13

A discretize–optimize approach for LDDMM registration

Thomas PolzinaMarc Niethammerb,cFrançois-Xavier VialarddJan Modersitzkia,e    aInstitute of Mathematics and Image Computing, University of Lübeck, Lübeck, Germany
bDepartment of Computer Science, University of North Carolina at Chapel Hill, Chapel Hill, NC, United States
cBiomedical Research Imaging Center (BRIC), Chapel Hill, NC, United States
dLaboratoire d'informatique Gaspard Monge, Université Paris-Est Marne-la-Vallée, UMR CNRS 8049, Champs sur Marne, France
eFraunhofer MEVIS, Lübeck, Germany

Abstract

Large deformation diffeomorphic metric mapping (LDDMM) is a popular approach for deformable image registration with nice mathematical properties. LDDMM encodes spatial deformations through time-varying velocity fields. Hence registration requires optimization over these time-varying velocity fields, resulting in a large-scale constrained optimization problem.

Typical numerical solution approaches for LDDMM use an optimize–discretize strategy, where optimality conditions are derived in the continuum and subsequently discretized and solved. Here we explore solution methods based on the discretize–optimize approach and discuss ramifications for popular LDDMM relaxation and shooting approaches. The focus is on a consistent method that uses the appropriate Runge–Kutta methods for the solution of all arising PDEs in the Eulerian frame. Additionally, we discuss both run-time and memory consumption requirements and present an approach that makes the registration suitable for standard PCs. We demonstrate the practicality of our proposed approach in the context of image registration applied to 3D computed tomography (CT) scans of the lung.

Keywords

LDDMM; Discretize–Optimize; Image Registration; Optimal Control; Lung; Computed Tomography; Runge–Kutta Methods

13.1 Introduction

The goal of image registration is to establish spatial correspondences between images. Image registration is a challenging but important task in image analysis. In particular, in medical image analysis image registration is a key tool to compare patient data in a common space, to allow comparisons between pre-, inter-, or post-intervention images, or to fuse data acquired from different and complementary imaging devices such as positron emission tomography (PET), computed tomography (CT), or magnetic resonance imaging (MRI) [56].

Image registration is of course not limited to medical imaging, but is important for a wide range of applications; for example, it is used in astronomy, biology/genetics, cartography, computer vision, and surveillance. Consequently, there exists a vast number of approaches and techniques; see, for example, [10,29,31,33,44,56,65,66,84,87,91,97,109] and references therein.

Optical flow approaches [48,55] are popular and commonly used in computer vision [12,100,72,106,11]. However, computer vision applications typically have different requirements regarding spatial transformations than medical image analysis applications. For example, when processing natural images, objects, such as cars or people, should typically be allowed to move independently of the background. However, in many medical applications it is natural to constrain the sought-for mapping to be bijective or even diffeomorphic. In early days of image registration, transformation models were (and sometimes still are) constricted to rigid mappings, which obviously fulfill this constraint. However, more flexible, for example, deformable, transformation models are required to capture subtle localized changes. Early work on deformable image registration includes variational approaches based on elasticity theory [30,9,6], where a spatial transformation is represented nonparametrically (via a deformation field) and is regularized through an elastic potential function.

The mathematical reason for introducing the regularizer is to ensure solutions of the variational formulation. It may also be interpreted as a penalty for nonelasticity. Elastic registration has been widely and successfully used. However, an elastic regularizer can in general not guarantee that a computed mapping is bijective [23]. To ensure bijectivity, one may add constraints to a registration formulation. Alternatively, one can formulate a registration model that by construction guarantees the regularity of the transformation. Examples for the former approach are [79,36,37]. While Rohlfing et al. [79] employed an additional penalty term on the determinant of the Jacobian of the transformation that improves transformation regularity but cannot guarantee bijective mappings, Haber and Modersitzki developed similar ideas, but with equality [36] or box constraints [37] thereby guaranteeing bijectivity. Nevertheless, Haber and Modersitzki still use a mathematical model based on linear elasticity theory, which is only appropriate for small deformations. As convex energies, such as the linear elastic potential, result in finite penalties, they are insufficient to guarantee one-to-one maps [13,24]. Therefore Burger et al. [13] used hyperelasticity and quasiconvex functions to obtain a registration model yielding bijective transformations while allowing for large deformations.

Another possibility to ensure diffeomorphic mappings (smooth mappings that are bijective and have a smooth inverse) is to express a transformation implicitly through velocity fields. The intuitive idea is that it is easy to ensure diffeomorphic transformations for small-scale displacements by using a sufficiently strong spatial regularization. Hence a complex diffeomorphic transform can be obtained, capturing large displacements, as the composition of a large or potentially infinite number of small-scale diffeomorphisms. These small-scale diffeomorphisms in turn can be obtained by integrating a static or time-dependent velocity field in time. Early work explored such ideas in the context of fluid-based image registration [23]. In fact, for approaches using velocity fields, the same regularizers as for the small displacement registration approaches (based on elasticity theory or the direct regularization of displacement fields) can be used. However, the regularizer is now applied to one or multiple velocity fields instead of a displacement field, and the transformation is obtained via time-integration of the velocity field.

The solution of the arising problems is computationally expensive, as in the most general case we now need to estimate a spatio-temporal velocity field instead of a static displacement field. Greedy solution approaches as well as solutions based on a stationary velocity field have been proposed to alleviate the computational burden (both in computational cost, but also with respect to memory requirements) [23,1,2,57]. However, these methods in general do not provide all the nice mathematical properties (metric, geodesics, etc.) of the nongreedy large deformation diffeomorphic metric mapping (LDDMM) registration approach [61], which we will explore numerically here.

Various numerical approaches have been proposed for LDDMM, for example, [23,61,96,7]. Traditionally, relaxation formulations [7] have been used to solve the LDDMM optimization problem. Here, one directly optimizes over spatio-temporal velocity fields, ensuring that the velocity field at any given time is sufficiently smooth. The resulting transformation, obtained via time-integration of the spatio-temporal velocity field, is used to deform the source image such that it becomes similar to the target image. Specifically, the LDDMM relaxation formulation is solved via an optimize–discretize approach [7], where a solution to the continuous time optimality conditions for the associated constrained optimization problem (the constraint being the relation between the spatial transformation and the velocity fields) is computed. These optimality conditions are the Euler–Lagrange equations corresponding to the constrained optimization problem and can be regarded as the continuous equivalent of the Karush–Kuhn–Tucker (KKT) [71, p. 321] equations of constrained optimization. To numerically determine a solution fulfilling the Euler–Lagrange equations, one uses an adjoint solution approach, which allows the efficient computation of the gradient of the LDDMM relaxation energy with respect to the spatio-temporal velocity field via a forward/backward sweep. This gradient can then be used within a gradient descent scheme or as the basis of sophisticated numerical solution approaches. Note that the adjoint solution approach is, for example, also at the core of the famous backpropagation algorithm for the training of neural networks [54] or the reverse mode of automatic differentiation [34, pp. 37].1 For an LDDMM relaxation solution, a geodesic path (fulfilling the Euler–Lagrange equations exactly) is only obtained at convergence [7].

More recently, shooting approaches like [3,98,70] have been proposed. Again, an optimize–discretize approach is used. Here, instead of numerically solving the Euler–Lagrange equations of the LDDMM energy (which can alternatively be represented via the Euler–Poincaré diffeomorphism equation [62] (EPDiff)), these Euler–Lagrange equations are imposed as a dynamical system, and the LDDMM energy is reformulated as an initial value problem. Hence all paths obtained during the optimization process are by construction geodesics—they may just not be the optimal ones. The price to pay for such a reduced parameterization is that one now optimizes over a second-order partial differential equation that describes the geodesic path. A simple analogy in the one-dimensional Euclidean space is that in relaxation, registration between two points is performed over all possible paths leading at convergence to a straight-line path. On the contrary, for shooting, we already know that the optimal solution should be a straight line, and consequentially optimization is only performed regarding the line slope and y-intercept.

Although the mathematical framework for LDDMM is very appealing, its numerical treatment is not trivial, may lead to nondiffeomorphic transformations, and is highly demanding both in terms of memory and computational costs. As discussed before, current LDDMM solutions are mainly based on optimize–discretize formulations, that is, optimality conditions are derived in continuous space and then discretized [7,62]. Not surprisingly, it is possible that a solution of the discrete equations is an optimizer neither for discrete nor for continuous energy [35].

Hence the goal of this chapter is to develop a discretize–optimize approach for LDDMM, where the starting point is the discretization of the energy functional. Consequentially, solutions to the associated optimality conditions are indeed optimizers of the discretized energy. Additionally, we integrate a suitable interpolation operator [51] into the LDDMM framework to reduce computational demands and memory consumption without losing registration accuracy. Furthermore, we use the so-called normalized gradient fields (NGF) distance measure, which is designed to align image edges [38,66]. NGF has been successfully applied to lung CT registration [81,51,7476,82], which is one of our example applications in Section 13.8.

The chapter is organized as follows. In Section 13.2 the LDDMM concept is introduced, and two approaches [7,45] used for solving the constrained optimization problems are discussed. In Section 13.3 we extend these models, discuss the shooting approach, and formulate them in the context of general distance measures for images. For simplicity, Section 13.4 discusses the discretization of the models in the one-dimensional case. This allows introducing the basic ingredients (grids, regularization, derivatives, transport equation, etc.) required for the 3D formulation in a compact manner. In Section 13.5 we address the discretization and solution of the partial differential equations constraining the LDDMM models (e.g., the transport equations) via Runge–Kutta methods. The extension of the described formalism to 3D is given in Section 13.6. In Section 13.7 details on the numerical optimization and the solution of the optimization problems in a multilevel approach are provided. Afterward, in Section 13.8 we evaluate the performance of the proposed methods and present experimental results. Finally, we conclude the chapter with a discussion of the results and possible extensions in Section 13.9.

13.2 Background and related work

LDDMM is used to automatically establish correspondences between two (or more) given datasets, which are referred to as source and target (fixed and moving or reference and template are also common names). The data typically consist of images, landmarks (point clouds), or surfaces. In the following we restrict ourselves to the registration of three-dimensional images. However, when considering landmarks or surfaces, the main change of the presented approaches would be the data term.

Specifically, LDDMM registration estimates a time- and space-dependent velocity field v:Ω×[0,1]R3Image, which flows the source image I0Image so that it matches as well as possible a target image I1Image. Conceptually, the data dimension and the time-horizon are arbitrary,2 but for ease of presentation, we assume that the data are three-dimensional, the images I0Image and I1Image are compactly supported on a domain ΩR3Image, and the time interval is [0,1]Image.

Assuming that all structures are visible in the source and target images, after an ideal registration I0(ϕ11(x))=I1(x)Image for all xΩImage for a transformation ϕ:Ω×[0,1]R3Image. This is, of course, not feasible in practice, either because there is no perfect structural correspondence or due to the presence of noise that precludes exact matching. Here we use the notation ϕt(x)ϕ(x,t)Image. This notation will be used for all variables depending on space and time. In particular, in this notation the subscript is not related to a partial derivative. The sought-for transformation and the velocity fields are related via ˙ϕt(x)=vt(ϕt(x))Image, ϕ0(x)=xImage, where vt(x)v(x,t)Image and ˙ϕ=tϕImage. In the remainder of the chapter we will often, for a cleaner and more compact notation, omit the spatial argument and assume that equations hold for arbitrary xΩImage. Furthermore, if t is not specified, then we assume that equations hold for all t[0,1]Image.

Following the work of Beg et al. [7], a solution of the registration problem is given by a minimizer of the energy

EBeg(v,ϕ)Reg(v)+SimSSD(I0ϕ11,I1)s.t. ˙ϕt=vtϕt,ϕ0=Id.}

Image(13.1)

The regularizer Reg enforces the smoothness of the velocity fields, and the distance measure (data fit) SimSSDImage is used to compute the similarity of the transformed source image I0ϕ11Image at t=1Image and the target image I1Image. The distance measure for images I and J is SimSSD(I,J)=12σ2Ω(I(x)J(x))2dxImage and is called the sum of squared differences (SSD). SSD is only one of many possibilities; see, for example, [66, p. 95]. We discuss further details and choices of distance measures in Section 13.3.4. The regularity of v is encouraged by employing a differential operator L in Reg:

Reg(v)1210vt2Ldtwithvt2LLvt,LvtΩvt(x)LLvt(x)dx,}

Image(13.2)

where LImage is the adjoint operator of L. We use the Helmholtz operator L(γIdαΔ)βImage, α, γ>0Image, and βNImage, which is a typical choice [7,45,107]. Here Δ denotes the spatial vectorial Laplacian, and vitImage, i=1,2,3Image, is the ith component of v at time t:

Δvt(x)(3i=1xi,xiv1t(x)3i=1xi,xiv2t(x)3i=1xi,xiv3t(x)).

Image(13.3)

Standard LDDMM models follow the optimize–discretize approach, that is, optimality conditions are computed for the continuous model through calculus of variations, and the resulting optimality conditions are then discretized. For example, in the derivation by Beg et al. [7] the variation of (13.1) with respect to v is computed resulting in optimality conditions of the form

ItI0ϕ1t,λt=|Jτ1t|λ1τt1,LLvt+λtIt=0,}

Image(13.4)

with initial and final conditions

I0=I0andλ1=1σ2(I0ϕ11I1),

Image(13.5)

where τtτ(t)Image is the flow for the negative velocity field (with τ1=IdImage), and ItImage is the transformed source image at time t. The adjoint variable λ:R3×[0,1]RImage is initialized at t=1Image with the negative image mismatch (also called residual; see Section 13.3.4 on distance measures) scaled by the weighting factor 1σ2>0Image, which balances regularizer and distance measure energy. The spatial gradient of the image at time t is referred to as ItImage, and the Jacobian of the spatial variables of τ1tImage is Jτ1tImage. Note that the adjoint variable λ is also called the scalar momentum in the literature [98], and the quantity λI=mImage is the vector-valued momentum [90]. In fact, the optimality conditions can be written solely with respect to the vector-valued momentum m, resulting in the EPDiff equation; see, for example, [67].

