This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Sparse Non-Negative Stencils for Anisotropic Diffusion thanks: This work was partly supported by ANR grant MESANGE ANR-08-BLAN-0198.

Jérôme Fehrenbach111 Institut de Mathématiques de Toulouse, Université Paul Sabatier, 31062 TOULOUSE CEDEX 9, France    Jean-Marie Mirebeau222CNRS, Laboratory CEREMADE, UMR 7534, University Paris Dauphine, Place du Maréchal De Lattre De Tassigny 75775 PARIS CEDEX 16, France
Abstract

We introduce a new discretization scheme for Anisotropic Diffusion, AD-LBR, on two and three dimensional cartesian grids. The main features of this scheme is that it is non-negative and has sparse stencils, of cardinality bounded by 66 in 2D, by 1212 in 3D, despite allowing diffusion tensors of arbitrary anisotropy. The radius of these stencils is not a-priori bounded however, and can be quite large for pronounced anisotropies. Our scheme also has good spectral properties, which permits larger time steps and avoids e.g. chessboard artifacts.

AD-LBR relies on Lattice Basis Reduction, a tool from discrete mathematics which has recently shown its relevance for the discretization on grids of strongly anisotropic Partial Differential Equations [14]. We prove that AD-LBR is in 2D asymptotically equivalent to a finite element discretization on an anisotropic Delaunay triangulation, a procedure more involved and computationally expensive. Our scheme thus benefits from the theoretical guarantees of this procedure, for a fraction of its cost. Numerical experiments in 2D and 3D illustrate our results.

keywords :

Anisotropic Diffusion, Non-Negative Numerical Scheme, Lattice Basis Reduction.

We consider throughout this paper a bounded smooth domain Ωd\Omega\subset{\mathbb{R}}^{d}, where d{2,3}d\in\{2,3\} denotes the dimension, equipped with a continuous diffusion tensor 𝐃\operatorname{\mathbf{D}}. We do not impose any bound on the diffusion tensor anisotropy, and we are in fact interested in pronounced, non axis-aligned anisotropies. Anisotropic diffusion is here understood in the sense of [25]: the diffusion tensor 𝐃(z)\operatorname{\mathbf{D}}(z), at a point zΩz\in\Omega, is a symmetric positive definite matrix whose eigenvalues may have different orders of magnitude. Our results are not relevant for isotropic diffusion with a variable scalar coefficient, as in the pioneering work of Perona and Malik [20].

We address the discretization of the following energy {\mathcal{E}}, defined for uH1(Ω)u\in H^{1}(\Omega):

(u):=Ωu(z)𝐃(z)2𝑑z.{\mathcal{E}}(u):=\int_{\Omega}\|\nabla u(z)\|^{2}_{\operatorname{\mathbf{D}}(z)}dz. (1)

We denote eM:=e,Me\|e\|_{M}:=\sqrt{\langle e,Me\rangle}, for any ede\in{\mathbb{R}}^{d}, and any MM in the set Sd+S_{d}^{+} of symmetric positive definite d×dd\times d matrices. Gradient descent for the energy (1) has the form of a parabolic PDE:

tu=div(𝐃u).\partial_{t}u=\operatorname{div}(\operatorname{\mathbf{D}}\nabla u). (2)

This equation, Anisotropic Diffusion, is with its variants at the foundation of powerful image processing techniques. Some variants include curvature terms [19], or diffusion-reaction terms [5]. Time varying and solution dependent diffusion tensors can also be considered. A general exposition can be found in [25], where various choices for the definition of the diffusion tensor 𝐃{\bf D} from the image uu, adapted to various applications, are proposed and discussed.

Our contribution in the discretization of the energy (1) results in improved numerical solutions of (2), in terms of accuracy and stability, for a minor increase in complexity. This extends to applications, such as Coherence Enhancing Diffusion and Edge Preserving Diffusion [25], see the numerical experiments in §4, which involve solving (2) using a solution dependent diffusion tensor 𝐃=𝐃(u){\bf D}={\bf D}(u). For that purpose, one fixes a time step ΔT\Delta T, and solves for each integer n0n\geq 0 the linear diffusion equation tu=div(𝐃nu)\partial_{t}u=\operatorname{div}(\operatorname{\mathbf{D}}_{n}\nabla u) on the interval [nΔT,(n+1)ΔT][n\Delta T,(n+1)\Delta T], with 𝐃n:=𝐃(u(nΔT))\operatorname{\mathbf{D}}_{n}:=\operatorname{\mathbf{D}}(u(n\Delta T)). In these applications, the diffusion tensor 𝐃(u){\bf D}(u) is typically defined in terms of the structure tensor [25] of uu, in such way that diffusion is pronounced within image homogeneous regions, and tangentially along image edges, but not across edges.

In two dimensions, AD-LBR strictly speaking is not the first non-negative scheme for anisotropic diffusion: the proof of Theorem 6 in [25] implicitly defines an alternative 6-point non-negative scheme. This alternative scheme does however lack many of the qualities of AD-LBR: it leads to axis aligned artifacts, spectral aberrations, stencils of larger radius, reduced numerical accuracy, and does not extend to 3D. A detailed description and comparison is presented in §4.1.

Consider a scale parameter h>0h>0, and a sampling Ωh\Omega_{h} of the domain Ω\Omega on the cartesian grid d{\mathbb{Z}}^{d}, rescaled by hh: with obvious notations

Ωh:=Ωhd.\Omega_{h}:=\Omega\cap h{\mathbb{Z}}^{d}.

We introduce a novel discretization of the energy (1), referred to as AD-LBR (Anisotropic Diffusion using Lattice Basis Reduction). It is a sum of squared differences of a discrete map uL2(Ωh)u\in L^{2}(\Omega_{h})

h(u):=hd2zΩheV(z)γz(e)|u(z+he)u(z)|2{\mathcal{E}}_{h}(u):=h^{d-2}\sum_{z\in\Omega_{h}}\sum_{e\in V(z)}\gamma_{z}(e)\,|u(z+he)-u(z)|^{2} (3)

The stencils V(z)dV(z)\subset{\mathbb{Z}}^{d}, zΩhz\in\Omega_{h}, are symmetric and have cardinality at most 66 in 2D, 1212 in 3D. The coefficients γz(e)0\gamma_{z}(e)\geq 0 are non-negative. They are constructed using a classical tool from discrete mathematics, Lattice Basis Reduction, which allows to cheaply build efficient stencils for grid discretizations of Partial Differential Equations (PDEs) involving strongly anisotropic diffusion tensors or Riemannian metrics. This approach has been applied to anisotropic static Hamilton-Jacobi PDEs in [14], resulting in a new numerical scheme: Fast Marching using Lattice Basis Reduction (FM-LBR). Substantial improvements were obtained in comparison with earlier methods, in terms of both accuracy and complexity.

The paper is organized as follows. We describe the stencils of the two dimensional AD-LBR in §1, and state our main 2D result: the asymptotic equivalence of AD-LBR with a finite element discretization on an Anisotropic Delaunay Triangulation. Section §2 provides additional details on the two dimensional stencils of AD-LBR, and describes the three dimensional ones. The more technical §3 details the proof of the 2D equivalence result stated in §1. Two and three dimensional numerical experiments are presented in §4, including qualitative and quantitative comparisons with five other numerical schemes.

1 Description of the scheme, and main results

Our numerical scheme, Anisotropic Diffusion using Lattice Basis Reduction (AD-LBR), involves the construction of stencils whose geometry is tailored after the local diffusion tensor. Its essential feature is non-negativity: the discrete energy h(u){\mathcal{E}}_{h}(u) is written as a sum (3) of squared differences of values of uu, with non-negative weights γz(e)0\gamma_{z}(e)\geq 0. This discretization is consistent if for each zΩhz\in\Omega_{h}, and any smooth uu,

hdu(z)𝐃(z)2=hd2eV(z)γz(e)u(z),he2.h^{d}\|\nabla u(z)\|_{\operatorname{\mathbf{D}}(z)}^{2}=h^{d-2}\sum_{e\in V(z)}\gamma_{z}(e)\,\langle\nabla u(z),he\rangle^{2}. (4)

Indeed, the left hand side approximates the contribution of the “voxel” z+[h/2,h/2]dz+[-h/2,h/2]^{d} to the integral (1), while the right hand side is obtained by inserting the first order approximation u(z+he)u(z)+u(z),heu(z+he)\simeq u(z)+\langle\nabla u(z),he\rangle in (3). The identity (4) is in turn equivalent to

𝐃(z)=eV(z)γz(e)eeT.\operatorname{\mathbf{D}}(z)=\sum_{e\in V(z)}\gamma_{z}(e)\,ee^{\mathrm{T}}. (5)

The next lemma shows how to obtain such a decomposition in 2D. We denote by u:=(b,a)u^{\perp}:=(-b,a) the rotation of a vector u=(a,b)2u=(a,b)\in{\mathbb{R}}^{2} by π/2\pi/2, in such way that for all v2v\in{\mathbb{R}}^{2}:

u,v=det(u,v).\langle u^{\perp},v\rangle=\det(u,v).
Lemma 1.

Let e0,e1,e22e_{0},e_{1},e_{2}\in{\mathbb{R}}^{2} be such that e0+e1+e2=0e_{0}+e_{1}+e_{2}=0, and |det(e1,e2)|=1|\det(e_{1},e_{2})|=1. Then for any DS2+D\in S_{2}^{+}, with the convention e3+i:=eie_{3+i}:=e_{i}:

D=0i2ei+1,Dei+2eieiT.D=-\sum_{0\leq i\leq 2}\langle e_{i+1}^{\perp},De_{i+2}^{\perp}\rangle\,e_{i}e_{i}^{\mathrm{T}}. (6)
Proof.

Note that 1=|det(e2,e0)|=|det(e0,e1)|1=|\det(e_{2},e_{0})|=|\det(e_{0},e_{1})|. Denoting by DD^{\prime} the right hand side of (6), we obtain

e1,De1\displaystyle\langle e_{1}^{\perp},D^{\prime}e_{1}^{\perp}\rangle =e0,De1e2,e12e1,De2e0,e12\displaystyle=-\langle e_{0}^{\perp},De_{1}^{\perp}\rangle\langle e_{2},e_{1}^{\perp}\rangle^{2}-\langle e_{1}^{\perp},De_{2}^{\perp}\rangle\langle e_{0},e_{1}^{\perp}\rangle^{2}
=e0+e2,De1=e1,De1.\displaystyle=-\langle e_{0}^{\perp}+e_{2}^{\perp},De_{1}^{\perp}\rangle=\langle e_{1}^{\perp},De_{1}^{\perp}\rangle.

Thus e1D=e1D\|e_{1}^{\perp}\|_{D^{\prime}}=\|e_{1}^{\perp}\|_{D}. Likewise e2D=e2D\|e_{2}^{\perp}\|_{D^{\prime}}=\|e_{2}^{\perp}\|_{D}, and e1+e2D=e0D=e0D=e1+e2D\|e_{1}^{\perp}+e_{2}^{\perp}\|_{D^{\prime}}=\|e_{0}^{\perp}\|_{D^{\prime}}=\|e_{0}^{\perp}\|_{D}=\|e_{1}^{\perp}+e_{2}^{\perp}\|_{D}. Since (e1,e2)(e_{1}^{\perp},e_{2}^{\perp}) is a basis of 2{\mathbb{R}}^{2}, the result follows. ∎

Refer to caption
Refer to caption
Figure 1: Right: the stencils associated to three matrices MM of anisotropy ratios κ(M)\kappa(M) equal to 1.11.1, 3.53.5, 88 respectively. The ellipses {z2;zM=1}\{z\in{\mathbb{R}}^{2};\,\|z\|_{M}=1\} are shown left; their principal axis is aligned with (cos(π/3),sin(π/3))(\cos(\pi/3),\sin(\pi/3)). More stencils are shown in [14].

The diffusion tensor 𝐃\operatorname{\mathbf{D}} is meant to measure gradients, as in (1). In order to measure angles between vectors, we introduce a Riemannian metric 333The Laplace Beltrami operator associated to 𝐌\operatorname{\mathbf{M}} does not coincide with div(𝐃)\operatorname{div}(\operatorname{\mathbf{D}}\nabla\cdot), unless 𝐃\operatorname{\mathbf{D}} is identically of determinant 11. This is not an issue for our application. 𝐌\operatorname{\mathbf{M}} on the domain Ω\Omega, which is proportional to the inverse of 𝐃\operatorname{\mathbf{D}}: for all zΩz\in\Omega

𝐌(z):=𝐝(z)𝐃(z)1, where𝐝(z):=det(𝐃(z))1d.\operatorname{\mathbf{M}}(z):=\operatorname{\mathbf{d}}(z)\operatorname{\mathbf{D}}(z)^{-1},\,\text{ where}\,\operatorname{\mathbf{d}}(z):=\det(\operatorname{\mathbf{D}}(z))^{\frac{1}{d}}. (7)

The normalizing factor 𝐝(z)\operatorname{\mathbf{d}}(z) was chosen so as to normalize the metric determinant: det(𝐌(z))=1\det(\operatorname{\mathbf{M}}(z))=1. This normalization reflects the fact that the construction of our stencil V(z)V(z) depends on the preferred direction of diffusion, and on the amount of anisotropy, whereas the absolute strength of diffusion is irrelevant. In dimension d=2d=2, one easily checks that for any zΩz\in\Omega and any e,f2e,f\in{\mathbb{R}}^{2}, one has

e,𝐃(z)f=𝐝(z)e,𝐌(z)f.\langle e^{\perp},\operatorname{\mathbf{D}}(z)f^{\perp}\rangle=\operatorname{\mathbf{d}}(z)\langle e,\operatorname{\mathbf{M}}(z)f\rangle. (8)

The AD-LBR is based on decompositions (5), given by the previous lemma, with a family of vectors (ei)i=02(e_{i})_{i=0}^{2} chosen so that the scalar products appearing in (6) are non-positive. The adequate concept is that of MM-obtuse superbase of d{\mathbb{Z}}^{d} [4].

Definition 1.
  • A basis of d{\mathbb{Z}}^{d} is a family (ei)i=1d(e_{i})_{i=1}^{d} of elements of d{\mathbb{Z}}^{d} such that |det(e1,,ed)|=1|\det(e_{1},\cdots,e_{d})|=1.

  • A superbase of d{\mathbb{Z}}^{d} is a family (ei)i=0d(e_{i})_{i=0}^{d} such that e0++ed=0e_{0}+\cdots+e_{d}=0, and (ei)i=1d(e_{i})_{i=1}^{d} is a basis of d{\mathbb{Z}}^{d}.

Definition 2.

Let MSd+M\in S_{d}^{+}. A family (ei)iI(e_{i})_{i\in I} of vectors in d{\mathbb{R}}^{d} is said to be MM-obtuse if ei,Mej0\langle e_{i},Me_{j}\rangle\leq 0 for all distinct i,jIi,j\in I.

In dimension d3d\leq 3, there exists for each MSd+M\in S_{d}^{+} at least one MM-obtuse superbase of d{\mathbb{Z}}^{d} [4]. The practical construction of such superbases is discussed in §2, and based on lattice basis reduction algorithms described in [11, 23, 17] (hence the name of our numerical scheme). This construction has a logarithmic numerical cost 𝒪(lnκ(M)){\mathcal{O}}(\ln\kappa(M)) in the anisotropy ratio of the matrix MM:

κ(M):=max|u|=|v|=1uMvM=MM1.\kappa(M):=\max_{|u|=|v|=1}\frac{\|u\|_{M}}{\|v\|_{M}}=\sqrt{\|M\|\|M^{-1}\|}. (9)

The AD-LBR energy h:L2(Ωh)+{\mathcal{E}}_{h}:L^{2}(\Omega_{h})\to{\mathbb{R}}_{+}, see (3), is in two dimensions written in terms of the following stencils and coefficients. Let zΩz\in\Omega, and let e0,e1,e2e_{0},e_{1},e_{2} be an 𝐌(z)\operatorname{\mathbf{M}}(z)-obtuse superbase of 2{\mathbb{Z}}^{2}. We set

V(z):={e0,e1,e2,e0,e1,e2},V(z):=\{e_{0},e_{1},e_{2},\,-e_{0},-e_{1},-e_{2}\},\\ (10)

and for 0i20\leq i\leq 2, with the convention ei+3:=eie_{i+3}:=e_{i},

γz(±ei):=12ei+1,𝐃(z)ei+2.\gamma_{z}(\pm e_{i}):=-\frac{1}{2}\langle e^{\perp}_{i+1},\operatorname{\mathbf{D}}(z)e^{\perp}_{i+2}\rangle. (11)

Lemma 1 implies the announced decomposition (5), and the weights γz\gamma_{z} are non-negative in view of (8). These weights γz:2+\gamma_{z}:{\mathbb{Z}}^{2}\to{\mathbb{R}}_{+}, extended by 0 outside V(z)V(z), do not depend on the choice of 𝐌(z)\operatorname{\mathbf{M}}(z)-obtuse superbase (e0,e1,e2)(e_{0},e_{1},e_{2}), see Lemma 11. Stencils of the three dimensional AD-LBR are described in §2, and involve a construction of Selling444The authors would like to thank Professor P. Q. Nguyen for pointing out this 1212 points 3D stencil, which is simpler and sparser than the 1414 points stencil proposed by the authors in an earlier version of the manuscript. [22]. The above description of the stencils V(z)V(z) is suitable for periodic, reflected, and Dirichlet boundary conditions (extending uu by zero outside Ωh\Omega_{h} in the latter case). In the case of Neumann boundary conditions, a slight modification is in order:

V(z;h):={eV(z);z+heΩh}.V(z;\,h):=\{e\in V(z);\,z+he\in\Omega_{h}\}.

We have so far established three strongpoints of the AD-LBR:

Non-negativity.

Off diagonal coefficients of the symmetric semi-definite N×NN\times N matrix, N=#(Ωh)N=\#(\Omega_{h}), associated to the energy h{\mathcal{E}}_{h} are non-positive, while diagonal coefficients are positive.

Sparsity.

Stencil cardinality is uniformly bounded, without restriction on the anisotropy ratio κ(𝐃(z))\kappa(\operatorname{\mathbf{D}}(z)) of the diffusion tensor.

Complexity.

The construction of the stencil V(z)V(z), and of the associated coefficients γz\gamma_{z}, has a logarithmic cost 𝒪(lnκ(𝐃(z))){\mathcal{O}}(\ln\kappa(\operatorname{\mathbf{D}}(z))) in the anisotropy ratio of the diffusion tensor.

The next result, Theorem 1, restricted to the two dimensional case, establishes that AD-LBR is asymptotically equivalent to a more involved and computationally intensive procedure: a finite element discretization of the energy (1), on an Anisotropic Delaunay Triangulation (ADT, see [10] and below) of the domain Ω\Omega. Under the assumptions of Theorem 1, AD-LBR benefits from two additional guarantees, that we state informally and without proof.

No chessboard artifacts.

Some numerical schemes for anisotropic diffusion suffer from chessboard artifacts, in the sense that periodic artifacts develop at the pixel level. Such artifacts cannot develop in finite element discretizations, since they would lead to high frequency oscillations of the finite element interpolant, and therefore to an increase of the energy (12). The asymptotic equivalence of the AD-LBR with a finite element discretization also rules out these defects.

Spectral correctness.

The nn-th smallest eigenvalue λn(h)\lambda_{n}(h) of the symmetric matrix associated to hdhh^{-d}{\mathcal{E}}_{h} (3), converges as h0h\to 0 towards the nn-th smallest eigenvalue λn\lambda_{n} of the continuous operator div(𝐃)-\operatorname{div}(\operatorname{\mathbf{D}}\nabla), for any given integer n0n\geq 0. See Figure 8 page 8 for an illustration. This follows from a similar property of the finite element energy h{\mathcal{E}}^{\prime}_{h} (12), and from the asymptotic equivalence (13).

