14

Spatially adaptive metrics for diffeomorphic image matching in LDDMM

Laurent Rissera,cFrançois-Xavier Vialardb,c    aInstitut de Mathématiques de Toulouse, CNRS, Université de Toulouse, UMR CNRS 5219, Toulouse, France
bLaboratoire d'informatique Gaspard Monge, Université Paris-Est Marne-la-Vallée, UMR CNRS 8049, Champs sur Marne, France
cBoth authors equally contributed to the chapter.

Abstract

Registering two medical images consists in computing a mapping between the organs of interest they contain. Although this mapping is dense in space, it can only be accurately estimated based on significant intensity variations in the images, which is a sparse information. Using deformation regularization properties that are physiologically meaningful is then one of the keys to estimate pertinent mappings. In the LDDMM framework these regularization properties are directly related to the right-invariant metric which controls the optimal deformation. In this chapter we then present different methodologies related to this degree of freedom. After briefly introducing the LDDMM framework, we present a simple strategy to regularize the mappings at different scales and a more advanced technique to make it possible to estimate a sliding motion at predefined locations. We then propose to switch the paradigm of right-invariant metrics to left-invariant ones, so that spatially adaptive metrics can be used in LDDMM. In the last part, we review different attempts to optimize these spatially adaptive metrics and propose a new evolution of LDDMM that incorporates spatially adaptive metrics.

Keywords

Large deformation diffeomorphic metric mapping (LDDMM); spatially adaptive metric; multiscale kernel; sliding motion constraints; left-invariant diffeomorphic metric (LIDM); semidirect product of groups

14.1 Introduction to LDDMM

14.1.1 Problem definition

The construction of the large deformation diffeomorphic metric mapping (LDDMM) framework is based on a variational setting and the choice of a Riemannian metric. Its goal is to estimate optimal smooth and invertible maps (diffeomorphisms) of the ambient space that represent a mapping between the points of a source image ISImage and those of a target image ITImage [9,6], see also Chapter 4. This diffeomorphic image registration formalism is particularly adapted to the registration of most 3D medical images, where the hypothesis that organ deformations are smooth is reasonable, and the topology of the represented organs is preserved. Note that this second property is mainly due to the fact that there is no occlusion or out-of-slice motion in such images. Image registration thus takes the form of an infinite-dimensional optimal control problem: Minimize the cost functional

J(ξ)=1210ξ(t)2Vdt+S(ISφ1)

Image(14.1)

under the constraints

tφ(t,x)=ξ(t,φ(t,x)),

Image(14.2)

φ(0,x)=x  xD.

Image(14.3)

The functional SImage represents the similarity measure between the registered images. For grey level images acquired using the same modality (e.g. a pair of MR images), the standard similarity metric is the so-called sum of squared differences between the deformed source image ISImage and the target image ITImage, that is, ISφ1IT2L2Image, both defined on a domain of the Euclidean space denoted by D. As summarized in Fig. 14.1, constraints (14.2) encode the trajectory of the points xDImage: At time t=0Image a point x of the source image ISImage is naturally at location φ(0,x)=xImage. Then its motion at times t[0,1]Image is defined by the integration of the time-dependent velocity field ξ(t,x)Image. The transformed location of x at time t=1Image is finally φ(1,x)Image and corresponds to the mapping of x in the target image ITImage.

Image
Figure 14.1 Transportation of the point x ∈ D through the diffeomorphism φ(t,x), where D is the domain of the source image IS. The point φ(1,x) is the mapping of x in the target image IT.

14.1.2 Properties

In Eq. (14.1), V is a Hilbert space of vector fields on a Euclidean domain DRdImage. A key technical assumption, which ensures that the computed maps are diffeomorphisms up to the numerical scheme accuracy, is that the inclusion map VW1,(D,Rd)Image, that is, the space of vector fields which are Lipschitz continuous, is continuous. The norm on V controls the W1,Image norm, and we call such a space V an admissible space of vector fields. In particular, these spaces are included in the family of reproducing kernel Hilbert spaces (RKHS) [3] since pointwise evaluations are a continuous linear map, which implies that such spaces are completely defined by their kernel. The kernel, denoted by kImage in this chapter, is a function from the product space D×DImage into RdImage that automatically satisfies the technical assumption mentioned if it is sufficiently smooth. Last, we denote by K:VVImage the isomorphism between VImage, the dual of V, and V.

Note that the contributions presented in this chapter build on the flexibility of the RKHS construction not only to accurately match the structure boundaries in the deformed source image ISφ1Image and the target image ITImage, but also to estimate physiologically plausible final deformation maps φ.

The direct consequence of the admissible hypothesis on V is that the flow of a time-dependent vector field in L2([0,1],V)Image is well defined; see [29, Appendix C]. Then the set of flows at time 1 defines a group of diffeomorphisms denoted by GVImage; that is, denoting

Fl1(ξ)=φ(1) where φ solves (14.2),

Image(14.4)

define

GVdef.={φ(1):ξL2([0,1],V) s.t. Fl1(ξ)},

Image(14.5)

which has been introduced by Trouvé [25]. On this group, Trouvé defines the metric

dist(ψ1,ψ0)2=inf{10ξ2Vdt:ξL2([0,1],V) s.t. ψ1=Fl1(ξ)ψ0},

Image(14.6)

under which he proves that GVImage is complete. In full generality very few mathematical properties of this group are known. However, in particular situations, such as where the space V is the space of Sobolev vector fields that satisfy the continuous injection property, then the group is also an infinite-dimensional Riemannian manifold (see [8]). Since the distance (14.6) is right-invariant, it is important to emphasize that for all ψ1,ψ2,ψ3GVImage, we have the following property:

dist(ψ1ψ3,ψ0ψ3)=dist(ψ1,ψ0).

