16

Diffeomorphic density registration

Martin BaueraSarang JoshibKlas Modinc    aFlorida State University, Department of Mathematics, Tallahassee, FL, United States
bUniversity of Utah, Department of Bioengineering, Scientific Computing and Imaging Institute, Salt Lake City, UT, United States
cChalmers University of Technology and the University of Gothenburg, Department of Mathematical Sciences, Göteborg, Sweden

Abstract

In this book chapter we study the Riemannian geometry of the density registration problem: Given two densities (not necessarily probability densities) defined on a smooth finite-dimensional manifold find a diffeomorphism which transforms one to the other. This problem is motivated by the medical imaging application of tracking organ motion due to respiration in thoracic CT imaging, where the fundamental physical property of conservation of mass naturally leads to modeling CT attenuation as a density.

We will study the intimate link between the Riemannian metrics on the space of diffeomorphisms and those on the space of densities. We finally develop novel computationally efficient algorithms and demonstrate their applicability for registering thoracic respiratory correlated CT imaging.

Keywords

density registration; information geometry; Fisher–Rao metric; optimal transport; image registration; diffeomorphism groups; random sampling

Acknowledgements

We thank Caleb Rottmann, who worked on the implementation of the weighted diffeomorphic density matching algorithm. We are grateful for valuable discussions with Boris Khesin, Peter Michor, and François-Xavier Vialard.

This work was partially supported by the grant NIH R01 CA169102, the Swedish Foundation for Strategic Research (ICA12-0052), an EU Horizon 2020 Marie Sklodowska-Curie Individual Fellowship (661482), and by the Erwin Schrödinger Institute program: Infinite-Dimensional Riemannian Geometry with Applications to Image Matching and Shape Analysis by the FWF-project P24625.

16.1 Introduction

Over the last decade image registration has received intense interest, both with respect to medical imaging applications and to the mathematical foundations of the general problem of estimating a transformation that brings two or more given medical images into a common coordinate system [18,22,41,27,39,2,1,13]. In this chapter we focus on a subclass of registration problems referred to as density registration. The primary difference between density registration and general image registration is in how the registration transformation acts on the image being transformed. In density registration the transformation not only deforms the underlying coordinate system, but also scales the image intensity by the local change in volume. In numerous medical imaging applications this is of critical importance and is a fundamental property of the registration problem. The primary motivating clinical application is that of estimating the complex changes in anatomy due to breathing as imaged via 4D respiratory correlated computed tomography (4DRCCT). Given the physical quantitative nature of CT imaging, the natural action of a transformation on a CT image is that of density action: Any local compression induces a corresponding change in local density, resulting in changes in the local attenuation coefficient. We will also see that this difference in action of the transformation on the image being registered has wide ranging implications to the structure of the estimation problem. In this chapter we will study the fundamental geometrical structure of the problem and exemplify its application. The basic outline is as follows: We will first study the abstract mathematical structure of the problem, precisely defining the space of densities and the space of transformation. We will also study the set of transformations that leave the density unchanged. We will see that the explicit characterization of this set of transformations plays a critical role in understanding the geometric structure of the density registration problem. We will then introduce the general (regularized) density matching problem and present efficient numerical algorithms for several specific choices of regularizers. Finally, we will present the before mentioned application to model breathing as imaged via 4D respiratory correlated computed tomography.

16.2 Diffeomorphisms and densities

Let M denote a smooth oriented Riemannian manifold of dimension n with (reference) volume form dxImage.

Definition 16.1

The space of smooth densities1 on M is given by

Dens(M)={ρC(M)|ρ(x)>0xM}.

Image

The mass of a subset ΩMImage with respect to ρDens(M)Image is given by

Massρ(Ω)=Ωρdx.

Image

As the focus of the chapter is the registration of densities via transformation, the group Diff(M)Image of smooth diffeomorphisms of the manifold plays a central role.

Definition 16.2

The set of diffeomorphisms on M, denoted Diff(M)Image, consists of smooth bijective mappings MMImage with smooth inverses. This set has a natural group structure by composition of maps. The Lie algebra of Diff(M)Image is given by the space X(M)Image of smooth vector fields (tangential if M has a boundary).2

The group of diffeomorphisms acts naturally on the space of densities via pullback and pushforward of densities. Indeed, pullback of densities is given by

Diff(M)×Dens(M)(φ,ρ)φρ=|Dφ|ρ(φ()).

Image

Notice that this is a right action, that is, (φη)ρ=η(φρ)Image. The corresponding left action is given by pushforward of densities

Diff(M)×Dens(M)(φ,ρ)φρ=(φ1)ρ=|Dφ1|ρ(φ1()).

Image

The action of Diff(M)Image on densities captures the notion of conservation of mass and is fundamentally different from the standard action of Diff(M)Image on functions given by composition (see chapter 4). Indeed, for the density action, we have, for any subset ΩMImage,

Massφρ(φ(Ω))=Massρ(Ω),

Image

which follows from the change-of-coordinates formula for integrals.

The isotropy subgroup of an element ρDens(M)Image is by definition the subgroup of Diff(M)Image that leaves the density ρ unchanged. It is given by

Diffρ(M)={φDiff(M)|φρ=ρ}.

Image

The particular case ρ1Image gives the subgroup of volume preserving diffeomorphisms denoted by SDiff(M)Image. In general, φDiffρ(M)Image implies that φ is mass preserving with respect to ρ. In particular, if ΩMImage, then

Massρ(Ω)=Massρ(φ(Ω)).

Image

The point of diffeomorphic density registration is to select a template density ρ0Dens(M)Image and then generate new densities by acting on ρ0Image by diffeomorphisms. In our framework we shall mostly use the left action (by pushforward), but analogous results are also valid for the right action (by pullback). One may ask “Which densities can be reached by acting on ρ0Image by diffeomorphisms?” In other words, find the range of the mapping

Diff(M)φφρ0.

Image

In the language of group theory it is called the Diff(M)Image-orbit of ρ0Image. This question was answered in 1965 by Moser [31] for compact manifolds: the result is that the Diff(M)Image-orbit of ρ0Image consists of all densities with the same total mass as ρ0Image. This result has been extended to noncompact manifolds [17] and manifolds with boundary [4]. For simplicity, we will only formulate the result in the compact case.

