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

Euler’s Elastica Based Cartoon-Smooth-Texture Image Decomposition

Roy Y. He111Department of Mathematics, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong. Email: [email protected]. Hao Liu222Department of Mathematics, Hong Kong Baptist University, Kowloon Tong, Kowloon, Hong Kong. Email: [email protected].
Abstract

We propose a novel model for decomposing grayscale images into three distinct components: the structural part, representing sharp boundaries and regions with strong light-to-dark transitions; the smooth part, capturing soft shadows and shades; and the oscillatory part, characterizing textures and noise. To capture the homogeneous structures, we introduce a combination of L0L^{0}-gradient and curvature regularization on level lines. This new regularization term enforces strong sparsity on the image gradient while reducing the undesirable staircase effects as well as preserving the geometry of contours. For the smoothly varying component, we utilize the L2L^{2}-norm of the Laplacian that favors isotropic smoothness. To capture the oscillation, we use the inverse Sobolev seminorm. To solve the associated minimization problem, we design an efficient operator-splitting algorithm. Our algorithm effectively addresses the challenging non-convex non-smooth problem by separating it into sub-problems. Each sub-problem can be solved either directly using closed-form solutions or efficiently using the Fast Fourier Transform (FFT). We provide systematic experiments, including ablation and comparison studies, to analyze our model’s behaviors and demonstrate its effectiveness as well as efficiency.

1 Introduction

In many tasks of image processing, such as denoising, segmentation, compression, and pattern analysis, a given image f:Ωf:\Omega\to\mathbb{R} defined on a compact domain Ω2\Omega\subset\mathbb{R}^{2} can be written as a sum of several components, each of which represents a distinctive feature of ff. For example, the well known Rudin-Osher-Fatemi (ROF) model [47] separates a noisy image f=u+nf=u+n into its noise-free part uu and the noisy remainder nn by considering

min(u,v)BV(Ω)×L2(Ω)/f=u+n[12λnL2(Ω)2+Ω|u|].\displaystyle\min_{(u,v)\in\text{BV}(\Omega)\times L^{2}(\Omega)/f=u+n}\left[\frac{1}{2\lambda}\|n\|^{2}_{L^{2}(\Omega)}+\int_{\Omega}|\nabla u|\right]. (1)

Here Ω|u|\int_{\Omega}|\nabla u| is the total variation (TV)-norm of a bounded variation function uBV(Ω)={uL1(Ω):Ω|u|<+}u\in\text{BV}(\Omega)=\{u\in L^{1}(\Omega):\int_{\Omega}|\nabla u|<+\infty\}, and λ>0\lambda>0 is a weight parameter. Various extensions of the ROF model are proposed in the literature mainly from two perspectives. The first branch of works is motivated by the fact that L2L^{2}-norm in (1) can fail to capture noise [41, 3], thus leaving visible oscillations in uu. Another class of works is based on the observation that TV-norm leads to various artifacts including staircase effects, and loss of geometry as well as contrast [12, 1, 38, 62]. See Figure 1 (e) for an illustration.

(a) (b) (c) (d) (e) (f)
Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
Figure 1: Motivation for curvature regularuzation on level lines. (a) Clean image. (b) Noisy image (σ=120/255\sigma=120/255). (c) Denoised image by L0L^{0}-gradient. (d) Denoised image by the MC penalty function [59]. (e) Denoised image by TV. (f) Denoised image by Euler’s elastica [42]. The noise is captured by H1H^{-1}-norm approximating the Meyer’s GG-norm. Observe that L0L^{0}-gradient and MC penalty function produce irregular boundaries, and TV-norm creates staircase effects. These artifacts are reduced in (f) when incorporating the curvature regularization on the level lines.

In his seminal work [41], Meyer suggested to characterize the oscillating patterns of an image including textures and noise by the dual space of the closure of the Schwartz class in BV(2)\text{BV}(\mathbb{R}^{2}). Later, Aubert and Aujol [3] refined the construction by restricting to the case of open and bounded domain Ω\Omega. They proposed the G(Ω)G(\Omega)-space as a subspace of the dual of the Sobolev space W01,1(Ω)W^{1,1}_{0}(\Omega), and formulated the Meyer’s model as

min(u,v)BV(Ω)×G(Ω)/f=u+n[1λvG(Ω)+Ω|u|],\displaystyle\min_{(u,v)\in\text{BV}(\Omega)\times G(\Omega)/f=u+n}\left[\frac{1}{\lambda}\|v\|_{G(\Omega)}+\int_{\Omega}|\nabla u|\right], (2)

where G(Ω)\|\cdot\|_{G(\Omega)} stands for the GG-norm. A key property of GG-norm is that, oscillating functions converging to 0 in G(Ω)G(\Omega) does not have to converge strongly in L2(Ω)L^{2}(\Omega) [3], which justifies the benefit of using GG-norm in place of L2L^{2}-norm in the ROF model for noise. As GG-norm is related to LL^{\infty}-norm, it is computationally challenging to solve (2). Vese and Osher [52] proposed the first practical solution by considering

min(u,𝐠)BV(Ω)×(Lp(Ω))2[Ω|u|+12λΩ|fuxg1yg2|2𝑑𝐱+μ𝐠Lp(Ω)],\displaystyle\min_{(u,\mathbf{g})\in\text{BV}(\Omega)\times\left(L^{p}(\Omega)\right)^{2}}\left[\int_{\Omega}|\nabla u|+\frac{1}{2\lambda}\int_{\Omega}|f-u-\partial_{x}g_{1}-\partial_{y}g_{2}|^{2}\,d\mathbf{x}+\mu\|\mathbf{g}\|_{L^{p}(\Omega)}\right], (3)

where d𝐱d\mathbf{x} stands for the Lebesgue measure in 2\mathbb{R}^{2}, λ>0\lambda>0 and μ>0\mu>0 are weight parameters, p1p\geq 1 is an integer index, and 𝐠Lp(Ω)=g12+g22Lp(Ω)\|\mathbf{g}\|_{L^{p}(\Omega)}=\|\sqrt{g_{1}^{2}+g_{2}^{2}}\|_{L^{p}(\Omega)} for 𝐠=(g1,g2)\mathbf{g}=(g_{1},g_{2}). They showed that as λ0\lambda\to 0 and pp\to\infty, xg1+yg2\partial_{x}g_{1}+\partial_{y}g_{2} belongs to G(Ω)G(\Omega). They derived the Euler-Lagrange equation related to (3) and solved the problem after discretization in the case of p=1p=1. In [45], a simplified version of (3) when p=2p=2 was studied, which leads to using H1(Ω)H^{-1}(\Omega) seminorm in place of Meyer’s GG-norm, where H1(Ω)H^{-1}(\Omega) is the dual of H01(Ω)H_{0}^{1}(\Omega). Approaching from a different perspective, Aujol et al. [4] proposed to tackle Meyer’s problem by restricting the function space G(Ω)G(\Omega) to a convex subset Gμ(Ω)={vG(Ω)|vG(Ω)μ}G_{\mu}(\Omega)=\{v\in G(\Omega)\leavevmode\nobreak\ |\leavevmode\nobreak\ \|v\|_{G(\Omega)}\leq\mu\} for some μ>0\mu>0. Analogous to Chambolle’s method [11] for (1), a projection algorithm was proposed in [4] to solve their model. In [53], Wen et al. considered the Legendre-Fenchel duality of the Meyer’s GG-norm and proposed a primal-dual algorithm for the cartoon-texture decomposition. In the literature, there are many other types of regularizations for capturing texture and noise. For instance, the Bounded Mean Oscillation (BMO) space was considered in [31, 22], the negative Hilbert-Sobolev space HsH^{-s} was applied in [33], and a general family of Hilbert spaces were studied in [6]. With a more refined decomposition, Aujol and Chambolle [5] suggested to separate a digital image ff as a sum of three components: the geometric part uBV(Ω)u\in\text{BV}(\Omega), the texture part vGμ(Ω)v\in G_{\mu}(\Omega), and the noise component ww in the Besov space [41] consisting of functions whose wavelet coefficients are in \ell^{\infty}. A fast filter method as well as a nice summary of variational models for cartoon++texture decomposition is available in [10]. These variational decomposition frameworks have also been extended and applied to color images [7, 32, 20, 54], RGB-d images [15], images on manifold [55], segmented image decomposition [48], and image registration [14].

To address the unsatisfying artifacts of TV-norm [12, 62, 50] in the ROF model, much efforts are devoted to constructing new regularizers that incorporate higher-order features or non-convex terms. The Euler’s elastica energy [42, 39, 49, 1]

Euler(v;a,b)=Ω(a+b(v|v|)2)|v|𝑑𝐱\displaystyle\mathcal{R}_{\text{Euler}}(v;a,b)=\int_{\Omega}\left(a+b\left(\nabla\cdot\frac{\nabla v}{|\nabla v|}\right)^{2}\right)\left|\nabla v\right|\,d\mathbf{x} (4)

has been used for disocclusion [42], inpainting [49], and denoising [1, 58], where a,b>0a,b>0 are weight parameters. See Figure 1 (f) for the reduction of staircase effects by the Euler’s elastica. In [38], Lysaker et al. proposed a regularizer based on the Frobenius norm of Hessian

H(v)=2vL1(Ω)=Ωvxx2+vxy2+vyx2+vyy2𝑑𝐱.\displaystyle\mathcal{R}_{H}(v)=\|\partial^{2}v\|_{L^{1}(\Omega)}=\int_{\Omega}\sqrt{v_{xx}^{2}+v_{xy}^{2}+v_{yx}^{2}+v_{yy}^{2}}\,d\mathbf{x}.

Regarding the image as a manifold in the Euclidean space with higher dimension, Zhu et al. [62] proposed to use the L1L^{1}-norm of the mean curvature as the regularizer

mean(v)=Ω|(v1+|v|2)|𝑑𝐱.\displaystyle\mathcal{R}_{\text{mean}}(v)=\int_{\Omega}\left|\nabla\cdot\left(\frac{\nabla v}{\sqrt{1+|\nabla v|^{2}}}\right)\right|\,d\mathbf{x}.

From the same geometric point of view, Liu et al. [35] considered the Gaussian curvature

Gauss(v)=Ωdet[vxxvxyvyxvyy](1+|v|2)3/2𝑑𝐱\displaystyle\mathcal{R}_{\text{Gauss}}(v)=\int_{\Omega}\frac{\text{det}\begin{bmatrix}v_{xx}&v_{xy}\\ v_{yx}&v_{yy}\end{bmatrix}}{(1+|\nabla v|^{2})^{3/2}}\,d\mathbf{x}

to reduce image noise while preserving geometric features. Other regularizers substituting the TV-norm include the Total Generalized Variation (TGV) [60, 56], total variation with adaptive exponent [9], non-local TV [23], locally modified TV [61], and Higher Degree TV (HDTV) [28].

The aforementioned two streams of research focus on improving different parts of the ROF model (1), and there are only a few works combining them. In [13], Chan et al. extended [5] by considering the inf-convolution of TV-norm and the squared L2L^{2}-norm of the Laplacian and arrived at the following model

min(v,w,n1,n2)BV(Ω)×H2(Ω)×Gμ(Ω)×B1,11\displaystyle\min_{(v,w,n_{1},n_{2})\in\text{BV}(\Omega)\times H^{2}(\Omega)\times G_{\mu}(\Omega)\times B_{1,1}^{1}} [Ω|v|+α2ΔwL2(Ω)2\displaystyle\bigg{[}\int_{\Omega}|\nabla v|+\frac{\alpha}{2}\|\Delta w\|_{L^{2}(\Omega)}^{2}
+J(n1μ)+B(n2δ)+12λΩ|fvwn1n2|2d𝐱].\displaystyle+J^{*}\left(\frac{n_{1}}{\mu}\right)+B^{*}\left(\frac{n_{2}}{\delta}\right)+\frac{1}{2\lambda}\int_{\Omega}|f-v-w-n_{1}-n_{2}|^{2}\,d\mathbf{x}\bigg{]}\;. (5)

Here α>0\alpha>0, J(n1μ)J^{*}\left(\frac{n_{1}}{\mu}\right) and B(n2δ)B^{*}\left(\frac{n_{2}}{\delta}\right) are the Legendre-Fenchel conjugate of the GG-norm and the B1,B_{-1,\infty}^{\infty}-norm, which enforce boundedness n1G(Ω)μ\|n_{1}\|_{G(\Omega)}\leq\mu and n2B1,δ\|n_{2}\|_{B_{-1,\infty}^{\infty}}\leq\delta, respectively, and B1,B_{-1,\infty}^{\infty} is the dual to the homogeneous Besov space B1,11B_{1,1}^{1} [41]. In (5), vv corresponds to the image’s piecewise constant part, and ww can be associated with the smoothly varying part such as soft lighting and blurry shadows. The n1n_{1} component corresponds to the textures, and the n2n_{2} component captures the noise. Recently, Huska et al. [29] proposed another three-component decomposition model, where the image f=v+w+nf=v+w+n consists of a cartoon part vv that captures the piecewise constant structures of the image, a smooth part ww that models the uneven lighting conditions in the scene, and an oscillatory component nn that includes textures and noise. Specifically, their model is

min(v,w,n)[β1v0+β22wL2(Ω)2+β3nH1(Ω)4+12λΩ|fvwn|2𝑑𝐱],\displaystyle\min_{(v,w,n)}\left[\beta_{1}\|\nabla v\|_{0}+\beta_{2}\|\partial^{2}w\|_{L^{2}(\Omega)}^{2}+\beta_{3}\|n\|^{4}_{H^{-1}(\Omega)}+\frac{1}{2\lambda}\int_{\Omega}|f-v-w-n|^{2}\,d\mathbf{x}\right], (6)

where v0\|\nabla v\|_{0} is the L0L^{0}-gradient energy [57] that counts the pixels with non-zero gradient, 2w(𝐱)2×2\partial^{2}w(\mathbf{x})\in\mathbb{R}^{2\times 2} for each 𝐱Ω\mathbf{x}\in\Omega is the Hessian matrix, 2wL2(Ω)2=(Ω(wxx2+2wxy2+wyy2)𝑑𝐱)2\|\partial^{2}w\|_{L^{2}(\Omega)}^{2}=\left(\int_{\Omega}(w_{xx}^{2}+2w^{2}_{xy}+w^{2}_{yy})\,d\mathbf{x}\right)^{2} , H1(Ω)\|\cdot\|_{H^{-1}(\Omega)} is the seminorm of the negative Sobolev space H1(Ω)H^{-1}(\Omega) as an approximation of the GG-norm, and βi\beta_{i}, i=1,2,3i=1,2,3 are positive weight parameters. For numerical computation, they replaced v0\|\nabla v\|_{0} by a modified minimax concave (MC) penalty function [59]. Note that the regularization on nn is the fourth power of the H1(Ω)H^{-1}(\Omega) seminorm, which was heuristically found to be useful in distinguishing ww and nn [29]. One of the major drawbacks of solely using the L0L^{0}-gradient regularization is visualized in Figure 1. Since L0L^{0}-gradient regularization is global and heights of boundary jumps do not affect its value, it can lead to strong staircase effects and loss of geometry [12].

