br />

use an image corrupted by a random noise realization. When neglecting this error introduced by the<br />

noise and other imaging artifacts we end up with results looking precise, but ignoring the influence<br />

<strong>of</strong> the noise. It is desirable to have image processing methods that are able to deal with information<br />

about the image noise, e.g. via the mentioned description <strong>of</strong> noise via the introduction <strong>of</strong> random<br />

variables for the image pixels. The next chapters deal exactly with this new idea for the processing<br />

<strong>of</strong> images and provide a theoretical background.<br />

2.6 Work Related to the <strong>Stochastic</strong> Framework<br />

In this section, we review work sounding similar to the work presented in this thesis and identify the<br />

differences and similarities. A lot <strong>of</strong> authors presented methods for image segmentation that take<br />

more or less stochasticity into account, e.g. via a modeling with Markov random fields or stochastic<br />

annealing, but none <strong>of</strong> the methods mentioned can propagate stochastic information from the input<br />

to the output <strong>of</strong> the segmentation process or model pixels as random variables.<br />

Markov Random Fields for Image <strong>Segmentation</strong><br />

The literature [39, 65, 153] uses Markov random fields (MRFs) for image segmentation frequently.<br />

MRFs are a possibility to model the noise in the input image, but the result <strong>of</strong> MRF segmentation<br />

is a deterministic segmentation result along with a map to remove the noise from the input image.<br />

Thus, this method tries to incorporate the noise via a stochastic modeling approach, but is not able<br />

to propagate uncertainty information from the input to the output.<br />

23


Chapter 2 Image <strong>Segmentation</strong> and Limitations<br />

Random Level Set Functions<br />

Stefanou et al. [141] presented a method to obtain a random level set, i.e. a polynomial chaos expansion<br />

<strong>of</strong> a level set function. The proposed work relates to this work, but [141] obtains the polynomial<br />

chaos expansion <strong>of</strong> the level set from a classical level set approach on the samples and an estimation<br />

<strong>of</strong> the polynomial chaos coefficient afterwards. Computing the level set equation on the samples is<br />

exactly the objective we want to overcome in this work by applying a stochastic level set equation<br />

on a stochastic image, obtained from the samples. This results in a significant speed-up, because in<br />

the stochastic framework the level set method has to be applied only once on the stochastic image,<br />

not on every sample.<br />

<strong>Stochastic</strong> Active Contours<br />

The work presented in this thesis should not be confounded with the method presented by Juan et<br />

al. [75] named “<strong>Stochastic</strong> Active Contour”. Although the title suggests a close relation between<br />

the stochastic level set equation and the work presented in [75], they are opposed. Juan et al. [75]<br />

proposed a technique to overcome drawbacks <strong>of</strong> the classical level set approach by adding fictive<br />

noise to the image. This relates to the simulated annealing technique [83]. The idea <strong>of</strong> the stochastic<br />

level set framework is the development <strong>of</strong> a stochastic active contour moving under an uncertain<br />

velocity, obtained from the uncertain gray values <strong>of</strong> the stochastic pixels.<br />

A Multiresolution <strong>Stochastic</strong> Level Set Method for Mumford-Shah Image <strong>Segmentation</strong><br />

Law et al. [89] use a method called “<strong>Stochastic</strong> Level Set Method” for the segmentation <strong>of</strong> objects.<br />

This also should not be confounded with the stochastic level set work presented in this thesis, because<br />

Law et al. proposed a method to overcome the drawback <strong>of</strong> the level set segmentation to run into local<br />

minima. They developed a method that “jumps” out <strong>of</strong> these local minima to get a global minimum<br />

<strong>of</strong> the solution. Again, this method is related to the simulated annealing technique [83].<br />

Error-in-Variables Likelihood Functions for Motion Estimation<br />

Nestares et al. [111,112] where able to compute confidence measures for error propagation in motion<br />

estimation [71]. Their method allows to estimate the influence <strong>of</strong> the image noise on the computed<br />

motion field. To achieve this, they combined Bayesian estimation [77] and likelihood functions to<br />

solve a total-least squares problem [58]. Nevertheless, their investigations are restricted to independent,<br />

identically distributed Gaussian random variables. Thus, their proposed framework can be used<br />

in rare situations only.<br />

Conclusion<br />

We presented the basics for mathematical image processing and gave a short overview over segmentation<br />

methods based on PDEs. All these methods produce good results on single images, but are<br />

unable to deal with error propagation. Furthermore, the robustness <strong>of</strong> the methods with respect to<br />

image noise and parameter changes is unknown. This highlights the need for error-aware methods<br />

for segmentation and image processing in general. Before we are able to present these methods, we<br />

have to provide background from stochasticity for the representation <strong>of</strong> random variables and about<br />

SPDEs. This is the task <strong>of</strong> the next chapter.<br />

24


Chapter 3<br />

SPDEs and Polynomial Chaos Expansions<br />

This chapter deals with the fundamentals required to develop stochastic images. First, we review<br />

notation and results from probability theory. Afterwards, we introduce SPDEs and the polynomial<br />

chaos expansion, the main ingredient for the numerical approximation <strong>of</strong> random variables.<br />

3.1 Basics from Probability Theory<br />

This section provides background from probability theory for the presentation <strong>of</strong> the stochastic images<br />

and SPDEs. First, we introduce the basic ingredients, probability measures, probability spaces<br />

and random variables.<br />

Definition A probability space (Ω,A ,Π) is a triple consisting <strong>of</strong> a sample space Ω containing all<br />

possible outcomes, a σ-algebra <strong>of</strong> events A ⊂ 2 Ω and a probability measure Π. The probability<br />

measure Π is defined on the σ-algebra A and has the following properties:<br />

• Π is non-negative: Π(A) ≥ 0 for all A ∈ A .<br />

• The measure <strong>of</strong> the sample space Ω is one: Π(Ω) = 1.<br />

• Π is countable additive, i.e. for a countable number <strong>of</strong> pairwise disjoint sets A i ⊂ A we have<br />

Π(∪A i ) = ∑(Π(A i )).<br />

On the probability space (Ω,A ,Π) we define functions from this space into the real numbers.<br />

Definition A random variable f : Ω → IR is a function from the sample space Ω into the real numbers<br />

that is measurable with respect to the σ-algebras A and B, where B is the Borel measure.<br />

Random variables are an important object for the definition <strong>of</strong> stochastic images. In Chapter 5, we<br />

will see that every pixel <strong>of</strong> a stochastic image is a random variable. For random variables, it is<br />

possible to define the probability density function (PDF):<br />

Definition The function ρ is called probability density function (PDF) <strong>of</strong> the random variable f if it<br />

satisfies Π(a < f < b) = ∫ b<br />

a ρ(x)dx for all a,b ∈ IR.<br />

Having the probability density at hand, we define further properties <strong>of</strong> random variables. The most<br />

important property <strong>of</strong> random variables is the expected value:<br />

Definition The expected value or first moment <strong>of</strong> a random variable X : Ω → IR with PDF ρ is<br />

∫<br />

∫<br />

∫<br />

E(X) = X(ω)dω = xρ(x)dx = xdΠ . (3.1)<br />

Ω<br />

In (3.1) we used dΠ = f dx to characterize integration with respect to the PDF.<br />

Knowing the probability density <strong>of</strong> a random variable allows us to transform the integral over the<br />

sample space Ω into an easier computable integral over the real numbers weighted by the probability<br />

density. Using this equality, it is also possible to compute higher-order moments <strong>of</strong> random variables:<br />

IR<br />

IR<br />

25


Chapter 3 SPDEs and Polynomial Chaos Expansions<br />

Definition The n-th central moment <strong>of</strong> a random variable X is<br />

M n (X) = E((X − E(X)) n ) . (3.2)<br />

A famous member <strong>of</strong> this class <strong>of</strong> moments is the second central moment, the variance:<br />

Var(X) = E ( (X − E(X)) 2) . (3.3)<br />

Later we have to evaluate the relation between random variables. A famous tool for this is the<br />

covariance.<br />

Definition The covariance <strong>of</strong> two random variables f ,g with finite second order moments is<br />

Cov( f ,g) = E( f g) − E( f )E(g) . (3.4)<br />

In what follows, it will be necessary to have a set <strong>of</strong> random variables indexed by a spatial position.<br />

This motivates the following definition:<br />

Definition A random field X is a collection <strong>of</strong> random variables indexed by a spatial position x ∈ IR n :<br />

X = {X x |x ∈ IR n } . (3.5)<br />

Random fields are elements <strong>of</strong> a tensor product space consisting <strong>of</strong> functions defined on the Cartesian<br />

product Ω × D. A random field is a function taking two arguments, a random event and a spatial<br />

position. We restrict the investigations to random fields satisfying smoothness assumptions.<br />

Definition Let D ⊂ IR n be the spatial domain <strong>of</strong> the random field and Ω the sample space. The<br />

tensor space L 2 (Ω) ⊗ H 1 (D) is the space <strong>of</strong> random fields satisfying u(ω,·) ∈ H 1 (D) almost sure<br />

and u(·,x) ∈ L 2 (Ω), where H 1 (D) is the usual Sobolev space and<br />

{<br />

∫<br />

}<br />

L 2 (Ω) = f : Ω → IR : f (ω) 2 dω < ∞ . (3.6)<br />

Ω<br />

This is a strong limitation which is typically not satisfied for random fields arising in financial problems<br />

[69, 108]. For image processing problems, this space is reasonable, because H 1 -regularity is<br />

typically assumed for classical image processing tasks [17] and L 2 -regularity <strong>of</strong> the stochastic part<br />

seems reasonable due to the limited energy an image acquisition device detects. Furthermore, the<br />

restriction to random fields with finite variance, i.e. satisfying (3.6), allows a discretization <strong>of</strong> the<br />

random fields <strong>using</strong> polynomial chaos expansions. Random fields will play a crucial role in the presentation<br />

<strong>of</strong> SPDEs, because the locally varying random coefficients <strong>of</strong> the SPDEs are random fields.<br />

3.2 <strong>Stochastic</strong> Partial Differential Equations<br />

We introduce SPDEs following [18] and use an elliptic model equation. The deterministic equation is<br />

−∇ · (a∇u) = f<br />

on D<br />

u = g on ∂D ,<br />

(3.7)<br />

where a is a diffusion coefficient, f a source term, g the boundary condition, and D the deterministic<br />

domain this equation holds in. In this equation, we assumed that we perfectly know the diffusion<br />

coefficient a and the right hand side f . In many applications, these quantities are not known exactly,<br />

but a description <strong>of</strong> the quantities through random fields is possible (see e.g. [8]). Let D be a bounded<br />

26


3.2 <strong>Stochastic</strong> Partial Differential Equations<br />

domain in IR d , (Ω,A ,Π) a complete probability space, and a : Ω× ¯D → IR a stochastic function with<br />

continuous and bounded covariance function that satisfies ∃a min ,a max ∈ (0,∞) with<br />

P(ω ∈ Ω : a(x,ω) ∈ [a min ,a max ],∀x ∈ ¯D) = 1 , (3.8)<br />

i.e. the diffusion coefficient is bounded away from zero and infinity for realizations ω ∈ Ω almost<br />

sure. In addition, let f : Ω × ¯D → IR be a stochastic function that satisfies<br />

∫ ∫<br />

( ∫ )<br />

f 2 (x,ω)dxdω = E f 2 (x,ω)dx < ∞ . (3.9)<br />

Ω D<br />

D<br />

Then the elliptic SPDE analog to 3.7 reads<br />

−∇ · (a(ω,·)∇u(ω,·)) = f (ω,·)<br />

almost sure on D<br />

u(·) = g(·) on ∂D .<br />

Applying this concept to other PDEs yields parabolic and hyperbolic SPDEs.<br />

3.2.1 Existence and Uniqueness <strong>of</strong> Solutions for Elliptic SPDEs<br />

(3.10)<br />

The pro<strong>of</strong> <strong>of</strong> the existence and uniqueness <strong>of</strong> solutions for elliptic SPDEs is closely related to the<br />

existence and uniqueness pro<strong>of</strong> <strong>of</strong> the classical problem. The Lax-Milgram theorem [37] is applicable<br />

in the stochastic context when we show continuity and coercivity <strong>of</strong> the related linear and<br />

bilinear forms. The main difficulty <strong>of</strong> the pro<strong>of</strong> is that the stochastic PDE requires the multiplication<br />

<strong>of</strong> stochastic quantities, because the expression a∇u has to be well-defined. For this, we introduce<br />

the Wick product [68, 155] and have to investigate conditions for its existence. Let us begin with<br />

notation for the definition <strong>of</strong> the Wick product. The presentation is based on [150]. In the following<br />

let {H α : α ∈ I}, where I is an index set, be an orthogonal basis <strong>of</strong> L 2 (Ω).<br />

Definition The Wick product <strong>of</strong> two random variables f ,g : Ω → IR is the formal series<br />

(<br />

f g = ∑ f α g β H α+β (ξ ) = ∑ ∑ f α g β<br />

)H γ (ξ ) , (3.11)<br />

α,β<br />

γ α+β=γ<br />

whereas a random variable is expressed in the orthogonal basis via f = ∑ α f α H α (ξ ). The H α depend<br />

on a vector ξ = (ξ 1 ,...) <strong>of</strong> basic random variables.<br />

The Wick product is not well-defined for all second order random variables, i.e. L 2 (Ω) is not closed<br />

under Wick multiplication (see [150]). Therefore, we introduce restrictions <strong>of</strong> the space L 2 (Ω) to<br />

ensure a well-defined Wick multiplication.<br />

Definition The Kondratiev-Hilbert spaces S ρ,k [85] are<br />

{<br />

}<br />

(S ) ρ,k := f = ∑ α<br />

f α H α : f α ∈ IR for α ∈ I and ‖ f ‖ ρ,k < ∞<br />

where −1 ≤ ρ ≤ 1 and k ∈ IR. We define the norm ‖ · ‖ ρ,k via the scalar product<br />

and the expression (2N) α via<br />

, (3.12)<br />

( f ,g) ρ,k := ∑ α<br />

f α g α (α!) 1+ρ (2N) αk (3.13)<br />

(2N) α :=<br />

∞<br />

∏<br />

i=1<br />

(2 d β (i)<br />

1 β (i)<br />

2<br />

...β<br />

(i)<br />

d )α i<br />

. (3.14)<br />

The product ∏ ∞ i=1 is the product over all possible multi-indices β. Kondratiev spaces are separable<br />

Hilbert spaces [150].<br />

27


Chapter 3 SPDEs and Polynomial Chaos Expansions<br />

Roughly speaking, a Kondratiev space S ρ,k is a subspace <strong>of</strong> L 2 (Ω), where the coefficients f α respect<br />

a decay condition such that (3.13) stays finite.<br />

The Kondratiev spaces from the previous definition are purely stochastic spaces. To add the spatial<br />

dependencies we have to make the coefficients f α functions that depend on a spatial variable and<br />

fulfill regularity assumptions.<br />

Definition The Hilbert space (S ) ρ,k,m (D) is<br />

{<br />

}<br />

(S ) ρ,k,m (D) := f (x) = ∑ α<br />

f α (x)H α : f α ∈ H m (D) ∀α ∈ I<br />

, (3.15)<br />

and the scalar product is defined in the same way as the scalar product <strong>of</strong> the space S ρ,k , where the<br />

scalar product <strong>of</strong> H m (D) ( f α ,g α ) H m replaces the expression f α g α .<br />

After the definition <strong>of</strong> the basic spaces for the Wick product, we define a Banach space such that the<br />

Wick product g → f g for every f from the Banach space is a continuous linear operator on (S ) −1,k,0<br />

(see [150], Proposition 4).<br />

Definition For D ⊂ IR d and l ∈ IR we define the space F l (D) via<br />

{<br />

F l (D) := f (x) = ∑ f α (x)H α : f α is measurable on D for every α and<br />

α<br />

(<br />

) } (3.16)<br />

‖ f ‖ l := esssup ∑ f α (x)|(2N) lα < ∞<br />

x∈D α<br />

and the space P l (D) via<br />

P l (D) := { f ∈ F l (D) : ∃A > 0 such that (E( f )g,g) 0,D ≥ A‖g‖ 2 0,D ∀g ∈ L 2 (D) } . (3.17)<br />

Using all the previous definitions it is possible to show existence and uniqueness <strong>of</strong> solutions <strong>of</strong><br />

SPDEs, because f ∈ F l ensures that the bilinear form is continuous and f ∈ P l the coerciveness <strong>of</strong><br />

the bilinear form. The existence and uniqueness result is originally by Vage [150]:<br />

Theorem 3.1. Let D ⊂ IR d be an open set <strong>of</strong> finite diameter and suppose a ∈ P l (D) for some l ∈ IR.<br />

Then there exists a constant K(a) ≤ 2l such that if k < K(a), (3.10) has a unique variational solution<br />

u ∈ (S ) −1,k,1 for every f ∈ (S ) −1,k,0 and g ∈ (S ) −1,k,1 .<br />

To sum up, we assure existence and uniqueness <strong>of</strong> solutions <strong>of</strong> SPDEs <strong>using</strong> the methods for PDEs,<br />

when the Wick product a∇u is well-defined and the SPDE fulfills (3.8) and (3.9).<br />

3.2.2 Parabolic SPDEs<br />

We construct parabolic SPDEs from elliptic SPDEs in the same way as for classical PDEs. We have<br />

to add time-dependence for the solution and incorporate an additional time derivative. We end up<br />

with a prototype for parabolic SPDEs given by<br />

u t (ω,x,t) − ∇ · (a(ω,x,t)∇u(ω,x,t)) = f (ω,x,t) almost sure on D × (0,T )<br />

u(x,t) = 0 on ∂D × (0,T )<br />

u(x,0) = u 0 on D × {0} .<br />

(3.18)<br />

Vage [150] proved existence and uniqueness <strong>of</strong> solutions for this kind <strong>of</strong> parabolic SPDEs. The<br />

findings are condensed in the following theorem (cf. Theorem 4 in [150]).<br />

Theorem 3.2. Let 0 < T < ∞, D ⊂ IR d be an open set <strong>of</strong> finite diameter, and a ∈ P l be given. Then<br />

there exists a constant K(a) ≤ 2l such that if ρ = −1 and k < K(a), (3.18) has a unique solution<br />

u ∈ W(0,T ) for any f ∈ L 2 (0,T ;((S ) −1,k,1<br />

0<br />

) ′ ) and u 0 ∈ (S ) −1,k,0 .<br />

28


3.3 Polynomial Chaos Expansions<br />

Figure 3.1: Relation between the stochastic spaces. We avoid the integration over Ω with respect to<br />

the measure Π. Instead, we transform the integral into integration over a subset <strong>of</strong> IR (the<br />

space Γ i ) with respect to the known PDF ρ <strong>of</strong> the basic random variables ξ i .<br />

3.2.3 Doob-Dynkin Lemma<br />

A famous result for the representation <strong>of</strong> the result <strong>of</strong> SPDEs is the Doob-Dynkin lemma. The<br />

version given here is cited from [132]:<br />

Lemma 1. Let (Ω,Σ) and (S,A ) be measurable spaces and f : Ω → S be a measurable function,<br />

i.e. f −1 (A ) ⊂ Σ. Then a function g : Ω → IR is measurable relative to the σ-algebra f −1 (A ) if and<br />

only if there is a measurable function h : S → IR such that g = h ◦ f .<br />

This lemma ensures that the solution <strong>of</strong> an SPDE is representable in the same random variables as<br />

the finite-dimensional input. This is due to the measurability <strong>of</strong> SPDEs when the coefficient is a<br />

linear combination <strong>of</strong> a finite number <strong>of</strong> random variables, because random variables are measurable<br />

by definition and the solution <strong>of</strong> an SPDE has to be continuous and thus is measurable. Furthermore,<br />

the product <strong>of</strong> a measurable function is measurable.<br />

Having the theory for existence and uniqueness <strong>of</strong> SPDE solutions at hand, we need a representation<br />

<strong>of</strong> stochastic quantities compatible with numerical schemes to compute approximations <strong>of</strong> the<br />

SPDE solutions. This approximation is based on the representation <strong>of</strong> random variables from (3.11).<br />

3.3 Polynomial Chaos Expansions<br />

The main contribution for the numerical treatment <strong>of</strong> SPDEs is the polynomial chaos expansion <strong>of</strong><br />

random variables. Based on the fundamental work <strong>of</strong> Wiener [156], who developed the polynomial<br />

chaos for Gaussian processes, leading to a basis formed by Hermite-polynomials, Cameron and<br />

Martin [27] proved that every random variable with a finite variance has a representation as Fourier-<br />

Hermite series. Later, Xiu and Karniadakis [160] developed the Wiener-Askey polynomial chaos<br />

or generalized polynomial chaos, which allows a representation <strong>of</strong> any random process with finite<br />

second-order moments in the polynomial chaos with an optimal basis.<br />

One main advantage <strong>of</strong> the representation <strong>of</strong> random variables in the polynomial chaos is the<br />

simplification <strong>of</strong> the calculation <strong>of</strong> integrals over the stochastic part. For arbitrary random variables<br />

with unknown probability density function, we have to calculate the integral over the abstract event<br />

space Ω. The use <strong>of</strong> the polynomial chaos expansions allows us to transform this integral into an<br />

integral over the real numbers by <strong>using</strong> the probability density function <strong>of</strong> the underlying random<br />

29


Chapter 3 SPDEs and Polynomial Chaos Expansions<br />

variables. Fig. 3.1 shows the situation. Instead <strong>of</strong> the direct computation <strong>of</strong> the integrals with the<br />

random variable X and the measure Π, we transform the integration into integration over the real<br />

numbers by <strong>using</strong> the polynomial chaos and the PDF ρ <strong>of</strong> the underlying random variables.<br />

3.3.1 Wiener Chaos<br />

In his seminal paper [156], Wiener developed the homogeneous (or Wiener) chaos formulated <strong>using</strong><br />

Hermite-polynomials in independent Gaussian random variables with zero mean and unit variance.<br />

Let ˜ξ = (ξ 1 ,...) be a vector <strong>of</strong> independent Gaussian random variables with zero mean, unit variance<br />

and PDFs ρ i , and V n (ξ i1 ,...,ξ in ) be Hermite-polynomials in n random variables. Cameron and<br />

Martin [27] proved that a random variable X with finite second-order moments has the representation<br />

X(ω) = a 0 V 0 +<br />

∞<br />

∑<br />

i 1 =1<br />

a i1 V 1 (ξ i1 (ω)) +<br />

∞ ∞<br />

∑ ∑<br />

i 1 =1 i 2 =1<br />

a i1 i 2<br />

V 2 (ξ i1 (ω),ξ i2 (ω)) + ... . (3.19)<br />

For notational convenience, this expression can be rewritten <strong>using</strong> multi-index notation<br />

X(ω) = ∑ ∞ α=1 a αΨ α ( ˜ξ (ω)) . (3.20)<br />

The functions V n and Ψ α have a one-to-one correspondence, i.e. every V n appears in the summation<br />

over j, but has a different index. In what follows, we do not denote the dependence <strong>of</strong> ξ on ω<br />

explicitly to ease notation when no integration over the stochastic space Ω is involved.<br />

The Hermite-polynomials Ψ α form an orthogonal basis <strong>of</strong> the space L 2 (Ω), i.e.<br />

∫<br />

Ω<br />

Ψ α ( ˜ξ (ω)<br />

)<br />

Ψ β ( ˜ξ (ω)<br />

)dω = 〈Ψ α ,Ψ β 〉 = 〈(Ψ α ) 2 〉δ αβ . (3.21)<br />

For a finite number <strong>of</strong> basic random variables ξ = (ξ 1 ,...,ξ n ) we simplify (3.21) by <strong>using</strong> (3.1). The<br />

scalar product 〈 f ,g〉 is<br />

∫<br />

∫<br />

〈Ψ α (ξ ),Ψ β (ξ )〉 = Ψ α (ξ (ω))Ψ β (ξ (ω))dω = Ψ α (x)Ψ β (x)dΠ, (3.22)<br />

Ω<br />

Γ<br />

where Γ = supp(ξ ) ⊂ IR n . It follows from (3.22) that the weighting function w that is needed to get<br />

orthonormal polynomials is<br />

1<br />

w(x) = √ . (3.23)<br />

(2π) n e − 1 2 xT x<br />

This weighting function is the key to understand the good approximation quality <strong>of</strong> the Hermiteexpansion,<br />

because the weighting function for the Hermite-polynomials is the same as the PDF <strong>of</strong><br />

an n-dimensional Gaussian random variable, i.e. w(x) = ∏ i ρ i = ρ(x). Xiu and Karniadakis [160]<br />

investigated this correspondence between the weighting functions for the orthogonal polynomial<br />

basis and the density functions <strong>of</strong> random variables. Section 3.3.3 summarizes the findings. Thus,<br />

the computation <strong>of</strong> the scalar product reduces to integration over a subset <strong>of</strong> IR n . For this, we use a<br />

quadrature rule. Since we are integrating polynomials, the usage <strong>of</strong> a suitable quadrature rule leads<br />

to exact results up to numerical inaccuracies.<br />

3.3.2 Cameron-Martin Theorem<br />

The Wiener chaos is an abstract representation for random variables, but it is unclear whether it converges<br />

to the desired random variable. The Cameron-Martin theorem [27] fills this gap <strong>of</strong> knowledge.<br />

We present the theorem in the version proposed in [27], but with the notation used in this thesis.<br />

30


3.3 Polynomial Chaos Expansions<br />

Theorem 3.3. The Wiener chaos representation <strong>of</strong> any random variable X ∈ L 2 (Ω) converges in the<br />

L 2 (Ω)-sense to X. This means, if X is any functional for which<br />

∫<br />

|X(ω)| 2 dω < ∞ , (3.24)<br />

Ω<br />

then<br />

∫<br />

lim |X(ω) −∑ N N→∞<br />

α=1 a αΨ α (ξ (ω))| 2 dω = 0 . (3.25)<br />

Ω<br />

The Fourier-Hermite coefficient a α is<br />

∫<br />

a α = X(ω)Ψ α (ξ (ω))dω . (3.26)<br />

Ω<br />

The Cameron-Martin theorem ensures that every random variable with finite variance has a representation<br />

in the Wiener chaos, but gives no information about the convergence rate <strong>of</strong> the representation.<br />

The convergence rate is important when the series expansion is cut after a finite number <strong>of</strong> terms.<br />

This is necessary for numerical algorithms dealing with polynomial chaos expansions. In fact, [160]<br />

showed that the convergence rate <strong>of</strong> the Wiener chaos is substantially less the optimal, exponential,<br />

convergence rate. The development <strong>of</strong> other chaos types leads to expansions that have better convergence<br />

properties. This is the topic <strong>of</strong> the next section, which introduces the generalized polynomial<br />

chaos expansion, originally proposed by Xiu and Karniadakis [160].<br />

3.3.3 Generalized Polynomial Chaos<br />

Xiu and Karniadakis [160] generalized the idea <strong>of</strong> the representation <strong>of</strong> random variables in an orthogonal<br />

basis formed by polynomials in random variables with known distribution. They proposed<br />

to use polynomials whose weighting functions correspond to the PDF <strong>of</strong> the underlying random variables.<br />

It turns out that these polynomials are the polynomials from the Askey-scheme [16]. Table 3.2<br />

shows the correspondence between important random variables and the associated polynomials. To<br />

summarize, a random variable with finite variance has a representation in the polynomial chaos by<br />

X(ω) = ∑ ∞ α=1 a αΨ α (ξ ) , (3.27)<br />

where the multi-dimensional polynomials are selected from the Askey-scheme [16]. The multidimensional<br />

polynomials are constructed from one-dimensional polynomials via<br />

ψ α = ∏ n i=1 H α i<br />

(ξ i ) , (3.28)<br />

whereas α is the index corresponding to the multiindex (α 1 ,...,α n ) and H αi , i = 1,...,n are polynomials<br />

in one random variable. Fig. 3.1 shows the first one-dimensional polynomials for the Legendrechaos<br />

and Fig. 3.3 the polynomials for the Hermite-chaos. We rescaled the Legendre- and Hermitepolynomials<br />

to get an orthonormal basis <strong>of</strong> L 2 (Ω) with respect to the weighted scalar product, i.e.<br />

〈Ψ α ,Ψ β 〉 = δ αβ , (3.29)<br />

because the weighting functions for the random variables are 0.5 and 1 √<br />

2π<br />

exp −x2<br />

2 , respectively.<br />

Ernst et al. [50] proved that the polynomial chaos expansion converges in quadratic mean, i.e. in<br />

the L 2 (Ω) sense [73], if and only if the basic random variables have finite moments <strong>of</strong> all orders and<br />

the probability density <strong>of</strong> the basic random variables is continuous. Furthermore, the moment problem<br />

(cf. [50]), i.e. the identification <strong>of</strong> the measure from the moments, has to be uniquely solvable.<br />

Nouy [116] showed that multimodal random variables are hard to approximate in a onedimensional<br />

polynomial chaos expansion. He solved this problem by introducing a special kind<br />

31


Chapter 3 SPDEs and Polynomial Chaos Expansions<br />

H 1 (x) = 1<br />

H 2 (x) = √ 3x<br />

H 3 (x) = √ 5 · (1.5 ∗ x 2 − 0.5)<br />

H 4 (x) = √ 7 · (2.5x 3 − 1.5 ∗ x)<br />

H 5 (x) = √ 9 · 1<br />

8 (35x4 − 30x 2 + 3.0)<br />

H 6 (x) = √ 11 · 1<br />

8 (63x5 − 70x 3 + 15x)<br />

H 7 (x) = √ 13 · 1<br />

16 (231x6 − 315x 4 + 105x − 5)<br />

H 8 (x) = √ 15 · 1<br />

16 (429x7 − 693x 5 + 315x 3 − 35x)<br />

H 9 (x) = √ 1<br />

17 ·<br />

128 (6435x8 − 12012x 6 + 6930x 4 − 1260x 2 + 35)<br />

H 10 (x) = √ 1<br />

19 ·<br />

128 (12155x9 − 25740x 7 + 18018x 5 − 4620x 3 + 315x)<br />

Table 3.1: The first ten one-dimensional Legendre-polynomials. The multi-dimensional polynomials<br />

up to degree nine are based on these polynomials and (3.40).<br />

<strong>of</strong> polynomial chaos expansion. In this expansion, one random variable acts as indicator function for<br />

the modes <strong>of</strong> the approximated random variable. Then the multimodal random variable is approximated<br />

on all modes independently. Wan and Karniadakis [152] introduced a similar approach called<br />

multi-element polynomial chaos (MEPC). The idea <strong>of</strong> this method is to decompose the stochastic<br />

space into smaller elements. Since we are approximating L 2 -functions in the stochastic space, we<br />

need no coupling condition between the stochastic elements, i.e. the solutions may have jumps across<br />

the elements. This allows for an efficient parallelization <strong>of</strong> the MEPC, because it is possible to perform<br />

the computations for elements in the stochastic space on different machines and there is no<br />

need for communication between the machines.<br />

To use the polynomial chaos expansion in numerical schemes makes it necessary to cut <strong>of</strong> the<br />

series expansion after a finite number <strong>of</strong> terms. This is done by choosing the number <strong>of</strong> random<br />

variables used for the approximation, denoted by n and by prescribing the maximal polynomial<br />

degree p in the expansion. As usual for a polynomial basis, the number <strong>of</strong> terms in the expansion is<br />

( ) n + p<br />

N = . (3.30)<br />

p<br />

Random variable Wiener-Askey chaos Support<br />

Gaussian Hermite-Polynomials (−∞,∞)<br />

Gamma Laguerre-Polynomials [0,∞)<br />

Beta Jacobi-Polynomials [a,b]<br />

Uniform Legendre-Polynomials [a,b]<br />

Poisson Charlier-Polynomials discrete<br />

Binomial Krawtchouk-Polynomials discrete<br />

Table 3.2: Important distributions and the corresponding polynomials for the expansion.<br />

32


3.3 Polynomial Chaos Expansions<br />

H 1 (x) = 1<br />

H 2 (x) = x<br />

H 3 (x) = 1 √<br />

2!<br />

(x 2 − 1)<br />

H 4 (x) = 1 √<br />

3!<br />

(x 3 − 3x)<br />

H 5 (x) = 1 √<br />

4!<br />

(x 4 − 6x 2 + 3)<br />

H 6 (x) = 1 √<br />

5!<br />

(x 5 − 10x 3 + 15x)<br />

H 7 (x) = 1 √<br />

6!<br />

(x 6 − 15x 4 + 45x 2 − 15)<br />

H 8 (x) = 1 √<br />

7!<br />

(x 7 − 21x 5 + 105x 3 − 105x)<br />

H 9 (x) = 1 √<br />

8!<br />

(x 8 − 28x 6 + 210x 4 − 420x 2 + 105)<br />

H 10 (x) = 1 √<br />

9!<br />

(x 9 − 36x 7 + 378x 5 − 1260x 3 + 945x)<br />

Table 3.3: The first ten one-dimensional Hermite-polynomials. The construction <strong>of</strong> the multidimensional<br />

polynomials up to degree 9 is based on these polynomials and (3.40).<br />

Thus, it is necessary to select a vector <strong>of</strong> random variables ξ = (ξ 1 ,...,ξ n ) and the polynomial<br />

degree p. Then, an approximation <strong>of</strong> a random variable in the polynomial chaos is<br />

X(ω) ≈ ∑ N α=1 a αΨ α (ξ ) . (3.31)<br />

The remaining part <strong>of</strong> this chapter deals with numerical methods for polynomial chaos expansions.<br />

Although the presented material is valid for all polynomials from the Askey-scheme, the numerical<br />

implementation is based on the Legendre-polynomials and uniform distributed random variables,<br />

because the support <strong>of</strong> the Legendre-polynomials is compact. This is advantageous for algorithms,<br />

especially when dealing with stochastic level sets. Chapter 4 discusses the combination <strong>of</strong> polynomial<br />

chaos expansions and SPDEs. There the information presented for the polynomial chaos is<br />

combined with finite element and finite difference schemes for the discretization <strong>of</strong> the equations.<br />

3.3.4 Calculations in the Polynomial Chaos<br />

To use the polynomial chaos in numerical schemes it is necessary to perform arithmetic operations<br />

in the polynomial chaos. In this section, we review the development <strong>of</strong> the basic operations like<br />

addition, subtraction, multiplication, division and the calculation <strong>of</strong> square roots. The presentation<br />

is based on the work <strong>of</strong> Debusschere et al. [38]. For the remaining part <strong>of</strong> this section let<br />

a = ∑ N α=1 a αΨ α (ξ ), b = ∑ N α=1 b αΨ α (ξ ), c = ∑ N α=1 c αΨ α (ξ ) (3.32)<br />

be three polynomial chaos variables. We compute the sum and the difference <strong>of</strong> quantities in the<br />

polynomial chaos by adding or subtracting the corresponding chaos coefficients, because the addition<br />

or subtraction <strong>of</strong> polynomials results in a polynomial with the same degree at most:<br />

c = a ± b = ∑ N α=1 a αΨ α (ξ ) ±∑ N α=1 b αΨ α (ξ ) = ∑ N α=1 (a α ± b α )Ψ α (ξ ) . (3.33)<br />

33


Chapter 3 SPDEs and Polynomial Chaos Expansions<br />

The multiplication <strong>of</strong> two polynomial chaos variables is more difficult. Since polynomials form<br />

the basis, the naive multiplication <strong>of</strong> polynomial chaos variables results in a polynomial with twice<br />

the degree <strong>of</strong> the factors. Thus, an additional projection step onto a polynomial with the same degree<br />

as the factors <strong>of</strong> the multiplication is necessary. This projection step is done by <strong>using</strong> the Galerkin<br />

or L 2 -projection, leading to a projection polynomial, whose error is orthogonal to the space spanned<br />

by the polynomial chaos. The idea <strong>of</strong> the projection is to multiply the naive product c = a · b with an<br />

element Ψ γ <strong>of</strong> the polynomial chaos basis, integrate over the stochastic dimensions and to compute<br />

the coefficient <strong>of</strong> the multiplication one after another from this expression:<br />

∫<br />

Γ<br />

N<br />

∑<br />

α=1<br />

∫<br />

c α Ψ α Ψ γ dΠ =<br />

Γ<br />

N<br />

∑<br />

α=1<br />

N<br />

∑<br />

β=1<br />

a α b β Ψ α Ψ β Ψ γ dΠ ⇒ c γ =<br />

N<br />

∑<br />

α=1<br />

N<br />

∑<br />

β=1<br />

〈Ψ α Ψ β Ψ γ 〉<br />

a α b β<br />

〈(Ψ γ ) 2 . (3.34)<br />

〉<br />

} {{ }<br />

C αβγ<br />

Note that we omit denoting the dependence <strong>of</strong> Ψ α from ξ and ω to simplify the notation. The<br />

quantity C αβγ is independent <strong>of</strong> the actual problem, it depends on the basis only. The values <strong>of</strong> C αβγ<br />

can be precomputed in a lookup table. The next section describes the generation <strong>of</strong> this table.<br />

The computation <strong>of</strong> the quotient <strong>of</strong> two random variables, a = c b<br />

is possible, too. To do this, we<br />

multiply the expression by b, yielding c = ab and use again the Galerkin projection for this equation:<br />

c γ = ∑ N α=1∑ N β=1 C αβγb β a α = ∑ N α=0 A γαa α . (3.35)<br />

This is a system <strong>of</strong> linear equations for the coefficients a α , which we solve by an iterative solver.<br />

In a similar manner, we compute the square root b = √ a <strong>of</strong> a polynomial chaos variable. First, we<br />

rewrite the equation in the form a = b 2 and then use the Galerkin projection to obtain<br />

a γ = ∑ N α=1∑ N β=1 C αβγb α b β . (3.36)<br />

This is a nonlinear system <strong>of</strong> equations for the unknown coefficients b α , which we solve <strong>using</strong><br />