Lemma 16.1

Moser [31]

Given ρ0,ρ1Dens(M)Image, where M is a compact manifold without boundary, there exists φDiff(M)Image such that φρ0=ρ1Image if and only if

Massρ1(M)=Massρ0(M).

Image

The diffeomorphism φ is unique up to right composition with elements in Diffρ0(M)Image or, equivalently, up to left composition with elements in Diffρ1(M)Image.

Since the total mass of a density is a positive real number, it follows from Moser's result that the set of Diff(M)Image-orbits in Dens(M)Image can be identified with R+Image. From a geometric point of view, this gives a fibration of Dens(M)Image as a fiber bundle over R+Image where each fiber corresponds to a Diff(M)Image-orbit. In turn, Moser's result also tells us that each orbit in itself is the base of a principal bundle fibration of Diff(M)Image. For example, the ρ0Image-orbit can be identified with the quotient Diff(M)/Diffρ0(M)Image through the projection

π:φφρ0.

Image

See references [29,5] for more details.

Remark 16.1

A consequence of the simple orbit structure of Dens(M)Image is that we can immediately check if the registration problem can be solved exactly by comparing the total mass of ρ0Image and ρ1Image. Furthermore, there is a natural projection from Dens(M)Image to any orbit simply by scaling by the total mass.

In diffeomorphic image registration, where the action on an image is given by composition with a diffeomorphism, the Diff(M)Image- orbits are much more complicated. Indeed, two generic images almost never belong to the same orbit. The problem of projecting from one orbit to another is ill-posed. On the other hand, because of the principal bundle structure of the space of densities, the exact registration problem of two densities with equal mass is well posed and has a complete geometric interpretation, which we will exploit to develop efficient numerical algorithms.

16.2.1 α-actions

The above mathematical development of diffeomorphisms acting on densities can be further generalized. By parameterizing the action by a positive constant α and define the α-action as follows: The group of diffeomorphisms Diff(Ω)Image acts from the left on densities by the α-action via

(φ,ρ)φαρ|Dφ1|αρφ1,

Image(16.1)

where |Dφ|Image denotes the Jacobian determinant of φ.

Remark 16.2

One theoretical motivation to study α-density action is that it enables the approximation of the standard action of Diff(M)Image on functions given by composition: formally, limα0φαI=Iφ1Image.

From a practical point of view, the motivation for the α-action stems from the fact that CT images do not transform exactly as densities. We will see in section 16.6.1 that for the application of density matching of thoracic CT images, the lungs behave as α-densities for α<1Image.

Analogous to the standard mass, we define the MassρpImage of a subset ΩMImage with respect to ρDensα(M)Image by

Massρp(Ω)=Ωρpdx.

Image

With this definition we immediately obtain the analogue of Lemma 16.1 for the α-action and thus also a similar principal fiber bundle picture.

Lemma 16.2

Given ρ0,ρ1Dens(M)Image, where M is a compact manifold, there exists φDiff(M)Image such that φαρ0=ρ1Image if and only if

Massρ1/α0(M)=Massρ11/α(M).

Image

16.3 Diffeomorphic density registration

In this part we will describe a general (Riemannian) approach to diffeomorphic density registration, that is, the problem of finding an optimal diffeomorphism φ that transports an α-density ρ0Image (source) to an α-density ρ1Image (target). By Moser's result (see Lemma 16.1 and Lemma 16.2) there always exists an infinite-dimensional set of solutions (diffeomorphisms) to this problem. Thus the main difficulty lies in the solution selection. Toward this aim, we introduce the regularized exact α-density registration problem:

Given a source density ρ0Image and a target density ρ1Image of the same total Massρ1/α0Image, find a diffeomorphism φ that minimizes

R(φ)under the constraintφαρ0=ρ1.

Image(16.2)

Here R(φ)Image is a regularization term.

Remark 16.3

Note that we have formulated the registration constraint using the left action of the diffeomorphism group, that is, φαρ=0ρ1Image. A different approach is to use the right action of Diff(M)Image with the constraint φαρ1=ρ0Image. These two approaches are conceptually different, as we aim to move the source to target using the left action while one moves the target to source using the right action. The resulting optimal deformations are however equal if the regularization term satisfies R(φ)=R(φ1)Image.

In later sections we will introduce several choices for RImage and discuss their theoretical and practical properties. In general, we aim to construct regularization terms such that the corresponding registration problem has the following desirable properties:

  1. 1.  Theoretical results on the existence and uniqueness of solutions;
  2. 2.  Fast and stable numerical computations of the minimizers;
  3. 3.  Meaningful optimal deformations.

Note that the notion of meaningful will depend highly on the specific application.

In practice we are sometimes not interested in enforcing the constraint, but are rather interested in a relaxed version of the above problem. Thus we introduce the inexact density registration problem:

Given a source density ρ0Image and a target density ρ1Image, find a diffeomorphism φ that minimizes

E(φ)=λd(φαρ0,ρ1)+R(φ).

Image(16.3)

Here λ>0Image is a scaling parameter, d(,)Image is a distance on the space of densities (the similarity measure), and R(φ)Image is a regularization term as before.

Remark 16.4

Note that we do not require the densities to have the same Mass1αImage in the inexact density matching framework. For densities that have the same Mass1αImage, we can retrieve the exact registration problem by considering the inexact registration problem as λImage.

On the space of probability densities there exists a canonical Riemannian metric, the Fisher–Rao metric, which allows for explicit formulas of the corresponding geodesic distance: it is given by the (spherical) Hellinger distance. For the purpose of this book chapter, we will often use this distance functional as a similarity measure.

16.4 Density registration in the LDDMM-framework

The LDDMM-framework is based on the idea of using a right-invariant metric on the diffeomorphism group to define the regularity measure, that is,

R(φ)=dist(id,φ),

Image(16.4)

where dist(,)Image denotes the geodesic distance of a right-invariant metric on Diff(M)Image. The resulting framework for inexact registration has been discussed for general shape spaces in chapter 4. Therefore we will keep the presentation in this chapter rather brief. Our focus will be on the geometric picture of the exact registration problem in this setup.

Remark 16.5

In the presentation of chapter 4 right-invariant metrics on Diff(M)Image have been defined using the theory of reproducing kernel Hilbert spaces (RKHS). We will follow a slightly different approach and equip the whole group of diffeomorphisms with a weak right-invariant metric; see [12] for a comparison of these two approaches.