Our convergence result, Theorem 1 below, is specialized to the case of a square periodic domain, which covers reflecting boundary conditions frequently used in image processing. Since the grid discretization must be compatible with the boundary conditions, any scale parameter hh appearing in the rest of the paper is assumed to be the inverse of a positive integer:

h{1/n;n1}.h\in\{1/n;\,n\geq 1\}.
Theorem 1.

Let Ω\Omega be the unit square [0,1[2[0,1[^{2}, equipped with periodic boundary conditions. Let 𝐃:Ω¯S2+\operatorname{\mathbf{D}}:\overline{\Omega}\to S_{2}^{+} be a (periodic) diffusion tensor with Lipschitz regularity, and let 𝐌\operatorname{\mathbf{M}} be the Riemannian metric defined by (7). When hh is sufficiently small, the periodic Riemannian domain (Ω,𝐌)(\Omega,\operatorname{\mathbf{M}}) admits an Anisotropic Delaunay Triangulation 𝒯h{\mathcal{T}}_{h}, with collection of vertices Ωh:=Ωh2\Omega_{h}:=\Omega\cap h{\mathbb{Z}}^{2}. For uL2(Ωh)u\in L^{2}(\Omega_{h}), define

h(u):=Ω(I𝒯hu)(z)𝐃(z)2𝑑z,{\mathcal{E}}^{\prime}_{h}(u):=\int_{\Omega}\|\nabla(\operatorname{I}_{{\mathcal{T}}_{h}}u)(z)\|^{2}_{\operatorname{\mathbf{D}}(z)}dz, (12)

where I𝒯\operatorname{I}_{\mathcal{T}} denotes the piecewise linear interpolation operator on a triangulation 𝒯{\mathcal{T}}. Then for some constant c=c(𝐃)c=c(\operatorname{\mathbf{D}}), independent of uu and hh,

(1ch)h(u)h(u)(1+ch)h(u).(1-ch){\mathcal{E}}_{h}(u)\leq{\mathcal{E}}^{\prime}_{h}(u)\leq(1+ch){\mathcal{E}}_{h}(u). (13)
Refer to caption
Refer to caption
Figure 2: The distance δp(q)\delta_{p}(q), from a grid point pp to q2q\in{\mathbb{R}}^{2}, is defined in terms of the local metric 𝐌(p)\operatorname{\mathbf{M}}(p), see (14). The level lines {q2;δp(q)=r}\{q\in{\mathbb{R}}^{2};\,\delta_{p}(q)=r\} are ellipses (left). The collection of points q2q\in{\mathbb{R}}^{2} closer to pp than to any other grid point is the Voronoi region of pp (left: the boundaries of Voronoi regions are shown dashed). The Voronoi diagram (right) is the collection of all Voronoi regions.
Refer to caption
Refer to caption
Figure 3: A Voronoi vertex qq is a point where the Voronoi regions of at least three grid points (pi)i=1k(p_{i})_{i=1}^{k} intersect (left: Voronoi region boundaries are shown dashed); here k=3k=3 (values k>3k>3 are non-generic). The dual Voronoi cell TqT_{q}, generically a triangle, is the convex envelope of the grid points {pi; 1ik}\{p_{i};\,1\leq i\leq k\} (left). The collection of dual Voronoi cells TqT_{q} defines a polygonization 𝒬h{\mathcal{Q}}_{h}, generically a triangulation, and the Anisotropic Delaunay Triangulation 𝒯h{\mathcal{T}}_{h} is obtained by arbitrarily triangulating (if necessary) the elements of 𝒬h{\mathcal{Q}}_{h}. Here 𝒯h=𝒬h{\mathcal{T}}_{h}={\mathcal{Q}}_{h} (right). The stencil Vh(p)V_{h}(p), of a vertex pp of 𝒯h{\mathcal{T}}_{h}, see (17), is represented by arrows (right).

Let us mention that the finite element discretization on an ADT is a more general procedure than AD-LBR, since it does not require the domain Ω\Omega to be sampled on a grid. This flexibility can be used to locally increase the density of vertices, in places where solution uu is expected to be less regular, or to insert vertices exactly on Ω\partial\Omega for a better discretization of boundary conditions. (Such refinements are however generally incompatible with image processing since the unknowns, the pixel values, lie by construction on a fixed and given cartesian grid.) Here and as often, the performance of AD-LBR is at the cost of its specialization.

The proof of Theorem 1 is postponed to §3, but for the sake of concreteness, we describe here the concept of Anisotropic Delaunay Triangulation (ADT) [10]. In the rest of this introduction, and in §3, we assume as in Theorem 1 that the diffusion tensor 𝐃\operatorname{\mathbf{D}} is defined on the square [0,1]2[0,1]^{2} and satisfies periodic boundary conditions. We extend it, as well as the metric 𝐌\operatorname{\mathbf{M}}, to the whole plane 2{\mathbb{R}}^{2} by periodicity.

We specialize the concept of ADT [10], to the domain 2{\mathbb{R}}^{2} and the collection of vertices h2h{\mathbb{Z}}^{2}. For that purpose, we introduce some notations. For all p,q2p,q\in{\mathbb{R}}^{2}, we denote by δp(q)\delta_{p}(q) the distance from pp to qq, as measured by the metric at the point pp:

δp(q):=qp𝐌(p).\delta_{p}(q):=\|q-p\|_{\operatorname{\mathbf{M}}(p)}. (14)

We denote by Δh(q)\Delta_{h}(q) the least distance from a point q2q\in{\mathbb{R}}^{2}, to the grid h2h{\mathbb{Z}}^{2}:

Δh(q):=minph2δp(q).\Delta_{h}(q):=\min_{p\in h{\mathbb{Z}}^{2}}\delta_{p}(q). (15)

We introduce the Voronoi cell Vorh(p)\operatorname{Vor}_{h}(p) of a grid point ph2p\in h{\mathbb{Z}}^{2}, which is the collection of points q2q\in{\mathbb{R}}^{2} closer to pp than to any other grid point:

Vorh(p):={q2;δp(q)=Δh(q)}.\operatorname{Vor}_{h}(p):=\{q\in{\mathbb{R}}^{2};\,\delta_{p}(q)=\Delta_{h}(q)\}. (16)

The collection of Voronoi cells is referred to as the Voronoi diagram, see Figure 2. A Voronoi vertex is a point q2q\in{\mathbb{R}}^{2} at which at least three distinct Voronoi regions intersect: (Vorh(pi))i=1k(\operatorname{Vor}_{h}(p_{i}))_{i=1}^{k}, k3k\geq 3, pih2p_{i}\in h{\mathbb{Z}}^{2}. We attach to qq a dual Voronoi cell TqT_{q}, defined as the convex hull of the points (pi)i=1k(p_{i})_{i=1}^{k}, see Figure 3.

The geometric dual 𝒬h{\mathcal{Q}}_{h}, of the Voronoi diagram, is defined as the collection of all dual Voronoi cells TqT_{q}. Note that, generically on the metric 𝐌\operatorname{\mathbf{M}}, no more than three Voronoi regions can intersect at any point in 2{\mathbb{R}}^{2}, thus the elements of 𝒬h{\mathcal{Q}}_{h} are generically triangles. If hh is small enough, we show in §3 (using the Dual Triangulation Theorem in [10]) that TqT_{q} is a strictly convex polygon, of vertices (pi)i=1k(p_{i})_{i=1}^{k} with the above notations, and that 𝒬h{\mathcal{Q}}_{h} is a polygonization (generically a triangulation) of 2{\mathbb{R}}^{2}, with vertices h2h{\mathbb{Z}}^{2}.

Since the metric 𝐌\operatorname{\mathbf{M}} and the vertices h2h{\mathbb{Z}}^{2} are periodic (recall that h=1/nh=1/n for some integer n1n\geq 1), arbitrarily triangulating the elements of 𝒬h{\mathcal{Q}}_{h}, respecting periodicity, yields a periodic triangulation 𝒯h{\mathcal{T}}_{h}.

Definition 3 (ADT, Labelle and Shewchuk [10]).

The triangulation 𝒯h{\mathcal{T}}_{h} obtained by the above construction is referred to as an ADT of the domain 2{\mathbb{R}}^{2}, with collection of vertices h2h{\mathbb{Z}}^{2}, and underlying Riemannian metric 𝐌\operatorname{\mathbf{M}}. Since 𝒯h{\mathcal{T}}_{h} is 2{\mathbb{Z}}^{2}-periodic, we also regard it as an ADT of the periodic unit square Ω\Omega.

We establish in §3.1 the existence of the ADT 𝒯h{\mathcal{T}}_{h}. Incidentally, we show in Lemma 7 (iii) page 7 that the angles of the elements of 𝒯h{\mathcal{T}}_{h}, measured with respect to the local metric 𝐌\operatorname{\mathbf{M}}, are asymptotically acute. This geometrical property (which holds thanks to our special choice of triangulation vertices, on a grid) is linked to the non-negativity of AD-LBR: indeed, it is known that finite elements discretizations such as (12) yield non-negative numerical schemes, and the discrete maximum principle, if the mesh satisfies a non-obtuse angle condition, see Lemma 3.1 in [9].

Subsection §3.2 is devoted to the study of MM-obtuse superbases of 2{\mathbb{Z}}^{2}, and their cousins MM-reduced bases of 2{\mathbb{Z}}^{2}, on which the AD-LBR relies: we discuss their characterization, uniqueness and stability properties. We study in §3.3 the finite element stencils, defined for ph2p\in h{\mathbb{Z}}^{2} by

Vh(p):={e2;[p,p+he] is an edge of 𝒯h},V_{h}(p):=\{e\in{\mathbb{Z}}^{2};[p,p+he]\text{ is an edge of }{\mathcal{T}}_{h}\}, (17)

see Figure 3 (right). We show that Vh(p)V_{h}(p) coincides with the AD-LBR stencil V(p)V(p), unless the lattice 2{\mathbb{Z}}^{2} admits a basis almost orthogonal with respect to the scalar product associated to 𝐌(p)\operatorname{\mathbf{M}}(p), see Lemma 13. This is tied to the fact that orthogonal grids admit several (usual) Delaunay triangulations. Overcoming this technical difficulty, we conclude the proof of Theorem 1.

Note that the construction of the ADT 𝒯h{\mathcal{T}}_{h} is not easy to parallelize, in particular when anisotropy is pronounced since the Voronoi regions of far away points interact. The construction of 𝒯h{\mathcal{T}}_{h} also involves solving polynomial equations of degree four, because Voronoi regions boundaries are conics, and Voronoi vertices must be identified at their intersections. In contrast, the AD-LBR stencils are independent of each other, and the numerical cost of their construction only grows logarithmically with the metric anisotropy.

2 Construction of obtuse superbases, and three dimensional stencils

Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 4: Right: stencil of the AD-LBR, for a symmetric matrix of eigenvector MM of anisotropy ratio κ(M)\kappa(M) equal to 22 (top) or 66 (bottom). The anisotropy is of “needle” type: the two largest eigenvalues of MM are equal, and the needle orientation is given by the vector (4,2,3)(4,2,3). The ellipsoid {z3;z1}\{z\in{\mathbb{R}}^{3};\,\|z\|\leq 1\} is shown left.

Algorithms for the construction of privileged bases of lattices, consisting of short and almost orthogonal vectors, have attracted an important research effort from the mathematical community, over a long period of time. The first such algorithm dates back to Lagrange [11], and is restricted to two dimensional lattices. Methods for high dimensional lattices, such as the LLL algorithm [12], are of key importance for integer programming and cryptography [18]. AD-LBR is based on the original algorithm of Lagrange [11], and on its recent extension to three dimensional lattices [23, 17]. These methods output a basis of d{\mathbb{Z}}^{d} reduced in the sense of Minkowski, which in dimension d4d\leq 4 is equivalent to the following definition.

We denote by e1++eke_{1}{\mathbb{Z}}+\cdots+e_{k}{\mathbb{Z}} the sub-lattice of d{\mathbb{Z}}^{d} generated by vectors e1,,ekde_{1},\cdots,e_{k}\in{\mathbb{Z}}^{d}. This sub-lattice equals {0}\{0\} by convention if k=0k=0.

Definition 4.

An MM-reduced basis of d{\mathbb{Z}}^{d}, where d4d\leq 4 and MSd+M\in S_{d}^{+}, is a basis (e1,,ed)(e_{1},\cdots,e_{d}) of d{\mathbb{Z}}^{d} such that

eiM=min{eM;ed(e1++ei1)}.\|e_{i}\|_{M}=\min\{\|e\|_{M};\,e\in{\mathbb{Z}}^{d}\setminus(e_{1}{\mathbb{Z}}+\cdots+e_{i-1}{\mathbb{Z}})\}. (18)

For each d4d\leq 4, and each MSd+M\in S_{d}^{+}, there exists at least one MM-reduced basis [17]. In contrast, there exists MS5+M\in S_{5}^{+} for which no basis of d{\mathbb{Z}}^{d} satisfies (18). The norms of the elements (ei)i=1d(e_{i})_{i=1}^{d} of an MM-reduced basis,

λi(M):=eiM,\lambda_{i}(M):=\|e_{i}\|_{M}, (19)

are called the Minkowski minima, and are independent of the choice of MM-reduced basis. In particular, e1e_{1} is the shortest vector of d{\mathbb{Z}}^{d}, with respect to the norm M\|\cdot\|_{M}, and e2e_{2} is the shortest linearly independent vector.

Lemma 2.

For any MSd+M\in S_{d}^{+}, 1id1\leq i\leq d,

M1212λi(M)M12.\|M^{-\frac{1}{2}}\|^{-\frac{1}{2}}\leq\lambda_{i}(M)\leq\|M\|^{\frac{1}{2}}.
Proof.

Note that M1212eeMM12e\|M^{-\frac{1}{2}}\|^{-\frac{1}{2}}\|e\|\leq\|e\|_{M}\leq\|M\|^{\frac{1}{2}}\|e\|, for any e2e\in{\mathbb{R}}^{2}. In addition: (i) any e2{0}e\in{\mathbb{Z}}^{2}\setminus\{0\} satisfies e1\|e\|\geq 1, and (ii) the set d(e1++ei1){\mathbb{Z}}^{d}\setminus(e_{1}{\mathbb{Z}}+\cdots+e_{i-1}{\mathbb{Z}}) appearing in (18) always contains at least one element ee of the canonical basis of d{\mathbb{R}}^{d}, so that e1\|e\|\leq 1. The announced result easily follows. ∎

We emphasize that obtaining an MM-reduced basis, i.e. solving the minimization problems (18), is both simple and cheap numerically. In dimension d=2d=2, this is the object of Lagrange’s algorithm [11] (later rediscovered by Gauss and often erroneously called Gauss’s algorithm, see [17]): initialize (e,f)(e,f) as the canonical basis of 2{\mathbb{Z}}^{2}, and

Do (e,f):=(f,eRound(e,Mf/fM2)f),\displaystyle\text{\bf Do }(e,\ f):=(f,\ e-\mathrm{Round}(\langle e,Mf\rangle/\|f\|_{M}^{2})\,f), (20)
while eM>fM.\displaystyle\text{\bf while }\|e\|_{M}>\|f\|_{M}.

This algorithm can be regarded as a two dimensional geometrical generalization of greatest common divisor computation. It can be extended to higher dimension and, in dimension up to four, outputs an MM-reduced basis after at most 𝒪(lnκ(M)){\mathcal{O}}(\ln\kappa(M)) iterations [17], each consisting of 𝒪(1){\mathcal{O}}(1) operations among reals.

The elements of an MM-reduced basis are heuristically never very far from being orthogonal, as illustrated by the following lemma.

Lemma 3.

Let MSd+M\in S_{d}^{+}, d4d\leq 4, and let (e1,,ed)(e_{1},\cdots,e_{d}) be an MM-reduced basis. Then for any i,j{1,,d}i,j\in\{1,\cdots,d\},

2|ei,Mej|eiM2.2|\langle e_{i},Me_{j}\rangle|\leq\|e_{i}\|_{M}^{2}. (21)
Proof.

Since ekM\|e_{k}\|_{M} is an increasing function of k{1,,d}k\in\{1,\cdots,d\}, we may assume that i<ji<j. If follows from (18) that ejMej+eiM\|e_{j}\|_{M}\leq\|e_{j}+e_{i}\|_{M}, and ejMejeiM\|e_{j}\|_{M}\leq\|e_{j}-e_{i}\|_{M}. Squaring these inequalities, and developing the scalar products, we obtain the announced result. ∎

Corollary 1.

Let MS2+M\in S_{2}^{+}, and let (e,f)(e,f) be an MM-reduced basis such that e,Mf0\langle e,Mf\rangle\leq 0. Then (e,f,g)(e,f,g) is an MM-obtuse superbase of 2{\mathbb{Z}}^{2}, with g:=efg:=-e-f. In addition

e,MgeM2/2,f,MgfM2/2.\langle e,Mg\rangle\leq-\|e\|_{M}^{2}/2,\quad\langle f,Mg\rangle\leq-\|f\|_{M}^{2}/2. (22)
Proof.

The previous Lemma implies e,M(e+f)eM2|e,Mf|12eM2.\langle e,M(e+f)\rangle\geq\|e\|_{M}^{2}-|\langle e,Mf\rangle|\geq\frac{1}{2}\|e\|_{M}^{2}. Likewise f,M(e+f)12fM2\langle f,M(e+f)\rangle\geq\frac{1}{2}\|f\|_{M}^{2}. The result follows. ∎

The practical construction of the two dimensional AD-LBR stencil at a point zΩz\in\Omega amounts to (i) compute an 𝐌(z)\operatorname{\mathbf{M}}(z)-reduced basis (e,f)(e,f) using Lagrange’s algorithm (20), (ii) replace ff with f-f, if necessary, so that e,Mf0\langle e,Mf\rangle\leq 0, and (iii) define the stencil V(z)V(z) and the weights γz\gamma_{z} in terms of the MM-obtuse superbase (e,f,g)(e,f,g) of 2{\mathbb{Z}}^{2}, where g=efg=-e-f, as described in (10) and (11).

The rest of this section is devoted to the description of the three dimensional AD-LBR stencils. In constrast with the two dimensional case, the construction of the 3D stencil V(z)V(z) at a point zΩz\in\Omega, involves a 𝐃(z)\operatorname{\mathbf{D}}(z)-obtuse basis, instead of an 𝐌(z)\operatorname{\mathbf{M}}(z)-obtuse basis.

Proposition 1.

Let DS3+D\in S_{3}^{+}, and let (e1,e2,e3)(e_{1},e_{2},e_{3}) be a DD-reduced basis. Let bi:=εieσ(i)b_{i}:=\varepsilon_{i}e_{\sigma(i)}, for all 1i31\leq i\leq 3, where the signs ε1,ε2,ε3{1,1}\varepsilon_{1},\varepsilon_{2},\varepsilon_{3}\in\{-1,1\}, and the permutation σ\sigma of {1,2,3}\{1,2,3\} are chosen so that

|b1,Db2|min{b1,Db3,b2,Db3}.|\langle b_{1},Db_{2}\rangle|\leq\min\{-\langle b_{1},Db_{3}\rangle,\,-\langle b_{2},Db_{3}\rangle\}. (23)

Then the following is a DD-obtuse superbase:

{(b1,b2,b3,b1b2b3)if b1,Db20,(b1,b2,b1+b3,b2b3)otherwise.\left\{\begin{array}[]{cl}(b_{1},b_{2},b_{3},-b_{1}-b_{2}-b_{3})&\text{if }\langle b_{1},Db_{2}\rangle\leq 0,\\ (-b_{1},b_{2},b_{1}+b_{3},-b_{2}-b_{3})&\text{otherwise.}\end{array}\right. (24)
Proof.

To achieve (23), one can choose σ\sigma such that bi:=eσ(i)b^{\prime}_{i}:=e_{\sigma(i)} satisfies |b1,Db2||b1,Db3||b2,Db3||\langle b^{\prime}_{1},Db^{\prime}_{2}\rangle|\leq|\langle b^{\prime}_{1},Db^{\prime}_{3}\rangle|\leq|\langle b^{\prime}_{2},Db^{\prime}_{3}\rangle|. Then choose the signs (εi)i=13(\varepsilon_{i})_{i=1}^{3} such that bi:=εibib_{i}:=\varepsilon_{i}b^{\prime}_{i} satisfies b1,Db30\langle b_{1},Db_{3}\rangle\leq 0 and b2,Db30\langle b_{2},Db_{3}\rangle\leq 0.