Image(14.7)

Instead of formulating the variational problem on the group of diffeomorphisms GVImage, it is often possible to rewrite the optimization problem on the space of images. More precisely, the minimization problem is taken to be

J(ξ)=10ξ(t)2Vdt+S(I(1))

Image(14.8)

under the constraints

tI(t,x)+I(t,x),ξ(t,x)=0,I(0,x)=IS(x)  xD.

Image

For S(I(1))=ϵ1I(1)IT2L2Image, the sum of squared differences and σ is a positive parameter, and using the Lagrange multiplier rule, we can write the gradient of this functional as

J(ξ)=2ξ(t)+K(I(t)P(t)),

Image(14.9)

where P(t)Image satisfies the continuity equation (the notation div stands for the divergence operator)

tP(t,x)+div(Pξ)=0

Image(14.10)

and the initial condition P(1)=2ϵ1(I(1)IT)Image. Therefore Eq. (14.10) has to be solved backward in time from t=1Image to t=0Image. Alternatively, using the solutions of continuity and advection equations in terms of the flow map, it is possible to rewrite the gradient as in (line 12 of) Algorithm 14.1, which will be discussed in Section 14.1.3.

Image
Algorithm 14.1 Interpreted LDDMM algorithm of [6] to register the images IS and IT.

More generally, it is possible to formulate the equivalent variational problem in the case where shapes are deformed rather than images, as, for instance, when registering point clouds or surfaces. Under mild conditions, it is also possible to prove that this approach induces a Riemannian metric on the orbit of the group action in some finite-dimensional cases (see also Chapter 4). We denote by Q the space of objects or shapes on which the deformation group is acting. When Q is an infinite-dimensional Riemannian manifold, the geometric picture is more complicated [5].

We now go back to the optimization problem. By first-order optimality and using again the notation JImage for the corresponding but different functional, a solution to formulation (14.1) can be written as

J(P0)=12DK(P0I0)(x)P0(x)I0(x)dx+S(I(1))

Image(14.11)

under the constraints

{tI+I,ξ=0,tP+div(Pξ)=0,ξ+K(P0I0)(x)=0,

Image(14.12)

with initial conditions P(t=0)=P0Image and I(t=0)=I0Image. The function P0:DRImage is sometimes called the momentum or scalar momentum, and we denoted

K(P0I0)(x)=Dk(x,y)P0(y)I0(y)dy;

Image(14.13)

in particular, this quantity can be reformulated as an L2Image norm of the quantity P0I0Image for the square root of the kernel k. Moreover, system (14.12) encodes the fact that the evolution of I(t)Image is geodesic in the LDDMM setting; see [28]. Therefore this formulation transforms the problem of optimizing on the time-dependent d-dimensional vector field ξ (sometimes called path-based optimization) into optimizing on a function P0Image defined on the domain D (sometimes called shooting method). At optimality the following fixed point equation has to be satisfied:

P(1)+IS(I(1))=0,

Image(14.14)

which can be used in practice for some optimization schemes [1].

14.1.3 Implementation

We now discuss different ideas related to the implementation of the LDDMM framework to register a source image ISImage onto a target image ITImage. Our discussion specifically builds on [6], where a practical algorithm of LDDMM for image matching was given. We then give hereafter an overview of this algorithm plus different numerical strategies we used to make it work efficiently. Note that our implementation of [6] and the extensions we developed are freely available on sourceforge.1

When registering two images, we have first to define a discrete domain on which φ(t,x)Image and v(t,x)Image are computed, where φ(t,x)Image is the mapping of x at time t through φ, and v(t,x)Image is the velocity field integrated in time to compute φ. A natural choice is to use a spatial grid defined by the pixel/voxel coordinates of ISImage. We denote by ˆDImage this discrete domain and recall that D is the dense image domain. Linear interpolation is recommended to estimate φ and v at point locations in D and outside ˆDImage. Note that ISImage and ITImage may have a different resolution or may not be aligned. We suppose here that they have already been aligned by a rigid deformation and that the final deformation φ(1,x)Image is composed with this deformation to reach the pixel/voxel coordinates of ITImage. In our implementation we also used an uniformly sampled grid to discretize t. The grid time step should also be sufficiently small to avoid generating noninvertible deformations when temporally integrating v. About 10 time steps are enough in most applications, but more time steps may be necessary when sharp deformations are computed [18].

We use the following notations to describe the registration algorithm: tθ,θ{1,,Θ}Image, are the discrete time points. For each tθImage, several vector fields are required to encode useful deformations based on the diffeomorphism φ: ϕtj,ti(x)Image first transports xˆDImage from time tiImage to time tjImage through φ. The images IS,tθImage and IT,tθImage also correspond to ISImage and ITImage transported at time tθImage using ϕ0,tθImage and ϕ1,tθImage respectively. Image registration is then a gradient descent algorithm where v is optimized with respect to ISImage, ITImage, and the smoothing kernel K as shown Algorithm 14.1.

We can first remark that the mappings ϕ1,tθ(x)Image and ϕtθ,1(x)Image are precomputed in the for loop at lines 5–7 of Algorithm 14.1. These mappings are indeed computed once for all and stored by using an Euler method from time tΘImage to time t0Image, whereas the mappings ϕ0,tθ(x)Image can be computed from time t0Image to time tΘImage in the energy gradients estimation loop.