From here on we assume that M is equipped with a smooth Riemannian metric g with volume density μ. To define a right-invariant metric on Diff(M)Image, we introduce the so-called inertia operator A:X(M)X(M)Image, where X(M)Image, the set of smooth vector fields, is the Lie algebra of Diff(M)Image. We will assume that A is a strictly positive elliptic differential operator, that is, self-adjoint with respect to the L2Image inner product on X(M)Image. For simplicity, we will only consider operators A that are defined via powers of the Laplacian of the Riemannian metric g, that is, we will only consider operators of the form

A=(1Δg)k

Image(16.5)

for some integer k. Here ΔgImage denotes the Hodge–Laplacian of the metric g. Most of the results discussed further are valid for a much larger class of (pseudo)differential operators; see [9]. Any such A defines the inner product GidImage on X(M)Image via

Gid(X,Y)=Mg(AX,Y)μ,

Image(16.6)

where μ denotes the induced volume density of g. We can extend this to a right-invariant metric on Diff(M)Image by right-translation:

Gφ(h,k)=Gid(hφ1,kφ1)=Mg(A(hφ1),kφ1)μ.

Image(16.7)

For an overwiew on right-invariant metrics on diffeomorphism groups, we refer to [28,12,7,8].

In this framework the exact density registration problem reads as follows.

Given a source density ρ0Image and a target density ρ1Image, find a diffeomorphism φ that minimizes

dist(id,φ)such that φαρ0=ρ1,

Image(16.8)

where dist(id,φ)Image is the geodesic distance on Diff(M)Image of the metric (16.7).

Using this particular regularization term provides an intuitive interpretation of the solution selection: we aim to find the transformation that is as close as possible to the identity under the constraint that it transports the source density to the target density.

In the following theorem we present a summary of the geometric picture that underlies the exact registration problem. To keep the presentation simple, we will only consider the case α=1Image, that is, the standard density action. A similar result can be obtained for general α.

Let π be the projection

π:Diff(M)Dens(M)Diffρ(M)Diff(M)

Image(16.9)

induced by the left action of the diffeomorphism group; see Lemma 16.1. By [9] we have the following:

Theorem 16.1

Let G be a right-invariant metric on Diff(M)Image of the form (16.7) with inertia operator A as in (16.5). Then there exists a unique Sobolev-type metric ˉGρImage on Dens(M)Image such that the projection π is a Riemannian submersion. The order of the induced metric ˉGImage on Dens(M)Image is k1Image, where k is the order of the metric G.

A direct consequence of the Riemannian submersion picture is the following characterization of the solutions of the exact density registration problem.

Corollary 16.1

Let ρ(t)Image, t[0,1]Image, be a minimizing geodesic connecting the given densities ρ0Image (source) and ρ1Image (target). Then the solution of the exact registration problem is given by the endpoint φ(1)Image of the horizontal lift of the geodesic ρ(t)Image.

Remark 16.6

This result describes an intriguing geometric interpretation of the solutions of the exact registration problem. Its applicability is however limited to the cases where there exist explicit solutions for the geodesic boundary value problem on the space of probability densities with respect to the metric ˉGImage. To our knowledge, the only such example is the so-called optimal information transport setting, which we will discuss in the next section. In the general case the solution of the exact density registration problem requires solving the horizontal geodesic boundary value problem on the group of diffeomorphisms, which is connected to the solution of a nonlinear PDE, the EPDiff equation. Various algorithms have been proposed for numerically solving the optimization problems [10,42,40]. We refer to Chapter 4 for more details.

16.5 Optimal information transport

In this section we describe an explicit way of solving the exact density registration problem. The framework in this section has been previously developed for random sampling from nonuniform arbitrary distributions [6]. For simplicity, we will restrict ourselves to the standard density action, that is, α=1Image. However, all the algorithms are easily generalized to general α. The specific setting uses deep geometric connections between the Fisher–Rao metric on the space of probability densities and a special right-invariant metric on the group of diffeomorphisms.

Definition 16.3

The Fisher–Rao metric is the Riemannian metric on Dens(M)Image given by

Gρ(˙ρ,˙ρ)=14M˙ρ2ρdx.

Image(16.10)

The main advantage of the Fisher–Rao metric is the existence of explicit formulas for the solution to the geodesic boundary value problem and thus also for the induced geodesic distance:

Proposition 16.1

Friedrich [14]

Given ρ0,ρ1Dens(M)Image with the same total mass, the Riemannian distance with respect to the Fisher–Rao metric is given by

dF(ρ0,ρ1)=arccos(Mρ1ρ0ρ0).

Image(16.11)

Furthermore, the geodesic between ρ0Image and ρ1Image is given by

ρ(t)=(sin(1t)θsinθ+sintθsinθρ1ρ0)2ρ0,

Image(16.12)

where θ=dF(ρ0,ρ1)Image.

Using formula (16.12) for geodesics, we will construct an almost explicit algorithm for solving an exact density registration problem of the form (16.2). To this end, we need to introduce a suitable regularization term. As in the LDDMM framework (see section 16.4), we will choose it as distance to the identity with respect to a right-invariant Riemannian metric on Diff(M)Image. However, to exploit the explicit formula (16.12), the right-invariant metric needs to communicate with the Fisher–Rao metric, as we now explain.

Definition 16.4

The information metric is the right-invariant Riemannian metric on Diff(M)Image given (at the identity) by

ˉGid(u,v)=MΔu,vdx+ki=1Mu,ξidxMv,ξidx,

Image(16.13)

where Δu denotes the Laplace–de Rham operator lifted to vector fields, and ξ1,,ξkImage are a basis of the harmonic fields on M. The Riemannian distance corresponding to ˉGImage is denoted dI(,)Image. Because of the Hodge decomposition theorem, the metric is independent of the choice of orthonormal basis for the harmonic fields.

Building on work by Khesin, Lenells, Misiolek, and Preston [24], Modin [29] showed that the metric ˉGImage descends to the Fisher–Rao metric on the space of densities. This fundamental property will serve as the basis for our algorithms.

We are now ready to formulate our special density registration problem, called the optimal information transport problem:

Optimal information transport (OIT) Given ρ0,ρ1Dens(M)Image and a Riemannian metric on M with volume form ρ0Image, find a diffeomorphism φ that minimizes

E(φ)=dI(id,φ)=dF(ρ0,φρ0)

Image(16.14)

under the constraint φρ0=ρ1Image.