The two families of vectors appearing in (24) are clearly superbases. We thus only need to show that they are DD-obtuse; in other words that e,Df0\langle e,Df\rangle\leq 0 for any two distinct elements e,fe,f of these families. Note that for all distinct i,j{1,2,3}i,j\in\{1,2,3\}, using (21),

2|bi,Dbj|biD2.2|\langle b_{i},Db_{j}\rangle|\leq\|b_{i}\|_{D}^{2}.

In the case where b1,Db20\langle b_{1},Db_{2}\rangle\leq 0, the pairwise scalar products between b1,b2,b3b_{1},b_{2},b_{3} are non-positive by construction. In addition

2b1+b2+b3,Db1\displaystyle 2\langle b_{1}+b_{2}+b_{3},Db_{1}\rangle
\displaystyle\geq (b1D22|b1,Db2|)+(b1D22|b1,Db3|)0.\displaystyle(\|b_{1}\|_{D}^{2}-2|\langle b_{1},Db_{2}\rangle|)+(\|b_{1}\|^{2}_{D}-2|\langle b_{1},Db_{3}\rangle|)\geq 0.

Likewise b1+b2+b3,Dbi0\langle b_{1}+b_{2}+b_{3},Db_{i}\rangle\geq 0 for all i{1,2,3}i\in\{1,2,3\}, which concludes the proof.

We next turn to the second case, where b1,Db20\langle b_{1},Db_{2}\rangle\geq 0. Enumerating all scalar products we obtain

b1,D(b1+b3)\displaystyle\langle b_{1},D(b_{1}+b_{3})\rangle b1D2|b1,Db3|0,\displaystyle\geq\|b_{1}\|_{D}^{2}-|\langle b_{1},Db_{3}\rangle|\geq 0,
b1,D(b2b3)\displaystyle\langle b_{1},D(-b_{2}-b_{3})\rangle =b1,Db2b1,Db30,\displaystyle=-\langle b_{1},Db_{2}\rangle-\langle b_{1},Db_{3}\rangle\geq 0,
b2,D(b1+b3)\displaystyle-\langle b_{2},D(b_{1}+b_{3})\rangle =b2,Db1b2,Db30,\displaystyle=-\langle b_{2},Db_{1}\rangle-\langle b_{2},Db_{3}\rangle\geq 0,
b2,D(b2+b3)\displaystyle\langle b_{2},D(b_{2}+b_{3})\rangle b2D2|b2,Db3|0,\displaystyle\geq\|b_{2}\|^{2}_{D}-|\langle b_{2},Db_{3}\rangle|\geq 0,

and finally

2b1+b3,D(b2+b3)2b1,Db2\displaystyle 2\langle b_{1}+b_{3},D(b_{2}+b_{3})\rangle\geq 2\langle b_{1},Db_{2}\rangle
+(b322|b1,Db3|)+(b322|b2,Db3|)0.\displaystyle+(\|b_{3}\|^{2}-2|\langle b_{1},Db_{3}\rangle|)+(\|b_{3}\|^{2}-2|\langle b_{2},Db_{3}\rangle|)\geq 0.

This concludes the proof. ∎

In view of the previous Proposition, obtaining a DD-obtuse superbase of 3{\mathbb{Z}}^{3} has numerical cost 𝒪(lnκ(D)){\mathcal{O}}(\ln\kappa(D)). Indeed a DD-reduced basis needs to be computed in a preliminary step, after what Proposition 1 is applied for a negligible 𝒪(1){\mathcal{O}}(1) cost. An alternative method for the construction of DD-obtuse superbases of 3{\mathbb{Z}}^{3} is presented in [4] and in appendix B of [3], but its numerical complexity is not known to the authors.

The three dimensional AD-LBR is defined by the following stencils and coefficients. Let zΩz\in\Omega, let D:=𝐃(z)D:=\operatorname{\mathbf{D}}(z), and let (ei)i=03(e_{i})_{i=0}^{3} be a DD-obtuse superbase of 3{\mathbb{Z}}^{3}. We set

V(z):={ek×el;k,l{0,1,2,3},kl},V(z):=\{e_{k}\times e_{l};\ k,l\in\{0,1,2,3\},\ k\neq l\},

and if {i,j,k,l}={0,1,2,3}\{i,j,k,l\}=\{0,1,2,3\}, iji\neq j and klk\neq l, then

γz(ek×el):=12ei,Dej.\gamma_{z}(e_{k}\times e_{l}):=-\frac{1}{2}\langle e_{i},De_{j}\rangle.

As announced, #(V(z))=12\#(V(z))=12, and the weights γz\gamma_{z} are non-negative. The proof of the scheme consistency (5), due to Selling [22], is reproduced in the next lemma for completeness. A generalization, appearing in Appendix B of [3], allows in arbitrary dimension to build a non-negative decomposition of the form (25) from a DD-obtuse superbase of d{\mathbb{Z}}^{d}. However the non existence of such a superbase, for some matrices DS4+D\in S_{4}^{+}, forbids a straightforward extension of AD-LBR to higher dimension.

Lemma 4 (Selling [22]).

Let (ei)i=03(e_{i})_{i=0}^{3} be a superbase of 3{\mathbb{Z}}^{3}. For all i,j,k,li,j,k,l such that {i,j,k,l}={0,1,2,3}\{i,j,k,l\}=\{0,1,2,3\}, i<ji<j, and k<lk<l, let cij:=ek×elc_{ij}:=e_{k}\times e_{l}. Then, for any DS3+D\in S_{3}^{+}:

D=0i<j3ei,DejcijcijT.D=-\sum_{0\leq i<j\leq 3}\langle e_{i},De_{j}\rangle c_{ij}c_{ij}^{\mathrm{T}}. (25)
Proof.

Let i,j,k,li,j,k,l be as in the definition of cijc_{ij}. Then

ei,cij=ei,ek×el=det(ei,ek,el){1,1},\langle e_{i},c_{ij}\rangle=\langle e_{i},e_{k}\times e_{l}\rangle=\det(e_{i},e_{k},e_{l})\in\{-1,1\},

since (ei,ek,el)(e_{i},e_{k},e_{l}) is a basis of 3{\mathbb{Z}}^{3}. Also

ej,cij\displaystyle\langle e_{j},c_{ij}\rangle =eiekel,ek×el\displaystyle=\langle-e_{i}-e_{k}-e_{l},\,e_{k}\times e_{l}\rangle
=ei,ek×el=ei,cij.\displaystyle=-\langle e_{i},\ e_{k}\times e_{l}\rangle=-\langle e_{i},c_{ij}\rangle.

In addition, clearly, ek,cij=el,cij=0.\langle e_{k},c_{ij}\rangle=\langle e_{l},c_{ij}\rangle=0. Denoting by DD^{\prime} the right hand side of (25), we obtain as a result

e0,De0\displaystyle\langle e_{0},D^{\prime}e_{0}\rangle =e0,De1e0,De2e0,De3\displaystyle=-\langle e_{0},De_{1}\rangle-\langle e_{0},De_{2}\rangle-\langle e_{0},De_{3}\rangle
=e0,D(e1e2e3)=e0,De0.\displaystyle=\langle e_{0},D(-e_{1}-e_{2}-e_{3})\rangle=\langle e_{0},De_{0}\rangle.
e0,De1\displaystyle\langle e_{0},D^{\prime}e_{1}\rangle =e0,De1e0,c01e1,c01=e0,De1.\displaystyle=-\langle e_{0},De_{1}\rangle\langle e_{0},c_{01}\rangle\langle e_{1},c_{01}\rangle=\langle e_{0},De_{1}\rangle.

Likewise ei,Dej=ei,Dej\langle e_{i},D^{\prime}e_{j}\rangle=\langle e_{i},De_{j}\rangle for all i,j{1,2,3,4}i,j\in\{1,2,3,4\}. It follows as announced that D=DD=D^{\prime}. ∎

3 Equivalence to a finite element discretization

This section is devoted to the proof of Theorem 1: the asymptotic equivalence of AD-LBR with a finite element discretization on an Anisotropic Delaunay Triangulation (ADT). We use the notations of §1. The existence of the ADT 𝒯h{\mathcal{T}}_{h} is established in the first subsection, for hh sufficiently small, as well as a few of its properties. The second subsection is devoted to the study of MM-reduced bases. Theorem 1 is proved in the third subsection, by comparing the stencils of the AD-LBR and of the finite element discretization.

We denote by 𝜿{\boldsymbol{\kappa}} the maximum anisotropy ratio (9) of the diffusion tensor

𝜿:=maxzΩκ(𝐃(z)).{\boldsymbol{\kappa}}:=\max_{z\in\Omega}\kappa(\operatorname{\mathbf{D}}(z)). (26)

Observing that κ(𝐃(z))=κ(𝐌(z))\kappa(\operatorname{\mathbf{D}}(z))=\kappa(\operatorname{\mathbf{M}}(z)), and recalling that det(𝐌(z))=1\det(\operatorname{\mathbf{M}}(z))=1, one easily checks that

𝜿12ee𝐌(z)𝜿12e,{\boldsymbol{\kappa}}^{-\frac{1}{2}}\|e\|\leq\|e\|_{\operatorname{\mathbf{M}}(z)}\leq{\boldsymbol{\kappa}}^{\frac{1}{2}}\|e\|, (27)

for all zΩz\in\Omega and all e2e\in{\mathbb{R}}^{2}.

3.1 Existence of an ADT

Our first lemma provides an uniform bound on the size of the Voronoi regions, see Figure 3, involved in the construction of the ADT.

Lemma 5.
  1. (i)

    For all r2r\in{\mathbb{R}}^{2}, one has Δh(r)𝜿12h\Delta_{h}(r)\leq{\boldsymbol{\kappa}}^{\frac{1}{2}}h.

  2. (ii)

    If p,qh2p,q\in h{\mathbb{Z}}^{2}, and rVorh(p)Vorh(q)r\in\operatorname{Vor}_{h}(p)\cap\operatorname{Vor}_{h}(q), then pr𝜿h{\|p-r\|\leq{\boldsymbol{\kappa}}h} and pq2𝜿h\|p-q\|\leq 2{\boldsymbol{\kappa}}h.

Proof.

Point (i). Rounding the coordinates of rr to a nearest multiple of hh, we obtain a point ph2p\in h{\mathbb{Z}}^{2} such that prh\|p-r\|\leq h. Recalling (27) we obtain δp(r)𝜿12h\delta_{p}(r)\leq{\boldsymbol{\kappa}}^{\frac{1}{2}}h, and therefore Δh(r)𝜿12h\Delta_{h}(r)\leq{\boldsymbol{\kappa}}^{\frac{1}{2}}h in view of (15).

Point (ii). We have 𝜿12prδp(r)=Δh(r)𝜿12h{\boldsymbol{\kappa}}^{-\frac{1}{2}}\|p-r\|\leq\delta_{p}(r)=\Delta_{h}(r)\leq{\boldsymbol{\kappa}}^{\frac{1}{2}}h. Thus pr𝜿h\|p-r\|\leq{\boldsymbol{\kappa}}h, and likewise qr𝜿h\|q-r\|\leq{\boldsymbol{\kappa}}h. Finally, by the triangle inequality, pqpr+qr2𝜿h\|p-q\|\leq\|p-r\|+\|q-r\|\leq 2{\boldsymbol{\kappa}}h. ∎

Following the notations of [10], we denote by τ(p,q)\tau(p,q), p,q2p,q\in{\mathbb{R}}^{2}, the smallest constant τ1\tau\geq 1 such that

τ1δp(r)δq(r)τδp(r),for all r2.\tau^{-1}\delta_{p}(r)\leq\delta_{q}(r)\leq\tau\delta_{p}(r),\quad\text{for all }r\in{\mathbb{R}}^{2}.

Equivalently, in the sense of symmetric matrices,

τ2𝐌(p)𝐌(q)τ2𝐌(p).\tau^{-2}\operatorname{\mathbf{M}}(p)\leq\operatorname{\mathbf{M}}(q)\leq\tau^{2}\operatorname{\mathbf{M}}(p). (28)

We also define a quantity τh1\tau_{h}\geq 1, closely related to the modulus of continuity of the metric 𝐌\operatorname{\mathbf{M}}:

τh:=max{τ(p,q);pq2𝜿h}.\tau_{h}:=\max\{\tau(p,q);\,\|p-q\|\leq 2{\boldsymbol{\kappa}}h\}. (29)

One has τh1\tau_{h}\to 1 as h0h\to 0, for any continuous metric 𝐌\operatorname{\mathbf{M}} (indeed 𝐌\operatorname{\mathbf{M}} is periodic and therefore uniformly continuous). If 𝐌\operatorname{\mathbf{M}} is Lipschitz, as assumed in Theorem 1, then τh=1+𝒪(h)\tau_{h}=1+{\mathcal{O}}(h).

We show in the next lemma the existence of an ADT, by applying the main result of [10], under the assumption that τh\tau_{h} is sufficiently small. More precisely, we assume in the rest of this subsection that

τh<1+𝜿2.\tau_{h}<\sqrt{1+{\boldsymbol{\kappa}}^{-2}}. (30)
Lemma 6.
  1. (i)

    If p,qh2p,q\in h{\mathbb{Z}}^{2}, pqp\neq q, and rVorh(p)Vorh(q)r\in\operatorname{Vor}_{h}(p)\cap\operatorname{Vor}_{h}(q), then δp(r)<δp(q)/τ(p,q)21.\delta_{p}(r)<\delta_{p}(q)/\sqrt{\tau(p,q)^{2}-1}.

  2. (ii)

    The geometric dual 𝒬h{\mathcal{Q}}_{h} of the Voronoi diagram is, as announced in §1, a polygonization of 2{\mathbb{R}}^{2} into strictly convex polygons, with vertices h2h{\mathbb{Z}}^{2}.

Proof.

Point (i). We may assume that τ(p,q)>1\tau(p,q)>1, otherwise there is nothing to prove. Point (ii) of Lemma 5 implies that pq2𝜿h\|p-q\|\leq 2{\boldsymbol{\kappa}}h, thus

τ(p,q)21τh21<𝜿1.\sqrt{\tau(p,q)^{2}-1}\leq\sqrt{\tau_{h}^{2}-1}<{\boldsymbol{\kappa}}^{-1}.

On the other hand δp(q)𝜿12qp𝜿12h\delta_{p}(q)\geq{\boldsymbol{\kappa}}^{-\frac{1}{2}}\|q-p\|\geq{\boldsymbol{\kappa}}^{-\frac{1}{2}}h, and δp(r)Δh(r)𝜿12h\delta_{p}(r)\leq\Delta_{h}(r)\leq{\boldsymbol{\kappa}}^{\frac{1}{2}}h. The announced inequality follows.

Point (ii). We apply Theorem 7 (Dual Triangulation Theorem) in [10]. Since the domain 2{\mathbb{R}}^{2} has no boundary, it suffices to check that all the Voronoi arcs and vertices are wedged, see [10]. This condition means that for any p,qh2p,q\in h{\mathbb{Z}}^{2} such that pqp\neq q, and any rVorh(p)Vorh(q)r\in\operatorname{Vor}_{h}(p)\cap\operatorname{Vor}_{h}(q), one has (rq)𝐌(q)(pq)>0(r-q)\operatorname{\mathbf{M}}(q)(p-q)>0, and likewise exchanging the roles of pp and qq. Heuristically, it expresses the acuteness of some angles measured in the local metric. Lemma 5 in [10] shows that this condition follows from point (i) of this lemma, which concludes the proof. ∎

We recall that 𝒯h{\mathcal{T}}_{h} is the triangulation obtained by arbitrarily triangulating the polygonization QhQ_{h} of the previous lemma, respecting periodicity, see Definition 3. Generically QhQ_{h} is already a triangulation, hence 𝒯h=Qh{\mathcal{T}}_{h}=Q_{h}, see §1. The Voronoi regions Vorh\operatorname{Vor}_{h}, and the triangulation 𝒯h{\mathcal{T}}_{h}, are illustrated in Figures 2 and 3.

The next lemma provides estimates of the diameter, the area, and the angles of the elements of 𝒯h{\mathcal{T}}_{h}. These geometrical properties also have an interpretation in the context of lattices: (ii) shows that the edges of any triangle T𝒯hT\in{\mathcal{T}}_{h} define a superbase (e,f,g)(e,f,g) of 2{\mathbb{Z}}^{2}, and (iii) that this superbase is almost 𝐌(z)\operatorname{\mathbf{M}}(z)-obtuse, for any zTz\in T.

Note that the vertices p,q,rp,q,r of any triangle T𝒯hT\in{\mathcal{T}}_{h} satisfy by construction

Vorh(p)Vorh(q)Vorh(r).\operatorname{Vor}_{h}(p)\cap\operatorname{Vor}_{h}(q)\cap\operatorname{Vor}_{h}(r)\neq\emptyset. (31)
Lemma 7.

Denote by he,hf,hghe,hf,hg the edges of a triangle T𝒯hT\in{\mathcal{T}}_{h}, where e,f,g2e,f,g\in{\mathbb{Z}}^{2} are oriented so that e+f+g=0e+f+g=0. Then

  1. (i)

    max{e,f,g}2𝜿\max\{\|e\|,\|f\|,\|g\|\}\leq 2{\boldsymbol{\kappa}}.

  2. (ii)

    |det(e,f)|=1|\det(e,f)|=1, thus |T|=h2/2|T|=h^{2}/2.

  3. (iii)

    e,𝐌(z)fθh\langle e,\operatorname{\mathbf{M}}(z)f\rangle\leq\theta_{h}, for any zTz\in T, where θh0\theta_{h}\to 0 as h0h\to 0. (Explicitly: θh=𝜿(3+9τ2h2)(τ2h21)\theta_{h}={\boldsymbol{\kappa}}(3+9\tau_{2h}^{2})(\tau_{2h}^{2}-1))

Proof.

Point (i). We denote by p,q,rp,q,r the vertices of TT, ordered in such way that p+he=qp+he=q, q+hf=rq+hf=r, r+hg=pr+hg=p. The announced estimate follows from (31), and from point (ii) of Lemma 5.

Point (ii). Since 𝒯h{\mathcal{T}}_{h} is a conforming triangulation, the intersection of TT with the collection h2h{\mathbb{Z}}^{2} of all vertices of 𝒯h{\mathcal{T}}_{h} consists of only three points: the vertices p,q,rp,q,r of TT. Thus the triangle of vertices e,0,f-e,0,f, homothetic to TT, contains no point of integer coordinates but its vertices. This implies that (e,f)(e,f) is a basis of 2{\mathbb{Z}}^{2}, hence |det(e,f)|=1|\det(e,f)|=1, as announced.

Point (iii). The pairwise distances between p,q,rp,q,r are bounded by 2𝜿h2{\boldsymbol{\kappa}}h, see point (i), and since zTz\in T so are the pairwise distances between p,q,r,zp,q,r,z. Defining s:=pq+rh2s:=p-q+r\in h{\mathbb{Z}}^{2}, and observing that sp=rq2𝜿h\|s-p\|=\|r-q\|\leq 2{\boldsymbol{\kappa}}h, we find that the pairwise distances between p,q,r,z,sp,q,r,z,s are bounded by 4𝜿h4{\boldsymbol{\kappa}}h.