Computing the variation leading to (13.4) and (13.5) in the optimize–discretize framework is rather involved. If the interest is only in I1Image, alternatively the flow equation in the form of a transport equation in Eulerian coordinates can be added as a constraint to the optimization problem. This has been proposed by Borzí et al. [8] for optical flow and by Hart et al. [45] for LDDMM. Following [45], the LDDMM formulation then becomes

EHart(v,I)=Reg(v)+SimSSD(I1,I1),s.t.˙It+Itvt=0,I0=I0,}

Image(13.6)

with the optimality conditions

˙It+Itvt=0,I0=I0,˙λt+div(λtvt)=0,λ1=1σ2(I1I1),LLvt+λtIt=0.}

Image(13.7)

The optimality conditions (13.4)(13.5) and (13.7) are the same, but (13.7) are written directly in terms of the image I and the Lagrange multiplier λ. In practice both formulations are solved by computing the flows ϕ and τ from which the solution of the transport equation (for I) can be obtained by interpolation.3 The solution for the scalar conservation law/the continuity equation for λ is obtained by interpolation and local weighting by the determinant of the Jacobian of the transformation.

The advantage of the second optimize–discretize approach (13.6) is that the optimality conditions (13.7) are (relatively) easy to derive and retain physical interpretability. For example, it is immediately apparent that the adjoint to the transport equation is a scalar conservation law. However, great care needs to be taken to make sure that all equations and their boundary conditions are consistently discretized in the optimize–discretize approach. For example, it cannot be guaranteed that for any given discretization, the gradients computed based on the discretized optimality conditions will minimize the overall energy [35]. In particular, these formulations require interpolation steps, which can cause trouble with the optimization, because it is unclear how they should affect a discretized gradient. Furthermore, the flow equations can be discretized in various ways (e.g., by a semi-Lagrangian scheme [7], upwind schemes [45], etc.). But, as these discretizations are not considered part of the optimization, their effect on the gradient remains unclear.

Therefore we follow the discretize–optimize approach in image registration [66, p. 12], which starts out with a discretization of the energy to be minimized and derives the optimality conditions from this discretized representation. Our approach is related to the work by Wirth et al. [101,83] on geodesic calculus for shapes but approaches the problem from an optimal control viewpoint.

In [5] the diffeomorphic matching is phrased as an optimal control problem, which is solved using a discretize–optimize approach. This method is related to the approach we will propose, but instead of matching images, in [5] surfaces and point clouds are registered. Another difference is the use of explicit first-order Runge–Kutta methods (i.e., forward Euler discretization), whereas we are using fourth-order methods (see Section 13.5.1) to fulfill numerical stability considerations. Recently, Mang and Ruthotto [58] also formulated LDDMM registration as an optimal control problem using a discretize–optimize approach. Specifically, they use the Gauss–Newton approach, which improves the convergence rate compared to the L-BFGS method we will apply but requires the solution of additional linear systems. However, the biggest difference between our proposed approach and the approach by Mang et al. [58] is that we use explicit Eulerian discretizations of the partial differential equation constraints of LDDMM via Runge–Kutta methods, whereas Mang et al. [58] use Lagrangian methods. Although such Lagrangian methods are attractive as they eliminate the risk of numerical instabilities and allow for arbitrary time step sizes, they require frequent interpolations and consequentially become computationally demanding. Instead, our formulation operates on a fixed grid and only requires interpolations for the computation of the image distance measure. To avoid numerical instabilities due to too large time-steps, we investigate how an admissible step size can be chosen under easily obtainable estimations for the maximal displacement when using a fourth-order Runge–Kutta method. To illustrate the difference in computational requirements, we note that although in [58] experiments were performed on a compute node with 40 CPUs, the computation for a 3D registration (with a number of voxels per image that is only about 10% of the voxel numbers of the images we are registering) takes between 25 and 120 minutes and thus is two to five times slower than our proposed methods on a single CPU. This dramatic difference in computational effort is most likely due to the tracking of the points/particles in the Lagrangian setting or the required interpolations.

An optimal solution vImage fulfills the optimality conditions (e.g., (13.7)). However, these conditions need to be determined numerically. The classical approach, proposed in [7], is the so-called relaxation approach. Here, for a given v, the transport equation for I is solved forward in time and the adjoint equation for λ backward in time. Given I and λ, the gradient of the energy with respect to the velocity v at any point in space and time can be computed, i.e.,

vtE(v,I)=LLvt+λtIt,

Image(13.8)

which can then, for example, be used in a gradient descent solution scheme. In practice typically the Hilbert-gradient is used [7]

HvtE(v,I)=vt+(LL)1(λtIt)

Image(13.9)

to improve numerical convergence.4 Upon convergence, a relaxation approach will fulfill the optimality conditions, which describe a geodesic path [7]. Unfortunately, the optimality conditions will in practice frequently not be fulfilled exactly, and the relaxation approach requires optimization over the full spatio-temporal velocity field v, whereas the geodesic path is completely specified by its initial conditions, that is, its initial velocity v0Image and its initial image I0Image. This has motivated the development of shooting approaches [3,98], where ones optimizes over the initial conditions of the geodesic directly instead of v. The numerical solution is similar to the relaxation approach in the sense that the adjoint system is derived to compute the gradient with respect to the initial conditions through a forward–backward sweep. Shooting approaches also allow for extensions to the LDDMM registration models, such as LDDMM-based geodesic regression [70].

A common problem with LDDMM is its large memory footprint and the high computational costs. For example, to register lung CT images, run times of up to three hours on a computer with 128 GB of RAM and 32 CPUs have been reported [86]. In particular, for the shooting approaches, several ideas to overcome these problems have been proposed. Marsland and McLachlan [59] and Durrleman et al. [27] use a limited number of control points for LDDMM and observe that the number of control points can be decreased substantially (i.e., much fewer control points than number of pixels or voxels are needed) without greatly impacting the registration result. Zhang et al. [107] exploit the smoothness of the velocities in the Fourier domain. The underlying idea is that the initial velocities of the geodesic shooting are band-limited and therefore can be well approximated by a limited number of elements of a finite-dimensional vector space.

In contrast to the brain MRI or synthetic examples discussed in most LDDMM publications (see, e.g., [7,27,107]), lung registration (our motivating registration application) requires aligning many small spatially distributed salient structures (such as vessels). Hence the spatial discretization for the images cannot be too coarse as it would otherwise risk ignoring these fine structures. We can still reduce computational costs and memory requirements by recognizing that velocity fields will be smooth. Specifically, as we assume that the structures of interest are well dispersed over the whole lung volume, we employ regular grids for the velocity fields (or momenta fields). However, as confirmed in our own previous work [76], velocity fields can be discretized more coarsely (about one quarter of resolution per dimension) than the images themselves due to their inherent smoothness. We thereby obtain a method that aligns images well without losing accuracy compared to higher-resolution approaches while reducing computational costs and memory requirements.

13.3 Continuous mathematical models

In this section we introduce the continuous models that will be solved in a discretize–optimize approach. Fig. 13.1 shows different options. To obtain a consistent discretization for relaxation and shooting, we would ideally like to start from a discretized LDDMM energy, derive the KKT conditions, impose the KKT conditions as a constraint, and then obtain a discretization of the shooting equations, which is consistent with the relaxation equations (i.e., the KKT equations of the discretized LDDMM energy). However, this is difficult in practice as it would no longer allow the use of explicit time-integration schemes. Instead, we suggest separate discretize–optimize approaches for the relaxation and the shooting formulations of LDDMM. For the discretize–optimize formulation of relaxation, we simply discretize the LDDMM energy including the transport equation constraint. We discuss in Section 13.3.1 the related continuous relaxation energies for the direct transport of images, whereas in Section 13.3.2 the corresponding energies for the transport of maps are considered. Alternatively, for the discretize–optimize approach for the shooting formulation of LDDMM, it is sensible to start with the discretized energy of the continuous shooting model including the associated constraints given by the EPDiff equation, which are described in Section 13.3.3.

Image
Figure 13.1 Different discretization options for LDDMM. Red (light gray in print version) squares indicate domains of numerical evaluation (all discrete), whereas the blue (dark gray in print version) squares indicate the domain of the optimality conditions. The standard approaches use an optimize–discretize approach. In particular, Beg et al. [7] use the relaxation approach (A), whereas Vialard et al. [98] use the shooting approach (C). An ideal approach would have consistent discretization for relaxation and shooting (E). However, this is difficult due to related requirements on the numerical integration approach. We therefore explore the discretize–optimize variants of relaxation (B) (see Section 13.5.3 and Section 13.5.4) and shooting (D) (see Section 13.5.5) in this chapter. In all discretize–optimize approaches the energy is consistent with the gradient used for optimization, which may not be the case for the optimize–discretize approaches. Furthermore, in shooting approaches the energy gets reformulated via the initial conditions of the geodesic equation, which is enforced and yields the adjoint geodesic system to compute the gradient with respect to the initial conditions.

In contrast to the methods presented in Section 13.2, we will use a general distance measure in what follows. Specifically, distance measures quantify the similarity of images by assigning a low (scalar) value to image pairs that are similar and a large value to image pairs that are different. Formally, we define the set of gray-value images as I{I:ΩR}Image. A distance measure is a mapping Sim:I×IRImage [65, p. 55]. We assume that the distance measure is differentiable and that ASim(A,B)Image is the derivative with respect to the first image. In Section 13.3.4 we give details of the used distance measures. Next, we describe the models we will use for registration. These are extensions of the methods in Section 13.2 and are called IBR, MBR, and MBS, respectively, where I means image, M map, B based, R relaxation, and S shooting.

13.3.1 Relaxation with transport of images (IBR)

We extend the relaxation model of Hart et al. [45] (see (13.6)) via a general differentiable distance measure Sim. This model is a relaxation model as the optimization is performed with respect to the full spatio-temporal velocity field in t[0,1]Image. This is in contrast to shooting approaches where images (or transformations) are following a flow determined completely by initial conditions at t=0Image; see Section 13.3.3. We do not explicitly model a transformation ϕ to transform the source image, but directly compute the image mismatch between the modeled image at t=1Image (I(1)=I1Image) and the target image I1IImage:

argminvE1(v,I)s.t.˙It+Itvt=0,I0=I0,withE1(v,I)=Reg(v)+Sim(I1,I1).}

Image(13.10)

13.3.2 Relaxation with transport of maps (MBR)

Typically, the spatial transform ϕ is of interest and needs to be computed, for example, to calculate local volume changes during respiration [95]. This transform can be computed via time-integration of the spatio-temporal velocity field. Hence, in Eulerian coordinates we now advect a deformation map instead of an image. When solving the transport equation this is advantageous because the deformations are smooth by design and thus suffer less from dissipation effects than images.

As our goal is to work with fixed computational grids (i.e., in Eulerian coordinates), we can also write the evolution equation for the inverse map ϕ11Image in Eulerian coordinates. The inverse transformation ϕ11Image is required to deform the source image to the coordinate system of the target image (i.e., I0ϕ11Image; see (13.1)).

The evolution of ϕ11Image is given by [63].

˙ϕ1t=Jϕ1tvt,ϕ01=Id,

Image(13.11)

which is nothing else than a set of independent transport equations for the coordinate functions (x, y, and z in 3D). Here, we use the spatial Jacobian (JϕJϕ(x,t)R3×3Image) of ϕ=(ϕ1,ϕ2,ϕ3)Image, which is defined as

Jϕ(x1ϕ1t(x)x2ϕt1(x)x3ϕ1t(x)x1ϕ2t(x)x2ϕ2t(x)x3ϕ2t(x)x1ϕ3t(x)x2ϕ3t(x)x3ϕ3t(x)).

Image(13.12)

Note, that we used superscripts to refer to the components of vectors such as ϕ(x,t)R3Image. As the point of view for the time direction of the transformations is arbitrary, we change the notation of the inverse transformation used in [7] (see also (13.1)) for convenience to ϕ. This results in the simplified notation

˙ϕt+Jϕtvt=0,ϕ0=Id.

Image(13.13)

We use (13.13) as the constraint for the map-based optimization problem:

argminvE2(ϕ,v)s.t. ˙ϕt+Jϕtvt=0,ϕ0=Id,withE2(ϕ,v)Reg(v)+Sim(I0ϕ1,I1).}

Image(13.14)

13.3.3 Shooting with maps using EPDiff (MBS)

Recall that the optimality conditions of the relaxation formulation can be written with respect to the vector-valued momentum m only. Hence, instead of obtaining a solution to these optimality conditions at convergence (as done in the relaxation approach), we simply enforce these optimality conditions as constraints in an optimization problem. This is the shooting approach [62,98]. Specifically, if we express these conditions with respect to the momentum, then we obtain the EPDiff equation (abbreviation for the Euler–Poincaré equation on the group of diffeomorphisms) [47,61,105]

˙mt=Jmtvtmtdiv(vt)Jvtmt,vt=Kmt.

Image(13.15)

Here m:Ω×[0,1]R3Image is the momentum, and K=(LL)1Image is a smoothing kernel. This equation describes a geodesic path and fully determines the transformation as it implies an evolution of the velocities v via vt=KmtImage. The total momentum will be constant for a geodesic due to the conservation of momentum [104,98]:

Km0,m0=Kmt,mt,t[0,1].