We also strongly recommend to compute IS,tθ(x)Image and IT,tθ(x)Image by resampling ISImage and ITImage using ϕ0,tθ(x)Image and ϕ1,tθ(x)Image, respectively. An alternative would be to compute iteratively the deformed images time point after time point, for example, to compute IS,tθ(x)Image using IS,tθ1(x)Image and v(tθ1,x)Image. This strategy would be far less memory consuming than the one we use, but it would also numerically diffuse the image intensities due to the iterative resamplings.

Another remark is that a simple and very efficient technique can be used to speed up the convergence of this registration algorithm. So-called momentum methods [15] are widely known in machine learning to speed up the convergence of gradient descent algorithms in high dimension. At each iteration it simply consists in updating the optimized variables with a linear combination of the current gradients and the previous update. Our (unpublished) experiences have shown that this technique is particularly efficient in image registration where, at a given iteration, the mapping can be already accurate in some regions and inaccurate in other regions.

The most important point to discuss to make the practical use of the LDDMM algorithm clear is that it depends on two parameters ϵ1Image and ϵ2Image, respectively the weight in front of the sum of squared differences (see discussion for Eqs. (14.9) and (14.10)) and the step length of the gradient descent. In practice ϵ1Image should be sufficiently large so that u(tθ,x)Image has much more influence than v(tθ,x)Image in line 14 of Algorithm 14.1. The vector field u(tθ,x)Image indeed pushes one image to the other and can be interpreted as a force field. The influence of v(tθ,x)Image should then be small but not negligible. This term is specific to LDDMM in the medical image registration community and indeed ensures the temporal consistency of the time-dependent deformations. The choice of ϵ2Image is more conventional in a gradient descent algorithm and controls the convergence speed. An empirical technique to tune it was given in [18]: At the first algorithm iteration we compute vmax=maxtθ,x||vE(tθ,x)||2Image. We then set ϵ2Image equal to 0.5/vmaxImage, where 0.5 is in pixels/voxels, so that the maximum update at the first iteration is half a pixel/voxel. The updates have then a reasonable and automatically controlled amplitude.

14.2 Sum of kernels and semidirect product of groups

14.2.1 Introduction

Hereafter we discuss the work presented in [19,7]. In most applications a Gaussian kernel is used to smooth the deformations. A kernel corresponding to the differential operator (Id+ηΔ)kImage for a well-chosen k with a single parameter η may also be used. The Gaussian width σ is commonly chosen to obtain a good matching accuracy. This means that small values, close to the image resolution, are used for σ. We can then wonder what is the effect of this parameter on the structure of the deformation. In [19] we have illustrated the influence of σ on the mapping obtained between two images of the grey matter acquired on a preterm baby at about 36 and 43 weeks of gestational age, as summarized in Fig. 14.2. Let us focus on the (B-top) subfigure of Fig. 14.2. The yellow (ligth gray in print version) isoline represents the cortex boundary in a 2D region of interest (ROI) out of a 3D segmented image S36Image, and the ROI is located in the red square of the (A-bottom) subfigure. The grey levels of the same (B-top) subfigure also represent the segmented cortex in the same preterm baby but 7 weeks later. It is obvious that the brain became globally larger as the brain and the skull strongly grow at this age. The shapes should be almost translated at the scale of this ROI to capture the amplitude of the deformation. It is important that existing cortex folds also became deeper and new folds appeared, which is normal during brain maturation because the cortex growth is faster than the skull growth. Capturing the folding process requires registering the images at a scale close to the image resolution here. To conclude, the registration of these images requires at a same time a large σ and a small σ. If only a small σ is used, then optimal path (and the optimization process) will lead to physiologically implausible deformations. This is obvious in Fig. 14.2(C), where the brown isoline represents the boundaries of the deformed voxels after registration. In this example the volume of some voxels becomes huge, and other voxels almost disappear. If this deformation was the real one, then the studied brain would have a strongly heterogeneous development in space, which is clearly not realistic. On the contrary, if only a large σ was used, then the optimal path would not capture fine deformations, as shown in Fig. 14.2(D). This justifies the use of multiscale kernels to establish geodesics between such follow-up medical images.

Image
Figure 14.2 (A) Grey matter extraction of the 3D MR image I36 (top) and resulting segmentation S36 (bottom). The red square indicates the 2D region of interest shown in (B,C,D). (B) The yellow (ligth gray in print version) and blue (dark gray in print version) isolines represent the cortical surface of S36 and S43, respectively. The grey levels are the segmented cortex of S43. (C,D) The brown isolines represent deformed cortical surfaces of S36 after LDDMM registration on S43 with σ = 1.5 and σ = 20, respectively. The grids represent the estimated dense deformations.

14.2.2 Multiscale kernels

As for the LDDMM model, we recall that the kernel spatially interpolates the rest of the information (i.e., the momentum) to drive the motion of the points where there is no gradient information, for example, in flat image regions. Therefore it is natural to introduce a sum of kernels to fill in the missing information while preserving the physiologically realistic matchings. Therefore more plausible deformations are obtained since the correlation of the motions of the points is higher.

In practice this method works really well, and the mathematical insight for its efficiency is probably the variational interpretation of the sum of kernel, explained hereafter. For simplicity, we only treat the case of a finite set of RKHS Hilbert spaces HiImage with kernels kiImage and Riesz isomorphisms KiImage between HiImage and HiImage for i=1,,nImage. For every i, HiImage is a subspace of the space of C1Image vector fields on the domain D. Denoting H=H1++HnImage, the space of all functions of the form v1++vnImage with viHiImage, the norm is defined by

v2H=inf{ni=1vi2Hi|ni=1vi=v}.