Newton’s method to find a root <strong>of</strong><br />

f (b) = b 2 − a . (3.37)<br />

The partial derivatives <strong>of</strong> this function are<br />

∂ f α (b)<br />

=<br />

∂b β<br />

N<br />

∑<br />

γ=1<br />

C βγα b γ . (3.38)<br />

As pointed out by Matthies and Rosic [98], it is possible to use a mild convergence criterion for<br />

Newton’s method depending on the expected value and the variance <strong>of</strong> the polynomial chaos variable.<br />

Using these building blocks, it is possible to construct numerical methods for nearly all possible<br />

calculations, e.g. the exponential <strong>of</strong> a random variable in the polynomial chaos is<br />

exp(a) = exp(a 1 )<br />

(<br />

1 +<br />

K<br />

∑<br />

n=1<br />

(<br />

∑<br />

N<br />

α=2 a α Ψ α) )<br />

n<br />

n!<br />

. (3.39)<br />

With the methods from this section, it is also possible to construct finite difference schemes for<br />

random variables. Chapter 4 investigates this further.<br />

3.3.5 The <strong>Stochastic</strong> Lookup Table<br />

We precompute the values <strong>of</strong> C αβγ in a lookup table to speed up the calculations in the polynomial<br />

chaos. It is possible to replace the calculation <strong>of</strong> the multi-dimensional integrals ∫ Γ Ψα Ψ β Ψ γ dΠ<br />

34


3.4 Relation to Interval Arithmetic<br />

Figure 3.2: Sparsity structure <strong>of</strong> the stochastic lookup table for n = 5 random variables and a polynomial<br />

degree p = 3. The gray dots indicate positions in the three-dimensional lookup<br />

table C αβγ that contain nonzero entries.<br />

by one-dimensional integration, because the basis functions are Ψ α = ∏ n j=1 H α j<br />

(ξ j ) whereas α corresponds<br />

to the multi-index (α 1 ,...,α n ) and H αi are polynomials in one random variable. Using<br />

the product representation <strong>of</strong> the polynomials, we simplify the equation by <strong>using</strong> that the random<br />

variables ξ i are statistically independent, i.e. E(ξ i ξ j ) = E(ξ i )E(ξ j ):<br />

)<br />

∫<br />

〈Ψ α Ψ β Ψ γ 〉 =<br />

dΠ<br />

=<br />

Γ<br />

n<br />

∏<br />

m=1<br />

(<br />

n<br />

∏<br />

m=1<br />

∫<br />

H (i) α m<br />

Γ m<br />

n<br />

H (i) α<br />

(ξ m ))(<br />

∏<br />

m<br />

m=1<br />

(ξ m )H β<br />

( j)<br />

m<br />

n<br />

H ( j) β<br />

(ξ m ))(<br />

∏<br />

m<br />

m=1<br />

(ξ m )H (k) γ<br />

(ξ m )dΠ m .<br />

m<br />

H (k) γ<br />

(ξ m )<br />

m<br />

(3.40)<br />

In (3.40) Π m = ρ m Π m ,i = 1,...n denotes integration with respect to the probability measures <strong>of</strong> the<br />

random variables ξ m ,m = 1,...n.<br />

3.4 Relation to Interval Arithmetic<br />

Interval arithmetic [64,78,102,104] is a possibility for reliable computations on a computer. Instead<br />

<strong>of</strong> <strong>using</strong> a single fixed number, this concept is based on intervals <strong>of</strong> numbers to provide an upper and<br />

a lower bound for the computation result. The result is considered to be uniformly distributed inside<br />

this interval. Arithmetic operations for these reliability intervals are defined via the lower and upper<br />

bounds <strong>of</strong> the intervals. Let x = [x, ¯x],y = [y,ȳ] be two intervals and ◦ one <strong>of</strong> the operations +,−,×,/.<br />

Then the resulting interval is defined as<br />

[x, ¯x] ◦ [y,ȳ] = [ min ( x ◦ y,x ◦ ȳ, ¯x ◦ y, ¯x ◦ ȳ ) ,max ( x ◦ y,x ◦ ȳ, ¯x ◦ y, ¯x ◦ ȳ )] . (3.41)<br />

The definition <strong>of</strong> the new interval bounds based on the old interval bounds is useful when dealing<br />

with monotonic functions only, e.g. computing the sine function <strong>of</strong> an interval fails, because<br />

35


Chapter 3 SPDEs and Polynomial Chaos Expansions<br />

Figure 3.3: PDFs <strong>of</strong> initial uniformly distributed input intervals (gray) and the PDFs <strong>of</strong> the results <strong>of</strong><br />

the polynomial chaos computation (black) for squaring an interval (left) and dividing an<br />

interval by itself (right).<br />

sin([30 ◦ ,150 ◦ ]) = [0.5,0.5] in the above arithmetic. Other problems are that there are no simple<br />

methods to link realizations <strong>of</strong> intervals, e.g. the naive multiplication yields [−2,2] 2 = [−4,4], because<br />

there is no information in the interval arithmetic calculus that the realizations <strong>of</strong> both intervals<br />

must be the same. This is also problematic for division, because dividing an interval by itself should<br />

result into an interval with zero width, but, e.g. 2x/x for x = [2,4] yields the interval [1,4], not the<br />

desired interval [2,2]. Furthermore, the result is forced to be uniformly distributed inside the resulting<br />

interval, which is not the case for nonlinear operations and the resulting intervals can become<br />

arbitrarily large. Nevertheless, interval arithmetic is used in applications [70].<br />

The polynomial chaos expansion can be thought as an extension <strong>of</strong> the interval arithmetic calculus.<br />

The results <strong>of</strong> polynomial chaos calculations are not forced to be uniformly distributed. Instead, they<br />

can have every distribution that can be represented in the chosen polynomial chaos basis. Results<br />

that can not be represented in this basis are projected onto the basis <strong>using</strong> the Galerkin projection<br />

introduced earlier. Fig. 3.3 shows the polynomial chaos result <strong>of</strong> the problematic operations squaring<br />

an interval and dividing by an interval. In both cases the polynomial chaos expansions yields the<br />

exact result, up to the machine precision. Furthermore, the problem <strong>of</strong> the huge resulting intervals<br />

<strong>of</strong> interval arithmetic operations is solved by <strong>using</strong> polynomial chaos expansions, because events at<br />

the tails <strong>of</strong> the intervals have a very low probability.<br />

Conclusion<br />

This chapter provided us with a possibility for the finite-dimensional approximation <strong>of</strong> arbitrary<br />

random variables, the polynomial chaos expansion. Even if the representation is possible for second<br />

order random variables only, this is sufficient for random variables arising in the image processing<br />

context. The finite-dimensional approximations and the associated closed computations provide a<br />

powerful toolbox for the discretization <strong>of</strong> SPDEs and a stochastic modeling <strong>of</strong> image processing<br />

problems. With the presented theoretical background, it is also possible to prove existence and<br />

uniqueness <strong>of</strong> solutions for the stochastic image processing models in the preceding chapters. It<br />

remains to show the few, easy to verify, assumptions.<br />

36


Chapter 4<br />

Discretization <strong>of</strong> SPDEs<br />

The discretization <strong>of</strong> SPDEs is an active research field [47, 159]. Besides the discretization based<br />

on sampling approaches like Monte Carlo simulation or stochastic collocation, there are methods for<br />

the intrusive computation <strong>of</strong> stochastic solutions in the literature. Intrusive means that we do not<br />

generate the solutions based on sampling strategies and we cannot reuse deterministic algorithms for<br />

the solution at the sampling points. This makes it necessary to develop new algorithms, but has the<br />

advantage that these algorithms are more efficient than the classical sampling approaches. This thesis<br />

focuses on intrusive methods. We present the sampling based approaches as well, but use them to<br />

verify the correctness <strong>of</strong> the intrusive algorithms and implementations only. The intrusive methods<br />

presented in this thesis range from the stochastic finite difference method based on polynomial chaos<br />

expansions to the generalized spectral decomposition [113–115, 117, 118], a method allowing to<br />

speed up the solution process <strong>of</strong> the stochastic finite element method (SFEM) [54].<br />

4.1 Sampling Based Discretization <strong>of</strong> SPDEs<br />

Since the mid <strong>of</strong> the 20 th century [100, 101] authors developed sampling based algorithms for the<br />

simulation <strong>of</strong> stochastic processes, starting with the development <strong>of</strong> the Monte Carlo method. Later,<br />

advanced sampling-based methods like the stochastic collocation method and improvements <strong>of</strong> these<br />

methods e.g. the combination <strong>of</strong> stochastic collocation and polynomial chaos expansions or the use<br />

<strong>of</strong> sparse grids based on a Smolyak construction [140] have been developed.<br />

4.1.1 Monte Carlo Simulation<br />

Monte Carlo simulation is the simplest technique for the discretization <strong>of</strong> random variables and<br />

SPDEs. A set <strong>of</strong> samples is generated randomly from the known distribution <strong>of</strong> the random variables<br />

via a pseudo random number generator like [97]. We can use the well-known deterministic<br />

algorithms on these samples and compute from the results approximations to stochastic quantities<br />

like expected value, variance, etc. <strong>using</strong> well-known formulas. For example, when we computed the<br />

solution <strong>of</strong> R samples, approximate expected value and variance are<br />

E(x) ≈ ¯x = 1 R ∑R i=1 x i and Var(x) ≈ 1<br />

R − 1 ∑R i=1 (x i − ¯x) 2 . (4.1)<br />

The main drawback <strong>of</strong> the Monte Carlo method is the slow convergence. In fact, Kendall [79] showed<br />

for the Monte Carlo method that the convergence <strong>of</strong> the samples mean towards the expected value<br />

is <strong>of</strong> order O(σ/ √ R). Despite the slow convergence rate, Monte Carlo methods are widely used (see<br />

e.g. [88, 94, 133]) due to the simple implementation and the possibility to reuse deterministic code.<br />

4.1.2 <strong>Stochastic</strong> Collocation<br />

During the last years, a variety <strong>of</strong> stochastic collocation (SC) techniques was developed. These<br />

techniques range from simple collocation techniques over sparse grid techniques to SC techniques<br />

allowing to obtain polynomial chaos coefficient (see [159] for a more detailed review).<br />

37


Chapter 4 Discretization <strong>of</strong> SPDEs<br />

Figure 4.1: Comparison between a sparse grid (left) constructed via Smolyak’s algorithm and a full<br />

tensor grid (right). The sparse grid contains significantly less nodes than the full tensor<br />

grid whose number <strong>of</strong> nodes growth exponentially with the dimension, but has nearly the<br />

same approximation order.<br />

SC is a non-intrusive approach for the discretization <strong>of</strong> SPDEs. In the simplest form, SC uses<br />

known points in the stochastic dimensions and performs runs <strong>of</strong> the deterministic problem for these<br />

points. The points are chosen following a quadrature rule, e.g. Gauss quadrature or Clenshaw-Curtis<br />

quadrature [33] where the points are selected based on the roots <strong>of</strong> the Chebyshev-polynomials [151].<br />

We construct higher-dimensional SC from the one-dimensional SC <strong>using</strong> tensor grids. This simple<br />

approach leads to a “curse <strong>of</strong> dimension” [119] when <strong>using</strong> the full tensor grids in higher dimension.<br />

To overcome this, we use Smolyak’s algorithm [122, 140], resulting in a sparse grid containing<br />

significant less nodes than the full tensor grid (see Fig. 4.1), but resulting in an approximation with<br />

nearly the same approximation order. The orders differ only by a logarithmic term, see [52].<br />

Following [158] it is possible to obtain a representation in the polynomial chaos from SC calculations.<br />

Having the usual polynomial chaos expansion (cf. Section 3.3)<br />

u(x,ω) = ∑ N α=0 a α(x)Ψ α (ξ (ω)) (4.2)<br />

in mind, we get the coefficients <strong>of</strong> the polynomial chaos from the collocation samples via a projection<br />

on the polynomial chaos<br />

a α (x) = ∑ Q j=1 u(x,y( j) )Ψ α (y ( j) )w j , (4.3)<br />

where y ( j) are the collocation points, w j the corresponding quadrature weights and Q the total number<br />

<strong>of</strong> collocation samples. This collocation approach allows an easy comparison <strong>of</strong> results obtained<br />

via SC and results from intrusive techniques presented in the following paragraphs. Besides the<br />

calculation <strong>of</strong> a polynomial chaos representation, the usual usage <strong>of</strong> SC is the computation <strong>of</strong> a<br />

Lagrange interpolation <strong>of</strong> the solution, i.e. to compute a representation like<br />

where L j are the Lagrange-polynomials.<br />

I u(x,ω) = ∑ Q j=1 u(x,y(i) )L j (ω) , (4.4)<br />

4.2 <strong>Stochastic</strong> Finite Difference Methods<br />

The sampling based approaches <strong>of</strong> the last section have the great advantage that calculations on<br />

the samples use classical methods for the solution <strong>of</strong> PDEs like finite element or finite difference<br />

38


4.3 <strong>Stochastic</strong> Finite Elements<br />

methods. In the following, we present an approach, where we discretize the SPDE directly. To make<br />

the approach more illustrative we demonstrate the method by <strong>using</strong> a parabolic SPDE<br />

∂ t u(t,x,ω) − u xx (t,x,ω) = f (t,x,ω) . (4.5)<br />

The temporal and spatial derivatives are determined <strong>using</strong> well-known approximations. Using the<br />

explicit Euler scheme for the discretization <strong>of</strong> the time derivative, we get<br />

u(t + τ,x,ω) = u(t,x,ω) + τ(u xx (t,x,ω) + f (t,x,ω)) . (4.6)<br />

Discretizing the spatial derivative <strong>using</strong> central differences, the fully discrete equation is<br />

( )<br />

u(t,x + h,ω) − 2u(t,x,ω) + u(t,x − h,ω)<br />

u(t + τ,x,ω) = u(t,x,ω) + τ<br />

+ f (t,x,ω)<br />

h 2<br />

. (4.7)<br />

The stochastic quantities in this equation are approximated by <strong>using</strong> a truncated polynomial chaos<br />

expansion leading to a numerical scheme that needs methods for the addition and multiplication <strong>of</strong><br />

polynomial chaos expansions. Section 3.3 presents numerical methods for this task.<br />

The main drawback <strong>of</strong> these methods is that computations in the polynomial chaos require the<br />

solution <strong>of</strong> linear systems <strong>of</strong> equations. Furthermore, the construction <strong>of</strong> unstructured or adaptive<br />

grids is complicated in comparison to the generation <strong>of</strong> adaptive grids for finite elements.<br />

The advantage <strong>of</strong> stochastic finite difference methods is the simple possibility to parallelize explicit<br />

stochastic finite difference schemes, because the computations on different nodes are independent.<br />

4.3 <strong>Stochastic</strong> Finite Elements<br />

It is well-known that the variational formulation <strong>of</strong> a deterministic PDE is<br />

find u ∈ V such that a(u,v) = b(v) ∀v ∈ V , (4.8)<br />

where a(·,·) is a bilinear form related to the PDE and b(·) a linear form related to the right hand side<br />

<strong>of</strong> the PDE. The space V is the space <strong>of</strong> all admissible functions, e.g. the Sobolev space H0 1 for the<br />

simple prototype equation −∇ · (k∇φ) = f in D, φ = 0 on ∂D.<br />

For stochastic coefficients, right hand sides or boundary conditions, the bilinear and/or linear<br />

form become stochastic quantities. Denote by a(u,v,ω), b(v,ω) the dependence <strong>of</strong> the forms on the<br />

stochastic event ω ∈ Ω. The aim <strong>of</strong> the stochastic problem is to find a random field, i.e. an element <strong>of</strong><br />

the tensor product space V ⊗S , u ∈ V ⊗S , where S is the space <strong>of</strong> random functions, e.g. L 2 (Ω),<br />

the space <strong>of</strong> all random variables with finite second order moments. The weak formulation <strong>of</strong> the<br />

stochastic problem is:<br />

find u ∈ V ⊗ S such that A(u,v) = B(v) ∀v ∈ V ⊗ S , (4.9)<br />

where<br />

∫<br />

A(u,v) = a(u,v,ω)dω = E(a(u,v,ω)) (4.10)<br />

Ω<br />

and<br />

∫<br />

B(v) = b(v,ω)dω = E(b(v,ω)) . (4.11)<br />

Ω<br />

The weak formulation <strong>of</strong> an SPDE is simply the expectation <strong>of</strong> the deterministic problem (4.8).. To<br />

ensure existence and uniqueness <strong>of</strong> a solution, we need the form A to be continuous and coercive and<br />

the form B to be continuous on the space V ⊗S . Hence, coercivity and continuity are ensured if the<br />

forms a and b are coercive, respectively continuous, for elementary events ω ∈ Ω almost sure and<br />

such that the Wick product is well-defined (cf. Section 3.2.1).<br />

39


Chapter 4 Discretization <strong>of</strong> SPDEs<br />

4.3.1 Discretization <strong>of</strong> the Spaces V and S<br />

We approximate the deterministic space V <strong>using</strong> the classical finite element approach. That means<br />

every u ∈ V ⊗ S is approximated by<br />

u(x,ω) ≈ ∑ n i=1 u i(ω)P i (x) , (4.12)<br />

where u i ∈ S and {P i } i=1,...,n is a basis <strong>of</strong> a finite dimensional subspace V h ⊂ V . We identify the<br />

space V h with IR n , because we have to store the coefficients for the basis elements only.<br />

We approximate the stochastic space S in two steps. First, we choose a finite set <strong>of</strong> random<br />

variables ζ =(ζ 1 ,...,ζ m ), span(ζ 1 ,...,ζ m )=S m ⊂S with finite variance and approximate u i ∈ S by<br />

u i (ω) ≈ ∑ m k=1 uk i ζ k (ω) , (4.13)<br />

where the coefficients u k i are deterministic coefficients for the random variables ζ k . Numerical calculations<br />

cannot use the space S m . Hence, we approximate the space S m by <strong>using</strong> the generalized<br />

polynomial chaos [160], cf. Section 3.3. We approximate the random variables ζ i with unknown<br />

distribution in the polynomial chaos by the same number <strong>of</strong> random variables and a prescribed polynomial<br />

degree p:<br />

u ∈ S m,p ⊂ S p : u = ∑ N i=1 u iΨ i (ξ ) . (4.14)<br />

The dimension <strong>of</strong> the space S m,p is N = ( )<br />

m+p<br />

m .<br />

For the finite dimensional subspace IR n ⊗ S p , the problem (4.9) is rewritten as<br />

E(v T Au) = E(v T b) ∀v ∈ IR n ⊗ S p , (4.15)<br />

where A ∈ Sp<br />

n×n is a stochastic matrix.<br />

Using the polynomial chaos basis, i.e. the space S m,p for the stochastic space and V h for the<br />

deterministic space we end up with a huge deterministic equation system to approximate the solution<br />

<strong>of</strong> ∇ · (a∇u) = f given by<br />

∑ N α=1<br />

where the matrices M α,β and L α,β are<br />

(M α,β ) i, j = E (Ψ α Ψ β ) ∫<br />

(<br />

L α,β ) i, j = ∑<br />

k<br />

(<br />

L α,β ) U α = ∑ N α=1 Mα,β F α , (4.16)<br />

(<br />

∑E<br />

γ<br />

D<br />

P i P j dx<br />

Ψ α Ψ β Ψ γ) a k γ<br />

∫<br />

D<br />

∇P i · ∇P j P k dx .<br />

(4.17)<br />

In (4.16) we used the notation F α = ( f i α) for the polynomial chaos representation <strong>of</strong> the quantities.<br />

4.4 Generalized Spectral Decomposition<br />

Selecting suitable subspaces <strong>of</strong> S m,p ⊗IR n and a special basis, which captures the dominant stochastic<br />

effects, we achieve a significant speed-up <strong>of</strong> the solution process and an enormous reduction <strong>of</strong><br />

the memory requirements. In the generalized spectral decomposition (GSD) [113], we approximate<br />

the solution u ∈ L 2 (Ω) ⊗ H 1 (D) by<br />

u(x,ξ ) ≈ ∑ K j=1 λ j(ξ )V j (x) , (4.18)<br />

where V j is a deterministic function, λ j a stochastic function and K the number <strong>of</strong> modes <strong>of</strong> the<br />

decomposition. Thus, the GSD computes a solution where the deterministic and the stochastic basis<br />

40


4.4 Generalized Spectral Decomposition<br />

functions are not fixed a priori. With the flexible basis functions we find a solution having significant<br />

fewer modes, i.e. K ≪ N, but nearly the same approximation quality.<br />

Nouy [113] showed how to compute the modes <strong>of</strong> an optimal approximation in the energy norm<br />

‖v‖ 2 A = E(vT Av) <strong>of</strong> the problem, i.e. such that<br />

∥<br />

∥<br />

∥u −∑ K ∥∥<br />

j=1 λ 2<br />

∥<br />

jU j = min ∥<br />

∥u −∑ K ∥∥<br />

A γ,V<br />

j=1 γ 2<br />

jV j . (4.19)<br />

A<br />

The next sections provide details about the GSD method, pro<strong>of</strong>s for the optimality <strong>of</strong> the approximation<br />

and implementation details. Further details about the GSD method can be found in [113].<br />

4.4.1 Best Approximation<br />

For deterministic linear systems <strong>of</strong> equations, it is possible to formulate an associated minimization<br />

problem, whose solution is the same as the solution <strong>of</strong> the weak formulation. For the discrete version<br />

<strong>of</strong> SPDEs, this minimization problem allows developing efficient methods for the solution <strong>of</strong> the<br />

weak formulation.<br />

The discrete version <strong>of</strong> the problem (4.15) is equivalent to the minimization problem<br />

( 1<br />

J (u) = min J (v), where J (v) = E<br />

v∈IR n ⊗S p 2 vT Au − v b)<br />

T<br />

. (4.20)<br />

This equivalence is well-known for deterministic problems, but holds for the expectation in stochastic<br />

equations, too. Using this relation, the best approximation <strong>of</strong> order M is<br />

J<br />

(<br />

∑<br />

M<br />

i=1 λ iU i<br />

)<br />

( )<br />

M<br />

= min J<br />

V 1 ,...V M ∈IR ∑ n i=1 γ iV i<br />

γ i ,...,γ M ∈S p<br />

. (4.21)<br />

It is well-known that in the deterministic setting the best approximation can be defined recursively:<br />

Let (λ 1 ,...,λ M−1 ),(U 1 ,...,U M−1 ) be the best approximation <strong>of</strong> order M − 1. Then the best approximation<br />

<strong>of</strong> order M is<br />

J<br />

(<br />

∑<br />

M<br />

i=1 λ iU i<br />

)<br />

(<br />

)<br />

= min J γV +∑ M−1<br />

V ∈IR n<br />

j=1 λ iU i<br />

γ∈S p<br />

. (4.22)<br />

This recursive definition is in general not true in the stochastic case (see the following calculations),<br />

but numerical tests show that we achieve good approximations for stochastic operators. With the<br />

recursive definition, we develop efficient numerical schemes for the solution <strong>of</strong> the minimization<br />

problem. The functional decomposes into two parts when we use the recursive definition:<br />

(<br />

λ M U M +∑ M−1<br />

i=1 λ iU i<br />

)<br />

J<br />

)<br />

1<br />

= E(<br />

2 (λ MU M ) T Au − (λ M U M ) T b<br />

( 1<br />

(<br />

M−1<br />

+ E<br />

2 ∑<br />

)<br />

i=1 λ iU i ) T b<br />

i=1<br />

) T λ iU i Au − ( ∑ M−1<br />

} {{ }<br />

already minimized<br />

. (4.23)<br />

The second summand <strong>of</strong> the equation is minimized already due to the recursive definition <strong>of</strong> the<br />

minimization. Introducing the residual values<br />

ũ = u −∑ M−1<br />

i=1 λ iU i and ˜b = b − ∑ M−1<br />

i=1 Aλ iU i (4.24)<br />

41


Chapter 4 Discretization <strong>of</strong> SPDEs<br />

and performing an additional calculation for the first term results in<br />

)<br />

1<br />

E(<br />

2 (λ MU M ) T Au − (λ M U M ) T b<br />

( 1<br />

= E<br />

2 (λ MU M ) T Aũ − (λ M U M ) T˜b 3<br />

( ) )<br />

+<br />

2 (λ MU M ) T M−1<br />

A ∑ i=1 λ iU i .<br />

(4.25)<br />

In the deterministic case, the product (λ M U M ) T A ( ∑ M−1<br />

i=1 λ )<br />

iU i is equal to zero because Ui are eigenvectors<br />

<strong>of</strong> the operator A and therefore U ⊥ AU i . In the stochastic case, this is not true, but we neglect<br />

this small error. However, the numerical results are reasonable. We introduce a functional J ˜ :<br />

J ˜<br />

1<br />

(λ M U M ) = E(<br />

2 (λ MU M ) T Aũ − (λ M U M ) T˜b<br />

)<br />

. (4.26)<br />

With the functional J ˜ , we transformed the minimization problem (4.20) into a series <strong>of</strong> simpler<br />

minimization problems. The next step is to find a method that allows an efficient solution <strong>of</strong> the<br />

problem series given by (4.26).<br />

Remark 6. The definition <strong>of</strong> the best approximation can be rewritten in matrix form as follows. Let<br />

W = (U 1 ,...,U M ) ∈ IR n×M be the matrix <strong>of</strong> all coefficients and Λ = (λ 1 ,...,λ M ) T ∈ IR M ⊗ S p the<br />

vector <strong>of</strong> stochastic functions. Then, (4.20) can equivalently be written<br />

J (WΛ) =<br />

min<br />

W∈IR n×M<br />

Λ∈IR M ⊗S p<br />

J (WΛ) , (4.27)<br />

and (4.26) can be written as<br />

˜ J (U M λ M ) =<br />

min<br />

V ∈IR n<br />

γ∈IR⊗S p<br />

˜ J (V γ) . (4.28)<br />

4.4.2 Stationary Conditions for the Functionals J and J ˜<br />

The simultaneous minimization <strong>of</strong> the functional J or J ˜ for deterministic W and stochastic Λ,<br />

respectively λ M and U M , is difficult due to the high dimension <strong>of</strong> the product space IR n ⊗ S p . A possibility<br />

to avoid the simultaneous minimization is to fix either the deterministic W or the stochastic Λ.<br />

For a fixed W, the stationary condition <strong>of</strong> J for Λ is<br />

and for fixed U the stationary condition <strong>of</strong><br />

E ( Λ ∗T (W T AW)Λ ) = E ( Λ ∗T W T b ) ∀Λ ∗ ∈ IR M ⊗ S p , (4.29)<br />

˜ J for λ is<br />

E ( λ ∗ (Ui T AU i )λ ) = E ( λ ∗ Ui<br />

T ˜b ) ∀λ ∗ ∈ S p . (4.30)<br />

Having the optimal Λ (or λ M ) at hand, we try to find a better W (or U M ) by solving stationary<br />

conditions for fixed Λ (or λ M ).<br />

For a fixed Λ, the stationary condition <strong>of</strong> J for W is<br />

and for fixed λ the stationary condition <strong>of</strong><br />

E ( Λ T (W ∗T AW)Λ ) = E ( Λ T W ∗T b ) ∀W ∗ ∈ IR n×M , (4.31)<br />

E ( λ(U ∗T<br />

i<br />

˜ J for U is<br />

AU i )λ ) = E ( λUi<br />

∗T ˜b ) ∀U ∗ ∈ IR n . (4.32)<br />

Iterating these stationary conditions leads to a solution <strong>of</strong> the coupled problem (4.27).<br />

42


4.4 Generalized Spectral Decomposition<br />

Algorithm 1 A GSD algorithm<br />

1: u ← 0, ˜b ← b<br />

2: for i = 1 to M do<br />

3: λ i ← λ 0<br />

4: for k = 1 to k max do<br />

5: U i ← E(Aλi 2)−1<br />

E(˜bλ i )<br />

6: U i ← U i /‖U i ‖<br />

7: λ i ← (Ui T AU i ) −1 Ui<br />

T ˜b<br />

8: end for<br />

9: u ← u + λ i U i<br />

10: ˜b ← ˜b − Aλ i U i<br />

11: end for<br />

4.4.3 An Algorithm for the GSD<br />

Having all the primarily presented results at hand, we combine them for a first algorithm for the<br />

numerical solution <strong>of</strong> an SPDE. The presented algorithm is a simplification <strong>of</strong> the power-type GSD<br />

algorithm presented by Nouy [113] and given in pseudo-code in Algorithm 1. The algorithm uses<br />

the presented iterative minimization <strong>of</strong> the functional J ˜ by iterating the stationary conditions (4.30)<br />

and (4.32). The expression in line 5 <strong>of</strong> the algorithm is a direct consequence <strong>of</strong> (4.32), the expression<br />

in line 7 a consequence <strong>of</strong> (4.30). As stated by Nouy [113], k max = 3,4 and M = 8 is sufficient for a<br />

good approximation <strong>of</strong> the solution, i.e. in a numerical algorithm we perform k max inner iterations <strong>of</strong><br />

(4.30) and (4.32) to find a new stochastic and a new deterministic basis function. Furthermore, we<br />

use a subspace spanned by M deterministic and M stochastic functions.<br />

4.4.4 Relation to the Karhunen-Loève Expansion<br />

The Karhunen-Loève expansion [95] is the classical way for the approximation <strong>of</strong> stochastic quantities.<br />

The expansion minimizes the distance between the solution u and an approximation <strong>of</strong> order<br />

M, i.e. we minimize the expression<br />

∥<br />

∥u −∑ M i=1 λ iU i<br />

∥ ∥∥<br />

2<br />

=<br />

∥ ∥ ∥∥u min −<br />

V 1 ,...V M<br />

∑ M ∥∥<br />

∈IR n i=1 γ 2<br />

iV i . (4.33)<br />

γ i ,...,γ M ∈S p<br />

The GSD minimizes the expression ∥ ∥u − ∑ M i=1 γ iV i<br />

∥ ∥<br />

2<br />

A , where the norm is ‖v‖2 A = E(vT Av), although<br />

the solution u is unknown before. Thus, the GSD computes the best approximation <strong>of</strong> a given order<br />

<strong>of</strong> the unknown solution, whereas we measure the distance between solution and approximation in<br />

the energy norm <strong>of</strong> the problem.<br />

Remark 7. The possibility to compute the best approximation <strong>of</strong> an unknown solution is the great<br />

advantage <strong>of</strong> the GSD, because the Karhunen-Loève expansion is able to compute an approximation<br />

with fewer modes to a known quantity only.<br />

4.4.5 Using the Polynomial Chaos Approximation with the GSD<br />

We cannot implement the GSD algorithm presented above directly, because it is formulated for random<br />

variables. Thus, an additional approximation step, e.g. the approximation <strong>of</strong> random variables in<br />

the polynomial chaos, is necessary to end up with a useful algorithm. Using the notation introduced<br />

earlier, we reformulate the steps 5 and 7 <strong>of</strong> the algorithm above. These steps are the complicated<br />

43


Chapter 4 Discretization <strong>of</strong> SPDEs<br />

Figure 4.2: Comparison <strong>of</strong> discretization methods with respect to implementational effort and speed.<br />

steps, in which we have to generate and solve equation systems. We reformulate the remaining steps<br />

in the same fashion. In step 5, we have to solve the system<br />

Using the polynomial chaos, the matrix is<br />

E(Aλ i λ j ) = ∑<br />

α<br />

E(Aλ 2<br />

i )U i = E(˜bλ i ) . (4.34)<br />

∑<br />

β<br />

E (˜bλ i<br />

)<br />

= ∑<br />

α<br />

(<br />

∑λ i,α λ j,β A γ E Ψ α Ψ β Ψ γ) . (4.35)<br />

γ<br />

The generation <strong>of</strong> the system matrix benefits from the generation <strong>of</strong> the lookup tables presented in<br />

Section 3.3.5. We generate the right hand side in a similar fashion:<br />

( ) Ψ α Ψ β . (4.36)<br />

∑<br />

˜b α λ i,β E<br />

β<br />

The value E ( Ψ α Ψ β ) is extracted from the lookup table by setting Ψ γ = 1. To sum up, the generation<br />

<strong>of</strong> the equation system requires the summation <strong>of</strong> values weighted by entries from the lookup table.<br />

The generation <strong>of</strong> the matrix and the right hand side can be parallelized.<br />

In Step 7, we have to solve the system<br />

E ( Ui T AU i Ψ α) λ i = E ( Ui<br />

T ˜bΨ α) . (4.37)<br />

Using the polynomial chaos for (4.37) results in a summation for the matrix and the right hand side:<br />

E ( Ui T AU i Ψ α) λ i = ∑∑Ui T A β U i λ i,γ E<br />

(Ψ α Ψ β Ψ γ)<br />

γ β<br />

E ( U T<br />

i<br />

˜bΨ α) = ∑<br />

β<br />

Ui<br />

T ˜b β E<br />

(<br />

Ψ α Ψ β ) .<br />

(4.38)<br />

To conclude, having a polynomial chaos approximation <strong>of</strong> the stochastic quantities, the GSD is<br />

implemented efficiently <strong>using</strong> the lookup table from Section 3.3.3. Furthermore, to improve the<br />

efficiency we skip the calculation as early as possible when the lookup table entry is zero.<br />

Fig. 4.2 compares the discretization methods presented in this chapter with respect to the implementational<br />

effort and the speed <strong>of</strong> the methods. The sampling based methods Monte Carlo simulation<br />

and stochastic collocation are easy to implement due to the possibility to reuse existing code.<br />

The drawback <strong>of</strong> these methods is the slow convergence <strong>of</strong> these methods towards the stochastic<br />

solution. The intrusive methods <strong>Stochastic</strong> FEM and the GSD need a lot <strong>of</strong> implementational effort<br />

because they cannot reuse existing deterministic code. The advantage <strong>of</strong> the these methods is the fast<br />

calculation <strong>of</strong> the stochastic result compared to the sampling based approaches.<br />

44


4.5 Adaptive Grids<br />

Figure 4.3: Refinement <strong>of</strong> a rectangular element <strong>of</strong> a finite element mesh. A single element on a<br />

coarser level splits up into four elements on the next finer level.<br />

4.5 Adaptive Grids<br />

To improve the efficiency <strong>of</strong> the GSD further, we combine the GSD with an adaptive grid approach<br />

for the spatial dimensions. Classically, images are represented by a regular grid, see Section 2.1.<br />

The discretization <strong>of</strong> stochastic images <strong>using</strong> regular image grids and the polynomial chaos will be<br />

described in detail in Section 5.1. Using adaptive grids for the spacial discretization we are able to use<br />

an optimal small basis in the stochastic dimensions through the GSD and a minimal set <strong>of</strong> nodes in<br />

the spatial dimensions, which reduces the memory requirements due to the tensor product structure.<br />

We adopt the adaptive grid approach from [129], which is based on rectangular elements and a<br />

quadtree structure for the refinement <strong>of</strong> the elements. Fig. 4.3 shows the refinement <strong>of</strong> a single<br />

element. The main idea is to start on the finest grid level and to coarsen an element if the error<br />

indicator S(x) <strong>of</strong> every node x <strong>of</strong> the element is smaller than a threshold ι.<br />

As error indicator, we used the gradient <strong>of</strong> the expected value <strong>of</strong> the solution, i.e.<br />

S(x) = |∇(E(u(x)))| . (4.39)<br />

The adaptive coarsening <strong>of</strong> rectangular elements leads to constrained or hanging nodes, i.e. nodes<br />

that are not vertices <strong>of</strong> all neighboring elements, see Fig. 4.4. These nodes need special handling<br />

when we assemble the FE-matrices, because these nodes are not usual degrees <strong>of</strong> freedom. Instead,<br />

they are constrained by the nodes which lie on the edges <strong>of</strong> the face the node lies on (see Fig. 4.4).<br />

For details about the assembling <strong>of</strong> the FE-matrices with hanging nodes, we refer to [120, 129].<br />

The error indicator S leads to problematic situations, in which the constraining node <strong>of</strong> a hanging<br />

node is also a hanging node on the next coarser level. Fig. 4.5 shows such a situation. To avoid this,<br />

the error indicator has to be saturated, as pointed out e.g. in [120, 129]. Following these references<br />

the saturation condition is as follows.<br />

Saturation condition. An error indicator value S(x) for x ∈ N (E) is always greater than every<br />

error indicator S(x C ) for x C ∈ N C (E). In this formula, N (E) are the nodes <strong>of</strong> the element E and<br />

N C (E) are the new nodes due to refinement <strong>of</strong> the element E.<br />

Figure 4.4: Refinement <strong>of</strong> elements leads to hanging nodes (circles) which are no degrees <strong>of</strong> freedom,<br />

instead the values <strong>of</strong> the constraining nodes (squares) restrict them.<br />

45


Chapter 4 Discretization <strong>of</strong> SPDEs<br />

Figure 4.5: For an unsaturated error indicator, the appearance <strong>of</strong> hanging nodes constrained by hanging<br />

nodes (due to level transitions <strong>of</strong> more than one between neighboring elements) is<br />

possible (left). The saturation <strong>of</strong> the error indicator ensures that there are level one transitions<br />

between neighboring elements only (right).<br />

This saturation condition ensures that there is a level one transition between neighboring elements<br />

only. Furthermore, we have to avoid the refinement <strong>of</strong> coarsened elements. Otherwise it is possible<br />

to end up in a situation where an element is refined in step n, coarsened in step n + 1 and so on. A<br />

slightly modified error indicator ˜S, which we define as the minimum <strong>of</strong> the actual error indicator and<br />

the error indicator <strong>of</strong> the previous iteration, achieves this. Alternatively, the refinement <strong>of</strong> coarsened<br />

elements can be avoided by <strong>using</strong> different thresholds for coarsening and refinement [34].<br />

4.5.1 Combining GSD and Adaptive Grids<br />

The combination <strong>of</strong> adaptive grids with the GSD method is straightforward. We assemble the<br />

stochastic matrices in the same way as the deterministic matrices. After the solution <strong>of</strong> the system<br />

is available, we interpolate the values on the hanging and inactive nodes. The only difficulty<br />

arises for the generation <strong>of</strong> the equation system for the new stochastic basis element (equation (4.30)<br />

respectively line 7 <strong>of</strong> the algorithm). There we have to compute the scalar product 〈U i ,AU i 〉 <strong>using</strong> the<br />

adaptive matrix A. The product AU i has a different weight for the constraining nodes than the vector<br />