Let xVorh(p)Vorh(q)Vorh(r)x\in\operatorname{Vor}_{h}(p)\cap\operatorname{Vor}_{h}(q)\cap\operatorname{Vor}_{h}(r). We have δp(x)=δq(x)=δr(x)=Δh(x)δs(x)\delta_{p}(x)=\delta_{q}(x)=\delta_{r}(x)=\Delta_{h}(x)\leq\delta_{s}(x), thus

δs(x)2δp(x)2δq(x)2+δr(x)2.\delta_{s}(x)^{2}\geq\delta_{p}(x)^{2}-\delta_{q}(x)^{2}+\delta_{r}(x)^{2}. (32)

(For intuition: in a classical Delaunay triangulation, xx would be the circumcenter of TT, and (32) would state that ss is outside the circumcircle of TT.) Denoting M:=𝐌(z)M:=\operatorname{\mathbf{M}}(z), and δ:=Δh(x)\delta:=\Delta_{h}(x), we obtain

|δp(x)2xpM2|\displaystyle|\delta_{p}(x)^{2}-\|x-p\|^{2}_{M}| δp(x)2(τ(p,z)21)\displaystyle\leq\delta_{p}(x)^{2}(\tau(p,z)^{2}-1)
δ2(τ2h21),\displaystyle\leq\delta^{2}(\tau_{2h}^{2}-1), (33)

using Lemma 5, and likewise for q,rq,r. We also have

δs(x)\displaystyle\delta_{s}(x) =pq+rx𝐌(s)\displaystyle=\|p-q+r-x\|_{\operatorname{\mathbf{M}}(s)}
px𝐌(s)+qx𝐌(s)+rx𝐌(s)\displaystyle\leq\|p-x\|_{\operatorname{\mathbf{M}}(s)}+\|q-x\|_{\operatorname{\mathbf{M}}(s)}+\|r-x\|_{\operatorname{\mathbf{M}}(s)}
3δτ2h.\displaystyle\leq 3\delta\tau_{2h}.

Thus, proceeding as in (33),

|δs(x)sxM2|δs(x)2(τ2h21)9δ2τ2h2(τ2h21).|\delta_{s}(x)-\|s-x\|^{2}_{M}|\leq\delta_{s}(x)^{2}(\tau_{2h}^{2}-1)\leq 9\delta^{2}\tau^{2}_{2h}(\tau_{2h}^{2}-1).

Inserting in (32) these estimates of δ(x)\delta_{\star}(x), {p,q,r,s}\star\in\{p,q,r,s\}, and using the fact that δ𝜿12h\delta\leq{\boldsymbol{\kappa}}^{\frac{1}{2}}h, see Lemma 5, we obtain after expansion the announced estimate of e,Mf\langle e,Mf\rangle. ∎

We next rewrite the finite element energy h{\mathcal{E}}^{\prime}_{h} (12) in a form similar to that of the AD-LBR energy h{\mathcal{E}}_{h} (3). Let φph:2\varphi_{p}^{h}:{\mathbb{R}}^{2}\to{\mathbb{R}} be the piecewise linear function on 𝒯h{\mathcal{T}}_{h} such that φph(p)=1\varphi_{p}^{h}(p)=1, and φph(q)=0\varphi_{p}^{h}(q)=0 for any vertex qh2q\in h{\mathbb{Z}}^{2} distinct from pp. This is the classical “hat function” encountered in finite element analysis. For all ph2p\in h{\mathbb{Z}}^{2}, e2{0}e\in{\mathbb{Z}}^{2}\setminus\{0\}, let

γph(e):=122φph(z),𝐃(z)φp+heh(z)𝑑z\gamma^{h}_{p}(e):=-\frac{1}{2}\int_{{\mathbb{R}}^{2}}\langle\nabla\varphi_{p}^{h}(z),\operatorname{\mathbf{D}}(z)\nabla\varphi_{p+he}^{h}(z)\rangle dz (34)

Clearly, γph(e)=0\gamma_{p}^{h}(e)=0 if [p,p+he][p,p+he] is not an edge of 𝒯h{\mathcal{T}}_{h}, in other words if ee does not belong to the stencil Vh(p)V_{h}(p), defined in (17). We express in the next lemma the finite element energy h{\mathcal{E}}^{\prime}_{h} (12) in terms of the stencils VhV_{h} and of the (potentially negative) weights γph\gamma^{h}_{p}.

Lemma 8.

For any uL2(Ωh)u\in L^{2}(\Omega_{h}), extended by periodicity to h2h{\mathbb{Z}}^{2}, one has

h(u)=pΩheVh(p)γph(e)|u(p+he)u(p)|2.{\mathcal{E}}^{\prime}_{h}(u)=\sum_{p\in\Omega_{h}}\sum_{e\in V_{h}(p)}\gamma_{p}^{h}(e)|u(p+he)-u(p)|^{2}. (35)
Proof.

For any triangle T𝒯hT\in{\mathcal{T}}_{h}, and any p,qh2p,q\in h{\mathbb{Z}}^{2}, we denote

sT(p,q):=Tφph(z),𝐃(z)φqh(z)𝑑z.s_{T}(p,q):=\int_{T}\langle\nabla\varphi_{p}^{h}(z),\,\operatorname{\mathbf{D}}(z)\nabla\varphi_{q}^{h}(z)\rangle\,dz.

Clearly sT(p,q)=0s_{T}(p,q)=0 if qq or pp is not a vertex of TT. The coefficient γph(e)\gamma_{p}^{h}(e), e2e\in{\mathbb{Z}}^{2}, is thus given by the following sum with at most two non-zero terms:

γph(e)=12T𝒯hsT(p,p+he).\gamma_{p}^{h}(e)=-\frac{1}{2}\sum_{T\in{\mathcal{T}}_{h}}s_{T}(p,p+he). (36)

Let p,q,rh2p,q,r\in h{\mathbb{Z}}^{2} be the vertices of a triangle T𝒯hT\in{\mathcal{T}}_{h}. Since the sum φph+φqh+φrh\varphi_{p}^{h}+\varphi_{q}^{h}+\varphi_{r}^{h} is constant on TT, equal to 11, it has a null gradient on TT, and therefore

sT(p,p)+sT(p,q)+sT(p,r)=0.s_{T}(p,p)+s_{T}(p,q)+s_{T}(p,r)=0.

Using this relation, and the two similar ones obtained by a cyclic permutation of p,q,rp,q,r, we obtain

T(I𝒯hu)(z)𝐃(z)2𝑑z\displaystyle\int_{T}\|\nabla(\operatorname{I}_{{\mathcal{T}}_{h}}u)(z)\|_{\operatorname{\mathbf{D}}(z)}^{2}dz
=\displaystyle= u(p)2sT(p,p)+u(q)2sT(q,q)+u(r)2sT(r,r)\displaystyle\,u(p)^{2}s_{T}(p,p)+u(q)^{2}s_{T}(q,q)+u(r)^{2}s_{T}(r,r)
+2u(p)u(q)sT(p,q)+2u(q)u(r)sT(q,r)\displaystyle+2u(p)u(q)s_{T}(p,q)+2u(q)u(r)s_{T}(q,r)
+2u(r)u(p)sT(r,p),\displaystyle+2u(r)u(p)s_{T}(r,p),
=\displaystyle= sT(p,q)(u(p)u(q))2sT(q,r)(u(q)u(r))2\displaystyle-s_{T}(p,q)(u(p)-u(q))^{2}-s_{T}(q,r)(u(q)-u(r))^{2}
sT(r,p)(u(r)u(p))2.\displaystyle-s_{T}(r,p)(u(r)-u(p))^{2}.

Summing this expression over all T𝒯hT\in{\mathcal{T}}_{h}, and combining it with (36), we obtain (35), which concludes the proof. ∎

Finally, we provide an approximation of the coefficients γph\gamma_{p}^{h} which will be easily compared with the AD-LBR weights γp\gamma_{p} (11).

Lemma 9.

Consider an edge [p,p+he][p,p+he] of 𝒯h{\mathcal{T}}_{h}, shared by the two distinct triangles T,T𝒯hT,T^{\prime}\in{\mathcal{T}}_{h}. Let hf,hghf,hg (resp. hf,hghf^{\prime},hg^{\prime}) be the two other vector edges of TT (resp. TT^{\prime}), oriented so that e+f+g=0e+f+g=0 (resp. e+f+g=0e+f^{\prime}+g^{\prime}=0). Then

|γph(e)+14(f,𝐃(p)g+f,𝐃(p)g)|εh,\left|\gamma_{p}^{h}(e)+\frac{1}{4}\left(\langle f^{\perp},\operatorname{\mathbf{D}}(p)\,g^{\perp}\rangle+\langle f^{\prime\perp},\operatorname{\mathbf{D}}(p)\,g^{\prime\perp}\rangle\right)\right|\leq\varepsilon_{h},

where εh:=2𝛋2max{𝐃(x)𝐃(y);xy2𝛋h}\varepsilon_{h}:=2{\boldsymbol{\kappa}}^{2}\max\{\|\operatorname{\mathbf{D}}(x)-\operatorname{\mathbf{D}}(y)\|;\,\|x-y\|\leq 2{\boldsymbol{\kappa}}h\}.

Proof.

We assume, up to exchanging ff and gg, that [p,phg][p,p-hg] is an edge of TT. Let α:=det(e,f){1,1}\alpha:=\det(e,f)\in\{-1,1\}, see point (ii) of Lemma 7; note that α=det(f,g)=det(g,e)\alpha=\det(f,g)=\det(g,e). Let γ\gamma be the constant value of φph\nabla\varphi_{p}^{h} on TT. Then γ,he=1\langle\gamma,he\rangle=-1 and γ,hg=1\langle\gamma,hg\rangle=1. These two independent linear identities are also satisfied by αf/h\alpha f^{\perp}/h, hence φph=γ=αf/h\nabla\varphi_{p}^{h}=\gamma=\alpha f^{\perp}/h on TT.

Denoting q:=p+hgq:=p+hg, we obtain likewise φqh=αg/h\nabla\varphi_{q}^{h}=\alpha g^{\perp}/h on TT. Hence recalling that |T|=h2/2|T|=h^{2}/2:

Tφph,𝐃(p)φqh\displaystyle\int_{T}\langle\nabla\varphi_{p}^{h},\operatorname{\mathbf{D}}(p)\nabla\varphi_{q}^{h}\rangle =h22αfh,𝐃(p)αgh\displaystyle=\frac{h^{2}}{2}\left\langle\frac{\alpha f^{\perp}}{h},\operatorname{\mathbf{D}}(p)\frac{\alpha g^{\perp}}{h}\right\rangle
=12f,𝐃(p)g\displaystyle=\frac{1}{2}\left\langle f^{\perp},\operatorname{\mathbf{D}}(p)g^{\perp}\right\rangle

Therefore, using point (i) of Lemma 7 in the last step,

|Tφph,𝐃(z)φqh𝑑z12f,𝐃(p)g|\displaystyle\left|\int_{T}\langle\nabla\varphi_{p}^{h},\operatorname{\mathbf{D}}(z)\nabla\varphi_{q}^{h}\rangle dz-\frac{1}{2}\langle f^{\perp},\operatorname{\mathbf{D}}(p)g^{\perp}\rangle\right|
=|Tφph,(𝐃(z)𝐃(p))φqh𝑑z|\displaystyle=\left|\int_{T}\langle\nabla\varphi_{p}^{h},(\operatorname{\mathbf{D}}(z)-\operatorname{\mathbf{D}}(p))\nabla\varphi_{q}^{h}\rangle dz\right|
h222𝜿h2𝜿hmax{𝐃(z)𝐃(p);zT}εh.\displaystyle\leq\frac{h^{2}}{2}\frac{2{\boldsymbol{\kappa}}}{h}\frac{2{\boldsymbol{\kappa}}}{h}\max\{\|\operatorname{\mathbf{D}}(z)-\operatorname{\mathbf{D}}(p)\|;\,z\in T\}\leq\varepsilon_{h}.

Proceeding likewise on TT^{\prime}, and recalling (34) (or (36)), we conclude the proof. ∎

3.2 Some properties of MM-reduced bases

We establish some technical properties of MM-reduced bases, thanks to which we will be able to compare in §3.3 the “geometric” construction of the ADT finite element stencils VhV_{h}, with the lattice based construction of the AD-LBR stencils VV.

Lemma 10.

Let MS2+M\in S_{2}^{+}, let e1,,en2e_{1},\cdots,e_{n}\in{\mathbb{Z}}^{2}, n>2n>2, and let ε{1,1}\varepsilon\in\{-1,1\}. Assume that for all 1in1\leq i\leq n, with the convention en+1:=e1e_{n+1}:=e_{1}:

det(ei,ei+1)\displaystyle\det(e_{i},e_{i+1}) =ε,\displaystyle=\varepsilon, (37)
ei,Mei+1\displaystyle\langle e_{i},Me_{i+1}\rangle >12min{eiM2,ei+1M2}.\displaystyle>-\frac{1}{2}\min\left\{\|e_{i}\|_{M}^{2},\|e_{i+1}\|_{M}^{2}\right\}. (38)

Then any MM-reduced basis (e,f)(e,f) satisfies

{e,f}{e1,,en}.\{e,f\}\subset\{e_{1},\cdots,e_{n}\}.
Proof.

Let z2{e1,,en}z\in{\mathbb{Z}}^{2}\setminus\{e_{1},\cdots,e_{n}\}. Our objective is to show that zz cannot be an element of an MM-reduced basis, and we may therefore assume that zz has co-prime coordinates.

It follows from (37) that the closed polygonal line of consecutive vertices e1,,ene_{1},\cdots,e_{n}, circles at least once around the origin, see Figure 5. Hence z=αei+βei+1z=\alpha e_{i}+\beta e_{i+1}, for some 1in1\leq i\leq n and some α,β0\alpha,\beta\geq 0. Since (ei,ei+1)(e_{i},e_{i+1}) is a basis of 2{\mathbb{Z}}^{2} (indeed |det(ei,ei+1)|=1|\det(e_{i},e_{i+1})|=1), the coefficients α\alpha and β\beta are integers. Since z{ei,ei+1}z\notin\{e_{i},e_{i+1}\}, α+β2\alpha+\beta\geq 2. Since zz has co-prime coordinates, αβ0\alpha\beta\neq 0.

Assuming without loss of generality that eiMei+1M\|e_{i}\|_{M}\geq\|e_{i+1}\|_{M}, we obtain using (38):

zM2=α2eiM2+β2ei+1M2+2αβei,Mei+1\displaystyle\|z\|_{M}^{2}=\alpha^{2}\|e_{i}\|_{M}^{2}+\beta^{2}\|e_{i+1}\|_{M}^{2}+2\alpha\beta\langle e_{i},Me_{i+1}\rangle
>α2eiM2+β2ei+1M2αβmin{eiM2,ei+1M2}\displaystyle>\alpha^{2}\|e_{i}\|_{M}^{2}+\beta^{2}\|e_{i+1}\|_{M}^{2}-\alpha\beta\min\{\|e_{i}\|_{M}^{2},\|e_{i+1}\|_{M}^{2}\}
eiM2+(α2+β21αβ)ei+1M2.\displaystyle\geq\|e_{i}\|^{2}_{M}+(\alpha^{2}+\beta^{2}-1-\alpha\beta)\|e_{i+1}\|_{M}^{2}.

Observing that α2+β21αβ0\alpha^{2}+\beta^{2}-1-\alpha\beta\geq 0 for all α,β[1,[\alpha,\beta\in[1,\infty[, we obtain zM>eiM\|z\|_{M}>\|e_{i}\|_{M}. Since eie_{i} and ei+1e_{i+1} are linearly independent, we have eiMλ2(M)\|e_{i}\|_{M}\geq\lambda_{2}(M). Finally zM>λ2(M)\|z\|_{M}>\lambda_{2}(M), hence zz cannot be an element of an MM-reduced basis, which concludes the proof. ∎

The next corollary reverses the construction, presented in Corollary 1, of an MM-obtuse superbase from an MM-reduced basis.

Corollary 2.

Let MS2+M\in S_{2}^{+} and let (e,f,g)(e,f,g) be an MM-obtuse superbase of 2{\mathbb{Z}}^{2}, ordered so that eMfMgM\|e\|_{M}\leq\|f\|_{M}\leq\|g\|_{M}. Then (e,f)(e,f) is an MM-reduced basis.

Proof.

The family (e,g,f,e,g,f)(e,-g,f,-e,g,-f) satisfies by construction the conditions of the previous lemma. Hence any 𝐌(z)\operatorname{\mathbf{M}}(z)-reduced basis (e,f)(e^{\prime},f^{\prime}) of 2{\mathbb{Z}}^{2} satisfies {e,f}{e,f,g,e,f,g}\{e^{\prime},f^{\prime}\}\subset\{e,f,g,-e,-f,-g\}. Observing that ee^{\prime} and ff^{\prime} are linearly independent, that eMfM\|e^{\prime}\|_{M}\leq\|f^{\prime}\|_{M}, and that eMfMgM\|e\|_{M}\leq\|f\|_{M}\leq\|g\|_{M}, we obtain that eMeM\|e\|_{M}\leq\|e^{\prime}\|_{M} and fMfM\|f\|_{M}\leq\|f^{\prime}\|_{M}. Recalling that MM-reduced bases are defined by the minimality of their M\|\cdot\|_{M}-norms, see Definition 4, we obtain as announced that (e,f)(e,f) is an MM-reduced basis. ∎

The previous lemma shows that for any zΩz\in\Omega, there exists an 𝐌(z)\operatorname{\mathbf{M}}(z)-reduced basis (e,f)(e,f) such that

V(z)={e,f,e+f,e,f,ef}.V(z)=\{e,f,e+f,-e,-f,-e-f\}. (39)
Refer to caption
Refer to caption
Refer to caption
Figure 5: (left) A family e1,,e11e_{1},\cdots,e_{11} satisfying condition (37) of Lemma 10: the closed polygonal line of vertices (e1,,e11)(e_{1},\cdots,e_{11}) circles (at least) once around the origin, and the triangles (0,ei,ei+1)^\widehat{(0,e_{i},e_{i+1})} have area 1/21/2. (Center and right) The lattice 2{\mathbb{Z}}^{2}, and an MM-reduced basis (e,f)(e,f), shown after a linear change of coordinates by AA, such that ATA=MS2+A^{\mathrm{T}}A=M\in S_{2}^{+}. Case μ(M)=0\mu(M)=0 (center), and case 2μ(M)=λ1(M)22\mu(M)=\lambda_{1}(M)^{2} (right).

Given MS2+M\in S_{2}^{+}, and an MM-reduced basis (e,f)(e,f) of 2{\mathbb{Z}}^{2}, we denote μ(M):=|e,Mf|\mu(M):=|\langle e,Mf\rangle|. This value can be expressed in terms of the Minkowski minima (19) and thus does not depend on the particular choice of MM-reduced basis. Indeed, recalling the identity

e,Mf2+det(M)det(e,f)2=eM2fM2,\langle e,Mf\rangle^{2}+\det(M)\det(e,f)^{2}=\|e\|_{M}^{2}\|f\|_{M}^{2},

we obtain

μ(M)=|e,Mf|=λ1(M)2λ2(M)2det(M).\mu(M)=|\langle e,Mf\rangle|=\sqrt{\lambda_{1}(M)^{2}\lambda_{2}(M)^{2}-\det(M)}. (40)

In addition one has

02μ(M)λ1(M)2,0\leq 2\mu(M)\leq\lambda_{1}(M)^{2}, (41)

where the right hand side follows from Lemma 3. A vanishing value, μ(M)=0\mu(M)=0, indicates that the lattice 2{\mathbb{Z}}^{2} admits an MM-orthogonal basis. In contrast, when the upper bound is met, 2μ(M)=λ1(M)22\mu(M)=\lambda_{1}(M)^{2}, one has fM=f+εeM\|f\|_{M}=\|f+\varepsilon e\|_{M} for ε:=signe,Mf\varepsilon:=-{\rm sign}\langle e,Mf\rangle, hence the reduced basis (e,f)(e,f) is not unique even up to sign changes. See Figure 5.

We next show that the stencils of the AD-LBR do not depend on the choices of reduced bases, as was announced in the introduction.

Lemma 11.

The weights γz:2+\gamma_{z}:{\mathbb{Z}}^{2}\to{\mathbb{R}}_{+} used in the AD-LBR at a point zΩz\in\Omega (defined on V(z)V(z) by (11) and extended to 2{\mathbb{Z}}^{2} by 0), do not depend on the choice of 𝐌(z)\operatorname{\mathbf{M}}(z)-obtuse superbase of 2{\mathbb{Z}}^{2}.