Refer to caption
Figure 2: Three-component image decomposition by the proposed method. A grayscale image ff is decomposed into a structural part vv^{*}, a smooth part ww^{*}, and an oscillatory part nn^{*}. The sum of vv^{*} and ww^{*}, denoted by uu^{*}, can be regarded as the denoised version of ff if the clean image does not have rich textures. Our effectively reduces the staircase effects.

In this paper, we propose a new three-component decomposition model to address the issue discussed above. Our model decomposes any grayscale image into a structural part, a smooth part, and an oscillatory part. For the structural part, our model enforces strong sparsity on the gradient while reducing the presence of staircase effects and encouraging shape-preserving boundary smoothing. We achieve this by combining the L0L^{0}-gradient energy with the Euler’s elastica (4) which controls the magnitude of level-line curvature. For the smooth part, instead of using the Frobenius norm of the Hessian as in (6), we substitute it with squared L2L^{2}-norm of the Laplacian, ΔwL2(Ω)2\|\Delta w\|_{L^{2}(\Omega)}^{2}, to encourage isotropic smoothness. It can be shown that ΔwL2(Ω)2=2wL2(Ω)2\|\Delta w\|^{2}_{L^{2}(\Omega)}=\|\partial^{2}w\|^{2}_{L^{2}(\Omega)} for compactly supported smooth functions [13]. Moreover, we prove that they have the same critical points under certain boundary conditions (see Appendix B). For the oscillatory part, we use the squared H1(Ω)H^{-1}(\Omega) semi-norm as a computationally practical substitute for the Meyer’s GG-norm. It yields not only effective extraction of the oscillations similarly to the 4-th power in (6) but also significant improvement on the computational efficiency.

Figure 2 illustrates the general performance of our model. The structural part, vv^{*}, contains definite contours and shades with sharp transition; the smooth part, ww^{*}, captures soft intensity variation on the cheeks and chin; and the oscillatory part, nn^{*}, includes noise and textures such as those of the hair. In case the input image does not contain fine-scale textures, the composition u=v+wu^{*}=v^{*}+w^{*} gives a denoised version of ff where the noisy perturbation is removed and sharp boundaries are preserved. The level lines of uu^{*} by our model are more regular compared to those without the curvature regularization. This shows that the proposed model effectively reduces the staircase effects caused by the L0L^{0}-gradient regularization.

Our model involves both non-convex and nonlinear high-order terms, thus designing an effective numerical scheme poses a particular challenge. For the non-convex L0L^{0}-gradient minimization, various algorithms and approximations are employed in the literature, such as hard-thresholding [57, 18], region-fusion [43], and penalty approximation [29]. In [44], the L0L^{0}-gradient projection was considered and solved by using a mixed 0,1\ell^{0,1} pseudo-norm. Another nontrivial part of our proposed model is the presence of the nonlinear high-order term for the curvatures of level lines. A classical approach is to derive the first variation and then discretize the associated Euler-Lagrange equation [49]. With higher efficiency, various Operator-Splitting Methods (OSMs) and Augmented Lagrangian Methods (ALMs) [24, 26, 50, 35, 27, 29, 37, 17, 58] have been proposed to handle variational models which contain curvature regularization. OSMs are commonly derived from associating an initial value problem with the formal optimality condition, and ALMs are usually developed via the augmented Lagrangian. These approaches are deeply related and sometimes equivalent [24, 26, 21]. The main advantage of these paradigms come with appropriately splitting the original problem into several sub-problems, each of which can be efficiently handled. Recently, connections between OSMs and neural networks are investigated in [30, 51, 34].

To efficiently complete the decomposition, we propose a fast operator-splitting scheme that leverages the Fast Fourier Transform (FFT) and explicit formulas for the involved sub-problems. For an image of size N×MN\times M, each iteration of the proposed algorithm has a computational complexity of 𝒪(MN(logM+logN))\mathcal{O}\left(MN(\log M+\log N)\right). In comparison, the computational complexity in [29] is 𝒪((MN)3)\mathcal{O}((MN)^{3}). We present various examples to validate our method. Specifically, we conduct ablation studies to verify the combination of regularization terms in our method. We design experiments to investigate the impact of regularization parameters. We also demonstrate the effectiveness and efficiency of our approach by comparing it to state-of-the-art models for three-component-decomposition.

Our main contributions in this work are

  • We propose a new model for decomposing a grayscale image into its structural part, smooth part, and oscillatory part. To achieve satisfying results, we incorporate curvature regularization on level lines along with the sparsity-inducing L0L^{0}-gradient. Our model effectively reduces the undesirable staircase effects and irregular boundary curves caused by the geometry-agnostic L0L^{0}-gradient, yielding more visually appealing components.

  • We develop a new operator-splitting scheme to tackle the non-convex non-smooth optimization problem associated with our model. The proposed algorithm is computationally efficient. Each sub-problem is directly solved or numerically addressed using FFT techniques. Remarkably, when compared to several state-of-the-art methods, our algorithm achieves a speedup of over 10 times.

  • We present systematic ablation and comparison studies to justify the effectiveness and efficiency of our proposed method. We also numerically investigate the effects of model parameters and discuss the problem of parameter selection.

The paper is organized as follows. In Section 2, we propose our model. In particular, we introduce our model accompanied by mathematical remarks in Section 2.1; and reformulate it in Section 2.2 for the subsequent algorithmic developments. In Section 3, we propose our operator-splitting scheme to address the non-convex non-smooth functional minimization problem associated with the reformulated model. In Section 4, we describe the discretization and implementation details. In Section 5, we present various experiments to justify the proposed model and compare it with other methods from different aspects. We conclude the paper in Section 6.

2 Proposed Cartoon-Smooth-Texture Decomposition Model

In this section, we propose a new model for cartoon-smooth-texture image decomposition integrating Euler’s elastica. We present a formal version of our model in Section 2.1 where each component is characterized by a functional. Considering the difficulty of the problem, we introduce a reformulation of our model in Section 2.1 and derive a formal optimality condition, which paves our way towards developing an efficient operator splitting scheme in Section 3.

2.1 Proposed model

Suppose Ω2\Omega\subset\mathbb{R}^{2} is a bounded open domain with a Lipschitz boundary. We assume that a grayscale image f:Ωf:\Omega\to\mathbb{R} is a sum of three components with distinctive features. In particular, we consider ff to be composed by a structural part v:Ωv^{*}:\Omega\to\mathbb{R} that defines objects with sharp boundaries or regions outlined by strong transitions between the light and the dark; a smooth part w:Ωw^{*}:\Omega\to\mathbb{R} that characterizes the soft intensity variation, such as shadows and uneven lighting; and an oscillatory part n:Ωn^{*}:\Omega\to\mathbb{R} that captures fine-scale textures and noise. Our objective is to extract these three components from ff, and we propose to achieve this by considering the following problem

{v,w,n}\displaystyle\{v^{*},w^{*},n^{*}\} argminv,w,nv+w+n=f[α0v0+αcurvΩ(v|v|)2|v|𝑑𝐱+αwΔwL2(Ω)2+αnnH1(Ω)2].\displaystyle\in\operatorname*{arg\,min}_{\begin{subarray}{c}v,w,n\\ v+w+n=f\end{subarray}}\left[\alpha_{0}\|\nabla v\|_{0}+\alpha_{\text{curv}}\int_{\Omega}\left(\nabla\cdot\frac{v}{|\nabla v|}\right)^{2}\left|\nabla v\right|\,d\mathbf{x}+\alpha_{w}\|\Delta w\|_{L^{2}(\Omega)}^{2}+\alpha_{n}\|n\|^{2}_{H^{-1}(\Omega)}\right]. (7)

In our proposed model (7), each component of the given image ff is characterized by different functionals. Particularly, for the structural part, we associate it with the L0L^{0}-gradient energy and the total squared curvature of the level-lines; for the smooth part, we consider the squared L2L^{2}-norm of the Laplacian; and for the oscillatory part, we use the squared inverse Sobolev norm. Each of these terms is weighted by a non-negative parameter, which controls various properties of the reconstructed components.

We design functionals in (7) for better visual effects and computational efficiency. The sparsity-inducing L0L^{0}-gradient energy enforces the gradient of vv to be piecewise constant. Compared to the well-known TV regularization, it punishes discontinuities independent of their jump sizes, yielding results with higher contrast; this also implies stronger stair-casing effects. To reduce such artifacts in vv, our second term regularizes the TV-weighted squared curvatures of its level-lines. It combines high-order derivative information of the image, which has been shown to be an effective strategy for improving the visual effects of the reconstruction [13]. The squared L2L^{2}-norm of the Laplacian encourages isotropic smoothness of ww, modeling the gradual changes of illumination over the scene. Its effect is similar to that of the squared L2L^{2}-norm of the second-order derivatives, i.e., 2wL2(Ω)2\|\partial^{2}w\|_{L^{2}(\Omega)}^{2} used in (6). Under certain boundary conditions, we show that they are equivalent in the sense of having the same critical points (see Appendix B), which generalizes the observation in [13]. Using this term, we gain significant computational efficiency thanks to the simpler first variation (see Section 3). As for the oscillatory part, we use the squared inverse Sobolev norm

nH1(Ω)2=Δ1nL2(Ω)2\|n\|_{H^{-1}(\Omega)}^{2}=\|\nabla\Delta^{-1}n\|_{L^{2}(\Omega)}^{2}

as a practical approximation for the Meyer’s GG-norm.

Remark 1.

We note that some regularizations in our model (7) are rather formal. For concreteness, it is natural to assume that wH2(Ω)w\in H^{2}(\Omega) and nH1(Ω)n\in H^{-1}(\Omega); thus there is no difficulty in understanding the regularizations on the smooth and oscillatory parts. Since a classical setting for piecewise constant images is the space of bounded variation functions BV(Ω)\text{BV}(\Omega) or de Giorgi and Ambrosio’s SBV(Ω)\text{SBV}(\Omega) [16], both terms associated with the structural part call for interpretations. For the L0L^{0}-gradient energy v0\|\nabla v\|_{0}, it is often understood in a discrete sense [57], then the energy is deduced from a counting measure supported on grid points. An alternative is to consider a sufficiently smooth approximation such as the minimax concave penalty function [59, 29]. For the Euler’s elastica part, the major technical problem is that level-lines of BV functions generally lack the sufficient regularity to define curvature in the classical sense. One strategy is to restrict vv to C2(Ω)C^{2}(\Omega) [40]; another approach is to define the weak curvature for BV functions after mollification [49]; a different paradigm involves considering the solution of the pp-Laplacian equation as p1+p\to 1^{+}; yet another method relies on the pairing theory [2] for distributions [46].

Remark 2.

No matter which mathematical definitions are used for interpretation, our model (7) is challenging to analyze for the existence of minimizers. The main difficulty is caused by the L0L^{0}-gradient energy and the Euler’s elastica energy. If we use the TV-norm for regularizing vv instead, the resulting model becomes convex, and the existence property can be derived similarly to [13]. With an additional restriction on the mean value of ww, then the uniqueness can be established. If we replace the L0L^{0}-gradient energy by the TV-norm, we recover the ordinary Euler’s elastica [39, 40, 49] for the vv component, and the related analysis on semi-continuity in the context of Γ\Gamma-convergence can be found in [8].

Remark 3.

In our proposed model (7), both of the regularization on the curvature of level-lines of vv^{*} and the regularization on the Laplacian of ww^{*} encourage smoothness. However, they emphasize different types of smoothing. The curvature regularization on vv^{*} admits a geometric prior that prefers straight level lines, whereas the regularization on ww^{*} has no preferences in diffusion directions, i.e., it is isotropic.

There exist functions ww defined on Ω2\Omega\subset\mathbb{R}^{2} whose curvatures of level lines are constantly zero but not harmonic. For example, consider w(x1,x2)=x12w(x_{1},x_{2})=x_{1}^{2} defined over (x1,x2)Ω=[1,1]×[1,1](x_{1},x_{2})\in\Omega=[-1,1]\times[-1,1]. There also exist harmonic functions whose level lines have non-zero curvatures, such as v(x1,x2)=x12x22v(x_{1},x_{2})=x_{1}^{2}-x_{2}^{2}. For general images, the amount of smoothness in a given image can be transferred between vv^{*} and ww^{*} depending on the ratio αcurv/αw\alpha_{\text{curv}}/\alpha_{w}. We investigate this special property of our model in Section 5 using real images.