U i , because the matrix has additional weights from the hanging nodes at the constraining nodes.<br />

We propose to add these weighting factors to the vector U i .<br />

Conclusion<br />

We presented methods for the discretization <strong>of</strong> SPDEs. Based on sampling strategies we presented<br />

Monte Carlo simulation and stochastic collocation with full or sparse grids constructed via Smolyak’s<br />

algorithm. This thesis uses the sampling based approaches to verify the implementations <strong>of</strong> the intrusive<br />

methods. Intrusive methods do not use a sampling strategy to solve the SPDEs. Instead, they<br />

are based on a development <strong>of</strong> numerical schemes acting on random variables. Intrusive methods<br />

are the key to the efficient numerical solution <strong>of</strong> SPDEs arising in image processing, because other<br />

methods are orders <strong>of</strong> magnitude too slow or provide inaccurate results after an adequate period. We<br />

presented the SFEM and the GSD method that tries to speed up the solution process on the SFEM<br />

by constructing an optimal, problem dependent, subspace.<br />

With this chapter, we have the fundamentals at hand to develop the concept <strong>of</strong> stochastic images<br />

and to design image processing operators acting on these stochastic images.<br />

46


Chapter 5<br />

<strong>Stochastic</strong> <strong>Images</strong><br />

As described in Section 2.5, noise corrupts classical images. The repeated acquisition <strong>of</strong> the same<br />

scene does not give identical images, because the noise typically is a stochastic quantity. Furthermore,<br />

applying segmentation methods to two randomly chosen samples from the same scene yield<br />

different results due to the noise. To model the noise <strong>of</strong> the acquisition process, we identify pixels<br />

by random variables, i.e. identify images by random fields. Assuming that these stochastic images<br />

fulfill mild regularity assumptions (H 1 -regularity in the spatial dimensions and L 2 -regularity in the<br />

stochastic dimensions), they are elements <strong>of</strong> the tensor product space H 1 (D) ⊗ L 2 (Ω) introduced in<br />

Section 3.1. We discretize this tensor product space <strong>using</strong> the polynomial chaos expansion introduced<br />

in Section 3.3 and finite elements or finite differences for the spatial dimensions. This chapter<br />

combines the methods presented so far to introduce the concept <strong>of</strong> stochastic images. Preusser et<br />

al. [130] introduced stochastic images, but used a pointwise product space, which is a subspace <strong>of</strong><br />

the tensor product space H 1 (D) ⊗ L 2 (Ω). We compare both approaches at the end <strong>of</strong> this chapter.<br />

5.1 Polynomial Chaos for <strong>Stochastic</strong> <strong>Images</strong><br />

It is popular in PDE based image processing to model an image f : D → IR on a domain D ⊂ IR d ,<br />

d = 2,3 <strong>using</strong> a finite element space and a representation<br />

f (x) = ∑ i∈I<br />

f i P i (x) , (5.1)<br />

where f i ∈ IR is the value <strong>of</strong> the ith pixel from the pixel set I and P i the shape function (e.g. tent<br />

function) <strong>of</strong> the ith pixel (see e.g. [17]). In a stochastic image, a single pixel no longer has a fixed<br />

value. Instead, it depends on a vector <strong>of</strong> random variables ξ (ω) = (ξ 1 (ω),...,ξ n (ω)) and on a<br />

random event ω ∈ Ω. Note that it is possible to combine the concept <strong>of</strong> stochastic images with<br />

other spatial discretizations, e.g. finite difference schemes. Then we have a pointwise representation<br />

f (x i ,ξ ) and apply an interpolation rule for positions located between pixel positions.<br />

Following [130], we obtain the representation <strong>of</strong> an image whose pixel values are random variables<br />

from (5.1) by replacing the fixed f i by random variables f i (ξ ):<br />

f (x,ξ ) = ∑ i∈I<br />

f i (ξ )P i (x) . (5.2)<br />

Fig. 5.1 shows a schematic sketch <strong>of</strong> this idea. Note that we omit denoting the dependence <strong>of</strong> ξ on<br />

ω to simplify the notation. The polynomial chaos expansion (3.31) approximates any second order<br />

random variable f i (ξ ) by a weighted sum <strong>of</strong> orthogonal multidimensional polynomials. This yields<br />

f (x,ξ ) = ∑ i∈I ∑ N α=1 f i αΨ α (ξ )P i (x) (5.3)<br />

as the representation <strong>of</strong> stochastic images, i.e. images whose pixels are random variables, discretized<br />

<strong>using</strong> finite elements for the spatial dimensions. Using finite differences, the value at a pixel is<br />

f (x i ,ξ ) = ∑ N α=1 f i αΨ α (ξ ) (5.4)<br />

47


Chapter 5 <strong>Stochastic</strong> <strong>Images</strong><br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

❳❳<br />

❳❳ ❳ supp ξ j , j = 1,...n<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

✘ ✘ ✘ ✘✘ ✘ x i<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

❳ ❳ ❳ ❳ ❳ supp Pi (x)<br />

<br />

<br />

Figure 5.1: Sketch <strong>of</strong> the ingredients <strong>of</strong> a stochastic image. We discretize the spatial dimensions<br />

<strong>using</strong> finite elements, but the coefficients <strong>of</strong> the FE basis functions are random variables.<br />

Every random variable has a support, which spans over the complete image, thus pixels<br />

depend on a random vector.<br />

and an interpolation rule provides the values at positions between neighboring pixels. For fixed α<br />

we call the coefficient f i α a stochastic mode <strong>of</strong> the pixel i. The set { f i α} i∈I collects the stochastic<br />

modes <strong>of</strong> all pixels for fixed α. Thus, it is possible to visualize the set as a classical image.<br />

From the polynomial chaos expansion <strong>of</strong> a stochastic image, we compute stochastic moments <strong>of</strong><br />

the image. With the use <strong>of</strong> the orthogonal set <strong>of</strong> basis functions we have E(Ψ 1 ) = 1, E(Ψ α ) = 0 for<br />

α > 1 and E(Ψ α Ψ β ) = 0 if α ≠ β. The expected value and the variance <strong>of</strong> a stochastic pixel are<br />

E( f (x i ,·)) = f i 1 ,<br />

Var( f (x i ,·)) = ∑ N α=2<br />

(<br />

f<br />

i<br />

α<br />

) 2<br />

E<br />

(<br />

(Ψ α ) 2) .<br />

(5.5)<br />

We obtain higher stochastic moments in a similar way. Furthermore, it is possible to visualize the<br />

complete stochastic information <strong>of</strong> every pixel, e.g. via a visualization <strong>of</strong> the PDFs <strong>of</strong> all pixels.<br />

Note that the representation <strong>of</strong> stochastic images presented here differs from the one discussed by<br />

Preusser et al. [130]. There, a space is used in which every pixel depends on one random variable<br />

only. However, for most <strong>of</strong> the image acquisition processes and image processing methods the<br />

assumption that the noise is independent for every pixel is not true. To represent these images in the<br />

ansatz space we let every pixel depend on a random vector ξ . Section 5.3 compares both concepts.<br />

The first step, required before the processing <strong>of</strong> the stochastic images starts, is the identification<br />

<strong>of</strong> the random variables in the input data. We estimate these random variables from data samples<br />

through the Karhunen-Loève expansion [41]. The Karhunen-Loève expansion, a stochastic version<br />

<strong>of</strong> the principal component analysis (PCA) [74], determines the eigenvalues and eigenvectors <strong>of</strong> the<br />

covariance matrix <strong>of</strong> the data samples and identifies the significant random variables in the data<br />

samples with these eigenvectors and eigenvalues.<br />

5.2 Generation <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> from Samples<br />

We obtain the polynomial chaos coefficients <strong>of</strong> the random variables X ∈ L 2 (Ω) by a maximum<br />

likelihood estimation [41], leading to a representation <strong>of</strong> X ∈ L 2 (Ω) in the polynomial chaos by<br />

X = ∑ N α=1 a αΨ α (ξ ) . (5.6)<br />

To use the notion <strong>of</strong> stochastic images developed in the previous sections for image processing, we<br />

need to obtain the coefficients <strong>of</strong> the representation (5.3) for the image undergoing the analysis. Let<br />

48


5.2 Generation <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> from Samples<br />

u (1) ,...,u (M) , with u (k) ∈ IR r , r = |I |, denote sample images, e.g. images resulting from repeated<br />

acquisitions. The goal is to identify these image samples as the samples <strong>of</strong> a vector <strong>of</strong> independent<br />

random variables X. To this end, the empirical Karhunen-Loève decomposition [95] yields<br />

u (k) = ū +∑ r √<br />

j=1 s j U j X (k)<br />

j , (5.7)<br />

where ū is the mean <strong>of</strong> the input samples. The pairs (s j ,U j ) for j = 1,...,r are the eigenpairs sorted<br />

in descending order <strong>of</strong> the r × r covariance matrix<br />

C := 1<br />

M − 1 ∑M k=1 (u(k) − ū) T (u (k) − ū) . (5.8)<br />

Moreover, the<br />

X (k)<br />

j = 1 √ s j<br />

U T j (u (k) − ū) (5.9)<br />

are samples <strong>of</strong> the desired vector <strong>of</strong> random variables X = (X 1 ,...,X n ), where n < r.<br />

The samples computed via the Karhunen-Loève expansion are samples <strong>of</strong> uncorrelated random<br />

variables, but the random variables are not necessarily independent. Using Gaussian random variables,<br />

we end up with independent random variables, because uncorrelated Gaussian random variables<br />

are independent. Gaussian random variables have the drawback <strong>of</strong> the infinite support <strong>of</strong> the<br />

density function, which causes problems in numerical schemes. Using other distributions, we assume<br />

the independence as well, leading to a small additional error, because we neglect correlation<br />

effects. Stefanou et al. [141] justified the assumption <strong>of</strong> independence by numerical experiments. In<br />

addition, they developed numerical methods for uncorrelated random variables.<br />

For a standard uniform random variable X it is possible to find a transformation g to an arbitrary<br />

distributed random variable with finite variance Y : Y = g(X). Since X(ω) ∈ [0,1] we transform it<br />

into a random variable Y with the desired distribution by applying the inverse cumulative distribution<br />

function (CDF) FY −1 <strong>of</strong> Y :<br />

Y = FY −1 (X) . (5.10)<br />

This mapping is a standard result used in textbooks about probability, e.g. in [62]. It can also be<br />

written as Y = FY<br />

−1 (F X(ω)), because X has a standard uniform distribution and X(ω) = F X (ω).<br />

Following [28], this result holds for arbitrary distributed random variables, too. In the next step, we<br />

project the random variable Y on an element <strong>of</strong> the polynomial chaos basis by multiplying with the<br />

basis element and taking the expected value.<br />

The estimation <strong>of</strong> the coefficients <strong>of</strong> the polynomial chaos expansion (5.6) <strong>of</strong> the random vector<br />

X from these samples is achieved by inverting the discrete empirical CDF F Xj , which is based on the<br />

samples X (k)<br />

j . This leads to a staircase-like approximation <strong>of</strong> the random variable X j . Following [141]<br />

we get X j,α from the projection on Ψ α via<br />

∫<br />

X j,α = E(X j Ψ α ) = F −1 (<br />

X Fξ j<br />

(y) ) Ψ α (ξ (y))dΠ . (5.11)<br />

Γ<br />

Note that the assumption <strong>of</strong> independence allows to use basis functions, which depend on one random<br />

variable only, i.e. Ψ α (ξ ) = Ψ α (ξ i ), i ∈ {1,...,n}. The empirical CDF and its empirical inverse are<br />

F Xj (x) = 1 M<br />

M<br />

∑<br />

k=1<br />

{<br />

FX −1<br />

j<br />

(y) = min x ∈<br />

( )<br />

I X (k)<br />

j ≤ x<br />

{<br />

X (k)<br />

j<br />

} M<br />

k=1<br />

,<br />

}<br />

∣ FXj (x) ≥ y<br />

,<br />

(5.12)<br />

49


Chapter 5 <strong>Stochastic</strong> <strong>Images</strong><br />

Figure 5.2: Decay <strong>of</strong> the sorted eigenvalues <strong>of</strong> the centered covariance matrix <strong>of</strong> 45 input samples<br />

from an ultrasound device.<br />

where I is the indicator function attaining value 1 for true arguments and 0 else. Note that the<br />

random variables X j are related to the eigenpairs (s j ,U j ) <strong>of</strong> the Karhunen-Loève decomposition via<br />

(5.9). With the expression for the inverse FX −1<br />

j<br />

and a numerical quadrature associated with the density<br />

ρ(ξ ) we compute the polynomial chaos coefficients Xj<br />

α independently from each other via<br />

X j,α ≈ ∑ R k=1 w kF −1<br />

X j<br />

(<br />

Fξ (y k ) ) Ψ α (y k ) , (5.13)<br />

where we used the notation R for the number <strong>of</strong> nodes <strong>of</strong> the quadrature rule and w k for the quadrature<br />

weights associated with the nodes.<br />

We emphasize that the assumption <strong>of</strong> independence <strong>of</strong> the random variables X j is strong and in<br />

general not true. However, following [141] in particular for a few input samples this assumption is<br />

reasonable. When the assumption <strong>of</strong> independence is not true, it is possible to get the polynomial<br />

chaos representation via methods presented by Stefanou et al. [141]. These methods require the<br />

resolution <strong>of</strong> an optimization problem on a Stiefel manifold [72], which is time-consuming. Desceliers<br />

[41] gives more details about the theoretical background <strong>of</strong> the presented method.<br />

Remark 8. It is necessary to store a few leading eigenvalues and eigenvectors <strong>of</strong> the covariance<br />

only to capture the significant stochastic effects in the input data. Fig. 5.2 shows the decay <strong>of</strong> the<br />

eigenvalues <strong>of</strong> the covariance matrix computed from 45 samples from an ultrasound device. The<br />

biggest eigenvalue is associated with the mean. The other two larger eigenvalues are most likely due<br />

to motion <strong>of</strong> objects in the images during the acquisition. The stochastic effects take place on scales<br />

that are orders lower than the expected value.<br />

5.2.1 Efficient Eigenpair Computation <strong>of</strong> the Covariance Matrix<br />

The computation <strong>of</strong> the covariance matrix <strong>of</strong> the input samples is a time-consuming and especially<br />

memory-consuming process, because the covariance matrix is typically dense and the memory consumption<br />

is the squared memory consumption <strong>of</strong> a single input sample. The storage <strong>of</strong> this matrix<br />

limits the usability for high-resolution images. To avoid the generation <strong>of</strong> the complete covariance<br />

matrix, we use the low rank approximation recently developed by Harbrecht et al. [63]. This approximation<br />

is based on the pivoted Cholesky decomposition and an additional post-processing step to<br />

generate a smaller matrix with the same leading eigenvalues.<br />

50


5.2 Generation <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> from Samples<br />

Figure 5.3: Left picture group: The first mode (=expected value), second mode, third mode and<br />

fourth mode <strong>of</strong> a stochastic CT image. Right: The sinogram, i.e. the raw data produced<br />

by the CT imaging device for the head phantom [139].<br />

Pivoted Cholesky Decomposition<br />

The pivoted Cholesky decomposition is based on the Cholesky decomposition, a decomposition for<br />

symmetric and nonsingular matrices [57]. The matrix A ∈ IR q×q is factorized into A = LL T , where L is<br />

a lower triangular matrix. The computation <strong>of</strong> the complete factorization requires O(q 3 ) operations.<br />

The pivoted Cholesky decomposition computes a rank m, m ≪ q, approximation <strong>of</strong> the matrix A,<br />

where the trace norm measures the difference between matrix A and low rank approximation A m :<br />

(√<br />

)<br />

‖A − A m ‖ tr = trace (A − A m ) T (A − A m ) . (5.14)<br />

We achieve this by a modification <strong>of</strong> the Cholesky decomposition by introducing a pivot search.<br />

This pivot search guarantees that the incomplete decomposition has the same leading eigenvalues as<br />

the original matrix A. A rank m approximation <strong>of</strong> the matrix A is given by the product <strong>of</strong> the two<br />

Cholesky factors L m and L T m, i.e.<br />

A m = L m L T m , (5.15)<br />

where the Cholesky factors are computed <strong>using</strong> Algorithm 1 from [63]. This algorithm needs access<br />

to the diagonal <strong>of</strong> the matrix A and m rows <strong>of</strong> the matrix only. The storage requirement decreases<br />

from q 2 to (m + 1)q and the number <strong>of</strong> operations from O(q 3 ) to O(m 3 ). However, this algorithm<br />

computes the exact values for the leading eigenvalues, not an approximation. Harbrecht [63] provides<br />

details about the theoretical background.<br />

The eigenvalue computation <strong>of</strong> the eigenvalues <strong>of</strong> A m benefits from the fact that the eigenvalues<br />

<strong>of</strong> A = L m L T m are the same as the eigenvalues <strong>of</strong> Ã = L T mL m . Thus, we transformed the computation<br />

<strong>of</strong> the m leading eigenvalues from a IR q×q matrix into the computation <strong>of</strong> the m eigenvalues <strong>of</strong> a<br />

IR m×m matrix, where m ≪ q. The eigenvectors <strong>of</strong> the initial matrix A are x = L m ˆx, where ˆx are the<br />

eigenvalues <strong>of</strong> the small matrix L T mL m (see [63]).<br />

5.2.2 Getting <strong>Stochastic</strong> <strong>Images</strong> from CT-data<br />

The construction <strong>of</strong> stochastic images from image samples requires the acquisition <strong>of</strong> a huge number<br />

<strong>of</strong> samples to get accurate results. For medical imaging techniques like US, or in other applications<br />

51


Chapter 5 <strong>Stochastic</strong> <strong>Images</strong><br />

like quality control, the repeated acquisition is possible. However, we cannot apply this technique to<br />

CT data, because the acquisition <strong>of</strong> CT data uses high-energy radiation [66]. Thus, the acquisition<br />

<strong>of</strong> multiple samples is unethical for medical applications. Therefore, we present another possibility<br />

for the generation <strong>of</strong> stochastic images from CT data based on the sinogram, the collection <strong>of</strong> rays<br />

through the object under different angles and directions [66].<br />

The approach is based on the hypothesis that the sinogram (see Fig. 5.3), the raw data <strong>of</strong> the<br />

acquisition process (see [21] for details), is free <strong>of</strong> noise and that the noise and the artifacts in the<br />

final CT images are due to the reconstruction step, which is necessary to transform the sinogram<br />

into the final data set. We use multiple reconstruction techniques and parameter settings to generate<br />

the input samples and use the technique described in the previous section to generate the stochastic<br />

images. The reconstruction techniques range from Fourier based methods to iterative methods with<br />

different settings for the data interpolation and the filter window for the low-pass filtering [154]. For<br />

the computation <strong>of</strong> the reconstructions, we use CTSim [134], for which source code is available.<br />

Thus, we combine the generation <strong>of</strong> input samples and the computation <strong>of</strong> the resulting stochastic<br />

image in one program that runs without user interaction.<br />

Another possibility to generate a stochastic image from the available CT image sample is to use a<br />

noise model.<br />

5.3 Comparison <strong>of</strong> the Space from [130] and the Space Used in this Thesis<br />

Preusser et al. [130] made a first step for the application <strong>of</strong> SPDEs in the image processing context.<br />

They proposed to use the space H h,p<br />

still<br />

:= V2 h ⊗ P p ⊂ H 1 (D) ⊗ L 2 (Γ) as ansatz space, where V2<br />

h<br />

is the classical finite element space spanned by multi-linear tent-functions P i and P p the space<br />

spanned by one-dimensional polynomials H 1 ,...,H p . Then, the authors identified a stochastic image<br />

f (x,ξ ) ∈ H h,p<br />

still<br />

with the polynomial chaos approximation:<br />

f (x,ξ ) = ∑ i∈I ∑ p α=1 f i αH α (ξ i )P i (x) . (5.16)<br />

In this representation, every pixel has its own random variable and the pixel is dependent on this<br />

random variable only. Remember, the space used in this thesis uses a limited number <strong>of</strong> random<br />

variables, but the support <strong>of</strong> these random variables ranges over the whole image.<br />

An SPDE having stochastic images as input or solution is discretized <strong>using</strong> the SFEM. The authors<br />

multiplied the equation by a test function <strong>of</strong> the form H β (ξ i )P i (x) ∈ H 1 (D) ⊗ L 2 (Γ), yielding to a<br />

block system matrix for the unknown polynomial chaos coefficients <strong>of</strong> the solution.<br />

The ansatz space and the discretization presented by Peusser et al. [130] have drawbacks in comparison<br />

to the space used in this thesis. These drawbacks are listed below:<br />

1. The authors used only test functions <strong>of</strong> the form H β (ξ i )P i (x), but functions <strong>of</strong> the form<br />

H β (ξ k )P i (x), k ≠ i, are also elements <strong>of</strong> the product space H 1 (D) ⊗ L 2 (Γ). This leads to a<br />

much too small system matrix <strong>of</strong> the SFEM method. Thus, the solution is computed in a<br />

subspace <strong>of</strong> the tensor product space H 1 (D) ⊗ L 2 (Γ) only.<br />

2. The dependence <strong>of</strong> pixels on independent random variables allows no propagation <strong>of</strong> stochastic<br />

information between the pixels. This is a serious problem when dealing with diffusion equations<br />

like in [130], because the diffusion transports stochastic information from a pixel into the<br />

surrounding region. The ansatz space chosen in [130] cannot store this information, because<br />

the neighboring pixels are independent <strong>of</strong> this specific random variable. Thus, the information<br />

is lost. To be more precise, the solution <strong>of</strong> the diffusion process <strong>of</strong> the random variables has<br />

to be projected on the ansatz space and the ansatz space is unable to store this information.<br />

Especially for diffusion equations, stochastic information is lost due to this projection step,<br />

leading to inaccurate results.<br />

52


5.4 Visualization <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong><br />

Figure 5.4: Second (left) and fifth (right) mode <strong>of</strong> a stochastic US image. The information encoded<br />

in these images is hard to interpret, because there is no deterministic equivalent.<br />

3. The ansatz space from [130] allows one basic random variable for the representation <strong>of</strong> arbitrary<br />

random variables in the polynomial chaos only. This is a strong limitation, because<br />

random variables reasonably representable in a polynomial chaos in one random variable can<br />

be properly approximated only. Other random variables with more complicated density functions<br />

have to be projected on this limited space, also leading to a loss <strong>of</strong> precision. This is due<br />

to the double limit in the Cameron-Martin theorem [27]. They showed the approximation <strong>of</strong><br />

L 2 -random variables when the number <strong>of</strong> basic random variables ξ i ,i = 1,...,n and the degree<br />

<strong>of</strong> the polynomials p goes to infinity.<br />

The ansatz space from [130] is useful only when the solution is independent for every pixel and<br />

the representation <strong>of</strong> the arbitrary random variable <strong>of</strong> a pixel through a polynomial in one random<br />

variable is sufficient. These applications are rare, especially the diffusion equations used for demonstration<br />

purposes in [130] and the segmentation methods presented in this thesis are critical.<br />

5.4 Visualization <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong><br />

During the last years, many authors developed methods for the visualization <strong>of</strong> uncertainty, see [61,<br />

125] and the references therein. The proposed visualization techniques are <strong>of</strong>ten limited to 1D or<br />

2D data. For 1D data, it is possible to draw additional information in the graph <strong>of</strong> the function,<br />

e.g. displaying the standard deviation and other stochastic quantities like kurtosis or skewness [125].<br />

The stochastic images introduced in this chapter are two- or three-dimensional. Furthermore, due<br />

to the polynomial chaos expansion, we have to visualize the additional stochastic dimensions.<br />

A stochastic image is given by (5.3) and thus, the visualization techniques for classical images are<br />

only partially feasible. One possibility for the visualization is via the images shown in Fig. 5.4. There<br />

the set fα,i i ∈ I for fixed α is visualized as a single image. The complete stochastic image can be<br />

visualized as N <strong>of</strong> such images, which is disappointing for images with high stochastic dimension.<br />

Another possibility, shown in Fig. 5.5, is to calculate the variance for pixels. The variance image is<br />

( ) (<br />

f<br />

i 2<br />

α E (Ψ α (ξ )) 2) P i (x) . (5.17)<br />

Var( f (x,ξ )) = ∑ i∈I ∑ N α=2<br />

Visualizing expected value and variance allows for getting an impression about the pixels variability.<br />

Another possibility for the visualization is to draw a set <strong>of</strong> samples from the computed output distribution,<br />

visualized in Fig. 5.6. With this sampling, we look at classical, well-known, pictures, but<br />

samples randomly drawn from the distribution highly influence the result. For a moderate number <strong>of</strong><br />

53


Chapter 5 <strong>Stochastic</strong> <strong>Images</strong><br />

Figure 5.5: Expected value (left) and variance (right) <strong>of</strong> a stochastic US-image. The expected value<br />

looks like a deterministic image and in the variance, regions with a high gray value<br />

uncertainty are visible as white dots.<br />

Figure 5.6: Two samples drawn from a stochastic image. The images differ due to realizations <strong>of</strong> the<br />

noise. In a printed version, these images look nearly the same.<br />

random variables it is also possible to generate selected samples from stochastic images by prescribing<br />

the values for every random variable. Then, we evaluate the basis functions from the polynomial<br />

chaos at these points and generate the image as the sum <strong>of</strong> the deterministic images, one for every<br />

basis function from the polynomial chaos. This can be automated by generating a dynamic image,<br />

which automatically loops over all possible realizations <strong>of</strong> the stochastic image [61].<br />

In the chapter dealing with stochastic level sets it is necessary to visualize stochastic contours,<br />

i.e. contours whose position and shape are dependent on random variables. The easiest possibility<br />

is to visualize realizations <strong>of</strong> the stochastic contour (see Fig. 5.7). Using this approach, we visually<br />

detect regions with a high uncertainty <strong>of</strong> the contour position, i.e. regions where the distance between<br />

realizations <strong>of</strong> the contour is greater than in other regions.<br />

For 3D stochastic surfaces, the visualization is even harder, because a slicing through 2D-images<br />

is cumbersome. Thus, a technique for the visualization <strong>of</strong> 3D stochastic surfaces is required. One<br />

possibility is to visualize the expected value and to color-code them by the variance [125]. Fig. 5.8<br />

shows such a visualization. The result is an image, which is comparable to the 2D result from<br />

Fig. 5.5, but combines the information into one image. Furthermore, Djurcilov [43] presented ideas<br />

for the volume rendering <strong>of</strong> stochastic images.<br />

54


5.4 Visualization <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong><br />

Figure 5.7: Visualization <strong>of</strong> realizations <strong>of</strong> a stochastic 2D contour. Every yellow line corresponds<br />

to a MC realization <strong>of</strong> the stochastic contour encoded in the stochastic image.<br />

Conclusion<br />

In this chapter, we presented the concept <strong>of</strong> stochastic images and introduced the polynomial chaos<br />

approximation <strong>of</strong> stochastic images. With the projection method from Section 5.2, we are able to construct<br />

stochastic images from samples. This is a crucial task, because without this projection method,<br />

stochastic images are a theoretical construct only, but applications cannot use them. Furthermore,<br />

we presented visualization techniques for stochastic images. The visualization is important to bring<br />

stochastic images into applications. Without an intuitive visualization <strong>of</strong> the additional stochastic<br />

content, it might be difficult to bring the concept <strong>of</strong> stochastic images into applications.<br />

Having the concept <strong>of</strong> stochastic images at hand, we investigate in the next chapters how segmentation<br />

methods can be extended to be able to accept stochastic images as input.<br />

Figure 5.8: Visualization <strong>of</strong> a 3D contour encoded in a 3D stochastic image. The expected value <strong>of</strong><br />

the 3D stochastic contour is color-coded by the variance. Regions with a high variance<br />

are red and regions with a low variance green.<br />

55


Chapter 6<br />

<strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong><br />

Using Elliptic SPDEs<br />

The task <strong>of</strong> this chapter is to combine the notion <strong>of</strong> stochastic images with the concept <strong>of</strong> SPDEs<br />

introduced in Chapter 3. SPDEs arise from variational formulations <strong>of</strong> image processing problems,<br />

when we apply these variational methods on stochastic images. In this chapter, we investigate segmentation<br />

methods based on elliptic SPDEs. Chapter 7 investigates parabolic SPDEs.<br />

Based on elliptic SPDEs we develop two segmentation methods for stochastic images, random<br />

walker segmentation and Ambrosio-Tortorelli segmentation <strong>of</strong> stochastic images. The segmentation<br />

methods differ in reference to user interaction and the number <strong>of</strong> parameters. The extension <strong>of</strong> the<br />

random walker segmentation is interactive. Thus, it is possible to improve the segmentation quality<br />

by adding additional seed regions interactively. On the other hand, the extension <strong>of</strong> the Ambrosio-<br />

Tortorelli segmentation is fully automatic. The user tunes the parameters only, but has no possibility<br />

to improve the quality <strong>of</strong> the segmentation afterwards, except for choosing a new set <strong>of</strong> parameters<br />

and trying to improve the quality this way.<br />

6.1 Random Walker <strong>Segmentation</strong> on <strong>Stochastic</strong> <strong>Images</strong><br />

Section 2.2 summarized random walker segmentation [59]. A stochastic extension <strong>of</strong> the random<br />

walker segmentation has to combine the notion <strong>of</strong> stochastic images developed in Chapter 5 with the<br />

concept <strong>of</strong> SPDEs from Chapter 3 and the discretization <strong>of</strong> SPDEs from Chapter 4.<br />

6.1.1 Deriving a <strong>Stochastic</strong> Random Walker Model<br />

The extension <strong>of</strong> the random walker segmentation [59] to a stochastic segmentation method is<br />

straightforward and follows the way for the generation <strong>of</strong> stochastic methods for image processing<br />

described by Preusser et al. [130] and by the author [1,3]. Furthermore, the author published the<br />

stochastic extension <strong>of</strong> the random walker method [5]. <strong>Stochastic</strong> images, described in Chapter 5,<br />

replace the classical images and all further steps are performed on the stochastic images.<br />

More precisely, we replace the classical image u : D → IR by a stochastic image v : D × Ω → IR<br />

as defined in (5.3). Random walker segmentation needs no assumptions about the regularity <strong>of</strong><br />

the input images, because it transforms the problem into a partition problem <strong>of</strong> a graph. To pro<strong>of</strong><br />

existence and uniqueness <strong>of</strong> the deduced SPDE related to the continuous formulation, we restrict the<br />

method to images with a H 1 -regularity in the spatial dimensions. This is the typical regularity for<br />

image processing tasks assumed for classical image processing [17]. To use the polynomial chaos<br />

expansion, we assume that the images are L 2 -regular in the stochastic dimensions. Thus, we use the<br />

tensor product space H 1 (D) ⊗ L 2 (Ω) introduced in Section 3.1. For the discretization we use the<br />

spaces V h ⊂ H 1 (D) consisting <strong>of</strong> multi-linear tent-functions for every pixel <strong>of</strong> the input image and<br />

S n,p ⊂ L 2 (Ω), a polynomial chaos expansion in n random variables with order p.<br />

We start by building a graph for the spatial dimensions <strong>of</strong> the stochastic image. On this graph,<br />

we define stochastic analogs <strong>of</strong> the edge weights and node degrees. The stochastic edge weight, the<br />

57


Chapter 6 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using Elliptic SPDEs<br />

edge weight is a random variable, is given by the same expression as the classical edge weight, but<br />

the quantities extracted from the image are random variables. Thus, the random variable describing<br />

the edge weight <strong>of</strong> the edge between neighboring pixels i and j is, cf. (2.5)<br />

(<br />

w i j (ξ ) = exp −β (g i (ξ ) − g j (ξ )) 2) . (6.1)<br />

Replacing the random variables by their polynomial chaos expansion, we have to compute<br />

w i j (ξ ) = exp<br />

( ( ) )<br />

N<br />

−β ∑ α=1 gi αΨ α (ξ ) −∑ N 2<br />

α=1 g αΨ j α (ξ )<br />

. (6.2)<br />

Section 3.3 describes how to perform calculations for random variables represented in the polynomial<br />

chaos. Note that we do not calculate the exponential <strong>of</strong> the polynomial chaos expansion explicitly.<br />

Instead, we compute a Galerkin projection <strong>of</strong> the exponential in the polynomial chaos via (3.39).<br />

From the definition <strong>of</strong> the stochastic edge weights, it is easy to generalize the node degrees to<br />

stochastic node degrees represented in the polynomial chaos:<br />

d i (ξ ) =<br />

∑<br />

{ j∈V :e i j ∈E}<br />

w i j (ξ ) =<br />

N<br />

∑ ∑<br />

{ j∈V :e i j ∈E} α=1<br />

w i, j<br />

α Ψ α (ξ ) . (6.3)<br />

The normalization step, to ensure that the maximal difference between g i and g j is one, is not straightforward<br />

because the quantities g i are random variables. A normalization <strong>of</strong> random variables is to<br />

ensure that the expected value <strong>of</strong> the random variable is one. This is achieved by dividing the difference<br />

<strong>of</strong> neighboring pixels by the maximal difference <strong>of</strong> the expected value <strong>of</strong> neighboring pixels:<br />

(g i (ξ ) − g j (ξ )) 2 (u i (ξ ) − u j (ξ )) 2<br />

=<br />

. (6.4)<br />

max k,l∈V,ek,l ∈E E<br />

((u k (ξ ) − u l (ξ ))<br />

2)<br />

From the stochastic edge weights and the stochastic node degrees it is easy to build the stochastic<br />

analog <strong>of</strong> the Laplacian matrix given by, cf. (2.9)<br />

⎧<br />

⎨ d i (ξ ) if i = j<br />

L i j (ξ ) = −w i j (ξ ) if v i and v j are adjacent nodes<br />

⎩<br />

0 otherwise<br />

(6.5)<br />

= ∑ N α=1 Lα Ψ α (ξ ) .<br />

The stochastic combinatorial Laplacian matrix has a representation in the polynomial chaos. The<br />

coefficient L α in this polynomial chaos expansion is a matrix containing at position Li α j the αth<br />

coefficient <strong>of</strong> the polynomial chaos expansion <strong>of</strong> either d i (ξ ) if i = j or <strong>of</strong> −w i j (ξ ) respectively zero.<br />

To define the linear system <strong>of</strong> equations to solve the stochastic random walker problem, we start<br />

with the stochastic analog <strong>of</strong> the weighted Dirichlet integral. It is given by taking the expected value<br />

<strong>of</strong> the classical weighted Dirichlet integral R w and inserting the stochastic quantities there:<br />

( ∫<br />

)<br />

1<br />

E(R w [u(ξ )]) = E w|∇u(ξ )| 2 dx . (6.6)<br />

2 D<br />

As for the classical energy (cf. Section 2.2), a minimizer is a harmonic function satisfying<br />

−∇ · (w(ξ )∇u(ξ )) = 0 in D × Ω<br />

u = 1 on V O<br />

u = 0 on V B .<br />

(6.7)<br />

58


6.1 Random Walker <strong>Segmentation</strong> on <strong>Stochastic</strong> <strong>Images</strong><br />

Remark 9. The methods from Section 3.2 ensure existence and uniqueness for the solution <strong>of</strong> (6.7).<br />

The coefficient fulfills w ∈ F l (D), because we use a truncated polynomial chaos representation.<br />

Using the polynomial chaos discretization <strong>of</strong> stochastic images from Chapter 5 and collecting all<br />

pixels in a vector x we get the discrete version <strong>of</strong> the Dirichlet integral<br />

) 1<br />

E(R w (x)) = E(<br />

2 xT Lx , (6.8)<br />

where L is the stochastic combinatorial matrix from (6.5). This equation requires a special ordering <strong>of</strong><br />

the vector x and the matrix L. The vector x is organized by grouping the coefficients for polynomials<br />

from the polynomial chaos together, x = (x1 1 h ,...,x|V |<br />

1<br />

,...,xN 1 h ,...,x|V |<br />

N ). The matrix L is a N × |V h |<br />

block matrix, with non-zero entries in the diagonal blocks only:<br />

L = diag(L 1 ,...,L N ) . (6.9)<br />

Reordering the vector x with respect to seeded and unseeded nodes (cf. Section 2.2) and <strong>using</strong> the<br />

same stochastic ordering scheme for the new vectors x U and x M yields<br />

( 1 [<br />

E(R w [x U ])=E x<br />

T<br />

2 M xU] [ ][ ]) ( T L M B xM 1 (<br />

B T<br />

=E x<br />

T<br />

L U x U 2 M L M x M + 2xUB T T x M + xUL T ) )<br />

U x U . (6.10)<br />

A stochastic minimizer <strong>of</strong> the discretized stochastic Dirichlet problem is given by<br />

L U (ξ )x U (ξ ) = −B(ξ ) T x M (ξ ) . (6.11)<br />

This system <strong>of</strong> linear equations is solved <strong>using</strong> the GSD. Section 4.4 describes the combination<br />

<strong>of</strong> the GSD with a discretization <strong>of</strong> the stochastic dimensions <strong>using</strong> the polynomial chaos. For<br />

the stochastic random walker segmentation, all quantities are available in their polynomial chaos<br />

approximation. The matrix is L = ∑ N α=1 Lα Ψ α (ξ ) and the vectors x U and x M are also available<br />

in their polynomial chaos approximation. Thus, we apply the algorithm presented in Section 4.4<br />

directly on these quantities. The next paragraph presents the results obtained from the GSD.<br />

Remark 10. Due to the construction <strong>of</strong> the solution via a problem on a graph, we end up with a<br />

much simpler stochastic problem in comparison to a direct solution <strong>of</strong> the SPDE (6.7). Using the<br />

definition <strong>of</strong> the solution via the graph, the matrix L has a representation L = diag(L 1 ,...,L N ). If we<br />

discretize the SPDE (6.7) via a SFEM approach, we end up with a matrix that has nonzero blocks<br />

away from the diagonal. This difference is due to a projection step <strong>of</strong> the graph representation,<br />

because the quantity w i j (ξ ) is projected back to the polynomial chaos at an early stage. When <strong>using</strong><br />

the SFEM this projection is at the end <strong>of</strong> the solution process.<br />