Image(14.15)

The minimum is achieved for a unique n-tuple of vector fields, and the space H endowed with the norm defined by (14.15) is complete. The result is the following: there exists a unique element pni=1HiImage for which we have vi=KipImage and

v=ni=1Kip,

Image(14.16)

the family (vi)i=1,,nImage realizing the (unique) infimum of the variational problem (14.15). Formula (14.15) induces a scalar product on H, which makes H an RKHS, and its associated kernel is k:=ni=1kiImage, where kiImage denotes the kernel of the space HiImage. This property was written in [3] and is standard in convex analysis. Indeed, note that this property is the particular case of an elementary result in convex analysis, at least in finite dimensions: the convex conjugate of an infimal convolution is equal to the sum of the convex conjugates [20].

Another phenomenon observed in practice is that a better quality of matching is obtained with a sum of kernels than with a single kernel of small width. Although we have no quantitative argument in this direction, we strongly believe that this is due to the convergence of the gradient descent algorithm to local minima. In standard image registration, coarse to fine techniques [11] are ubiquitous. They consist in first registering two images with a strong regularization level and then iteratively decreasing the regularization level when the algorithm has converged at the current scale. At each considered scale, gradient descent-based registration is then likely to be performed in a stable orbit with respect to the compared shape scale. In LDDMM, using the sum of kernels at different scales instead of small scales only may then have a similar effect from an optimization point of view.

Based on the practical implementation of LDDMM for images of [6] and summarized Algorithm 14.1, we have proposed to use smoothing kernels constructed as the sum of several Gaussian kernels [19]. These kernels, denoted by MK, that are the weighted sums of N Gaussian kernels KσnImage, each of them being parameterized by its standard deviation σnImage:

MK(x)=Nn=1anKσn(x)=Nn=1an(2π)3/2|Σn|1/2exp(12xTΣ1nx),

Image(14.17)

where ΣnImage and anImage are respectively the covariance matrix and the weight of the nth Gaussian function. Each ΣnImage is only defined by a characteristic scale σnImage: Σn=σnIdRdImage. Once this kernel is defined, the registration algorithm is the same as in Algorithm 14.1.

A tricky aspect of this kernel construction for practical applications is however the tuning of their weights anImage. Although the choice of the σnImage has a rather intuitive influence on the optimal deformations, the tuning of the anImage strongly depends on the representation and the spatial organization of the registered shapes at the scales σnImage, n[1,N]Image. As described in [19], it depends on: (1) Representation and spatial organization of the structures: A same shape can be encoded in various ways. For instance, it can be a binary or a grey-level image. This representation has first a nonlinear influence on the similarity metric (the sum of squared differences in LDDMM) forces (unsmoothed gradients) as shown line 12 of Algorithm 14.1. The choice of optimal parameters anImage is even more complicated to do as the spatial relation between the shape structures should also be taken into account when smoothing the forces (line 13 of Algorithm 14.1). (2) Prior knowledge: Prior knowledge about the amplitude of the structures displacement at each scale σnImage may be incorporated in anImage.

In [17] we have proposed to semiautomatically tune the anImage as follows:

an=an/g(Kσn,IS,IT),

Image

where g(Kσn,IS,IT)Image represents the typical amplitude of the forces when registering ISImage to ITImage at a scale σnImage. This amplitude is related to (1) and cannot therefore be computed analytically. An empirical technique to tune it is the following: for each KσnImage, the value of g(Kσn,IS,IT)Image can be estimated by observing the maximum update of the velocity field v in a preiteration of registration of ISImage on ITImage using only the kernel KσnImage with an=1Image. The apparent weights anImage, n[1,N]Image, provide an intuitive control of the amplitude of the displacements and are related to (2). To deform the largest features of ISImage and ITImage with a similar amplitude at each scale σnImage, the user should tune all the apparent weights anImage with the same value. Typical results we obtained in [19] on the example of Fig. 14.2 are shown in Fig. 14.3. They make clear the fact that multiscale kernels with automatically tuned anImage following our method efficiently solved the problem we initially described.

Image
Figure 14.3 Registration results obtained on the example of Fig. 14.2 using multiscale kernels. MKN stands for the sum of N kernels. Here MK4 and MK7 were automatically designed with apparent weights aiImage having the same value.

14.2.3 Distinguishing the deformations at different scales

It is interesting to remark that the influence of each subkernel of the multiscale kernels we defined can be measured. Distinguishing scale-dependent deformations is indeed useful for further statistical analysis. A first attempt to characterize this influence has been presented in [19] and was strongly developed in [7]. The main contribution of [7] was to formulate the multiscale LDDMM registration with a semidirect product. Registering ISImage on ITImage is then done by minimizing a registration energy ENImage with respect to the N-tuple (v1,,vN)Image where each time-dependent velocity field vnImage is associated with scale-dependent deformations. Then the minimized energy is

EN(v1,,vN)=12Nn=110vn(t)2Hndt+S(IS,IT,φ),

Image(14.18)

where the space HnImage corresponds to the kernel KσnImage, and φn(t)Image is defined by

tφk(t)=(vk(t)+(IdAdφk(t))nn=k+1vn(t))φk(t).

Image(14.19)

Here AdφvImage also denotes the adjoint action of the group of diffeomorphisms on the Lie algebra of vector fields:

Adφv(x)=(Dφ.v)φ1(x)=Dφ1(x)φ.v(φ1(x)).