We emphasize that the vv^{*} component is also regularized by the L0L^{0}-gradient, which is crucial in distinguishing the structural features in vv^{*} from the soft shading and blurry shapes represented by ww^{*}. Consider a Gaussian blurred smooth function fε(x1,x2)=Gεh(x1,x2)f_{\varepsilon}(x_{1},x_{2})=G_{\varepsilon}*h(x_{1},x_{2}) defined on [0,1]×[0,1][0,1]\times[0,1] where h(x1,x2)=1h(x_{1},x_{2})=1 if 0x10.5,0x210\leq x_{1}\leq 0.5,0\leq x_{2}\leq 1, and h(x1,x2)=0h(x_{1},x_{2})=0 otherwise, GεG_{\varepsilon} is the Gaussian kernel with standard deviation ε>0\varepsilon>0, and * is the convolution. The level lines of fεf_{\varepsilon} remain straight and ΔfL2(Ω)2=𝒪(ε4)>0\|\Delta f\|_{L^{2}(\Omega)}^{2}=\mathcal{O}(\varepsilon^{-4})>0 which decreases in ε\varepsilon. If vv^{*} is only penalized by the curvature of its level line and we ignore the oscillatory component for simplicity, then the optimal decomposition is v=fεv^{*}=f_{\varepsilon} and w=0w^{*}=0 for any choice of ε\varepsilon. This is contradictory to the visual perception of homogeneous structures with clear boundaries. To consider the L0L^{0}-gradient, we first note that the Gaussian kernel has a compact support in the discrete sense. In particular, for a fixed threshold γ>0\gamma>0, we note that

1ε2πexp(x122ε2)>γ\frac{1}{\varepsilon\sqrt{2\pi}}\exp\left(-\frac{x_{1}^{2}}{2\varepsilon^{2}}\right)>\gamma\;

only if

2ε2ln(γε2π)x12ε2ln(γε2π).-\sqrt{-2\varepsilon^{2}\ln(\gamma\varepsilon\sqrt{2\pi})}\leq x_{1}\leq\sqrt{-2\varepsilon^{2}\ln(\gamma\varepsilon\sqrt{2\pi})}\;.

Hence, fε022ε2ln(γε2π)\|\nabla f_{\varepsilon}\|_{0}\approx 2\sqrt{-2\varepsilon^{2}\ln(\gamma\varepsilon\sqrt{2\pi})}, which increases when ε\varepsilon is bigger, i.e., fεf_{\varepsilon} is blurrier. With appropriate choices of the parameter α0\alpha_{0} and αn\alpha_{n}, the optimal decomposition will be v=0v^{*}=0 and w=fεw^{*}=f_{\varepsilon} if ε\varepsilon is sufficiently large, and the optimal decomposition becomes v=fεv^{*}=f_{\varepsilon} and w=0w^{*}=0 if ε\varepsilon is sufficiently small. By considering the L0L^{0}-gradient of vv^{*} component, the decomposition is compatible with visual perception.

2.2 Model reformulation and formal optimality condition

As discussed above, Model (7) is non-trivial to analyze and solve. To arrive at a practical form that can be numerically addressed, we first write the constraint as a quadratic penalization term to convert (7) as follows

minvH1(Ω),wH2(Ω),nH1(Ω)[\displaystyle\min_{\begin{subarray}{c}v\in H^{1}(\Omega),w\in H^{2}(\Omega),n\in H^{-1}(\Omega)\end{subarray}}\bigg{[} α0v0+αcurvΩ(v|v|)2|v|𝑑𝐱+αwΔwL2(Ω)2+αnnH1(Ω)2\displaystyle\alpha_{0}\|\nabla v\|_{0}+\alpha_{\text{curv}}\int_{\Omega}\left(\nabla\cdot\frac{v}{|\nabla v|}\right)^{2}\left|\nabla v\right|\,d\mathbf{x}+\alpha_{w}\|\Delta w\|_{L^{2}(\Omega)}^{2}+\alpha_{n}\|n\|^{2}_{H^{-1}(\Omega)}
+12f(v+w+n)L2(Ω)2].\displaystyle+\frac{1}{2}\|f-(v+w+n)\|_{L^{2}(\Omega)}^{2}\bigg{]}. (8)

We then resolve the high-order and non-linear term associated with the curvature by introducing auxiliary variables 𝐪=v\mathbf{q}=\nabla v and 𝝁=v|v|\boldsymbol{\mu}=\frac{\nabla v}{|\nabla v|}, which satisfy the obvious constraints 𝐪𝝁=|𝐪|\mathbf{q}\cdot\boldsymbol{\mu}=|\mathbf{q}| and |𝝁|1|\boldsymbol{\mu}|\leq 1 as in [17]. Moreover, if we let 𝐠=(g1,g2)=Δ1vL2(Ω)×L2(Ω)\mathbf{g}=(g_{1},g_{2})=\nabla\Delta^{-1}v\in L^{2}(\Omega)\times L^{2}(\Omega) as in [45], then (8) is equivalent to

{minvH1(Ω)𝐪(L2(Ω))2,𝝁(H1(Ω))2,wH2(Ω),𝐠(H1(Ω))2[α0𝐪0+αcurvΩ(𝝁)2|𝐪|d𝐱+αwΔwL2(Ω)2+αn𝐠L2(Ω)2+12f(v+w+𝐠)L2(Ω)2],𝐪=v,𝝁=v|v|,|𝝁|1.\displaystyle\begin{cases}\displaystyle\min_{\begin{subarray}{c}v\in H^{1}(\Omega)\\ \mathbf{q}\in(L^{2}(\Omega))^{2},\boldsymbol{\mu}\in(H^{1}(\Omega))^{2},\\ w\in H^{2}(\Omega),\mathbf{g}\in(H^{1}(\Omega))^{2}\end{subarray}}\bigg{[}\alpha_{0}\|\mathbf{q}\|_{0}+\alpha_{\text{curv}}\int_{\Omega}\left(\nabla\cdot\boldsymbol{\mu}\right)^{2}\left|\mathbf{q}\right|\,d\mathbf{x}+\alpha_{w}\|\Delta w\|_{L^{2}(\Omega)}^{2}+\alpha_{n}\|\mathbf{g}\|^{2}_{L^{2}(\Omega)}\\ \hskip 56.9055pt+\frac{1}{2}\|f-(v+w+\nabla\cdot\mathbf{g})\|_{L^{2}(\Omega)}^{2}\bigg{]},\\ \mathbf{q}=\nabla v,\\ \boldsymbol{\mu}=\frac{\nabla v}{|\nabla v|},\\ |\boldsymbol{\mu}|\leq 1.\end{cases} (9)

We note that the problem can be further simplified by observing the following characterization of minimizers of Model (8), whose proof is in Appendix A.

Proposition 2.1.

Let (v,w,n)(v^{*},w^{*},n^{*}) be a minimizer of (8), then we have

Ωv+w+nd𝐱=Ωf𝑑𝐱.\displaystyle\int_{\Omega}v^{*}+w^{*}+n^{*}\,d\mathbf{x}=\int_{\Omega}f\,d\mathbf{x}. (10)

This implies that if (v,𝐪,w,𝐠)(v^{*},\mathbf{q}^{*},w^{*},\mathbf{g}^{*}) is a minimizer of Model (9), vv^{*} is in fact uniquely determined by the solution of the following problem

{2v=𝐪 in Ω,(v𝐪)𝐧=0 on Ω,Ωv𝑑𝐱=Ωfw𝐠d𝐱.\displaystyle\begin{cases}\nabla^{2}v=\nabla\cdot\mathbf{q}^{*}&\mbox{ in }\Omega,\\ (\nabla v-\mathbf{q}^{*})\cdot\mathbf{n}=0&\mbox{ on }\partial\Omega,\\ \int_{\Omega}v\,d\mathbf{x}=\int_{\Omega}f-w^{*}-\nabla\cdot\mathbf{g}^{*}\,d\mathbf{x}.\end{cases} (11)

where 𝐧\mathbf{n} denotes the outward normal direction along Ω\partial\Omega. Therefore, we consider an approximation of (9) as

min𝐪(L2(Ω))2,𝝁(H1(Ω))2,wH2(Ω),𝐠(H1(Ω))2[\displaystyle\min_{\begin{subarray}{c}\mathbf{q}\in(L^{2}(\Omega))^{2},\boldsymbol{\mu}\in(H^{1}(\Omega))^{2},\\ w\in H^{2}(\Omega),\mathbf{g}\in(H^{1}(\Omega))^{2}\end{subarray}}\bigg{[} α0𝐪0+αcurvΩ|𝝁|2|𝐪|𝑑𝐱+αwΩ|2w|2𝑑𝐱\displaystyle\alpha_{0}\|\mathbf{q}\|_{0}+\alpha_{\rm curv}\int_{\Omega}\left|\nabla\cdot\boldsymbol{\mu}\right|^{2}\left|\mathbf{q}\right|\,d\mathbf{x}+\alpha_{w}\int_{\Omega}|\nabla^{2}w|^{2}d\mathbf{x}
+αnΩ|𝐠|2d𝐱+12Ω|f(v𝐪,w,𝐠+w+𝐠)|2d𝐱+IΣf(𝐪,w,𝐠)+IS(𝐪,𝝁)],\displaystyle+\alpha_{n}\int_{\Omega}|\mathbf{g}|^{2}d\mathbf{x}+\frac{1}{2}\int_{\Omega}|f-(v_{\mathbf{q},w,\mathbf{g}}+w+\nabla\cdot\mathbf{g})|^{2}d\mathbf{x}+I_{\Sigma_{f}}(\mathbf{q},w,\mathbf{g})+I_{S}(\mathbf{q},\boldsymbol{\mu})\bigg{]}, (12)

where v𝐪,w,𝐠v_{\mathbf{q},w,\mathbf{g}} is uniquely determined by 𝐪,w,𝐠\mathbf{q},w,\mathbf{g} via (11); and the indicator functions

IΣf(𝐪,w,𝐠)={0 if (𝐪,w,𝐠)Σf,+ otherwise,IS(𝐪,𝝁)={0 if (𝐪,𝝁)S,+ otherwise.\displaystyle I_{\Sigma_{f}}(\mathbf{q},w,\mathbf{g})=\begin{cases}0&\mbox{ if }(\mathbf{q},w,\mathbf{g})\in\Sigma_{f},\\ +\infty&\mbox{ otherwise},\end{cases}\quad I_{S}(\mathbf{q},\boldsymbol{\mu})=\begin{cases}0&\mbox{ if }(\mathbf{q},\boldsymbol{\mu})\in S,\\ +\infty&\mbox{ otherwise}.\end{cases}

with

Σf={(𝐪,w,𝐠)(L2(Ω))2×L2(Ω)×(H1(Ω))2|\displaystyle\Sigma_{f}=\bigg{\{}(\mathbf{q},w,\mathbf{g})\in(L^{2}(\Omega))^{2}\times L^{2}(\Omega)\times(H^{1}(\Omega))^{2}\leavevmode\nobreak\ |\leavevmode\nobreak\
vH1(Ω)such that𝐪=vandΩ(v+w+𝐠f)d𝐱=0},\displaystyle\hskip 85.35826pt\exists v\in H^{1}(\Omega)\leavevmode\nobreak\ \text{such that}\leavevmode\nobreak\ \mathbf{q}=\nabla v\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ \int_{\Omega}(v+w+\nabla\cdot\mathbf{g}-f)\,d\mathbf{x}=0\bigg{\}},
S={(𝐪,𝝁)(L2(Ω))2×(L2(Ω))2|𝐪𝝁=|𝐪|,|𝝁|1}.\displaystyle S=\left\{(\mathbf{q},\boldsymbol{\mu})\in(L^{2}(\Omega))^{2}\times(L^{2}(\Omega))^{2}\leavevmode\nobreak\ |\leavevmode\nobreak\ \mathbf{q}\cdot\boldsymbol{\mu}=|\mathbf{q}|,|\boldsymbol{\mu}|\leq 1\right\}.

Suppose we find {𝐪,𝝁,w,𝐠}\{\mathbf{q}^{*},\boldsymbol{\mu}^{*},w^{*},\mathbf{g}^{*}\} as a minimizer of (12), then we have v=v𝐪,w,𝐠v^{*}=v_{\mathbf{q}^{*},w^{*},\mathbf{g}^{*}} for the structural part, ww^{*} for the smooth part, and n=𝐠n^{*}=\nabla\cdot\mathbf{g}^{*} for the oscillatory part. Moreover, if we denote

J0(𝐪,w,𝐠)=12Ω|f(v𝐪,w,𝐠+w+𝐠)|2𝑑𝐱,\displaystyle J_{0}(\mathbf{q},w,\mathbf{g})=\frac{1}{2}\int_{\Omega}|f-(v_{\mathbf{q},w,\mathbf{g}}+w+\nabla\cdot\mathbf{g})|^{2}d\mathbf{x},
J1(𝐪)=α0𝐪0,\displaystyle J_{1}(\mathbf{q})=\alpha_{0}\|\mathbf{q}\|_{0},
J2(𝐪,𝝁)=αcurvΩ|𝝁|2|𝐪|𝑑𝐱,\displaystyle J_{2}(\mathbf{q},\boldsymbol{\mu})=\alpha_{\rm curv}\int_{\Omega}\left|\nabla\cdot\boldsymbol{\mu}\right|^{2}\left|\mathbf{q}\right|\,d\mathbf{x},
J3(w)=αwΩ|2w|2𝑑𝐱,\displaystyle J_{3}(w)=\alpha_{w}\int_{\Omega}|\nabla^{2}w|^{2}d\mathbf{x},
J4(𝐠)=αnΩ|𝐠|2𝑑𝐱,\displaystyle J_{4}(\mathbf{g})=\alpha_{n}\int_{\Omega}|\mathbf{g}|^{2}d\mathbf{x},

the minimizer {𝐪,𝝁,w,𝐠}\{\mathbf{q}^{*},\boldsymbol{\mu}^{*},w^{*},\mathbf{g}^{*}\} satisfies the formal optimality condition

{𝐪J1(𝐪)+𝐪J2(𝐪,𝝁)+𝐪J0(𝐪,w,𝐠)+𝐪IΣf(𝐪,w,𝐠)+𝐪IS(𝐪,𝝁)0,𝝁J2(𝐪,𝝁)+𝝁IS(𝐪,𝝁)0,wJ3(w)+wJ0(𝐪,w,𝐠)+wIΣf(𝐪,w,𝐠)0,𝐠J4(𝐠)+𝐠J0(𝐪,w,𝐠)+𝐠IΣf(𝐪,w,𝐠)0.\displaystyle\begin{cases}\partial_{\mathbf{q}}J_{1}(\mathbf{q}^{*})+\partial_{\mathbf{q}}J_{2}(\mathbf{q}^{*},\boldsymbol{\mu}^{*})+\partial_{\mathbf{q}}J_{0}(\mathbf{q}^{*},w^{*},\mathbf{g}^{*})+\partial_{\mathbf{q}}I_{\Sigma_{f}}(\mathbf{q}^{*},w^{*},\mathbf{g}^{*})+\partial_{\mathbf{q}}I_{S}(\mathbf{q}^{*},\boldsymbol{\mu}^{*})\ni 0,\\ \partial_{\boldsymbol{\mu}}J_{2}(\mathbf{q}^{*},\boldsymbol{\mu}^{*})+\partial_{\boldsymbol{\mu}}I_{S}(\mathbf{q}^{*},\boldsymbol{\mu}^{*})\ni 0,\\ \partial_{w}J_{3}(w^{*})+\partial_{w}J_{0}(\mathbf{q}^{*},w^{*},\mathbf{g}^{*})+\partial_{w}I_{\Sigma_{f}}(\mathbf{q}^{*},w^{*},\mathbf{g}^{*})\ni 0,\\ \partial_{\mathbf{g}}J_{4}(\mathbf{g}^{*})+\partial_{\mathbf{g}}J_{0}(\mathbf{q}^{*},w^{*},\mathbf{g}^{*})+\partial_{\mathbf{g}}I_{\Sigma_{f}}(\mathbf{q}^{*},w^{*},\mathbf{g}^{*})\ni 0.\end{cases} (13)

where \partial is the subdifferential.