6.1.2 Results<br />

In the following, we demonstrate the benefits <strong>of</strong> the stochastic extension <strong>of</strong> the random walker segmentation<br />

on three data sets. The first data set consists <strong>of</strong> M = 5 samples with a resolution <strong>of</strong><br />

100 × 100 pixels from the artificial “street sequence” [99]. Note that we do not consider the images<br />

as a sequence, instead we treat them as five samples <strong>of</strong> the noisy and uncertain acquisition <strong>of</strong> the<br />

same scene. The second data set consists <strong>of</strong> 45 samples with a resolution <strong>of</strong> 300 × 300 pixels from<br />

an ultrasound device 1 . The third data set is a liver mask on a varying background with resolution<br />

129 × 129. The whole image is corrupted by uniform noise and 25 samples with different noise<br />

realizations are treated as input. We computed a stochastic image containing n = 5 random variables<br />

1 Thanks to Dr. Darko Ojdanić for providing the data set.<br />

59


Chapter 6 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using Elliptic SPDEs<br />

Figure 6.1: Expected value (top row) and variance (bottom row) <strong>of</strong> the street image (left) and the US<br />

image (right). Color-coded are the seed regions for interior (yellow) and exterior (red).<br />

for the ultrasound device, n = 3 random variables for the liver samples, and n = 2 random variables<br />

for the street scene. The number <strong>of</strong> random variables is chosen in dependence on the decay <strong>of</strong> the<br />

eigenvalues <strong>of</strong> the covariance matrix <strong>of</strong> the input samples (cf. Section 5.2). The eigenvalues <strong>of</strong> the<br />

covariance matrix show an exponential decay. Thus, it is sufficient to store a few <strong>of</strong> them to capture<br />

the important stochastic effects. For the three data sets, we used a polynomial degree p = 3 and<br />

computed a stochastic image from these samples <strong>using</strong> the method presented in Section 5.2. The<br />

polynomial degree <strong>of</strong> three for the polynomial chaos expansion is a good balance between accuracy<br />

<strong>of</strong> the polynomial expansion and computational effort. The user defines the seed points for the segmentation<br />

<strong>of</strong> the image on the expected value <strong>of</strong> the stochastic image. Fig. 6.1 shows the expected<br />

value, the variance, and the seed points. With the stochastic image and the seed points as input, we<br />

perform the stochastic random walker segmentation. The only free parameter β varies during the<br />

experiments. Fig. 6.2 shows the expected value <strong>of</strong> the segmented object for different values <strong>of</strong> β.<br />

Together with the expected value, we are able to show the variance <strong>of</strong> the segmentation result for<br />

every pixel. The variance <strong>of</strong> the pixels gives information, how the gray value uncertainty in the<br />

stochastic input image influences the segmentation results. Thus, the variance, respectively the polynomial<br />

chaos coefficients <strong>of</strong> the result, contains the information how the gray value uncertainty in<br />

the input image propagates through the segmentation process and influences the result. Regions with<br />

a high variance indicate regions where the input gray value uncertainty influences the detection <strong>of</strong><br />

the object. It is obvious from the images that the uncertainty changes from the input to the output. In<br />

the input data, the gray value uncertainty spreads over the whole image, whereas in the segmentation<br />

result, the gray value uncertainty concentrates at the object boundary.<br />

60


6.1 Random Walker <strong>Segmentation</strong> on <strong>Stochastic</strong> <strong>Images</strong><br />

<strong>Stochastic</strong> <strong>Images</strong><br />

Mean only<br />

β = 3 β = 5 β = 5<br />

E<br />

Ultrasound Image<br />

Var<br />

n/a<br />

E<br />

Cartoon Image<br />

Var<br />

n/a<br />

Figure 6.2: Mean and variance <strong>of</strong> the probabilities for pixels to belong to the object. Furthermore, we<br />

show in red Monte Carlo realizations <strong>of</strong> the object boundary sampled from the stochastic<br />

result. A high variance indicates pixels where the gray value uncertainty highly influences<br />

the result. For comparison we added a classical random walker segmentation result<br />

in the last column. There the variance image is not available, because the method acts on<br />

a classical image.<br />

61


Chapter 6 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using Elliptic SPDEs<br />

Figure 6.3: MC-realizations <strong>of</strong> the stochastic object boundary for the stochastic liver image segmented<br />

with the stochastic random walker approach with β = 10. On the right we highlight<br />

a region <strong>of</strong> the image, where the noise in the input image influences the result.<br />

Remark 11. The classical random walker result can be interpreted as a probability map, i.e. the<br />

result <strong>of</strong> random walker segmentation is a probability for every pixel to belong to the object or not.<br />

When we apply the stochastic random walker method, the first interpretation <strong>of</strong> the result is that we<br />

compute “probabilities <strong>of</strong> probabilities”, because we are computing the probability distribution <strong>of</strong><br />

the values at every pixel. Let us emphasize that the probability interpretation <strong>of</strong> the classical result is<br />

one possible interpretation only. Mathematically, we are computing the result <strong>of</strong> a diffusion problem,<br />

and the stochastic extension is, in this interpretation, a diffusion problem with stochastic diffusion<br />

coefficient. The results can be interpreted as the stochastic solution <strong>of</strong> the stochastic diffusion problem.<br />

The analog to the probability interpretation is that we computed the probabilities for belonging<br />

to the object in dependence on the input gray value uncertainty.<br />

The stochastic object boundary can be visualized by tracking the deterministic object boundary (the<br />

value 0.5 in the result image) for realizations <strong>of</strong> the random variables. The work <strong>of</strong> Prassni et al. [126]<br />

inspired this kind <strong>of</strong> visualization. The difference is that Prassni et al. [126] visualized the iso-lines<br />

<strong>of</strong> different probabilities, whereas we visualize the same iso-line for realizations <strong>of</strong> the stochastic<br />

image. Fig. 6.3 shows the result <strong>of</strong> such a visualization.<br />

It is possible to compute and visualize other quantities extracted from the segmentation, e.g. the<br />

volume <strong>of</strong> the segmented object. The obvious visualization <strong>of</strong> the stochastic volume is to draw the<br />

PDF <strong>of</strong> the volume. The PDF <strong>of</strong> the segmented volume can be computed from the segmentation by<br />

summing up the random variables at every pixel, because they specify the “probability” that the pixel<br />

belongs to the object. Thus, the random variable v(ξ ) specifying the objects volume is<br />

v(ξ ) =<br />

N<br />

∑<br />

α=1<br />

v α Ψ α (ξ ) := ∑ x i (ξ ) . (6.12)<br />

i∈I<br />

Having a look at the PDF, it is easy to decide whether the image noise influences the segmented<br />

volume strongly or not. If the segmented volume is strongly influenced the PDF is broad, otherwise<br />

the function is narrow. Fig. 6.4 shows the PDF <strong>of</strong> the segmented volume from the street image.<br />

Moreover, the choice <strong>of</strong> the parameter β influences the pr<strong>of</strong>ile <strong>of</strong> the PDF. A smaller β leads to a<br />

diffuse object boundary and to a broader PDF.<br />

62


6.1 Random Walker <strong>Segmentation</strong> on <strong>Stochastic</strong> <strong>Images</strong><br />

Figure 6.4: PDF <strong>of</strong> the area <strong>of</strong> the segmented person from the street image for β = 25 (black) and<br />

β = 50 (gray). From the PDF we judge the reliability <strong>of</strong> the segmentation, a narrow PDF<br />

indicates that the image noise influences the segmentation marginally.<br />

Remark 12. Another possibility to calculate the object volume is to count pixels with a value above<br />

0.5 only. Thus, we compute image samples from the stochastic result via a Monte Carlo approach,<br />

threshold these samples, count the number <strong>of</strong> object pixels, and calculate the volume PDF. This<br />

method is time-consuming. The proposed method has the advantage to include the partial volume<br />

effect [21] at the boundary, because it considers pixels with a probability less than 0.5 partially.<br />

6.1.3 Comparison with Monte Carlo Simulation and <strong>Stochastic</strong> Collocation<br />

To verify the intrusive solution <strong>of</strong> the resulting SPDE via polynomial chaos, stochastic finite elements,<br />

and the GSD method, we compared this solution with the solutions obtained via Monte Carlo<br />

sampling and a stochastic collocation approach. Fig. 6.5 shows the comparison <strong>of</strong> the expected value<br />

and the variance computed via GSD, stochastic collocation, and Monte Carlo sampling. The small<br />

difference between the variances <strong>of</strong> the three solutions might be due to the projection <strong>of</strong> the Laplacian<br />

matrix on the polynomial chaos. However, the great benefit <strong>of</strong> the GSD method is the significantly<br />

better performance. We now investigate this in detail.<br />

6.1.4 Performance Evaluation<br />

Due to the availability <strong>of</strong> the implementation possibilities for the solution <strong>of</strong> SPDEs, we are able<br />

to compare the execution times <strong>of</strong> the approaches. We did the detailed comparison for the random<br />

walker segmentation in this thesis only, but the results generalize to the Ambrosio-Tortorelli<br />

approach, because it uses the same methods.<br />

Table 6.1 shows the comparison <strong>of</strong> the execution times <strong>of</strong> the GSD method, the Monte Carlo<br />

method, and the stochastic collocation method with Smolyak and full grid. It is easy to see that<br />

the GSD method outperforms the sampled based approaches. This supports the decision to prefer<br />

the GSD method and the finite difference method for random variables throughout this thesis. The<br />

stochastic collocation methods suffer from the “curse <strong>of</strong> dimension” [119], because the execution<br />

times grow exponentially with the number <strong>of</strong> random variables in the stochastic images.<br />

63


Chapter 6 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using Elliptic SPDEs<br />

GSD Monte Carlo <strong>Stochastic</strong> Collocation<br />

E<br />

Var<br />

Figure 6.5: Comparison <strong>of</strong> the discretization methods for the computation <strong>of</strong> the stochastic random<br />

walker result to verify the intrusive discretization. The small difference between the<br />

intrusive discretization via the GSD method and the two other sampling based approaches<br />

might be due to the projection <strong>of</strong> the Laplacian matrix on the polynomial chaos.<br />

The great benefit <strong>of</strong> the Monte Carlo method is that the method is independent on the number <strong>of</strong><br />

random variables. Nevertheless, the 1000 samples used in this comparison are a lower bound for the<br />

number <strong>of</strong> runs needed to get accurate results. Recall that the rate <strong>of</strong> convergence is O(( √ R −1 )) and<br />

even with this number <strong>of</strong> runs, the Monte Carlo method is slower than the GSD method.<br />

6.1.5 <strong>Segmentation</strong> and Volumetry <strong>of</strong> an Object<br />

In many applications, the noise <strong>of</strong> every pixel in the image is independent <strong>of</strong> the noise <strong>of</strong> the neighboring<br />

pixels. It is possible to model this kind <strong>of</strong> stochastic images with our approach, too. In this<br />

case, we have to use one basic random variable for every pixel, i.e. we end up with n = |I | basic<br />

random variables for the polynomial chaos.<br />

Street (n = 2) Liver (n = 3) Ultrasound (n = 5)<br />

Monte Carlo 76 113 1814<br />

Stoch. Collocation (full grid) 16 390 ≈ 1 400 000<br />

Stoch. Collocation (sparse grid) 6 18 634<br />

GSD 9 15 437<br />

Table 6.1: Comparison <strong>of</strong> the execution times (in sec) <strong>of</strong> the discretization methods.<br />

64


6.1 Random Walker <strong>Segmentation</strong> on <strong>Stochastic</strong> <strong>Images</strong><br />

Figure 6.6: Input “doughnut” without noise (left) and noisy input image treated as expected value <strong>of</strong><br />

the stochastic image (right).<br />

To demonstrate the possibility to model such images, we used an artificial test image, a “doughnut”<br />

with an area <strong>of</strong> 60 pixels in front <strong>of</strong> a constant background with resolution 20 × 20 pixels. Fig. 6.6<br />

shows the noise-free initial image. We corrupted the image by uniform noise (see Fig. 6.6) and treated<br />

the noisy image as the expected value <strong>of</strong> our stochastic image. This modeling is close to the situation<br />

in real applications. There, the real noise-free image is not available and thus, the sample at hand is<br />

the best available estimate <strong>of</strong> the expected value. Due to the high number <strong>of</strong> random variables, we<br />

restricted the polynomial chaos to a degree <strong>of</strong> one, i.e. we are able to capture the effects expressible<br />

in<br />

(<br />

uniform random variables only. Using a polynomial degree <strong>of</strong> order one the polynomial chaos has<br />

401<br />

) (<br />

1 = 401 coefficients, <strong>using</strong> a polynomial degree <strong>of</strong> two we would end up with 402<br />

)<br />

2 = 80601.<br />

An up-to-date personal computer cannot store such a high number <strong>of</strong> stochastic modes. A solution<br />

could be the sparse polynomial chaos introduced by Blatman and Sudret [22].<br />

After initialization <strong>of</strong> the expected value with the noisy image, we have to prescribe values for the<br />

remaining polynomial chaos coefficients <strong>of</strong> the input image. Since we assume that the noise at every<br />

pixel is independent, we have to prescribe a value for the coefficient corresponding to the random<br />

variable <strong>of</strong> the pixel. We set this coefficient to 0.5/ √ 3, modeling a uniform distributed random<br />

variable with support [w − 0.5,w + 0.5] around the expected value w given by the noisy input image.<br />

The result <strong>of</strong> the random walker on this stochastic image is a stochastic image with the same<br />

number <strong>of</strong> random variables. Since the random walker method requires the solution <strong>of</strong> a stochastic<br />

diffusion equation, stochastic information is transported between the pixels. Thus, a pixel in<br />

the result image depends on all basic random variables <strong>of</strong> the input image. The visualization <strong>of</strong><br />

polynomial chaos coefficients <strong>of</strong> the solution is unintuitive and cumbersome, because we have 401<br />

coefficients per pixel. Consequently, we use the visualization techniques from Section 5.4. Fig. 6.7<br />

shows realizations <strong>of</strong> the stochastic object boundary and the seed points for the segmentation.<br />

In applications, features <strong>of</strong> the segmented object are <strong>of</strong> interest, e.g. in medical applications the<br />

volume <strong>of</strong> the object is <strong>of</strong> interest to get information about the growth or shrinkage <strong>of</strong> the segmented<br />

lesion. The volume <strong>of</strong> the segmented object in the stochastic image is a stochastic quantity, because<br />

it depends on the particular noise realization. Thus, it is possible to visualize the PDF <strong>of</strong> the object<br />

volume. We investigated two possibilities to compute the volume PDF from the stochastic segmentation<br />

result. Section 6.1.2 introduced the first method. There the polynomial chaos expansions <strong>of</strong><br />

all pixels are added, and the PDF <strong>of</strong> the resulting random variable is computed via Monte Carlo<br />

sampling from this random variable. This method is comparable with methods that consider partial<br />

volume effects, because there is no binary decision whether the pixel belongs to the object or not. In<br />

fact, we add all the stochastic possibilities <strong>of</strong> the pixels to belong to the object.<br />

65


Chapter 6 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using Elliptic SPDEs<br />

Figure 6.7: Left: The object seed points (yellow) and background seed points (red) used as initialization<br />

<strong>of</strong> the stochastic random walker method. Right: The MC-realizations <strong>of</strong> the<br />

stochastic segmentation result differ significantly for different noise realizations.<br />

The other possibility to compute the volume PDF from the stochastic result is inspired by the<br />

classical method to compute the random walker result. We generate samples from the stochastic<br />

segmentation result via Monte Carlo sampling and estimate the volume <strong>of</strong> the object given by pixels<br />

with value above 0.5 on every sample. Fig. 6.8 compares the two approaches for the computation <strong>of</strong><br />

the object’s volume. Having in mind that the “real” object volume is 60 pixels, both methods slightly<br />

overestimate the object’s volume, but the real object volume is close to the expected value (60.39 for<br />

the summation <strong>of</strong> the random variables and 60.83 for the object thresholding) <strong>of</strong> both PDFs.<br />

Figure 6.8: The PDF for both possibilities <strong>of</strong> the volume computation, the summation <strong>of</strong> the random<br />

variables (gray) and the thresholding (black). The true volume is 60 pixels.<br />

66


6.2 Ambrosio-Tortorelli <strong>Segmentation</strong> on <strong>Stochastic</strong> <strong>Images</strong><br />

6.2 Ambrosio-Tortorelli <strong>Segmentation</strong> on <strong>Stochastic</strong> <strong>Images</strong><br />

In the following, we focus on the combination <strong>of</strong> the notion <strong>of</strong> stochastic images with the segmentation<br />

approach in the spirit <strong>of</strong> Ambrosio and Tortorelli [14]. Section 2.3 introduced this approach.<br />

The author published the stochastic Ambrosio-Tortorelli extension in [3].<br />

For the segmentation <strong>of</strong> stochastic images by the phase field approach <strong>of</strong> Ambrosio and Tortorelli,<br />

we replace the deterministic u and φ by their stochastic analogs. The stochastic energy components<br />

are then defined as the expectations <strong>of</strong> the classical energy components (cf. Section 2.3), i.e.<br />

∫<br />

Efid s (u) := E(E fid) =<br />

E s reg(u,φ) := E(E reg ) =<br />

E s phase (φ) := E(E phase) =<br />

∫<br />

Γ D<br />

∫ ∫<br />

Γ D<br />

∫ ∫<br />

Γ<br />

(u(x,ξ ) − u 0 (x,ξ )) 2 dxdΠ<br />

and we define the stochastic energy as the sum <strong>of</strong> these, i.e.<br />

D<br />

µ<br />

(φ (x,ξ ) 2 + k ε<br />

)<br />

|∇u(x,ξ )| 2 dxdΠ<br />

ν ε |∇φ(x,ξ )| 2 + ν 4ε (1 − φ (x,ξ ))2 dxdΠ<br />

(6.13)<br />

EAT(u,φ) s = Efid s (u) + Es reg(u,φ) + Ephase s (φ) . (6.14)<br />

The Euler-Lagrange equations <strong>of</strong> the stochastic Ambrosio-Tortorelli energy are obtained from the<br />

first variation <strong>of</strong> (6.13). Since the stochastic energies (6.13) are the expectations <strong>of</strong> the classical<br />

energies (2.16), the computations are analog. For example, we get for a test function θ : D × Γ → IR<br />

d<br />

∣<br />

dt Es fid (u +t θ) ∣∣t=0<br />

= d ∫ ∫ (<br />

) 2 ∣<br />

∣∣t=0<br />

u(x,ξ ) +tθ(x,ξ ) − u 0 (x,ξ ) dx dΠ<br />

dt<br />

Γ D<br />

∫ ∫ (<br />

)<br />

= 2 u(x,ξ ) − u 0 (x,ξ ) θ(x,ξ ) dx dΠ .<br />

Γ<br />

D<br />

(6.15)<br />

With analog computations for the remaining energy contributions, we arrive at the following system<br />

<strong>of</strong> SPDEs: We seek for u,φ : D × Γ → IR as the weak solutions <strong>of</strong><br />

−∇ · (µ(φ(x,ξ<br />

) 2 + k ε )∇u(x,ξ ) ) + u(x,ξ ) = u 0 (x,ξ )<br />

( 1<br />

−ε∆φ(x,ξ ) +<br />

4ε + µ )<br />

2ν |∇u(x,ξ )|2 φ(x,ξ ) = 1<br />

4ε . (6.16)<br />

This system is analog to the classical system (2.18) in which stochastic images replace the classical<br />

images. The equations are SPDEs, because the coefficients φ(x,ξ ) 2 and |∇u(x,ξ )| 2 are random<br />

fields. Moreover, the right hand side <strong>of</strong> the first equation, u 0 (x,ξ ), is a random field. We use random<br />

fields from the tensor product space H 1 (D) ⊗ L 2 (Ω). This space enables us to use finite elements for<br />

the discretization <strong>of</strong> the spatial part and the polynomial chaos expansion for the stochastic part.<br />

Remark 13. Recently, Krajsek et al. [86] developed an extension <strong>of</strong> the Ambrosio-Tortorelli model<br />

based on Bayesian estimation theory [77]. This concept is related to the approach presented here,<br />

but limited to Gaussian random variables, whereas the approach presented here deals with arbitrary<br />

distributions with finite variance. The investigation <strong>of</strong> a link between the approaches is future work.<br />

67


Chapter 6 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using Elliptic SPDEs<br />

6.2.1 Γ-Convergence <strong>of</strong> the <strong>Stochastic</strong> Ambrosio-Tortorelli Model<br />

Ambrosio and Tortorelli [14] showed the Γ-convergence <strong>of</strong> their model towards the Mumford-Shah<br />

model. It is possible to extend this result to show the Γ-convergence <strong>of</strong> the stochastic extension <strong>of</strong><br />

the Ambrosio-Tortorelli model towards a stochastic Mumford-Shah model. For the formulation <strong>of</strong><br />

the result, we use the stochastic analog <strong>of</strong> the space D h,n from [14], the space D h,n ⊗ L 2 (Ω), which<br />

contains admissible functions for the energies. In the notation <strong>of</strong> [14], n is the space dimension and<br />

h = 1/ √ ε. Thus, letting the scale <strong>of</strong> the phase field ε tend to zero is equivalent to letting h → ∞.<br />

Theorem 6.1. The stochastic Ambrosio-Tortorelli model E(E AT ) Γ-converges to the stochastic<br />

Mumford-Shah model E(E MS ) as ε → 0. More precisely let (u h ,φ h ) ∈ D h,n ⊗ L 2 (Ω) be a sequence<br />

that converges to (u,φ) in D h,n ⊗ L 2 (Ω). Then we have<br />

∫<br />

∫<br />

E MS (u(ω),K(ω))dω ≤ liminf E AT (u h (ω),φ h (ω))dω (6.17)<br />

h→∞<br />

Ω<br />

and for every (u,φ) there exists a sequence (u h ,φ h ) ∈ D h,n converging to (u,φ) such that<br />

∫<br />

∫<br />

E MS (u(ω),K(ω))dω ≥ limsup E AT (u h (ω),φ h (ω))dω . (6.18)<br />

Ω<br />

h→∞ Ω<br />

In both inequalities, the edge set K is defined accordingly as the discontinuity set <strong>of</strong> u.<br />

Pro<strong>of</strong>. We begin the pro<strong>of</strong> by citing a famous theorem for the interchange <strong>of</strong> a limit process and<br />

integration, Fatou’s lemma (see [32]):<br />

Theorem 6.2 (Fatou’s lemma). For a sequence <strong>of</strong> nonnegative measurable functions f n ,<br />

∫<br />

∫<br />

liminf f n ≤ liminf f n . (6.19)<br />

We have to show that we can interchange the limit process and the integration. Let us assume that<br />

this interchange is possible (all requirements <strong>of</strong> Fatou’s lemma are satisfied). Then, we have<br />

∫<br />

∫<br />

∫<br />

liminf E AT (u h ,φ h )dω ≥ liminf E AT (u h (ω),φ h (ω))dω = E MS (u(ω),K(ω))dω (6.20)<br />

h→∞ Ω<br />

Ω h→∞ Ω<br />

and by <strong>using</strong> the reverse <strong>of</strong> Fatou’s lemma we get<br />

∫<br />

∫<br />

∫<br />

limsup<br />

h→∞<br />

Ω<br />

E AT (u h ,φ h )dω ≤<br />

Ω<br />

limsupE AT (u h (ω),φ h (ω))dω =<br />

h→∞<br />

Ω<br />

Ω<br />

E MS (u(ω),K(ω))dω , (6.21)<br />

because for every realization ω ∈ Ω we have the Γ-convergence <strong>of</strong> the Ambrosio-Tortorelli model to<br />

the Mumford-Shah model initially proved by Ambrosio and Tortorelli [14]. Thus, we have to show<br />

that the interchange <strong>of</strong> the limit process and the integration is possible.<br />

The existence <strong>of</strong> the deterministic series ensures the existence <strong>of</strong> a series for which the limit<br />

superior is less than the Γ-limit. For every ω ∈ Ω we choose the deterministic series constructed by<br />

Ambrosio and Tortorelli [14]. The inequality is ensured because Fatou’s lemma yields<br />

∫<br />

∫<br />

∫<br />

limsup<br />

h→∞<br />

Ω<br />

E AT (u h ,φ h )dω ≤<br />

Ω<br />

limsupE AT (u h (ω),φ h (ω))dω =<br />

h→∞<br />

Ω<br />

E MS (u(ω),K(ω))dω . (6.22)<br />

We justify the applicability <strong>of</strong> Fatou’s lemma in the following.<br />

To use Fatou’s lemma we have to show that E AT is nonnegative and measurable. The first condition<br />

is trivially ensured, because E AT is the sum <strong>of</strong> integrals <strong>of</strong> positive (squared) functions and thus<br />

nonnegative. The second condition is also ensured, because <strong>of</strong> the following theorem from [142]:<br />

Theorem 6.3. Any lower semicontinuous function f is measurable.<br />

Following [14], the functional E AT is semicontinuous when we use the space D h,n . Thus, the<br />

Ambrosio-Tortorelli functional is nonnegative and measurable and Fatou’s lemma can be applied.<br />

68


6.2 Ambrosio-Tortorelli <strong>Segmentation</strong> on <strong>Stochastic</strong> <strong>Images</strong><br />

✛<br />

✛<br />

j<br />

✻<br />

i<br />

❄<br />

✲<br />

α<br />

✲<br />

✻ + L α,β<br />

✟ ✟✟✟✟✟Mα,β β<br />

❄<br />

Figure 6.9: Structure <strong>of</strong> the block system <strong>of</strong> an SPDE. Every block has the sparsity structure <strong>of</strong> a<br />

classical finite element matrix and the block structure <strong>of</strong> the matrix is sparse, meaning<br />

that some <strong>of</strong> the blocks are zero. The sparsity structure on the block level depends on the<br />

number <strong>of</strong> random variables and the polynomial chaos degree used in the discretization.<br />

6.2.2 Weak Formulation and Discretization<br />

The system (6.16) contains two elliptic SPDEs, which are supposed to be interpreted in the weak<br />

sense. To this end, we multiply the equations by a test function θ : H 1 (D) × L 2 (Γ) → IR, integrate<br />

over Γ with respect to the corresponding probability measure, and integrate by parts over the physical<br />

domain D. For the first equation in (6.16) this leads us to<br />

∫<br />

Γ<br />

∫<br />

D<br />

)<br />

∫<br />

µ<br />

(φ (x,ξ ) 2 + k ε ∇u(x,ξ ) · ∇θ(x,ξ ) + u(x,ξ )θ(x,ξ )dxdΠ =<br />

Γ<br />

∫<br />

D<br />

u 0 (x,ξ )θ(x,ξ )dxdΠ<br />

(6.23)<br />

and to an analog expression for the second part <strong>of</strong> (6.16). Here we assume homogeneous Neumann<br />

boundary conditions for u and φ such that no boundary terms appear in the weak form. For the existence<br />

<strong>of</strong> solutions <strong>of</strong> this SPDE, the constant k ε is supposed to ensure the positivity <strong>of</strong> the diffusion<br />

coefficient µ(φ 2 + k ε ). In fact, there must exist c min ,c max ∈ (0,∞) and I = [c min ,c max ] such that<br />

(<br />

P ω ∈ Ω ∣ )<br />

µ<br />

(φ (x,ξ (ω)) 2 + k ε ∈ I<br />

)<br />

∀x ∈ D = 1 , (6.24)<br />

i.e. the coefficient is bounded almost sure by c min and c max .<br />

The Doob-Dynkin lemma (see Section 3.2.3) ensures that the solutions <strong>of</strong> the SPDEs have a representation<br />

in the same basis as the input, allowing us to use the same polynomial chaos approximation<br />

for the input and the solution <strong>of</strong> the SPDEs. This is due to the continuity and measurability <strong>of</strong> the<br />

stochastic partial differential operators.<br />

The weak system (6.23) is discretized by a substitution <strong>of</strong> the polynomial chaos expansion (5.3) <strong>of</strong><br />

the image and the phase field. As test functions, products P j (x)Ψ β (ξ ) <strong>of</strong> spatial basis functions and<br />

stochastic basis functions are used. Denoting the vectors <strong>of</strong> coefficients by U α = (u i α) i∈I ∈ IR |I |<br />

and similarly for the phase field φ and the initial image u 0 we get the fully discrete systems<br />

N (<br />

∑<br />

) M α,β + L α,β U α =<br />

α=1<br />

(<br />

N<br />

∑<br />

α=1<br />

εS α,β + T α,β ) Φ α =<br />

N<br />

∑<br />

α=1<br />

N<br />

∑<br />

α=1<br />

M α,β (U 0 ) α<br />

A α (6.25)<br />

69


Chapter 6 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using Elliptic SPDEs<br />

Figure 6.10: Nonzero pattern <strong>of</strong> the SFEM matrix for the smoothed stochastic image <strong>using</strong> n = 5<br />

random variables and a polynomial degree p = 3. A black dot denotes a block that has<br />

a nonzero stochastic part, thus having the sparsity structure <strong>of</strong> a classical FEM matrix.<br />

for all β ∈ {1,...,N}, where M α,β ,L α,β ,S α,β and T α,β are blocks <strong>of</strong> the system matrix, defined as<br />

(M α,β ) i, j = E (Ψ α Ψ β ) ∫<br />

(S α,β ) i, j = E (Ψ α Ψ β ) ∫<br />

D<br />

D<br />

P i P j dx ,<br />

∇P i · ∇P j dx<br />

(6.26)<br />

and (<br />

(<br />

L α,β ) i, j = ∑<br />

k<br />

T α,β ) i, j = ∑<br />

k<br />

(<br />

∑E<br />

Ψ α Ψ β Ψ γ) ∫<br />

(˜φ 2 ) k γ<br />

γ<br />

∫<br />

∑E<br />

γ<br />

(<br />

Ψ α Ψ β Ψ γ) u k γ<br />

D<br />

D<br />

∇P i · ∇P j P k dx ,<br />

P i P j P k dx .<br />

(6.27)<br />

Here, (˜φ 2 ) k γ denotes a coefficient <strong>of</strong> the polynomial chaos expansion <strong>of</strong> the Galerkin projection <strong>of</strong> φ 2<br />

onto the image space (cf. Section 3.3.3). The right hand side vector <strong>of</strong> the phase field equation is<br />

∫<br />

(A α ) i =<br />

Γ<br />

∫<br />

Ψ α dΠ<br />

D<br />

⎧ ∫<br />

1<br />

⎪⎨<br />

4ε P i dx =<br />

D ⎪⎩<br />

1<br />

4ε P i dx if α = 1 ,<br />

0 else .<br />

(6.28)<br />

Note that the expectations <strong>of</strong> the products <strong>of</strong> stochastic basis functions involved above are again<br />

the components <strong>of</strong> the lookup table introduced in Section 3.3.5. The deterministic integrals can be<br />

precomputed, because they are needed several times during the assembling <strong>of</strong> the system matrix.<br />

Analog to the classical finite element method the systems <strong>of</strong> linear equations can be treated by an<br />

iterative solver like the method <strong>of</strong> conjugate gradients [67].<br />

The general block structure <strong>of</strong> an SFEM matrix is depicted in Fig. 6.9 and the sparsity structure on<br />

the block level for five random variables and a polynomial degree <strong>of</strong> three is depicted in Fig. 6.10.<br />

In addition, the matrix generation can be accelerated <strong>using</strong> lookup tables. The memory consumption<br />

is enormous, because the matrix has N 2 -times the storage requirement <strong>of</strong> the deterministic<br />

matrix, where N is the dimension <strong>of</strong> the polynomial chaos. Thus, we use the GSD method for the<br />

solution to avoid the generation <strong>of</strong> the SFEM matrix.<br />

70


6.2 Ambrosio-Tortorelli <strong>Segmentation</strong> on <strong>Stochastic</strong> <strong>Images</strong><br />

<strong>Stochastic</strong> Generalization <strong>of</strong> the Edge Linking Step<br />

The edge linking step from Section 2.3 can be applied on the stochastic Ambrosio-Tortorelli model,<br />

too. We introduce an additional coefficient c for the image equation. This coefficient is a random<br />

field, i.e. c ∈ H 1 (D) × L 2 (Ω). The modified image equation in the stochastic context reads<br />

−∇ · (µc(x,ξ<br />

)(φ(x,ξ ) 2 + k ε )∇u(x,ξ ) ) + u(x,ξ ) = u 0 (x,ξ ) . (6.29)<br />

The random field c is composed <strong>of</strong> the stochastic generalizations <strong>of</strong> the edge continuity and the edge<br />

consistency step. Thus, c is<br />

c(x,ξ ) = c dc (x,ξ ) · c h (x,ξ ) , (6.30)<br />

whereas these quantities are<br />

(<br />

(c dc (ξ )) i = ζ (ξ ) dc) + 1 − ( ζ (ξ ) dc) i<br />

i φ(ξ ) i<br />

( (<br />

))<br />

(<br />

ζ (ξ ) dc) = exp ε dc 1<br />

i |η s | ∑ ∇u i (ξ ) · ∇u j (ξ ) − 1<br />

j∈η s<br />

1<br />

(c h (ξ )) i =<br />

1 + α ( )<br />

φ(ξ ) i − φ(ξ ) 2 .<br />

i<br />

(6.31)<br />

To calculate c dc and c h , it is necessary to use the calculations for random variables approximated in<br />

the polynomial chaos presented in Section 3.3.3. The only quantity for which a stochastic generalization<br />

is not obvious is the orthogonal edge direction ∇u ⊥ i . This direction is needed, because we<br />

have to sum up over pixels perpendicular to the image gradient in the second equation <strong>of</strong> (6.31). This<br />

perpendicular direction is also a stochastic quantity, but it is not possible to sum up in a stochastic direction.<br />

To overcome this, we use the direction E ( (∇u) ⊥) and neglect the error due to the inaccurate<br />

direction. This is similar to the upwinding problem for stochastic equations [92].<br />

Remark 14. Erdem et al. [49] proposed additional feedback measures for textures and the local<br />

scale. These measures are not included here, but can be generalized in a similar fashion.<br />

6.2.3 Results<br />

In the following, we demonstrate the performance and advantages <strong>of</strong> the stochastic extension <strong>of</strong><br />

the Ambrosio-Tortorelli segmentation approach. We use three data sets which cover a broad range<br />

<strong>of</strong> possible input data. Furthermore, we compare the results <strong>of</strong> the stochastic Ambrosio-Tortorelli<br />

model with the adaptive extension for the spatial dimensions and the stochastic version <strong>of</strong> the edge<br />

linking step. Thus, the organization <strong>of</strong> this section is the following: First, we demonstrate the method<br />

on three data sets. Then we show the results <strong>of</strong> the combination <strong>of</strong> the stochastic method with an<br />

adaptive grid approach for the spatial dimensions. Finally, we demonstrate that the stochastic method<br />

benefits from the idea <strong>of</strong> edge linking [49].<br />

The first input image data set consists <strong>of</strong> M = 5 samples from the artificial “street sequence” [99].<br />

The second data set consists <strong>of</strong> M = 45 image samples from ultrasound (US) imaging <strong>of</strong> a structure in<br />

the forearm, acquired within two seconds. The third data set contains ten images <strong>of</strong> a scene acquired<br />

with a digital camera 2 . Note that we do not consider the street sequence as an image sequence<br />

here. Instead, we use five frames as samples <strong>of</strong> the noisy and uncertain acquisition <strong>of</strong> the same<br />

object. From the samples, we compute the polynomial chaos representation <strong>using</strong> n = 5 (digital<br />

camera), n = 10 (US), respectively n = 4 (street scene) random variables with the method described<br />

2 Thanks to PD Dr. Christoph S. Garbe for providing the data set.<br />

71


Chapter 6 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using Elliptic SPDEs<br />

Figure 6.11: Mean value <strong>of</strong> the three data sets used to demonstrate the stochastic Ambrosio-Tortorelli<br />

method. For the second data set, we denoted image regions the text refers to.<br />

in Section 5.2. The images have a resolution <strong>of</strong> 100 × 100 pixels for the street sequence, 129 × 129<br />

pixels for the US data set, and 513 × 513 pixels for the digital camera data set. We use a polynomial<br />

degree <strong>of</strong> p = 3 for the street scene and the US data and p = 2 for the digital camera sequence. This<br />

leads to a polynomial chaos dimension <strong>of</strong> N = 56 (digital camera), N = 286 (US), and N = 35 (street<br />

scene), respectively. For the reduction <strong>of</strong> the complexity by the GSD, we set K = 6. Furthermore,<br />

we use ν = 0.00075 and k ε = 2.0h in all computations, where h is the grid spacing. To show the<br />

influence <strong>of</strong> the random variables, we used the US data <strong>using</strong> the expected value only (n = 0).<br />

Before we proceed with the presentation and interpretation <strong>of</strong> the results, let us remember the<br />

power <strong>of</strong> the method. For the stochastic image and stochastic phase field it is possible to visualize<br />

the PDF <strong>of</strong> every pixel (see Fig. 6.12), because we compute with this method the coefficients which<br />

describe these random variables in a basis spanned by orthogonal polynomials in random variables.<br />

Street Image Data Set<br />

We use five samples <strong>of</strong> the street sequence to compute the stochastic image. Fig. 6.14 shows the<br />

expected value and the variance <strong>of</strong> the stochastic input image computed <strong>using</strong> the method presented<br />

Figure 6.12: PDF <strong>of</strong> a pixel from the phase field computed from the polynomial chaos expansion <strong>of</strong><br />

the pixel via a sampling approach. Although we use uniform basic random variables for<br />

the polynomial chaos, the resulting random variables have skewed and Gaussian like<br />

distributions due to the use <strong>of</strong> higher order polynomials in the basic random variables.<br />