Image(14.20)

These equations then allow us to quantify scale-dependent deformations φnImage in the whole deformation φ. We can also sum over all scales to form v(t)=nk=1vk(t)Image and compute the flow φ(t)Image of v(t)Image. A simple calculation finally shows that

φ(t)=φ1(t)φn(t).

Image(14.21)

Results and algorithmic description of the solution for 3D images were given [7]. An illustration of this paper, where the deformations between two brain images were split into 7 scales, is given Fig. 14.4. Note also that in [21] the authors build on these ideas of multiscale kernels and incorporate some sparsity prior. On the other hand, we can extend the space of kernels as done in [24], in which the authors construct multiscale kernels based on wavelet frames and with an apparent improvement of the registration results, although the corresponding group structure interpretation is possibly lost.

Image
Figure 14.4 Representation of scale-dependent deformations φk out of a deformation φ obtained between two brain images using [7]. The colors represent the amplitude of the scale-dependent deformations at the brain surface.

14.3 Sliding motion constraints

14.3.1 Introduction

Now we focus on how to model sliding constraints in the LDDMM formalism. Such constraints are observed, for example, at the lung boundaries, as emphasized in Fig. 14.5. In [18] we have developed a smoothing strategy to solve this problem by using Algorithm 14.1 (of [6]) with specific smoothing properties. The central idea is to predefine different regions of interest ΩkImage in the domain Ω of the registered images at the boundary of which discontinuous deformations will be potentially estimated. Note first that these regions of interest are fixed so the source image ISImage and the target image ITImage should be aligned at the boundaries of the regions ΩkImage, which is done in Algorithm 14.1 by using a standard registration strategy with large amount of smoothing. This domain decomposition is illustrated Fig. 14.6.

Image
Figure 14.5 Illustration of the sliding motion at the lung boundary in the coronal view of two CT volumes acquired on the same subject. The motion of the emphasized vessel with respect to ribs 1 and 2 clearly demonstrate the sliding motion at the lung boundary. Images out of the EMPIRE10 challenge [13].
Image
Figure 14.6 (Left) Subdivision of the registration domain Ω into Ω1 (inside the lung) and Ω2. Subdomain boundaries are represented by ∂Ω1 and ∂Ω2. (Right) Velocity field v which can be obtained in Ω after enforcing sliding conditions in the neighborhoods of ∂Ω1 and ∂Ω2.

14.3.2 Methodology

Instead of considering a reproducing kernel Hilbert apace (RKHS) V embedded in C1(Ω,Rn)Image or W1,Image as in the previous section, here we use N RKHS of vector fields VkC1(Ωk,[0,1])Image, which can capture sliding motion, that is, with an orthogonal component to the boundary that vanishes at any point of ΩkImage. The set of admissible vector fields is therefore defined by V:=Nk=1VkImage, the direct sum of the Hilbert spaces (Vk)k1,NImage. In particular, the norm on V of a vector field vtImage is given by

vt2V=Nk=1vkt2Vk,

Image(14.22)

where vktImage is the restriction of vtImage to ΩkImage. The flow of any vL2([0,1],V)Image is then well defined although the resulting deformations are piecewise diffeomorphic and not diffeomorphic. As a consequence, the deformation is a diffeomorphism on each subdomain and allows for sliding motion along the boundaries.

Now that an admissible RKHS is defined, let us focus on the strategy we used to mimic the Gaussian smoothing of the updates u (see line 13 in Algorithm 14.1) with the desired properties. We use the heat equation to smooth u in each region ΩkImage: u/τ=ΔuImage, where τ[0,Γ]Image is a virtual diffusion time. We denote by ΩkImage the boundaries of ΩkImage. Here Γ controls the amount of smoothing. To prevent from information exchange between the different regions, Neumann boundary conditions are additionally modeled at each point x of ΩkImage: u(x)n(x)=0Image, where n(x)Image is normal to ΩkImage at x. Independent Gaussian based convolution in each region ΩkImage, would have been a quicker alternative in terms of computations but would not take into account the intrinsic region geometry. Then, to ensure that the orthogonal component to the boundary vanishes at any point of ΩkImage, we use a projection strategy of the updates before and after smoothing so that they respect this constraint.

To do so, we consider the vector field T so that for each point xΩImage, x+T(x)Image is the nearest boundary between two subdomains in a limited neighborhood around the boundaries ΩkImage. For the registration of pulmonary images, we empirically use a neighborhood of about γ=20Image millimeters. Consider a velocity field w defined on Ω. We use T to enforce the sliding conditions around ΩkImage by reducing the contributions of w(x)Image in the direction of T(x)Image when ||T(x)||L2<γImage:

w(x)=w(x)α(x)T(x)<w(x),T(x)>L2||T(x)||2L2,

Image(14.23)

where the weight α(x)Image equals (γ||T(x)||)2/γImage. For numerical stability, w(x)Image is set to 0 if ||T(x)||2L2=0Image. The registration algorithm is then the same as Algorithm 14.1 except line 13, where u is first projected using Eq. (14.23), then smoothed using the heat (diffusion) equation, and then projected again using Eq. (14.23).

14.3.3 Results and discussion

Results shown in [18] make clear the impact of this strategy compared with standard smoothing kernels. Fig. 14.7 shows the impact of such a piecewise diffeomorphic kernel when registering lung image where a sliding motion is clearly required at the lung boundaries. Note that to make this strategy tractable on large medical images (as in Fig. 14.7), we also coded it in the LogDemons formalism of [26]. This formalism is indeed less memory consuming than LDDMM, as the diffeomorphisms are encoded in stationary velocity fields and not time-dependent ones as in LDDMM. Computational burden would be too high in the LDDMM framework. However, both LogDemons and LDDMM with the proposed sliding motion estimation strategy led to similar results on smaller images.