In general, the formula for dI(id,φ)Image is not available explicitly; we would have to solve a nonlinear PDE (the EPDiff equation). However, because of the special relation between dIImage and dFImage, we have the following result, which is the key to an efficient algorithm.

Theorem 16.2

[29,5]

The OIT problem has a unique solution, that is, there is a unique diffeomorphism φDiff(M)Image minimizing dI(id,φ)Image under the constraint φρ0=ρ1Image. The solution is explicitly given by φ(1)Image, where φ(t)Image is the solution to the problem

Δf(t)=˙ρ(t)ρ(t)φ(t),v(t)=(f(t)),ddtφ(t)1=v(t)φ(t)1,φ(0)=id,

Image(16.15)

and ρ(t)Image is the Fisher–Rao geodesic connecting ρ0Image and ρ1Image:

ρ(t)=(sin(1t)θsinθ+sintθsinθρ1ρ0)2ρ0,cosθ=Mρ1ρ0ρ0.

Image(16.16)

Based on Theorem 16.2, we now give a semiexplicit algorithm for numerical computation of the solution to the optimal information transport problem. The algorithm assumes that we have a numerical way to represent functions, vector fields, and diffeomorphisms on M and numerical methods for

  • •  composing functions and vector fields with diffeomorphisms,
  • •  computing the ∇ of functions, and
  • •  computing solutions to Poisson's equation on M.

Numerical algorithm for optimal information transport

  1. 1.  Choose a step size ε=1/KImage for some positive integer K and calculate the Fisher–Rao geodesic ρ(t)Image and its derivative ˙ρ(t)Image at all time points tk=kKImage using equation (16.16).
  2. 2.  Initialize φ0=idImage. Set k0Image.
  3. 3.  Compute sk=˙ρ(tk)ρ(tk)φkImage and solve the Poisson equation

Δfk=sk.

Image(16.17)

  1. 4.  Compute the gradient vector field vk=fkImage.
  2. 5.  Construct approximations ψkImage to exp(εvk)Image, for example,

ψk=idεvk.

Image(16.18)

  1. 6.  Update the diffeomorphism

φk+1=φkψk.

Image(16.19)

  1. If needed, we may also compute the inverse by φ1k+1=φ1k+εvφ1kImage.
  2. 7.  Set kk+1Image and continue from step 3 unless k=KImage.

Although it is possible to use optimal information transport and the algorithm above for medical image registration problems, the results so obtained are typically not satisfactory; the diffeomorphism obtained tends to compress and expand matter instead of moving it (see, e.g., [5, Sec. 4.2]). Another problem is that the source and target densities are required to be strictly positive, which is typically not the case for medical images. In section 16.6 we will develop a gradient flow-based approach, which will lead to much better results for these applications. However, in applications where either the source or the target density is uniform (with respect to the natural Riemannian structure of the manifold at hand), the OIT approach can be very competitive, which yields a natural application for random sampling.

16.5.1 Application: random sampling from nonuniform distribution

In this section we describe an application of OIT to random sampling from nonuniform distributions, that is, the following problem.

Random sampling problem

Let ρ1Dens(M)Image. Generate N random samples from the probability distribution ρ1Image.

The classic approach to sample from a probability distribution on a higher-dimensional space is to use Markov chain Monte Carlo (MCMC) methods, for example, the Metropolis–Hastings algorithm [20]. An alternative idea is to use diffeomorphic density registration between the density ρ1Image and the standard density ρ0Image from which samples can be drawn easily. Indeed, we can then draw samples from ρ0Image and transform them via the computed diffeomorphism to generate samples from ρ1Image. A benefit of transport-based methods over traditional MCMC methods is cheap computation of additional samples; it amounts to drawing uniform samples and then evaluating the transformation. On the other hand, unlike MCMC, transport-based methods scale poorly with increasing dimensionality of M.

Moselhy and Marzouk [30] and Reich [33] proposed to use optimal mass transport (OMT) to construct the desired diffeomorphism φ, thereby enforcing φ=cImage for some convex function c. The OMT approach implies solving, in one form or another, the heavily nonlinear Monge–Ampere equation for c. A survey of the OMT approach to random sampling is given by Marzouk et al. [26]. Using OIT instead of OMT, the problem simplifies significantly, as the OIT-algorithm only involves solving linear Poisson problems.

As a specific example, consider M=T2(R/2πZ)2Image with distribution defined in Cartesian coordinates x,y[π,π)Image by

ρ3exp(x210(yx2/2+1)2)+1/10,

Image(16.20)

normalized so that the ratio between the maximum and mimimum of ρ is 100. The resulting density is depicted in Fig. 16.1 (left).

Image
Figure 16.1 Application of OIT to random sampling(Left) The probability density ρ of (16.20). The maximal density ratio is 100. (Right) 105 samples from ρ calculated using our OIT-based random sampling algorithm.

We draw 105 samples from this distribution using a MATLAB implementation of our algorithm, available under MIT license at https://github.com/kmodin/oit-random

The implementation can be summarized as follows. To solve the Poisson problem, we discretize the torus by a 256×256Image mesh and use the fast Fourier transform (FFT) to invert the Laplacian. We use 100 time steps. The resulting diffeomorphism is shown as a mesh warp in Fig. 16.2. We then draw 105 uniform samples on [π,π]2Image and apply the diffeomorphism on each sample (applying the diffeomorphism corresponds to interpolation on the warped mesh). The resulting random samples are depicted in Fig. 16.1 (right). Drawing new samples is very efficient. For example, another 107 samples can be drawn in less than a second.

Image
Figure 16.2 Application of OIT to random samplingThe computed diffeomorphism φK shown as a warp of the uniform 256 × 256 mesh (every 4th mesh-line is shown). Notice that the warp is periodic. The ratio between the largest and smallest warped volumes is 100.

16.6 A gradient flow approach

In the optimal information transport described in the previous section the fundamental restriction is that the volume form of Riemannian metric of the base manifold is compatible with the density being transformed in that it has to be conformally related to the source density ρ0Image. In most medical imaging applications this modeling assumption is not applicable. In this section we will develop more general algorithms that relax the requirement for the metric to be compatible with the densities to be registered. We will consider the natural extension of the Fisher–Rao metric to the space of all densities and the case where dx(Ω)=Image, for which it is given by

d2F(I0dx,I1dx)=Ω(I0I1)2dx.