72


6.2 Ambrosio-Tortorelli <strong>Segmentation</strong> on <strong>Stochastic</strong> <strong>Images</strong><br />

Samples E(φ) Var(φ)<br />

Monte Carlo<br />

GSD<br />

Figure 6.13: <strong>Segmentation</strong> result <strong>of</strong> the street scene. On the left we show the five samples the stochastic<br />

input image is computed from. On the right we compare the results computed via<br />

the GSD method and a Monte Carlo sampling.<br />

in Section 5.2. It is visible from the pictures that the gray value uncertainty is high close to the edges<br />

<strong>of</strong> moving objects. Thus, we expect the highest phase field variance in these regions. The results<br />

depicted in Fig. 6.13 match with these expectations. Indeed, in the region around the wheels <strong>of</strong> the<br />

car and around the right shoulder <strong>of</strong> the person, the edge detection is most influenced by the moving<br />

camera, respectively the varying gray values between the samples at the edges. Also around the<br />

edges in the background, the variance increases due to the moving camera. However, the stochastic<br />

method can detect the edges in the image properly. The result <strong>of</strong> the stochastic method contains<br />

much more information than the deterministic method. The expected value <strong>of</strong> the stochastic method<br />

is comparable to the result <strong>of</strong> the classical method, the stochastic information like chaos coefficients,<br />

variance, etc. are the real benefit <strong>of</strong> the method. Thus, we use the variance, indicating the robustness<br />

<strong>of</strong> the detected edges to get information, which is not available in the classical model.<br />

To verify the intrusive GSD method, we compared the results <strong>of</strong> the GSD implementation with<br />

a simple Monte Carlo method with 10000 sample computations. Fig. 6.13 shows the results and<br />

Figure 6.14: Expected value and variance <strong>of</strong> the stochastic input image <strong>of</strong> the street scene.<br />

73


Chapter 6 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using Elliptic SPDEs<br />

10 random variables mean only<br />

ε = 0.2h, µ = 1/300 ε = 0.4h, µ = 1/300 ε = 0.4h, µ = 1/400 ε = 0.4h, µ = 1/400<br />

E<br />

Image<br />

Var<br />

n/a<br />

E<br />

Phase Field<br />

Var<br />

n/a<br />

Figure 6.15: Mean and variance <strong>of</strong> the image and phase field for varying ε and µ <strong>using</strong> the US data.<br />

For comparison, we added the result from the deterministic method applied on the mean.<br />

reveals that both approaches lead to similar results, but again, the GSD implementation is 100 times<br />

faster than performing Monte Carlo simulation with a suitable number <strong>of</strong> samples.<br />

Ultrasound Samples<br />

The conversion <strong>of</strong> the input samples into the polynomial chaos as described in Section 5.2 leads to<br />

the representation <strong>of</strong> the stochastic US image with 286 coefficients per pixel. Thus, a visualization <strong>of</strong><br />

this stochastic image via stochastic moments like expected value and variance is necessary. Fig. 6.15<br />

shows the expected value and the variance <strong>of</strong> the phase field φ and the smoothed image u for settings<br />

<strong>of</strong> the smoothing coefficient µ and the phase field width ε. The algorithm needs about 100 iterations,<br />

i.e. alternating solutions <strong>of</strong> (6.16) for u and φ, to compute a solution. However, in the first steps, the<br />

convergence is fast and after about 10 iterations, there is no visible difference in u and φ.<br />

From the variance image <strong>of</strong> the phase field, the identification <strong>of</strong> regions where the input distribution<br />

has a strong influence on the segmentation result (areas with high variance) is straightforward.<br />

A benefit <strong>of</strong> the new stochastic edge detection via the phase field φ is that it allows for an identifi-<br />

74


6.2 Ambrosio-Tortorelli <strong>Segmentation</strong> on <strong>Stochastic</strong> <strong>Images</strong><br />

without edge linking edge linking edge linking + adaptive<br />

E<br />

Image<br />

Var<br />

E<br />

Phase Field<br />

Var<br />

Figure 6.16: Comparison <strong>of</strong> the stochastic Ambrosio-Tortorelli model (left column) with the extended<br />

model <strong>using</strong> the edge linking procedure described in Section 2.3.3 (middle column)<br />

and a combination <strong>of</strong> the edge linking and adaptive grid approach (right column).<br />

Note that these results are computed with the same parameter set. The differences in<br />

the results are due to the additional edge linking parameter c only.<br />

75


Chapter 6 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using Elliptic SPDEs<br />

cation <strong>of</strong> edges in a way that is robust with respect to parameter changes. In particular, within the<br />

four regions marked in Fig. 6.11 the expectation <strong>of</strong> the phase field is highly influenced by the choice<br />

<strong>of</strong> µ and ν as visible in Fig. 6.15. The blurred edge at position 1 can be seen in the expectation <strong>of</strong><br />

the phase field only when we use a narrow phase field. In region 2, we have a different situation in<br />

which the edge can be identified <strong>using</strong> a widish phase field. In addition, the edges at positions 3 and<br />

4 can be identified <strong>using</strong> adjusted parameters. However, note that one <strong>of</strong> these edges is not seen in<br />

the expectation <strong>of</strong> φ because <strong>of</strong> the particular choice <strong>of</strong> parameters; a high variance <strong>of</strong> φ indicates<br />

the possible existence <strong>of</strong> an edge. In particular, this is true for the regions 1 and 2.<br />

Moreover, the algorithm estimates the reliability <strong>of</strong> detected edges: A low expected value <strong>of</strong> the<br />

phase field and a low variance indicate that the edge is robust and not influenced by the noise and<br />

uncertainty <strong>of</strong> the acquisition process. This is true for the upper edges <strong>of</strong> the structure. In contrast, a<br />

high variance in regions with a high or low expected value <strong>of</strong> the phase field, e.g. the labeled regions<br />

1–4, indicates regions where the detected edge is sensitive to the noise.<br />

In addition, we can easily extract the distribution <strong>of</strong> the gray values for any pixel location inside<br />

the image and the phase field from the polynomial chaos expansion obtained via the GSD method.<br />

Fig. 6.12 shows the PDF <strong>of</strong> a pixel from the phase field computed via the GSD.<br />

Adaptive Grids<br />

With a combination <strong>of</strong> the stochastic method and the adaptive grid approach described in Section 4.5,<br />

we decrease the memory requirements and increase the performance further. Fig. 6.17 shows the<br />

results <strong>of</strong> the adaptive method and compares them with the results from the uniform grid. For the<br />

computations, we used a threshold for the error indicator <strong>of</strong> ι = 0.005. Thus, elements where the<br />

error indicator S(x) is smaller than ι at every node are not further refined. The choice <strong>of</strong> a suitable<br />

value for the error indicator threshold is important, because too small values lead to unnecessary<br />

fine grids, whereas too high values for the threshold lead to coarse grids even close to edges. This<br />

causes over-shootings in the numerical computation <strong>of</strong> the phase field, i.e. the phase field has no<br />

longer values between zero and one. The results shown in Fig. 6.17 use a maximal level <strong>of</strong> 7, leading<br />

to a uniform grid <strong>of</strong> size 129 × 129 at the beginning. The final adaptive grid depicted on the right<br />

<strong>of</strong> Fig. 6.17 has about 70% fewer degrees <strong>of</strong> freedom, but yields nearly the same results. Fig. 6.17<br />

shows no visible difference between the uniform grid and the adaptive grid solution.<br />

From Fig. 6.17 it is visible that the saturation condition, required to build admissible grids, leads<br />

to an area <strong>of</strong> increasing element size around detected edges. In flat regions, i.e. where the image<br />

gradient magnitude |∇u| is small, the elements are coarser compared to regions close to edges.<br />

Edge Continuity and Edge Consistency<br />

The combination <strong>of</strong> the stochastic Ambrosio-Tortorelli segmentation with an edge linking step is a<br />

great advance on the way to a detection and volumetry <strong>of</strong> objects in stochastic images. Fig. 6.16<br />

shows the results <strong>of</strong> the stochastic extension <strong>of</strong> the edge linking step. We used α = 10, which turns<br />

out to be a good weighting between smoothing out unwanted edges and sharpening regions to get<br />

closed contours, respectively linked edges. We used s = 4, i.e. we used the two neighboring pixels in<br />

the directions perpendicular to the image gradient to compute the feedback measure ζ dc (ξ ).<br />

The figures indicate that the use <strong>of</strong> the modified diffusivity in the image equation <strong>of</strong> (6.16) leads<br />

to a better detection <strong>of</strong> closed contours in the stochastic images. These closed contours lead to<br />

cartoon-like smoothed images, because they avoid the smoothing over undetected edges.<br />

It is possible to combine the edge linking step with the adaptive grid approach, leading to a fast and<br />

accurate extension <strong>of</strong> the initial stochastic Ambrosio-Tortorelli model. The last column <strong>of</strong> Fig. 6.16<br />

shows the result <strong>of</strong> such a combination.<br />

76


6.2 Ambrosio-Tortorelli <strong>Segmentation</strong> on <strong>Stochastic</strong> <strong>Images</strong><br />

full grid adaptive grid full vs. adaptive grid<br />

E<br />

Image<br />

Var<br />

E<br />

Phase Field<br />

Var<br />

Figure 6.17: Comparison <strong>of</strong> the full grid and adaptive grid solution. The full grid and adaptive grid<br />

solution are visually identical, but the computation <strong>of</strong> the adaptive grid solution needs<br />

significantly less DOFs. Thus, it can be applied on high-resolution images.<br />

Conclusion<br />

We presented extensions <strong>of</strong> the random walker and the Ambrosio-Tortorelli model to stochastic images<br />

and applied the methods on different data sets. Especially, the intuitive visualization <strong>of</strong> the<br />

stochastic random walker method via the visualization <strong>of</strong> contour realizations and the objects volume<br />

PDF can be useful to convince the image processing community <strong>of</strong> stochastic modeling.<br />

Furthermore, we presented a detailed theoretical foundation <strong>of</strong> the stochastic Ambrosio-Tortorelli<br />

extension. The availability <strong>of</strong> the theoretical foundation along with the intuitive visualization <strong>of</strong> the<br />

results is the key to a widely accepted method in image processing. The acceleration <strong>of</strong> the algorithm<br />

via an adaptive grid approach and the integration <strong>of</strong> the edge linking step shows the potential <strong>of</strong> the<br />

proposed methods to be extended to the users’ needs easily.<br />

77


Chapter 7<br />

<strong>Stochastic</strong> Level Sets<br />

Level sets are widely used in applications ranging from computer vision [148] over material science<br />

to computer-aided design [138] for the tracking and representation <strong>of</strong> moving interfaces arising<br />

e.g. in the simulation <strong>of</strong> radi<strong>of</strong>requency ablation [13]. Dervieux and Thomasset [40] and Sethian<br />

and Osher [121, 138] introduced level sets in the form used today. The main idea is to embed the<br />

moving interface as the zero level set <strong>of</strong> a higher-dimensional function φ. The moving boundary is<br />

then equivalent to a propagation <strong>of</strong> the level sets <strong>of</strong> the function φ over time. The actual position <strong>of</strong><br />

the boundary at time t is reconstructed from the function φ by tracking the zero level set at time t.<br />

Level sets are used for the segmentation <strong>of</strong> images as well. They are more flexible in comparison<br />

to a parametrization <strong>of</strong> the boundary used e.g. for snakes [76]. In addition, advanced segmentation<br />

methods like geodesic active contours [30, 82], an energy minimization method, are based on<br />

level sets.<br />

When we try to combine a level set based segmentation approach with stochastic images, we end<br />

up with a stochastic velocity for the level set propagation, i.e. we have to solve a hyperbolic SPDE.<br />

The development <strong>of</strong> numerical methods for hyperbolic SPDEs is an active research field. To the<br />

best <strong>of</strong> the authors knowledge, there is no method available in the literature that can be applied to<br />

the stochastic level set equation. The use <strong>of</strong> classical methods, like upwinding schemes [138], is<br />

not possible, because they are based on the sign <strong>of</strong> the propagation speed, which is in the stochastic<br />

context a random variable, too. Thus, we use a parabolic approximation <strong>of</strong> the level set equation.<br />

This enables us to use the methods developed in the previous chapters.<br />

Due to the importance <strong>of</strong> the level set equation in other applications besides the segmentation <strong>of</strong><br />

images, this chapter is split into two parts. First, we present the derivation <strong>of</strong> the parabolic approximation<br />

<strong>of</strong> the stochastic level set equation along with the numerical discretization. Furthermore,<br />

we present numerical tests showing the applicability <strong>of</strong> the discretization. The second part <strong>of</strong> this<br />

chapter deals with the application <strong>of</strong> the stochastic level set equation for image segmentation. We<br />

introduce stochastic extensions <strong>of</strong> three widely used segmentation methods based on the level set<br />

equation: gradient-based segmentation, geodesic active contours, and Chan-Vese segmentation.<br />

7.1 Derivation <strong>of</strong> a <strong>Stochastic</strong> Level Set Equation<br />

The discretization <strong>of</strong> the classical level set equation is based on techniques for the discretization <strong>of</strong><br />

hyperbolic conservation laws. The discretization <strong>of</strong> hyperbolic SPDEs is still a challenging task. To<br />

the best <strong>of</strong> the authors knowledge, there are two possibilities, which are less accurate [92] or timeconsuming<br />

[147]. Thus, we focus on a parabolic approximation <strong>of</strong> the level set equation to avoid<br />

the numerical problems related to the hyperbolic level set version. The parabolic stochastic level set<br />

equation is based on the work <strong>of</strong> Sun and Beckermann [143] for the classical level set equation. The<br />

stochastic level set equation is derived from the equation<br />

φ(y(t,ω),t,ω) = 0 almost sure in Ω , (7.1)<br />

79


Chapter 7 <strong>Stochastic</strong> Level Sets<br />

where t is the time, ω a stochastic event, and y(t,ω) the path <strong>of</strong> a particle on the interface. Using the<br />

chain rule, we get the stochastic version <strong>of</strong> the advection equation<br />

φ t (t,x,ω) + v(t,x,ω) · ∇φ(t,x,ω) = 0 , (7.2)<br />

where v = ∂y(t,ω)<br />

∂t<br />

is the speed <strong>of</strong> the level set propagation. The speed decomposes in a component in<br />

the normal direction N and in the tangential directions T <strong>of</strong> the interface:<br />

where v N and v T are<br />

v(t,x,ω) = v N (t,x,ω) + v T (t,x,ω) , (7.3)<br />

v N (t,x,ω) = (v(t,x,ω) · N(t,x,ω))N(t,x,ω) resp. v T (t,x,ω) = v(t,x,ω) − v N (t,x,ω) . (7.4)<br />

Note that the decomposition is dependent on the stochastic event ω, because for every realization<br />

ω ∈ Ω <strong>of</strong> the level set φ we get a different normal N(t,x,ω) and a different decomposition <strong>of</strong> the<br />

stochastic quantity v(t,x,ω). Substituting (7.3) and (7.4) into (7.2) and <strong>using</strong> the relations<br />

v T (t,x,ω) · ∇φ(t,x,ω) = 0 and v N (t,x,ω) · ∇φ(t,x,ω) = v n (t,x,ω)|∇φ(t,x,ω)| , (7.5)<br />

where v n is the speed in the normal direction, yields the stochastic extension <strong>of</strong> the level set equation:<br />

φ t (t,x,ω) + v n (t,x,ω)|∇φ(t,x,ω)| = 0 . (7.6)<br />

As already mentioned, the discretization <strong>of</strong> this deterministic equation uses methods for hyperbolic<br />

conservation laws, e.g. upwinding schemes. To the best <strong>of</strong> the authors knowledge, there is no accurate<br />

and fast upwinding scheme for SPDEs available. To avoid the use <strong>of</strong> a numerical upwinding<br />

scheme for hyperbolic SPDEs, we modify the stochastic level set equation in the spirit <strong>of</strong> Sun and<br />

Beckermann [143]. We start with a decomposition <strong>of</strong> the speed v n into a component independent and<br />

a component dependent on the interface curvature κ:<br />

v n (t,x,ω) = a(t,x,ω) − b(x,t,ω)κ(t,x,ω) . (7.7)<br />

The curvature κ is expressed <strong>using</strong> the level set φ, this is a standard approach for deterministic level<br />

sets [138], and rewritten <strong>using</strong> the quotient rule:<br />

( )<br />

∇φ(t,x,ω)<br />

κ(t,x,ω) = ∇ · N(t,x,ω) = ∇ ·<br />

|∇φ(t,x,ω)|<br />

(<br />

) (7.8)<br />

1<br />

(∇φ(t,x,ω)) · ∇(|∇φ(t,x,ω)|)<br />

=<br />

∆φ(t,x,ω) − .<br />

|∇φ(t,x,ω)|<br />

|∇φ(t,x,ω)|<br />

The previous modeling is valid for sufficiently smooth level set functions. If we prescribe a special<br />

behavior <strong>of</strong> the level set in the normal direction <strong>of</strong> the level set, quantities like the gradient or the<br />

curvature can be computed easily. For the special choice <strong>of</strong> the level set function<br />

( ) n(t,x,ω)<br />

φ(t,x,ω) = −tanh √ , (7.9)<br />

2W<br />

where n is the distance to the interface in the normal direction and W ∈ IR an additional parameter<br />

controlling the width <strong>of</strong> the tangential pr<strong>of</strong>ile, we get for the norm <strong>of</strong> the gradient:<br />

|∇φ(t,x,ω)| = − ∂φ(t,x,ω)<br />

∂n<br />

This is because the derivative <strong>of</strong> the hyperbolic tangent is<br />

= 1 − φ(t,x,ω)2 √<br />

2W<br />

. (7.10)<br />

(tanhx) ′ = 1 − tanh 2 x . (7.11)<br />

80


7.1 Derivation <strong>of</strong> a <strong>Stochastic</strong> Level Set Equation<br />

Remark 15. Prescribing a special behavior <strong>of</strong> the level set function, the hyperbolic tangent pr<strong>of</strong>ile,<br />

is a standard technique in the level set context. A typical choice for classical level sets is the signed<br />

distance function, i.e. to ensure that |∇φ| = 1 (see [138]).<br />

The last term in (7.8) is the second derivative <strong>of</strong> φ in the normal direction, i.e.<br />

(∇φ(t,x,ω)) · ∇(|∇φ(t,x,ω)|)<br />

|∇φ(t,x,ω)|<br />

= ∂ 2 φ(t,x,ω)<br />

∂n 2 , (7.12)<br />

we use the special pr<strong>of</strong>ile <strong>of</strong> the level set from (7.9), the derivation rule for the hyperbolic tangent<br />

and the chain rule to simplify the expression:<br />

∂ 2 φ(t,x,ω)<br />

∂n 2 = − φ(t,x,ω)( 1 − φ(t,x,ω) 2)<br />

W 2 . (7.13)<br />

Substituting this into (7.8), we get<br />

κ(t,x,ω) =<br />

1<br />

(∆φ(t,x,ω) + φ(t,x,ω)( 1 − φ(t,x,ω) 2) )<br />

|∇φ(t,x,ω)|<br />

W 2<br />

. (7.14)<br />

Now we are able to substitute the findings from (7.3) to (7.14) into the level set equation (7.6). First,<br />

we substitute the decomposition <strong>of</strong> the speed from (7.3) and the expressions for the level set gradient<br />

from (7.10) and the curvature from (7.14) into the level set equation (7.6):<br />

φ t (t,x,ω) + a 1 − φ(t,x,ω)2 √<br />

2W<br />

= b<br />

(∆φ(t,x,ω) + φ(t,x,ω)( 1 − φ(t,x,ω) 2) )<br />

W 2<br />

. (7.15)<br />

This equation is parabolic for b > 0 and the hyperbolic term a|∇φ| is converted into a nonlinear term<br />

in φ. Sun and Beckermann [143] derived the deterministic equivalent <strong>of</strong> (7.15). They showed that<br />

it is possible to implement (7.15), but the resulting phase field has the prescribed tangential pr<strong>of</strong>ile<br />

across the interface. Hence, it is not a signed distance function, which is preferred in applications.<br />

A possibility to sustain the parabolic term in the absence <strong>of</strong> a curvature dependent speed, i.e. if<br />

b = 0, is to subtract the curvature from (7.8) from the reformulated level set equation (7.15):<br />

φ t + a 1 − φ 2 (<br />

√ = b ∆φ + φ(1 − φ 2 )<br />

2W W 2 − |∇φ|∇ · ∇φ )<br />

|∇φ|<br />

. (7.16)<br />

This subtraction is based on the idea <strong>of</strong> the counter term approach developed by Folch et al. [51] and<br />

should not be confounded with setting b = 0, because the first term on the right hand side conserves<br />

the tangential pr<strong>of</strong>ile <strong>of</strong> the level set. If we set b = 0, the equation moves an arbitrary shaped level set,<br />

instead <strong>of</strong> producing the tangential shaped level set. Note that in (7.16) and the following equations<br />

we use the notation φ for the phase field instead <strong>of</strong> writing φ(t,x,ω) to ease notation. Of course,<br />

the phase field stays dependent on time t, spatial position x and stochastic event ω. Furthermore, we<br />

omit denoting the dependence <strong>of</strong> other quantities from t, x, and ω, when this is obvious.<br />

To summarize the modifications <strong>of</strong> the level set equation: We have a stochastic parabolic level<br />

set equation. When the speed is curvature dependent, we use (7.15). Otherwise, we use (7.16).<br />

Furthermore, the hyperbolic term a|∇φ| becomes a term nonlinear in φ.<br />

In a last step, we use the nonlinear preconditioning technique from [56]. With the substitution<br />

(<br />

φ = −tanh ψ/( √ )<br />

2W)<br />

, (7.17)<br />

81


Chapter 7 <strong>Stochastic</strong> Level Sets<br />

Figure 7.1: <strong>Stochastic</strong> level sets do not have a fixed position where φ(x) = 0. Instead, there is a band<br />

with positive probability that the level set is equal to zero, i.e. the position <strong>of</strong> the zero<br />

level set is random and it is possible to estimate the PDF <strong>of</strong> the interface location in the<br />

normal direction <strong>of</strong> the expected value <strong>of</strong> the interface (lower right corner).<br />

which ensures that ψ is a signed distance function to the interface because <strong>of</strong> φ = −tanh √<br />

2W<br />

, we<br />

get the final version <strong>of</strong> the stochastic level set equation<br />

( (<br />

1 − |∇ψ|<br />

2 )√ 2<br />

ψ t + a|∇ψ| = b ∆ψ +<br />

tanh<br />

ψ<br />

( ) ) ∇ψ<br />

√ − |∇ψ|∇ ·<br />

, (7.18)<br />

W<br />

2W |∇ψ|<br />

where we omitted the dependence <strong>of</strong> the function ψ on time t, spatial position x, and random event ω.<br />

Following [143], where the deterministic equivalent <strong>of</strong> this equation was derived, the right hand side<br />

<strong>of</strong> this function serves as an integrated reinitialization scheme for the level set ψ. Thus, further<br />

reinitialization is not required for deterministic level sets.<br />

7.1.1 Interpretation <strong>of</strong> <strong>Stochastic</strong> Level Sets<br />

Having (7.18) at hand, we have to interpret the result <strong>of</strong> the level set motion with random speed.<br />

Due to the random variable/field that controls the speed <strong>of</strong> the level set motion the position <strong>of</strong> the<br />

zero (and all other) level sets is a random quantity, too. A possibility to estimate the influence <strong>of</strong> the<br />

random speed component on the level set motion is to calculate the probability that the zero level set<br />

is at a specific position. Furthermore, we can calculate the whole band with positive probability that<br />

the zero level set is located there, i.e. where<br />

P(φ(x) = 0) > 0 (7.19)<br />

holds. In the normal direction <strong>of</strong> the expected value E(φ) = 0 <strong>of</strong> the zero level set location, we can<br />

estimate the PDF <strong>of</strong> the interface position (see Fig. 7.1).<br />

Remark 16. Using Gaussian random variables in combination with stochastic level sets, we end up<br />

with a nonzero probability for the interface location in the whole domain. This is due to the infinite<br />

support <strong>of</strong> Gaussian random variables. Thus, we limit the following investigations to a polynomial<br />

chaos in uniform random variables. We denote a random variable X uniformly distributed in the<br />

interval [a,b] by X ∼ U [a,b]. Uniform random variables have a compact support, leading to a band<br />

with finite thickness for the potential interface location.<br />

n<br />

82


7.2 Discretization <strong>of</strong> the <strong>Stochastic</strong> Level Set Equation<br />

<strong>Stochastic</strong> Signed Distance Functions<br />

It is desirable to use signed distance functions as level sets in the stochastic context, too. A stochastic<br />

signed distance function has to be a classical signed distance function for every realization ω ∈ Ω.<br />

Theorem 7.1. A stochastic signed distance function fulfills E(|∇φ|) = 1 and Var(|∇φ|) = 0.<br />

Pro<strong>of</strong>. The first property is ensured by<br />

the second property by<br />

∫<br />

∫<br />

E(|∇φ(x)|) = |∇φ(x,ω)|dω = 1 dω = 1 , (7.20)<br />

Ω<br />

Ω<br />

∫<br />

∫<br />

Var(|∇φ(x)|) = (|∇φ(x,ω)| − 1) 2 dω = 0 dω = 0 . (7.21)<br />

Ω<br />

Ω<br />

7.2 Discretization <strong>of</strong> the <strong>Stochastic</strong> Level Set Equation<br />

For the numerical tests, we discretize (7.18) <strong>using</strong> the explicit Euler scheme for the discretization <strong>of</strong><br />

the time derivative. The time discrete version <strong>of</strong> (7.18) is<br />

(<br />

( √<br />

2<br />

ψ t+τ = ψ t +τ −a|∇ψ t | + b ∆ψ t +<br />

W<br />

(<br />

1 − |∇ψ t | 2) ( ) ))<br />

tanh √ ψt<br />

∇ψ<br />

− |∇ψ t t<br />

|∇ ·<br />

2W |∇ψ t , (7.22)<br />

|<br />

where ψ t is the phase field at time t. The spatial discretization is done via a uniform grid like in<br />

Section 5.1. The stochastic part is discretized <strong>using</strong> the polynomial chaos. Thus, we have to build<br />

numerical schemes for the gradient norm, the curvature, and the hyperbolic tangent that deal with<br />

polynomial chaos expansions.<br />

Gradient Norm and Laplacian<br />

The gradient norm is computed <strong>using</strong> finite difference schemes. The directional derivatives are computed<br />

<strong>using</strong> central differences in the interior <strong>of</strong> the domain and forward resp. backward differences<br />

at the domain boundary. The necessary computations <strong>of</strong> the square and the square root are performed<br />

<strong>using</strong> the methods from Section 3.3.<br />

The Laplacian is computed as the sum <strong>of</strong> the second directional derivatives, which we compute<br />

<strong>using</strong> central differences in the interior and forward resp. backward differences at the boundary.<br />

Hyperbolic Tangent<br />

We compute the hyperbolic tangent <strong>using</strong> tanhx = 1 − 2(exp(2x) + 1) −1 . Thus, we use the computation<br />

<strong>of</strong> the exponential function in the polynomial chaos from Section 3.3 and other methods from<br />

there, i.e. we compute the Galerkin projection <strong>of</strong> the hyperbolic tangent on the polynomial chaos.<br />

Curvature<br />

The computation <strong>of</strong> the curvature is the most critical process for the computation <strong>of</strong> the update,<br />

because the update is done in the whole domain, not in a narrow band around the zero level set. This<br />

makes it necessary to compute a stable curvature even in regions with a high curvature. These regions<br />

arise in simple settings, e.g. when the level set is initialized as a circle. The curvature in the midpoint<br />

<strong>of</strong> the circle goes to infinity. We use a method for the stable curvature computation proposed by Sun<br />

83


Chapter 7 <strong>Stochastic</strong> Level Sets<br />

and Beckermann [143] based on an idea by Echebarria et al. [46]. It is given by<br />

( ) ∇ψ<br />

∇ ·<br />

|∇ψ|<br />

⎛<br />

≈ 1 ⎝<br />

h<br />

ψ i+1, j − ψ i, j<br />

√<br />

(ψ i+1, j − ψ i, j ) 2 + (ψ i+1, j+1 + ψ i, j+1 − ψ i+1, j−1 − ψ i, j−1 ) 2 /16<br />

ψ i, j − ψ i−1, j<br />

− √<br />

(ψ i, j − ψ i−1, j ) 2 + (ψ i−1, j+1 + ψ i, j+1 − ψ i−1, j−1 − ψ i, j−1 ) 2 /16<br />

+ √<br />

ψ i, j+1 − ψ i, j<br />

(ψ i, j+1 − ψ i, j ) 2 + (ψ i+1, j+1 + ψ i+1, j − ψ i−1, j+1 − ψ i−1, j ) 2 /16<br />

⎞<br />

−√<br />

ψ i, j − ψ i, j−1<br />

⎠ .<br />

(ψ i, j − ψ i, j−1 ) 2 + (ψ i+1, j−1 + ψ i+1, j − ψ i−1, j−1 − ψ i−1, j ) 2 /16<br />

(7.23)<br />

Due to the independence <strong>of</strong> the update for the spatial positions, this finite difference scheme can be<br />

parallelized on multiple processor cores easily.<br />

Remark 17. Due to the hyperbolic tangent pr<strong>of</strong>ile <strong>of</strong> the level sets across the interface, we have<br />

to respect a condition on the maximal curvature <strong>of</strong> the represented object. For a high curvature,<br />

the hyperbolic tangent pr<strong>of</strong>iles overlap for points on the interface. This leads to instabilities <strong>of</strong> the<br />

numerical schemes for the discretization.<br />

7.3 Reinitialization <strong>of</strong> <strong>Stochastic</strong> Level Sets<br />

The right hand side <strong>of</strong> (7.18) is an integrated reinitialization <strong>of</strong> the level set function. Following<br />

[143], this reinitialization is sufficient to get accurate results for deterministic level sets. When<br />

<strong>using</strong> a stochastic velocity, we have to reinitialize all polynomial chaos coefficients, which are on different<br />

scales. Typically, the first coefficient, the expected value, is orders <strong>of</strong> magnitude bigger than<br />

the remaining coefficients. Furthermore, the coefficients <strong>of</strong> polynomials in uncoupled random variables<br />

are close to zero. During the numerical experiments, we observed that the reinitialization via<br />

(7.18) is not sufficient. Thus, we need an additional reinitialization to get accurate stochastic results.<br />

The classical reinitialization methods for level sets are not applicable in the stochastic context.<br />

The Fast Marching method [138] is based on an upwinding scheme. As discussed in Section 7.1,<br />

a stochastic upwinding scheme is not available. Iterative reinitialization via φ t = sign(φ)(1 − |∇φ|)<br />

is not possible because the signature <strong>of</strong> a stochastic quantity is not well-defined. The equation for<br />

energy minimization [90], i.e. α|E(|∇φ| − 1) 2 | + β|Var(|∇φ|)| → min, is unstable if φ converges to<br />

the stochastic signed distance function.<br />

To get a working reinitialization scheme for stochastic level sets, we use a modification <strong>of</strong> the<br />

stochastic level set equation (7.18). As already mentioned, the right hand side <strong>of</strong> this function is<br />

an integrated reinitialization. We use this equation, set the speed to zero, i.e. a = 0, and solve the<br />

equation for artificial time T . Doing this, the reinitialization equation is<br />

( (<br />

1 − |∇ψ|<br />

2 )√ 2<br />

ψ t = b ∆ψ +<br />

tanh<br />

ψ<br />

( ) ) ∇ψ<br />

√ − |∇ψ|∇ ·<br />

. (7.24)<br />

W<br />

2W |∇ψ|<br />

In all numerical experiments, we apply this reinitialization equation. We use for every time step <strong>of</strong><br />

(7.18) ten reinitialization time steps <strong>of</strong> (7.24) with a time step size for the reinitialization <strong>of</strong> 0.5τ,<br />

where τ is the time step size <strong>of</strong> the original problem.<br />

84


7.4 Numerical Experiments<br />

cosine inward<br />

cosine outward<br />

E Var E Var<br />

PC<br />

SC<br />

MC<br />

MCL<br />

Figure 7.2: Comparison <strong>of</strong> expected value and variance <strong>of</strong> the resulting phase field for the cosine<br />

test <strong>of</strong> (7.18) <strong>using</strong> the polynomial chaos (PC), stochastic collocation (SC), Monte Carlo<br />

simulation (MC), and Monte Carlo simulation <strong>of</strong> the original level set equation (MCL).<br />

7.4 Numerical Experiments<br />

In this section, we present numerical experiments for the verification <strong>of</strong> the proposed algorithm and<br />

for the implementation <strong>of</strong> the algorithm. To validate the intrusive implementation in the polynomial<br />

chaos, we verify the results with Monte Carlo experiments and a stochastic collocation approach.<br />

To show that the phase field equation is comparable with the native level set equation, we added a<br />

Monte Carlo experiment based on the original level set equation<br />

φ t + a|∇φ| = 0 . (7.25)<br />

We are able to compare four implementations <strong>of</strong> the stochastic level set evolution: The intrusive<br />

implementation <strong>of</strong> the preconditioned phase field in the polynomial chaos (PC), a stochastic collocation<br />

approach based on the preconditioned phase field (SC), a Monte Carlo simulation <strong>of</strong> the<br />

85


Chapter 7 <strong>Stochastic</strong> Level Sets<br />

rarefaction fan<br />

shock<br />

E Var E Var<br />

PC<br />

SC<br />

MC<br />

MCL<br />

Figure 7.3: Comparison <strong>of</strong> the expected value and variance <strong>of</strong> the resulting phase field for the rarefaction<br />

fan and the shock, two classical tests for level set propagation. The figure shows<br />

the comparison <strong>of</strong> the four discretizations <strong>of</strong> the stochastic phase field equation.<br />

preconditioned phase field (MC), and a Monte Carlo simulation <strong>of</strong> the original level set implementation<br />

(MCL). The comparison is performed on two typical tests for level set evolution, the evolution<br />

<strong>of</strong> a cosine curve in the inward and outward direction and the evolution <strong>of</strong> an edge <strong>of</strong> a square in the<br />

inward and outward direction. Furthermore, we demonstrate the extension <strong>of</strong> the proposed method<br />

to three spatial dimensions on the Stanford bunny data set [149]. In contrast to other publications<br />

dealing with mean curvature motion [138], we use the Stanford bunny and apply the preconditioned<br />

phase field equation with stochastic speed on it. In all numerical experiments, we set W = 2.5h,<br />

where h is the grid spacing and in the absence <strong>of</strong> a curvature dependent speed, we set b = 1.25h.<br />

For the evolution <strong>of</strong> the cosine (see Fig. 7.2), the challenge is the development <strong>of</strong> a shock [138],<br />

when the curve moves inward. Due to the stochastic velocity, we used a uniformly distributed speed a<br />

with E(a) = 1.0 and Var(a) = 0.04, i.e. a ∼ U [1 − 0.2 √ 3,1 + 0.2 √ 3], the position <strong>of</strong> the shock is<br />

uncertain, and the discretization has to be adequate in a vicinity <strong>of</strong> the possible shock positions.<br />

For the numerical experiments, we use a spatial resolution <strong>of</strong> 129 × 129, a polynomial chaos in one<br />

86


7.4 Numerical Experiments<br />

Figure 7.4: Expected value color-coded by the variance for the Stanford bunny after shrinkage under<br />

an uncertain speed in the normal direction. Red indicates regions with a high variance<br />

and green regions with low variance. In addition, we show one slice <strong>of</strong> the variance.<br />

random variable with order two and apply 30 steps with step size 0.1h. Furthermore, we computed<br />

20 time steps <strong>of</strong> the reinitialization equation (7.24) with step size 0.2h after every time step. The<br />

polynomial chaos coefficients <strong>of</strong> the speed are set to a 1 = 1, a 2 = 0.2, and a 3 = 0, such that the<br />

expansion fulfills E(a) = 1 and Var(a) = 0.04.<br />

It is visible from Fig. 7.2 that the methods based on the stochastic preconditioned phase field<br />

formulation lead to the same results. Only the discretization <strong>of</strong> the level set method leads to deviating<br />

results. This is due to the reinitialization <strong>of</strong> the level set via Fast Marching [138]. The Fast Marching<br />

method assumes that the level set values at a grid point near the interface, the trial nodes (see [138])<br />

is the signed distance to the interface. This is not true in the presence <strong>of</strong> a shock due to the crossing<br />

<strong>of</strong> the zero level set. For deterministic level sets this error can be neglected, even for stochastic level<br />

sets the expected value is accurate. However, for the stochastic part <strong>of</strong> the solution (the variance in<br />

Fig. 7.2), which is orders <strong>of</strong> magnitude smaller than the expected value, this error becomes relevant.<br />

Thus, it is more precise to use the reinitialization via (7.24) in the presence <strong>of</strong> a shock.<br />

When the interface moves outward, we have a rarefaction fan (see [138]), because one point on<br />

the zero level set is the closest point for multiple points away from the interface. The same problem<br />

as for the shock arises and again, the reinitialization via (7.24) is the better method.<br />

The second test is the evolution <strong>of</strong> one edge <strong>of</strong> a square in the inward and outward direction<br />

depicted in Fig. 7.3. Again, we have the development <strong>of</strong> a shock when the curve moves inward and<br />

<strong>of</strong> a rarefaction fan when the curve moves outward.<br />

The last test is the contraction <strong>of</strong> the Stanford bunny under an uncertain velocity. Again, we used<br />

the speed E(a) = 1.0 and Var(a) = 0.04 from the previous tests. Fig. 7.4 shows the results. For the<br />

Stanford bunny, we have the evolution <strong>of</strong> a 3D object. We use a method for the visualization <strong>of</strong> 3D<br />

stochastic images from Section 5.4 by visualizing the expected value color-coded by the variance.<br />

As expected, we see a high variance <strong>of</strong> the contour in regions with high curvature. This is due to the<br />

development <strong>of</strong> shocks, when the contour moves inside these regions.<br />