Image
Figure 14.7 Deformation magnitude and deformed grids obtained when registering I1 to I5 using LogDemons using sliding motion modeling (S-LogD) or not (LogD MR). Color bar is from 0 to 5 cm. The black dots represent the thoracic cage boundary.

14.4 Left-invariant metrics

14.4.1 Introduction

In this section, we describe the results obtained in [22,23]. A natural extension of the sum of kernels consists in having a kernel that may depend on the location. However, the right-invariant point of view is meant for a homogeneous material whose properties are translation invariant although this is not required by the theory. In practice the kernel used in diffeomorphic methods has always been chosen to be translationally invariant and isotropic. In LDDMM spatially adaptive or nonisotropic (“direction-dependent”) kernels have no obvious interpretation, because the norm is defined in Eulerian coordinates, so that as t varies during the deformation, a fixed point in the source image moves through space, and conversely, a fixed point in space will correspond to different points in the source image. Similarly, the directions in a direction-dependent kernel are defined with respect to Eulerian coordinates, not the coordinates of the moving source image. Nonetheless, spatially adaptive kernels are potentially of great interest in medical applications if they can be made to represent spatially variable (or nonisotropic) deformability of tissue. This is indeed already done in [16] to model sliding conditions between the lungs and the ribs. In this section we present a slightly different registration framework than LDDMM, which naturally supports the use of spatially adaptive kernels.

14.4.2 Methodology

The proposed framework is based on a left-invariant metric on the group of deformations, where its name LIDM (left-invariant diffeomorphic metrics) comes from. Left-invariance means that this metric satisfies, in a smooth setting, the following property: For all elements ψ1Image, ψ2Image, ψ3Image in the group,

dist(ψ3ψ1,ψ3ψ0)=dist(ψ1,ψ0),

Image(14.24)

which is in contrast with formula (14.7). In fact, such a left-invariant metric is based on a choice of norm in the body (Lagrangian) coordinates of the source image. This means that instead of the V norm applied to the spatial velocity defined by (14.12), it is applied to the convective velocity v(t)Image implicitly defined by

tφ(t)=dφ(t)v(t),

Image(14.25)

where dφ(t)Image is the spatial derivative of φ(t)Image.

It is well known that left- and right-invariant metrics are isometric by the inverse map. Therefore, as expected, we obtain the following result.

Corollary 14.1

[Equivalence between LIDM and LDDMM] Consider the problem of minimizing

JIT(φ)=1210v(t)2Vdt+E(ISφ11,IT)

Image(14.26)

for φ0=IdΩImage, and with constraint either

tφt=dφtvt(LIDM constraint)

Image(14.27)

or

tφt=vtφt(LDDMM constraint).

Image(14.28)

Then

  1. 1.  The optimal endpoint φ1Image is the same with either constraint.
  2. 2.  If ϕtImage minimizes JImage in LIDM, then φt:=ϕ1t1ϕ1Image minimizes JImage in LDDMM.
  3. 3.  If φtImage minimizes JImage in LDDMM, then ϕt:=φ1φ1t1Image minimizes JImage in LIDM.

Although not surprising, this result gives a mathematical interpretation to the use of spatially adaptive kernels that can be defined using a variational approach. Let us consider, as in the previous section, a family of RKHS (Hi)i=1,,nImage and an operator A:H1HnH=H1++HnImage. On the space H we introduce

vH2=inf{i=1nviHi2|A(v1,,vn)=v}.

Image(14.29)

Using again duality and under mild assumptions, the kernel associated with H is Hpi=1nKi(Ap)iHImage.

Let us give an instance of it in the context of biomedical images. Suppose we have a partition of unity ((χi)i=1,,nImage) of the domain of interest (a manual segmentation of the biological shape) where we have some knowledge of the deformability properties of the shape modeled by the kernel KiImage. The map A can be chosen as i=1nχiviImage, and the corresponding kernel is

K=i=1nχiKiχi.

Image(14.30)

This kernel satisfies the embedding condition under mild conditions on the element of the partition of unity (χi)i=1,,nImage.

14.4.3 Results and discussion

The experiment of Fig. 14.8 is adapted from [22], and it shows registration results for a synthetic example, which includes features at different scales. LIDM shows the results of the registration using a kernel defined accordingly to the partition of unity shown in the figure (two Gaussian kernels with a large σ on the white and a small σ on the black). As expected, it performs better than the sum of kernel because it captures the small scale deformations.

Image
Figure 14.8 Results of image registration tests on a synthetic example.

The use of spatially adaptive kernels provably improves the registration results on real data. However, the shortcoming of this approach is that the kernel does not evolve with the deformed shape. For small/medium deformations, it may not be a problem, but it cannot be applied in the case of large deformations. In such a case the kernel has to depend on the shape itself. Such approaches have actually been developed in [30,4,2], where the operator A depends on the shape itself, but developing models for images associated with an efficient implementation remains open.

14.5 Open directions

14.5.1 Learning the metric

There are two different point of views that can be developed to improve on the results of LDDMM: On the one hand, incorporating mechanical or biological constraints to reproduce more realistic results is what we have described so far in this chapter. On the other hand, it is also natural to learn the metric parameters using data driven methods if no mechanical model is well established for the data of interest. Instead of having a partition of unity drawn by the user, it is also natural to ask whether the smoothing kernel can be learned from data. We summarize here an approach proposed in [27] to learn the parameters from a given population and a given template.