3 Operator-Splitting Algorithm for the Proposed Model

In this section, we propose an efficient algorithm for Model (12) based on an operator-splitting framework. In Section 2.2, we associate the model with a gradient flow based on its formal optimality condition. In Section 3.1, we present a novel operator-splitting scheme to solve the reformulated model. This non-convex non-smooth minimization problem is decomposed into several subproblems, whose solutions are described in Section 3.2-3.4. In Section 3.5, we adapt the proposed algorithm to periodic boundary condition. We summarize the proposed scheme in Algorithm 1.

3.1 Proposed scheme

We associate (13) with the following initial value problem in analogous to the gradient flow

{𝐩t+𝐪J1(𝐩)+𝐪J2(𝐩,𝝀)+𝐪J0(𝐩,r,𝐬)+𝐪IΣf(𝐩)+𝐪IS(𝐩,𝝀)0,γ1𝝀t+𝝁J2(𝐩,𝝀)+𝝁IS(𝐩,𝝀)0,γ2rt+wJ3(r)+wJ0(𝐩,r,𝐬)=0,γ3𝐬t+𝐠J4(𝐬)+𝐠J0(𝐩,w,𝐬)=0,{𝐩(0),𝝀(0),r(0),𝐬(0)}={𝐩0,𝝀0,r0,𝐬0},\displaystyle\begin{cases}\frac{\partial\mathbf{p}}{\partial t}+\partial_{\mathbf{q}}J_{1}(\mathbf{p})+\partial_{\mathbf{q}}J_{2}(\mathbf{p},\boldsymbol{\lambda})+\partial_{\mathbf{q}}J_{0}(\mathbf{p},r,\mathbf{s})+\partial_{\mathbf{q}}I_{\Sigma_{f}}(\mathbf{p})+\partial_{\mathbf{q}}I_{S}(\mathbf{p},\boldsymbol{\lambda})\ni 0,\\ \gamma_{1}\frac{\partial\boldsymbol{\lambda}}{\partial t}+\partial_{\boldsymbol{\mu}}J_{2}(\mathbf{p},\boldsymbol{\lambda})+\partial_{\boldsymbol{\mu}}I_{S}(\mathbf{p},\boldsymbol{\lambda})\ni 0,\\ \gamma_{2}\frac{\partial r}{\partial t}+\partial_{w}J_{3}(r)+\partial_{w}J_{0}(\mathbf{p},r,\mathbf{s})=0,\\ \gamma_{3}\frac{\partial\mathbf{s}}{\partial t}+\partial_{\mathbf{g}}J_{4}(\mathbf{s})+\partial_{\mathbf{g}}J_{0}(\mathbf{p},w,\mathbf{s})=0,\\ \{\mathbf{p}(0),\boldsymbol{\lambda}(0),r(0),\mathbf{s}(0)\}=\{\mathbf{p}_{0},\boldsymbol{\lambda}_{0},r_{0},\mathbf{s}_{0}\},\end{cases} (14)

where {𝐩0,𝝀0,r0,𝐬0}\{\mathbf{p}_{0},\boldsymbol{\lambda}_{0},r_{0},\mathbf{s}_{0}\} is a given initial condition, and γ1,γ2,γ3>0\gamma_{1},\gamma_{2},\gamma_{3}>0 are parameters that control the evolution speed of 𝐩,𝝀,r,𝐬\mathbf{p},\boldsymbol{\lambda},r,\mathbf{s}, respectively. We note that the steady state solution of (14) satisfies the optimality condition (13) for Model (12). To avoid confusion, we distinguish the time-dependent variables in (14) from the optimization variables in (12), and Table 1 shows the correspondences.

Correspondence
General variables 𝐪\mathbf{q} 𝝁\boldsymbol{\mu} ww 𝐠\mathbf{g}
PDE solutions 𝐩\mathbf{p} 𝝀\boldsymbol{\lambda} rr 𝐬\mathbf{s}
Table 1: Notation of general variables and PDE solutions and their correspondences.

We find the steady state solution of (14) by employing the operator-splitting framework [26, 25]. Our proposed scheme proceeds as follows:

Initialization

Initialize{𝐩0,𝝀0,r0,𝐬0}={𝐩0,𝝀0,r0,𝐬0}.\displaystyle\text{Initialize}\leavevmode\nobreak\ \{\mathbf{p}^{0},\boldsymbol{\lambda}^{0},r^{0},\mathbf{s}^{0}\}=\{\mathbf{p}_{0},\boldsymbol{\lambda}_{0},r_{0},\mathbf{s}_{0}\}.

Fix some τ>0\tau>0 as the time step. For k=0,1,2,k=0,1,2,\dots, denote tk=kτt^{k}=k\tau for k=0,1,2k=0,1,2\dots; and we update {𝐩k,𝝀k,rk,𝐬k}{𝐩k+1,𝝀k+1,rk+1,𝐬k+1}\{\mathbf{p}^{k},\boldsymbol{\lambda}^{k},r^{k},\mathbf{s}^{k}\}\rightarrow\{\mathbf{p}^{k+1},\boldsymbol{\lambda}^{k+1},r^{k+1},\mathbf{s}^{k+1}\} via four fractional steps:

Fractional step 1: Solve

{{𝐩t+𝐪J1(𝐩)0,γ1𝝀t=0,γ2rt=0,γ3𝐬t=0, in Ω×(tk,tk+1),{𝐩(tk),𝝀(tk),r(tk),𝐬(tk)}={𝐩k,𝝀k,rk,𝐬k},\displaystyle\begin{cases}\begin{cases}\frac{\partial\mathbf{p}}{\partial t}+\partial_{\mathbf{q}}J_{1}(\mathbf{p})\ni 0,\\ \gamma_{1}\frac{\partial\boldsymbol{\lambda}}{\partial t}=0,\\ \gamma_{2}\frac{\partial r}{\partial t}=0,\\ \gamma_{3}\frac{\partial\mathbf{s}}{\partial t}=0,\end{cases}\mbox{ in }\Omega\times(t^{k},t^{k+1}),\\ \{\mathbf{p}(t^{k}),\boldsymbol{\lambda}(t^{k}),r(t^{k}),\mathbf{s}(t^{k})\}=\{\mathbf{p}^{k},\boldsymbol{\lambda}^{k},r^{k},\mathbf{s}^{k}\},\end{cases}

and set

{𝐩k+1/4,𝝀k+1/4,rk+1/4,𝐬k+1/4}={𝐩(tk+1),𝝀(tk+1),r(tk+1),𝐬(tk+1)}.\displaystyle\{\mathbf{p}^{k+1/4},\boldsymbol{\lambda}^{k+1/4},r^{k+1/4},\mathbf{s}^{k+1/4}\}=\{\mathbf{p}(t^{k+1}),\boldsymbol{\lambda}(t^{k+1}),r(t^{k+1}),\mathbf{s}(t^{k+1})\}.

Fractional step 2: Solve

{{𝐩t+𝐪J2(𝐩,𝝀)=0,γ1𝝀t+𝝁J2(𝐩,𝝀)0,γ2rt=0,γ3𝐬t=0, in Ω×(tk,tk+1),{𝐩(tk),𝝀(tk),r(tk),𝐬(tk)}={𝐩k+1/4,𝝀k+1/4,rk+1/4,𝐬k+1/4},\displaystyle\begin{cases}\begin{cases}\frac{\partial\mathbf{p}}{\partial t}+\partial_{\mathbf{q}}J_{2}(\mathbf{p},\boldsymbol{\lambda})=0,\\ \gamma_{1}\frac{\partial\boldsymbol{\lambda}}{\partial t}+\partial_{\boldsymbol{\mu}}J_{2}(\mathbf{p},\boldsymbol{\lambda})\ni 0,\\ \gamma_{2}\frac{\partial r}{\partial t}=0,\\ \gamma_{3}\frac{\partial\mathbf{s}}{\partial t}=0,\end{cases}\mbox{ in }\Omega\times(t^{k},t^{k+1}),\\ \{\mathbf{p}(t^{k}),\boldsymbol{\lambda}(t^{k}),r(t^{k}),\mathbf{s}(t^{k})\}=\{\mathbf{p}^{k+1/4},\boldsymbol{\lambda}^{k+1/4},r^{k+1/4},\mathbf{s}^{k+1/4}\},\end{cases}

and set

{𝐩k+2/4,𝝀k+2/4,rk+2/4,𝐬k+2/4}={𝐩(tk+1),𝝀(tk+1),r(tk+1),𝐬(tk+1)}.\displaystyle\{\mathbf{p}^{k+2/4},\boldsymbol{\lambda}^{k+2/4},r^{k+2/4},\mathbf{s}^{k+2/4}\}=\{\mathbf{p}(t^{k+1}),\boldsymbol{\lambda}(t^{k+1}),r(t^{k+1}),\mathbf{s}(t^{k+1})\}.

Fractional step 3: Solve

{{𝐩t+𝐪IS(𝐩,𝝀)0,γ1𝝀t+𝝁IS(𝐩,𝝀)0,γ2rt=0,γ3𝐬t=0, in Ω×(tk,tk+1),{𝐩(tk),𝝀(tk),r(tk),𝐬(tk)}={𝐩k+2/4,𝝀k+2/4,rk+2/4,𝐬k+2/4},\displaystyle\begin{cases}\begin{cases}\frac{\partial\mathbf{p}}{\partial t}+\partial_{\mathbf{q}}I_{S}(\mathbf{p},\boldsymbol{\lambda})\ni 0,\\ \gamma_{1}\frac{\partial\boldsymbol{\lambda}}{\partial t}+\partial_{\boldsymbol{\mu}}I_{S}(\mathbf{p},\boldsymbol{\lambda})\ni 0,\\ \gamma_{2}\frac{\partial r}{\partial t}=0,\\ \gamma_{3}\frac{\partial\mathbf{s}}{\partial t}=0,\end{cases}\mbox{ in }\Omega\times(t^{k},t^{k+1}),\\ \{\mathbf{p}(t^{k}),\boldsymbol{\lambda}(t^{k}),r(t^{k}),\mathbf{s}(t^{k})\}=\{\mathbf{p}^{k+2/4},\boldsymbol{\lambda}^{k+2/4},r^{k+2/4},\mathbf{s}^{k+2/4}\},\end{cases}

and set

{𝐩k+3/4,𝝀k+3/4,rk+3/4,𝐬k+3/4}={𝐩(tk+1),𝝀(tk+1),r(tk+1),𝐬(tk+1)}.\displaystyle\{\mathbf{p}^{k+3/4},\boldsymbol{\lambda}^{k+3/4},r^{k+3/4},\mathbf{s}^{k+3/4}\}=\{\mathbf{p}(t^{k+1}),\boldsymbol{\lambda}(t^{k+1}),r(t^{k+1}),\mathbf{s}(t^{k+1})\}.

Fractional step 4: Solve

{{𝐩t+𝐪J0(𝐩,r,𝐬)+𝐪IΣf(𝐩,r,𝐬)0,γ1𝝀t=0,γ2rt+wJ0(𝐩,r,𝐬)+wJ3(r)+wIΣf(𝐩,r,𝐬)0,γ3𝐬t+𝐠J0(𝐩,r,𝐬)+𝐠J4(𝐬)+𝐠IΣf(𝐩,r,𝐬)0, in Ω×(tk,tk+1),{𝐩(tk),𝝀(tk),r(tk),𝐬(tk)}={𝐩k+3/4,𝝀k+3/4,rk+3/4,𝐬k+3/4},\displaystyle\begin{cases}\begin{cases}\frac{\partial\mathbf{p}}{\partial t}+\partial_{\mathbf{q}}J_{0}(\mathbf{p},r,\mathbf{s})+\partial_{\mathbf{q}}I_{\Sigma_{f}}(\mathbf{p},r,\mathbf{s})\ni 0,\\ \gamma_{1}\frac{\partial\boldsymbol{\lambda}}{\partial t}=0,\\ \gamma_{2}\frac{\partial r}{\partial t}+\partial_{w}J_{0}(\mathbf{p},r,\mathbf{s})+\partial_{w}J_{3}(r)+\partial_{w}I_{\Sigma_{f}}(\mathbf{p},r,\mathbf{s})\ni 0,\\ \gamma_{3}\frac{\partial\mathbf{s}}{\partial t}+\partial_{\mathbf{g}}J_{0}(\mathbf{p},r,\mathbf{s})+\partial_{\mathbf{g}}J_{4}(\mathbf{s})+\partial_{\mathbf{g}}I_{\Sigma_{f}}(\mathbf{p},r,\mathbf{s})\ni 0,\end{cases}\mbox{ in }\Omega\times(t^{k},t^{k+1}),\\ \{\mathbf{p}(t^{k}),\boldsymbol{\lambda}(t^{k}),r(t^{k}),\mathbf{s}(t^{k})\}=\{\mathbf{p}^{k+3/4},\boldsymbol{\lambda}^{k+3/4},r^{k+3/4},\mathbf{s}^{k+3/4}\},\end{cases}

and set

{𝐩k+1,𝝀k+1,rk+1,𝐬k+1}={𝐩(tk+1),𝝀(tk+1),r(tk+1),𝐬(tk+1)}.\displaystyle\{\mathbf{p}^{k+1},\boldsymbol{\lambda}^{k+1},r^{k+1},\mathbf{s}^{k+1}\}=\{\mathbf{p}(t^{k+1}),\boldsymbol{\lambda}(t^{k+1}),r(t^{k+1}),\mathbf{s}(t^{k+1})\}.