Image(16.21)

Notice that d2F(,)Image in this case is the Hellinger distance. For details, see [5].

The Fisher–Rao metric is the unique Riemannian metric on the space of probability densities that is invariant under the action of the diffeomorphism group [7,3]. This invariance property extends to the induced distance function, so

d2F(I0dx,I1dx)=d2F(φ(I0dx),φ(I1dx))φDiff(Ω).

Image(16.22)

Motivated by the aforementioned properties, we develop a weighted diffeomorphic registration algorithm for registration of two density images. The algorithm is based on the Sobolev H1Image gradient flow on the space of diffeomorphisms that minimizes the energy functional

E(φ)=d2F(φ(fdx),(fφ1)dx)+d2F(φ(I0dx),I1dx).

Image(16.23)

This energy functional is only a slight modification of the energy functional studied in [5]. Indeed, if f in the equation is a constant σ>0Image, then (16.23) reduces to the energy functional of Bauer, Joshi, and Modin [5, § 5.1]. Moreover, the geometry described in [5, § 5.3] is valid also for the functional (16.23), and, consequently, the algorithm developed in [5, § 5.2] can be used also for minimizing (16.23). There the authors view the energy functional as a constrained minimization problem on the product space Dens(Ω)×Dens(Ω)Image equipped with the product distance; see Fig. 16.3 and [5, § 5] for details on the resulting geometric picture. Related work on diffeomorphic density registration using the Fisher Rao metric can be found in [37,36].

Image
Figure 16.3 Geometry of the density registration problemIllustration of the geometry associated with the density registration problem. The gradient flow on Diff(Ω) descends to a gradient flow on the orbit Orb(fdx,I0dx). When constrained to Orb(fdx,I0dx)⊂Dens(Ω)×Dens(Ω), this flow strives to minimize the product Fisher–Rao distance to ((fφ) dx,I1dx).

Using the invariance property of the Fisher–Rao metric and assuming infinite volume, the main optimization problem associated with the energy functional (16.23) is the following.

Given densities I0dxImage, I1dxImage, and fdxImage, find φDiff(Ω)Image minimizing

E(φ)=Ω(|Dφ1|1)2fφ1dxE1(φ)+Ω(|Dφ1|I0φ1I1)2dxE2(φ).

Image(16.24)

The invariance of the Fisher–Rao distance can be seen with a simple change of variables xφ(y)Image, dx|Dφ|dyImage, and |Dφ1|1|Dφ|Image. Then, Equation (16.24) becomes

E(φ)=Ω(1|Dφ|)2fdy+Ω(I0|Dφ|I1φ)2dy.

Image(16.25)

To better understand the energy functional E(φ)Image, we consider the two terms separately. The first term E1(φ)Image is a regularity measure for the transformation. It penalizes the deviation of the diffeomorphism φ from being volume preserving. The density fdxImage acts as a weighting on the domain Ω; that is, change of volume (compression and expansion of the transformation φ) is penalized more in regions of Ω where f is large. The second term E2(φ)Image penalizes dissimilarity between I0dxImage and φ(I1dx)Image. It is the Fisher–Rao distance between the initial density I0dxImage and the transformed target density φ(I1dx)Image. Because of the invariance (16.22) of the Fisher–Rao metric, this is the same as the Fisher–Rao distance between I1dxImage and φ(I0dx)Image.

Solutions to problem (16.24) are not unique. To see this, let DiffI(Ω)Image denote the space of all diffeomorphisms preserving the volume form IdxImage:

DiffI(Ω)={φDiff(Ω)||Dφ|(Iφ)=I}.

Image(16.26)

If φ is a minimizer of E()Image, then ψφImage for any

ψDiff1,I0(Ω)Diff1(Ω)DiffI0(Ω)

Image(16.27)

is also a minimizer. Notice that this space is not trivial. For example, any diffeomorphism generated by a Nambu–Poisson vector field (see [32]), with I0Image as one of its Hamiltonians, will belong to it. A strategy to handle the degeneracy was developed in [5, § 5]: the fact that the metric is descending with respect to the H1Image metric on Diff(Ω)Image can be used to ensure that the gradient flow is infinitesimally optimal, that is, always orthogonal to the null-space. We employ the same strategy in this paper. The corresponding geometric picture can be seen in Fig. 16.3.

To derive an gradient algorithm to optimize the energy functional the natural metric on the space of diffeomorphisms to use is the H1Image-metric due to its intimate link with the Fisher–Rao metric as described previously. The H1Image-metric on the space of diffeomorphisms is defined using the Hodge Laplacian on vector fields and is given by

GIφ(U,V)=ΩΔu,vdx.

Image(16.28)

Due to its connections to information geometry, we also refer to this metric as the information metric. Let GIEImage denote the gradient with respect to the information metric defined previously. Our approach for minimizing the functional of (16.25) is to use a simple Euler integration of the time discretization of the gradient flow:

˙φ=GIE(φ).

Image(16.29)

The resulting final algorithm is one order of magnitude faster than LDDMM, since we are not required to time integrate the geodesic equations, as is necessary in LDDMM [42].

In the following theorem we calculate the gradient of the energy functional.

Theorem 16.3

The GIImage-gradient of the registration functional (16.25) is given by

GIE=Δ1(grad(fφ1(1|Dφ1|))|Dφ1|I0φ1grad(I1)+grad(|Dφ1|I0φ1)I1).

Image(16.30)

Proof

We first calculate the variation of the energy functional. Therefore let φsImage be a family of diffeomorphisms parameterized by the real variable s such that

φ0=φanddds|s=0φs=vφ.

Image(16.31)

We use the following identity derived in [21]:

dds|s=0|Dφs|=12|Dφ|div(v)φ.

Image(16.32)

The variation of the first term of the energy functional is

dds|s=0E1(φ)=Ωf(y)(|Dφ(y)|1)|Dφ(y)|div(v)φ(y)dy.