Building on the LIDM model, we aim at designing a set of kernels expressing spatially adaptive metrics. We use (symmetric) positive definite matrices M as a parameterization of this set of kernels. To ensure the smoothness of the deformations, any kernel of this set has to satisfy the constraint that the Hilbert space of vector fields is embedded in the Banach space of C1Image vector fields. To enforce this constraint, we propose the following parameterization:

K={KˆMKˆ|M SDP operator on L2(Rd,Rd)},

Image(14.31)

where KˆImage is a spatially homogeneous smoothing kernel (typically Gaussian). Now the variational model consists in minimizing the functional with a positive real β:

F(M)=β2dS++2(M,Id)+1Nn=1NminvJIn(v,M),

Image(14.32)

where M is symmetric. The first term is a regularizer of the kernel parameters so that the minimization problem is well posed. Here, it favors parameterizations of M close to the identity matrix, but other a priori correlation matrix could be used. The term dS++2(Id,M)Image can be chosen as the squared distance on the space of positive definite matrices given by log(M)2Image. Here again other choices of regularizations could have been used such as the log-determinant divergence. This model has been implemented in [27], where a simple method of dimension reduction was used since the matrix M is of size n2Image, where n is the number of voxels, and it gave promising results on the 40 subjects of the LONI Probabilistic Brain Atlas (LPBA40). An illustration of the matrix M we computed in this paper is given in Fig. 14.9.

Image
Figure 14.9 Values out of M learnt on 40 subjects of the LONI Probabilistic Brain Atlas (LPBA40). The values are represented at their corresponding location in the template image T. (DiagM): Values M(j,j) for j ∈ [1,…,N]. Color bar ranges from 1 (black) to 1.04 (white). (GridM): M(i,j) for a fixed i and j ∈ [1,…,L]. The red (mid gray in print version) point corresponds to i = j and has an intensity of 1.03. Color bar ranges from −0.05 (black) to 0.05 (white) for other points. Yellow (ligth gray in print version) curves represent the boundary between white and grey matter in T.

Another possible direction as done in [14] consists in learning the partition of unity (χi)iIImage with some smoothness constraints on the χiImage such as H1Image or TV. Moreover, since there is an interplay in the optimality equations between the gradient of the deformed image and the deformation, it is possible to introduce some information on the regularization of the partition so that it takes into account this interplay.

14.5.2 Other models

From a methodological point of view, the main weakness of the previously presented methods is probably the fact that the metric does not evolve with the shape in Lagrangian coordinates. To make the practical impact of this property clear, assume that a part of the shape follows a fixed volume transformation. In that case the models proposed are clearly unadapted as they cannot incorporate this property. This is why constrained diffeomorphic evolutions have been introduced in the literature [30,2]. Most of the time these constraints are incorporated in Lagrangian coordinates such as in [10]. However, fast computational methods for diffeomorphic image matching are designed in Eulerian coordinates; see, for instance, [31,12]. We propose an Eulerian-based PDE model, which can be seen as a mild modification of the LDDMM framework presented, incorporating the modeling assumption that the metric is naturally image dependent.

The standard formulation of shape-dependent metric registration is similar to the formulation in (14.1) when the norm on the vector field v depends on the current shape I. It is important that it is often possible to preserve the metric property in this type of modification. The manifold Q that we are going to consider consists in augmenting the template image I with a partition of unity (χi)iIImage. The definition of the action of φ in the group of diffeomorphisms is as follows:

{φI=def.Iφ1,φχi=def.χiφ1.

Image(14.33)

In other words, we consider the partition of unity as additional images that are advected by the flow. The variational problem is then the following:

min1201vH2dt+S(φI),

Image(14.34)

where the norm is as in Eq. (14.29), that is,

vH2=inf{i=1nviHi2|v(x)=i=1nχi(x)vi(x)xD},

Image(14.35)

and the flow is defined by

{tφ(t,x)=v(t,φ(t,x)),φ(0,x)=x.

Image(14.36)

Alternatively, functional (14.34) can be rewritten using a shape-dependent metric using the Lagrange multiplier method, and it is similar to the constrained evolutions proposed in [30]. The optimality equation can be written as

{I˙+I,v=0,P˙+(Pv)=0,χ˙i+χi,v=0i=1,,n,λ˙i+(λiv)=PI+λiχi,vii=1,,n,v=i=1nχiKiχi(PI+λiχi).

Image(14.37)

Note that the Lagrange multiplier associated with the partition evolves accordingly to the fourth equation in (14.37), which has a source term on its right-hand side that differs from the optimality equations (14.12).

References

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

2. S. Arguillère, M. Miller, L. Younes, Diffeomorphic surface registration with atrophy constraints, SIAM Journal on Imaging Sciences 2016;9(3):975–1003.

3. N. Aronszajn, Theory of reproducing kernels, Transactions of the American Mathematical Society 1950;68:337–404.

4. S. Arguillere, E. Trélat, A. Trouvé, L. Younes, Shape deformation analysis from the optimal control viewpoint, Journal de Mathématiques Pures et Appliquées 2015;104.

5. Sylvain Arguillere, Emmanuel Trélat, Alain Trouvé, Laurent Younes, Shape deformation analysis from the optimal control viewpoint, Journal de Mathématiques Pures et Appliquées 2015;104:139–178.