We time discretize the systems above by the Marchuk-Yanenko scheme and get

{𝐩k+1/4𝐩kτ+𝐪J1(𝐩k+1/4)0,γ1𝝀k+1/4𝝀kτ=0,γ2rk+1/4rkτ=0,γ3𝐬k+1/4𝐬kτ=0,\displaystyle\begin{cases}\frac{\mathbf{p}^{k+1/4}-\mathbf{p}^{k}}{\tau}+\partial_{\mathbf{q}}J_{1}(\mathbf{p}^{k+1/4})\ni 0,\\ \gamma_{1}\frac{\boldsymbol{\lambda}^{k+1/4}-\boldsymbol{\lambda}^{k}}{\tau}=0,\\ \gamma_{2}\frac{r^{k+1/4}-r^{k}}{\tau}=0,\\ \gamma_{3}\frac{\mathbf{s}^{k+1/4}-\mathbf{s}^{k}}{\tau}=0,\end{cases} (15)
{𝐩k+2/4𝐩k+1/4τ+𝐪J2(𝐩k+2/4,𝝀k+1/4)0,γ1𝝀k+2/4𝝀k+1/4τ+𝝁J2(𝐩k+2/4,𝝀k+2/4)0,γ2rk+2/4rk+1/4τ=0,γ3𝐬k+2/4𝐬k+1/4τ=0,\displaystyle\begin{cases}\frac{\mathbf{p}^{k+2/4}-\mathbf{p}^{k+1/4}}{\tau}+\partial_{\mathbf{q}}J_{2}(\mathbf{p}^{k+2/4},\boldsymbol{\lambda}^{k+1/4})\ni 0,\\ \gamma_{1}\frac{\boldsymbol{\lambda}^{k+2/4}-\boldsymbol{\lambda}^{k+1/4}}{\tau}+\partial_{\boldsymbol{\mu}}J_{2}(\mathbf{p}^{k+2/4},\boldsymbol{\lambda}^{k+2/4})\ni 0,\\ \gamma_{2}\frac{r^{k+2/4}-r^{k+1/4}}{\tau}=0,\\ \gamma_{3}\frac{\mathbf{s}^{k+2/4}-\mathbf{s}^{k+1/4}}{\tau}=0,\end{cases} (16)
{𝐩k+3/4𝐩k+2/4τ+𝐪IS(𝐩k+3/4,𝝀k+3/4)0,γ1𝝀k+3/4𝝀k+2/4τ+𝝁IS(𝐩k+3/4,𝝀k+3/4)0,γ2rk+3/4rk+2/4τ=0,γ3𝐬k+3/4𝐬k+2/4τ=0,\displaystyle\begin{cases}\frac{\mathbf{p}^{k+3/4}-\mathbf{p}^{k+2/4}}{\tau}+\partial_{\mathbf{q}}I_{S}(\mathbf{p}^{k+3/4},\boldsymbol{\lambda}^{k+3/4})\ni 0,\\ \gamma_{1}\frac{\boldsymbol{\lambda}^{k+3/4}-\boldsymbol{\lambda}^{k+2/4}}{\tau}+\partial_{\boldsymbol{\mu}}I_{S}(\mathbf{p}^{k+3/4},\boldsymbol{\lambda}^{k+3/4})\ni 0,\\ \gamma_{2}\frac{r^{k+3/4}-r^{k+2/4}}{\tau}=0,\\ \gamma_{3}\frac{\mathbf{s}^{k+3/4}-\mathbf{s}^{k+2/4}}{\tau}=0,\end{cases} (17)
{𝐩k+1𝐩k+3/4τ+𝐪J0(𝐩k+1,rk+1,𝐬k+1)+𝐪IΣf(𝐩k+1,rk+1,𝐬k+1)0,γ1𝝀k+1𝝀k+3/4τ=0,γ2rk+1rk+3/4τ+wJ0(𝐩k+1,rk+1,𝐬k+1)+wJ3(rk+1)+wIΣf(𝐩k+1,rk+1,𝐬k+1)0,γ3𝐬k+1𝐬k+3/4τ+𝐠J0(𝐩k+1,rk+1,𝐬k+1)+𝐠J4(𝐬k+1)+𝐠IΣf(𝐩k+1,rk+1,𝐬k+1)0.\displaystyle\begin{cases}\frac{\mathbf{p}^{k+1}-\mathbf{p}^{k+3/4}}{\tau}+\partial_{\mathbf{q}}J_{0}(\mathbf{p}^{k+1},r^{k+1},\mathbf{s}^{k+1})+\partial_{\mathbf{q}}I_{\Sigma_{f}}(\mathbf{p}^{k+1},r^{k+1},\mathbf{s}^{k+1})\ni 0,\\ \gamma_{1}\frac{\boldsymbol{\lambda}^{k+1}-\boldsymbol{\lambda}^{k+3/4}}{\tau}=0,\\ \gamma_{2}\frac{r^{k+1}-r^{k+3/4}}{\tau}+\partial_{w}J_{0}(\mathbf{p}^{k+1},r^{k+1},\mathbf{s}^{k+1})+\partial_{w}J_{3}(r^{k+1})+\partial_{w}I_{\Sigma_{f}}(\mathbf{p}^{k+1},r^{k+1},\mathbf{s}^{k+1})\ni 0,\\ \gamma_{3}\frac{\mathbf{s}^{k+1}-\mathbf{s}^{k+3/4}}{\tau}+\partial_{\mathbf{g}}J_{0}(\mathbf{p}^{k+1},r^{k+1},\mathbf{s}^{k+1})+\partial_{\mathbf{g}}J_{4}(\mathbf{s}^{k+1})+\partial_{\mathbf{g}}I_{\Sigma_{f}}(\mathbf{p}^{k+1},r^{k+1},\mathbf{s}^{k+1})\ni 0.\end{cases} (18)

for k=0,1,2,k=0,1,2,\dots. We note that each system above corresponds to a subproblem which can be solved explicitly or approximated efficiently. In the following sections, we describe the details.

3.2 On the solution to sub-problem (15) and (16)

Given 𝐩k\mathbf{p}^{k} for k=0,1,2,k=0,1,2,\dots, the vector field 𝐩k+1/4\mathbf{p}^{k+1/4} solves

𝐩k+1/4=argmin𝐪(L2(Ω))2[12Ω|𝐪𝐩k|2𝑑𝐱+τα0𝐪0].\displaystyle\mathbf{p}^{k+1/4}=\operatorname*{arg\,min}_{\mathbf{q}\in(L^{2}(\Omega))^{2}}\left[\frac{1}{2}\int_{\Omega}|\mathbf{q}-\mathbf{p}^{k}|^{2}d\mathbf{x}+\tau\alpha_{0}\|\mathbf{q}\|_{0}\right]\;.

This problem has the closed-form solution [57]

𝐩k+1/4={𝟎 if |𝐩k|2τα0,𝐩k otherwise.\displaystyle\mathbf{p}^{k+1/4}=\begin{cases}\mathbf{0}&\mbox{ if }|\mathbf{p}^{k}|^{2}\leq\tau\alpha_{0},\\ \mathbf{p}^{k}&\mbox{ otherwise}.\end{cases} (19)

The first equation in (16) is the Euler-Lagrange equation of the following minimization problem

𝐩k+2/4=argmin𝐪(L2(Ω))2[12Ω|𝐪𝐩k+1/4|2𝑑𝐱+ταcurvΩ|𝝀k+1/4|2|𝐪|𝑑𝐱],\displaystyle\mathbf{p}^{k+2/4}=\operatorname*{arg\,min}_{\mathbf{q}\in(L^{2}(\Omega))^{2}}\left[\frac{1}{2}\int_{\Omega}|\mathbf{q}-\mathbf{p}^{k+1/4}|^{2}d\mathbf{x}+\tau\alpha_{\rm curv}\int_{\Omega}\left|\nabla\cdot\boldsymbol{\lambda}^{k+1/4}\right|^{2}|\mathbf{q}|d\mathbf{x}\right]\;,

whose closed-form solution is given as

𝐩k+2/4=max{0,1ταcurv|𝝀k+1/4|2|𝐩k+1/4|}𝐩k+1/4.\displaystyle\mathbf{p}^{k+2/4}=\max\left\{0,1-\frac{\tau\alpha_{\rm curv}\left|\nabla\cdot\boldsymbol{\lambda}^{k+1/4}\right|^{2}}{|\mathbf{p}^{k+1/4}|}\right\}\mathbf{p}^{k+1/4}. (20)

For the second equation of (16), we note that 𝝀k+2/4\boldsymbol{\lambda}^{k+2/4} solves

𝝀k+2/4=argmin𝝁(H1(Ω))2[γ12Ω|𝝁𝝀k+1/4|2𝑑𝐱+ταcurvΩ|𝝁|2|𝐩k+2/4|𝑑𝐱],\displaystyle\boldsymbol{\lambda}^{k+2/4}=\operatorname*{arg\,min}_{\boldsymbol{\mu}\in(H^{1}(\Omega))^{2}}\left[\frac{\gamma_{1}}{2}\int_{\Omega}|\boldsymbol{\mu}-\boldsymbol{\lambda}^{k+1/4}|^{2}d\mathbf{x}+\tau\alpha_{\rm curv}\int_{\Omega}\left|\nabla\cdot\boldsymbol{\mu}\right|^{2}|\mathbf{p}^{k+2/4}|d\mathbf{x}\right], (21)

and the corresponding Euler-Lagrange equation is

{γ1𝝀k+2/4𝝀k+1/4τ2αcurv(|𝐩k+2/4|𝝀k+2/4)=0 in Ω,|𝐩k+2/4|𝝀k+2/4=0 on Ω.\displaystyle\begin{cases}\gamma_{1}\frac{\boldsymbol{\lambda}^{k+2/4}-\boldsymbol{\lambda}^{k+1/4}}{\tau}-2\alpha_{\rm curv}\nabla(|\mathbf{p}^{k+2/4}|\nabla\cdot\boldsymbol{\lambda}^{k+2/4})=0\mbox{ in }\Omega,\\ |\mathbf{p}^{k+2/4}|\nabla\cdot\boldsymbol{\lambda}^{k+2/4}=0\mbox{ on }\partial\Omega.\end{cases} (22)

3.3 On the solution to sub-problem (17)

We update 𝐩k+3/4\mathbf{p}^{k+3/4} and 𝝀k+3/4\boldsymbol{\lambda}^{k+3/4} using the method proposed in [17, Section 3.5]. Note that 𝐩k+3/4\mathbf{p}^{k+3/4} and 𝝀k+3/4\boldsymbol{\lambda}^{k+3/4} solve

(𝐩k+3/4,𝝀k+3/4)=argmin(𝐪,𝝁)S[12Ω|𝐪𝐩k+2/4|2𝑑𝐱+γ12Ω|𝝁𝝀k+2/4|2𝑑𝐱].\displaystyle(\mathbf{p}^{k+3/4},\boldsymbol{\lambda}^{k+3/4})=\operatorname*{arg\,min}_{(\mathbf{q},\boldsymbol{\mu})\in S}\left[\frac{1}{2}\int_{\Omega}|\mathbf{q}-\mathbf{p}^{k+2/4}|^{2}d\mathbf{x}+\frac{\gamma_{1}}{2}\int_{\Omega}|\boldsymbol{\mu}-\boldsymbol{\lambda}^{k+2/4}|^{2}d\mathbf{x}\right]. (23)

Problem (23) can be solved in a pointwise manner. For each given 𝐱\mathbf{x}, (𝐩k+3/4(𝐱),𝝀k+3/4(𝐱))(\mathbf{p}^{k+3/4}(\mathbf{x}),\boldsymbol{\lambda}^{k+3/4}(\mathbf{x})) satisfies

(𝐩k+3/4(𝐱),𝝀k+3/4(𝐱))=argmin(𝐪(𝐱),𝝁(𝐱))χ[|𝐪(𝐱)𝐪k+2/4(𝐱)|2+γ1|𝝁(𝐱)𝝀k+2/4(𝐱)|2]\displaystyle(\mathbf{p}^{k+3/4}(\mathbf{x}),\boldsymbol{\lambda}^{k+3/4}(\mathbf{x}))=\operatorname*{arg\,min}_{(\mathbf{q}(\mathbf{x}),\boldsymbol{\mu}(\mathbf{x}))\in\chi}\left[|\mathbf{q}(\mathbf{x})-\mathbf{q}^{k+2/4}(\mathbf{x})|^{2}+\gamma_{1}|\boldsymbol{\mu}(\mathbf{x})-\boldsymbol{\lambda}^{k+2/4}(\mathbf{x})|^{2}\right] (24)

with χ={(𝐚,𝐛)2:𝐚𝐛=|𝐚|,|𝐛|1}\chi=\{(\mathbf{a},\mathbf{b})\in\mathbb{R}^{2}:\mathbf{a}\cdot\mathbf{b}=|\mathbf{a}|,|\mathbf{b}|\leq 1\}. Define

χ1={(𝐚,𝐛)2:|𝐚|=0,|𝐛|1},andχ2={(𝐚,𝐛)2:|𝐚|0,𝐚𝐛=|𝐪|,|𝐛|=1}.\displaystyle\chi_{1}=\{(\mathbf{a},\mathbf{b})\in\mathbb{R}^{2}:|\mathbf{a}|=0,|\mathbf{b}|\leq 1\},\;\text{and}\;\chi_{2}=\{(\mathbf{a},\mathbf{b})\in\mathbb{R}^{2}:|\mathbf{a}|\neq 0,\mathbf{a}\cdot\mathbf{b}=|\mathbf{q}|,|\mathbf{b}|=1\}.

We have χ=χ1χ2\chi=\chi_{1}\cup\chi_{2}. We next discuss the solution of (24) on χ1\chi_{1} and χ2\chi_{2} separately.