Image(13.16)

If we integrate (13.15) and (13.16) into (13.14), then the minimization problem has the following form:

argminm0E3(m,ϕ,v)s.t.tϕ+Jϕv=0,ϕ0=Id,tm+Jmv+mdiv(v)+Jvm=0,v=Km,with E31210mt,Kmtdt+Sim(I0ϕ1,I1)(13.16)=12m0,Km0+Sim(I0ϕ1,I1).}

Image(13.17)

13.3.4 Distance measures

Registration quality depends on the appropriateness of the deformation model and the distance measure. Different distance measures can easily be integrated into LDDMM; see [22,4]. Note that we formulate our models in generality and that the following derivations hold for a large class of differentiable distance measures [66, p. 109]:

Sim(A,B)1σ2ψ(r(A,B)),

Image(13.18)

where r is the residuum or image mismatch of A and B, and ψ is typically an integral of the residual values. The functions r and ψ are assumed to be at least once differentiable. This is necessary to allow derivative-based optimization, which we are using in our discretize–optimize schemes. Data fidelity and regularity are balanced via the weight σ>0Image, which can be interpreted as the image noise level. We use two different differentiable distance measures in our experiments. The L2Image-based distance measure sum of squared differences (SSD) [65, p. 56] is used in most LDDMM publications (see, e.g., [7,45,107]):

SimSSD(A,B)=12σ2Ω(A(x))B(x))2dx.

Image(13.19)

Hence ψSSD(r)=12Ωr(x)2dxImage, and the residuum is the difference rSSD(A,B)=ABImage. As motivated in Section 13.1, we will use the normalized gradient field (NGF) distance measure for the registration of lung CT images. NGF aligns image edges [38,66] and is popular for lung CT registration [81,51,7476]:

SimNGF(A,B)1σ2Ω1A,B2εA2εB2εdx

Image(13.20)

with u,vεε2+3i=1uiviImage and u2εu,uεImage for u, vR3Image. Throughout this chapter we set ε=50Image for lung CT registrations. Rewriting SimNGFImage with ψ and r yields:

ψNGF(r)=Ω1r(x)2dx,rNGF(A,B)=A,BεAεBε.

Image

This concludes the description of the continuous LDDMM models. In the following sections we present their discretization.

13.4 Discretization of the energies

In this section we describe the numerical computation of the continuous energies introduced in Section 13.3. We start with the description of the different regular grids, which are used for discretization of the velocities, images, transformations, et cetera in Section 13.4.1. The discretization of the regularizer and the distance measures is addressed in Section 13.4.2 and Section 13.4.3 respectively.

13.4.1 Discretization on grids

We use regular grids to discretize v, I, ϕ, et cetera in time and space. To properly handle boundary conditions, it is important to discuss on which grid the velocities and images are given. In particular, we choose nodal grids for the velocities v and transformations ϕ and cell-centered grids for the images I. For simplicity, we start with the discretization for the one-dimensional space. We assume that the discrete source and target images have MNImage values located at equidistant centers of intervals whose union is the closure of the domain ˉΩ=[ω1,ω2]RImage. For the nodal grids nNImage, n2Image points are used. Details on both types of grids are given, for example, in [66, p. 125].

Fig. 13.2 shows two example grids. The nodal grid is depicted as blue (dark gray in print version) squares, whereas the cell-centered grid is visualized as red (light gray in print version) dots. Note that the number of cells for images and velocities does not need to be equal. In fact, it is the key idea of our approach to speed up computations and reduce memory requirements by discretizing velocity and transformation fields on a lower resolution grid than the images, that is, n<MImage. Consequently, we have different spatial step sizes hv1ω2ω1n1Image and hI1ω2ω1MImage. As we consider only one dimension first, we will omit the indices and write hvImage and hIImage. The time axis [0,1]Image is discretized for v, ϕ, and I on a nodal grid with N time steps, and accordingly the time step size is ht1N1Image. The resulting nodal grid is xndRnNImage, and the cell-centered image grid is a vector xccRMNImage. The discretized velocity field is then given as vv(xnd)Rn×NImage. Accordingly, the images are discretized as II(xcc)RM×NImage.

Image
Figure 13.2 Examples for 1D grids for discrete points in time. The image grid points are cell-centered and plotted as red (light gray in print version) circles, whereas the grid points for v and ϕ are nodal and plotted as blue (dark gray in print version) squares. The parameters for these grids are N = 5, M = 7, and n = 4.

When solving the constraint equations (e.g., the transport equation for images in (13.10)), we need images and velocities discretized at the same resolution. We therefore use a linear grid interpolation (bi/trilinear for 2D/3D data). The appropriate weights are stored in the components of the interpolation matrix PRM×nImage. See [51,80] on how to compute the weights and how to implement a matrix-free computation of the matrix–matrix product. By multiplying the velocity matrix v with the matrix P we obtain a cell-centered discretization PvRM×NImage, which can be used to transport the intensities located at the image grid points. When solving the adjoint equation of, for example, (13.7), the transposed matrix has to be used to interpolate the derivative of the distance measure to the nodal grid to update the velocities.

13.4.2 Discretization of the regularizer

Two main computations are required to discretize the regularization term. First, we compute the differential operator L, which is based on second derivatives, and its effect on the velocity field to be regularized. Second, we approximate the integral of the regularizer using numerical quadrature. In particular, we use the trapezoidal rule. Here we formulate everything initially in 1D for simplicity and assume Neumann boundary conditions for the velocities v and Dirichlet boundary conditions for the images throughout. The discretized regularizer has the form

Regh(ˉv)hvht2ˉvˉLWˉLˉv,

Image(13.21)

and we proceed with details on the computation of ˉLImage, ˉvImage, and W.

We use a standard approach to discretize the Helmholtz operator L=(γIdαΔ)βImage with Neumann boundary conditions; see, for example, [92]. The derivatives are approximated with central finite differences. The proper choice of the parameters α, γ>0Image and βNImage depends on the application and in particular on the spatial dimension kNImage. For instance, to obtain diffeomorphic solutions for k=3Image, at least β=2Image is needed [26]. To be more precise, using Sobolev embedding theorems [108, p. 62], the following inequality has to be fulfilled (see [60] and references therein):

s>k2+1,

Image(13.22)

where s is the order of the HsImage metric used for the velocity fields vtImage, t[0,1]Image. For vRn×NImage, the discrete operator is LRn×nImage:

L(γIdnαΔhv)β,

Image(13.23)

where IdnRn×nImage is the identity matrix, and

Δhv1(hv)2(1112112111)

Image(13.24)

is a discrete Laplacian operator. We use a trapezoidal quadrature for integration in space and time. The different weights for inner and outer grid points are assigned by multiplication with the diagonal matrix WRnN×nNImage. Kronecker products are used for a compact notation and to compute ˉLRnN×nNImage, which enables regularization for all time steps at once:

WWNWn,Wpdiag(12,1,,1,12)Rp×p,ˉLIdNL.}

Image(13.25)

If we use the column-vector representation of v,

ˉv(v1,1,,vn,1,v1,2,,vn,2,,v1,N,,vn,N),

Image

and (13.25), then we obtain (13.21) for the discretized regularizer.

13.4.3 Discretization of the distance measures

For simplicity, we only give details of the discretization of the SSD distance measure defined in (13.19). For two discrete one-dimensional images A=(ai)Mi=1Image and B=(bi)Mi=1Image (e.g., the discrete source (I0Image) and target (I1Image) images), we use a midpoint quadrature to compute the discrete distance measure value

Simh,SSD(A,B)=hI2σ2Mi=1(aibi)2=:ψ(r).

Image(13.26)

Hence the discrete residual for SSD is rr(A,B)=(aibi)Mi=1Image, and the discrete outer function is given as ψ(r)=hI2σ2Mi=1r2iImage. The numerical gradient, which will be needed for numerical optimization, consists of the following derivatives:

ψ(r)=hIσ2rRM,irj=δi,j,i,j=1,,M.

Image(13.27)

Hence the numerical Jacobian of r is just the identity matrix, Ar=IdMRM×MImage. Based on these results, (13.27), and the chain rule, we obtain the numerical gradient for the distance measure:

ASimh,SSD(A,B)=(Ar)ψ(r)=hIσ2r.

Image(13.28)

We transposed the inner and outer derivative in (13.28) as we assume column vector gradients throughout the chapter. For discretizations of SSD in multiple dimensions or other distance measures, such as NGF, we refer to [66, p. 107].

13.5 Discretization and solution of PDEs

The partial differential equations we need to solve are the transport/advection equation, the EPDiff equation, and the adjoint equations of the models (13.10), (13.14), and (13.17).

The standard LDDMM registrations use semi-Lagrangian methods [7] or upwinding for the solution of the transport equation [45]. In this work semi-Lagrangian methods are not used as they require interpolations at every time step and hence have substantially increased computational costs. Upwinding includes logical switches, which interfere with our model, which requires differentiable constraints. We therefore focus on methods that use central differences to compute spatial derivatives. However, this then requires appropriate schemes for time-integration as, for example, an Euler forward scheme, which is known to be unconditionally unstable in such a case [92] for the solution of the transport equation. An implicit scheme generates stable solutions, but solving many linear equation systems is too expensive for our image registration purposes. As a compromise, we use explicit fourth-order Runge–Kutta methods, which have a sufficiently large stability region (and in particular include part of the imaginary axis) and have acceptable computational requirements; see Section 13.5.1. In Section 13.5.2 we describe how to consistently solve the adjoint equations (e.g., scalar conservation/continuity equations). We then apply these approaches to the IBR (Section 13.5.3), MBR (Section 13.5.4), and MBS (Section 13.5.5) models.

First, consider the one-dimensional transport equation, which is part of the IBR model:

˙It+(xIt)vt=0,I0=I0.

Image(13.29)

The spatial derivative is approximated with central finite differences and homogeneous Dirichlet boundary conditions are assumed for the images. The matrix to compute the discrete image derivatives is

DI1=12hI(1110110111)Rm×m.

Image(13.30)

For t=ht[0,1]Image, =0,,N1Image, we approximate the spatial derivative for all image grid points xccImage as

xI(xcc,t)DI1I,

Image(13.31)

where IImage is the (+1)Imageth column of I, that is, the image at time tImage.

13.5.1 Runge–Kutta methods

In the following we will use notation that is common in optimal control publications. In particular, x is not a spatial variable, but it is the state we are interested in, for example, an evolving image or transformation map. Solving the transport equation (13.29) is an initial value problem of the form

˙x(t)=f(x(t),u(t)),x0x(0)=x0,t[0,1],

Image(13.32)

where x(t)RpImage with initial value x0=x0RpImage is the state variable, and u(t)RqImage is an external input called control. The change over time is governed by the right-hand side function f:Rp×RqRpImage.

To avoid excessive smoothing when solving (13.29) numerically, we do not use methods that introduce additional diffusion to ensure stability for the Euler forward method such as Lax–Friedrichs [92]. Instead, we use a Runge–Kutta method, which includes at least part of the imaginary axis in its stability region. We can then discretize the spatial derivative of the transport equation using central differences as given in (13.30) and (13.31) while still maintaining numerical stability.

Runge–Kutta methods have been investigated in the context of optimal control, for example, by Hager [39,40]. Since we are dealing with a Hamiltonian system, we would ideally like to pick a symplectic Runge–Kutta scheme [43, p. 179] as proposed for image registration in [59]. Consistent approximations have been explored in [89]. In particular, we are interested in a symplectic symmetric Runge–Kutta scheme [20], which preserves the energy of the system (here the Sobolev norm of the velocity) [85]. Energy-preserving Runge–Kutta methods have been investigated in [19]. The simplest such method is the implicit midpoint method [59,19]. However, implicit Runge–Kutta methods require the solution of an equation system at every iteration step. Although this can be accomplished by Newton's method or a fixed point iteration, it may compromise the symplectic property [93] and can become computationally costly.

We therefore restrict ourselves to explicit Runge–Kutta methods, which are nonsymplectic, but easy to implement, fast, and (at sufficiently high order) are only mildly dissipative and hence, experimentally, do not lose a significant amount of energy for the short time-periods we are integrating over for image registration.

We restrict the discussion to Runge–Kutta methods with sNImage stages. See [42, p. 132] for details. In the general case considered now, f has an explicit time dependence f:[0,1]×Rp×RqRpImage. Given the state variable x0=x(t0)=x0RpImage with tN1=htImage and the control uiRqImage, =0Image, 1,,N1Image, i=1,,sImage, the evolution of the state over time (xImage, =1,,N1Image) is computed with these methods.

One-step Runge–Kutta methods can be written as [42, p. 134]

yi=x+htsj=1aijf(t+cjht,yj,uj)x+1=x+htsi=1bif(t+ciht,yi,ui),i=1,,s,=0,,N2.}

Image(13.33)

According to (13.32) and our definition of f, x+1x(t+1)Image is the approximated state, yix(t+ciht)Image are the intermediate discrete states, and uiu(t+ciht)Image are the given discrete control variables. The matrix ARs×sImage and the vectors c, bRsImage depend on the chosen Runge–Kutta method. If c1=0Image and A is lower triangular, then the Runge–Kutta method is explicit; otherwise, it is implicit [14, p. 98]. In Table 13.1, A, b, and c are given in the so-called Butcher tableau for the fourth-order Runge–Kutta methods used for solving the state and adjoint equations in this work.

Table 13.1

Butcher tableaux of the fourth-order explicit Runge–Kutta methods used in this chapter. For the adjoint system, the matrix ˉAImage with ˉaij=bjajibiImage, i,j = 1,…,s, is given.