87


Chapter 7 <strong>Stochastic</strong> Level Sets<br />

Figure 7.5: Mean <strong>of</strong> the CT data set (left) and the liver data set (right) for the segmentation test.<br />

7.5 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using <strong>Stochastic</strong> Level Sets<br />

Level set evolution with an uncertain speed can be useful in applications, e.g. in physical applications<br />

where level sets track the interface between materials. Often, the material parameters are not known<br />

exactly and the interface speed can be modeled dependent on the random material parameters.<br />

The focus <strong>of</strong> this thesis is on the application <strong>of</strong> segmentation methods. This is why we use this new<br />

concept for stochastic level sets at the moment for segmentation only. The author used level sets for<br />

the modeling <strong>of</strong> physical effects, the evaporation <strong>of</strong> water during radi<strong>of</strong>requency ablation [10,11,13].<br />

It is possible to use stochastic level sets in this context, because the material parameters can be<br />

modeled as random variables [87, 128] which leads to an uncertain interface speed.<br />

For segmentation, we investigate three segmentation methods based on level sets for stochastic<br />

extensions: gradient-based segmentation, geodesic active contours, and Chan-Vese segmentation.<br />

Other segmentation methods based on level sets can also be suitable for stochastic extensions.<br />

7.5.1 Gradient-Based <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong><br />

Gradient-based segmentation <strong>of</strong> an image u : D → IR is introduced in Section 2.4.2 and given by<br />

φ t + v(1 − bκ)|∇φ| = 0 , (7.26)<br />

where the function v is called stopping function, because this function controls the stopping <strong>of</strong> the<br />

level set evolution at the desired boundaries. Often, the function v is given by<br />

v = 1/(1 + |∇u|) , (7.27)<br />

where u is the image to segment. Typically, the level set is initialized as the signed distance function<br />

<strong>of</strong> a small circle inside the object to segment. There is no theoretical justification <strong>of</strong> this method<br />

besides the observation that the level set speed is close to zero at sharp edges due to the reciprocal<br />

dependence between image gradient and speed. We replace the classical image u(x) by a stochastic<br />

image u(x,ω). The equation for stochastic gradient-based segmentation is<br />

and the speed is a stochastic quantity, too:<br />

φ t (t,x,ω) + v(t,x,ω)(1 − bκ(t,x,ω))|∇φ(t,x,ω)| = 0 (7.28)<br />

v(t,x,ω) =<br />

1<br />

1 + |∇u(t,x,ω)|<br />

. (7.29)<br />

88


7.5 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using <strong>Stochastic</strong> Level Sets<br />

Figure 7.6: Left: Mean contour during the evolution <strong>of</strong> the stochastic level set. The iso-contours<br />

are drawn on the variance image <strong>of</strong> the final, magenta contour. The contour detection is<br />

influenced by the image noise on the bottom and the right <strong>of</strong> the object (high variance).<br />

Right: Contour realizations <strong>of</strong> the stochastic gradient-based segmentation <strong>of</strong> the CT data.<br />

This method can be implemented by <strong>using</strong> the stochastic preconditioned phase field implementation<br />

introduced in the last section by rearranging the equation to<br />

where the stochastic speed ṽ is<br />

φ t (t,x,ω) + ṽ(t,x,ω)|∇φ(t,x,ω)| = 0 , (7.30)<br />

ṽ(t,x,ω) = 1 − bκ(t,x,ω)<br />

1 + |∇φ(t,x,ω)|<br />

. (7.31)<br />

Using the decomposition into curvature dependent and independent parts, we end up with<br />

Numerical Results<br />

φ t +<br />

1<br />

1 + |∇φ| |∇φ| − b<br />

κ|∇φ| = 0 . (7.32)<br />

1 + |∇φ|<br />

For the presentation <strong>of</strong> the results <strong>of</strong> the gradient-based segmentation <strong>of</strong> stochastic images we use<br />

two data sets. The first data set consists <strong>of</strong> 289 reconstructions <strong>of</strong> a CT data set with 100 × 100<br />

pixels. Section 5.2.2 gives details about the generation <strong>of</strong> the reconstructions. These reconstructions<br />

are treated as independent realizations <strong>of</strong> a stochastic image and the polynomial chaos expansion <strong>of</strong><br />

the stochastic image is calculated <strong>using</strong> the methods from Section 5.2. The second data set consists<br />

<strong>of</strong> a liver mask embedded into a 129 × 129 pixel image with a varying gradient strength to the<br />

background. This image is corrupted by uniform noise and 25 samples, i.e. noise realizations, <strong>of</strong> this<br />

image are treated as input for the generation <strong>of</strong> the stochastic image. In both data sets, the generated<br />

stochastic image contains two random variables, and we use a polynomial degree <strong>of</strong> two, i.e. n = 2<br />

and p = 2. Fig. 7.5 shows the expected value <strong>of</strong> the stochastic image <strong>of</strong> both data sets. The parameter<br />

b is set to the grid spacing h and the level set is initialized as a circle centered in the center <strong>of</strong> the<br />

image with a radius <strong>of</strong> 0.15 <strong>of</strong> the image width.<br />

Fig. 7.6 shows the expected value during the level set evolution (the colored contours) and the<br />

variance after 280 iterations <strong>of</strong> the gradient-based segmentation with time step τ = 0.2h <strong>of</strong> the second<br />

data set. It shows the typical behavior <strong>of</strong> a rapid propagation <strong>of</strong> the level set towards the object<br />

boundary and the influence <strong>of</strong> the stopping function that tries to stop the evolution at the boundary.<br />

89


Chapter 7 <strong>Stochastic</strong> Level Sets<br />

MC SC PC<br />

E<br />

Var<br />

Figure 7.7: Resulting image with the expected value <strong>of</strong> the contour (red) <strong>of</strong> the segmented object<br />

and the phase field variance with the expected value <strong>of</strong> the contour for gradient-based<br />

segmentation <strong>of</strong> a stochastic CT image. The variance is constant in the normal direction<br />

<strong>of</strong> the expected value <strong>of</strong> zero level set.<br />

In Fig. 7.6, we depicted realizations <strong>of</strong> the stochastic contour encoded in the stochastic result <strong>of</strong><br />

the segmentation <strong>of</strong> the first data set. The figure shows that the noise in the input image influences<br />

the segmentation in regions with a low gradient, i.e. in the bone regions <strong>of</strong> the head phantom. In<br />

regions where the level set has not entered the bone or in regions where the evolution reached the<br />

outer bone boundary, the segmentation is more stable with respect to noise. This is visible from the<br />

realizations <strong>of</strong> the contour lines, which are close together in these regions.<br />

A comparison <strong>of</strong> the intrusive implementation <strong>using</strong> the polynomial chaos expansion and the sampling<br />

based implementations <strong>using</strong> Monte Carlo simulation and stochastic collocation is depicted in<br />

Fig. 7.7 and shows a good consistency <strong>of</strong> the implementations. The expected value is visually at the<br />

same position for the three methods, and the variance is similar, too.<br />

7.5.2 <strong>Stochastic</strong> Geodesic Active Contours<br />

Geodesic active contours try to minimize the energy <strong>of</strong> a curve (cf. Section 2.4.3). For a stochastic<br />

curve C(q,ω) : [0,1] × Ω → IR 2 and a stochastic edge indicator g(x,ω) : IR × Ω → IR the expected<br />

value <strong>of</strong> the geodesic curve energy is<br />

∫ ∫ 1<br />

∫ ∫ 1<br />

E(B(C)) = βg u (|∇u(C(q,ω))|)dqdω + α|C ′ (q,ω)|dqdω . (7.33)<br />

Ω 0<br />

Ω 0<br />

This energy tries to minimize the expected value <strong>of</strong> the curve length ∫ Ω<br />

∫ 1<br />

0 |C′ (q,ω)|dqdω weighted<br />

by the edge indicator g, i.e. the functional is minimal when we found a short path along an edge<br />

90


7.5 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using <strong>Stochastic</strong> Level Sets<br />

MC SC PC<br />

E<br />

Var<br />

Figure 7.8: Mean and variance <strong>of</strong> the stochastic geodesic active contour segmentation <strong>of</strong> the stochastic<br />

CT data set. The variance is constant in the normal direction <strong>of</strong> the zero level set.<br />

inside the image. Typically, the edge indicator is<br />

g u =<br />

1<br />

(1 − εκ) , (7.34)<br />

1 + |∇G σ ∗ u| p<br />

where G σ is a Gaussian smoothing kernel with width σ and p ∈ {1,2}. Computing the stochastic<br />

Euler-Lagrange equation as necessary condition for a minimum <strong>of</strong> the function is done in the same<br />

fashion as in [30, 82], but we have to respect the outer integration over Ω. We end up with the<br />

stochastic Euler-Lagrange equation<br />

φ t (t,x,ω) = g u (t,x,ω)β|∇φ(t,x,ω)| − α∇g u (t,x,ω)∇φ(t,x,ω) + εκ|∇φ(t,x,ω)| , (7.35)<br />

which is analog to the deterministic one. The parameters α,β, and γ can be freely chosen to optimize<br />

the segmentation result. The meaning <strong>of</strong> the parameters is the following:<br />

• α: The parameter α controls the attraction <strong>of</strong> the minima <strong>of</strong> the edge indicator g u when it is<br />

positive. Otherwise, the level set is pushed away from the minima.<br />

• β: The parameter β controls the shrinkage or expansion <strong>of</strong> the level set. A negative value <strong>of</strong> β<br />

leads to a shrinkage <strong>of</strong> the level set and positive β to an expansion <strong>of</strong> the level set. Thus, this<br />

parameter controls whether the initial level set is inside or outside <strong>of</strong> the desired contour.<br />

• ε: The parameter ε acts as a weighting term for the curvature smoothing.<br />

The stochastic geodesic active contour level set equation is discretized <strong>using</strong> the methods presented<br />

in Section 3.3 and by <strong>using</strong> the stochastic preconditioned phase field equation.<br />

91


Chapter 7 <strong>Stochastic</strong> Level Sets<br />

Figure 7.9: Left: Evolution <strong>of</strong> the expected value contour <strong>of</strong> the stochastic geodesic active contour<br />

method. The shown variance corresponds to the contour after 240 iterations (the magenta<br />

contour). Right: Mean value <strong>of</strong> the stochastic image to be segmented and the contours at<br />

time points <strong>of</strong> the level set evolution. The final contour matches the object boundary.<br />

Numerical Results<br />

The application <strong>of</strong> the stochastic geodesic active contour method is demonstrated on the same data<br />

sets as the gradient-based segmentation on stochastic images, namely the stochastic CT image and<br />

the liver mask with two random variables and a polynomial degree <strong>of</strong> two.<br />

A comparison <strong>of</strong> the intrusive implementation <strong>using</strong> the polynomial chaos expansion and the sampling<br />

based implementations <strong>using</strong> Monte Carlo simulation and stochastic collocation is depicted in<br />

Fig. 7.8 for the CT data and shows a good consistency <strong>of</strong> the implementations.<br />

Fig. 7.9 shows the expected value contour for time points during the level set evolution together<br />

with the variance <strong>of</strong> the final level set <strong>of</strong> the segmentation <strong>of</strong> the second data set. Again, the regions<br />

with a high variance are the bottom and the upper-right side <strong>of</strong> the object. This is consistent with the<br />

results from the gradient-based segmentation (cf. Fig. 7.6). Furthermore, the right picture <strong>of</strong> Fig. 7.9<br />

shows the evolution <strong>of</strong> the expected value contour on the expected value image. The expected value<br />

contour after 240 iterations with a time step <strong>of</strong> 0.2h is aligned to the object boundary. The variance<br />

corresponding to this contour is the same as the variance in the left picture <strong>of</strong> the same figure.<br />

The advantage <strong>of</strong> the stochastic geodesic active contour approach over the stochastic gradientbased<br />

segmentation is that a running over the edges is mostly avoided (cf. Fig. 7.11).<br />

7.5.3 <strong>Stochastic</strong> Chan-Vese segmentation<br />

We derive the stochastic Chan-Vese model from classical Chan-Vese model by replacing all quantities<br />

by their stochastic counterparts:<br />

( ( ) )<br />

∇φ<br />

φ t = δ ε (φ) µ∇ · − ν − λ 1 (u 0 − c 1 ) 2 + λ 2 (u 0 − c 2 ) 2<br />

|∇φ|<br />

, (7.36)<br />

where the phase field φ, the initial image u 0 , the mean values c 1 and c 2 , and the smooth delta<br />

approximation δ ε are stochastic quantities, i.e. they are dependent on the random event ω ∈ Ω. The<br />

function δ ε is the derivative <strong>of</strong> the stochastic smooth Heaviside approximation<br />

H ε (z(ω)) = 1 2<br />

(<br />

1 + 2 ( )) z(ω)<br />

π arctan ε<br />

. (7.37)<br />

92


7.5 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using <strong>Stochastic</strong> Level Sets<br />

Figure 7.10: Mean (left) <strong>of</strong> the stochastic CT and the variance (right) <strong>of</strong> the stochastic Chan-Vese<br />

solution. Additionally, we show the expected value contour at different time steps.<br />

The regularized stochastic δ-function δ ε is<br />

1<br />

δ ε (z(ω)) =<br />

πε + π . (7.38)<br />

ε<br />

z(ω)2<br />

The mean value <strong>of</strong> the object and the background is a stochastic quantity because we have to average<br />

over a collection <strong>of</strong> random variables. The mean values are<br />

∫<br />

D<br />

c 1 (φ) =<br />

u ∫<br />

0(x)H ε (φ(x))dx<br />

∫<br />

D<br />

D H and c 2 (φ) =<br />

u 0(x)(1 − H ε (φ(x)))dx<br />

∫<br />

ε(φ(x))dx<br />

D (1 − H . (7.39)<br />

ε(φ(x)))dx<br />

Note that we average over the spatial dimensions, i.e. over the deterministic image domain only.<br />

Thus, c 1 and c 2 are random variables. In (7.39) we have to evaluate the Heaviside approximation,<br />

which involves the computation <strong>of</strong> the inverse tangent <strong>of</strong> a stochastic quantity. To avoid the necessity<br />

to develop a numerical scheme for the stochastic inverse tangent, we use a well-known approximation,<br />

see e.g. [131]:<br />

{<br />

x<br />

arctan(x) ≈<br />

1+0.28x 2 if |x| ≤ 1<br />

else<br />

π<br />

2 − x<br />

x 2 +0.28<br />

. (7.40)<br />

This is not a real drawback <strong>of</strong> the stochastic discretization, because it can be interpreted as an alternative<br />

approximation <strong>of</strong> the Heaviside function and is not as bad as an approximation <strong>of</strong> an approximation<br />

as it might look. The remaining part <strong>of</strong> the Chan-Vese model is generalized to stochastic<br />

quantities by <strong>using</strong> Debusschere’s methods for the computation with polynomial chaos quantities<br />

(see Section 3.3 and [38]). The main driving force <strong>of</strong> the stochastic Chan-Vese model is the difference<br />

between the mean value <strong>of</strong> the separated region and the actual gray value. The mean value<br />

<strong>of</strong> the image regions is computed via an averaging <strong>of</strong> a collection <strong>of</strong> random variables. Thus, the<br />

stochastic information cancels out <strong>of</strong> the stochastic Chan-Vese model, because we are approximating<br />

the “real”, noise-free, mean value when we average over a huge number <strong>of</strong> random variables.<br />

The Variance as Homogenization Criterion for <strong>Stochastic</strong> Chan-Vese <strong>Segmentation</strong><br />

Up to now, we have used the (spatial) mean value <strong>of</strong> the stochastic image as homogenization criterion<br />

only. Thus, we ignore stochastic information, e.g. the variance, <strong>of</strong> the stochastic image. Homogenizing<br />

the variance <strong>of</strong> the segmented object and background can improve the segmentation result<br />

further. For example, in medical images different organs or tissue components can have different<br />

93


Chapter 7 <strong>Stochastic</strong> Level Sets<br />

Figure 7.11: Left: MC-realizations <strong>of</strong> the stochastic contour from the stochastic Chan-Vese segmentation<br />

applied on the CT data set. Right: Realizations <strong>of</strong> the stochastic contour from the<br />

stochastic geodesic active contour approach applied on the CT data set.<br />

noise levels. Thus, they can be separated by homogenizing the variance. To include the homogenization<br />

<strong>of</strong> the variance we add additional terms to the stochastic Chan-Vese model that are inspired by<br />

the terms for the homogenization <strong>of</strong> the mean value. The inclusion <strong>of</strong> stochastic moments in functionals<br />

has been investigated e.g. by Tiesler et al. [146]. To be more precise, we add two additional<br />

components to the Chan-Vese energy leading to<br />

( ( )<br />

∇φ<br />

φ t = δ ε (φ) µ∇·<br />

)−ν −λ 1 (u 0 − c 1 ) 2 +λ 2 (u 0 − c 2 ) 2 −ρ 1 (Var(u 0 ) − v 1 ) 2 +ρ 2 (Var(u 0 )−v 2 ) 2 .<br />

|∇φ|<br />

(7.41)<br />

In (7.41) we added two parameters ρ 1 and ρ 2 to weight the additional components. Furthermore, the<br />

new components v 1 and v 2 are defined as<br />

∫<br />

D<br />

v 1 (φ) =<br />

Var(u 0(x))H ε (φ(x))dx<br />

∫<br />

D H ε(φ(x))dx<br />

∫<br />

and v 2 (φ) =<br />

D Var(u 0(x))(1 − H ε (φ(x)))dx<br />

∫<br />

D (1 − H ε(φ(x)))dx<br />

. (7.42)<br />

Remember, the variance can be computed from the polynomial chaos expansion <strong>of</strong> the stochastic image<br />

easily. Moreover, it is possible to homogenize every polynomial chaos coefficient independently,<br />

leading to various additional constraints.<br />

Numerical Results<br />

We apply the stochastic Chan-Vese model on the same data sets as the other methods: the stochastic<br />

CT and the stochastic liver mask. Fig. 7.10 shows the expected value <strong>of</strong> the liver data set along<br />

with the expected value contour at stages <strong>of</strong> the evolution. The stochastic Chan-Vese model slightly<br />

overestimates the object, because the final (green) contour is not perfectly aligned with the boundary.<br />

This is due to the homogenization criterion the stochastic Chan-Vese model tries to fulfill. Fig. 7.10<br />

shows the variance <strong>of</strong> the final level set along with the contours at stages <strong>of</strong> the evolution. The<br />

variance indicated that the segmentation is uncertain in the critical areas at the object’s bottom and<br />

top. Furthermore, the variance identifies two critical regions on the right and the left <strong>of</strong> the object.<br />

Fig. 7.11 shows realizations <strong>of</strong> the final contour via a MC-sampling from the stochastic result.<br />

For the liver data set, we show the level set evolution on the expected value <strong>of</strong> the initial image and<br />

on the variance <strong>of</strong> the final level set in Fig. 7.12. The data set is constructed by adding artificial noise<br />

to a noise-free image. This noise nearly cancels out due to the averaging process for the computa-<br />

94


7.5 <strong>Segmentation</strong> <strong>of</strong> <strong>Stochastic</strong> <strong>Images</strong> Using <strong>Stochastic</strong> Level Sets<br />

Figure 7.12: Mean (left) <strong>of</strong> the stochastic liver image and the variance <strong>of</strong> the stochastic Chan-Vese<br />

solution. In addition, we show the expected value contour at different time steps.<br />

Figure 7.13: Variance <strong>of</strong> the stochastic image to segment (left), the expected value is not depicted,<br />

because the expected value is an image with the same gray value at every pixel. On<br />

the right, the segmentation result is depicted on one realization (one sample) <strong>of</strong> the<br />

stochastic image to segment.<br />

tion <strong>of</strong> the random variable for the mean inside the regions. The realizations drawn from the final<br />

stochastic level set fit to each other and to the final contour <strong>of</strong> the level set evolution from Fig. 7.12.<br />

The extension <strong>of</strong> the stochastic Chan-Vese approach that tries to homogenize the variance <strong>of</strong> the<br />

object and the background allows to segment objects in images with constant mean, i.e. it allows to<br />

segment objects from constant images where the classical method fails. Fig. 7.13 shows the result<br />

<strong>of</strong> the segmentation <strong>of</strong> an image with constant mean, but non-constant variance. Drawing samples<br />

<strong>of</strong> this image (cf. Fig. 7.13) the object is visible on samples through the different variance levels, but<br />

again, the classical Chan-Vese approach cannot segment the object in the image due to the constant<br />

mean value, whereas the variance extension <strong>of</strong> the stochastic Chan-Vese approach yields the correct<br />

result. In fact, the Chan-Vese approach without variance homogenization would not move the initial<br />

contour because the driving force is zero due to the constant mean value.<br />

Conclusion<br />

We presented an extension <strong>of</strong> the level set approach to use random variables or random fields as<br />

propagation speed. The use <strong>of</strong> this uncertain speed leadss to an uncertain interface position char-<br />

95


Chapter 7 <strong>Stochastic</strong> Level Sets<br />

acterizing the influence <strong>of</strong> the uncertain propagation speed. The resulting stochastic level set equation,<br />

a hyperbolic SPDE, is transformed into a parabolic SPDE to end up with an equation that can<br />

be discretized with intrusive numerical methods for SPDEs. The extension <strong>of</strong> the classical level<br />

set equation is important in many applications, because the modeling <strong>of</strong> imprecise known material<br />

parameters, boundaries, or source terms through random fields [53, 84, 110, 128, 161] is a rapidly<br />

growing field <strong>of</strong> research. Furthermore, we presented a method for the reinitialization <strong>of</strong> stochastic<br />

level sets and showed that the commonly used classical reinitialization methods like fast marching<br />

cannot be applied in the stochastic context.<br />

Based on the stochastic level set equation, we extended three segmentation methods. Using a<br />

stochastic image as input for these methods, we end up with an uncertain speed as driving term for<br />

the segmentation that depends on information extracted from the stochastic image. Using gradient information<br />

only, as in the first presented method, we end up with a method that uses local information<br />

only. Thus, this method is highly sensitive to the noise characterized by the stochastic components <strong>of</strong><br />

the stochastic image. Using additional global stochastic information, as in the stochastic Chan-Vese<br />

approach, weakens the influence <strong>of</strong> the input uncertainty on the segmentation result.<br />

Additional to the stochastic extensions <strong>of</strong> the classical segmentation methods, we presented an<br />

extension <strong>of</strong> the Chan-Vese approach that tries to homogenize the variance <strong>of</strong> the segmented object.<br />

Thus, this method is not an extension in the spirit <strong>of</strong> the other method extensions, where we replaced<br />

classical images by their stochastic counterparts. Instead, this extension allows to use stochastic<br />

information as driving force <strong>of</strong> the segmentation. This enables us to segment images that cannot be<br />

segmented with the classical methods. For example, we are able to segment objects in an image with<br />

constant mean, when they have different noise properties, i.e. a different variance.<br />

96


Chapter 8<br />

<strong>Segmentation</strong> <strong>of</strong> Classical <strong>Images</strong> Using<br />

<strong>Stochastic</strong> Parameters<br />

In the previous chapters, we presented methods for the segmentation <strong>of</strong> stochastic images. All these<br />

methods are based on the solution <strong>of</strong> SPDEs and we get a stochastic segmentation result characterizing<br />

the influence <strong>of</strong> the gray value uncertainty on the segmentation result. This chapter uses a<br />

different approach that leads also to SPDEs for the segmentation <strong>of</strong> images.<br />

In applications, the user tweaks the parameters <strong>of</strong> the segmentation methods to get satisfying<br />

results. Often, the user performs this tweaking for every single data set. Thus, the segmentation<br />

result is not dependent on the input image and the selected segmentation methods only, but also on<br />

the particular choice <strong>of</strong> the parameters by the user. This yields the problem that the segmentation<br />

result is not reproducible among users. The influence <strong>of</strong> the user parameters on the segmentation<br />

results is important e.g. in medical applications, when different users segment base and follow-up<br />

scans <strong>using</strong> different segmentation parameters. It is difficult to decide whether the segmentation<br />

result is different due to a growth <strong>of</strong> the tumor or due to the different segmentation parameters.<br />

In cancer therapy the further treatment <strong>of</strong> the patient is based on the segmentation results <strong>using</strong><br />

RECIST [48, 145]. Thus, information about the stability <strong>of</strong> the segmentation with respect to the<br />

parameters might be useful to come to an informed decision.<br />

In this chapter, we try to investigate the influence <strong>of</strong> the segmentation parameters on the segmentation<br />

result. This task is known as sensitivity analysis [26, 135]. The main idea is to replace the<br />

deterministic segmentation parameters by random variables and apply the segmentation methods on<br />

deterministic images. The stochastic segmentation result is comparable to the result <strong>of</strong> the segmentation<br />

<strong>of</strong> stochastic images. The difference is that the components are stochastic due to the stochastic<br />

parameters instead <strong>of</strong> the stochastic image. We visualize the results <strong>using</strong> the same techniques as for<br />

stochastic images, showing the influence <strong>of</strong> the segmentation parameters on the segmentation result.<br />

With this approach, we detect regions in the image highly influenced by the choice <strong>of</strong> the segmentation<br />

parameters and regions, where the segmentation is robust with respect to parameter changes.<br />

In addition, we investigate which segmentation parameters have a strong influence on the segmentation<br />

result. For geodesic active contours, the influence <strong>of</strong> the smoothing term should be nearly the<br />

same on the whole image, whereas the weight related to the edge detector is important on the edges<br />

in the image. This approach needs few random variables only, typically one for every segmentation<br />

parameter. Hence, this approach is suitable for a discretization via the methods presented in this<br />

thesis without the need to reduce the number <strong>of</strong> random variables necessary for stochastic images<br />

via the Karhunen-Loève decomposition.<br />

In the following, we investigate the use <strong>of</strong> stochastic segmentation parameters for random walker<br />

segmentation, Ambrosio-Tortorelli segmentation, gradient-based segmentation, and geodesic active<br />

contours. We discretize the methods <strong>using</strong> slightly adopted versions <strong>of</strong> the stochastic segmentation<br />

methods for the segmentation <strong>of</strong> stochastic images presented in Chapter 6 and Chapter 7.<br />

97


Chapter 8 <strong>Segmentation</strong> <strong>of</strong> Classical <strong>Images</strong> Using <strong>Stochastic</strong> Parameters<br />

8.1 Random Walker <strong>Segmentation</strong> with <strong>Stochastic</strong> Parameter<br />

The random walker segmentation has one free parameter that the user has to choose during the<br />

segmentation process. This parameter, denoted by β, controls the influence <strong>of</strong> the image gradient on<br />

the matrix entries because the edge weights for random walker segmentation (cf. Section 2.2) are<br />

(<br />

w i j = exp −β (g i − g j ) 2) . (8.1)<br />

Making the parameter β a random variable and approximating this random variable in the polynomial<br />

chaos (cf. Section 3.3), the stochastic edge weights for the sensitivity analysis are<br />

w i j (ξ ) = exp<br />

(<br />

−<br />

( N∑<br />

β α Ψ α (ξ )<br />

α=1<br />

)(g i − g j ) 2 )<br />

. (8.2)<br />

Note that the parameter β is not restricted, but can be, the standard random variable for the construction<br />

<strong>of</strong> the polynomial chaos expansion, e.g. a uniform random variable. In fact, we use the power <strong>of</strong><br />

the polynomial chaos approximation by making the parameter β dependent on a couple <strong>of</strong> standard<br />

random variables with an adequate polynomial degree in the polynomial chaos expansion.<br />

Using the stochastic edge weights from (8.2), we define the node degree analog to Section 6.1:<br />

d i (ξ ) =<br />

∑<br />

{ j∈V :e i j ∈E}<br />

w i j (ξ ) =<br />

∑<br />

N<br />

∑<br />

{ j∈V :e i j ∈E} α=1<br />

(w i j ) α Ψ α (ξ ) . (8.3)<br />

Note that for the sensitivity analysis we use the exact normalization <strong>of</strong> the image gradient given<br />

by (2.6), because the pixel values are deterministic values in this setting. From the stochastic edge<br />

weights (8.2) and the stochastic node degrees (8.3), we construct the stochastic Laplacian matrix via<br />

⎧<br />

⎨<br />

L i j (ξ ) =<br />

⎩<br />

=<br />

N<br />

∑<br />

α=1<br />

d i (ξ ) if i = j<br />

−w i j (ξ ) if v i and v j are adjacent nodes<br />

0 otherwise<br />

L α Ψ α (ξ ) .<br />

Finally, we end up with the same stochastic equation system as in Section 6.1, but the stochastic<br />

components are due to the stochastic parameter instead <strong>of</strong> stochastic pixels inside the image:<br />

(8.4)<br />

L U (ξ )x U (ξ ) = −B(ξ ) T x M (ξ ) . (8.5)<br />

We have to use stochastic images to store the stochastic solution. The stochastic images have to<br />

contain the same random variables the parameter depends on.<br />

Remark 18. The discretization <strong>of</strong> the random walker segmentation with a stochastic parameter uses<br />

the generalized spectral decomposition. The only small variation in the implementation is that we<br />

have to use a polynomial chaos approximation <strong>of</strong> the parameter β for the calculation <strong>of</strong> the edge<br />

weights. The edge weights themselves are already random quantities in the stochastic random walker<br />

implementation <strong>of</strong> Section 6.1.<br />

98


8.1 Random Walker <strong>Segmentation</strong> with <strong>Stochastic</strong> Parameter<br />

Figure 8.1: Left: Realizations <strong>of</strong> the stochastic contour obtained from the random walker segmentation<br />

with stochastic parameter. Right: Mean and variance <strong>of</strong> the stochastic contour<br />

obtained from the random walker segmentation with stochastic parameter.<br />

Results<br />

We perform random walker segmentation on the well-known data sets from the last chapters. Because<br />

all stochasticity is due to the parameters, we use the expected value image <strong>of</strong> the stochastic<br />

input data set only, i.e. we use a deterministic input image. Thus, this method is, in contrast to all<br />

other methods presented so far, usable for classical images. To be able to capture the stochasticity<br />

introduced by the stochastic parameters, we identify the deterministic input image with a stochastic<br />

image containing one random variable (the random variable the stochastic parameter depends on)<br />

and use a maximal polynomial degree <strong>of</strong> four. The stochastic components <strong>of</strong> the input are set to zero.<br />

The only parameter in the random walker method is the parameter β for the estimation <strong>of</strong> the<br />

graph weights. In the following, we model this parameter as a uniformly distributed random variable<br />

with expected value ten, i.e. the first coefficient <strong>of</strong> the polynomial chaos expansion is β 1 = 10. For<br />

the experiments we used a variance <strong>of</strong> the uniform random variable <strong>of</strong> three, resulting in a random<br />

variable that is uniformly distributed between 7 and 13, i.e. β ∼ U [7,13]. For the polynomial chaos<br />

expansion <strong>of</strong> the parameter β, setting β 2 = √ 3 and β i = 0,i > 2 models this behavior.<br />

Fig. 8.1 shows the image to segment, realizations <strong>of</strong> the stochastic contour, the expected value<br />

contour and the variance for the US data set. Note that there is no direct relation between regions<br />

with a high variance and regions where the contour realizations are far from each other as one might<br />

suggest. In fact, the distance between the contour realizations depends on the gradient <strong>of</strong> the underlying<br />

probability map (the expected value <strong>of</strong> the result) and the variance. In regions where the<br />

expected value is around 0.5 and has a low gradient, a small variance can influence the contour position<br />

significantly, whereas in regions with a high gradient, even a high variance cannot influence<br />

the contour position visually. The upper right corner <strong>of</strong> Fig. 8.1, where a low variance corresponds<br />

to varying contour positions, shows this effect. Furthermore, this is visible in Fig. 8.2, where the<br />

random walker segmentation <strong>of</strong> the liver data set is shown. There, the highest uncertainty in the<br />

contour position is in the quadrant with the lowest gradient between object and background and the<br />

highest uncertainty in the expected value <strong>of</strong> the probability map is at the bottom <strong>of</strong> the object.<br />

Furthermore, the result depicted in Fig. 8.2 shows two problems <strong>of</strong> the random walker segmentation<br />

method. First, the method needs a strong gradient between object and background. Otherwise,<br />

99


Chapter 8 <strong>Segmentation</strong> <strong>of</strong> Classical <strong>Images</strong> Using <strong>Stochastic</strong> Parameters<br />

Figure 8.2: Left: Realizations <strong>of</strong> the stochastic contour obtained from the random walker segmentation<br />

with stochastic parameter. Right: Mean and variance <strong>of</strong> the stochastic contour<br />

obtained from the random walker segmentation with stochastic parameter.<br />

the segmentation fails, like in the upper left part <strong>of</strong> the segmentation result in Fig. 8.2. The second<br />

problem is due to sharp corners <strong>of</strong> the object, like the corner in the middle <strong>of</strong> Fig. 8.2. The random<br />

walker method tries to identify smooth objects, because internally it solves a diffusion equation that<br />

prefers smooth solutions. Both problems can be reduced by defining additional seed points for the<br />

segmentation close to the problematic regions.<br />

Another observation from the stochastic segmentation result is that the PDF <strong>of</strong> the segmented volume<br />

(cf. Section 6.1.2) is not uniformly distributed, even though the input (the stochastic parameter)<br />

has a uniform distribution. Fig. 8.3 shows the PDF <strong>of</strong> the segmented areas for both test examples.<br />

For the US data set, the resulting PDF is close to a uniform distribution (the left picture <strong>of</strong> Fig. 8.3),<br />

for the liver data set the area PDF is concentrated around a peak (right picture <strong>of</strong> Fig. 8.3). Both<br />

PDF are computed <strong>using</strong> the method described in Section 6.1.5 given by summing up the random<br />

variables <strong>of</strong> all pixels, cf. (6.12).<br />

Figure 8.3: Volume <strong>of</strong> the stochastic contour obtained from the random walker segmentation with<br />

stochastic parameter. The left curve shows the PDF <strong>of</strong> the object in the US image, the<br />

right curve the PDF <strong>of</strong> the liver in the liver data set. The PDFs are obtained <strong>using</strong> (6.12).<br />

100


8.2 Ambrosio-Tortorelli <strong>Segmentation</strong> with <strong>Stochastic</strong> Parameters<br />

8.2 Ambrosio-Tortorelli <strong>Segmentation</strong> with <strong>Stochastic</strong> Parameters<br />

Ambrosio-Tortorelli segmentation on stochastic images requires the solution <strong>of</strong> a system <strong>of</strong> two<br />

coupled SPDEs and involves four parameters the user has to choose:<br />

−∇ · (µ(φ(x,ξ<br />

) 2 + k ε )∇u(x,ξ ) ) + u(x,ξ ) = u 0 (x,ξ )<br />

( 1<br />

−ε∆φ(x,ξ ) +<br />

4ε + µ )<br />

2ν |∇u(x,ξ )|2 φ(x,ξ ) = 1<br />

4ε . (8.6)<br />

The parameter µ controls the influence <strong>of</strong> the phase field value on the image smoothing process and<br />

ε controls the width <strong>of</strong> the phase field. The influence <strong>of</strong> the image gradient on the phase field is<br />

controled by ν and k ε is an additional regularization parameter that ensures ellipticity <strong>of</strong> the first<br />

equation. By making all four parameters random variables, it is possible to investigate which parameter<br />

has the strongest influence on the segmentation result. The adoption <strong>of</strong> (8.6) is straightforward.<br />

We replace the classical parameters µ,ν,k ε and ε by their stochastic counterparts µ(ξ ),ν(ξ ),k ε (ξ )<br />

and ε(ξ ) and approximate these random variables in the polynomial chaos. We end up with<br />

−∇ · (µ(ξ<br />

)(φ(x,ξ ) 2 + k ε (ξ ))∇u(x,ξ ) ) + u(x,ξ ) = u 0 (x,ξ )<br />

( 1<br />

−∇ · (ε(ξ )∇φ(x,ξ )) +<br />

4ε(ξ ) + µ(ξ ) )<br />

2ν(ξ ) |∇u(x,ξ )|2 φ(x,ξ ) = 1<br />

4ε(ξ ) . (8.7)<br />

The discretization <strong>of</strong> this SPDE system is analog to the discretization <strong>of</strong> the SPDE system for stochastic<br />

images. The differences are that the coefficient <strong>of</strong> the Laplacian in the second equation is a<br />

stochastic quantity and that the right hand side <strong>of</strong> the second equation is a stochastic quantity, too.<br />

∫D 1<br />

4ε(ω)<br />

Thus, we integrate ∫ Ω dxdω to compute the right hand side by an integration rule, and we<br />

have to use an assembling method for the inhomogeneous stiffness matrix to discretize ∇ε(ω)∇φ.<br />

The discretization <strong>of</strong> (8.7) <strong>using</strong> finite elements for the deterministic dimensions and the polynomial<br />

chaos for the stochastic dimensions is<br />

N<br />

∑<br />

α=1<br />

(<br />

) M α,β + L α,β U α =<br />

N<br />

∑<br />

α=1<br />

M α,β (U 0 ) α ,<br />

N (<br />

∑<br />

α=1<br />

S α,β + T α,β ) Φ α =<br />

N<br />

∑<br />

α=1<br />

A α (8.8)<br />

for all β ∈ {1,...,N}, where M α,β ,L α,β ,S α,β and T α,β are blocks <strong>of</strong> the system matrix, defined as<br />

(M ) (Ψ ) ∫<br />

α,β = E α Ψ β P i P j dx<br />

i, j D<br />

( ) (<br />

S α,β = ∑∑E<br />

Ψ α Ψ β Ψ γ) ∫<br />

(˜ε) k (8.9)<br />

γ ∇P i · ∇P j P k dx ,<br />

i, j<br />

k γ<br />

and (<br />

(<br />

L α,β ) i, j = ∑<br />

k<br />

T α,β ) i, j = ∑<br />

k<br />

(<br />

∑E<br />

γ<br />

∑E<br />

γ<br />

Ψ α Ψ β Ψ γ) (˜φ 2 ) k γ<br />

(<br />

Ψ α Ψ β Ψ γ) u k γ<br />

D<br />

∫<br />

D<br />

∫<br />

D<br />

∇P i · ∇P j P k dx,<br />

P i P j P k dx .<br />

(8.10)<br />

Here, (˜φ 2 ) k γ and u k γ denote the coefficients <strong>of</strong> the polynomial chaos expansion <strong>of</strong> the Galerkin projection<br />