Image(16.33)

We do a change of variables yφ1(x)Image, dy|Dφ1(x)|dxImage, using the fact that |Dφ(y)|=1|Dφ1(x)|Image:

=Ωfφ1(x)(1|Dφ1(x)|)div(v(x))dx

Image(16.34)

=fφ1(1|Dφ1|),div(v)L2(R3)

Image(16.35)

=grad(fφ1(1|Dφ1|)),vL2(R3)

Image(16.36)

using the fact that the adjoint of the divergence is the negative gradient. For the second term of the energy functional, we expand the square

E2(φ)=ΩI0(y)2I0(y)I1φ(y)|Dφ(y)|+I1φ(y)|Dφ(y)|dy.

Image(16.37)

Now ΩI1φ(y)|Dφ(y)|dyImage is constant (conservation of mass), so we only need to minimize over the middle term. Then the derivative is

dds|s=0E2(φ)=Ω2I0(y)(gradI1Tv)φ(y)|Dφ(y)|I0(y)I1φ(y)|Dφ(y)|div(v)φ(y)dy.

Image(16.38)

We do the same change of variables as before:

=ΩI0φ1(x)|Dφ1(x)||Dφ1(x)|(2gradI1(x)Tv(x)+I1(x)div(v)(x))

Image(16.39)

=2|Dφ1|I0φ1gradI1,vL2(R3)|Dφ1|I0φ1I1,div(v)L2(R3)

Image(16.40)

=|Dφ1|I0φ1gradI1,vL2(R3)+grad(|Dφ1|I0φ1)I1,vL2(R3).

Image(16.41)

From these equations we conclude that

Δ(GIE)=grad(fφ1(1|Dφ1|))|Dφ1|I0φ1gradI1+grad(|Dφ1|I0φ1)I1

Image(16.42)

Since we are taking the Sobolev gradient of E, we apply the inverse Laplacian to the right-hand side of Equation (16.42) to solve for GIEImage. □

Remark 16.7

Notice that in the formula for GIEImage we never need to compute φ, so in practice we only compute φ1Image. We update this directly via φ1(x)φ1(x+ϵGIE)Image for some step size ϵ.

16.6.1 Thoracic density registration

We now present an application of the developed theory to the problem of estimating complex anatomical deformations associated with the breathing cycle as imaged via computed tomography (CT) [16]. This problem has wide-ranging medical applications, in particular, radiation therapy of the lung, where accurate estimation of organ deformations during treatment impacts dose calculation and treatment decisions [35,23,38,15]. The current state-of-the-art radiation treatment planning involves the acquisition of a series of respiratory correlated CT (RCCT) images to build 4D (three spatial and one temporal) treatment planning data sets. Fundamental to the processing and clinical use of these 4D data sets is the accurate estimation of registration maps that characterize the motion of organs at risk and the target tumor volumes.

The 3D image produced from X-ray CT is an image of linear attenuation coefficients. For narrow beam X-ray, the linear attenuation coefficient (LAC) for a single material (units cm1Image) is defined as μ(x)=mρ(x)Image, where m is a material-specific property called the mass attenuation coefficient (units cm2/gImage) that depends on the energy of the X-ray beam. The linear attenuation coefficient is proportional to the true density and therefore exhibits conservation of mass. Unfortunately, CT image intensities do not represent true narrow beam linear attenuation coefficients. Instead, modern CT scanners use wide beams that yield secondary photon effects at the detector. CT image intensities reflect effective linear attenuation coefficients as opposed to the true narrow beam linear attenuation coefficient.

To see the relationship between effective LAC and true narrow beam LAC, we ran a Monte Carlo simulation using an X-ray spectrum and geometry from a Philips CT scanner at various densities of water (since lung tissue is very similar to a mixture between water and air) [11]. The nonlinear relationship between effective LAC and narrow beam LAC relationship is clear (see Fig. 16.4).

Image
Figure 16.4 Massρ1/αImage conservation in lung density matchingEffective LAC from Monte Carlo simulation (solid line) and NIST reference narrow beam LAC (dashed line). The true relationship between effective LAC and narrow beam LAC is nonlinear.

If we have conservation of mass within a single subject in a closed system, then we expect an inverse relationship between average density in a region Ω and volume of that region: Dt=MVtImage. Here Vt=Ωt1dxImage, Dt=ΩtIt(x)dx/VtImage, ΩtImage is the domain of the closed system (that moves over time), and t is a phase of the breathing cycle. This relationship becomes linear in log space with a slope of −1:

ln(Dt)=ln(M)ln(Vt)

Image(16.43)

Our experimental results confirm the Monte Carlo simulation in that lungs imaged under CT do not follow this inverse relationship. Rather, the slope found in these datasets in log space is consistently greater than −1 (see Fig. 16.6). This implies that for real clinical CT data sets, the lung tissue is acted on by an α-density action. Using the isomorphism between α-densities and 1-densities, we estimate the power transformation I(x)I(x)αImage and estimate the α that yields the best conservation of mass property.

For each subject, we perform a linear regression of the measured LAC density in the homogeneous lung region and the calculated volume in log space. Let d(α)=log(ΩtIt(x)αdx/Ωt1dx)Image (the log density) and v=log(Ωt1dx)Image (the log volume), where again t is a breathing cycle timepoint. The linear regression then models the relationship in log space as d(α)av+bImage. Let aj(α)Image be the slope solved for in this linear regression for the jth subject. To find the optimal α for the entire dataset, we solve

α=arg minαj(aj(α)+1)2,

Image(16.44)

which finds the value of α that gives us an average slope closest to −1.

Applying this power function to the CT data allows us to perform our density registration algorithm based on the theory developed. We therefore seek to minimize the energy functional described in the previous section, given by

E(φ)=d2F(φ(fdx),(fφ1)dx)+d2F(φ(I0dx),I1dx))

Image(16.45)

=Ω(|Dφ1|1)2fφ1dxE1(φ)+Ω(|Dφ1|I0φ1I1)2dxE2(φ).

Image(16.46)

We construct the density f(x)dxImage, a positive weighting on the domain Ω, to model the physiology of the thorax: regions where f(x)Image is high have a higher penalty on nonvolume-preserving deformations and regions where f(x)Image is low have a lower penalty on nonvolume-preserving deformations. Physiologically, we know that the lungs are quite compressible as air enters and leaves. Surrounding tissue including bones and soft tissue, on the other hand, is essentially incompressible. Therefore our penalty function f(x)Image is low inside the lungs and outside the body and high elsewhere. For our penalty function, we simply implement a sigmoid function of the original CT image: f(x)=sig(I0(x))Image.