General Butcher tableauRunge–Kutta 4Adjoint Runge–Kutta 4
cA
b
00000
12Image12Image000
12Image012Image00
10010
16Image13Image13Image16Image
00100
12Image0012Image0
12Image00012Image
10000
16Image13Image13Image16Image

Image

Because all considered constraints can be written in the form of (13.32) and thus do not have an explicit dependence on time, we can simplify (13.33) to

yi=x+htsj=1aijf(yj,uj),x+1=x+htsi=1bif(yi,ui),i=1,,s,=0,,N2.}

Image(13.34)

13.5.2 Runge–Kutta methods for the adjoint system

As mentioned before, a consistent discretization of the energies and constraints is desirable. Therefore when using Runge–Kutta integrations for time-dependent constraints, we need to compute the adjoint model of the chosen Runge–Kutta integrator. This was worked out by [40]. For completeness, we give the result using our notation. Note that the optimal control problems for the relaxation approaches considered in this chapter are Bolza problems of the form

argminx(1),uEB(x(1),u)s.t. ˙x(t)=f(x,u),x(0)=x0=x0,EB(x(1),u)C1(x(1))+10C2(x(t),u(t))dt.}

Image(13.35)

Here C1:RpRImage is only depending on the final state (corresponding to the distance measure in the image registration), and C2:Rp×RqRImage is a cost function depending on all intermediate states and controls (which is the regularization part in LDDMM). In [40] Mayer problems with

argminx(1)EM(x(1))s.t. ˙x(t)=f(x,u),x0=x(0)=x0}

Image(13.36)

are solved. Mayer and Bolza problems are equivalent as they can be converted into each other [25, p. 159], and hence the results of Hager still apply to our problems. For the sake of a compact notation, we omit the superscripts M and B. Next, we briefly summarize how to obtain the adjoint equations of (13.35); for a complete derivation, see [40].

We introduce the arrays

X(x)N1=0Rp×N,Y(yi)N1,s=0,i=1Rp×N×s,U(ui)N1,s=0,i=1Rq×N×s,

Image

which contain all discrete states, intermediate states, and control variables. Following [40] and given the discretized energy E(xN1,U)Image, Eqs. (13.34) are added as constraints through the Lagrangian multipliers

Λ(λ)N1=1Rp×N1 and Ξ(ξi)N2,s=0,i=1,

Image

yielding the Lagrange function

L(X,Y,U,Λ,Ξ)=E(xN1,U)+N2=0[λ+1(x+1xhtsi=1bif(yi,ui))+si=1(ξi)(yixhtsj=1ai,jf(yj,uj))].}

Image(13.37)

By computing the partial derivatives of (13.37) with respect to xImage, yiImage, and uiImage and substituting

χi=λ+1+sj=1ajibiξj,

Image(13.38)

we obtain after some algebra the adjoint system and the gradient with respect to uiImage as [40]

λN1=xE(xN1,U),χi=λ+1+htsj=1bjajibi(xf(yj,uj))χj,λ=λ+1+htsi=1bi(xf(yi,ui))χi,uiL=uiEht(uf(yi,ui))χi.}

Image(13.39)

In the case of shooting with the EPDiff equation everything is determined by the initial momentum. Hence the derivatives with respect to x0Image of (13.37) also involve the derivative of E. The partial derivative of (13.37) with respect to the initial condition x0Image can be shown to be

Lx0=Ex0λ1htsi=1bi(xf(yi0,ui0))χi0.

Image(13.40)

The latter part is (formally) the negative of the adjoint variable λ evaluated for =0Image:

Lx0=Ex0λ0.

Image(13.41)

Note that λImage is strictly only defined for =1,,N1Image and λ0Image is simply a convenience notation. The adjoint equations given in (13.39) represent a Runge–Kutta method with b and c unchanged and aij=bjajibiImage. In particular, this implies that an explicit (implicit) Runge–Kutta method yields an explicit (implicit) Runge–Kutta method for the adjoint equations (with time reversed). The adjoint of the considered Runge–Kutta method is also shown in Table 13.1. When applied in reverse time direction, it is identical to the Runge–Kutta method used for the numerical integration of the state. Next, we will explicitly compute the equations for the 1D LDDMM registration using a transport equation on the images (i.e., the IBR model).

13.5.3 Application to the IBR model

The states we are interested in are the discrete images X=IRm×NImage. The control sequence that influences the image transport is the velocity matrix U=vRn×NImage.

For the three-dimensional registration problems we are addressing later, memory consumption cannot be neglected. We therefore propose reducing the number of stored control variables uiImage. We found empirically that vv+1/vImage for all =0,,N2Image was small, that is, the velocities v only change marginally over time. Hence we approximate the velocity fields within one Runge–Kutta step as piecewise constant: uiu1=:uImage, i=2,,sImage. This is similar to the classical relaxation approach by Beg et al. [7], where velocities within a time-step are also assumed to be constant, and maps are then computed via a semi-Lagrangian scheme based on these constant velocity fields. Assuming piecewise constant velocity fields for our approach reduces the memory requirement for storing the control variables in U by a factor s. Empirically, this does not affect the results greatly and still constitutes a possible discretization choice for our discretize–optimize approach. Memory requirements are reduced further by the fact that the update in the last equation of (13.39) now only needs to be computed for one instead of s controls for each time tImage. Additionally, the χiImage are then only needed in the update for λImage and do not have to be stored for later computations to update the controls as χ1=λImage for the explicit methods considered for numerical integration.

As motivated in Section 13.4.1, n and m might be different, and we use the interpolation matrix P to change from the v given on a nodal grid to the resolution of I that is given on a cell-centered grid. The right-hand side function for our Runge–Kutta method then follows from (13.29) to (13.31):

f1(I,v)diag(DI1I)Pv=(DI1I)(Pv),

Image(13.42)

where ⊙ denotes the Hadamard product. The resulting derivatives needed are:

If1(I,v)=diag(Pv)DI1,

Image(13.43)

vf1(I,v)=diag(DI1I)P.

Image(13.44)

The concrete Bolza energy (13.35), that is, the discretized objective function of (13.10), for our problem is given as

E1(IN1,v)Regh(ˉv)+Simh(IN1,I1).

Image(13.45)

Taking a closer look at (13.21), we easily compute

vE1(IN1,v)=hvhtˉLWˉLˉv.

Image(13.46)

For the update of vImage, =0,,N1Image, the components n+1,,(+1)nImage of vE1(IN1,v)Image have to be used.

To compute the final adjoint state λN1RmImage, the derivative of the discrete distance measure is needed:

λN1=ASimh(IN1,I1),

Image(13.47)

where ASimhImage denotes the derivative with respect to the first argument of SimhImage.

13.5.4 Application to the MBR model

As a smooth transition at the boundary of the domain is expected for the transformations, we assume homogeneous Neumann boundary conditions. The corresponding matrix to compute the discrete derivative for one component of the transformations given on a nodal grid is

Dv1=12hv(0010110100)Rn×n.

Image(13.48)

The discrete transformation maps ϕϕ(xnd)Rn×NImage are discretized exactly like the velocities. As before, ϕImage is the (+1)Imageth column of ϕ and {0,1,,N1}Image. To include the full information provided by the images, we will interpolate I0Image linearly using PϕN1Image and write ˜I0I0(PϕN1)Image for the resulting image. The discretized objective function of (13.14) then has the following form:

E2(ϕN1,v)=Regh(ˉv)+Simh(˜I0,I1).

Image(13.49)

Prolongating ϕN1Image with P also has an impact on the size of the adjoint variable λRm×NImage, and the final state is

λN1=ddPϕN1Simh(˜I0,I1)=ASimh(˜I0,I1)h˜I0.

Image(13.50)

For the second equation, the chain rule was applied yielding the product of the discrete gradient h˜I0RmImage and ASimhImage.

The discrete right-hand side function f2Image for the transport of the transformation maps is almost identical to (13.42), only the discrete derivatives are changed, and no interpolation is needed:

f2(ϕ,v)=diag(Dv1ϕ)v=(Dv1ϕ)v.

Image(13.51)

The derivatives of (13.51) needed for computing the adjoint equations are:

ϕf2(ϕ,v)=diag(v)Dv1,

Image(13.52)

vf2(ϕ,v)=diag(Dv1ϕ).

Image(13.53)

The regularization, which depends on the control variable, in this model (13.49) is the same as (13.45), and hence the partial derivative is equal to (13.46). However, the distance measure depends on the interpolated image I0(PϕN1)Image, and it is possible that ϕN1Image has a lower resolution than I0Image. Because the high resolution of the image mismatch (this is essentially the role of the adjoint λ) should be retained during the solution of the adjoint equations, the matrix P is used to connect Eqs. (13.52) and (13.53) to (13.50). The evolution of λ backward in time is then determined by (13.43) instead of (13.52). Also, before using λ to compute the update for the control v using (13.53), the grid change has to be undone by computing PλImage. By doing computations stepwise (i.e., storing only λImage and λ1Image on the high image resolution, =1,,N1Image) and keeping the shorter vectors PλImage the memory requirements can be reduced.

13.5.5 Application to the MBS model using EPDiff

We briefly repeat the constraints of the MBS model:

˙ϕt+Jϕtvt=0,ϕ0=Id,

Image(13.54)

˙mt+Jmtvt+mtdiv(vt)+Jvtmt=0,

Image(13.55)

vt=Kmt,K=(LL)1.

Image(13.56)

How (13.54) is solved numerically is described in Section 13.5.4. In 1D the EPDiff equation (13.55) simplifies to

˙mt+(xmt)vt+2mt(xvt)=0.

Image(13.57)

Because of (13.56), the momentum m is discretized on the same grid as the velocities. Hence our discrete representation is given on the nodal grid xndImage: m=m(xnd)Rn×NImage, and the matrix Dv1Image is also used for numerical derivation of the momentum.

The adjoint of the discrete differential operator L defined in (13.23) is the transposed matrix LImage. As L is symmetric, L=LImage. Furthermore, for a positive weight γ, the matrix L is positive definite, and we can deduce that LLImage is invertible. We can deduce that the kernel K(LL)1Rn×nImage is positive definite and (13.56) can be discretized as v=KmImage. Due to the positive definiteness of K, the latter equation system has a unique solution for all {0,,N1}Image. Note that the equation system v=KmImage can be solved efficiently using the fast Fourier transform; see, for example, [7,107]. The discrete energy for the MBS model can be simplified as the regularizer only depends on the initial momentum (see the last row of (13.17)):

E3(m,ϕ)=hv2(m0)Km0+Simh(˜I0,I1),

Image(13.58)

where we used again the notation ˜I0=I0(PϕN1)Image. The transformations ϕ are updated by the following right-hand side function, which is obtained by substituting v=KmImage into (13.51):

f13(ϕ,m)=diag(Dv1ϕ)Km.

Image(13.59)

The second discrete right-hand side that is used in the Runge–Kutta method is the discrete version of (13.57), where again v=KmImage is used:

f23(m)=diag(Dv1m)Km2diag(Dv1Km)m.

Image(13.60)

Again, for solving the adjoint equations, we need the derivatives. The difference is that we now have two state variables ϕ and m and accordingly two adjoint variables λ1λϕRn×NImage and λ2λMRn×NImage with intermediate stages χ1,iχϕ,iRn×NImage and χ2,iχM,iRn×NImage, i=1,,sImage, respectively. We will now summarize the derivatives of the right-hand side function f3=(f13,f23)Image with respect to the state variables ϕImage and mImage and omit the function arguments for convenience:

ϕf13=diag(Km)Dv1,

Image(13.61)

mf13=diag(D1vϕ)K,

Image(13.62)

ϕf23=0,

Image(13.63)

mf23=2diag(Dv1Km)diag(Km)Dv12diag(m)Dv1Kdiag(Dv1m)K.

Image(13.64)

Contrary to the two models before, we do not need to update our control u=vImage as it is directly available by (13.56), and therefore we do not need to compute ufImage. Substituting (13.61)(13.64) into the adjoint system (13.39) yields:

λ1=λ1+1+htsi=1bi(ϕf13)χ1,i,

Image(13.65)

χ1,i=λ1+1+htsj=1bjajibi(ϕf13)χ1,j,

Image(13.66)

λ2=λ2+1+htsi=1bi(mf13mf23)(χ1,iχ2,i),

Image(13.67)

χ2,i=λ2+1+htsj=1bjajibi(mf13mf23)(χ1,jχ2,j).

Image(13.68)

The final states of the adjoints λ1Image and λ2Image are given by the partial derivatives with respect to the final states ϕN1Image and mN1Image, respectively:

λ1N1=PϕN1E3(m,ϕ)=ASimh(˜I0,I1)h˜I0,

Image(13.69)

λ2N1=mN1E3(m,ϕ)=0.

Image(13.70)

The update of the initial momentum m0Image, which is the initial value influencing the complete model via Eqs. (13.54)(13.56) is given by adapting (13.41):

Lm0=E3(m,ϕ)m0λ20=hvKm0λ20.

Image(13.71)

13.6 Discretization in multiple dimensions

This section generalizes the equations from Section 13.4 and Section 13.5 to multidimensional LDDMM registration problems. Multidimensional formulations can easily be obtained from the one-dimensional representation by defining appropriate discrete operators. We illustrate this here for the three-dimensional case, but the approach is adaptable to arbitrary dimensions. For the computation of distance measures and their derivatives, we refer to [66, p. 95].

13.6.1 Discretization of the regularizer

The number of image voxels per dimension is denoted as MiImage, i=1,,3Image, and the number of nodal grid points as niImage. We define the total number of voxels and grid points as

M3i=1Mi,n3i=1ni.

Image(13.72)

The discrete images are thus I0,I1RMImage, and the velocities are discretized as vR3n×NImage. Our domain has a box shape: ˉΩ=[ω1,ω2]×[ω3,ω4]×[ω5,ω6]R3Image. The resulting cell widths are hI1=ω2ω1M1Image, hI2=ω4ω3M2Image, hI3=ω6ω5M3Image for the images and hv1=ω2ω1n11Image, hv2=ω4ω3n21Image, hv3=ω6ω5n31Image for the velocities and transformations. For the computation of the regularizer, we need the volume of one cell that is defined as

hvhv1hv2hv3.

Image(13.73)

The construction for the discrete Laplacian is achieved using Kronecker products (see [66, p. 122]):

Δ=Idn3Idn2Δhv1+Idn3Δhv2Idn1+Δhv3Idn2Idn1,

Image(13.74)

where ΔhivImage is the finite difference matrix given in (13.24), which is used to compute the second derivative in the ith dimension of one component of the velocities v. Hence we can write the multidimensional regularization matrix as an extension of (13.23):

L=(γIdnαΔ)βRn×n.