<strong>of</strong> µ(ξ )(φ(ξ ) 2 1<br />

+ k ε (ξ )) respectively<br />

4ε(ξ ) + µ(ξ )<br />

onto the image space (cf. [38]).<br />

2ν(ξ )|∇u(ξ )| 2<br />

Finally, the right hand side vector <strong>of</strong> the phase field equation is<br />

∫ ∫<br />

(A α 1<br />

) i =<br />

4ε P i(x)dxΨ α (ξ )dΠ . (8.11)<br />

Γ<br />

D<br />

We solve the SPDE system for the sensitivity analysis with the same methods as the SPDE system<br />

for stochastic images from Section 6.2, i.e. it is possible to use the GSD for the resolution process.<br />

101


Chapter 8 <strong>Segmentation</strong> <strong>of</strong> Classical <strong>Images</strong> Using <strong>Stochastic</strong> Parameters<br />

Figure 8.4: Ambrosio-Tortorelli model applied on the expected value <strong>of</strong> the liver data set <strong>using</strong> a<br />

stochastic parameter µ. The upper row shows the expected value (left) and the variance<br />

(right) <strong>of</strong> the smoothed image, the lower row the expected value (left) and the variance<br />

(right) <strong>of</strong> the phase field.<br />

Results<br />

We applied the Ambrosio-Tortorelli segmentation with stochastic parameters on the liver data set.<br />

Again, we use the expected value <strong>of</strong> the stochastic data set as deterministic input and construct a<br />

stochastic input image that contains one random variable and a maximal polynomial chaos degree<br />

<strong>of</strong> four. As in the random walker tests, the remaining stochastic dimensions are filled up with zeros.<br />

To separate the influence <strong>of</strong> the stochasticity <strong>of</strong> the parameters in the Ambrosio-Tortorelli model, we<br />

use one stochastic parameter for the first tests and keep the other parameters deterministic.<br />

Fig. 8.4 shows the result for a uniformly distributed parameter µ. To be precise, µ is uniformly<br />

distributed between 200 and 600, i.e. µ ∼ U [200,600]. The parameter µ controls the influence <strong>of</strong> the<br />

smoothing term in the image equation. For large µ we get sharper images with sharp edges. Thus,<br />

a stochastic parameter µ influences the smoothing <strong>of</strong> the image. This is visible from the variance <strong>of</strong><br />

102


8.2 Ambrosio-Tortorelli <strong>Segmentation</strong> with <strong>Stochastic</strong> Parameters<br />

Figure 8.5: Ambrosio-Tortorelli model applied on the expected value <strong>of</strong> the liver data set <strong>using</strong> a<br />

stochastic parameter ε. The upper row shows the expected value (left) and the variance<br />

(right) <strong>of</strong> the smoothed image and the lower row the expected value (left) and the variance<br />

(right) <strong>of</strong> the phase field.<br />

the smoothed image in Fig. 8.4, where a smoothing across the object boundaries leads to a variance<br />

that looks similar to the original image. This is due to the cartoon-like initial image. Once energy is<br />

transported across the edge, it is equally distributed in the whole region due to the smoothing term.<br />

The smooth image resulting from the image equation influences the phase field, because it leads to<br />

diffuse boundaries and to a wide phase field that is visible in the phase field variance in Fig. 8.4.<br />

In Fig. 8.5 we used a stochastic parameter ε uniformly distributed between 0.0015 and 0.0035,<br />

i.e. ε ∼ U [0.0015,0.0035]. The parameter ε influences the width <strong>of</strong> the phase field, but has no<br />

influence on the smoothing parts <strong>of</strong> the equations. We observe changes in the variance around the<br />

edges in Fig. 8.5. Directly, the parameter ε influences the width <strong>of</strong> the phase field and due to the<br />

wider phase field, the image is smoothed differently close to edges.<br />

103


Chapter 8 <strong>Segmentation</strong> <strong>of</strong> Classical <strong>Images</strong> Using <strong>Stochastic</strong> Parameters<br />

8.3 Gradient-Based <strong>Segmentation</strong> with <strong>Stochastic</strong> Parameter<br />

Gradient-based segmentation via a level set formulation contains one parameter b that controls the<br />

influence <strong>of</strong> the curvature κ. Making this parameter a random variable, we end up with<br />

φ t (t,x,ω) + v(1 − b(ω)κ(t,x,ω))|∇φ(t,x,ω)| = 0 . (8.12)<br />

The stopping function v is v = 1<br />

1+|∇u|<br />

. Additionally to the necessary Galerkin projection in the<br />

numerical scheme for the solution <strong>of</strong> the gradient-based segmentation, we have to project bκ back to<br />

the polynomial chaos. For this, we use the standard methods presented in Section 3.3. The remaining<br />

part <strong>of</strong> the discretization is analog to the discretization <strong>of</strong> the stochastic gradient-based segmentation<br />

with stochastic image.<br />

Results<br />

We present the gradient-based segmentation with stochastic parameter <strong>using</strong> the CT data set and the<br />

liver data set. As usual, we used the expected value as input and used one random variable and a<br />

polynomial degree <strong>of</strong> four. For the experiment, we used a stochastic parameter b that is uniformly<br />

distributed between 0.75 and 1.25, i.e. b ∼ U [0.75,1.25]. The parameter b controls the influence <strong>of</strong><br />

the curvature smoothing. A higher parameter b leads to smoother contours. This is shown in Fig. 8.6<br />

where the contour realizations vary with respect to the curvature.<br />

Figure 8.6: Result <strong>of</strong> the gradient-based segmentation with stochastic parameter b, i.e. with a stochastic<br />

curvature smoothing. The upper row shows the results for the CT data set, expected<br />

value <strong>of</strong> the image and contour realizations (left) and variance <strong>of</strong> the level set with contour<br />

realizations (right). The lower row shows the same results for the liver data set. In<br />

all figures we add Monte Carlo realizations <strong>of</strong> the stochastic object boundary. The red<br />

contour corresponds to a b = 0.75, yellow to b = 1.0 and blue to b = 1.25.<br />

104


8.4 Geodesic Active Contours with <strong>Stochastic</strong> Parameters<br />

8.4 Geodesic Active Contours with <strong>Stochastic</strong> Parameters<br />

The sensitivity analysis for the geodesic active contour approach follows the procedure for the sensitivity<br />

analysis <strong>of</strong> the other segmentation methods. Geodesic active contours are given by<br />

φ t (t,x) = γg(t,x)κ(t,x)|∇φ(t,x)| + α∇g(t,x)∇φ(t,x) − βg|∇φ| . (8.13)<br />

The parameters α,β, and γ can be chosen to optimize the segmentation result. The parameter α<br />

controls the attraction <strong>of</strong> the minima <strong>of</strong> the speed function v. The parameter β controls the shrinkage<br />

(negative β) or expansion (positive β) <strong>of</strong> the level set and the parameter γ acts as a weighting term<br />

for the curvature smoothing.<br />

Making the segmentation parameters random variables, we end up with<br />

φ t = γ(ω)g(t,x,ω)κ(t,x,ω)|∇φ(t,x,ω)| + α(ω)∇g(t,x,ω)∇φ(t,x,ω) − β(ω)g|∇φ| . (8.14)<br />

This equation is nearly identical to the stochastic geodesic active contour equation, but requires an<br />

additional projection step during the discretization to projects the products γg,α∇g, and βg back to<br />

the polynomial chaos. Besides this additional projection step, we use the same numerical methods<br />

as for the discretization <strong>of</strong> the stochastic geodesic active contour equation in Section 7.5.2, i.e. we<br />

use an explicit time step discretization via the Euler method and a uniform spatial grid.<br />

Results<br />

The geodesic active contour method with stochastic parameters is performed on the same data sets as<br />

in the previous sections. Due to the smooth objects that we try to segment in the images, we ignore<br />

the smoothing term by setting γ = 0. The parabolic approximation and the attraction term ∇g∇φ<br />

ensure that we get smooth results in this setting, too. The parameters α and β are chosen by setting<br />

α 1 = 0.08, α 2 = 0.002, β 1 = 1.0, and β 2 = 0.02. Thus, we use two stochastic parameters at the same<br />

time and make them both dependent on the same random variable. Since we set the expected value<br />

and the first coefficient to a nonzero value, we end up with uniformly distributed parameters.<br />

Fig. 8.7 shows the result for the CT data set. The image is easy to segment due to the homogeneous<br />

gradient between the inner parts <strong>of</strong> the head phantom and the bone. The problematic parts are the<br />

Figure 8.7: Result <strong>of</strong> the geodesic active contour segmentation with stochastic parameters for the CT<br />

data set. On the left the expected value <strong>of</strong> the image and contour realizations and on the<br />

right the variance <strong>of</strong> the level set with contour realizations are shown.<br />

105


Chapter 8 <strong>Segmentation</strong> <strong>of</strong> Classical <strong>Images</strong> Using <strong>Stochastic</strong> Parameters<br />

Figure 8.8: Result <strong>of</strong> the geodesic active contour segmentation with stochastic parameters for the<br />

liver data set. On the left the expected value <strong>of</strong> a detail <strong>of</strong> the image and contour realizations<br />

and on the right the variance <strong>of</strong> the level set with contour realizations are shown.<br />

regions, where the object to segment does not have the “elliptic” contour behavior. In these regions,<br />

the gradient differs from the remaining parts <strong>of</strong> the image. The geodesic active contour method has<br />

different attractors depending on the particular value <strong>of</strong> α and β in these regions. This is visible<br />

from Fig. 8.7, because the contour realizations are far from each other, and the variance is high in a<br />

region in the upper part <strong>of</strong> the object boundary.<br />

Remark 19. Note that for level sets there is, in contrast to the random walker method, a one-to-one<br />

correspondence between the distance between the contour realizations and the variance, because we<br />

use a stochastic equivalent <strong>of</strong> the signed distance function. Thus, deviations in the level set position<br />

are related to the variance directly.<br />

For the liver data set, Fig. 8.8 shows the results with the same stochastic parameters. Again, the<br />

results are close together for parameter realizations, because we have one attractor for the level set<br />

only (the object boundary). The differences in the lower part <strong>of</strong> the object are due to the weak<br />

attraction <strong>of</strong> the liver boundary for some realizations <strong>of</strong> the parameter α.<br />

Conclusion<br />

The presented sensitivity analysis is a natural extension <strong>of</strong> the stochastic image processing framework<br />

presented in this thesis. With the sensitivity analysis, we investigate the robustness <strong>of</strong> the<br />

classical image segmentation methods with respect to parameter changes. This additional stochastic<br />

information is available for the costs <strong>of</strong> a few Monte Carlo runs. However, we do not use the Monte<br />

Carlo method, but have to solve a couple <strong>of</strong> deterministic problems when <strong>using</strong> the GSD method.<br />

A possible application <strong>of</strong> this kind <strong>of</strong> sensitivity analysis is to warn the user, when the segmentation<br />

result is sensitive to parameter changes. This can be done via background calculations, i.e. the system<br />

computes the stochastic solution while the user examines the deterministic result. When necessary,<br />

the system informs the user about the stochastic result and makes additional information like the<br />

variance or contour realizations available.<br />

106


Chapter 9<br />

Summary, Discussion, and Conclusion<br />

In this thesis, we presented extensions <strong>of</strong> PDE-based segmentation methods to stochastic images,<br />

i.e. images whose pixels are random variables. The characterization <strong>of</strong> such stochastic images is<br />

based on the recently developed generalized polynomial chaos expansion. With this expansion, we<br />

developed extensions <strong>of</strong> the well-known finite element and finite difference schemes for the discretization<br />

<strong>of</strong> the PDE to the stochastic dimensions, leading to stochastic PDEs. To demonstrate the<br />

power <strong>of</strong> <strong>using</strong> stochastic images, we extended the well-known segmentation methods proposed by<br />

Mumford-Shah and the related approximation by Ambrosio-Tortorelli as well as the random walker<br />

method and three methods based on a level set formulation. The input for the stochastic segmentation<br />

is constructed via computing the leading random variables via a principal component analysis<br />

<strong>of</strong> samples <strong>of</strong> the input scene and a projection on the polynomial chaos basis. Furthermore, we<br />

used the stochastic images and model extensions to perform a sensitivity analysis <strong>of</strong> the methods by<br />

identifying the parameters with random variables.<br />

9.1 Discussion<br />

The work presented in this thesis is a complete framework for the important task <strong>of</strong> error propagation<br />

in mathematical image processing [36, 106]. For every step <strong>of</strong> the mathematical image processing<br />

pipeline (data acquisition, data representation, operator modeling, discretization, solution strategies<br />

and visualization) methods for the solution <strong>of</strong> the particular problems are presented. Besides the<br />

development <strong>of</strong> the framework, theoretical justifications <strong>of</strong> the methods are presented as well. In<br />

particular, these are the extensions <strong>of</strong> the Γ-convergence pro<strong>of</strong> for the stochastic Ambrosio-Tortorelli<br />

model and the pro<strong>of</strong> <strong>of</strong> the existence <strong>of</strong> SPDE solutions used in this thesis.<br />

This thesis applies the error propagation framework to mathematical operators for image segmentation,<br />

but the thesis can also be seen as a case study to demonstrate the applicability <strong>of</strong> the methods<br />

in image processing. Other image processing operators based on a PDE formulation can be extended<br />

by the presented methods easily, because the framework and the implementation <strong>of</strong> all steps around<br />

the operator extension are available. The only step that remains is the stochastic operator extension.<br />

Furthermore, the stochastic parameter study presented in this thesis sensitizes users to be skeptic<br />

about the segmentation results if these are not robust with respect to parameter changes.<br />

9.2 Conclusion<br />

We presented methods for all tasks along the stochastic image processing pipeline, but some <strong>of</strong> the<br />

methods presented in this thesis can be improved to get more stable and more accurate results. For<br />

example, the projection step for the estimation <strong>of</strong> the input distribution (cf. Section 5.2) is based<br />

on a Monte Carlo sampling (it is based on the uncorrelated image samples) and the method has<br />

the poor convergence speed O(1/ √ N) <strong>of</strong> the Monte Carlo method. Stefanou et al. [141] presented<br />

two methods based on an optimization problem. These methods are computationally much more<br />

expensive, but lead to a better convergence speed. Furthermore, the complete stochastic pipeline is<br />

restricted to a few basic random variables, which might be problematic, because image noise has<br />

107


Chapter 9 Summary, Discussion, and Conclusion<br />

Figure 9.1: Visualization <strong>of</strong> the PDF <strong>of</strong> the object boundary in the case <strong>of</strong> the segmentation <strong>of</strong> the<br />

stochastic US image.<br />

a short correlation length in many applications. Thus, multiple random variables are required to<br />

characterize the input adequately. A possibility to deal with huge numbers <strong>of</strong> random variables,<br />

and therefore a high dimensional polynomial chaos, is to use adaptive methods for the stochastic<br />

dimensions. A starting point are the methods presented in [22, 80, 152]. In addition, the use <strong>of</strong> the<br />

parabolic approximation <strong>of</strong> the level set equation might be problematic in applications where sharp<br />

corners and shocks are important. A direct implementation <strong>of</strong> the stochastic level set equation could<br />

be based on the work for hyperbolic SPDEs in [92, 147], but these methods are computationally too<br />

expensive and not accurate enough for this task, but the discretization <strong>of</strong> hyperbolic SPDEs is still<br />

an active field <strong>of</strong> research.<br />

Another important task is the development <strong>of</strong> intuitive visualization techniques for the high dimensional<br />

stochastic output. This thesis presented ideas for the visualization <strong>of</strong> the stochastic results, but<br />

in cooperation with visualization experts, these techniques can be improved. The availability <strong>of</strong> such<br />

visualization techniques is helpful to convince the image processing community <strong>of</strong> this kind <strong>of</strong> error<br />

propagation and to use the error propagation in applications. The user, e.g. a physician, needs<br />

an intuitive access to the stochastic data. A starting point might be the visualization <strong>of</strong> stochastic<br />

boundaries depicted in Fig. 9.1.<br />

9.3 Outlook and Future Work<br />

Besides the improvement <strong>of</strong> the methods presented in this thesis, there are possibilities for future<br />

research in the field <strong>of</strong> image processing with SPDEs. For example, advanced segmentation methods<br />

based on level set formulations can be investigated for stochastic extensions. Furthermore, it is<br />

planned to investigate registration methods [103] for stochastic extensions. Moreover, the stochastic<br />

extensions presented in [130] have to be adapted to the new ansatz space.<br />

Further work directions are the development <strong>of</strong> efficient methods for stochastic finite difference<br />

schemes, especially for nonlinear operations, because the calculation <strong>of</strong> the square root <strong>of</strong> a polynomial<br />

chaos expansion is a bottleneck for most algorithms in this thesis. In addition, the investigation<br />

108


9.3 Outlook and Future Work<br />

<strong>of</strong> level set schemes, which do not need a reinitialization step, is important for efficient stochastic<br />

methods, because at the moment 80% <strong>of</strong> the computation time is spent for the reinitialization.<br />

In addition, the emerging field <strong>of</strong> tensor-structured methods [19, 60, 81] is important for the efficient<br />

solution <strong>of</strong> the presented SPDEs. Tensor-structured methods represent the data and the operators<br />

in a compressed form with a storage requirement linear in the number <strong>of</strong> dimensions, instead <strong>of</strong><br />

the exponential dependence when storing the uncompressed data. Up to now, there are first numerical<br />

examples available in the literature [19, 81] and the methods are not applied on problems arising<br />

in applications like image processing.<br />

A big challenge for the future is to bring this error-aware image processing pipeline into applications.<br />

To be able to achieve this, it is necessary to use problem-dependent basic random variables for<br />

the polynomial chaos. For example, for the modeling <strong>of</strong> magnetic resonance images it is advantageous<br />

to use Rice distributed basic random variables, because the noise <strong>of</strong> gradient magnitude images<br />

is Rice distributed. To use a compatible basis leads to more accurate results with fewer basic random<br />

variables. Other input data require different basic random variables. Therefore, it might be a good<br />

idea to construct the basis on the fly if the input data is available based on the method from [157].<br />

109


List <strong>of</strong> Figures<br />

1.1 Left: CT image <strong>of</strong> a lung lesion (the small roundish structure in the middle <strong>of</strong> the<br />

image). Right: The segmentation mask computed via region growing [127]. . . . . . 1<br />

1.2 Noisy images from an ultrasound device (left) showing a structure in the forearm and<br />

a computed tomography (right) <strong>of</strong> a vertebra in a human spine. . . . . . . . . . . . . 2<br />

1.3 This thesis combines findings from image processing with findings about SPDEs to<br />

yield segmentation algorithms acting on stochastic images. . . . . . . . . . . . . . . 3<br />

2.1 Sketch <strong>of</strong> the ingredients <strong>of</strong> a digital image. At every intersection <strong>of</strong> the regular grid<br />

lines a pixel is located and for every pixel the corresponding FE basis function has<br />

its support in the elements around this pixel. . . . . . . . . . . . . . . . . . . . . . . 8<br />

2.2 The graph generated from a 3 × 3 image contains 9 nodes and 12 edges. The edges<br />

e mn connect the nodes (the black dots) v l . Every edge e mn has a weight w mn describing<br />

the costs for traveling along this edge. . . . . . . . . . . . . . . . . . . . . . . . 10<br />

2.3 Left: Definition <strong>of</strong> the seed regions for the object (yellow) and the background (red).<br />

Middle: The probability that a random walker reaches an object seed. Black denotes<br />

probability zero, white probability one. Right: Random walker segmentation result<br />

<strong>of</strong> the ultrasound image. As input we used the seed regions from the left image and<br />

β = 200. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

2.4 From left to right: Three steps <strong>of</strong> the interactive random walker segmentation. We<br />

show the seeds and the image to segment in the upper row and the segmentation<br />

corresponding to this particular choice <strong>of</strong> the seeds in the lower row. The addition<br />

<strong>of</strong> seed regions for the object and the background yield an iterative refinement <strong>of</strong> the<br />

segmentation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12<br />

2.5 Left: The initial (noisy) US image treated as input for the Ambrosio-Tortorelli approach.<br />

Middle: The smooth Ambrosio-Tortorelli approximation <strong>of</strong> the initial image.<br />

Right: The corresponding phase field, i.e. the approximation <strong>of</strong> the edge set <strong>of</strong> the<br />

smoothed image. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

2.6 Comparison <strong>of</strong> the Ambrosio-Tortorelli model (left) and the extended model <strong>using</strong><br />

the edge linking procedure (right). Data set provided by PD Dr. Christoph S. Garbe. . 17<br />

2.7 <strong>Segmentation</strong> <strong>of</strong> a medical image based on a level set propagation with gradientbased<br />

speed function. The time increases from left to right and the zero level set (red<br />

line) approximates the boundary <strong>of</strong> the object (a liver mask) at the end. . . . . . . . 19<br />

2.8 <strong>Segmentation</strong> <strong>using</strong> geodesic active contours. Left: The initial image. Right: Solution<br />

<strong>of</strong> the geodesic active contour method initialized with small circles inside the<br />

object. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20<br />

2.9 <strong>Segmentation</strong> <strong>of</strong> an object without sharp edges <strong>using</strong> the Chan-Vese approach. In red,<br />

we show the steady-state solution <strong>of</strong> the Chan-Vese segmentation method initialized<br />

with a small circle inside the object. . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

2.10 A test pattern corrupted by uniform (left), Gaussian (middle), and speckle noise (right). 22<br />

3.1 Relation between the stochastic spaces. We avoid the integration over Ω with respect<br />

to the measure Π. Instead, we transform the integral into integration over a subset <strong>of</strong><br />

IR (the space Γ i ) with respect to the known PDF ρ <strong>of</strong> the basic random variables ξ i . . 29<br />

111


List <strong>of</strong> Figures<br />

3.2 Sparsity structure <strong>of</strong> the stochastic lookup table for n = 5 random variables and a<br />

polynomial degree p = 3. The gray dots indicate positions in the three-dimensional<br />

lookup table C αβγ that contain nonzero entries. . . . . . . . . . . . . . . . . . . . . 35<br />

3.3 PDFs <strong>of</strong> initial uniformly distributed input intervals (gray) and the PDFs <strong>of</strong> the results<br />

<strong>of</strong> the polynomial chaos computation (black) for squaring an interval (left) and<br />

dividing an interval by itself (right). . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

4.1 Comparison between a sparse grid (left) constructed via Smolyak’s algorithm and a<br />

full tensor grid (right). The sparse grid contains significantly less nodes than the full<br />

tensor grid whose number <strong>of</strong> nodes growth exponentially with the dimension, but<br />

has nearly the same approximation order. . . . . . . . . . . . . . . . . . . . . . . . 38<br />

4.2 Comparison <strong>of</strong> discretization methods with respect to implementational effort and<br />

speed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

4.3 Refinement <strong>of</strong> a rectangular element <strong>of</strong> a finite element mesh. A single element on a<br />

coarser level splits up into four elements on the next finer level. . . . . . . . . . . . . 45<br />

4.4 Refinement <strong>of</strong> elements leads to hanging nodes (circles) which are no degrees <strong>of</strong><br />

freedom, instead the values <strong>of</strong> the constraining nodes (squares) restrict them. . . . . 45<br />

4.5 For an unsaturated error indicator, the appearance <strong>of</strong> hanging nodes constrained by<br />

hanging nodes (due to level transitions <strong>of</strong> more than one between neighboring elements)<br />

is possible (left). The saturation <strong>of</strong> the error indicator ensures that there are<br />

level one transitions between neighboring elements only (right). . . . . . . . . . . . 46<br />

5.1 Sketch <strong>of</strong> the ingredients <strong>of</strong> a stochastic image. We discretize the spatial dimensions<br />

<strong>using</strong> finite elements, but the coefficients <strong>of</strong> the FE basis functions are random variables.<br />

Every random variable has a support, which spans over the complete image,<br />

thus pixels depend on a random vector. . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

5.2 Decay <strong>of</strong> the sorted eigenvalues <strong>of</strong> the centered covariance matrix <strong>of</strong> 45 input samples<br />

from an ultrasound device. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

5.3 Left picture group: The first mode (=expected value), second mode, third mode and<br />

fourth mode <strong>of</strong> a stochastic CT image. Right: The sinogram, i.e. the raw data produced<br />

by the CT imaging device for the head phantom [139]. . . . . . . . . . . . . . 51<br />

5.4 Second (left) and fifth (right) mode <strong>of</strong> a stochastic US image. The information encoded<br />

in these images is hard to interpret, because there is no deterministic equivalent. 53<br />

5.5 Expected value (left) and variance (right) <strong>of</strong> a stochastic US-image. The expected<br />

value looks like a deterministic image and in the variance, regions with a high gray<br />

value uncertainty are visible as white dots. . . . . . . . . . . . . . . . . . . . . . . . 54<br />

5.6 Two samples drawn from a stochastic image. The images differ due to realizations<br />

<strong>of</strong> the noise. In a printed version, these images look nearly the same. . . . . . . . . . 54<br />

5.7 Visualization <strong>of</strong> realizations <strong>of</strong> a stochastic 2D contour. Every yellow line corresponds<br />

to a MC realization <strong>of</strong> the stochastic contour encoded in the stochastic image. 55<br />

5.8 Visualization <strong>of</strong> a 3D contour encoded in a 3D stochastic image. The expected value<br />

<strong>of</strong> the 3D stochastic contour is color-coded by the variance. Regions with a high<br />

variance are red and regions with a low variance green. . . . . . . . . . . . . . . . . 55<br />

6.1 Expected value (top row) and variance (bottom row) <strong>of</strong> the street image (left) and the<br />

US image (right). Color-coded are the seed regions for interior (yellow) and exterior<br />

(red). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />

112


List <strong>of</strong> Figures<br />

6.2 Mean and variance <strong>of</strong> the probabilities for pixels to belong to the object. Furthermore,<br />

we show in red Monte Carlo realizations <strong>of</strong> the object boundary sampled from<br />

the stochastic result. A high variance indicates pixels where the gray value uncertainty<br />

highly influences the result. For comparison we added a classical random<br />

walker segmentation result in the last column. There the variance image is not available,<br />

because the method acts on a classical image. . . . . . . . . . . . . . . . . . . 61<br />

6.3 MC-realizations <strong>of</strong> the stochastic object boundary for the stochastic liver image segmented<br />

with the stochastic random walker approach with β = 10. On the right we<br />

highlight a region <strong>of</strong> the image, where the noise in the input image influences the<br />

result. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62<br />

6.4 PDF <strong>of</strong> the area <strong>of</strong> the segmented person from the street image for β = 25 (black)<br />

and β = 50 (gray). From the PDF we judge the reliability <strong>of</strong> the segmentation, a<br />

narrow PDF indicates that the image noise influences the segmentation marginally. . 63<br />

6.5 Comparison <strong>of</strong> the discretization methods for the computation <strong>of</strong> the stochastic random<br />

walker result to verify the intrusive discretization. The small difference between<br />

the intrusive discretization via the GSD method and the two other sampling based approaches<br />

might be due to the projection <strong>of</strong> the Laplacian matrix on the polynomial<br />

chaos. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

6.6 Input “doughnut” without noise (left) and noisy input image treated as expected value<br />

<strong>of</strong> the stochastic image (right). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

6.7 Left: The object seed points (yellow) and background seed points (red) used as initialization<br />

<strong>of</strong> the stochastic random walker method. Right: The MC-realizations <strong>of</strong><br />

the stochastic segmentation result differ significantly for different noise realizations. 66<br />

6.8 The PDF for both possibilities <strong>of</strong> the volume computation, the summation <strong>of</strong> the<br />

random variables (gray) and the thresholding (black). The true volume is 60 pixels. . 66<br />

6.9 Structure <strong>of</strong> the block system <strong>of</strong> an SPDE. Every block has the sparsity structure<br />

<strong>of</strong> a classical finite element matrix and the block structure <strong>of</strong> the matrix is sparse,<br />

meaning that some <strong>of</strong> the blocks are zero. The sparsity structure on the block level<br />

depends on the number <strong>of</strong> random variables and the polynomial chaos degree used<br />

in the discretization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69<br />

6.10 Nonzero pattern <strong>of</strong> the SFEM matrix for the smoothed stochastic image <strong>using</strong> n = 5<br />

random variables and a polynomial degree p = 3. A black dot denotes a block that<br />

has a nonzero stochastic part, thus having the sparsity structure <strong>of</strong> a classical FEM<br />

matrix. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

6.11 Mean value <strong>of</strong> the three data sets used to demonstrate the stochastic Ambrosio-<br />

Tortorelli method. For the second data set, we denoted image regions the text refers<br />

to. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

6.12 PDF <strong>of</strong> a pixel from the phase field computed from the polynomial chaos expansion<br />

<strong>of</strong> the pixel via a sampling approach. Although we use uniform basic random<br />

variables for the polynomial chaos, the resulting random variables have skewed and<br />

Gaussian like distributions due to the use <strong>of</strong> higher order polynomials in the basic<br />

random variables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

6.13 <strong>Segmentation</strong> result <strong>of</strong> the street scene. On the left we show the five samples the<br />

stochastic input image is computed from. On the right we compare the results computed<br />

via the GSD method and a Monte Carlo sampling. . . . . . . . . . . . . . . . 73<br />

6.14 Expected value and variance <strong>of</strong> the stochastic input image <strong>of</strong> the street scene. . . . . 73<br />

6.15 Mean and variance <strong>of</strong> the image and phase field for varying ε and µ <strong>using</strong> the US<br />

data. For comparison, we added the result from the deterministic method applied on<br />

the mean. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

113


List <strong>of</strong> Figures<br />

6.16 Comparison <strong>of</strong> the stochastic Ambrosio-Tortorelli model (left column) with the extended<br />

model <strong>using</strong> the edge linking procedure described in Section 2.3.3 (middle<br />

column) and a combination <strong>of</strong> the edge linking and adaptive grid approach (right<br />

column). Note that these results are computed with the same parameter set. The<br />

differences in the results are due to the additional edge linking parameter c only. . . . 75<br />

6.17 Comparison <strong>of</strong> the full grid and adaptive grid solution. The full grid and adaptive<br />

grid solution are visually identical, but the computation <strong>of</strong> the adaptive grid solution<br />

needs significantly less DOFs. Thus, it can be applied on high-resolution images. . . 77<br />

7.1 <strong>Stochastic</strong> level sets do not have a fixed position where φ(x) = 0. Instead, there is<br />

a band with positive probability that the level set is equal to zero, i.e. the position <strong>of</strong><br />

the zero level set is random and it is possible to estimate the PDF <strong>of</strong> the interface<br />

location in the normal direction <strong>of</strong> the expected value <strong>of</strong> the interface (lower right<br />

corner). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

7.2 Comparison <strong>of</strong> expected value and variance <strong>of</strong> the resulting phase field for the cosine<br />

test <strong>of</strong> (7.18) <strong>using</strong> the polynomial chaos (PC), stochastic collocation (SC), Monte<br />

Carlo simulation (MC), and Monte Carlo simulation <strong>of</strong> the original level set equation<br />

(MCL). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

7.3 Comparison <strong>of</strong> the expected value and variance <strong>of</strong> the resulting phase field for the<br />

rarefaction fan and the shock, two classical tests for level set propagation. The figure<br />

shows the comparison <strong>of</strong> the four discretizations <strong>of</strong> the stochastic phase field equation. 86<br />

7.4 Expected value color-coded by the variance for the Stanford bunny after shrinkage<br />

under an uncertain speed in the normal direction. Red indicates regions with a high<br />

variance and green regions with low variance. In addition, we show one slice <strong>of</strong> the<br />

variance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

7.5 Mean <strong>of</strong> the CT data set (left) and the liver data set (right) for the segmentation test. . 88<br />

7.6 Left: Mean contour during the evolution <strong>of</strong> the stochastic level set. The iso-contours<br />

are drawn on the variance image <strong>of</strong> the final, magenta contour. The contour detection<br />

is influenced by the image noise on the bottom and the right <strong>of</strong> the object (high<br />

variance). Right: Contour realizations <strong>of</strong> the stochastic gradient-based segmentation<br />

<strong>of</strong> the CT data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89<br />

7.7 Resulting image with the expected value <strong>of</strong> the contour (red) <strong>of</strong> the segmented object<br />

and the phase field variance with the expected value <strong>of</strong> the contour for gradientbased<br />

segmentation <strong>of</strong> a stochastic CT image. The variance is constant in the normal<br />

direction <strong>of</strong> the expected value <strong>of</strong> zero level set. . . . . . . . . . . . . . . . . . . . . 90<br />

7.8 Mean and variance <strong>of</strong> the stochastic geodesic active contour segmentation <strong>of</strong> the<br />

stochastic CT data set. The variance is constant in the normal direction <strong>of</strong> the zero<br />

level set. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91<br />

7.9 Left: Evolution <strong>of</strong> the expected value contour <strong>of</strong> the stochastic geodesic active contour<br />

method. The shown variance corresponds to the contour after 240 iterations (the<br />

magenta contour). Right: Mean value <strong>of</strong> the stochastic image to be segmented and<br />

the contours at time points <strong>of</strong> the level set evolution. The final contour matches the<br />

object boundary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92<br />

7.10 Mean (left) <strong>of</strong> the stochastic CT and the variance (right) <strong>of</strong> the stochastic Chan-Vese<br />

solution. Additionally, we show the expected value contour at different time steps. . . 93<br />

7.11 Left: MC-realizations <strong>of</strong> the stochastic contour from the stochastic Chan-Vese segmentation<br />

applied on the CT data set. Right: Realizations <strong>of</strong> the stochastic contour<br />

from the stochastic geodesic active contour approach applied on the CT data set. . . . 94<br />

114


List <strong>of</strong> Figures<br />

7.12 Mean (left) <strong>of</strong> the stochastic liver image and the variance <strong>of</strong> the stochastic Chan-Vese<br />

solution. In addition, we show the expected value contour at different time steps. . . 95<br />

7.13 Variance <strong>of</strong> the stochastic image to segment (left), the expected value is not depicted,<br />

because the expected value is an image with the same gray value at every pixel. On<br />

the right, the segmentation result is depicted on one realization (one sample) <strong>of</strong> the<br />

stochastic image to segment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95<br />

8.1 Left: Realizations <strong>of</strong> the stochastic contour obtained from the random walker segmentation<br />

with stochastic parameter. Right: Mean and variance <strong>of</strong> the stochastic<br />

contour obtained from the random walker segmentation with stochastic parameter. . . 99<br />

8.2 Left: Realizations <strong>of</strong> the stochastic contour obtained from the random walker segmentation<br />

with stochastic parameter. Right: Mean and variance <strong>of</strong> the stochastic<br />

contour obtained from the random walker segmentation with stochastic parameter. . . 100<br />

8.3 Volume <strong>of</strong> the stochastic contour obtained from the random walker segmentation<br />

with stochastic parameter. The left curve shows the PDF <strong>of</strong> the object in the US<br />

image, the right curve the PDF <strong>of</strong> the liver in the liver data set. The PDFs are obtained<br />

<strong>using</strong> (6.12). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100<br />

8.4 Ambrosio-Tortorelli model applied on the expected value <strong>of</strong> the liver data set <strong>using</strong><br />

a stochastic parameter µ. The upper row shows the expected value (left) and the<br />

variance (right) <strong>of</strong> the smoothed image, the lower row the expected value (left) and<br />

the variance (right) <strong>of</strong> the phase field. . . . . . . . . . . . . . . . . . . . . . . . . . 102<br />

8.5 Ambrosio-Tortorelli model applied on the expected value <strong>of</strong> the liver data set <strong>using</strong><br />

a stochastic parameter ε. The upper row shows the expected value (left) and the<br />

variance (right) <strong>of</strong> the smoothed image and the lower row the expected value (left)<br />

and the variance (right) <strong>of</strong> the phase field. . . . . . . . . . . . . . . . . . . . . . . . 103<br />

8.6 Result <strong>of</strong> the gradient-based segmentation with stochastic parameter b, i.e. with a<br />

stochastic curvature smoothing. The upper row shows the results for the CT data set,<br />

expected value <strong>of</strong> the image and contour realizations (left) and variance <strong>of</strong> the level<br />

set with contour realizations (right). The lower row shows the same results for the<br />

liver data set. In all figures we add Monte Carlo realizations <strong>of</strong> the stochastic object<br />

boundary. The red contour corresponds to a b = 0.75, yellow to b = 1.0 and blue to<br />

b = 1.25. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104<br />

8.7 Result <strong>of</strong> the geodesic active contour segmentation with stochastic parameters for<br />

the CT data set. On the left the expected value <strong>of</strong> the image and contour realizations<br />

and on the right the variance <strong>of</strong> the level set with contour realizations are shown. . . 105<br />

8.8 Result <strong>of</strong> the geodesic active contour segmentation with stochastic parameters for<br />

the liver data set. On the left the expected value <strong>of</strong> a detail <strong>of</strong> the image and contour<br />

realizations and on the right the variance <strong>of</strong> the level set with contour realizations are<br />

shown. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106<br />

9.1 Visualization <strong>of</strong> the PDF <strong>of</strong> the object boundary in the case <strong>of</strong> the segmentation <strong>of</strong><br />

the stochastic US image. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108<br />

115


List <strong>of</strong> Tables<br />

3.1 The first ten one-dimensional Legendre-polynomials. The multi-dimensional polynomials<br />

up to degree nine are based on these polynomials and (3.40). . . . . . . . . . 32<br />

3.2 Important distributions and the corresponding polynomials for the expansion. . . . . 32<br />

3.3 The first ten one-dimensional Hermite-polynomials. The construction <strong>of</strong> the multidimensional<br />

polynomials up to degree 9 is based on these polynomials and (3.40). . . 33<br />