Proof.

We denote M:=𝐌(z)M:=\operatorname{\mathbf{M}}(z) and D:=𝐃(z)D:=\operatorname{\mathbf{D}}(z). Let (e,f,g)(e,f,g) and (e,f,g)(e^{\prime},f^{\prime},g^{\prime}) be two MM-obtuse superbases, and let V,VV,V^{\prime} and γ,γ:2+\gamma,\gamma^{\prime}:{\mathbb{Z}}^{2}\to{\mathbb{R}}_{+} be the corresponding AD-LBR stencils and weights defined by (10) and (11). We may assume, using Corollary 2 and up to reordering, that (e,f)(e,f) and (e,f)(e^{\prime},f^{\prime}) are MM-reduced bases.

Corollary 1 states that the scalar products e,Mg\langle e,Mg\rangle, f,Mg\langle f,Mg\rangle, e,Mg\langle e^{\prime},Mg^{\prime}\rangle and f,Mg\langle f^{\prime},Mg^{\prime}\rangle are (strictly) negative. On the other hand

e,Mf=e,Mf=μ(M)0.\langle e,Mf\rangle=\langle e^{\prime},Mf^{\prime}\rangle=-\mu(M)\leq 0. (42)

Applying Lemma 10 to the family

(e,g,f,e,g,f)(e^{\prime},-g^{\prime},f^{\prime},-e^{\prime},g^{\prime},-f^{\prime})

we obtain that

{e,f}{e,f,g,e,f,g}.\{e,f\}\subset\{e^{\prime},f^{\prime},g^{\prime},-e^{\prime},-f^{\prime},-g^{\prime}\}. (43)

If μ(M)0\mu(M)\neq 0, then e,Mf\langle e,Mf\rangle and e,Mf\langle e^{\prime},Mf^{\prime}\rangle are negative, and not merely non-positive, thus {e,f}{e,f,g}\{e,f\}\subset\{e^{\prime},f^{\prime},g^{\prime}\}, or {e,f}{e,f,g}\{e,f\}\subset\{-e^{\prime},-f^{\prime},-g^{\prime}\}. Since e+f+g=0=e+f+ge+f+g=0=e^{\prime}+f^{\prime}+g^{\prime}, it follows that {e,f,g}={e,f,g}\{e,f,g\}=\{e^{\prime},f^{\prime},g^{\prime}\}, or {e,f,g}={e,f,g}\{e,f,g\}=\{-e^{\prime},-f^{\prime},-g^{\prime}\}. The stencils V,VV,V^{\prime} are thus identical, see (10), and so are the weights γ,γ\gamma,\gamma^{\prime}.

If μ(M)=0\mu(M)=0, then the stencils V,VV,V^{\prime} may not be identical. Observe however that e,Df=0=e,Df\langle e^{\perp},Df^{\perp}\rangle=0=\langle e^{\prime\perp},Df^{\prime\perp}\rangle, using (8). Hence using the weights expression (11):

γ(±g)\displaystyle\gamma(\pm g) =e,Df/2=0,\displaystyle=-\langle e^{\perp},Df^{\perp}\rangle/2=0, (44)
γ(±e)\displaystyle\gamma(\pm e) =fD2/2,\displaystyle=\|f^{\perp}\|_{D}^{2}/2, γ(±f)\displaystyle\gamma(\pm f) =eD2/2,\displaystyle=\|e^{\perp}\|_{D}^{2}/2,

and likewise for γ,e,f,g\gamma^{\prime},e^{\prime},f^{\prime},g^{\prime}. Note also that gM2=eM2+fM2>λ2(M)2\|g^{\prime}\|^{2}_{M}=\|e^{\prime}\|_{M}^{2}+\|f^{\prime}\|_{M}^{2}>\lambda_{2}(M)^{2}, hence ee and ff are different from gg^{\prime} and g-g^{\prime}. It follows from (43) that {e,f}={ε1e,ε2f}\{e,f\}=\{\varepsilon_{1}e^{\prime},\varepsilon_{2}f^{\prime}\} for some ε1,ε2{1,1}\varepsilon_{1},\varepsilon_{2}\in\{-1,1\}. This implies γ=γ\gamma=\gamma^{\prime} in view of (44), and concludes the proof. ∎

The next lemma establishes weak uniqueness and stability properties for MM-reduced bases, in the case of a strict inequality 2μ(M)<λ1(M)2.2\mu(M)<\lambda_{1}(M)^{2}.

Lemma 12.

Consider M,MS2+M,M^{\prime}\in S_{2}^{+}, an MM-reduced basis (e,f)(e,f), and an MM^{\prime}-reduced basis (e,f)(e^{\prime},f^{\prime}). Let τ1\tau\geq 1 be such that τ2MMτ2M\tau^{-2}M\leq M^{\prime}\leq\tau^{2}M, in the sense of symmetric matrices. Assume either:

  1. (i)

    2μ(M)<λ1(M)22\mu(M)<\lambda_{1}(M)^{2}, and τ=1\tau=1 (i.e. M=MM^{\prime}=M).

  2. (ii)

    4μ(M)λ1(M)24\mu(M)\leq\lambda_{1}(M)^{2}, and τ41+13κ(M)2\tau^{4}\leq 1+\frac{1}{3}{\kappa(M)^{-2}}.

Then {e,f}{e,f,e,f}.\{e^{\prime},f^{\prime}\}\subset\{e,f,-e,-f\}.

Proof.

Denoting α:=2μ(M)/λ1(M)2\alpha:=2\mu(M)/\lambda_{1}(M)^{2}, we obtain:

4e,Mf=e+fM2efM2\displaystyle 4\langle e,M^{\prime}f\rangle=\|e+f\|_{M^{\prime}}^{2}-\|e-f\|_{M^{\prime}}^{2}
τ2e+fM2τ2efM2\displaystyle\leq\tau^{2}\|e+f\|_{M}^{2}-\tau^{-2}\|e-f\|_{M}^{2}
=(τ2τ2)(eM2+fM2)+2(τ2+τ2)e,Mf\displaystyle=(\tau^{2}-\tau^{-2})(\|e\|_{M}^{2}+\|f\|_{M}^{2})+2(\tau^{2}+\tau^{-2})\langle e,Mf\rangle
((τ2τ2)(1+κ(M)2)+α(τ2+τ2))eM2\displaystyle\leq((\tau^{2}-\tau^{-2})(1+\kappa(M)^{2})+\alpha(\tau^{2}+\tau^{-2}))\|e\|_{M}^{2}
((τ41)(1+κ(M)2)+α(τ4+1))eM2.\displaystyle\leq((\tau^{4}-1)(1+\kappa(M)^{2})+\alpha(\tau^{4}+1))\|e\|_{M^{\prime}}^{2}.

In the fourth line we used Lemma 2, which implies that fM=λ2(M)κ(M)λ1(M)=κ(M)eM\|f\|_{M}=\lambda_{2}(M)\leq\kappa(M)\lambda_{1}(M)=\kappa(M)\|e\|_{M}, and Lemma 3 to bound 2e,Mf2\langle e,Mf\rangle. Replacing α\alpha and τ\tau with their assumed upper bounds, we obtain 2e,Mf<eM22\langle e,M^{\prime}f\rangle<\|e\|_{M^{\prime}}^{2}. Proceeding likewise, we obtain 2|e,Mf|<min{eM2,fM2}2|\langle e,M^{\prime}f\rangle|<\min\{\|e\|_{M^{\prime}}^{2},\|f\|_{M^{\prime}}^{2}\}. We may therefore apply Lemma 10 to MM^{\prime} and (e,f,e,f)(e,f,-e,-f), which implies {e,f}{e,f,e,f}\{e^{\prime},f^{\prime}\}\subset\{e,f,-e,-f\} as announced. ∎

3.3 Comparison of the stencils

We assume in this subsection that the scale parameter hh is sufficiently small. Our assumption is stronger than the one used in §3.1, see (30), hence in particular there exists an Anisotropic Delaunay Triangulation 𝒯h{\mathcal{T}}_{h}. More precisely we assume that

τh1+1/(3𝜿2)4 and θhθ0:=1/(4𝜿).\tau_{h}\leq\sqrt[4]{1+1/(3{\boldsymbol{\kappa}}^{2})}\ \text{ and }\ \theta_{h}\leq\theta_{0}:=1/(4{\boldsymbol{\kappa}}). (45)

See (26), (29), and Lemma 7 for the definition of 𝜿{\boldsymbol{\kappa}}, τh\tau_{h} and θh\theta_{h} respectively. For Lipschitz metrics, τh=1+𝒪(h)\tau_{h}=1+{\mathcal{O}}(h) and θh=𝒪(h)\theta_{h}={\mathcal{O}}(h).

Our objective is to compare the stencils V(p)V(p), Vh(p)V_{h}(p), of the AD-LBR (10) and of the ADT finite element discretization (17) respectively, at a point ph2p\in h{\mathbb{Z}}^{2}. The next lemma shows that they are equal unless the lattice 2{\mathbb{Z}}^{2} is almost orthogonal with respect to the local metric; a property quantified via μ(𝐌(p))\mu(\operatorname{\mathbf{M}}(p)), see (40).

Lemma 13.

Let ph2p\in h{\mathbb{Z}}^{2}, and let M:=𝐌(p)M:=\operatorname{\mathbf{M}}(p). If μ(M)>θh\mu(M)>\theta_{h}, then Vh(p)=V(p)V_{h}(p)=V(p). In any case, one has for any MM-reduced basis (e,f)(e,f):

Vh(p)\displaystyle V_{h}(p)\supset {e,f,e,f}\displaystyle\{e,f,-e,-f\} (46)
Vh(p)\displaystyle V_{h}(p)\subset {e,f,e+f,ef,e,f,ef,fe}\displaystyle\{e,f,e+f,e-f,\ -e,-f,-e-f,f-e\} (47)
Proof.

We assume that e,Mf0\langle e,Mf\rangle\leq 0, up to replacing ff with f-f. Let T𝒯hT\in{\mathcal{T}}_{h} be a triangle containing pp, and let he1,he2,he3he_{1},he_{2},he_{3} be the edges of TT, oriented so that e1+e2+e3=0e_{1}+e_{2}+e_{3}=0. Using point (iii) of Lemma 7, and (27), we obtain for all 1i31\leq i\leq 3, with the convention e4:=e1e_{4}:=e_{1}

ei,Mei+1θhθ0<12𝜿12min{eiM2,ei+1M2}.\langle e_{i},Me_{i+1}\rangle\leq\theta_{h}\leq\theta_{0}<\frac{1}{2{\boldsymbol{\kappa}}}\leq\frac{1}{2}\min\{\|e_{i}\|_{M}^{2},\|e_{i+1}\|_{M}^{2}\}. (48)

Denote E:={e1,e2,e3}E:=\{e_{1},e_{2},e_{3}\}, and E:={e1,e2,e3}-E:=\{-e_{1},-e_{2},-e_{3}\}. Applying Lemma 10 to MM and the points (e1,e3,e2,e1,e3,e2)(e_{1},-e_{3},e_{2},\linebreak-e_{1},e_{3},-e_{2}), we obtain that {e,f}E(E)\{e,f\}\subset E\cup(-E). Up to exchanging EE with E-E, we thus have {e,f}E\{e,f\}\subset E or {e,f}E\{e,-f\}\subset E. Since the elements of EE sum to zero, we conclude that

E={e,f,ef} or E={e,f,e+f},E=\{e,f,-e-f\}\ \text{ or }\ E=\{e,-f,-e+f\}, (49)

which implies (47).

If μ(M)=|e,Mf|>θh\mu(M)=|\langle e,Mf\rangle|>\theta_{h}, then (48) forbids the second case in (49). Thus E={e,f,ef}E=\{e,f,-e-f\}, and therefore Vh(p)V(p)V_{h}(p)\subset V(p), using (39).

Let T𝒯hT\in{\mathcal{T}}_{h} be a triangle containing pp and intersecting the half line L:={p+re;r>0}L:=\{p+re;\,r>0\}. We know (49) that hehe is a vector edge of TT (i.e. the difference between two vertices of TT). The corresponding edge segment must be [p,p+he][p,p+he], since otherwise TLT\cap L would be empty. Thus eVh(p)e\in V_{h}(p). Applying the same argument to e,f,f-e,f,-f, we obtain (46).

If μ(M)>θh\mu(M)>\theta_{h}, then h(e+f)h(e+f) is also a vector edge of any triangle T𝒯hT\in{\mathcal{T}}_{h} containing pp, since we eliminated the second case in (49). Reasoning as above we find that {e+f,ef}Vh(p)\{e+f,-e-f\}\subset V_{h}(p), and therefore V(p)Vh(p)V(p)\subset V_{h}(p). Thus V(p)=Vh(p)V(p)=V_{h}(p). This concludes the proof. ∎

We introduce new stencils W(p),W(p)W(p),W^{\prime}(p), for p2p\in{\mathbb{R}}^{2}, defined as follows. Let M:=𝐌(p)M:=\operatorname{\mathbf{M}}(p). If μ(M)θ0\mu(M)\leq\theta_{0}, then denoting by (e,f)(e,f) an MM-reduced basis,

W(p)\displaystyle W(p) :={e,f,e,f},\displaystyle:=\{e,f,-e,-f\}, (50)
W(p)\displaystyle W^{\prime}(p) :={e,f,e+f,ef,e,f,ef,fe}.\displaystyle:=\{e,f,e+f,e-f,\ -e,-f,-e-f,f-e\}. (51)

On the other hand, if μ(M)>θ0\mu(M)>\theta_{0}, then

W(p):=V(p)=:W(p).W(p):=V(p)=:W^{\prime}(p). (52)

The previous lemma implies that W(p)Vh(p)W(p)W(p)\subset V_{h}(p)\subset W^{\prime}(p) for any ph2p\in h{\mathbb{Z}}^{2}.

Refer to caption
Figure 6: Consider a point ph2p\in h{\mathbb{Z}}^{2}, and denote M:=𝐌(p)M:=\operatorname{\mathbf{M}}(p). From left to right: ellipse {zM1}\{\|z\|_{M}\leq 1\}, AD-LBR stencil V(p)V(p), stencils W(p)Vh(p)W(p)W(p)\subset V_{h}(p)\subset W^{\prime}(p). For W(p)W(p) and W(p)W^{\prime}(p) we assumed that μ(M)<θ0\mu(M)<\theta_{0}, otherwise they are equal to V(p)V(p). Note that V(p)V(p), W(p)W(p), W(p)W^{\prime}(p) only depend on MM, while Vh(p)V_{h}(p) depends on the structure of the triangulation 𝒯h{\mathcal{T}}_{h}.
Lemma 14.

The stencils W(p)W(p), W(p)W^{\prime}(p), do not depend on the choice of 𝐌(p)\operatorname{\mathbf{M}}(p)-reduced basis.

Proof.

Let M:=𝐌(p)M:=\operatorname{\mathbf{M}}(p). If μ(M)>θ0\mu(M)>\theta_{0}, then W(p)W(p), W(p)W^{\prime}(p) are defined by (52), hence there is nothing to prove. Otherwise we obtain μ(M)θ01/(4𝜿)λ1(M)2/4\mu(M)\leq\theta_{0}\leq 1/(4{\boldsymbol{\kappa}})\leq\lambda_{1}(M)^{2}/4. Hence, by Lemma 12, any two MM-reduced bases (e,f)(e,f), (e,f)(e^{\prime},f^{\prime}), need to satisfy {e,f}{e,f,e,f}\{e^{\prime},f^{\prime}\}\subset\{e,f,-e,-f\}. In view of (50) and (51), they thus yield the same stencils W(p)W(p), W(p)W^{\prime}(p). ∎

Let h,h{\mathcal{F}}_{h},{\mathcal{F}}^{\prime}_{h} be the energies associated to the stencils W,WW,W^{\prime}: for uL2(Ωh)u\in L^{2}(\Omega_{h}), extended to h2h{\mathbb{Z}}^{2} by periodicity,

h(u)\displaystyle{\mathcal{F}}_{h}(u) :=zΩhgW(z)|u(z+hg)u(z)|2,\displaystyle:=\sum_{z\in\Omega_{h}}\sum_{g\in W(z)}|u(z+hg)-u(z)|^{2},
h(u)\displaystyle{\mathcal{F}}^{\prime}_{h}(u) :=zΩhgW(z)|u(z+hg)u(z)|2.\displaystyle:=\sum_{z\in\Omega_{h}}\sum_{g\in W^{\prime}(z)}|u(z+hg)-u(z)|^{2}.

The outline of the proof of Theorem 1 is as follows. We prove in Lemmas 16, 17 and 15 respectively that for any uL2(Ωh)u\in L^{2}(\Omega_{h}):

|h(u)h(u)|\displaystyle|{\mathcal{E}}^{\prime}_{h}(u)-{\mathcal{E}}_{h}(u)| (εh+C0θh)h(u)\displaystyle\leq(\varepsilon_{h}+C_{0}\theta_{h}){\mathcal{F}}^{\prime}_{h}(u) (53)
h(u)\displaystyle{\mathcal{F}}^{\prime}_{h}(u) C1h(u)\displaystyle\leq C_{1}{\mathcal{F}}_{h}(u) (54)
h(u)\displaystyle{\mathcal{F}}_{h}(u) C2h(u),\displaystyle\leq C_{2}{\mathcal{E}}_{h}(u), (55)

where the constants C0,C1,C2C_{0},C_{1},C_{2} only depend on the metric 𝐌\operatorname{\mathbf{M}}. Combining these inequalities, and recalling that θh=𝒪(h)\theta_{h}={\mathcal{O}}(h) and εh=𝒪(h)\varepsilon_{h}={\mathcal{O}}(h) for Lipschitz metrics (εh\varepsilon_{h} is defined in Lemma 9), we obtain

|h(u)h(u)|chh(u),|{\mathcal{E}}^{\prime}_{h}(u)-{\mathcal{E}}_{h}(u)|\leq ch{\mathcal{E}}_{h}(u),

for some constant c=c(𝐌)c=c(\operatorname{\mathbf{M}}). This establishes (13), and concludes the proof of Theorem 1.

For each p2p\in{\mathbb{R}}^{2}, we denote by ηp\eta_{p}, ηp:2{0,1}\eta^{\prime}_{p}:{\mathbb{Z}}^{2}\to\{0,1\}, the characteristic functions of W(p)W(p) and W(p)W^{\prime}(p) respectively. The proofs of (53) and (55) immediately result from the comparison, in Lemmas 16 and 15 respectively, of the coefficients γp\gamma_{p}, γph\gamma^{h}_{p}, ηp\eta_{p}, ηp\eta^{\prime}_{p} appearing in the expressions of h,h,h,h{\mathcal{E}}_{h},{\mathcal{E}}^{\prime}_{h},{\mathcal{F}}_{h},{\mathcal{F}}^{\prime}_{h}.

In the following, it will be convenient to express the AD-LBR weights, and others, in terms of the scalar product associated to the Riemannian metric. We thus recall (8): for any zΩz\in\Omega, and any e,f2e,f\in{\mathbb{R}}^{2},

e,𝐃(z)f=𝐝(z)e,𝐌(z)f.\langle e^{\perp},\operatorname{\mathbf{D}}(z)f^{\perp}\rangle=\operatorname{\mathbf{d}}(z)\langle e,\operatorname{\mathbf{M}}(z)f\rangle.

We also define the bounds (0<𝐝¯𝐝¯<0<\underline{\operatorname{\mathbf{d}}}\leq\overline{\operatorname{\mathbf{d}}}<\infty)

𝐝¯:=minzΩ𝐝(z),𝐝¯:=maxzΩ𝐝(z).\underline{\operatorname{\mathbf{d}}}:=\min_{z\in\Omega}\operatorname{\mathbf{d}}(z),\quad\overline{\operatorname{\mathbf{d}}}:=\max_{z\in\Omega}\operatorname{\mathbf{d}}(z).
Lemma 15.