Over χ1\chi_{1}, denote the solution to (24) by (𝐩^k+3/4(𝐱),𝝀^k+3/4(𝐱))(\widehat{\mathbf{p}}^{k+3/4}(\mathbf{x}),\widehat{\boldsymbol{\lambda}}^{k+3/4}(\mathbf{x})). Problem (24) reduces to

𝝀^(𝐱)k+3/4=argmin𝝁(𝐱)2:|𝝁(𝐱)|1|𝝁(𝐱)𝝀k+2/4(𝐱)|2,\displaystyle\widehat{\boldsymbol{\lambda}}(\mathbf{x})^{k+3/4}=\operatorname*{arg\,min}_{\boldsymbol{\mu}(\mathbf{x})\in\mathbb{R}^{2}:|\boldsymbol{\mu}(\mathbf{x})|\leq 1}|\boldsymbol{\mu}(\mathbf{x})-\boldsymbol{\lambda}^{k+2/4}(\mathbf{x})|^{2}\;,

and 𝐩^k+3/4(𝐱)=𝟎\widehat{\mathbf{p}}^{k+3/4}(\mathbf{x})=\mathbf{0}. The closed-form solution for 𝝀^k+3/4(𝐱)\widehat{\boldsymbol{\lambda}}^{k+3/4}(\mathbf{x}) is given as

𝝀^k+3/4(𝐱)=𝝀k+2/4(𝐱)max{1,|𝝀k+2/4(𝐱)|}.\displaystyle\widehat{\boldsymbol{\lambda}}^{k+3/4}(\mathbf{x})=\frac{\boldsymbol{\lambda}^{k+2/4}(\mathbf{x})}{\max\{1,|\boldsymbol{\lambda}^{k+2/4}(\mathbf{x})|\}}. (25)

Over χ2\chi_{2}, denote the solution to (24) by (𝐩~k+3/4(𝐱),𝝀~k+3/4(𝐱))(\widetilde{\mathbf{p}}^{k+3/4}(\mathbf{x}),\widetilde{\boldsymbol{\lambda}}^{k+3/4}(\mathbf{x})), and it can be written as

(𝐩~k+3/4(𝐱),𝝀~k+3/4(𝐱))=argmin(𝐪(𝐱),𝝁(𝐱))2×2,𝐪(𝐱)𝟎,𝐪(𝐱)𝝁(𝐱)=|𝐪(𝐱)|,|𝝁(𝐱)|=1[|𝐪(𝐱)𝐩k+2/4(𝐱)|2+γ1|𝝁(𝐱)𝝀k+2/4(𝐱)|2].\displaystyle(\widetilde{\mathbf{p}}^{k+3/4}(\mathbf{x}),\widetilde{\boldsymbol{\lambda}}^{k+3/4}(\mathbf{x}))=\operatorname*{arg\,min}_{\begin{subarray}{c}(\mathbf{q}(\mathbf{x}),\boldsymbol{\mu}(\mathbf{x}))\in\mathbb{R}^{2}\times\mathbb{R}^{2},\\ \mathbf{q}(\mathbf{x})\neq\mathbf{0},\mathbf{q}(\mathbf{x})\cdot\boldsymbol{\mu}(\mathbf{x})=|\mathbf{q}(\mathbf{x})|,\\ |\boldsymbol{\mu}(\mathbf{x})|=1\end{subarray}}\left[|\mathbf{q}(\mathbf{x})-\mathbf{p}^{k+2/4}(\mathbf{x})|^{2}+\gamma_{1}|\boldsymbol{\mu}(\mathbf{x})-\boldsymbol{\lambda}^{k+2/4}(\mathbf{x})|^{2}\right]. (26)

Denote θ=|𝐪(𝐱)|\theta=|\mathbf{q}(\mathbf{x})|, 𝐲(𝐱)=𝐩k+2/4(𝐱)\mathbf{y}(\mathbf{x})=\mathbf{p}^{k+2/4}(\mathbf{x}), and 𝐰(𝐱)=𝝁k+2/4(𝐱)\mathbf{w}(\mathbf{x})=\boldsymbol{\mu}^{k+2/4}(\mathbf{x}) for each 𝐱Ω\mathbf{x}\in\Omega. Problem (26) is equivalent to solving

min(θ,𝐪(𝐱))×2,θ>0,|𝐪(𝐱)|=1[θ22θ𝐪(𝐱)𝐲(𝐱)2γ1𝐪(𝐱)𝐰(𝐱)].\displaystyle\min_{(\theta,\mathbf{q}(\mathbf{x}))\in\mathbb{R}\times\mathbb{R}^{2},\theta>0,|\mathbf{q}(\mathbf{x})|=1}\left[\theta^{2}-2\theta\mathbf{q}(\mathbf{x})\cdot\mathbf{y}(\mathbf{x})-2\gamma_{1}\mathbf{q}(\mathbf{x})\cdot\mathbf{w}(\mathbf{x})\right]\;. (27)

For a fixed θ\theta, the solution 𝐪(𝐱,θ)\mathbf{q}^{*}(\mathbf{x},\theta) of (27) is given by

𝐪(𝐱,θ)=θ𝐲(𝐱)+γ1𝐰(𝐱)|θ𝐲(𝐱)+γ1𝐰(𝐱)|.\displaystyle\mathbf{q}^{*}(\mathbf{x},\theta)=\frac{\theta\mathbf{y}(\mathbf{x})+\gamma_{1}\mathbf{w}(\mathbf{x})}{|\theta\mathbf{y}(\mathbf{x})+\gamma_{1}\mathbf{w}(\mathbf{x})|}\;. (28)

Substituting (28) to (27) leads to

minθ>0[θ22|θ𝐲(𝐱)+γ1𝐰(𝐱)|].\displaystyle\min_{\theta>0}\left[\theta^{2}-2|\theta\mathbf{y}(\mathbf{x})+\gamma_{1}\mathbf{w}(\mathbf{x})|\right]. (29)

We solve (29) by a fixed point method. Denote

E(𝐱,θ)=θ22|θ𝐲(𝐱)+γ1𝐰(𝐱)|,\displaystyle E(\mathbf{x},\theta)=\theta^{2}-2|\theta\mathbf{y}(\mathbf{x})+\gamma_{1}\mathbf{w}(\mathbf{x})|\;,

and thus

dEdθ(𝐱,θ)=2θ2𝐲(𝐱)(θ𝐲(𝐱)+γ1𝐰(𝐱))|θ𝐲(𝐱)+γ1𝐰(𝐱)|.\displaystyle\frac{dE}{d\theta}(\mathbf{x},\theta)=2\theta-\frac{2\mathbf{y}(\mathbf{x})\cdot(\theta\mathbf{y}(\mathbf{x})+\gamma_{1}\mathbf{w}(\mathbf{x}))}{|\theta\mathbf{y}(\mathbf{x})+\gamma_{1}\mathbf{w}(\mathbf{x})|}.

We initialize θ0=|𝐲(𝐱)|\theta^{0}=|\mathbf{y}(\mathbf{x})|. For m=0,1,2,m=0,1,2,\dots, we update θm+1\theta^{m+1} from θm\theta^{m} via

θm+1=max{0,𝐲(𝐱)(θm𝐲(𝐱)+γ1𝐰(𝐱))|θm𝐲(𝐱)+γ1𝐰(𝐱)|}\displaystyle\theta^{m+1}=\max\left\{0,\frac{\mathbf{y}(\mathbf{x})\cdot(\theta^{m}\mathbf{y}(\mathbf{x})+\gamma_{1}\mathbf{w}(\mathbf{x}))}{|\theta^{m}\mathbf{y}(\mathbf{x})+\gamma_{1}\mathbf{w}(\mathbf{x})|}\right\} (30)

until |θm+1θm|<ε|\theta^{m+1}-\theta^{m}|<\varepsilon for some small ε>0\varepsilon>0.

Denote the converged θ\theta by θ(𝐱)\theta^{*}(\mathbf{x}), then we compute

𝝀~k+3/4(𝐱)=θ(𝐱)𝐩k+2/4(𝐱)+γ1𝝀k+2/4(𝐱)|θ(𝐱)𝐩k+2/4(𝐱)+γ1𝝀k+2/4(𝐱)|,𝐩~k+3/4(𝐱)=θ𝝀~k+3/4(𝐱).\displaystyle\widetilde{\boldsymbol{\lambda}}^{k+3/4}(\mathbf{x})=\frac{\theta^{*}(\mathbf{x})\mathbf{p}^{k+2/4}(\mathbf{x})+\gamma_{1}\boldsymbol{\lambda}^{k+2/4}(\mathbf{x})}{|\theta^{*}(\mathbf{x})\mathbf{p}^{k+2/4}(\mathbf{x})+\gamma_{1}\boldsymbol{\lambda}^{k+2/4}(\mathbf{x})|},\quad\widetilde{\mathbf{p}}^{k+3/4}(\mathbf{x})=\theta^{*}\widetilde{\boldsymbol{\lambda}}^{k+3/4}(\mathbf{x})\;. (31)

Finally, we define

G(𝐪,𝝁;𝐱)=|𝐪(𝐱)𝐪k+2/4(𝐱)|2+γ1|𝝁(𝐱)𝝀k+2/4(𝐱)|2,\displaystyle G(\mathbf{q},\boldsymbol{\mu};\mathbf{x})=|\mathbf{q}(\mathbf{x})-\mathbf{q}^{k+2/4}(\mathbf{x})|^{2}+\gamma_{1}|\boldsymbol{\mu}(\mathbf{x})-\boldsymbol{\lambda}^{k+2/4}(\mathbf{x})|^{2},

and compare the minimizers in χ1\chi_{1} and χ2\chi_{2} to obtain (𝐩k+3/4(𝐱),𝝀k+3/4(𝐱))(\mathbf{p}^{k+3/4}(\mathbf{x}),\boldsymbol{\lambda}^{k+3/4}(\mathbf{x})) as

(𝐩k+3/4(𝐱),𝝀k+3/4(𝐱))={(𝐩^k+3/4(𝐱),𝝀^k+3/4(𝐱)) if G(𝐩^k+3/4,𝝀^k+3/4;𝐱)G(𝐩~k+3/4,𝝀~k+3/4;𝐱),(𝐩~k+3/4(𝐱),𝝀~k+3/4(𝐱)) otherwise.\displaystyle(\mathbf{p}^{k+3/4}(\mathbf{x}),\boldsymbol{\lambda}^{k+3/4}(\mathbf{x}))=\begin{cases}(\widehat{\mathbf{p}}^{k+3/4}(\mathbf{x}),\widehat{\boldsymbol{\lambda}}^{k+3/4}(\mathbf{x}))&\mbox{ if }G(\widehat{\mathbf{p}}^{k+3/4},\widehat{\boldsymbol{\lambda}}^{k+3/4};\mathbf{x})\\ &\hskip 28.45274pt\leq G(\widetilde{\mathbf{p}}^{k+3/4},\widetilde{\boldsymbol{\lambda}}^{k+3/4};\mathbf{x}),\\ (\widetilde{\mathbf{p}}^{k+3/4}(\mathbf{x}),\widetilde{\boldsymbol{\lambda}}^{k+3/4}(\mathbf{x}))&\mbox{ otherwise}.\end{cases} (32)

3.4 On the solution to sub-problem (18)

In (18), (𝐩k+1,rk+1,𝐬k+1)(\mathbf{p}^{k+1},r^{k+1},\mathbf{s}^{k+1}) solves

min(𝐪,w,𝐠)Σf,wH2(Ω),𝐠(H1(Ω))2\displaystyle\min_{\begin{subarray}{c}(\mathbf{q},w,\mathbf{g})\in\Sigma_{f},w\in H^{2}(\Omega),\\ \mathbf{g}\in(H^{1}(\Omega))^{2}\end{subarray}} [12Ω|𝐪𝐩k+3/4|2d𝐱+γ22Ω|wrk+3/4|2d𝐱+γ32Ω|𝐠𝐬k+3/4|2d𝐱\displaystyle\bigg{[}\frac{1}{2}\int_{\Omega}|\mathbf{q}-\mathbf{p}^{k+3/4}|^{2}d\mathbf{x}+\frac{\gamma_{2}}{2}\int_{\Omega}|w-r^{k+3/4}|^{2}d\mathbf{x}+\frac{\gamma_{3}}{2}\int_{\Omega}|\mathbf{g}-\mathbf{s}^{k+3/4}|^{2}d\mathbf{x}
+ταwΩ|2w|2d𝐱+ταnΩ|𝐠|2d𝐱+τ2Ω|f(v𝐪,w,𝐠+w+𝐠)|2d𝐱],\displaystyle+\tau\alpha_{w}\int_{\Omega}|\nabla^{2}w|^{2}d\mathbf{x}+\tau\alpha_{n}\int_{\Omega}|\mathbf{g}|^{2}d\mathbf{x}+\frac{\tau}{2}\int_{\Omega}|f-(v_{\mathbf{q},w,\mathbf{g}}+w+\nabla\cdot\mathbf{g})|^{2}d\mathbf{x}\bigg{]}, (33)

or equivalently

{(vk+1,rk+1,𝐬k+1)=argminvH1(Ω),wH2(Ω),𝐠(H1(Ω))2[12Ω|v𝐩k+3/4|2d𝐱+γ22Ω|wrk+3/4|2d𝐱+γ32Ω|𝐠𝐬k+3/4|2𝑑𝐱+ταwΩ|2w|2𝑑𝐱+ταnΩ|𝐠|2𝑑𝐱+τ2Ω|f(v+w+𝐠)|2d𝐱],𝐩k+1=vk+1.\displaystyle\begin{cases}(v^{k+1},r^{k+1},\mathbf{s}^{k+1})=\displaystyle\operatorname*{arg\,min}_{\begin{subarray}{c}v\in H^{1}(\Omega),w\in H^{2}(\Omega),\\ \mathbf{g}\in(H^{1}(\Omega))^{2}\end{subarray}}\bigg{[}\frac{1}{2}\int_{\Omega}|\nabla v-\mathbf{p}^{k+3/4}|^{2}d\mathbf{x}+\frac{\gamma_{2}}{2}\int_{\Omega}|w-r^{k+3/4}|^{2}d\mathbf{x}\\ \hskip 113.81102pt+\frac{\gamma_{3}}{2}\displaystyle\int_{\Omega}|\mathbf{g}-\mathbf{s}^{k+3/4}|^{2}d\mathbf{x}+\tau\alpha_{w}\int_{\Omega}|\nabla^{2}w|^{2}d\mathbf{x}+\tau\alpha_{n}\int_{\Omega}|\mathbf{g}|^{2}d\mathbf{x}\\ \hskip 113.81102pt+\frac{\tau}{2}\displaystyle\int_{\Omega}|f-(v+w+\nabla\cdot\mathbf{g})|^{2}d\mathbf{x}\bigg{]},\\ \mathbf{p}^{k+1}=\nabla v^{k+1}.\end{cases} (34)