Image(13.75)

This makes multiplication of L with one component of the vector-valued velocities at each discretization time point possible. As in ReghImage, the velocities are integrated over time after multiplication with the Helmholtz operator, and L has to be multiplied with all components and for all times. We use copies of L for a short description but do not keep the copies in memory. In fact, as L is sparse, we implemented the multiplication with L without building the matrix. The corresponding matrix ˉLImage is

ˉLId3NL,

Image(13.76)

and the matrix for the proper weighting of cells between grid points that are located at the boundaries of the box is realized as in (13.25), yielding

WWNId3Wn3Wn2Wn1.

Image(13.77)

The discrete regularization energy is then given by applying the trapezoidal rule:

Regh(ˉv)hvht2ˉvˉLWˉLˉv.

Image(13.78)

The vector ˉvR3nNImage is the linearized matrix v, that is, the velocities are stacked up dimension-by-dimension for one time tImage: First, all x-components of the velocities (v1RnImage), then all y-components (v2RnImage), and finally all z-components (v3RnImage) are stored. Then the same is done with the next time point, and so on. In the following equation we summarize this procedure using subvectors viImage, i=1Image, 2, 3, =0,1,,N1Image, of v:

v(v10v11v1N1v20v21v2N1v30v31v3N1).

Image(13.79)

For the shooting model, the kernel K is needed. As in the one-dimensional case, it is defined as

K=(LL)1Rn×n.

Image(13.80)

Note that the matrix K is not stored, but equation systems are solved to obtain the momentum fields from the velocity fields; see Section 13.5.5.

13.6.2 Integrands and associated gradients for the Runge–Kutta methods

To solve the Runge–Kutta integration (13.34) and in particular the adjoint equations (13.39), the specifications of the right-hand side function f and its derivatives are required. We distinguish here solutions of the transport equation (for a relaxation approach) and direct solutions of the LDDMM optimality conditions (for shooting). In the first case we consider image-based and map-based solutions. As the map-based solution has the advantage that considerably less image artifacts occur during numerical solution of the transport equation (see also Section 13.8.1), we will focus on this adaption for the shooting approach. Additionally, the shooting approach eliminates the dependency on the velocity v as it can be computed from the state variables. To evaluate f and solve the adjoint equations, we will need matrices to numerically compute the first derivative. In three dimensions these matrices are [66, p. 122]:

ˉD1Id3Id2D1,

Image(13.81)

ˉD2Id3D2Id1,

Image(13.82)

ˉD3D3Id2Id1,

Image(13.83)

where the subscripts denote the directions of the partial derivatives (x, y, and z), and (,){(I,M),(v,n)}Image.

13.6.2.1 IBR model

We start with the discretization of the relaxation approach with transportation of images, that is, model (13.10). This is the simplest case as the state X corresponds to the vectorized images I=(I)N1=0RM×NImage only. The control variables are the discretized velocities vR3n×NImage. We want to transport the images with all their details and bring the velocities to the image resolution using the trilinear interpolation matrix PRM×nImage; see Section 13.4.1 and references therein. With the notation Pvi=:ˆviImage, i=1Image, 2, 3, =0,,N1Image, the discretized transport equation becomes

f1(I,v)=3i=1diag(ˉDIiI)ˆvi.

Image(13.84)

The corresponding derivatives are:

If1(I,v)=3i=1diag(ˆvi)ˉDIi,

Image(13.85)

vf1(I,v)=diag((ˉDI1+ˉDI2+ˉDI3)I)P.

Image(13.86)

For the adjoint system, we need the final state of λRM×NImage, which is already given in (13.47) but is repeated here for completeness:

λN1=ASimh(IN1,I1).

Image(13.87)

As our objective function is essentially the same as for the one-dimensional case given in (13.45), combining (13.46) and (13.78) yields for the derivative of the regularizer with respect to the velocities:

vE1(IN1,v)=hvhtˉLWˉLˉv.

Image(13.88)

13.6.2.2 MBR model

In the map-based case we transport all the coordinates (ϕ1Image, ϕ2Image, ϕ3Image) individually, that is, we have as many transport equations as there are spatial dimensions. The discrete transformation maps are ordered as

ϕ(ϕ10ϕ11ϕ1N1ϕ20ϕ21ϕ2N1ϕ30ϕ31ϕ3N1)R3n×N.

Image(13.89)

Thus ϕ and v have the same discretization, and the transport equations do not need any further interpolations and hence no matrix/operator P:

f2(ϕ,v)=(3i=1diag(ˉDviϕ1)vi3i=1diag(ˉDviϕ2)vi3i=1diag(ˉDviϕ3)vi).

Image(13.90)

The gradients with respect to the state variable ϕ for the individual components of (13.90) can be expressed by the Kronecker delta as the transport in the different spatial directions is decoupled:

ϕlfj2(ϕ,v)=δj,l3i=1diag(vi)ˉDvi,j,l=1,2,3.

Image(13.91)

The derivatives with respect to the control variable v are

vf2(ϕ,v)=(diag(ˉDvjϕi))3i,j=1.

Image(13.92)

To complete our adjoint system, we also need the final state of the adjoint variables λN1Image. Again, we abbreviate the transformed template images as ˜I0=I0(ˉPϕN1)Image, and ∘ denotes trilinear interpolation on the grid ˉPϕN1Image. Here the matrix ˉPImage contains three copies of P and converts each dimension of ϕN1Image to cell-centered points:

ˉP=(PPP)R3ˉm×3n.

Image(13.93)

The final state is similar to the one-dimensional case given in (13.50), but each transformation component is treated individually:

λN1=diag(h˜I0)(ASimh(˜I0,I1)ASimh(˜I0,I1)ASimh(˜I0,I1)).

Image(13.94)

Here h˜I0R3MImage is the discrete gradient of the image ˜I0RMImage, where the first M components are the partial derivatives in the x-direction, then the y- and finally the z-directions follow. The arguments given at the end of Section 13.5.4 still apply, and the adjoint equations are solved on the image resolution using the matrix P appropriately and storing the adjoint only when necessary at high resolution. As the regularization energy is the same as in the image-based model:

E2(ϕN1,v)=Regh(ˉv)+Simh(˜I0,I1).

Image(13.95)

The derivative with respect to the velocities is also the same:

vE2(ϕN1,v)=hvhtˉLWˉLˉv.

Image(13.96)

13.6.2.3 MBS model

Instead of solving map-based equations to decouple spatial transformations from the objects to be transported, we can directly solve the EPDiff equation (13.55) with the constraint (13.56) to optimize the discrete energy

E3(m,ϕ)=hv23i=1(mi0)Kmi0+Simh(˜I0,I1).

Image(13.97)

To compute the transformed source image, we still need the transformation ϕ and thus (13.90) to compute its evolution. The resulting discrete right-hand side function has six components:

fj3(ϕ,m)=3i=1diag(ˉDviϕj)Kmi,

Image(13.98)

fj+33(m)=3i=1(diag(ˉDvimj)Kmi+diag(ˉDvjKmi)mi+diag(mj)ˉDviKmi)

Image(13.99)

for j=1,2,3Image, where we used vi=KmiImage, i=1,2,3Image and =0,,N1Image. A useful relation when computing the derivatives of the right-hand side function is

mlvj=δj,lK={0,lj,K,l=j.

Image(13.100)

The resulting derivatives of f3Image are

ϕlfj3=δj,l3i=1diag(Kmi)ˉDvi,

Image(13.101)

mlfj3=diag(ˉDvlϕj)K,

Image(13.102)

ϕlfj+33=0,

Image(13.103)

mlfj+33=(diag(ˉDvlmj)+diag(ml)ˉDvj+diag(mj)ˉDvl)Kdiag(ˉDvjKml)δj,l3i=1(diag(ˉDviKmi)+diag(Kmi)ˉDvi),

Image(13.104)

for j, l=1,2,3Image.

As we have six state variables ϕ1Image, ϕ2Image, ϕ3Image and m1Image, m2Image, m3Image that are updated using Runge–Kutta methods with the right-hand sides f13,,f63Image, there are also six adjoint variables λ1,,λ6Image with intermediate variables χ1,i,,χ6,iImage for i=1,,sImage and =0,,N1Image. The derivatives of f3Image, given in (13.101)(13.104), are used in the adjoint Runge–Kutta system (13.39) to update the adjoint variables. It remains to give a complete description are the final states of the adjoints λ1N1,,λ6N1Image. They are given by the partial derivatives of E3Image with respect to the final states ϕN1Image and mN1Image, respectively. The first three equations are the same as in (13.94), but we repeat them for completeness, the last λ4N1Image, λ5N1Image, λ6N1Image vanish as the energy depends on the initial momentum and not on the final one:

(λ1N1λ2N1λ3N1)=diag(h˜I0)(ASimh(˜I0,I1)ASimh(˜I0,I1)ASimh(˜I0,I1)),

Image(13.105)

λjN1=0,j=4,5,6.

Image(13.106)

The update of the initial momentum m0Image that is the initial value influencing the complete model via Eqs. (13.98) and (13.99) is given by adapting (13.41):

Lm0=E3(m,ϕ)m0(λ40,λ50,λ60)=(hvKm10λ40hvKm20λ50hvKm30λ60).

Image(13.107)

13.7 Multilevel registration and numerical optimization

A key concept of most image registration algorithms is the registration on multiple levels. By computing solutions to registration problems with images whose resolution is gradually increased the risk of getting stuck in a local minimum is reduced [66, p. 13]. Additionally, the numerical optimization on the coarse levels with images consisting of fewer pixels/voxels is faster but gives a good approximation for the solution on the next level, further reducing runtime.

Because of these two reasons, we also employed a multilevel registration. The number of points in space varied during the multilevel optimization such that the grid spacing was halved on each finer level. After choosing n1j,j=1,2,3Image, the number of grid points is hence given as nij(n1j1)2i1+1Image. Given images I0,I1RM1×M2×M3Image, smoothed and downsampled versions I0i,I1iRMi1×Mi2×Mi3Image with Mij=Mj2F+iImage were computed for j=1,2,3Image and i=1,,FImage [66, p. 40]. Problems (13.10), (13.14), and (13.17) were then solved using a coarse-to-fine resolution multilevel strategy with FNImage levels. For the (bi/tri)linear interpolation, which is needed to compute the transformed source image I˜0Image, we used the Matlab code provided by [66].

The discretization resolution of the velocities, momenta, and transformations can be chosen by the user. One drawback is that (for the 3D lung registration, we are tackling) choosing nji=MjiImage exceeds common memory limitations and results in an extremely expensive solution of the constraint (13.34) and adjoint Eqs. (13.39). As motivated before, by LDDMM design v and ϕ are assumed to be smooth functions, and it is usually not necessary to use a high resolution for v and ϕ. This motivates our choice of nji<MjiImage.

To maintain accurate registrations, we use all given image information (at the highest resolution) by solving the adjoint equations at image resolution as given in the final adjoint state Eqs. (13.87), (13.94), and (13.105) for the different approaches, respectively. Empirically, Mji4njiImage worked well for lung CT registrations in our previous work [76]. The appropriate (trilinear) interpolations for the change between grids were realized with P. As mentioned before, P is sparse but large, but P does not need to be kept in memory [51,80]. We use matrix notation but implemented a matrix-free operator following [51,80].

It is useful to start a registration with the result of a preregistration ψR3nImage. Given a (preferably diffeomorphic) transformation ψ only the initial conditions of the problems (13.10), (13.14), and (13.17) have to be adapted: I(0)=I0ψImage or ϕ(0)=ψImage, respectively. In discrete notation these equations read I0=I0(Pψ)Image and ϕ0=ψR3nImage.

The number of simulated points in time N is very important for the LDDMM registration. From the stability function [41, p. 16] for the explicit Runge–Kutta 4th-order the maximal admissible time step for solving the discrete transport equation can be derived as [73, p. 175]

ht22(i=13aihi)1,

Image(13.108)

where aiImage is the expected maximal magnitude of movement in direction xiImage. An intermediate temporal step size was determined as

h˜t=2.8(i=13aihi)1.

Image(13.109)

Using 2.8 instead of 22Image adds a safety margin shielding against inaccuracies due to discretization errors. Additionally, as we are interested in integer values N, we compute:

N=h˜t1+1,ht=1N1,

Image(13.110)

which results in an even larger safety margin. According to (13.108), the time step has to decrease when the grid step size decreases, and therefore we determine htImage at the finest level (i=FImage). The main difficulty left is the estimation of the maximal motion that occurred to obtain aiImage, i=1,2,3Image. However, for typical medical registration scenarios in which we are interested, upper bounds for the motion that should be recovered by the registration are available. In the case of lung CT registration, even from full inspiration to full expiration, the maximal amplitude can be estimated (with a large safety margin) as 100 mm; see [17] for example datasets.

The discretized objective functions (13.45), (13.95), and (13.97) were optimized with an L-BFGS approach [71, p. 176] saving the last 5 iteration vectors for approximation of the inverse Hessian. The maximum number of iterations was set to kmax=50Image. Additional stopping criteria proposed in [66, p. 78] were used to terminate the optimization on each level. During the optimization, an Armijo line search with parameters βj=0.5j1Image, kmaxLS=30Image, and c1=106Image was used to guarantee a decrease of the objective function [71, p. 33]. The gradients used for the optimization are given in the last equation of (13.39) for the relaxation approaches and in (13.107) for the shooting approach. They are detailed for IBR in Section 13.6.2.1, Section 13.6.2.2, and Section 13.6.2.3 for the IBR, MBR, and MBS models, respectively. Before using the gradients of the relaxation methods for optimization, they are smoothed to obtain the Hilbert gradients (13.9) by N-fold application of K. As K1Image coincides with the second order derivative of ReghImage, this can be viewed as a Newton step only with respect to the regularization term.