6. M. Faisal Beg, Michael I. Miller, Alain Trouvé, Laurent Younes, Computing large deformation metric mappings via geodesic flow of diffeomorphisms, International Journal of Computer Vision 2005;61:139–157.

7. Martins Bruveris, Laurent Risser, François-Xavier Vialard, Mixture of kernels and iterated semidirect product of diffeomorphisms groups, Multiscale Modeling & Simulation 2012;10(4):1344–1368.

8. Bruveris Martins, François-Xavier Vialard, On completeness of groups of diffeomorphisms, Journal of the European Mathematical Society 2017;19(5):1507–1544.

9. P. Dupuis, U. Grenander, M.I. Miller, Variational problems on flows of diffeomorphisms for image matching, Quarterly of Applied Mathematics 1998;56:587–600.

10. Barbara Gris, Stanley Durrleman, Alain Trouvé, A sub-Riemannian modular approach for diffeomorphic deformations, Frank Nielsen, Frédéric Barbaresco, eds. Geometric Science of Information. Cham: Springer International Publishing; 2015:39–47.

11. B.D. Lucas, T. Kanade, An iterative image registration technique with an application to stereo vision, Proceedings of the 7th International Joint Conference on Artificial Intelligence – Vol. 2, IJCAI'81. 1981:674–679.

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

13. K. Murphy, B. Van Ginneken, J.M. Reinhardt, S. Kabus, K. Ding, X. Deng, J.P.W. Pluim, et al., Evaluation of registration methods on thoracic CT: the EMPIRE10 challenge, IEEE Transactions on Medical Imaging 2011;30(10):1901–1920.

14. Marc Niethammer, François-Xavier Vialard, Roland Kwitt, Metric Learning for Image Registration. 2018.

15. D.E. Rumelhart, G.E. Hinton, R.J. Williams, Learning representations by back-propagating errors, Nature 1986;323(6088):533–536.

16. 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;17:182–193.

17. L. Risser, F.X. Vialard, M. Murgasova, D.D. Holm, D. Rueckert, Large diffeomorphic registration using fine and coarse strategies. application to the brain growth characterization, International Workshop on Biomedical Image Registration (WBIR). Lecture Notes in Computer Science. 2010;vol. 6204:186–197.

18. L. Risser, F.X. Vialard, n.H.Y. Baluwala, J.A. Schnabel, Piecewise-diffeomorphic image registration: application to the motion estimation between 3D CT lung images with sliding conditions, Medical Image Analysis 2012;17(2):182–193.

19. L. Risser, F.-X. Vialard, R. Wolz, M. Murgasova, D.D. Holm, D. Rueckert, Simultaneous multi-scale registration using large deformation diffeomorphic metric mapping, IEEE Transactions on Medical Imaging 2011;30(10):1746–1759.

20. R.T. Rockafellar, R.J-B. Wets, Variational Analysis, vol. 317. Springer Science & Business Media; 2009.

21. Stefan Sommer, Francois Lauze, Mads Nielsen, Xavier Pennec, Sparse multi-scale diffeomorphic registration: the kernel bundle framework, Journal of Mathematical Imaging and Vision 2012;46:07.

22. Tanya Schmah, Laurent Risser, François-Xavier Vialard, Left-invariant metrics for diffeomorphic image registration with spatially-varying regularisation, Kensaku Mori, Ichiro Sakuma, Yoshinobu Sato, Christian Barillot, Nassir Navab, eds. Medical Image Computing and Computer-Assisted Intervention – MICCAI 2013: 16th International Conference, Nagoya, Japan, September 22–26, 2013, Proceedings, Part I. Berlin, Heidelberg: Springer; 2013:203–210.

23. Tanya Schmah, Laurent Risser, François-Xavier Vialard, Diffeomorphic Image Matching With Left-Invariant Metrics. New York, NY: Springer; 2015:373–392.

24. Mingzhen Tan, Anqi Qiu, Multiscale frame-based kernels for large deformation diffeomorphic metric mapping, IEEE Transactions on Medical Imaging 2018;37(10):2344–2355.

25. Alain Trouvè, Action de groupe de dimension infinie et reconnaissance de formes (Infinite-dimensional group action and pattern recognition), 1995.

26. Tom Vercauteren, Xavier Pennec, Aymeric Perchant, Nicholas Ayache, Symmetric log-domain diffeomorphic registration: a demons-based approach, Int. Conference on Medical Image Computing and Computer Assisted Intervention (MICCAI). LNCS. 2008:754–761.

27. François-Xavier Vialard, Laurent Risser, Spatially-varying metric learning for diffeomorphic image registration: a variational framework, Polina Golland, Nobuhiko Hata, Christian Barillot, Joachim Hornegger, Robert Howe, eds. Medical Image Computing and Computer-Assisted Intervention – MICCAI 2014: 17th International Conference, Boston, MA, USA, September 14–18, 2014, Proceedings, Part I. Cham: Springer International Publishing; 2014:227–234.

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

29. Laurent Younes, Shapes and Diffeomorphisms. Springer; 2010.

30. Laurent Younes, Constrained diffeomorphic shape evolution, Foundations of Computational Mathematics Jun 2012;12(3):295–325.

31. Miaomiao Zhang, Fletcher P. Thomas, Finite-dimensional Lie algebras for fast diffeomorphic image registration, Sebastien Ourselin, Daniel C. Alexander, Carl-Fredrik Westin, Cardoso M. Jorge, eds. Information Processing in Medical Imaging. Cham: Springer International Publishing; 2015:249–260.

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

You can't read the all page of ebook, please click here login for view all page.
Reset
3.145.2.184