The optimality condition for (vk+1,rk+1,𝐬k+1)(v^{k+1},r^{k+1},\mathbf{s}^{k+1}) is

{{2vk+1+𝐩k+3/4τ+vk+1(frk+1𝐬k+1)=0 in Ω,(vk+1𝐩k+3/4)𝐧=0 on Ω,{γ2rk+1rk+3/4τ+rk+1(fvk+1𝐬k+1)+2αw(4rk+1)=0 in Ω,2rk+1=0,3rk+1𝐧=0 on Ω,{γ3𝐬k+1𝐬k+3/4τ(𝐬k+1)+2αn𝐬k+1+(fvk+1rk+1)=0 in Ω,f(vk+1+rk+1+𝐬k+1)=0 on Ω.\displaystyle\begin{cases}\begin{cases}\frac{-\nabla^{2}v^{k+1}+\nabla\cdot\mathbf{p}^{k+3/4}}{\tau}+v^{k+1}-(f-r^{k+1}-\nabla\cdot\mathbf{s}^{k+1})=0&\mbox{ in }\Omega,\\ (\nabla v^{k+1}-\mathbf{p}^{k+3/4})\cdot\mathbf{n}=0&\mbox{ on }\partial\Omega,\end{cases}\\ \begin{cases}\gamma_{2}\frac{r^{k+1}-r^{k+3/4}}{\tau}+r^{k+1}-(f-v^{k+1}-\nabla\cdot\mathbf{s}^{k+1})+2\alpha_{w}(\nabla^{4}r^{k+1})=0&\mbox{ in }\Omega,\\ \nabla^{2}r^{k+1}=0,\ \nabla^{3}r^{k+1}\cdot\mathbf{n}=0&\mbox{ on }\partial\Omega,\end{cases}\\ \begin{cases}\gamma_{3}\frac{\mathbf{s}^{k+1}-\mathbf{s}^{k+3/4}}{\tau}-\nabla(\nabla\cdot\mathbf{s}^{k+1})+2\alpha_{n}\mathbf{s}^{k+1}+\nabla(f-v^{k+1}-r^{k+1})=0&\mbox{ in }\Omega,\\ f-(v^{k+1}+r^{k+1}+\nabla\cdot\mathbf{s}^{k+1})=0&\mbox{ on }\partial\Omega.\end{cases}\end{cases} (35)

The proposed operator splitting scheme is summarized in Algorithm 1.

0:  Gray image ff; model parameters: α0,αcurv,αw,αn\alpha_{0},\alpha_{\text{curv}},\alpha_{w},\alpha_{n}; algorithmic parameters: time step τ\tau, evolution speed factor γi\gamma_{i}, i=1,2,3i=1,2,3, threshold ρ\rho for termination, maximal number of iteration Itermax\text{Iter}_{\max}
1:  for k=1,,Itermaxk=1,\dots,\text{Iter}_{\max} do
2:     Update 𝐩k𝐩k+1/4\mathbf{p}^{k}\to\mathbf{p}^{k+1/4} by (19), and set 𝝀k+1/4=𝝀k\boldsymbol{\lambda}^{k+1/4}=\boldsymbol{\lambda}^{k}, rk+1/4=rkr^{k+1/4}=r^{k}, 𝐬k+1/4=𝐬k\mathbf{s}^{k+1/4}=\mathbf{s}^{k}.
3:     Update 𝐩n+1/4𝐩k+2/4\mathbf{p}^{n+1/4}\to\mathbf{p}^{k+2/4} by (20), 𝝀k+1/4𝝀k+2/4\boldsymbol{\lambda}^{k+1/4}\to\boldsymbol{\lambda}^{k+2/4} by solving (22), and set rk+2/4=rk+1/4r^{k+2/4}=r^{k+1/4}, 𝐬k+2/4=𝐬k+1/4\mathbf{s}^{k+2/4}=\mathbf{s}^{k+1/4}.
4:     Update 𝐩k+2/4𝐩k+3/4\mathbf{p}^{k+2/4}\to\mathbf{p}^{k+3/4} and 𝝀k+2/4𝝀k+3/4\boldsymbol{\lambda}^{k+2/4}\to\boldsymbol{\lambda}^{k+3/4} by (32), and set rk+3/4=rk+2/4r^{k+3/4}=r^{k+2/4}, 𝐬k+3/4=𝐬k+2/4\mathbf{s}^{k+3/4}=\mathbf{s}^{k+2/4}.
5:     Compute vk+1v^{k+1}, rk+1r^{k+1}, and 𝐬k+1\mathbf{s}^{k+1} by solving (35), and set 𝐩k+1=vk+1\mathbf{p}^{k+1}=\nabla v^{k+1}, nk+1=𝐬k+1n^{k+1}=\nabla\cdot\mathbf{s}^{k+1}, 𝝀k+1=𝝀k+3/4\boldsymbol{\lambda}^{k+1}=\boldsymbol{\lambda}^{k+3/4}.
6:     if max{rk+1rk2rk2,vk+1vn2vk2}<ρ\max\{\frac{\|r^{k+1}-r^{k}\|_{2}}{\|r^{k}\|_{2}},\frac{\|v^{k+1}-v^{n}\|_{2}}{\|v^{k}\|_{2}}\}<\rho then
7:        break.
8:     end if
9:  end for
10:  return  the converged homogeneous structure vv^{*}, smooth part ww^{*}, and oscillatory part nn^{*}, such that ff is approximated by v+w+nv^{*}+w^{*}+n^{*}.
Algorithm 1 Proposed operator-splitting algorithm

3.5 Adapt to periodic boundary conditions for efficiency

We note that while (22) and (35) contain high-order derivatives, they are linear. Therefore, we can leverage the Fast Fourier Transform (FFT) to convert these PDE systems to pointwise algebraic systems, for which we can deduce explicit solutions. To garnish such computational efficiency, we extend our model (7) to periodic data, and this requires slight modifications on some boundary conditions.

Without loss of generality, consider a rectangular domain Ω=[0,L1]×[0,L2]\Omega=[0,L_{1}]\times[0,L_{2}] and define the spaces

HPk(Ω)={vHk(Ω):v(0,y)=v(L1,y),v(:,0)=v(:,L2)}\displaystyle H^{k}_{P}(\Omega)=\{v\in H^{k}(\Omega):v(0,y)=v(L_{1},y),\ v(:,0)=v(:,L_{2})\} (36)

for k=1,2k=1,2. We revise the set Σf\Sigma_{f} to

Σf={(𝐪,w,𝐠)(\displaystyle\Sigma_{f}=\big{\{}(\mathbf{q},w,\mathbf{g})\in( L2(Ω))2×L2(Ω)×(H1(Ω))2|\displaystyle L^{2}(\Omega))^{2}\times L^{2}(\Omega)\times(H^{1}(\Omega))^{2}|
vHP1(Ω)such that𝐪=vandΩ(v+w+𝐠f)d𝐱=0}.\displaystyle\exists v\in H_{P}^{1}(\Omega)\leavevmode\nobreak\ \text{such that}\leavevmode\nobreak\ \mathbf{q}=\nabla v\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ \int_{\Omega}(v+w+\nabla\cdot\mathbf{g}-f)\,d\mathbf{x}=0\big{\}}. (37)

Problem (12) and (11) are replaced by

min𝐪(HP1(Ω))2,𝝁(HP1(Ω))2,wHP2(Ω),𝐠(HP1(Ω))2[\displaystyle\min_{\begin{subarray}{c}\mathbf{q}\in(H_{P}^{1}(\Omega))^{2},\boldsymbol{\mu}\in(H^{1}_{P}(\Omega))^{2},\\ w\in H^{2}_{P}(\Omega),\mathbf{g}\in(H^{1}_{P}(\Omega))^{2}\end{subarray}}\bigg{[} α0Ω𝐪0𝑑𝐱+αcurvΩ|𝝁|2|𝐪|𝑑𝐱+αwΩ|2w|2𝑑𝐱\displaystyle\alpha_{0}\int_{\Omega}\|\mathbf{q}\|_{0}\,d\mathbf{x}+\alpha_{\rm curv}\int_{\Omega}\left|\nabla\cdot\boldsymbol{\mu}\right|^{2}\left|\mathbf{q}\right|\,d\mathbf{x}+\alpha_{w}\int_{\Omega}|\nabla^{2}w|^{2}d\mathbf{x}
+αnΩ|𝐠|2d𝐱+12Ω|f(v𝐪+w+𝐠)|2d𝐱+IΣf(𝐪,w,𝐠)+IS(𝐪,𝝁)]\displaystyle+\alpha_{n}\int_{\Omega}|\mathbf{g}|^{2}d\mathbf{x}+\frac{1}{2}\int_{\Omega}|f-(v_{\mathbf{q}}+w+\nabla\cdot\mathbf{g})|^{2}d\mathbf{x}+I_{\Sigma_{f}}(\mathbf{q},w,\mathbf{g})+I_{S}(\mathbf{q},\boldsymbol{\mu})\bigg{]} (38)

and

{2v𝐪,w,𝐠=𝐪 in Ω,Ωv𝐪,w,𝐠𝑑𝐱=Ωfw𝐠d𝐱,v𝐪,w,𝐠(0,x2)=v𝐪,w,𝐠(L1,x2) for 0x2L2,v𝐪,w,𝐠(x1,0)=v𝐪,w,𝐠(x1,L2) for 0x1L1,(v𝐪,w,𝐠x1q1)(0,x2)=(v𝐪,w,𝐠x1q1)(L1,x2) for 0x2L2,(v𝐪,w,𝐠x2q2)(x1,0)=(v𝐪,w,𝐠x2q2)(x1,L2) for 0x1L1.\displaystyle\begin{cases}\nabla^{2}v_{\mathbf{q},w,\mathbf{g}}=\nabla\cdot\mathbf{q}&\mbox{ in }\Omega,\\ \int_{\Omega}v_{\mathbf{q},w,\mathbf{g}}\,d\mathbf{x}=\int_{\Omega}f-w-\nabla\cdot\mathbf{g}\,d\mathbf{x},\\ v_{\mathbf{q},w,\mathbf{g}}(0,x_{2})=v_{\mathbf{q},w,\mathbf{g}}(L_{1},x_{2})&\mbox{ for }0\leq x_{2}\leq L_{2},\\ v_{\mathbf{q},w,\mathbf{g}}(x_{1},0)=v_{\mathbf{q},w,\mathbf{g}}(x_{1},L_{2})&\mbox{ for }0\leq x_{1}\leq L_{1},\\ \left(\frac{\partial v_{\mathbf{q},w,\mathbf{g}}}{\partial x_{1}}-q_{1}\right)(0,x_{2})=\left(\frac{\partial v_{\mathbf{q},w,\mathbf{g}}}{\partial x_{1}}-q_{1}\right)(L_{1},x_{2})&\mbox{ for }0\leq x_{2}\leq L_{2},\\ \left(\frac{\partial v_{\mathbf{q},w,\mathbf{g}}}{\partial x_{2}}-q_{2}\right)(x_{1},0)=\left(\frac{\partial v_{\mathbf{q},w,\mathbf{g}}}{\partial x_{2}}-q_{2}\right)(x_{1},L_{2})&\mbox{ for }0\leq x_{1}\leq L_{1}.\end{cases} (39)

For subproblem solvers, we replace (21) and (22) by

𝝀k+2/4=argmin𝝁(HP1(Ω))2[γ12Ω|𝝁𝝀k+1/4|22𝑑𝐱+ταwΩ|𝝁|2|𝐩k+2/4|𝑑𝐱]\displaystyle\boldsymbol{\lambda}^{k+2/4}=\operatorname*{arg\,min}_{\boldsymbol{\mu}\in(H_{P}^{1}(\Omega))^{2}}\left[\frac{\gamma_{1}}{2}\int_{\Omega}|\boldsymbol{\mu}-\boldsymbol{\lambda}^{k+1/4}|_{2}^{2}d\mathbf{x}+\tau\alpha_{w}\int_{\Omega}\left|\nabla\cdot\boldsymbol{\mu}\right|^{2}|\mathbf{p}^{k+2/4}|d\mathbf{x}\right] (40)

and

{γ1𝝀k+2/4𝝀k+1/4τ2αw(|𝐩k+2/4|𝝀k+2/4)=0 in Ω,(|𝐩k+2/4|𝝀k+2/4)(0,x2)=(|𝐩k+2/4|𝝀k+2/4)(L1,x2) for 0x2L2,(|𝐩k+2/4|𝝀k+2/4)(x1,0)=(|𝐩k+2/4|𝝀k+2/4)(x1,L1) for 0x1L1,\displaystyle\begin{cases}\gamma_{1}\frac{\boldsymbol{\lambda}^{k+2/4}-\boldsymbol{\lambda}^{k+1/4}}{\tau}-2\alpha_{w}\nabla(|\mathbf{p}^{k+2/4}|\nabla\cdot\boldsymbol{\lambda}^{k+2/4})=0&\mbox{ in }\Omega,\\ \left(|\mathbf{p}^{k+2/4}|\nabla\cdot\boldsymbol{\lambda}^{k+2/4}\right)(0,x_{2})=\left(|\mathbf{p}^{k+2/4}|\nabla\cdot\boldsymbol{\lambda}^{k+2/4}\right)(L_{1},x_{2})&\mbox{ for }0\leq x_{2}\leq L_{2},\\ \left(|\mathbf{p}^{k+2/4}|\nabla\cdot\boldsymbol{\lambda}^{k+2/4}\right)(x_{1},0)=\left(|\mathbf{p}^{k+2/4}|\nabla\cdot\boldsymbol{\lambda}^{k+2/4}\right)(x_{1},L_{1})&\mbox{ for }0\leq x_{1}\leq L_{1},\end{cases} (41)

respectively.