The determinant of the Jacobian of the transformation can be used to find singularities (foldings of the grid or grid cells with vanishing volume) or to evaluate whether local volume expansion or contraction occurs. Therefore it is a good criterion to check for local invertibility [24, p. 222] and to find out if the computed transformation is a diffeomorphism. We computed Jacobian determinant values JiRImage, i=1,2,,=1k(nF1)Image on the last registration level F with the dedicated method used for hyperelastic regularization [13] in the FAIR [66] Matlab toolbox.5 The computed transformation ϕN1Image contained sporadically negative Jacobians for some parameter choices (and thus was not diffeomorphic), which resulted from inaccuracies that occur when solving the transport equation for iterated time steps with the fourth-order Runge–Kutta method. Therefore, we used the obtained velocities v to compute the final transformation φR3nImage in a slightly different manner. The fourth-order Runge–Kutta method (13.34) with the right-hand side given in (13.90) was solved for one time step starting from the identity grid xndImage with varying velocities vImage for =0,,N2Image to obtain small deformations φImage. These steps were then concatenated (starting from the preregistration ψ) using trilinear interpolation to obtain the total transformation

φφN2φN3φ0ψ.

Image(13.111)

If each of the transformations ψ, φImage for =0,,N2Image is diffeomorphic, then also φ is diffeomorphic [69,50,104]. However, repeated interpolations are time-consuming and difficult to handle when solving the adjoint system consistently. Therefore, during the numerical optimization, the methods presented in Section 13.5 and Section 13.6 are used to estimate the discrete velocity fields v. After the optimization is finished, the final transformation φ is computed within a postprocessing step according to (13.111) to achieve a transformation that is guaranteed to be diffeomorphic.

13.8 Experiments and results

This section shows experimental results on 2D radiographs of hands (Section 13.8.1) and 3D CT lung images (Section 13.8.2).

13.8.1 2D registration of hand radiographs

We illustrate our approach on a 2D dataset. The dataset consists of two 128×128Image pixel radiographs of human right hands of two different patients (see Fig. 13.3(A) and Fig. 13.3(F)). The dataset has previously been used in [65] and is part of the FAIR Matlab toolbox [66]. Before applying the LDDMM methods, the images were affinely preregistered using FAIR with SSD distance measure on the coarsest level of the image pyramid. SSD was also used for the LDDMM methods. According to (13.23), HsImage regularity with s>2Image is necessary for k=2Image (and thus β2Image) to guarantee diffeomorphic solutions. However, s=2Image is the limit case, and we therefore decided to employ β=1Image. The other parameters were determined empirically and fixed for all three approaches to α=10Image, γ=3Image, σ=1Image, F=3Image and M11=M21=32Image, n11=n21=9Image, N=10Image. The second column of Fig. 13.3 shows the results of the preregistration. The following columns display the results of the IBR, MBR, and MBS approaches. Properties of the computed transformation maps are most obvious in the Jacobian determinant visualizations in the second row of Fig. 13.3. The shooting approach clearly produces smoother transformations than both relaxation approaches. Furthermore, the volume change for the relaxation approaches is locally quite different indicating that advecting the non-smooth images (via the transport equation) introduces more numerical problems than advecting the comparatively smooth spatial transformations. The resulting distortion in the deformed moving image is apparent when comparing Fig. 13.3(C) and Fig. 13.3(D). In the former oscillations next to sharp edges in the images arise, whereas the latter image has the same quality as the undeformed moving image.

Image
Figure 13.3 Results for the hand examples. First row: Fixed image I1 and transformed moving images. Second row: Moving image I0 and Jacobians of the deformation grids. Third row: Absolute differences of fixed and (transformed) moving image. Second column: Results for the affine pre-registration. Third column: Results for the IBR approach. Fourth column: Results for the MBR approach. Fifth column: Results for the MBS approach. The three different LDDMM schemes were applied with the same parameters: α = 10, β = 1, γ = 3, σ = 1, F = 3 and m11=m21=32Image, n11=n21=9Image, N = 10.

Due to their representational flexibility, the relaxation approaches reach a slightly better image match as visible in the last row in the difference images. In particular the ring and the baby finger of the shooting result lack some width compared to the fixed image and are closer to the width of the fingers in the undeformed source image I0Image. The inferior image match is quantified by the three SSD values for the different approaches. These values also indicate that due to the image deterioration induced by the IBR approach, the image match is not as good as with the MBR approach. However, although the differences in image mismatch when comparing IBR and MBR to MBS are relatively small, the difference in transformation field smoothness is remarkable.

13.8.2 3D registration of lung CT data

To demonstrate the suitability of our method for 3D lung CT data, we make use of the publicly available DIR-Lab datasets [16,17,15]. These datasets were acquired from 20 subjects (with or without pulmonary diseases) during the inhale and exhale phases and are either part of the COPDgene study [77] or 3D volumes from 4DCT scans. The number of voxels per axial slice is 512×512Image for all COPD datasets [17] and 4DCT datasets 6 to 10 [16]. For 4DCT datasets 1 to 5 [15], the number of voxels per axial slice is 256×256Image. Each slice has a thickness of 2.5 mm, and the number of slices ranges between 94 and 135. The axial resolution varies from 0.586mm×0.586mmImage to 0.742mm×0.742mmImage for the COPD data [17], is 0.97mm×0.97mmImage for 4DCT datasets 6 to 10 [16] and ranges from 0.97mm×0.97mmImage to 1.16mm×1.16mmImage for 4DCT datasets 1 to 5 [15]. Additionally to the images, 300 expert-placed landmarks per image are provided. These landmarks can be used for evaluation of registration accuracy.

Statistics of the Jacobian of the computed deformation fields are often used to evaluate the plausibility of the registration see, for example, [68] for the application to lung CT registration. Therefore we evaluate the Jacobians for the computed final transformation φ to estimate the reconstructed volume change of local regions of the lung and to assess for unreasonable expansions, contractions, or foldings of lung tissue.

In many clinical applications only information within the lungs is required, for example, for the staging and phenotyping of COPD [32]. Although the motion of the lung induced by respiration is smooth (making LDDMM a suitable registration model), motion discontinuities occur at the interface with the ribcage [88]. These discontinuities interfere with the LDDMM model, although in [78] a modification is presented, which applies special kernels to allow for discontinuous motions. However, as we are only interested in the alignment of the lungs, we generated lung segmentations using the method proposed in [53] and masked the CT scans to remove unnecessary image information. Hence an adaptation of our models is no longer necessary. In addition to the masking the images were cropped such that in each direction approximately 5 mm of margin were retained. For all registrations, the inhale scans were used as target images and the exhale scans as source images.

In the following experiments we used a thin-plate spline (TPS) preregistration. The TPS is based on keypoints acquired with the method in [46], which were also used for the registration of difficult COPD data with remarkably accurate results [82]. However, the TPS preregistration with these keypoints may produce transformations with foldings (i.e., negative Jacobians). Therefore, to ensure that ψ is diffeomorphic, we removed keypoints whose set of six nearest neighbors changes after TPS computation by more than one point, suggesting a violation of topology preservation. This procedure reduced the number of keypoints to approximately one quarter of the original number and results in decreased accuracy of the preregistration; see Table 13.2. The subsequent registration with our LDDMM methods MBR and MBS retained the diffeomorphic property of the estimated transformation, which might not be the case for small deformation methods; see the discussion in [23].

Table 13.2

Mean landmark distances of the 300 landmark pairs after registration with state-of-the-art methods. All values are given in mm and were computed after rounding to the next voxel center.

blank cellMean landmark distance after registration in mm
CaseInitialMILOMRFLMPDIS-COisoPTVPrereg.MBRMBS
[18][46][74][82][99]
4DCT13.890.740.760.800.740.76
4DCT24.340.780.770.750.800.82
4DCT36.940.910.900.950.930.95
4DCT49.831.241.241.391.241.31
4DCT57.481.171.121.251.101.11
4DCT610.890.900.851.040.860.87
4DCT711.030.870.801.090.830.88
4DCT814.991.041.341.201.111.10
4DCT97.920.980.921.100.960.93
4DCT107.300.890.821.090.880.90
Average8.460.950.951.070.950.96
COPD126.340.931.001.210.790.771.151.120.91
COPD221.791.771.621.971.462.222.181.491.49
COPD312.640.991.001.060.840.821.190.970.96
COPD429.581.141.081.640.740.851.320.890.97
COPD530.081.020.961.460.710.771.180.790.92
COPD628.460.991.011.340.640.861.270.840.92
COPD721.601.031.051.160.790.741.320.920.93
COPD826.461.311.081.540.770.811.471.031.03
COPD914.860.860.790.990.620.831.020.750.73
COPD1021.811.231.181.390.860.921.511.021.09
Average23.361.131.081.380.820.961.360.981.00
Total Average15.911.131.081.160.820.961.210.960.98

Image

Due to the deteriorated quality of the deformed image of the IBR approach (see Section 13.8.1) and the interest in the volume change and plausibility of the transformation maps, we only applied the MBR and MBS approaches to the 3D lung datasets. We chose NGF as the distance measure, as it has been shown to work well on lung CT data; see Section 13.1. The same model parameters were used for all registrations: α=10Image, β=2Image, γ=1Image, σ=0.05Image, n1=(13,13,17)Image, and F=3Image. For the estimation of the time step htImage, we need an estimate for the maximal motion that may have occurred, which was denoted as aiImage, i=1,2,3Image, in Section 13.7. As the employed preregistration achieves a very good initial alignment and dramatically reduces the maximal motion required, we chose ai=30Image, i=1,2,3Image, and the number of time steps according to (13.110).

Registration results for dataset 4 of the COPD datasets are shown as coronal views in Fig. 13.4. The initial unaligned images are shown in the leftmost column; note the differences in noise level and volume for the exhale scan (I0Image) and the inhale scan (I1Image). The results for the preregistration ψ alone are very good and shown in the second column. Only smaller vessels and some lung boundaries show visually obvious mismatches. Our subsequent LDDMM registrations MBR and MBS, shown in the third and fourth columns, respectively, successfully handle these mismatches and produce visually very similar results. The second row shows that the Jacobian determinant maps for all registration approaches are smooth and positive. The largest volume changes occur near the mediastinum and might be influenced by the beating heart. Taking a closer look at the image of the Jacobian determinants of MBS (Fig. 13.4(H)) reveals that it is smoother than that of MBR (Fig. 13.4(G)) regarding variations in the left-right direction. Both approaches yield very smooth results in the anterior-superior direction. In particular, the lung volume decreases (Jacobian determinant values smaller than 1), which is what we expect during exhalation, which is modeled by our choice of using the inhale phase as the fixed image and the exhale phase as the moving image.

Image
Figure 13.4 Exemplary result for the DIR-Lab COPD lung CT dataset 4 [17]. First row: Fixed image I1 and transformed moving images. Second row: Moving image I0 and Jacobians of the deformation grids overlaid on the fixed scan. Third row: Overlays of fixed (orange (light gray in print version)) and (transformed) moving image (blue (dark gray in print version)); due to addition of RGB values, aligned structures appear gray or white. Second column: Results for the preregistration. Third column: Results for the MBR approach. Fourth column: Results for the MBS approach. MBR and MBS schemes were applied with the same parameters: α = 10, β = 2, γ = 1, σ = 0.05, F = 3 and n11=n21=13Image, n31=17Image, N = 9.

Table 13.2 shows the mean landmark distances for all 20 datasets after MBR and MBS registration and allows for a comparison to reported results of state-of-the-art methods participating in the DIR-Lab benchmark.6 For the 4DCT data, all compared algorithms achieve similar results regarding landmark accuracy. Results on the DIR-Lab 4DCT data were not reported in the literature for MILO [18], MRF [46], and DIS-CO [82] and are therefore not listed. DIS-CO [82] achieves the best mean landmark distance (also called target registration error (TRE)) on the more challenging COPD data. The second best method regarding TRE is [99] but is closely followed by the MBR and the MBS approaches. The algorithm proposed in [99] does not use lung segmentations for image masking, but its total variation-based regularization allows for sliding motions. However, this also has the drawback that there is no guarantee that the estimated transformations contain no foldings.

The advantage of the MBR and MBS methods (ranking third and fourth regarding TRE) is that they produce transformation maps without foldings or implausible volume changes within the lungs; see Table 13.3. Only two other methods, LMP [74] and DIS-CO [82], can guarantee this regularity, and LMP achieves a worse TRE on the COPD data than both MBR and MBS.

Table 13.3

Minimum, mean, and standard deviation (SD) of the Jacobian determinant values of the computed transformations φR3n¯Image restricted to the lung volume. The mean value should be similar to the volume ratio (VR), which was computed as the ratio of exhale lung volume to inhale lung volume using the lung masks.

MBRMBS
CaseVRMin.Mean ± SDMin.Mean±SD
4DCT10.910.480.92±0.100.500.92±0.10
4DCT20.920.450.92±0.090.420.92±0.11
4DCT30.900.470.90±0.100.520.90±0.11
4DCT40.860.460.86±0.150.410.86±0.16
4DCT50.910.460.92±0.130.440.92±0.15
4DCT60.770.420.77±0.090.440.77±0.09
4DCT70.820.430.79±0.090.440.79±0.09
4DCT80.830.340.80±0.090.340.80±0.10
4DCT90.860.460.87±0.100.420.87±0.10
4DCT100.860.310.87±0.130.320.87±0.14
COPD10.620.120.65±0.150.170.65±0.14
COPD20.730.150.75±0.240.170.75±0.25
COPD30.800.290.80±0.130.290.80±0.14
COPD40.490.180.50±0.120.160.50±0.12
COPD50.490.150.49±0.110.150.49±0.11
COPD60.620.180.64±0.160.060.64±0.16
COPD70.750.390.75±0.110.380.75±0.12
COPD80.580.210.59±0.130.230.59±0.13
COPD90.730.320.74±0.120.320.74±0.13
COPD100.610.110.61±0.120.110.61±0.13

Image

The Jacobians given in Table 13.3 show that the computed transformation field is indeed very smooth, as the standard deviation is small. Furthermore, the mean volume change is well correlated with the expected volume ratio, and no foldings of the grid occurred as the minimal Jacobian is always positive. To obtain diffeomorphic transformations φ from the previously computed ϕN1Image in the discrete framework, we used the procedure described in Section 13.7. The transformations ϕN1Image and φ show only small differences. The mean value of ϕN1φImage computed over all 20 DIR-Lab datasets is 3.15 mm for the MBR method and 5.80 mm for the MBS method.