For any p2p\in{\mathbb{R}}^{2}, one has on 2{\mathbb{Z}}^{2}

ηpC2γp, with C2:=2𝐝¯/θ0.\eta_{p}\leq C_{2}\gamma_{p},\quad\text{ with }C_{2}:=2\overline{\operatorname{\mathbf{d}}}/\theta_{0}.
Proof.

Let M:=𝐌(p)M:=\operatorname{\mathbf{M}}(p), and let (e,f,g)(e,f,g) be an MM-obtuse superbase of 2{\mathbb{Z}}^{2}. We can assume, thanks to Corollary 2, that (e,f)(e,f) is an MM-reduced basis. Then using (22)

2𝐝(p)γp(±f)\displaystyle 2\operatorname{\mathbf{d}}(p)\gamma_{p}(\pm f) 12eM212𝜿=2θ0,\displaystyle\geq\frac{1}{2}\|e\|_{M}^{2}\geq\frac{1}{2{\boldsymbol{\kappa}}}=2\theta_{0},

hence γp(±f)θ0/𝐝¯\gamma_{p}(\pm f)\geq\theta_{0}/\overline{\operatorname{\mathbf{d}}}, and likewise γp(±e)θ0/𝐝¯\gamma_{p}(\pm e)\geq\theta_{0}/\overline{\operatorname{\mathbf{d}}}. If μ(M)θ0\mu(M)\leq\theta_{0}, then W(p)={e,f,e,f}W(p)=\{e,f,-e,-f\}, and this concludes the proof.

Assume now that μ(M)>θ0\mu(M)>\theta_{0}. Then

2𝐝(p)γp(±g)=e,Mf=μ(M)θ0,2\operatorname{\mathbf{d}}(p)\gamma_{p}(\pm g)=-\langle e,Mf\rangle=\mu(M)\geq\theta_{0},

hence γp(±g)θ0/(2𝐝¯)\gamma_{p}(\pm g)\geq\theta_{0}/(2\overline{\operatorname{\mathbf{d}}}). The result follows since W(p)={e,f,g,e,f,g}W(p)=\{e,f,g,-e,-f,-g\}. ∎

Let ph2p\in h{\mathbb{Z}}^{2} and let e1,,eke_{1},\cdots,e_{k} be the consecutive elements of Vh(p)V_{h}(p), in trigonometric order. We define for all 1ik1\leq i\leq k, denoting M:=𝐌(p)M:=\operatorname{\mathbf{M}}(p),

γ~ph(ei):=𝐝(p)4(eiei1,Mei1+eiei+1,Mei+1),\tilde{\gamma}_{p}^{h}(e_{i}):=-\frac{\operatorname{\mathbf{d}}(p)}{4}(\langle e_{i}-e_{i-1},\,Me_{i-1}\rangle+\langle e_{i}-e_{i+1},\,Me_{i+1}\rangle),

with the periodic conventions ek+1:=e1e_{k+1}:=e_{1}, e0:=eke_{0}:=e_{k}. We also set γ~ph=0\tilde{\gamma}_{p}^{h}=0 on 2{e1,,ek}{\mathbb{Z}}^{2}\setminus\{e_{1},\cdots,e_{k}\}.

Lemma 16.

For any ph2p\in h{\mathbb{Z}}^{2}, one has on 2{\mathbb{Z}}^{2}

|γphγ~ph|εhηp, and |γ~phγp|C0θhηp,|\gamma^{h}_{p}-\tilde{\gamma}^{h}_{p}|\leq\varepsilon_{h}\eta^{\prime}_{p},\quad\text{ and }\quad|\tilde{\gamma}^{h}_{p}-\gamma_{p}|\leq C_{0}\theta_{h}\eta^{\prime}_{p}, (56)

where εh\varepsilon_{h} is given in Lemma 9, and C0=1/𝐝¯C_{0}=1/\underline{\operatorname{\mathbf{d}}}.

Proof.

The coefficients γp\gamma_{p}, γph\gamma^{h}_{p}, γ~ph\tilde{\gamma}^{h}_{p}, are all equal to zero outside of W(p)W^{\prime}(p). This holds by construction of γp\gamma_{p}, and by Lemma 13 for γph\gamma^{h}_{p}, γ~ph\tilde{\gamma}^{h}_{p}. We may therefore forget about the presence of ηp\eta^{\prime}_{p} in (56).

First inequality. Lemma 9 states that |γphγ~ph|εh|\gamma^{h}_{p}-\tilde{\gamma}^{h}_{p}|\leq\varepsilon_{h} on 2{\mathbb{Z}}^{2}, which concludes the proof.

Second inequality. If μ(M)>θh\mu(M)>\theta_{h}, then Vh(p)=V(p)V_{h}(p)=V(p). Comparing the definition of γ~ph\tilde{\gamma}^{h}_{p} with that of γp\gamma_{p} (11) we observe that γ~ph=γp\tilde{\gamma}^{h}_{p}=\gamma_{p} on 2{\mathbb{Z}}^{2}, which concludes the proof in this case.

Assume now that μ(M)θh\mu(M)\leq\theta_{h}. Let (e,f,g)(e,f,g) be an MM-obtuse superbase of 2{\mathbb{Z}}^{2}. We can assume, thanks to Corollary 2 that (e,f)(e,f) is an MM-reduced basis. Looking at (11) and denoting δ:=2𝐝(p)\delta:=2\operatorname{\mathbf{d}}(p), we find that

|δγp(±e)fM2|=|e,Mf|=μ(M)θh.|\delta\gamma_{p}(\pm e)-\|f\|_{M}^{2}|=|\langle e,Mf\rangle|=\mu(M)\leq\theta_{h}.

Likewise |δγp(±f)eM2|θh|\delta\gamma_{p}(\pm f)-\|e\|_{M}^{2}|\leq\theta_{h}. In addition

δγp(±(e+f))=μ(M)θh, and γp(±(ef))=0.\delta\gamma_{p}(\pm(e+f))=\mu(M)\leq\theta_{h},\text{ and }\gamma_{p}(\pm(e-f))=0.

Combining the definition of γ~ph\tilde{\gamma}^{h}_{p} with the description of the stencil Vh(p)V_{h}(p) in Lemma 13, we obtain that

2δγ~ph(e)={fe,Mforf+e,Mf}+{fe,Mforf+e,Mf}.2\delta\,\tilde{\gamma}^{h}_{p}(e)=\left\{\begin{array}[]{c}\langle f-e,Mf\rangle\\ \text{or}\\ \langle f+e,Mf\rangle\end{array}\right\}+\left\{\begin{array}[]{c}\langle f-e,Mf\rangle\\ \text{or}\\ \langle f+e,Mf\rangle\end{array}\right\}.

In any case |δγ~ph(e)fM2|θh|\delta\,\tilde{\gamma}^{h}_{p}(e)-\|f\|_{M}^{2}|\leq\theta_{h}. The expressions and estimates of γ~ph\tilde{\gamma}^{h}_{p} at the points e,f,f-e,f,-f are obtained similarly. Likewise, using Lemma 13,

2δγ~ph(e+f)={e,Mf+e,Mf if e+fVh(p),0 otherwise.2\delta\,\tilde{\gamma}^{h}_{p}(e+f)=\left\{\begin{array}[]{ll}\langle e,Mf\rangle+\langle e,Mf\rangle&\text{ if }e+f\in V_{h}(p),\\ 0&\text{ otherwise.}\end{array}\right.

In any case |δγ~ph(e+f)|θh|\delta\,\tilde{\gamma}^{h}_{p}(e+f)|\leq\theta_{h}. The expressions and estimates of γ~ph\tilde{\gamma}_{p}^{h} at the points (e+f),ef,(ef)-(e+f),e-f,-(e-f) are similar. Comparing the above estimates of γp\gamma_{p}, γ~ph\tilde{\gamma}^{h}_{p}, we obtain that δ|γpγ~ph|2θh\delta|\gamma_{p}-\tilde{\gamma}^{h}_{p}|\leq 2\theta_{h} on {e,f,e+f,ef,e,f,ef,fe}=W(p)\{e,f,e+f,e-f,\linebreak-e,-f,-e-f,f-e\}=W^{\prime}(p). Since δ=2𝐝(p)2𝐝¯=2/C0\delta=2\operatorname{\mathbf{d}}(p)\geq 2\underline{\operatorname{\mathbf{d}}}=2/C_{0}, this concludes the proof. ∎

In the last lemma of this section, we control the contribution to the energy h{\mathcal{F}}^{\prime}_{h} of a stencil W(p)W^{\prime}(p), ph2p\in h{\mathbb{Z}}^{2}, in terms of the contributions to h{\mathcal{F}}_{h} of W(p)W(p) and of the neighboring stencils W(p+he)W(p+he), eW(p)e\in W(p). This leads to an estimate of h{\mathcal{F}}^{\prime}_{h} in terms of h{\mathcal{F}}_{h}, which concludes the proof of Theorem 1.

Lemma 17.

One has h(u)C1h(u){\mathcal{F}}^{\prime}_{h}(u)\leq C_{1}{\mathcal{F}}_{h}(u), for any uL2(Ωh)u\in L^{2}(\Omega_{h}), with C1:=17C_{1}:=17.

Proof.

Consider a grid point ph2p\in h{\mathbb{Z}}^{2}, and denote M:=𝐌(p)M:=\operatorname{\mathbf{M}}(p). Assume first that μ(M)θ0\mu(M)\leq\theta_{0}, so that W(p)W(p)W(p)\subsetneq W^{\prime}(p). Consider also an arbitrary gW(p)W(p)g\in W^{\prime}(p)\setminus W(p), and observe that g=e+fg=e+f for some MM-reduced basis (e,f)(e,f).

We set p:=p+ep^{\prime}:=p+e and M:=𝐌(p)M^{\prime}:=\operatorname{\mathbf{M}}(p^{\prime}). Applying point (ii) of Lemma 12, we find that (e,f)(e,f) is also an MM^{\prime}-reduced basis. Indeed we have as required

4μ(M)4θ0=𝜿1λ1(M)2,4\mu(M)\leq 4\theta_{0}={\boldsymbol{\kappa}}^{-1}\leq\lambda_{1}(M)^{2},

using (27), and the assumption on τ\tau follows from (45) and (28). Therefore

fW(p), and h1(pp)=eW(p).f\in W(p^{\prime}),\text{ and }h^{-1}(p^{\prime}-p)=e\in W(p^{\prime}). (57)

We obtain

|u(p+g)u(p)|2\displaystyle|u(p+g)-u(p)|^{2} (58)
=|u(p+e+f)u(p)|2\displaystyle=|u(p+e+f)-u(p)|^{2}
2(|u(p+e+f)u(p+e)|2+|u(p+e)u(p)|2)\displaystyle\leq 2(|u(p+e+f)-u(p+e)|^{2}+|u(p+e)-u(p)|^{2})
=2(|u(p+f)u(p)|2+|u(p+e)u(p)|2).\displaystyle=2(|u(p^{\prime}+f)-u(p^{\prime})|^{2}+|u(p+e)-u(p)|^{2}).

Denote, for all qh2q\in h{\mathbb{Z}}^{2},

h(u;q)\displaystyle{\mathcal{F}}_{h}(u;q) :=gW(q)|u(q+hg)u(q)|2,\displaystyle:=\sum_{g\in W(q)}|u(q+hg)-u(q)|^{2},
h(u;q)\displaystyle{\mathcal{F}}^{\prime}_{h}(u;q) :=gW(q)|u(q+hg)u(q)|2.\displaystyle:=\sum_{g\in W^{\prime}(q)}|u(q+hg)-u(q)|^{2}.

Using (58), we obtain

h(u;p)h(u;p)𝒢h(u;p){\mathcal{F}}^{\prime}_{h}(u;p)-{\mathcal{F}}_{h}(u;p)\leq{\mathcal{G}}_{h}(u;p) (59)

where 𝒢h(u;p){\mathcal{G}}_{h}(u;p) is given by

{4h(u;p)+2gW(p)h(u;p+g), if μ(𝐌(p))θ00, if μ(𝐌(p))>θ0\left\{\begin{array}[]{ll}\displaystyle 4{\mathcal{F}}_{h}(u;p)+2\sum_{g\in W(p)}{\mathcal{F}}_{h}(u;p+g),&\text{ if }\mu(\operatorname{\mathbf{M}}(p))\leq\theta_{0}\\ 0,&\text{ if }\mu(\operatorname{\mathbf{M}}(p))>\theta_{0}\end{array}\right.

When h(u;p){\mathcal{F}}_{h}(u;p^{\prime}) appears in 𝒢h(u;p){\mathcal{G}}_{h}(u;p), with p,ph2p,p^{\prime}\in h{\mathbb{Z}}^{2}, ppp\neq p^{\prime}, we have h1(pp)W(p)h^{-1}(p^{\prime}-p)\in W(p^{\prime}), see (57). For each ph2p^{\prime}\in h{\mathbb{Z}}^{2}, there are thus at most #(W(p))6\#(W(p^{\prime}))\leq 6 points ph2{p}p\in h{\mathbb{Z}}^{2}\setminus\{p^{\prime}\} such that h(u;p){\mathcal{F}}_{h}(u;p^{\prime}) appears in 𝒢h(u;p){\mathcal{G}}_{h}(u;p). Summing (59) over pΩhp\in\Omega_{h}, we thus obtain h(u)h(u)(4+2×6)h(u){\mathcal{F}}^{\prime}_{h}(u)-{\mathcal{F}}_{h}(u)\leq(4+2\times 6){\mathcal{F}}_{h}(u) (the constant could easily be improved), which concludes the proof. ∎

4 Numerical experiments

We compare our scheme AD-LBR with a family of other schemes: finite difference, finite elements, and two schemes from the image processing literature. We begin with a quantitative comparison for the discretization of the restoration equation, in a synthetic case where the exact solution is analytically available for reference. The second test case is a qualitative comparison of Coherence-Enhancing Diffusion (CED) [25], on a real image and the quality assessment is by visual inspection. Finally we present a 3D implementation of AD-LBR for proof of feasibility, featuring a synthetic CED experiment, and a application of Edge Enhancing Diffusion to MRI data.

4.1 The different schemes

Our two dimensional numerical experiments feature the following six numerical schemes for anisotropic diffusion.

  1. AD-LBR: the scheme presented in this work.

  2. Finite Differences (FD). The gradient and the divergence are discretized using standard centered finite differences [15], see Remark 3 for details. This approach, arguably the most straightforward, leads to a 9 point stencil.

  3. Bilinear Finite Elements (Q1). Bilinear finite elements, also referred to as Q1 finite elements, are linear with respect to each space direction. This amounts to use a 9 points stencil, where the coefficients are different from the previous scheme.

  4. Weickert-Scharr scheme (WS). This scheme, introduced in [27], is based on a second order approximation of the gradient using a 3×33\times 3 centered stencil. As a result, it offers good accuracy and rotation invariance when applied to sufficiently smooth functions, but lacks robustness guarantees such as the discrete maximum principle and spectral correctness (see §1), even for 𝐃=Id\operatorname{\mathbf{D}}={\rm Id}. The stencil for this scheme has size 5×55\times 5.

  5. Weickert’s Non-Negative scheme (W-NN). The coefficients of this scheme, detailed in [25] page 95, are non-negative as long as the anisotropy ratio (9) satisfies κ1+22.41\kappa\leq 1+\sqrt{2}\sim 2.41.

  6. Axes-directed Non-Negative scheme (A-NN). This six point non-negative scheme is implicitly defined in the proof Theorem 6 in [25], and can be regarded as a generalisation of W-NN. See Remark 1 below for details. Among the 6 points of the stencil, 4 points are along the axes of coordinates.

Note that other schemes exist, see for instance [16, 28]. While an exhaustive comparison is in principle desirable, it could not be done here due to time and space constraints.

To fix the ideas and illustrate the difference between the schemes, we propose to compute the stencil and the coefficients for different constant diffusion tensors 𝐃\operatorname{\mathbf{D}}, in isotropic and anisotropic cases. Denoting by RR the matrix of rotation by the angle θ=π/6\theta=\pi/6, and by κ1\kappa\geq 1 the chosen anisotropy ratio, we set, identically on 2{\mathbb{R}}^{2}:

𝐃:=R(100κ2)RT.\operatorname{\mathbf{D}}:=R\left(\begin{array}[]{cc}1&0\\ 0&\kappa^{-2}\end{array}\right)R^{\mathrm{T}}. (60)

The results are presented in Tables 1 and 2. Note that for the two last cases (anisotropy κ=10\kappa=\sqrt{10} and κ=50\kappa=\sqrt{50}) the AD-LBR stencil contains points that are outside the 3×33\times 3 neighborhood of the pixel. However the stencil contains 6 points, as expected. This contrasts with the schemes FD, Q1, W-NN where only the 3×33\times 3 neighborhood is involved. Another observation is that the off-center stencil coefficients of the AD-LBR are non-positive (this gives non-negative off-diagonal coefficients for div(𝐃)\operatorname{div}({\bf D}\nabla)), in contrast with schemes FD, Q1, WS, and with scheme W-NN for anisotropy κ>1+2\kappa>1+\sqrt{2}. This is an essential property of AD-LBR (and A-NN), and as a consequence our scheme satisfies, unconditionally, the discrete maximum principle [1, 6].

The largest eigenvalue of the discrete operatordiv(𝐃)-\operatorname{div}({\bf D}\nabla) is given in Table 3, for the different schemes. It turns out that AD-LBR has in most cases the smallest eigenvalues among all schemes, except for scheme WS and occasionally A-NN. This property allows (although this was not done in our numerical experiments) to use larger time steps for AD-LBR than for the other schemes, when solving parabolic equations (2) or (64) with an explicit time discretization.

Operator splitting is a classical approach to further increase the timestep in (potentially anisotropic and non-linear) diffusion PDEs [25, 26, 2]. The AD-LBR is compatible with Additive Operator Splitting, by applying Remark (e) page 111 in [25], although the efficiency of this technique is here compromised by the potentially large number of directions in our adaptive stencils. Let us also mention Multiplicative Operator Splittings, and Additive-Multiplicative Operator Splittings, which allow to combine different time-steps [2, 8]. None of these methods was used in our experiments.

Remark 1 (Axes-directed non negative six point scheme).

The following six point scheme A-NN is, in our belief, the best possible implementation of the constructive proof in [25] of the existence of non-negative schemes for two dimensional anisotropic diffusion. Like AD-LBR, this scheme is defined by the data at each point zΩz\in\Omega of a stencil V(z)V(z), and of non-negative weights γz\gamma_{z}.

Let 𝐃(z)=(abbc).\operatorname{\mathbf{D}}(z)=\left(\begin{array}[]{cc}a&b\\ b&c\end{array}\right). In the diagonal case b=0b=0, the scheme A-NN relies on the classical four points stencil. Otherwise note that

a|b||b|c=acb2|b|c>0.\frac{a}{|b|}-\frac{|b|}{c}=\frac{ac-b^{2}}{|b|c}>0.

Let p,q{0}p,q\in{\mathbb{Z}}\setminus\{0\} be such that bpq0bpq\geq 0,

|b|c|pq|a|b|,\frac{|b|}{c}\leq\left|\frac{p}{q}\right|\leq\frac{a}{|b|}, (61)

and max(|p|,|q|)\max(|p|,|q|) is minimal. The scheme A-NN is defined by the six point stencil

V(z):={(±1,0),(0,±1),±(p,q)}V(z):=\{(\pm 1,0),(0,\pm 1),\pm(p,q)\}

and the non-negative weights

2γ(±1,0)\displaystyle 2\gamma(\pm 1,0) =apqb,\displaystyle=a-\frac{p}{q}b, 2γ(0,±1)\displaystyle 2\gamma(0,\pm 1) =cqpb,\displaystyle=c-\frac{q}{p}b,
2γ(±(p,q))\displaystyle 2\gamma(\pm(p,q)) =bpq.\displaystyle=\frac{b}{pq}.

These coefficients are non-negative by construction, and consistency (5) is easily checked. Contrary to AD-LBR, the coordinate axes play a privileged role in A-NN. This introduces axis aligned artifacts which are visible in Figure 12 (g).

Remark 2 (Stencil radius).