Recall the Sobolev gradient calculated in Theorem 16.30 with respect to the energy functional given by

δE=Δ1((fφ1(1|Dφ1|))|Dφ1|I0φ1(I1)+(|Dφ1|I0φ1)I1).

Image(16.47)

Then the current estimate of φ1Image is updated directly via an Euler integration of the gradient flow [34]:

φ1j+1(x)=φ1j(x+ϵδE)

Image(16.48)

for some step size ϵ. Since we take the Sobolev gradient, the resulting deformation is guaranteed to be invertible with a sufficiently small ϵ. Also notice that the gradient only depends on φ1Image, so there is no need to keep track of both φ and φ1Image. The exact numerical algorithm is as follows:

Numerical algorithm for weighted diffeomorphic density registration

Image

The algorithm was implemented using the PyCA package and can be downloaded at https://bitbucket.org/crottman/pycaapps/src/master/ See the application Weighted Diffeomorphic Density Registration.

For the DIR dataset, we solved for the exponent that yields conservation of mass, which yielded α=0.60Image giving us the best fit. Without using the exponential fit, the average slope of log density log volume plot was −0.66 (SD 0.048). After applying the exponential to the CT intensities, the average slope is −1.0 (SD 0.054). The log–log plots of all ten patients in the DIR dataset and the box plots of the slope are shown in Fig. 16.5.

Image
Figure 16.5 Application to lung density matchingDensity and volume log–log plots. Upper left: log–log plots without applying the exponential correction for all ten DIR subjects. The best fit line to each dataset is in red (light gray in print version), and the mass-preserving line (slope = −1) is in black. Upper right: log–log plots after applying the exponential correction I(x)α to the CT images. In this plot the best fit line matches very closely to the mass-preserving line. Bottom row: the corresponding box plots of the slopes found in the regression.

For the 30 subject dataset, we solved for α=0.52Image, which gives us conservation of mass. Without using the exponential fit, the average slope of the log–log plot was −0.59 (SD 0.11).

We applied our proposed weighted density registration algorithm to the first subject from the DIR dataset. This subject has images at 10 timepoints and has a set of 300 corresponding landmarks between the full inhale image and the full exhale image. These landmarks were manually chosen by three independent observers. Without any deformation, the landmark error is 4.01 mm (SD 2.91 mm). Using our method, the landmark error is reduced to 0.88 mm (SD 0.94 mm), which is only slightly higher than the observer repeat registration error of 0.85 mm (SD 1.24 mm).

We implement our algorithm on the GPU and plot the energy and the Fisher–Rao metric with and without applying the deformation. These results are shown in Fig. 16.6. In this figure we show that we have excellent data match, whereas the deformation remains physiologically realistic: inside the lungs there is substantial volume change due to respiration, but the deformation outside the lungs is volume preserving. With a 256×256×94Image voxel dataset, our algorithm takes approximately nine minutes running for four thousand iterations on a single nVidia Titan Z GPU.

Image
Figure 16.6 Application of to lung density matchingRegistration results. Top row: full inhale, full exhale, and the deformed exhale density estimated using our method. Middle row: Jacobian determinant of the transformation, initial Fisher–Rao metric, and Fisher–Rao metric after applying the density action. Notice that outside the lungs the estimated deformation is volume preserving. Bottom row: Energy as a function of iterations, and penalty function.

References

1. John Ashburner, A fast diffeomorphic image registration algorithm, NeuroImage 2007;38(1):95–113.

2. Brian B. Avants, Nicholas J. Tustison, Gang Song, Philip A. Cook, Arno Klein, James C. Gee, A reproducible evaluation of ants similarity metric performance in brain image registration, NeuroImage 2011;54(3):2033–2044.

3. Nihat Ay, Jürgen Jost, Hong Le Van, Lorenz Schwachhöfer, Information geometry and sufficient statistics, The Annals of Statistics 2014.

4. Augustin Banyaga, Formes-volume sur les variétés à bord, L'Enseignement Mathématique 1974;20(2):127–131.

5. M. Bauer, S. Joshi, K. Modin, Diffeomorphic density matching by optimal information transport, SIAM Journal on Imaging Sciences 2015;8(3):1718–1751.

6. M. Bauer, S. Joshi, K. Modin, Diffeomorphic random sampling using optimal information transport, F. Nielsen, F. Barbaresco, eds. Geometric Science of Information, GSI 2017. Lecture Notes in Computer Science. Cham: Springer; 2017;vol. 10589.

7. Martin Bauer, Martins Bruveris, Peter W. Michor, Uniqueness of the Fisher–Rao metric on the space of smooth densities, Bulletin of the London Mathematical Society 2016;48(3):499–506.

8. Martin Bauer, Joachim Escher, Boris Kolev, Local and global well-posedness of the fractional order EPDiff equation on RdImage, Journal of Differential Equations 2015;258(6):2010–2053.

9. Martin Bauer, Sarang Joshi, Klas Modin, On geodesic completeness for Riemannian metrics on smooth probability densities, Calculus of Variations and Partial Differential Equations 2017;56(4), 113 18.

10. Mirza Faisal Beg, Michael I. Miller, Alain Trouvé, Laurent Younes, Computing large deformation metric mappings via geodesic flows of diffeomorphisms, International Journal of Computer Vision 2005;61(2):139–157.

11. John M. Boone, J. Anthony Seibert, An accurate method for computer-generating tungsten anode x-ray spectra from 30 to 140 kV, Medical Physics 1997;24(11):1661–1670.

12. Bruveris Martins, François-Xavier Vialard, On completeness of groups of diffeomorphisms, Journal of the European Mathematical Society 2017;19:1507–1544 10.4171/JEMS/698.

13. Brad C. Davis, P. Thomas Fletcher, Elizabeth Bullitt, Sarang Joshi, Population shape regression from random design data, International Journal of Computer Vision 2010;90(2):255–266.

14. Thomas Friedrich, Die Fisher-Information und symplektische Strukturen, Mathematische Nachrichten 1991;153(1):273–296.

15. Sarah E. Geneser, J.D. Hinkle, Robert M. Kirby, Brian Wang, Bill Salter, S. Joshi, Quantifying variability in radiation dose due to respiratory-induced tumor motion, Medical Image Analysis 2011;15(4):640–649.