All experiments were run on a PC with a single 3.4 GHz Intel i7-2600 quad-core CPU, 16 GB of RAM, and Ubuntu 14.04. Runtimes for the MBS approach ranged between 16 and 55 min with an average of 28 min. The peak memory consumption was 4.3 GB over all datasets. Analogously, for MBR, the runtimes ranged between 5 and 24 min with an average of 14 min and a peak memory consumption of 2.1 GB.

13.9 Discussion and conclusion

LDDMM is a popular registration method with nice mathematical properties that has successfully been applied to various applications; see, for example, [64] and references therein. The estimated deformation fields are guaranteed to be diffeomorphic, the deformation paths are geodesics, large displacements can be captured, and LDDMM induces a distance measure on images [7]. These properties, however, come with the drawbacks of a large computational burden, both with respect to runtime and memory consumption.

It is interesting that numerical approaches for LDDMM have not been explored in great depth, and the predominant solution approach is still optimize–discretize via gradient descent as proposed in [7]. Hence we chose to develop discretize–optimize schemes for relaxation and shooting, offering consistent numerical optimizations for the discretized energy, which, as discussed in [35], cannot be guaranteed for optimize–discretize approaches such as [7]. Formulating a discretized energy included the discretization of PDE constraints (mainly transport and scalar conservation equations), which we accomplished via an explicit fourth-order Runge–Kutta method as a good compromise between efficiency and accuracy. We also reduced memory consumption and runtime by representing velocities and transformation maps at lower spatial resolutions compared to the images. This is justified due to the smoothness of velocity fields and transformation maps and still allowed for high-quality registrations. Finally, our proposed methods can also be used with different image similarity measures, and we explored the use of NGF for lung registration.

We demonstrated the characteristics of the proposed registration methods for a two-dimensional registration example of human hand radiographs. This experiment confirmed that the relaxation approaches yield a better image match due to more degrees of freedom compared to the shooting approach, which is much more constrained (solutions are geodesics by construction and not only at absolute convergence) and yields a smoother deformation field.

The methods were also tested on the challenging lung CT datasets provided by DIR-Lab [16,17,15], where we integrated the NGF distance measure to well align edges located at, for example, vessels and lung boundaries. We compared results based on the distances of 300 expert-annotated landmark pairs per inhale-exhale scan to the state-of-the-art. With an average error of 0.98 mm on the DIR-Lab COPD datasets the relaxation approach achieves the third-best result. The average error is only 0.02 mm higher than for the second-best method [99], which is almost negligible. The shooting approach performs slightly worse with an average landmark distance of 1.00 mm for the COPD datasets whilst yielding desirable smooth transformations. Nevertheless, the relaxation approach may be preferable for registration of lung CT scans from full-inspiration to full-expiration due to its reduced runtime.

In future work an extended evaluation that includes a comparison with publicly available diffeomorphic registration software packages like Deformetrica7 [28], MRICloud8 [49], LDDMM with frame-based kernel9 [94], FshapesTk10 [21], and Lagrangian LDDMM11 [58] would be beneficial.

For fast prototyping, the usage of automatic differentiation methods could be employed to derive a DO scheme from a discretized objective function. This has been done, for example, in [52] for shape matching. However, it is an interesting question how the computation time competes with the dedicated methods developed in this chapter. Automatic differentiation is also very popular in the field of machine learning. Recently, deep learning approaches for image registration have been proposed. Yang et al. [102,103] developed such a method for the shooting formulation of LDDMM. To refine the output of the deep-learning approach, it could be used as a fast initialization for a subsequent registration with our discretize–optimize approaches and/or the deep network could be trained based on our LDDMM approaches.

References

1. Vincent Arsigny, Olivier Commowick, Xavier Pennec, Nicholas Ayache, A log-Euclidean framework for statistics on diffeomorphisms, Rasmus Larsen, Mads Nielsen, Jon Sporring, eds. Medical Image Computing and Computer-Assisted Intervention – MICCAI 2006. Springer; 2006:924–931.

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

3. John Ashburner, Karl J. Friston, Diffeomorphic registration using geodesic shooting and Gauss–Newton optimisation, NeuroImage 2011;55(3):954–967.

4. Brian B. Avants, Charles L. Epstein, Murray Grossman, James C. Gee, Symmetric diffeomorphic image registration with cross-correlation: evaluating automated labeling of elderly and neurodegenerative brain, Medical Image Analysis 2008;12(1):26–41.

5. Robert Azencott, Roland Glowinski, Jiwen He, Aarti Jajoo, Yipeng Li, Andrey Martynenko, Ronald H.W. Hoppe, Sagit Benzekry, Stuart H. Little, William A. Zoghbi, Diffeomorphic matching and dynamic deformable surfaces in 3D medical imaging, Computational Methods in Applied Mathematics 2010;10(3):235–274.

6. Ruzena Bajcsy, Robert Lieberson, Martin Reivich, A computerized system for the elastic matching of deformed radiographic images to idealized atlas images, Journal of Computer Assisted Tomography 1983;7:618–625.

7. 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.

8. Alfio Borzì, Kazufumi Ito, Karl Kunisch, Optimal control formulation for determining optical flow, SIAM Journal on Scientific Computing 2003;24(3):818–847.

9. Chaim Broit, Optimal Registration of Deformed Images. [PhD thesis] USA: Computer and Information Science, University of Pensylvania; 1981.

10. Lisa Gottesfeld Brown, A survey of image registration techniques, ACM Computing Surveys 1992;24(4):325–376.

11. Thomas Brox, Jitendra Malik, Large displacement optical flow: descriptor matching in variational motion estimation, IEEE Transactions on Pattern Analysis and Machine Intelligence 2011;33(3):500–513.

12. Andrés Bruhn, Joachim Weickert, Christoph Schnörr, Lucas/Kanade meets Horn/Schunck: combining local and global optic flow methods, International Journal of Computer Vision 2005;61(3):211–231.

13. Martin Burger, Jan Modersitzki, Lars Ruthotto, A hyperelastic regularization energy for image registration, SIAM Journal on Scientific Computing 2013;35(1):B132–B148.

14. John C. Butcher, Numerical Methods for Ordinary Differential Equations. 3rd edition Chichester, UK: John Wiley & Sons Ltd.; 2016.

15. Richard Castillo, Edward Castillo, Rudy Guerra, Valen E. Johnson, Travis McPhail, et al., A framework for evaluation of deformable image registration spatial accuracy using large landmark point sets, Physics in Medicine and Biology 2009;54(7):1849–1870.

16. Edward Castillo, Richard Castillo, Josue Martinez, Maithili Shenoy, Thomas Guerrero, Four-dimensional deformable image registration using trajectory modeling, Physics in Medicine and Biology 2010;55(1):305–327.

17. Richard Castillo, Edward Castillo, David Fuentes, Moiz Ahmad, Abbie M. Wood, et al., A reference dataset for deformable image registration spatial accuracy evaluation using the COPDgene study archive, Physics in Medicine and Biology 2013;58(9):2861–2877.

18. Edward Castillo, Richard Castillo, David Fuentes, Thomas Guerrero, Computing global minimizers to a constrained B-spline image registration problem from optimal l1 perturbations to block match data, Medical Physics 2014;41(4), 041904.

19. Elena Celledoni, Robert I. McLachlan, David I. McLaren, Brynjulf Owren, G. Reinout W. Quispel, William M. Wright, Energy-preserving Runge–Kutta methods, ESAIM: Mathematical Modelling and Numerical Analysis 2009;43:645–649.

20. Robert P.K. Chan, On symmetric Runge–Kutta methods of high order, Computing 1990;45:301–309.

21. Benjamin Charlier, Nicolas Charon, Alain Trouvé, The fshape framework for the variability analysis of functional shapes, Foundations of Computational Mathematics 2017;17(2):287–357.

22. Christophe Chefd'hotel, Gerardo Hermosillo, Olivier Faugeras, Flows of diffeomorphisms for multimodal image registration, IEEE International Symposium on Biomedical Imaging – ISBI 2002. IEEE; 2002:753–775.

23. Gary E. Christensen, Richard D. Rabbitt, Michael I. Miller, Deformable templates using large deformation kinetics, IEEE Transactions on Image Processing 1996;5(10):1435–1447.

24. Philippe G. Ciarlet, Mathematical Elasticity, Volume I: Three-Dimensional Elasticity. North-Holland: Elsevier Science; 1988.

25. Domenico D'Alessandro, Introduction to Quantum Control and Dynamics. Boca Raton, FL, USA: Chapman & Hall/CRC; 2007.

26. Paul Dupuis, Ulf Grenander, Michael I. Miller, Variational problems on flows of diffeomorphisms for image matching, Quarterly of Applied Mathematics 1998;56(3):1–20.

27. Stanley Durrleman, Marcel Prastawa, Guido Gerig, Sarang Joshi, Optimal data-driven sparse parameterization of diffeomorphisms for population analysis, Information Processing in Medical Imaging 2011. Berlin/Heidelberg: Springer; 2011:123–134.

28. Stanley Durrleman, Marcel Prastawa, Nicolas Charon, Julie R. Korenberg, Sarang Joshi, Guido Gerig, Alain Trouvé, Morphometry of anatomical shape complexes with dense deformations and sparse parameters, NeuroImage 2014;101:35–49.

29. Bernd Fischer, Jan Modersitzki, Ill-posed medicine – an introduction to image registration, Inverse Problems 2008;24, 034008.

30. Martin A. Fischler, Robert A. Elschlager, The representation and matching of pictorial structures, IEEE Transactions on Computers 1973;22(1):67–92.

31. J. Michael Fitzpatrick, David L.G. Hill, Calvin R. Maurer Jr, Image registration, Jacob Beutel, J. Michael Fitzpatrick, Steven C. Horii, Yongmin Kim, Harold L. Kundel, Milan Sonka, Richard L. Van Metter, eds. Handbook of Medical Imaging, Volume 2: Medical Image Processing and Analysis. Bellingham, Washington, USA: SPIE; 2000:447–514.

32. Craig J. Galbán, Meilan K. Han, Jennifer L. Boes, Komal Chughtai, Charles R. Meyer, et al., Computed tomography-based biomarker provides unique signature for diagnosis of COPD phenotypes and disease progression, Nature Medicine 2012;18:1711–1715.

33. A. Ardeshir Goshtasby, Image Registration: Principles, Tools and Methods. New York: Springer; 2012.

34. Andreas Griewank, Andrea Walther, Evaluating Derivatives: Principles and Techniques of Algorithmic Differentiation. 2nd edition Philadelphia, PA, USA: SIAM; 2008.

35. Eldad Haber, Lauren Hanson, Model Problems in PDE-Constrained Optimization. [Technical report, TR-2007-009] Emory University; 2007.

36. Eldad Haber, Jan Modersitzki, Numerical methods for volume preserving image registration, Inverse Problems 2004;20(5):1621–1638.

37. Eldad Haber, Jan Modersitzki, Image registration with a guaranteed displacement regularity, International Journal of Computer Vision 2007;71(3).

38. Eldad Haber, Jan Modersitzki, Intensity gradient based registration and fusion of multi-modal images, Methods of Information in Medicine 2007;46(3):292–299.

39. William W. Hager, Rates of convergence for discrete approximations to unconstrained control problems, SIAM Journal on Numerical Analysis 1976;13(4):449–472.

40. William W. Hager, Runge–Kutta methods in optimal control and the transformed adjoint system, Numerische Mathematik 2000;87(2):247–282.

41. Ernst Hairer, Gerhard Wanner, Solving Ordinary Differential Equations II – Stiff and Differential–Algebraic Problems. 2nd edition Berlin/Heidelberg, Germany: Springer; 1996.

42. Ernst Hairer, Syvert P. Nørsett, Gerhard Wanner, Solving Ordinary Differential Equations I: Nonstiff Problems. Berlin/Heidelberg, Germany: Springer; 1993.

43. Ernst Hairer, Christian Lubich, Gerhard Wanner, Geometric Numerical Integration: Structure-Preserving Algorithms for Ordinary Differential Equations. 2nd edition Berlin/Heidelberg, Germany: Springer; 2006.

44. Joseph V. Hajnal, Derek L.G. Hill, David J. Hawkes, Medical Image Registration. Boca Raton, FL, USA: CRC Press; 2001.

45. Gabriel L. Hart, Christopher Zach, Marc Niethammer, An optimal control approach for deformable registration, Computer Vision and Pattern Recognition Workshops. IEEE; 2009:9–16.

46. Mattias Paul Heinrich, Heinz Handels, Ivor J.A. Simpson, Estimating large lung motion in COPD patients by symmetric regularised correspondence fields, Nassir Navab, Joachim Hornegger, William M. Wells, Alejandro F. Frangi, eds. Medical Image Computing and Computer-Assisted Intervention – MICCAI 2015. Springer; 2015:338–345.

47. Darryl D. Holm, Jerrold E. Marsden, Tudor S. Ratiu, The Euler–Poincaré equations and semidirect products with applications to continuum theories, Advances in Mathematics 1998;137:1–81.

48. Berthold K.P. Horn, Brian G. Schunck, Determining optical flow, Artificial Intelligence 1981;17(1–3):185–203.

49. Saurabh Jain, Daniel J. Tward, David S. Lee, Anthony Kolasny, Timothy Brown, J. Tilak Ratnanather, Michael I. Miller, Laurent Younes, Computational anatomy gateway: leveraging XSEDE computational resources for shape analysis, ACM International Conference Proceeding Series 2014:54:1–54:6.