The two dimensional stencils of AD-LBR coincide with those of FM-LBR, a numerical scheme for anisotropic static Hamilton-Jacobi PDEs introduced in [14] the second author. As shown in Proposition 1.6 of [14], the euclidean radius

r=max{v;vV(z)}r=\max\{\|v\|;v\in V(z)\}

of this stencil is bounded by κ(𝐃(z))\kappa(\operatorname{\mathbf{D}}(z)).

In contrast, consider for 0<ε<1/40<\varepsilon<1/4 the matrix

D:=(112ε12ε13ε).D:=\left(\begin{array}[]{cc}1&1-2\varepsilon\\ 1-2\varepsilon&1-3\varepsilon\end{array}\right).

If follows from (61) that 1+εp/q+𝒪(ε2)1+2ε1+\varepsilon\leq p/q+{\mathcal{O}}(\varepsilon^{2})\leq 1+2\varepsilon. From this point, one easily obtains that qε1κ(D)2q\gtrsim\varepsilon^{-1}\approx\kappa(D)^{2}. The radius of the A-NN stencil, at a point zΩz\in\Omega, may thus be of the order of κ(𝐃(z))2\kappa(\operatorname{\mathbf{D}}(z))^{2}. The radii of the AD-LBR and A-NN stencils, computed for diffusion tensors of anisotropy κ=10\kappa=10 and of various orientations, are illustrated on Figure 7.

Refer to caption
Figure 7: Radius of the AD-LBR stencil (plain), and of the A-NN stencil (dashed), for a matrix DθD_{\theta} of anisotropy ratio κ=10\kappa=10 and eigenvector (cosθ,sinθ)(\cos\theta,\sin\theta). The AD-LBR stencil is here always the smallest, and its radius does not exceed 5.15.1, versus 36.136.1 for A-NN.
Refer to caption D=(19/2664)D=\left(\begin{array}[]{cc}19/2&6\\ 6&4\end{array}\right)
Figure 8: The seven smallest eigenvalues of the operator div(D)-\operatorname{div}(D\nabla), on [0,1]2[0,1]^{2} with periodic boundary conditions, discretized on a n×nn\times n grid with 4n124\leq n\leq 12. Plain: AD-LBR discretization; Dashed: A-NN discretization. Some eigenvalues have multiplicities, hence less than 14 graphs are visible. Eigenvalues of the A-NN discretization here “oscillate” with the dimension. In contrast, thanks to its asymptotic equivalence with a finite element scheme, the smallest eigenvalues of the AD-LBR discretization converge towards those of the continuous partial differential operator.
Remark 3 (Scheme FD).

The operator div(𝐃)\operatorname{div}(\operatorname{\mathbf{D}}\nabla\cdot) is discretized using centered finite differences [15]. This involves quantities defined at half integer indices, and in particular the diffusion tensor is here given on the offsetted grid (i+1/2,j+1/2)(i+1/2,\,j+1/2), (i,j)2(i,j)\in{\mathbb{Z}}^{2}. For the sake of readability, we thus define i+:=i+1/2i^{+}:=i+1/2 and i:=i1/2i^{-}:=i-1/2. The gradient operator is discretized by:

(xu)i+,j=ui+1,jui,j,(yu)i,j+=ui,j+1ui,j.(\partial_{x}u)_{i^{+},j}=u_{i+1,j}-u_{i,j},\quad(\partial_{y}u)_{i,j^{+}}=u_{i,j+1}-u_{i,j}.

The divergence is defined as follows:

div(𝐃u)i,j\displaystyle\operatorname{div}({\bf D}\nabla u)_{i,j} =x(𝐃11xu+𝐃12yu)i,j\displaystyle=\partial_{x}({\bf D}^{11}\partial_{x}u+{\bf D}^{12}\partial_{y}u)_{i,j}
+y(𝐃21xu+𝐃22yu)i,j,\displaystyle+\partial_{y}({\bf D}^{21}\partial_{x}u+{\bf D}^{22}\partial_{y}u)_{i,j},

with

(𝐃11xu)i+,j\displaystyle({\bf D}^{11}\partial_{x}u)_{i^{+},j} =12(𝐃i+,j+11+𝐃i+,j11)(xu)i+,j,\displaystyle=\frac{1}{2}\left({\bf D}^{11}_{i^{+},j^{+}}+{\bf D}^{11}_{i^{+},j^{-}}\right)(\partial_{x}u)_{i^{+},j},
x(𝐃11xu)i,j\displaystyle\partial_{x}({\bf D}^{11}\partial_{x}u)_{i,j} =(𝐃11xu)i+,j(𝐃11xu)i,j\displaystyle=({\bf D}^{11}\partial_{x}u)_{i^{+},j}-({\bf D}^{11}\partial_{x}u)_{i^{-},j}
(𝐃21xu)i+,j+\displaystyle({\bf D}^{21}\partial_{x}u)_{i^{+},j^{+}} =12𝐃i+,j+21((xu)i+,j+(xu)i+,j+1),\displaystyle=\dfrac{1}{2}{\bf D}^{21}_{i^{+},j^{+}}\left((\partial_{x}u)_{i^{+},j}+(\partial_{x}u)_{i^{+},j+1}\right),
y(𝐃21xu)i,j\displaystyle\partial_{y}({\bf D}^{21}\partial_{x}u)_{i,j} =12((𝐃21xu)i+,j+(𝐃21xu)i+,j\displaystyle=\dfrac{1}{2}\left(({\bf D}^{21}\partial_{x}u)_{i^{+},j^{+}}-({\bf D}^{21}\partial_{x}u)_{i^{+},j^{-}}\right.
+(𝐃21xu)i,j+(𝐃21xu)i,j),\displaystyle\phantom{=}\left.+({\bf D}^{21}\partial_{x}u)_{i^{-},j^{+}}-({\bf D}^{21}\partial_{x}u)_{i^{-},j^{-}}\right),

and similar terms involving yu\partial_{y}u.

Table 1: The stencil coefficients for different constant diffusion tensors, and the different schemes presented. The value of the anisotropy ratio κ\kappa is given in the second row, and the orientation of the principal axis is θ=π/6\theta=\pi/6, see (60). The bold coefficient indicates the center node. In some examples we present for clarity reasons only half of the stencil (the other half can be deduced by symmetry). Stencil entries are highlighted when they are positive and off-center - an undesirable property which gives rise to stability issues. For small anisotropies, κ1+2\kappa\leq 1+\sqrt{2}, one has AD-LBR = W-NN = A-NN.
κ\kappa κ=1\kappa=1 (𝐃=Id\operatorname{\mathbf{D}}={\rm Id}) κ=2\kappa=\sqrt{2}
stencil for AD-LBR 0101𝟒1010\begin{array}[]{ccc}0&-1&0\\ -1&{\bf 4}&-1\\ 0&-1&0\end{array} 00.410.220.662.570.660.220.410\begin{array}[]{ccc}0&-0.41&-0.22\\ -0.66&{\bf 2.57}&-0.66\\ -0.22&-0.41&0\end{array}
stencil for FD 0101𝟒1010\begin{array}[]{ccc}0&-1&0\\ -1&{\bf 4}&-1\\ 0&-1&0\end{array} 0.110.630.110.88𝟑0.880.110.630.11\begin{array}[]{ccc}{\color[rgb]{0.72,0,0}0.11}&-0.63&-0.11\\ -0.88&{\bf 3}&-0.88\\ -0.11&-0.63&{\color[rgb]{0.72,0,0}0.11}\end{array}
stencil for Q1 13(1111𝟖1111)\dfrac{1}{3}\left(\begin{array}[]{ccc}-1&-1&-1\\ -1&{\bf 8}&-1\\ -1&-1&-1\end{array}\right) 0.140.130.360.38𝟐0.380.360.130.14\begin{array}[]{ccc}-0.14&-0.13&-0.36\\ -0.38&{\bf 2}&-0.38\\ -0.36&-0.13&-0.14\end{array}
stencil for WS 0.10.060.020.1200.060.460.120.10.1200.060.10.060.02\begin{array}[]{ccc}-0.1&-0.06&-0.02\\ {\color[rgb]{0.72,0,0}0.12}&0&-0.06\\ {\bf 0.46}&{\color[rgb]{0.72,0,0}0.12}&-0.1\\ {\color[rgb]{0.72,0,0}0.12}&0&-0.06\\ -0.1&-0.06&-0.02\end{array} 0.060.050.020.010.040.060.350.070.090.010.040.040.060.020.01\begin{array}[]{ccc}-0.06&-0.05&-0.02\\ {\color[rgb]{0.72,0,0}0.01}&-0.04&-0.06\\ {\bf 0.35}&{\color[rgb]{0.72,0,0}0.07}&-0.09\\ {\color[rgb]{0.72,0,0}0.01}&{\color[rgb]{0.72,0,0}0.04}&-0.04\\ -0.06&-0.02&-0.01\end{array}
stencil for W-NN 0101𝟒1010\begin{array}[]{ccc}0&-1&0\\ -1&{\bf 4}&-1\\ 0&-1&0\end{array} 00.410.220.662.570.660.220.410\begin{array}[]{ccc}0&-0.41&-0.22\\ -0.66&{\bf 2.57}&-0.66\\ -0.22&-0.41&0\end{array}
stencil for A-NN 0101𝟒1010\begin{array}[]{ccc}0&-1&0\\ -1&{\bf 4}&-1\\ 0&-1&0\end{array} 00.410.220.662.570.660.220.410\begin{array}[]{ccc}0&-0.41&-0.22\\ -0.66&{\bf 2.57}&-0.66\\ -0.22&-0.41&0\end{array}
Table 2: The stencil coefficients for different metrics and the different schemes presented, similarly to Table 1 but with more pronounced anisotropies. For the scheme A-NN some points of the stencil are too far from the center node to be represented here, so we indicate the coordinates of these points and the associated coefficient.
κ\kappa κ=10\kappa=\sqrt{10} κ=50\kappa=\sqrt{50}
stencil for AD-LBR 00.260.061.160.260\begin{array}[]{ccc}0&-0.26&-0.06\\ {\bf 1.16}&-0.26&0\end{array} 00.110.160.550.010\begin{array}[]{ccc}0&-0.11&-0.16\\ {\bf 0.55}&-0.01&0\end{array}
stencil for FD 0.190.320.190.772.20.770.190.320.19\begin{array}[]{ccc}{\color[rgb]{0.72,0,0}0.19}&-0.32&-0.19\\ -0.77&{\bf 2.2}&-0.77\\ -0.19&-0.32&{\color[rgb]{0.72,0,0}0.19}\end{array} 0.210.270.210.762.040.760.210.270.21\begin{array}[]{ccc}{\color[rgb]{0.72,0,0}0.21}&-0.27&-0.21\\ -0.76&{\bf 2.04}&-0.76\\ -0.21&-0.27&{\color[rgb]{0.72,0,0}0.21}\end{array}
stencil for Q1 0.010.040.380.411.470.410.380.040.01\begin{array}[]{ccc}{\color[rgb]{0.72,0,0}0.01}&{\color[rgb]{0.72,0,0}0.04}&-0.38\\ -0.41&{\bf 1.47}&-0.41\\ -0.38&{\color[rgb]{0.72,0,0}0.04}&{\color[rgb]{0.72,0,0}0.01}\end{array} 0.040.080.380.421.360.420.380.080.04\begin{array}[]{ccc}{\color[rgb]{0.72,0,0}0.04}&{\color[rgb]{0.72,0,0}0.08}&-0.38\\ -0.42&{\bf 1.36}&-0.42\\ -0.38&{\color[rgb]{0.72,0,0}0.08}&{\color[rgb]{0.72,0,0}0.04}\end{array}
stencil for WS 0.020.040.020.090.080.070.250.040.080.090.080.020.020.0040.003\begin{array}[]{ccc}-0.02&-0.04&-0.02\\ {\color[rgb]{0.72,0,0}0.09}&-0.08&-0.07\\ {\bf 0.25}&{\color[rgb]{0.72,0,0}0.04}&-0.08\\ {\color[rgb]{0.72,0,0}0.09}&{\color[rgb]{0.72,0,0}0.08}&-0.02\\ -0.02&{\color[rgb]{0.72,0,0}0.004}&-0.003\end{array} 0.020.040.020.090.080.070.240.030.080.090.080.020.020.010.002\begin{array}[]{ccc}-0.02&-0.04&-0.02\\ {\color[rgb]{0.72,0,0}0.09}&-0.08&-0.07\\ {\bf 0.24}&{\color[rgb]{0.72,0,0}0.03}&-0.08\\ {\color[rgb]{0.72,0,0}0.09}&{\color[rgb]{0.72,0,0}0.08}&-0.02\\ -0.02&{\color[rgb]{0.72,0,0}0.01}&-0.002\end{array}
stencil for W-NN 00.060.390.391.420.390.390.060\begin{array}[]{ccc}0&{\color[rgb]{0.72,0,0}0.06}&-0.39\\ -0.39&{\bf 1.42}&-0.39\\ -0.39&{\color[rgb]{0.72,0,0}0.06}&0\end{array} 00.160.420.331.190.330.420.160\begin{array}[]{ccc}0&{\color[rgb]{0.72,0,0}0.16}&-0.42\\ -0.33&{\bf 1.19}&-0.33\\ -0.42&{\color[rgb]{0.72,0,0}0.16}&0\end{array}
stencil for A-NN 0.0700.640.19\begin{array}[]{ccc}-0.07&0\\ {\bf 0.64}&-0.19\end{array} 0.0100.170.05\begin{array}[]{ccc}-0.01&0\\ {\bf 0.17}&-0.05\end{array}
γ(3,2)=0.06\gamma(3,2)=-0.06 γ(5,3)=0.03\gamma(5,3)=-0.03
Table 3: Largest eigenvalue of the discretized operator div(𝐃)-\operatorname{div}({\bf D}\nabla), for the constant metric 𝐃=D\operatorname{\mathbf{D}}=D, where the matrix DD is given on Tables 1 and 2. The time step, in the explicit discretization of (64), should not exceed the inverse of this value.
κ\kappa κ=1\kappa=1 κ=2\kappa=\sqrt{2} κ=10\kappa=\sqrt{10} κ=50\kappa=\sqrt{50}
eigenvalue AD-LBR 8 4.27 2.06 1.06
eigenvalue FD 8 6.22 5.06 4.85
eigenvalue Q1 5.7 4.94 4.32 4.20
eigenvalue WS 1 1 1 1
eigenvalue W-NN 8 4.27 3.1 3.02
eigenvalue A-NN 8 4.27 1.04 0.3

4.2 A test case with an explicit solution

Consider an image vL2(Ω)v\in L^{2}(\Omega), defined on a domain Ω\Omega, and a diffusion tensor field 𝐃:ΩS2+{\bf D}:\Omega\to S_{2}^{+}. A classical approach to restore the image vv, if it has been corrupted by additive noise, is to find uH1(Ω)u\in H^{1}(\Omega) which minimizes:

j(u)=Ω|uv|2+λΩu𝐃2.j(u)=\int_{\Omega}|u-v|^{2}+\lambda\int_{\Omega}||\nabla u||^{2}_{{\bf D}}. (62)

In other words, uu is a penalized least squares approximation of vv. The parameter λ>0\lambda>0 should be adjusted so as to avoid excessive smoothing (for large λ\lambda), or insufficient denoising (for small λ\lambda). The solution uu can be characterized as the solution to the static elliptic PDE:

{λdiv(𝐃u)+u=v, on Ω.u,n=0, on Ω.\left\{\begin{array}[]{ll}-\lambda\operatorname{div}({\bf D}\nabla u)+u=v,&\text{ on }\Omega.\\ \langle\nabla u,n\rangle=0,&\text{ on }\partial\Omega.\end{array}\right. (63)

In applications [20, 24] the diffusion tensor 𝐃\operatorname{\mathbf{D}} is usually adapted to the local image structure, in order to avoid smoothing the edges of vv. We construct below a test case (image vv and tensor field 𝐃{\bf D}), for which the solution uu is known analytically.

In order to obtain an analytic solution, we first consider a separable problem where the image is invariant by translation along the horizontal axis, and the metric is constant with axes parallel to the coordinate axes. This first problem is invariant under translations along the xx-axis, and therefore boils down to a 1-dimensional problem. This separable problem is then transported by a diffeomorphism in order to obtain a new problem where the axes of the metric are no more parallel to the coordinate axes.

The analytical image is composed of a black and a white stripe: v0(x,y)=𝟏y<0.5v_{0}(x,y)={\bf 1}_{y<0.5}, see Figure 9. Given κ1\kappa\geq 1, we consider the constant diffusion tensor

𝐃0=(100κ2).{\bf D}_{0}=\left(\begin{array}[]{cc}1&0\\ 0&\kappa^{-2}\end{array}\right).

The analytical solution u0u_{0} of (63), applied to 𝐃0\operatorname{\mathbf{D}}_{0} and v0v_{0}, is known in the case of the infinite domain Ω=2\Omega={\mathbb{R}}^{2}. In Fourier domain all the coefficients are real and:

u0^(ξ)=v0^(ξ)/(1+ξ,𝐃0ξ).\widehat{u_{0}}(\xi)=\widehat{v_{0}}(\xi)/(1+\langle\xi,{\bf D}_{0}\xi\rangle).

This separable problem is transformed using the following diffeomorphism: for (x,y)Ω(x,y)\in\Omega

f(x,y)=(x,y+αcos(2πx)).f(x,y)=(x,y+\alpha\cos(2\pi x)).

The Jacobian of ff is

J(x,y)=(102παsin(2πx)1)J(x,y)=\left(\begin{array}[]{cc}1&0\\ -2\pi\alpha\sin(2\pi x)&1\end{array}\right)

We apply the different restoration schemes to the image v=v0fv=v_{0}\circ f, and the following diffusion tensor:

𝐃(z)\displaystyle{\bf D}(z) =|detJ(z)|J(z)1𝐃0(J(z)1)T\displaystyle=|\det J(z)|\ J(z)^{-1}\,{\bf D}_{0}\,(J(z)^{-1})^{\mathrm{T}}
=J(z)1𝐃0(J(z)1)T=(1sss2+κ2),\displaystyle=J(z)^{-1}\,{\bf D}_{0}\,(J(z)^{-1})^{\mathrm{T}}=\left(\begin{array}[]{cc}1&s\\ s&s^{2}+\kappa^{-2}\end{array}\right),

where we denoted z=(x,y)Ωz=(x,y)\in\Omega and s=2παsin(2πx)s=2\pi\alpha\sin(2\pi x). The numerical solution is compared to the analytical function u=u0fu=u_{0}\circ f, which is the exact solution in the case of the infinite domain Ω=2\Omega={\mathbb{R}}^{2}. This numerical solution was obtained on the bounded domain Ω=[0,1[2\Omega=[0,1[^{2}, equipped with reflecting boundary conditions. Numerical evidence suggests that this change of domain and of boundary conditions has only an anecdotic impact on the solution of (63), with the parameters chosen in this test case.

We used α:=1/3\alpha:=1/3 in the numerical experiments. The maximum value of κ(𝐃(x))\kappa(\operatorname{\mathbf{D}}(x)), among all xΩx\in\Omega, is equivalent to κmax:=κ1+(2πα)22.3κ\kappa_{\max}:=\kappa\sqrt{1+(2\pi\alpha)^{2}}\simeq 2.3\kappa.

Refer to caption
Refer to caption
Figure 9: Left: image vanav_{\rm ana}. Right: image v=vanafv=v_{\rm ana}\circ f transformed by the diffeomorphism ff.

4.3 Results for the synthetic test case

We present in Figure 10 the performance results of the different schemes, for different values of the anisotropy κ\kappa, obtained on a series of grids of size ranging from 100×100100\times 100 to 1200×12001200\times 1200. The anisotropy varies from κ=2\kappa=2 to κ=10\kappa=10, which are relevant values for imaging applications, see the numerical experiments in §4.4. The quality of a scheme is measured by the L2L^{2} difference and the H1H^{1} semi-norm difference between the numerical solution and the analytical solution. Note that the error is concentrated close to the discontinuity, since the solution tends rapidly to a constant (0 or 1) far from the discontinuity. We chose the smoothing parameter λ=103\lambda=10^{-3} in (62). The linear equation obtained by the discretization of (63) is solved using Conjugate Gradient.