6.1 Comparison <strong>of</strong> the execution times (in sec) <strong>of</strong> the discretization methods. . . . . . . 64<br />

117


Appendix A<br />

Publications Written During the Course<br />

<strong>of</strong> the Thesis<br />

Parts <strong>of</strong> the results <strong>of</strong> this thesis are already published or submitted for publication. Besides the<br />

publications related to this thesis, the author published results about the simulation <strong>of</strong> radio frequency<br />

(RF) ablation. We give a short introduction into RF ablation before we list the papers.<br />

A.1 Publications Related to <strong>Stochastic</strong> <strong>Images</strong><br />

[1] T. Pätz, R. M. Kirby, and T. Preusser. Ambrosio-Tortorelli segmentation <strong>of</strong> stochastic images:<br />

Model extensions, theoretical investigations and numerical methods. Submitted to International<br />

Journal <strong>of</strong> Computer Vision, 2011.<br />

[2] T. Pätz, R. M. Kirby, and T. Preusser. <strong>Segmentation</strong> <strong>of</strong> stochastic images <strong>using</strong> stochastic extensions<br />

<strong>of</strong> the Ambrosio-Tortorelli and the random walker model. PAMM, 11(1):859–860, 2011.<br />

[3] T. Pätz and T. Preusser. Ambrosio-Tortorelli segmentation <strong>of</strong> stochastic images. In K. Daniilidis,<br />

P. Maragos, and N. Paragios, editors, Computer Vision - ECCV 2010, volume 6315 <strong>of</strong> Lecture<br />

Notes in Computer Science, pages 254–267. Springer Berlin / Heidelberg, 2010. (This paper<br />

received the ECCV 2010 Best Student Paper Award.).<br />

[4] T. Pätz and T. Preusser. <strong>Segmentation</strong> <strong>of</strong> stochastic images <strong>using</strong> level set propagation with<br />

uncertain speed. In preparation, 2011.<br />

[5] T. Pätz and T. Preusser. <strong>Segmentation</strong> <strong>of</strong> stochastic images with a stochastic random walker<br />

method. Submitted to IEEE Transactions on Image Processing, 2011.<br />

[6] T. Pätz and T. Preusser. Variational image segmentation <strong>using</strong> stochastic parameters. In preparation,<br />

2011.<br />

A.2 Publications Related to Radi<strong>of</strong>requency Ablation<br />

RF ablation is a minimally invasive technique for a local ablation <strong>of</strong> abnormal tissue, like primary or<br />

metastatic cancer. During the last years, RF ablation has become an alternative to the surgical resection<br />

<strong>of</strong> the tumor. At the beginning <strong>of</strong> the treatment, an internally cooled RF probe is percutaneously<br />

placed inside the tissue and connected to an RF generator. The generator delivers an electric current<br />

in the radio-frequency range (typically 500 kHz) with a power between 25W and 200W. Due to the<br />

electric impedance, the tissue close to the probe is heated and above 60 ◦ C it is destroyed.<br />

The modeling and simulation <strong>of</strong> RF ablation is a multiple investigated research topic (see [20] for<br />

a review). Many scientists presented simulations with varying detail, because multiple biophysical<br />

effects take place during the ablations. Another challenge is the modeling <strong>of</strong> the physical parameters<br />

influencing the ablation outcome, because these parameters are (nonlinearly) influenced by biophysical<br />

effects. For example, the electric conductivity is nonlinearly dependent on the temperature, the<br />

119


vaporization state, and the coagulation state <strong>of</strong> the tissue. The simulation <strong>of</strong> RF ablation typically<br />

uses a coupled system <strong>of</strong> PDEs for the electric potential and the heat transfer.<br />

[7] I. Altrogge, T. Pätz, T. Kröger, H.-O. Peitgen, and T. Preusser. Optimization and fast estimation<br />

<strong>of</strong> vessel cooling for RF ablation. In World Congress on Medical Physics and Biomedical<br />

Engineering, September 2009, Munich, Germany, volume 25/4 <strong>of</strong> IFMBE Proceedings, pages<br />

1202–1205. Springer, 2010.<br />

[8] I. Altrogge, T. Preusser, T. Kröger, S. Haase, T. Pätz, and R. M. Kirby. Sensitivity analysis for<br />

the optimization <strong>of</strong> radi<strong>of</strong>requency ablation in the presence <strong>of</strong> material parameter uncertainty.<br />

Submitted to International Journal for Uncertainty Quantification, 2011.<br />

[9] T. Kröger, T. Pätz, I. Altrogge, A. Schenk, K. S. Lehmann, B. B. Frericks, J.-P. Ritz, H.-O.<br />

Peitgen, and T. Preusser. Fast estimation <strong>of</strong> the vascular cooling in RFA based on numerical<br />

simulation. Open Biomed Eng J, 4:16–26, 2010.<br />

[10] T. Pätz, T. Kröger, and T. Preusser. Simulation <strong>of</strong> radi<strong>of</strong>requency ablation including water evaporation.<br />

In World Congress on Medical Physics and Biomedical Engineering, September 2009,<br />

Munich, Germany, volume 25/4 <strong>of</strong> IFMBE Proceedings, pages 1287–1290. Springer, 2010.<br />

[11] T. Pätz and T. Preusser. Simulation <strong>of</strong> water evaporation during radi<strong>of</strong>requency ablation <strong>using</strong><br />

composite finite elements. In Proceedings <strong>of</strong> the 1st Conference on Multiphysics Simulation –<br />

Advanced Methods for Industrial Engineering, 2010.<br />

[12] T. Pätz and T. Preusser. Composite finite elements for a phase-change model. Submitted to<br />

SIAM Journal on Scientific Computing, 2011.<br />

[13] T. Pätz and T. Preusser. Simulation <strong>of</strong> water evaporation during radi<strong>of</strong>requency ablation <strong>using</strong><br />

composite finite elements. The International Journal <strong>of</strong> Multiphysics, Special Edition: Multiphysics<br />

Simulations – Advanced Methods for Industrial Engineering, pages 145–156, 2011.<br />

120


Bibliography<br />

[14] L. Ambrosio and M. Tortorelli. Approximation <strong>of</strong> functionals depending on jumps by elliptic<br />

functionals via Gamma-convergence. Communications on Pure and Applied Mathematics,<br />

43(8):999–1036, 1990.<br />

[15] L. Ambrosio and M. Tortorelli. On the approximation <strong>of</strong> free discontinuity problems. Bollettino<br />

della Unione Matematica Italiana, Sezione B, 6:105–123, 1992.<br />

[16] R. Askey and J. Wilson. Some basic hypergeometric polynomials that generalize Jacobi polynomials.<br />

Memoirs <strong>of</strong> the American Mathematical Society, 319, 1985.<br />

[17] G. Aubert and P. Kornprobst. Mathematical Problems in Image Processing: Partial Differential<br />

Equations and the Calculus <strong>of</strong> Variations (second edition), volume 147 <strong>of</strong> Applied<br />

Mathematical Sciences. Springer-Verlag, 2006.<br />

[18] I. Babuška, R. Tempone, and G. E. Zouraris. Solving elliptic boundary value problems with<br />

uncertain coefficients by the finite element method: the stochastic formulation. Computer<br />

Methods in Applied Mechanics and Engineering, 194(12-16):1251–1294, 2005. Special Issue<br />

on Computational Methods in <strong>Stochastic</strong> Mechanics and Reliability Analysis.<br />

[19] J. Ballani and L. Grasedyck. A projection method to solve linear systems in tensor format.<br />

Preprint 46, DFG-Schwerpunktprogramm 1324, 2010.<br />

[20] E. J. Berjano. Theoretical modeling for radi<strong>of</strong>requency ablation: state-<strong>of</strong>-the-art and challenges<br />

for the future. BioMedical Engineering OnLine, 5(1):24, 2006.<br />

[21] J. Beutel. Handbook <strong>of</strong> medical imaging: Physics and psychophysics. SPIE Press, 2000.<br />

[22] G. Blatman and B. Sudret. Sparse polynomial chaos expansions and adaptive stochastic finite<br />

elements <strong>using</strong> a regression approach. Comptes Rendus Mécanique, 336(6):518–523, 2008.<br />

[23] W. J. Boettinger, J. A. Warren, C. Beckermann, and A. Karma. Phase-field simulation <strong>of</strong><br />

solidification. Annual Review <strong>of</strong> Materials Research, 32(1):163–194, 2002.<br />

[24] S. Boyaval, C. L. Bris, Y. Maday, N. C. Nguyen, and A. T. Patera. A reduced basis approach<br />

for variational problems with stochastic parameters: Application to heat conduction<br />

with variable Robin coefficient. Computer Methods in Applied Mechanics and Engineering,<br />

198(41-44):3187–3206, 2009.<br />

[25] Y. Boykov and G. Funka-Lea. Graph cuts and efficient n-d image segmentation. International<br />

Journal <strong>of</strong> Computer Vision, 70:109–131, 2006.<br />

[26] D. Cacuci, M. Ionescu-Bujor, and I. Navon. Sensitivity and Uncertainty Analysis: Applications<br />

to large-scale systems. Chapman & Hall/CRC Press, 2005.<br />

[27] R. H. Cameron and W. T. Martin. The orthogonal development <strong>of</strong> non-linear functionals in<br />

series <strong>of</strong> Fourier-Hermite functionals. The Annals <strong>of</strong> Mathematics, 48(2):385–392, 1947.<br />

[28] G. Casella and R. Berger. Statistical inference. Thomson Learning, 2002.<br />

121


Bibliography<br />

[29] V. Caselles, F. Catte, B. Coll, and F. Dibos. A geometric model for edge detection. Numerische<br />

Mathematik, 66:1–31, 1993.<br />

[30] V. Caselles, R. Kimmel, and G. Sapiro. Geodesic active contours. International Journal <strong>of</strong><br />

Computer Vision, 22(1):61–79, 1997.<br />

[31] T. Chan and L. Vese. Active contours without edges. IEEE Transactions on Image Processing,<br />

10(2):266–277, 2001.<br />

[32] W. Cheney. Analysis for Applied Mathematics. Springer, 2001.<br />

[33] C. W. Clenshaw and A. R. Curtis. A method for numerical integration on an automatic computer.<br />

Numerische Mathematik, 2:197–205, 1960.<br />

[34] J. Condell, B. Scotney, and P. Morrow. Adaptive grid refinement procedures for efficient<br />

optical flow computation. International Journal <strong>of</strong> Computer Vision, 61:31–54, 2005.<br />

[35] E. De Giorgi, M. Carriero, and A. Leaci. Existence theorem for a minimum problem with free<br />

discontinuity set. Archive for Rational Mechanics and Analysis, 108:195–218, 1989.<br />

[36] M. De Santo, C. Liguori, and A. Pietrosanto. Uncertainty characterization in image-based<br />

measurements: a preliminary discussion. IEEE Transactions on Instrumentation and Measurement,<br />

49(5):1101–1107, 2000.<br />

[37] L. Debnath and P. Mikusinski. Introduction to Hilbert Spaces with Applications. Academic<br />

Press, 2006.<br />

[38] B. J. Debusschere, H. N. Najm, P. P. Pébay, O. M. Knio, R. G. Ghanem, and O. P. Le Maître.<br />

Numerical challenges in the use <strong>of</strong> polynomial chaos representations for stochastic processes.<br />

SIAM Journal on Scientific Computing, 26(2):698–719, 2005.<br />

[39] H. Deng and D. A. Clausi. Unsupervised image segmentation <strong>using</strong> a simple MRF model<br />

with a new implementation scheme. Pattern Recognition, 37(12):2323–2335, 2004.<br />

[40] A. Dervieux and F. Thomasset. A finite element method for the simulation <strong>of</strong> a Rayleigh-<br />

Taylor instability. In R. Rautmann, editor, Approximation Methods for Navier-Stokes Problems,<br />

volume 771 <strong>of</strong> Lecture Notes in Mathematics, pages 145–158. Springer, 1980.<br />

[41] C. Desceliers, R. Ghanem, and C. Soize. Maximum likelihood estimation <strong>of</strong> stochastic chaos<br />

representations from experimental data. International Journal for Numerical Methods in Engineering,<br />

66:978–1001, 2006.<br />

[42] P. Dirac. The principles <strong>of</strong> quantum mechanics. International series <strong>of</strong> monographs on physics.<br />

Clarendon Press, 1981.<br />

[43] S. Djurcilov, K. Kim, P. F. J. Lermusiaux, and A. Pang. Volume rendering data with uncertainty<br />

information. Data Visualization (Proceedings <strong>of</strong> the EG+IEEE VisSym), pages 243–<br />

252, 2001.<br />

[44] G. Dougherty. Digital Image Processing for Medical Applications. Cambridge <strong>University</strong><br />

Press, 2009.<br />

[45] P. G. Doyle and J. L. Snell. Random walks and electric networks. Carus Mathematical<br />

Monographs., 22, 1984.<br />

122


Bibliography<br />

[46] B. Echebarria, R. Folch, A. Karma, and M. Plapp. Quantitative phase-field model <strong>of</strong> alloy<br />

solidification. Physical Review E, 70(6):061604, 2004.<br />

[47] M. Eiermann, O. G. Ernst, and E. Ullmann. Computational aspects <strong>of</strong> the stochastic finite<br />

element method. Computing and Visualization in Science, 10(1):3–15, 2007.<br />

[48] E. Eisenhauer, P. Therasse, J. Bogaerts, L. Schwartz, D. Sargent, R. Ford, J. Dancey, S. Arbuck,<br />

S. Gwyther, M. Mooney, L. Rubinstein, L. Shankar, L. Dodd, R. Kaplan, D. Lacombe,<br />

and J. Verweij. New response evaluation criteria in solid tumours: Revised recist guideline<br />

(version 1.1). European Journal <strong>of</strong> Cancer, 45(2):228–247, 2009.<br />

[49] E. Erdem, A. Sancar-Yilmaz, and S. Tari. Mumford-Shah regularizer with spatial coherence.<br />

In F. Sgallari, A. Murli, and N. Paragios, editors, Scale Space and Variational Methods<br />

in Computer Vision, volume 4485 <strong>of</strong> Lecture Notes in Computer Science, pages 545–555.<br />

Springer Berlin / Heidelberg, 2007.<br />

[50] O. G. Ernst, A. Mugler, H.-J. Starkl<strong>of</strong>f, and E. Ullmann. On the convergence <strong>of</strong> generalized<br />

polynomial chaos expansions. ESAIM: Mathematical Modelling and Numerical Analysis,<br />

46:317–339, 2012.<br />

[51] R. Folch, J. Casademunt, A. Hernandez-Machado, and L. Ramirez-Piscina. Phase-field model<br />

for Hele-Shaw flows with arbitrary viscosity contrast. I. Theoretical approach. Physical Review<br />

E, 60:1724, 1999.<br />

[52] K. Frank and S. Heinrich. Computing discrepancies <strong>of</strong> Smolyak quadrature rules. Journal <strong>of</strong><br />

Complexity, 12(4):287–314, 1996.<br />

[53] D. Fransos. <strong>Stochastic</strong> Numerical Models for Wind Engineering. PhD thesis, Politecnico di<br />

Torino, 2008.<br />

[54] R. G. Ghanem and P. D. Spanos. <strong>Stochastic</strong> finite elements: a spectral approach. Springer-<br />

Verlag, New York, 1991.<br />

[55] F. Gibou, L. Chen, D. Nguyen, and S. Banerjee. A level set based sharp interface method<br />

for the multiphase incompressible Navier-Stokes equations with phase change. Journal <strong>of</strong><br />

Computational Physics, 222(2):536–555, 2007.<br />

[56] K. Glasner. Nonlinear preconditioning for diffuse interfaces. Journal <strong>of</strong> Computational<br />

Physics, 174(2):695–711, 2001.<br />

[57] G. H. Golub and C. F. V. Loan. Matrix Computations. The Johns Hopkins <strong>University</strong> Press,<br />

3rd edition, 1996.<br />

[58] G. H. Golub and C. F. van Loan. An analysis <strong>of</strong> the total least squares problem. SIAM Journal<br />

on Numerical Analysis, 17(6):883–893, 1980.<br />

[59] L. Grady. Random walks for image segmentation. IEEE Transactions on Pattern Analysis<br />

and Machine Intelligence, 28(11):1768–1783, 2006.<br />

[60] L. Grasedyck. Hierarchical singular value decomposition <strong>of</strong> tensors. SIAM Journal on Matrix<br />

Analysis and Applications, 31(4):2029–2054, 2010.<br />

[61] H. Griethe and H. Schumann. The visualization <strong>of</strong> uncertain data: Methods and problems. In<br />

T. Schulze, G. Horton, B. Preim, and S. Schlechtweg, editors, Simulation und Visualisierung<br />

2006 (SimVis 2006), Magdeburg, pages 143–156. SCS Publishing House e.V., 2006.<br />

123


Bibliography<br />

[62] J. Gubner. Probability and random processes for electrical and computer engineers. Cambridge<br />

<strong>University</strong> Press., 2006.<br />

[63] H. Harbrecht, M. Peters, and R. Schneider. On the low-rank approximation by the pivoted<br />

Cholesky decomposition. Applied Numerical Mathematics, 2011. In press.<br />

[64] B. Hayes. A lucid interval. American Scientist, 91(6):484–488, 2003.<br />

[65] K. Held, E. Kops, B. Krause, W. Wells, R. Kikinis, and H.-W. Müller-Gärtner. Markov random<br />

field segmentation <strong>of</strong> brain MR images. IEEE Transactions on Medical Imaging, 16(6):878–<br />

886, 1997.<br />

[66] G. Herman. Fundamentals <strong>of</strong> Computerized Tomography: Image Reconstruction from Projections.<br />

Advances in pattern recognition. Springer, 2009.<br />

[67] M. R. Hestenes and E. Stiefel. Methods <strong>of</strong> conjugate gradients for solving linear systems.<br />

Journal <strong>of</strong> Research <strong>of</strong> the National Bureau <strong>of</strong> Standards, 49(6):409–436, 1952.<br />

[68] T. Hida and N. Ikeda. Analysis on Hilbert space with reproducing kernel arising from multiple<br />

Wiener integral. In L. M. Le Cam and J. Neyman, editors, Proceedings <strong>of</strong> the Fifth Berkeley<br />

Symposium on Mathematical Statistics and Probability. <strong>University</strong> <strong>of</strong> California Press, 1967.<br />

[69] H. Holden, B. Øksendal, J. Ubøe, and T. Zhang. <strong>Stochastic</strong> partial differential equations. A<br />

modeling, white noise functional approach. 2nd ed. Universitext. New York, Springer, 2010.<br />

[70] O. Holzmann, B. Lang, and H. Schütt. Newton’s constant <strong>of</strong> gravitation and verified numerical<br />

quadrature. Reliable Computing, 2:229–239, 1996.<br />

[71] B. K. P. Horn and B. G. Schunck. Determining optical flow. Artificial Intelligence, 17:185–<br />

203, 1981.<br />

[72] I. James. The topology <strong>of</strong> Stiefel manifolds. London Mathematical Society lecture note series.<br />

Cambridge <strong>University</strong> Press, 1976.<br />

[73] S. Janson. Gaussian Hilbert Spaces. Cambridge <strong>University</strong> Press, 1997.<br />

[74] I. Jolliffe. Principal Component Analysis. Springer, 2002.<br />

[75] O. Juan, R. Keriven, and G. Postelnicu. <strong>Stochastic</strong> motion and the level set method in computer<br />

vision: <strong>Stochastic</strong> active contours. International Journal <strong>of</strong> Computer Vision, 69(1):7–<br />

25, 2006.<br />

[76] M. Kass, A. Witkin, and D. Terzopoulos. Snakes: Active contour models. International<br />

Journal <strong>of</strong> Computer Vision, 1(4):321–331, 1988.<br />

[77] S. Kay. Fundamentals Of Statistical Signal Processing. Prentice Hall, 2001.<br />

[78] R. B. Kearfott. Interval computations: Introduction, uses, and resources. Euromath Bulletin,<br />

2:95–112, 1996.<br />

[79] M. Kendall, A. Stuart, J. Ord, and A. O’Hagan. Kendall’s advanced theory <strong>of</strong> statistics,<br />

volume 1. Edward Arnold, 1994.<br />

[80] B. Khoromskij and I. Oseledets. Quantics-TT collocation approximation <strong>of</strong> parameterdependent<br />

and stochastic elliptic PDEs. Computational Methods in Applied Mathematics,<br />

10(4):376–394, 2010.<br />

124


Bibliography<br />

[81] B. N. Khoromskij and C. Schwab. Tensor-structured Galerkin approximation <strong>of</strong> parametric<br />

and stochastic elliptic PDEs. SIAM Journal on Scientific Computing, 33(1):364–385, 2011.<br />

[82] S. Kichenassamy, A. Kumar, P. Olver, A. Tannenbaum, and A. Yezzi. Gradient flows and<br />

geometric active contour models. In Fifth International Conference on Computer Vision,<br />

1995. Proceedings., pages 810–815, 1995.<br />

[83] S. Kirkpatrick, C. D. Gelatt, and M. P. Vecchi. Optimization by simulated annealing. Science,<br />

220(4598):671–680, 1983.<br />

[84] O. M. Knio and O. P. Le Maître. Uncertainty propagation in CFD <strong>using</strong> polynomial chaos<br />

decomposition. Fluid Dynamics Research, 38(9):616–640, 2006.<br />

[85] Y. G. Kondratiev, P. Leukert, and L. Streit. Wick calculus in Gaussian analysis. Acta Applicandae<br />

Mathematicae, 44:269–294, 1996.<br />

[86] K. Krajsek, I. Dedovic, and H. Scharr. An estimation theoretical approach to Ambrosio-<br />

Tortorelli image segmentation. In R. Mester and M. Felsberg, editors, Pattern Recognition,<br />

volume 6835 <strong>of</strong> Lecture Notes in Computer Science, pages 41–50. Springer, 2011.<br />

[87] T. Kröger, I. Altrogge, O. Konrad, R. M. Kirby, and T. Preusser. Estimation <strong>of</strong> probability density<br />

functions for parameter sensitivity analyses. In H. Hauser, S. Strassburger, and H. Theisel,<br />

editors, SimVis, pages 61–74. SCS Publishing House e.V., 2008.<br />

[88] D. Landau and K. Binder. A guide to Monte Carlo simulations in statistical physics. Cambridge<br />

<strong>University</strong> Press, 2005.<br />

[89] Y. Law, H. Lee, and A. Yip. A multiresolution stochastic level set method for Mumford-Shah<br />

image segmentation. IEEE Transactions on Image Processing, 17(12):2289–2300, 2008.<br />

[90] C. Li, C. Xu, C. Gui, and M. D. Fox. Level set evolution without re-initialization: A new<br />

variational formulation. In 2005 IEEE Computer Society Conference on Computer Vision and<br />

Pattern Recognition (CVPR 2005), 20-26 June 2005, pages 430–436, 2005.<br />

[91] Z. Liang, P. Lauterbur, I. E. in Medicine, and B. Society. Principles <strong>of</strong> magnetic resonance<br />

imaging: a signal processing perspective. IEEE Press Series in Biomedical Engineering. SPIE<br />

Optical Engineering Press, 2000.<br />

[92] G. Lin, C.-H. Su, and G. Karniadakis. Predicting shock dynamics in the presence <strong>of</strong> uncertainties.<br />

Journal <strong>of</strong> Computational Physics, 217(1):260–276, 2006.<br />

[93] G. Lin, X. Wan, C.-H. Su, and G. E. Karniadakis. <strong>Stochastic</strong> computational fluid mechanics.<br />

Computing in Science and Engineering, 9(2):21–29, 2007.<br />

[94] M. Ljungberg, S. Strand, and M. King. Monte Carlo calculations in nuclear medicine: applications<br />

in diagnostic imaging. Medical Science Series. Taylor & Francis, 1998.<br />

[95] M. Loève. Probability theory. Springer-Verlag, New York, 4th edition, 1977.<br />

[96] R. Malladi, J. A. Sethian, and B. C. Vemuri. Evolutionary fronts for topology-independent<br />

shape modeling and recovery. In ECCV ’94: Proceedings <strong>of</strong> the third European conference<br />

on Computer vision (vol. 1), pages 3–13. Springer-Verlag New York, 1994.<br />

125


Bibliography<br />

[97] M. Matsumoto and T. Nishimura. Mersenne twister: a 623-dimensionally equidistributed<br />

uniform pseudo-random number generator. ACM Transactions on Modeling and Computer<br />

Simulation, 8(1):3–30, 1998.<br />

[98] H. G. Matthies and B. V. Rosic. Inelastic media under uncertainty: <strong>Stochastic</strong> models and<br />

computational approaches. In B. D. Reddy, editor, IUTAM Symposium on Theoretical, Computational<br />

and Modelling Aspects <strong>of</strong> Inelastic Media, volume 11 <strong>of</strong> IUTAM Bookseries, pages<br />

185–194. Springer Netherlands, 2008.<br />

[99] B. McCane, K. Novins, D. Crannitch, and B. Galvin. On benchmarking optical flow. Computer<br />

Vision and Image Understanding, 84(1):126–143, 2001.<br />

[100] N. Metropolis. The beginning <strong>of</strong> the Monte Carlo method. Los Alamos Science, pages 125–<br />

130, 1987.<br />

[101] N. Metropolis and S. Ulam. The Monte Carlo method. Journal <strong>of</strong> the American Statistical<br />

Association, 44(247):335–341, 1949.<br />

[102] T. Minamoto and K. Aoki. A blind digital image watermarking method <strong>using</strong> interval wavelet<br />

decomposition. International Journal <strong>of</strong> Signal Processing, Image Processing and Pattern<br />

Recognition, 3(2):59–72, 2010.<br />

[103] J. Modersitzki. Numerical methods for image registration. Numerical mathematics and scientific<br />

computation. Oxford <strong>University</strong> Press, 2004.<br />

[104] R. Moore, R. Kearfott, and M. Cloud. Introduction to interval analysis. Society for Industrial<br />

and Applied Mathematics, 2009.<br />

[105] F. Morgan and J. F. Bredt. Geometric Measure Theory – A Beginner’s Guide. Elsevier, 2000.<br />

[106] D. Mumford. The Dawning <strong>of</strong> the Age <strong>of</strong> <strong>Stochastic</strong>ity. In Mathematics: frontiers and<br />

perspectives, pages 197–218, 2000.<br />

[107] D. Mumford and J. Shah. Optimal approximations by piecewise smooth functions and associated<br />

variational problems. Communications on Pure and Applied Mathematics, 42(5):577–<br />

685, 1989.<br />

[108] M. Musiela and T. Zariphopoulou. <strong>Stochastic</strong> partial differential equations and portfolio<br />

choice. In Contemporary Quantitative Finance, pages 195–216. Springer, 2010.<br />

[109] H. N. Najm. Uncertainty quantification and polynomial chaos techniques in computational<br />

fluid dynamics. Annual Review <strong>of</strong> Fluid Mechanics, 41(1):35–52, 2009.<br />

[110] V. A. B. Narayanan and N. Zabaras. Variational multiscale stabilized FEM formulations<br />

for transport equations: stochastic advection-diffusion and incompressible stochastic Navier-<br />

Stokes equations. Journal <strong>of</strong> Computational Physics, 202(1):94–133, 2005.<br />

[111] O. Nestares and D. Fleet. Error-in-variables likelihood functions for motion estimation. In<br />

International Conference on Image Processing 2003 (ICIP 2003). Proceedings, volume 3,<br />

pages 77–80, 2003.<br />

[112] O. Nestares, D. J. Fleet, and D. J. Heeger. Likelihood functions and confidence bounds for<br />

total-least-squares problems. IEEE Computer Society Conference on Computer Vision and<br />

Pattern Recognition, 1:523–530, 2000.<br />

126


Bibliography<br />

[113] A. Nouy. A generalized spectral decomposition technique to solve a class <strong>of</strong> linear stochastic<br />

partial differential equations. Computer Methods in Applied Mechanics and Engineering,<br />

196(45-48):4521–4537, 2007.<br />

[114] A. Nouy. Generalized spectral decomposition method for solving stochastic finite element<br />

equations: invariant subspace problem and dedicated algorithms. Computer Methods in Applied<br />

Mechanics and Engineering, 197(51-52):4718–4736, 2008.<br />

[115] A. Nouy. Recent developments in spectral stochastic methods for the numerical solution <strong>of</strong><br />

stochastic partial differential equations. Archives <strong>of</strong> Computational Methods in Engineering,<br />

16(3):251–285, 2009.<br />

[116] A. Nouy. Identification <strong>of</strong> multi-modal random variables through mixtures <strong>of</strong> polynomial<br />

chaos expansions. Comptes Rendus Mécanique, 338(12):698–703, 2010.<br />

[117] A. Nouy. Proper generalized decompositions and separated representations for the numerical<br />

solution <strong>of</strong> high dimensional stochastic problems. Archives <strong>of</strong> Computational Methods in<br />

Engineering, 17(4):403–434, 2010.<br />

[118] A. Nouy and O. P. Le Maître. Generalized spectral decomposition for stochastic nonlinear<br />

problems. Journal <strong>of</strong> Computational Physics, 228(1):202–235, 2009.<br />

[119] E. Novak and K. Ritter. The curse <strong>of</strong> dimension and a universal method for numerical integration.<br />

In Multivariate Approximation and Splines, pages 177–187, 1998.<br />

[120] M. Ohlberger and M. Rumpf. Adaptive projection operators in multiresolutional scientific<br />

visualization. IEEE Transactions on Visualization and Computer Graphics, 5(4):74–94, 1999.<br />

[121] S. Osher and J. A. Sethian. Fronts propagating with curvature dependent speed: Algorithms<br />

based on Hamilton-Jacobi formulations. Journal <strong>of</strong> Computational Physics, 79:12–49, 1988.<br />

[122] K. Petras. Asymptotically minimal Smolyak cubature. Technical report, Technische Universität<br />

Braunschweig, 1999.<br />

[123] T. Pock and H. Bisch<strong>of</strong>. A probabilistic multi-phase model for variational image segmentation.<br />

In K. Franke, K.-R. Müller, B. Nickolay, and R. Schäfer, editors, Pattern Recognition, 28th<br />

DAGM Symposium, Berlin, Germany, September 12–14, 2006, Proceedings, volume 4174 <strong>of</strong><br />

Lecture Notes in Computer Science, pages 71–80. Springer, 2006.<br />

[124] A. Polycarpou. Introduction to the finite element method in electromagnetics. Synthesis<br />

lectures on computational electromagnetics. Morgan & Claypool Publishers, 2006.<br />

[125] K. Potter, J. Krüger, and C. Johnson. Towards the visualization <strong>of</strong> multi-dimensional stochastic<br />

distribution data. In Proceedings <strong>of</strong> The International Conference on Computer Graphics<br />

and Visualization (IADIS) 2008, 2008.<br />

[126] J.-S. Prassni, T. Ropinski, and K. Hinrichs. Uncertainty-aware guided volume segmentation.<br />

IEEE Transactions on Visualization and Computer Graphics, 16:1358–1365, 2010.<br />

[127] W. Pratt. Digital image processing: PIKS Scientific inside. Wiley-Interscience, 2007.<br />

[128] T. Preusser and H.-O. Peitgen. Patient-specific planning for radio-frequency ablation <strong>of</strong> tumors<br />

in the presence <strong>of</strong> uncertainty. it - Information Technology, 52:265–271, 2010.<br />

127


Bibliography<br />

[129] T. Preusser and M. Rumpf. An adaptive finite element method for large scale image processing.<br />

Journal <strong>of</strong> Visual Communication and Image Representation, 11(2):183–195, 2000.<br />

[130] T. Preusser, H. Scharr, K. Krajsek, and R. Kirby. Building blocks for computer vision with<br />

stochastic partial differential equations. International Journal <strong>of</strong> Computer Vision, 80(3):375–<br />

405, 2008.<br />

[131] S. Rajan, S. Wang, R. Inkol, and A. Joyal. Efficient approximations for the arctangent function.<br />

IEEE Signal Processing Magazine, 23(3):108–111, 2006.<br />

[132] M. M. Rao and R. J. Swift. Probability Theory with Applications. Springer-Verlag, 2006.<br />

[133] D. W. O. Rogers. Fifty years <strong>of</strong> Monte Carlo simulations for medical physics. Physics in<br />

Medicine and Biology, 51(13):R287–R301, 2006.<br />

[134] K. M. Rosenberg. CTSim – Open Source Computed Tomography Simulator. http://ctsim.org.<br />

[135] A. Saltelli, K. Chan, and E. Scott. Sensitivity analysis. Wiley series in probability and statistics.<br />

Wiley, 2000.<br />

[136] J. Serra. Image analysis and mathematical morphology. Number 1. Academic Press, 1982.<br />

[137] J. Sethian and P. Smereka. Level set methods for fluid interfaces. Annual Review <strong>of</strong> Fluid<br />

Mechanics, 35:341–372, 2003.<br />

[138] J. A. Sethian. Level Set Methods and Fast Marching Methods. Cambridge <strong>University</strong> Press,<br />

1999.<br />

[139] L. Shepp and B. Logan. The Fourier reconstruction <strong>of</strong> a head section. IEEE Transactions on<br />

Nuclear Science, 21(3):21–43, 1974.<br />

[140] S. Smolyak. Quadrature and interpolation formulas for tensor products <strong>of</strong> certain classes <strong>of</strong><br />

functions. Soviet Mathematics - Doklady, 4:240–243, 1963.<br />

[141] G. Stefanou, A. Nouy, and A. Clement. Identification <strong>of</strong> random shapes from images through<br />

polynomial chaos expansion <strong>of</strong> random level-set functions. International Journal for Numerical<br />

Methods in Engineering, 79(2):127–155, 2009.<br />

[142] J. Stubbe. Measure and integration 2009. Lecture notes, Institut de mathématiques d’analyse<br />

et applications, École polytechnique fédérale de Lausanne.<br />

[143] Y. Sun and C. Beckermann. Sharp interface tracking <strong>using</strong> the phase-field equation. Journal<br />

<strong>of</strong> Computational Physics, 220(2):626–653, 2007.<br />

[144] J. S. Suri, S. Laxminarayan, J. Gao, and L. Reden. Image segmentation via PDEs. In<br />

E. Micheli-Tzanakou, J. S. Suri, and S. Laxminarayan, editors, PDE and Level Sets: Algorithmic<br />

Approaches to Static and Motion Imagery, pages 153–223. Springer, 2002.<br />

[145] P. Therasse, S. G. Arbuck, E. A. Eisenhauer, J. Wanders, R. S. Kaplan, L. Rubinstein, J. Verweij,<br />

M. Van Glabbeke, A. T. van Oosterom, M. C. Christian, and S. G. Gwyther. New guidelines<br />

to evaluate the response to treatment in solid tumors. Journal <strong>of</strong> the National Cancer<br />

Institute, 92(3):205–216, 2000.<br />

[146] H. Tiesler, R. M. Kirby, D. Xiu, and T. Preusser. <strong>Stochastic</strong> collocation for optimal control<br />

problems with stochastic PDE constraints. Submitted to SIAM Journal on Optimization, 2010.<br />

128


Bibliography<br />

[147] J. Tryoen, O. Le Maître, M. Ndjinga, and A. Ern. Intrusive Galerkin methods with upwinding<br />

for uncertain nonlinear hyperbolic systems. Journal <strong>of</strong> Computational Physics, 229(18):6485–<br />

6511, 2010.<br />

[148] R. Tsai and S. Osher. Level set methods and their applications in image science. Communications<br />

in Mathematical Sciences, 1(4):623–656, 2003.<br />

[149] G. Turk and M. Levoy. Zippered polygon meshes from range images. In Proceedings <strong>of</strong> the<br />

21st annual conference on Computer graphics and interactive techniques, SIGGRAPH ’94,<br />

pages 311–318, 1994.<br />

[150] G. Vage. Variational methods for PDEs aplied to stochastic partial differential equations.<br />

Mathematica Scandinavica, 82:113–137, 1998.<br />

[151] G. G. Walter. Wavelets and Other Orthogonal Systems with Applications. CRC Press, 1994.<br />

[152] X. Wan and G. E. Karniadakis. An adaptive multi-element generalized polynomial chaos<br />

method for stochastic differential equations. Journal <strong>of</strong> Computational Physics, 209(2):617–<br />

642, 2005.<br />

[153] R. Weisenseel, W. Karl, D. Castanon, and R. Brewer. MRF-based algorithms for segmentation<br />

<strong>of</strong> SAR images. In Proceedings <strong>of</strong> the International Conference on Image Processing 1998<br />

(ICIP 98), volume 3, pages 770–774, 1998.<br />

[154] M. Wernick and J. Aarsvold. Emission Tomography: The Fundamentals <strong>of</strong> PET and SPECT.<br />

Elsevier Academic Press, 2004.<br />

[155] G. C. Wick. The evaluation <strong>of</strong> the collision matrix. Physical Review, 80(2):268–272, 1950.<br />

[156] N. Wiener. The homogeneous chaos. American Journal <strong>of</strong> Mathematics, 60(4):897–936,<br />

1938.<br />

[157] J. Witteveen and H. Bijl. Modeling arbitrary uncertainties <strong>using</strong> Gram-Schmidt polynomial<br />

chaos. In 44th AIAA Aerospace Sciences Meeting and Exhibit, 2006.<br />

[158] D. Xiu. Efficient collocational approach for parametric uncertainty analysis. Communications<br />

in Computational Physics, 2(2):293–309, 2007.<br />

[159] D. Xiu. Fast numerical methods for stochastic computations: A review. Communications in<br />

Computational Physics, 5(2–4):242–272, 2009.<br />

[160] D. Xiu and G. E. Karniadakis. The Wiener–Askey polynomial chaos for stochastic differential<br />

equations. SIAM Journal on Scientific Computing, 24(2):619–644, 2002.<br />

[161] D. Xiu and G. E. Karniadakis. Modeling uncertainty in flow simulations via generalized<br />

polynomial chaos. Journal <strong>of</strong> Computational Physics, 187(1):137–167, 2003.<br />

129

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!