16. Vladlena Gorbunova, Jon Sporring, Pechin Lo, Martine Loeve, Harm A Tiddens, Mads Nielsen, Asger Dirksen, Marleen de Bruijne, Mass preserving image registration for lung ct, Medical Image Analysis 2012;16(4):786–795.

17. Robert E. Greene, Katsuhiro Shiohama, Diffeomorphisms and volume-preserving embeddings of noncompact manifolds, Transactions of the American Mathematical Society 1979;255:403–414.

18. Ulf Grenander, Michael I. Miller, Computational anatomy: an emerging discipline, Quarterly of Applied Mathematics 1998;56(4):617–694.

19. Richard S. Hamilton, The inverse function theorem of Nash and Moser, Bulletin of the American Mathematical Society 1982;7(1):65–222.

20. W. Keith Hastings Monte, Carlo sampling methods using Markov chains and their applications, Biometrika 1970;57(1):97–109.

21. Jacob Hinkle, Sarang Joshi, Idiff: irrotational diffeomorphisms for computational anatomy, Information Processing in Medical Imaging. Springer; 2013:754–765.

22. Sarang Joshi, Brad Davis, Matthieu Jomier, Guido Gerig, Unbiased diffeomorphic atlas construction for computational anatomy, NeuroImage 2004;23:S151–S160.

23. Paul J. Keall, Sarang Joshi, S. Sastry Vedam, Jeffrey V. Siebers, Vijaykumar R. Kini, Radhe Mohan, Four-dimensional radiotherapy planning for DMLC-based respiratory motion tracking, Medical Physics 2005;32(4):942–951.

24. B. Khesin, J. Lenells, G. Misiołek, S.C. Preston, Geometry of diffeomorphism groups, complete integrability and geometric statistics, Geometric and Functional Analysis 2013;23(1):334–366.

25. Serge Lang, Fundamentals of Differential Geometry. Graduate Texts in Mathematics. New York: Springer; 1999;vol. 191.

26. Youssef Marzouk, Tarek Moselhy, Matthew Parno, Alessio Spantini, Sampling via measure transport: an introduction, Roger Ghanem, David Higdon, Houman Owhadi, eds. Handbook of Uncertainty Quantification. Cham: Springer International Publishing; 2016.

27. Michael I. Miller, Alain Trouvé, Laurent Younes, On the metrics and Euler–Lagrange equations of computational anatomy, Annual Review of Biomedical Engineering 2002;4(1):375–405.

28. G. Misiołek, S.C. Preston, Fredholm properties of Riemannian exponential maps on diffeomorphism groups, Inventiones Mathematicae 2010;179(1):191–227.

29. Klas Modin, Generalized Hunter–Saxton equations, optimal information transport, and factorization of diffeomorphisms, The Journal of Geometric Analysis 2015;25(2):1306–1334.

30. Tarek A. El Moselhy, Youssef M. Marzouk, Bayesian inference with optimal maps, Journal of Computational Physics 2012;231(23):7815–7850.

31. Jürgen Moser, On the volume elements on a manifold, Transactions of the American Mathematical Society 1965;120:286–294.

32. N. Nakanishi, A survey of Nambu–Poisson geometry, Lobachevskii Journal of Mathematics 1999;4:5–11 (electronic).

33. Sebastian Reich, A nonparametric ensemble transform method for Bayesian inference, SIAM Journal on Scientific Computing 2013;35(4):A2013–A2024.

34. Caleb Rottman, Ben Larson, Pouya Sabouri, Amit Sawant, Sarang Joshi, Diffeomorphic density registration in thoracic computed tomography, Sebastien Ourselin, Leo Joskowicz, Mert R. Sabuncu, Gozde Unal, William Wells, eds. Medical Image Computing and Computer-Assisted Intervention – MICCAI 2016. Cham: Springer International Publishing; 2016:46–53.

35. Amit Sawant, Paul Keall, Kim Butts Pauly, Marcus Alley, Shreyas Vasanawala, Billy W. Loo Jr, Jacob Hinkle, Sarang Joshi, Investigating the feasibility of rapid MRI for image-guided motion management in lung cancer radiotherapy, BioMed Research International 2014:2014.

36. D. Seo, J. Ho, B.C. Vemuri, Computing diffeomorphic paths for large motion interpolation, Computer Vision and Pattern Recognition (CVPR), 2013 IEEE Conference on. June 2013:1227–1232.

37. D.H. Seo, J. Ho, J.H. Traverse, J. Forder, B. Vemuri, Computing diffeomorphic paths with application to cardiac motion analysis, 4th MICCAI Workshop on Mathematical Foundations of Computational Anatomy. 2013:83–94.

38. Yelin Suh, Walter Murray, Paul J. Keall, IMRT treatment planning on 4D geometries for the era of dynamic MLC tracking, Technology in Cancer Research and Treatment 2014;13(6):505–515.

39. Tom Vercauteren, Xavier Pennec, Aymeric Perchant, Nicholas Ayache, Diffeomorphic demons: efficient non-parametric image registration, NeuroImage 2009;45(1):S61–S72.

40. François-Xavier Vialard, Laurent Risser, Daniel Rueckert, Colin J. Cotter, Diffeomorphic 3D image registration via geodesic shooting using an efficient adjoint calculation, International Journal of Computer Vision Apr. 2012;97(2):229–241.

41. Laurent Younes, Shapes and Diffeomorphisms, vol. 171. Springer Science & Business Media; 2010.

42. Laurent Younes, Felipe Arrate, Michael I. Miller, Evolutions equations in computational anatomy, NeuroImage 2009;45(1, Supplement 1):S40–S50 Mathematics in Brain Imagingo.


1  “If M is compact, then Dens(M)Image is an infinite-dimensional Fréchet manifold[19], that is, a manifold modeled on a Fréchet space. For the purpose of analysis, it is often useful to instead work with the Sobolev completion Denss(M)Image for a Sobolev index s>n/2Image. Denss(M)Image is then a Banach manifold[25]. The benefit of Banach over Fréchet manifolds is that most standard results from finite dimensions, such as the inverse function theorem, are valid.”

2  “If M is compact, then Diff(M)Image is a Fréchet Lie group[19, § I.4.6], i.e., a Fréchet manifold where the group operations are smooth mappings.”

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.221.165.246