We also tested extreme anisotropies, κ100\kappa\geq 100 (thus κmax230\kappa_{\max}\geq 230), which can be relevant in physics related applications. None of the tested schemes showed convincing results: methods based on fixed stencils fail because the discrete operator looses positivity, while the AD-LBR (and A-NN even more) suffers from under-sampling due to the large radius of its stencils. We thus refer to [7] for a radically different approach tailored for this setting. This method introduces an auxiliary one-dimensional unknown, which is constant on the field lines (obtained in a preprocessing step) of the anisotropy direction field, and varies orthogonally to them.

The performance advantage of the AD-LBR is particularly clear when the error is measured in the H1H^{1} semi-norm: for the anisotropy κ=10\kappa=10 and the resolution 500×500500\times 500, which are relevant values in image processing, AD-LBR outperforms its alternatives by a factor ranging from 3 to 5.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 10: Numerical results for the synthetic test case, with different values of the anisotropy factor: κ=2\kappa=2 (top row), κ=5\kappa=5 (middle row), κ=10\kappa=10 (bottom row). Vertical axis: relative error in L2L^{2} norm (left column), or H1H^{1} semi-norm (right column), for the six schemes tested. Horizontal axis: integer NN, where the image resolution is N×NN\times N. Since the tested schemes are first order, numerical error is expected to be proportional to N1N^{-1}. Log-log scale.

4.4 Coherence-enhancing diffusion

In order to document the interest of our discretization, we implement Coherence-Enhancing Diffusion [25] using the different numerical schemes at our disposal. The following parabolic equation is considered:

tu=div(𝐃(Jρ(uσ))u).\partial_{t}u=\operatorname{div}({\bf D}(J_{\rho}(\nabla u_{\sigma}))\nabla u). (64)

This equation is non-linear since the diffusion tensor depends on the solution uu. This tensor also depends on four user defined parameters σ,ρ,C+\sigma,\rho,C\in{\mathbb{R}}_{+}, α]0,1[\alpha\in]0,1[. Let KσK_{\sigma} (resp. KρK_{\rho}), be the Gaussian kernel of variance σ\sigma (resp. ρ\rho). Define the convolution uσ:=Kσuu_{\sigma}:=K_{\sigma}\star u, and the structure tensor Jρ:=Kρ(uσuσT)J_{\rho}:=K_{\rho}\star(\nabla u_{\sigma}\nabla u_{\sigma}^{T}). The diffusion tensor 𝐃(Jρ){\bf D}(J_{\rho}) possesses the same eigenvectors (v1,v2)(v_{1},v_{2}) as JρJ_{\rho}, and if the eigenvalues of JρJ_{\rho} are μ1μ2\mu_{1}\geq\mu_{2} then the eigenvalues of 𝐃(Jρ){\bf D}(J_{\rho}) are

λ1\displaystyle\lambda_{1} :=α\displaystyle:=\alpha
λ2\displaystyle\lambda_{2} :=α+(1α)exp(C(μ1μ2)2).\displaystyle:=\alpha+(1-\alpha)\exp\left(\dfrac{-C}{(\mu_{1}-\mu_{2})^{2}}\right).

This ensures that one smoothes preferably along the coherence direction v2v_{2}, with a diffusivity that increases with respect to the coherence (μ1μ2)2(\mu_{1}-\mu_{2})^{2}. When the time parameter tt becomes large, the image tends to a constant image, therefore it is necessary to stop the process at some finite time TT. The ratio of the eigenvalues is bounded by λ2/λ11/α\lambda_{2}/\lambda_{1}\leq 1/\alpha, hence κ1/α\kappa\leq 1/\sqrt{\alpha}.

We used an explicit time discretization for (64), with time step Δt\Delta t. The image un+1u^{n+1} at time (n+1)Δt(n+1)\Delta t is defined by the explicit equation:

un+1unΔt=div(𝐃(Jρ(uσn))un).\dfrac{u^{n+1}-u^{n}}{\Delta t}=\operatorname{div}({\bf D}(J_{\rho}(\nabla u_{\sigma}^{n}))\nabla u^{n}).

The parameters used in our simulation were: σ=0.5\sigma=0.5, ρ=4\rho=4, C=105C=10^{-5}, α=102\alpha=10^{-2} and Δt=0.02\Delta t=0.02. This gives a maximum anisotropy of κ=10\kappa=10. The algorithm was applied to a fingerprint image. The results obtained for T=10T=10 are shown in Figures 11 and 12, and they document the ability of our scheme to close interrupted lines more efficiently than the other schemes. The largest eigenvalue of the discrete operator div(𝐃)-\operatorname{div}({\bf D}\nabla) at t=0t=0 is given in Table 4 for the different schemes. As was already noticed in the constant metric case, it turns out that AD-LBR has the smallest eigenvalues among all schemes, except for scheme WS. This property allows (although this was not done in our numerical experiments) to use larger time steps for AD-LBR than for the other schemes.

Note also that ridges are clearer, and valleys are darker, using AD-LBR than with the other schemes. (Gray-scale range is the same for all images, see also Figure 13). This reflects the fact that AD-LBR avoids, better than the other schemes, smoothing transversally to the orientation encoded in the continuous anisotropic PDE (64).

Remark 4 (Computation time).

Numerical solvers of the parabolic PDE (64) combine three main components: (i) Constructing the diffusion tensor. (ii) Assembling the discretization stencils and the operator sparse matrix. (iii) Performing an explicit time step. Components (i) and (ii) are executed exactly the same number of times, while step (iii) is generally more frequent: in order to save CPU time, one typically does not update the diffusion operator at each time step. We produced a C++ implementation of AD-LBR, within the Insight Toolkit open source library. Although our code is neither parallel nor aggressively optimized, we believe that comparing the CPU times for steps (i), (ii) and (iii) is informative, and allows to estimate the additional cost of AD-LBR which is essentially contained in step (ii).

For our 2D Coherence-Enhancing Diffusion (CED) experiment, on the 512×512512\times 512 fingerprint image, (i) takes 0.21s, (ii) 0.027s, (iii) 0.005s. For our 3D CED Experiment, on 100×100×100100\times 100\times 100 synthetic data, (i) takes 1.35s, (ii) 0.51s, (iii) 0.035s. In both cases, the AD-LBR specific step (ii) is dominated by the construction of the diffusion tensor (i). Step (ii) may also be dominated by the mere cost (iii) of iterations, provided the operator is updated less than once every 6 explicit steps in 2D (14 in 3D). To our eyes, the limited additional cost (ii) of AD-LBR is acceptable in view of the strong theoretical guarantees, and qualitative improvements, brought by this scheme.

Table 4: Largest eigenvalue of the discretized operator div(𝐃)-\operatorname{div}({\bf D}\nabla), where 𝐃=𝐃(Jρ(uσ)){\bf D}={\bf D}(J_{\rho}(\nabla u_{\sigma})) at t=0t=0.
scheme AD-LBR FD Q1 WS W-NN A-NN
eigenvalue 3.75 5.67 5.09 0.96 3.83 6.23
Refer to caption
(a) Original image
Refer to caption
(b) AD-LBR
Refer to caption
(c) FD
Refer to caption
(d) Q1
Refer to caption
(e) WS
Refer to caption
(f) W-NN
Refer to caption
(g) A-NN
Figure 11: From top to bottom and from left to right: Original image (with two regions highlighted); diffused image using AD-LBR; FD; Q1; WS; W-NN; A-NN. Here T=10T=10.
Refer to caption
(a) Original image
Refer to caption
(b) AD-LBR
Refer to caption
(c) FD
Refer to caption
(d) Q1
Refer to caption
(e) WS
Refer to caption
(f) W-NN
Refer to caption
(g) A-NN
Figure 12: Detail of the region on the right. From top to bottom and from left to right: original image; diffused image using AD-LBR; FD; Q1; WS; W-NN; A-NN.
Refer to caption

Refer to caption Refer to caption Refer to caption

Figure 13: Evolution under CED of a section of the fingerprint image. The ridges in the evolved image are higher, and the valleys are deeper, with AD-LBR than with the other schemes. This illustrates the fact that AD-LBR, respecting the continuous PDE, diffuses more along the structure and less in the orthogonal direction. From top to bottom: location of the section of the image; section at T=10T=10; section at T=50T=50; section at T=100T=100.

4.5 3-dimensional experiments

In order to illustrate the feasibility of our scheme in 3D space, we present the action of anisotropic diffusion PDEs on two examples. The first example is a 3D analog of the synthetic test case presented in [27], featuring Coherence-Enhancing Diffusion. The second one is the application of Edge-Enhancing Diffusion to a MRI scan.

Synthetic example

The original, radially varying image is defined on the cube [0,1]3[0,1]^{3}. The gray-level at a point xx is defined by

u0(x)=cos(2(r/R)3),u^{0}(x)=\cos\left(2(r/R)^{3}\right),

where r:=|x|r:=|x| and R:=1/2R:=1/2. This image presents a series of concentric level-sets. We present in Figure 14 the level sets {u0=0}\{u^{0}=0\}, and a slice through the plane z=0.7z=0.7.

The image u0u^{0} is perturbed by

u:=u0+n,u:=u^{0}+n,

where nn is an additive Gaussian noise of variance σ=0.5\sigma=0.5. The reconstructed image is obtained using a 3D Coherence-Enhancing Diffusion PDE [25], similar to the 2D one in section 4.4:

tu=div(𝐃(Jρ(uσ))u),\partial_{t}u=\operatorname{div}({\bf D}(J_{\rho}(\nabla u_{\sigma}))\nabla u),

where JρJ_{\rho} is the structure tensor defined by Jρ:=Kρ(uσuσT)J_{\rho}:=K_{\rho}\star(\nabla u_{\sigma}\nabla u_{\sigma}^{T}), uσ:=Kσuu_{\sigma}:=K_{\sigma}\star u. The tensor 𝐃(Jρ){\bf D}(J_{\rho}) possesses the same eigenvectors (v1,v2,v3)(v_{1},v_{2},v_{3}) as JρJ_{\rho}, and if the eigenvalues of JρJ_{\rho} are μ1μ2μ3\mu_{1}\geq\mu_{2}\geq\mu_{3} then the eigenvalues of 𝐃(Jρ){\bf D}(J_{\rho}) are

λ1:=α\lambda_{1}:=\alpha
λ2:=α+(1α)exp(C(μ1μ2)2),\lambda_{2}:=\alpha+(1-\alpha)\exp\left(\dfrac{-C}{(\mu_{1}-\mu_{2})^{2}}\right),
λ3:=α+(1α)exp(C(μ1μ3)2),\lambda_{3}:=\alpha+(1-\alpha)\exp\left(\dfrac{-C}{(\mu_{1}-\mu_{3})^{2}}\right),

where α=102\alpha=10^{-2}. The anisotropy ratio is bounded by κ=1/α=10\kappa=1/\sqrt{\alpha}=10. We used the values σ=0.5\sigma=0.5, ρ=4\rho=4. The problem is discretized using 1003100^{3} voxels. We present in Figure 14 the noisy image uu (levelset 0 and planar slice) and the result after 20 time-steps of Δt=103\Delta t=10^{-3}.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 14: Levelset 0 (top) and slice (bottom) of a 3D image. Original (left), noisy (center), and reconstructed (right) images. Slice in the plane z=0.7z=0.7, with values clipped to the range [1.3,1.3][-1.3,1.3].

3D MRI data

The data is a 256×256×100256\times 256\times 100 Magnetic Resonance Imaging scan of a skull, and was obtained from the ”University of North Carolina Volume Rendering Test Data Set” archive.

The reconstructed image is obtained using a 3D Edge-Enhancing Diffusion PDE [25], which differs from the above Coherence-Enhancing Diffusion one by the choice of the diffusion tensor eigenvalues. The optimal choice of these eigenvalues indeed depends on the application, and is still an active subject of research [13]. With the above notations, the eigenvalues of 𝐃(Jρ){\bf D}(J_{\rho}) are

λ1\displaystyle\lambda_{1} :=1exp(Cμ12)\displaystyle:=1-\exp\left(\dfrac{-C}{\mu_{1}^{2}}\right)
λ2\displaystyle\lambda_{2} :=1exp(Cμ22),\displaystyle:=1-\exp\left(\dfrac{-C}{\mu_{2}^{2}}\right),
λ3\displaystyle\lambda_{3} :=1.\displaystyle:=1.

We used the values σ=0.5\sigma=0.5, ρ=4\rho=4. In our experiment, the maximum anisotropy ratio was κ=11.2\kappa=11.2. We present in Figure 15 the original image and two slices of the result after 10 time-steps of Δt=104\Delta t=10^{-4}.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 15: Top: MRI data. Bottom: Data processed via 3D Edge Enhancing diffusion, using AD-LBR. Left: 3D rendering of the original volume data and the processed volume, the 3D rendering was obtained using ImageJ 3D viewer [21] (the effect of anisotropic diffusion is not much visible in this first representation). Center: slice of the original and processed volume. Right: details of the original and processed slices.

Conclusion

We introduced in this paper a new numerical scheme, AD-LBR, for anisotropic diffusion in image processing. This scheme is non-negative, and its stencils have a limited support: 6 points in 2D, 12 points in 3D. The former property implies that our scheme respects the maximum principle of Alvarez, Guichard, Lions and Morel, which is an essential feature of parabolic PDEs.

AD-LBR outperformed all tested alternatives in a quantitative numerical experiment: a test case in which approximate numerical solutions are compared against a known analytical solution. In a second qualitative test case, different schemes were used to enhance a fingerprint image. Our scheme appears here to close more efficiently the lines of the fingerprint, and to diffuse less orthogonally to the lines. This is precisely the purpose of the implemented PDE, coherence enhancing diffusion. We also presented a 3-dimensional implementation as a proof of feasibility.

The construction of the stencils of the AD-LBR is both original and non-trivial. The computational load for this aspect of the algorithm is fortunately not dominant, thanks to the use of a tool from discrete geometry: lattice basis reduction. The AD-LBR also allows to use larger time steps than most of its counterparts, in explicit discretizations of parabolic equations.

AD-LBR trivially extends to vector valued and matrix valued images, by applying it on each image component independently. (In other words, the coupling between image components lies in the construction of the common diffusion tensor 𝐃\operatorname{\mathbf{D}}, which AD-LBR regards as user input.) Future work will be devoted to the application of AD-LBR to the regularization of diffusion tensor fields, arising for instance from diffusion MRI, for which we expect it to be particularly appropriate: thanks to the scheme non-negativity, positive-definiteness is naturally preserved.

References

  • [1] L. Alvarez, F. Guichard, P.-L. Lions, J.-M. Morel, Axioms and fundamental equations of Image processing, Arch. Rational Mech. Anal., vol. 123, 199–257 (1993)
  • [2] D. Barash, T. Schlick, M. Israeli, and R. Kimmel, Multiplicative operator splittings in non-linear diffusion: from spatial splitting to multiplicative timesteps, Journal of Mathematical Imaging and Vision, 19:33-48, (2003).
  • [3] J.-B. Bost and K. Künnemann, Hermitian vector bundles and extension groups on arithmetic schemes. I. Geometry of numbers, 2010
  • [4] J. H. Conway, N. J. A. Sloane, Low-dimensional lattices. VI. Voronoi reduction of three-dimensional lattices., Proceedings of the Royal Society of London. Series A: Mathematical and Physical Sciences 436.1896 (1992): 55-68.
  • [5] G.-H. Cottet, L. Germain, Image processing through reaction combined with nonlinear diffusion, Math. Comp., vol. 61, 659–673 (1993).
  • [6] L. Dascal, A. Ditkowski, and N. Sochen, On the Discrete Maximum Principle for the Beltrami Color Flow, J. Math. Imaging Vision, vol. 29, 63–77 (2007)
  • [7] P. Degond, A. Lozinski, J. Narski, C. Negulescu, An Asymptotic-Preserving method for highly anisotropic elliptic equations based on a micro-macro decomposition, J. Comput. Phys., vol. 231(7), 2724–2740 (2012)
  • [8] S. Grewenig, J. Weickert, and A. Bruhn, From box filtering to fast explicit diffusion, Pattern Recognition 533-542, (2010).
  • [9] W. Huang, Discrete maximum principle and a Delaunay-type mesh condition for linear finite element approximations of two-dimensional anisotropic diffusion problems, arXiv preprint arXiv:1008.0562, (2010)
  • [10] F. Labelle, J. R. Shewchuk, Anisotropic Voronoi Diagrams and Guaranteed-Quality Anisotropic Mesh Generation, Proceedings of the Nineteenth Annual Symposium on Computational Geometry, 191-200 (2003)
  • [11] J. L. Lagrange, Recherches d’arithmétique, Nouveaux Mémoires de l’Académie de Berlin, (1773)
  • [12] A. K. Lenstra, H. W. Lenstra, and L. Lovász, Factoring polynomials with rational coefficients, Mathematische Annalen 261, 513–534, (1982)
  • [13] A. M. Mendrik, E. J. Vonken, A. Rutten, M. A. Viergever, and B. van Ginneken, Noise reduction in computed tomography scans using 3-d anisotropic hybrid diffusion with continuous switch, Medical Imaging, IEEE Transactions on, 28(10), 1585-1594. (2009)
  • [14] J.-M. Mirebeau, Anisotropic Fast Marching on Cartesian Grids, using Lattice Basis Reduction, preprint, 2012.
  • [15] A. Mitchell and D. Griffiths The Finite Difference Method in Partial Differential Equations. Chichester: Wiley (1980).
  • [16] P. Mrázek and M. Navara, Consistent positive directional splitting of anisotropic diffusion, Proc. Sixth Computer Vision Winter Workshop, (2001)
  • [17] P. Q. Nguyen, and D. Stehlé, Low-dimensional lattice basis reduction revisited, ACM Transactions on Algorithms, Article 46 (2009).
  • [18] P. Q. Nguyen and J. Stern, The two faces of lattices in cryptology, In Proceedings of the 2001 Cryptography and Lattices Conference (CALC’01). Lecture Notes in Computer Science, vol. 2146. Springer-Verlag, 146–180, (2001)
  • [19] S. Osher, L. Rudin Feature-oriented image enhancement using shock filters, SIAM J. Numer. Anal., vol. 27, 919–940 (1990)
  • [20] P. Perona and J. Malik, Scale-Space and Edge Detection Using Anisotropic Diffusion, IEEE Trans. Patt. Anal. Mach. Int., vol. 12, 629–639 (1990)
  • [21] B. Schmid, J. Schindelin, A. Cardona, M. Longair, M. Heisenberg, A high-level 3D visualization API for Java and ImageJ. BMC Bioinformatics, 11:274 (2010)
  • [22] E. Selling. über die binären und ternären quadratischen formen, J. reine angew. Math., 77:143–229, (1874).
  • [23] I. Semaev, A 3-dimensional lattice reduction algorithm, In Proceedings of the 2001 Cryp- tography and Lattices Conference (CALC’01). Lecture Notes in Computer Science, vol. 2146. Springer-Verlag, 181–193, (2001)
  • [24] J. Weickert, Theoretical foundations of anisotropic diffusion in image processing, Computing, SUppl. 11, 221–236 (1996)
  • [25] J. Weickert, Anisotropic Diffusion in Image Processing, Teubner, Stuttgart (1998)
  • [26] J. Weickert, B. Romeny, and M. Viergever, Efficient and reliable schemes for nonlinear diffusion filtering, IEEE Trans. Image Proc., vol. 7, 398–410 (1998)
  • [27] J. Weickert, and H. Scharr, A scheme for coherence-enhancing diffusion filtering with optimized rotation invariance, J. Visual Comm. Image Rep., Vol. 13, 103–118, (2002).
  • [28] M. Welk and G. Steidl and J. Weickert, Locally analytic schemes: A link between diffusion filtering and wavelet shrinkage, Applied and Computational Harmonic Analysis, 24, pp 195–224, (2008)