50. Bilge Karaçalı, Christos Davatzikos, Estimating topology preserving and smooth displacement fields, IEEE Transactions on Medical Imaging 2004;23(7):868–880.

51. Lars König, Jan Rühaak, A fast and accurate parallel algorithm for non-linear image registration using normalized gradient fields, International Symposium on Biomedical Imaging – ISBI 2014. IEEE; 2014:580–583.

52. Line Kühnel, Stefan Sommer, Computational anatomy in Theano, Graphs in Biomedical Image Analysis, Computational Anatomy and Imaging Genetics. LNCS. Springer; 2017;vol. 10551:164–176.

53. Bianca Lassen, Jan-Martin Kuhnigk, Michael Schmidt, Stefan Krass, Heinz-Otto Peitgen Lung, Lung Lobe, Segmentation methods at fraunhofer MEVIS, Proceedings of the Fourth International Workshop on Pulmonary Image Analysis. 2011:185–199.

54. Yann LeCun, A theoretical framework for back-propagation, David S. Touretzky, Geoffrey Hinton, Terrence J. Sejnowski, eds. Proceedings of the 1988 Connectionist Models Summer School. Morgan Kaufmann; 1988:21–28.

55. Bruce D. Lucas, Takeo Kanade, An iterative image registration technique with an application to stereo vision, Proc. Seventh International Joint Conference on Artificial Intelligence. 1981;vol. 130:674–679.

56. J.B. Antoine Maintz, Max A. Viergever, A survey of medical image registration, Medical Image Analysis 1998;2(1):1–36.

57. Andreas Mang, George Biros, An inexact Newton–Krylov algorithm for constrained diffeomorphic image registration, SIAM Journal on Imaging Sciences 2015;8(2):1030–1069.

58. Andreas Mang, Lars Ruthotto, A Lagrangian Gauss–Newton–Krylov solver for mass- and intensity-preserving diffeomorphic image registration, SIAM Journal on Scientific Computing 2017;39(5):B860–B885.

59. Stephen Marsland, Robert McLachlan, A Hamiltonian particle method for diffeomorphic image registration, Nico Karssemeijer, Boudewijn Lelieveldt, eds. Information Processing in Medical Imaging – IPMI 2007. Berlin/Heidelberg: Springer; 2007:396–407.

60. Michael I. Miller, Computational anatomy: shape, growth, and atrophy comparison via diffeomorphisms, NeuroImage 2004;23(Suppl. 1):S13–S33.

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

62. Michael I. Miller, Alain Trouvé, Laurent Younes, Geodesic shooting for computational anatomy, Journal of Mathematical Imaging and Vision 2006;24:209–228.

63. Michael I. Miller, Laurent Younes, Alain Trouvé, Diffeomorphometry and geodesic positioning systems for human anatomy, Technology 2014;2(1):36–43.

64. Michael I. Miller, Alain Trouvé, Laurent Younes, Hamiltonian systems and optimal control in computational anatomy: 100 years since D'Arcy Thompson, Annual Review of Biomedical Engineering 2015;17(1):447–509.

65. Jan Modersitzki, Numerical Methods for Image Registration. New York: Oxford University Press; 2004.

66. Jan Modersitzki, FAIR: Flexible Algorithms for Image Registration. Philadelphia: SIAM; 2009.

67. David Mumford, Peter W. Michor, On Euler's equation and ‘EPDiff’, Journal of Geometric Mechanics 2013;5(3):319–344.

68. Keelin Murphy, Bram van Ginneken, Joseph M. Reinhardt, Sven Kabus, Kai Ding, et al., Evaluation of registration methods on thoracic CT: the EMPIRE10 challenge, IEEE Transactions on Medical Imaging 2011;30(11):1901–1920.

69. Oliver Musse, Fabrice Heitz, Jean-Paul Armspach, Topology preserving deformable image matching using constrained hierarchical parametric models, IEEE Transactions on Image Processing 2001;10(7):1081–1093.

70. Marc Niethammer, Yang Huang, François-Xavier Vialard, Geodesic regression for image time-series, Gabor Fichtinger, Anne Martel, Terry Peters, eds. Medical Image Computing and Computer-Assisted Intervention – MICCAI 2011. Springer; 2011:655–662.

71. Jorge Nocedal, Stephen Wright, Numerical Optimization. New York, NY, USA: Springer; 2006.

72. Nils Papenberg, Andrés Bruhn, Thomas Brox, Stephan Didas, Joachim Weickert, Highly accurate optic flow computation with theoretically justified warping, International Journal of Computer Vision 2006;67(2):141–158.

73. Thomas Polzin, Large Deformation Diffeomorphic Metric Mappings – Theory, Numerics, and Applications. [PhD thesis] University of Lübeck, Institute of Mathematics and Image Computing; 2018.

74. Thomas Polzin, Jan Rühaak, René Werner, Jan Strehlow, Stefan Heldmann, Heinz Handels, Jan Modersitzki, Combining automatic landmark detection and variational methods for lung ct registration, Proceedings of the Fifth International MICCAI Workshop on Pulmonary Image Analysis. 2013:85–96.

75. Thomas Polzin, Jan Rühaak, René Werner, Heinz Handels, Jan Modersitzki, Lung registration using automatically detected landmarks, Methods of Information in Medicine 2014;4:250–256.

76. Thomas Polzin, Marc Niethammer, Mattias Paul Heinrich, Heinz Handels, Jan Modersitzki, Memory efficient LDDMM for Lung CT, Sebastien Ourselin, Leo Joskowicz, Mert R. Sabuncu, Gozde Unal, William M. Wells, eds. Medical Image Computing and Computer-Assisted Intervention – MICCAI 2016. Springer; 2016:28–36.

77. Elizabeth A. Regan, John E. Hokanson, James R. Murphy, David A. Lynch, Terri H. Beaty, et al., Genetic epidemiology of COPD (COPDGene) study design, COPD 2011;7:32–43.

78. Laurent Risser, François-Xavier Vialard, Habib Y. Baluwala, Julia A. Schnabel, Piecewise-diffeomorphic image registration: application to the motion estimation between 3D CT lung images with sliding conditions, Medical Image Analysis 2013;17(2):182–193.

79. Torsten Rohlfing, Calvin R. Maurer Jr., David A. Bluemke, Michael A. Jacobs, Volume-preserving nonrigid registration of MR breast images using free-form deformation with an incompressibility constraint, IEEE Transactions on Medical Imaging 2003;22(6):730–741.

80. Jan Rühaak, Matrix-Free Techniques for Efficient Image Registration and Their Application to Pulmonary Image Analysis. [PhD thesis] Bremen, Germany: Jacobs University; 2017.

81. Jan Rühaak, Stefan Heldmann, Till Kipshagen, Bernd Fischer, Highly accurate fast lung CT registration, David R. Haynor, Sebastien Ourselin, eds. SPIE Medical Imaging 2013: Image Processing. 2013:86690Y1–86690Y9.

82. Jan Rühaak, Thomas Polzin, Stefan Heldmann, Ivor J.A. Simpson, Heinz Handels, Jan Modersitzki, Mattias Paul Heinrich, Estimation of large motion in lung CT by integrating regularized keypoint correspondences into dense deformable registration, IEEE Transactions on Medical Imaging 2017;36(8):1746–1757.

83. Martin Rumpf, Benedikt Wirth, Discrete geodesic calculus in shape space and applications in the space of viscous fluidic objects, SIAM Journal on Imaging Sciences 2013;6(4):2581–2602.

84. Lars Ruthotto, Jan Modersitzki, Non-linear image registration, Otmar Scherzer, ed. Handbook of Mathematical Methods in Imaging. New York: Springer; 2015:2005–2051.

85. Jesús María Sanz-Serna, Runge–Kutta schemes for Hamiltonian systems, BIT Numerical Mathematics 1988;28:877–883.

86. Ryo Sakamoto, Susumu Mori, Michael I. Miller, Tomohisa Okada, Kaori Togashi, Detection of time-varying structures by large deformation diffeomorphic metric mapping to aid reading of high-resolution CT images of the lung, PLoS ONE 2014;9(1):1–11.

87. Otmar Scherzer, Mathematical Models for Registration and Applications to Medical Imaging. New York: Springer; 2006.

88. Alexander Schmidt-Richberg, René Werner, Heinz Handels, Jan Ehrhardt, Estimation of slipping organ motion by registration with direction-dependent regularization, Medical Image Analysis 2012;16(1):150–159.

89. Adam L. Schwartz, Elijah Polak, Consistent approximations for optimal control problems based on Runge–Kutta integration, SIAM Journal on Control and Optimization 1996;34(4):1235–1269.

90. Nikhil Singh, Jacob Hinkle, Sarang C. Joshi, P. Thomas Fletcher, A vector momenta formulation of diffeomorphisms for improved geodesic regression and atlas constructuion, 2013 IEEE 10th International Symposium on Biomedical Imaging: From Nano to Macro. San Francisco. 2013:1219–1222.

91. Aristeidis Sotiras, Christos Davatzikos, Nikos Paragios, Deformable medical image registration: a survey, IEEE Transactions on Medical Imaging 2013;32(7):1153–1190.

92. John C. Strikwerda, Finite Difference Schemes and Partial Differential Equations. 2nd edition Philadelphia: SIAM; 2004.

93. Xiaobo Tan, Almost symplectic Runge–Kutta schemes for Hamiltonian systems, Journal of Computational Physics 2005;203(1):250–273.

94. Mingzhen Tan, Anqi Qiu, Large deformation multiresolution diffeomorphic metric mapping for multiresolution cortical surfaces: a coarse-to-fine approach, IEEE Transactions on Image Processing 2016;25(9):4061–4074.

95. Nicholas J. Tustison, Tessa S. Cook, Gang Song, James C. Gee, Pulmonary kinematics from image data: a review, Academic Radiology 2011;18(4):402–417.

96. Carole J. Twining, Stephen Marsland, Constructing diffeomorphic representations for the groupwise analysis of nonrigid registrations of medical images, IEEE Transactions on Medical Imaging 2004;23:1006–1020.

97. Petra A. van den Elsen, Evert-Jan D. Pol, Max A. Viergever, Medical image matching – a review with classification, IEEE Engineering in Medicine and Biology 1993;12(1):26–39.

98. 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 2012;97(2):229–241.

99. Valery Vishnevskiy, Tobias Gass, Gábor Székely, Christine Tanner, Orcun Goksel, Isotropic total variation regularization of displacements in parametric image registration, IEEE Transactions on Medical Imaging 2017;36(2):385–395.

100. Joachim Weickert, Andrés Bruhn, Thomas Brox, Nils Papenberg, A survey on variational optic flow methods for small displacements, O. Scherzer, ed. Mathematical Models for Registration and Applications to Medical Imaging. Mathematics in Industry. Berlin: Springer; 2006;vol. 10.

101. Benedikt Wirth, Leah Bar, Martin Rumpf, Guillermo Sapiro, A continuum mechanical approach to geodesics in shape space, International Journal of Computer Vision 2011;93(3):293–318.

102. Xiao Yang, Roland Kwitt, Marc Niethammer, Fast predictive image registration, International Workshop on Deep Learning and Data Labeling for Medical Applications – LABELS/DLMIA 2016. Springer International Publishing; 2016:48–57.

103. Xiao Yang, Roland Kwitt, Martin Styner, Marc Niethammer Quicksilver, Fast predictive image registration – a deep learning approach, NeuroImage 2017;158:378–396.

104. Laurent Younes, Shapes and Diffeomorphisms. Berlin/Heidelberg, Germany: Springer; 2010.

105. Laurent Younes, Felipe Arrate, Michael I. Miller, Evolutions equations in computational anatomy, NeuroImage 2009;45(1 Suppl 1):S40–S50.

106. Christopher Zach, Thomas Pock, Horst Bischof, A duality based approach for realtime TV-l 1 optical flow, Fred A. Hamprecht, Christoph Schnörr, Bernd Jähne, eds. Joint Pattern Recognition Symposium – DAGM 2007: Pattern Recognition. LNCS. Berlin, Heidelberg: Springer; 2007;vol. 4713:214–223.

107. Miaomiao Zhang, P. Thomas Fletcher, Finite-dimensional lie algebras for fast diffeomorphic image registration, Information Processing in Medical Imaging – IPMI 2015. 2015:249–260.

108. William P. Ziemer, Weakly Differentiable Functions: Sobolev Spaces and Functions of Bounded Variation. 1st edition New York, NY, USA: Springer Science+Business Media; 1989.

109. Barbara Zitová, Jan Flusser, Image registration methods: a survey, Image and Vision Computing 2003;21(11):977–1000.


1  “Automatic differentiation is heavily used in modern approaches for deep learning to avoid manual computations of adjoint equations altogether.”

2  “Although the spatial dimension is arbitrary, care has to be taken to ensure that the regularization is strong enough to ensure diffeomorphic results.”

3  “Whereas it is possible to solve the transport equation directly for the images (see Section 13.3.1), it is numerically much easier to solve for ϕ and τ as they will be spatially smooth; see Section 13.3.2.”

4  “Note that (LL)1Image amounts to a smoothing operation, and hence the Hilbert gradient is a spatially smoothed gradient.”

5  “Available at https://github.com/C4IR/FAIR.m.”

6  http://www.dir-lab.com/Results.html.”

7  http://www.deformetrica.org.”

8  https://mricloud.org/.”

9  http://www.bioeng.nus.edu.sg/cfa/brainmapping.html.”

10  https://github.com/fshapes/fshapesTk/.”

11  https://github.com/C4IR/FAIR.m/tree/master/add-ons/LagLDDMM.”

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

You can't read the all page of ebook, please click here login for view all page.
Reset
3.19.56.45