Replace (33), (34) and (35) by

min(𝐪,w,𝐠)Σf,wHP2(Ω),𝐠(HP1(Ω))2\displaystyle\min_{\begin{subarray}{c}(\mathbf{q},w,\mathbf{g})\in\Sigma_{f},w\in H_{P}^{2}(\Omega),\\ \mathbf{g}\in(H_{P}^{1}(\Omega))^{2}\end{subarray}} [12Ω|𝐪𝐩k+3/4|2d𝐱+γ22Ω|wrk+3/4|2d𝐱+γ32Ω|𝐠𝐬k+3/4|2d𝐱\displaystyle\bigg{[}\frac{1}{2}\int_{\Omega}|\mathbf{q}-\mathbf{p}^{k+3/4}|^{2}d\mathbf{x}+\frac{\gamma_{2}}{2}\int_{\Omega}|w-r^{k+3/4}|^{2}d\mathbf{x}+\frac{\gamma_{3}}{2}\int_{\Omega}|\mathbf{g}-\mathbf{s}^{k+3/4}|^{2}d\mathbf{x}
+ταwΩ|2w|2d𝐱+ταnΩ|𝐠|2d𝐱+τ2Ω|f(v𝐪,w,𝐠+w+𝐠)|2d𝐱],\displaystyle+\tau\alpha_{w}\int_{\Omega}|\nabla^{2}w|^{2}d\mathbf{x}+\tau\alpha_{n}\int_{\Omega}|\mathbf{g}|^{2}d\mathbf{x}+\frac{\tau}{2}\int_{\Omega}|f-(v_{\mathbf{q},w,\mathbf{g}}+w+\nabla\cdot\mathbf{g})|^{2}d\mathbf{x}\bigg{]}, (42)
{(vk+1,rk+1,𝐬k+1)=argminvHP1(Ω),wHP2(Ω),𝐠(HP1(Ω))2[12Ω|v𝐩k+3/4|2d𝐱+γ22Ω|wrk+3/4|2d𝐱+γ32Ω|𝐠𝐬k+3/4|2𝑑𝐱+ταwΩ|2w|2𝑑𝐱+ταnΩ|𝐠|2𝑑𝐱+τ2Ω|f(v+w+𝐠)|2d𝐱],𝐩k+1=vk+1,\displaystyle\begin{cases}(v^{k+1},r^{k+1},\mathbf{s}^{k+1})=\displaystyle\operatorname*{arg\,min}_{\begin{subarray}{c}v\in H_{P}^{1}(\Omega),w\in H_{P}^{2}(\Omega),\\ \mathbf{g}\in(H_{P}^{1}(\Omega))^{2}\end{subarray}}\bigg{[}\frac{1}{2}\int_{\Omega}|\nabla v-\mathbf{p}^{k+3/4}|^{2}d\mathbf{x}+\frac{\gamma_{2}}{2}\int_{\Omega}|w-r^{k+3/4}|^{2}d\mathbf{x}\\ \hskip 113.81102pt+\frac{\gamma_{3}}{2}\displaystyle\int_{\Omega}|\mathbf{g}-\mathbf{s}^{k+3/4}|^{2}d\mathbf{x}+\tau\alpha_{w}\int_{\Omega}|\nabla^{2}w|^{2}d\mathbf{x}+\tau\alpha_{n}\int_{\Omega}|\mathbf{g}|^{2}d\mathbf{x}\\ \hskip 113.81102pt+\frac{\tau}{2}\displaystyle\int_{\Omega}|f-(v+w+\nabla\cdot\mathbf{g})|^{2}d\mathbf{x}\bigg{]},\\ \mathbf{p}^{k+1}=\nabla v^{k+1},\end{cases} (43)

and

{{2vk+1+𝐩k+3/4τ+vk+1(frk+1𝐬k+1)=0, in Ω,(vk+1x1p1k+3/4)(0,x2)=(vk+1x1p1k+3/4)(L1,x2) for 0x2L2,(vk+1x2p2k+3/4)(x1,0)=(vk+1x2p2k+3/4)(x1,L2) for 0x1L1,{γ2rk+1rk+3/4τ+rk+1(fvk+1𝐬k+1)+2αw(4rk+1)=0 in Ω,(2rk+1)(0,x2)=(2rk+1)(L1,x2),((2rk+1)x1)(0,x2)=((2rk+1)x1)(L1,x2) for 0x2L2,(2rk+1)(x1,0)=(2rk+1)(x1,L2),((2rk+1)x2)(x1,0)=((2rk+1)x2)(x1,L2) for 0x1L1,{γ3𝐬k+1𝐬k+3/4τ(𝐬k+1)+2αn𝐬k+1+(fvk+1rk+1)=0 in Ω,(f(vk+1+rk+1+𝐬k+1))(0,x2)=(f(vk+1+rk+1+𝐬k+1))(L1,x2) for 0x2L2,(f(vk+1+rk+1+𝐬k+1))(x1,0)=(f(vk+1+rk+1+𝐬k+1))(x1,L2) for 0x1L1,\displaystyle\begin{cases}\begin{cases}\frac{-\nabla^{2}v^{k+1}+\nabla\cdot\mathbf{p}^{k+3/4}}{\tau}+v^{k+1}-(f-r^{k+1}-\nabla\cdot\mathbf{s}^{k+1})=0,&\mbox{ in }\Omega,\\ \left(\frac{\partial v^{k+1}}{\partial x_{1}}-p_{1}^{k+3/4}\right)(0,x_{2})=\left(\frac{\partial v^{k+1}}{\partial x_{1}}-p_{1}^{k+3/4}\right)(L_{1},x_{2})&\mbox{ for }0\leq x_{2}\leq L_{2},\\ \left(\frac{\partial v^{k+1}}{\partial x_{2}}-p_{2}^{k+3/4}\right)(x_{1},0)=\left(\frac{\partial v^{k+1}}{\partial x_{2}}-p_{2}^{k+3/4}\right)(x_{1},L_{2})&\mbox{ for }0\leq x_{1}\leq L_{1},\end{cases}\\ \begin{cases}\gamma_{2}\frac{r^{k+1}-r^{k+3/4}}{\tau}+r^{k+1}-(f-v^{k+1}-\nabla\cdot\mathbf{s}^{k+1})+2\alpha_{w}(\nabla^{4}r^{k+1})=0&\mbox{ in }\Omega,\\ (\nabla^{2}r^{k+1})(0,x_{2})=(\nabla^{2}r^{k+1})(L_{1},x_{2}),\ \left(\frac{\partial(\nabla^{2}r^{k+1})}{\partial x_{1}}\right)(0,x_{2})=\left(\frac{\partial(\nabla^{2}r^{k+1})}{\partial x_{1}}\right)(L_{1},x_{2})&\mbox{ for }0\leq x_{2}\leq L_{2},\\ (\nabla^{2}r^{k+1})(x_{1},0)=(\nabla^{2}r^{k+1})(x_{1},L_{2}),\ \left(\frac{\partial(\nabla^{2}r^{k+1})}{\partial x_{2}}\right)(x_{1},0)=\left(\frac{\partial(\nabla^{2}r^{k+1})}{\partial x_{2}}\right)(x_{1},L_{2})&\mbox{ for }0\leq x_{1}\leq L_{1},\end{cases}\\ \begin{cases}\gamma_{3}\frac{\mathbf{s}^{k+1}-\mathbf{s}^{k+3/4}}{\tau}-\nabla(\nabla\cdot\mathbf{s}^{k+1})+2\alpha_{n}\mathbf{s}^{k+1}+\nabla(f-v^{k+1}-r^{k+1})=0&\mbox{ in }\Omega,\\ (f-(v^{k+1}+r^{k+1}+\nabla\cdot\mathbf{s}^{k+1}))(0,x_{2})=(f-(v^{k+1}+r^{k+1}+\nabla\cdot\mathbf{s}^{k+1}))(L_{1},x_{2})&\mbox{ for }0\leq x_{2}\leq L_{2},\\ (f-(v^{k+1}+r^{k+1}+\nabla\cdot\mathbf{s}^{k+1}))(x_{1},0)=(f-(v^{k+1}+r^{k+1}+\nabla\cdot\mathbf{s}^{k+1}))(x_{1},L_{2})&\mbox{ for }0\leq x_{1}\leq L_{1},\end{cases}\end{cases} (44)

respectively, with 𝐩=(p1,p2)\mathbf{p}=(p_{1},p_{2}). In the rest of this paper, we always assume these adaptations and consider the periodic boundary condition.

4 Numerical Discretization

In this section, we provide implementation details for solving the proposed minimization problem.

4.1 Synopsis

Let Ω\Omega be a rectangular domain with M×NM\times N pixels. Denote the two spatial directions by x1,x2x_{1},x_{2} with step size Δx1=Δx2=h\Delta x_{1}=\Delta x_{2}=h for some h>0h>0. For an image ff defined on Ω\Omega, we denote f(i,j)=f(ih,jh)f(i,j)=f(ih,jh). Assume that ff satisfies the periodic boundary condition. We define the backward (-) and forward (++) approximation for f/x1\partial f/\partial x_{1} and f/x2\partial f/\partial x_{2} as

1f(i,j)={f(i,j)f(i1,j)h 1<iM,f(1,j)f(M,j)hi=1,\displaystyle\partial_{1}^{-}f(i,j)=\begin{cases}\frac{f(i,j)-f(i-1,j)}{h}&\ 1<i\leq M,\\ \frac{f(1,j)-f(M,j)}{h}&\ i=1,\end{cases} 1+f(i,j)={f(i+1,j)f(i,j)h 1i<M,f(1,j)f(M,j)hi=M,\displaystyle\partial_{1}^{+}f(i,j)=\begin{cases}\frac{f(i+1,j)-f(i,j)}{h}&\ 1\leq i<M,\\ \frac{f(1,j)-f(M,j)}{h}&\ i=M,\end{cases}
2f(i,j)={f(i,j)f(i,j1)h 1<iN,f(i,1)f(i,N)hi=1,\displaystyle\partial_{2}^{-}f(i,j)=\begin{cases}\frac{f(i,j)-f(i,j-1)}{h}&\ 1<i\leq N,\\ \frac{f(i,1)-f(i,N)}{h}&\ i=1,\end{cases} 2+f(i,j)={f(i,j+1)f(i,j)h 1i<N,f(i,1)f(i,N)hi=N.\displaystyle\partial_{2}^{+}f(i,j)=\begin{cases}\frac{f(i,j+1)-f(i,j)}{h}&\ 1\leq i<N,\\ \frac{f(i,1)-f(i,N)}{h}&\ i=N.\end{cases}

For any scalar valued function ff and vector valued function 𝐩=(p1,p2)\mathbf{p}=(p_{1},p_{2}), the backward (-) and forward (++) approximation for gradient and divergence are defined as

±f(i,j)=(1±f(i,j),2±f(i,j)),div±𝐩(i,j)=1±p1(i,j)+2±p2(i,j).\displaystyle\nabla^{\pm}f(i,j)=(\partial^{\pm}_{1}f(i,j),\partial^{\pm}_{2}f(i,j)),\quad\text{div}^{\pm}\mathbf{p}(i,j)=\partial_{1}^{\pm}p_{1}(i,j)+\partial^{\pm}_{2}p_{2}(i,j).

The discretized Laplacian is defined as

2f(i,j)=div(+f(i,j))=11+f(i,j)+22+f(i,j),\displaystyle\nabla^{2}f(i,j)=\text{div}^{-}(\nabla^{+}f(i,j))=\partial_{1}^{-}\partial_{1}^{+}f(i,j)+\partial_{2}^{-}\partial_{2}^{+}f(i,j)\;,

which recovers the central difference scheme.

Define the shifting and identity operator as

(𝒮1±f)(i,j)=f(i±1,j),(𝒮2±f)(i,j)=f(i,j±1),(f)(i,j)=f(i,j),\displaystyle(\mathcal{S}_{1}^{\pm}f)(i,j)=f(i\pm 1,j),\quad(\mathcal{S}_{2}^{\pm}f)(i,j)=f(i,j\pm 1),\quad(\mathcal{I}f)(i,j)=f(i,j)\;,

where the periodic boundary condition is assumed. Denote the discrete Fourier transform and its inverse by \mathcal{F} and 1\mathcal{F}^{-1}, respectively. We have

(𝒮1±f)(i,j)=e±2π1(i1)/M(f)(i,j),and(𝒮2±f)(i,j)=e±2π1(j1)/N(f)(i,j).\displaystyle\mathcal{F}(\mathcal{S}_{1}^{\pm}f)(i,j)=e^{\pm 2\pi\sqrt{-1}(i-1)/M}\mathcal{F}(f)(i,j)\;,\text{and}\leavevmode\nobreak\ \mathcal{F}(\mathcal{S}_{2}^{\pm}f)(i,j)=e^{\pm 2\pi\sqrt{-1}(j-1)/N}\mathcal{F}(f)(i,j)\;.

We use Real()\text{Real}(\cdot) to denote the real part of a complex quantity.

4.2 Discrete analogue of 𝐩k+1/4\mathbf{p}^{k+1/4}, 𝐩k+2/4\mathbf{p}^{k+2/4} and 𝝀k+2/4\boldsymbol{\lambda}^{k+2/4}

For 𝐩k+1/4\mathbf{p}^{k+1/4}, we discretize (19) as

𝐩k+1/4(i,j)={𝟎 if (p1k(i,j))2+(p2k(i,j))2α0τ,𝐩k(i,j) otherwise,\displaystyle\mathbf{p}^{k+1/4}(i,j)=\begin{cases}\mathbf{0}&\mbox{ if }(p_{1}^{k}(i,j))^{2}+(p_{2}^{k}(i,j))^{2}\leq\alpha_{0}\tau,\\ \mathbf{p}^{k}(i,j)&\mbox{ otherwise},\end{cases} (45)

and the computational cost is 𝒪(MN)\mathcal{O}(MN).

According to (20), 𝐩k+2/4\mathbf{p}^{k+2/4} is updated pointwisely. We thus have

𝐩k+2/4(i,j)=max{0,1ταw|div𝝀k+1/4(i,j)|2(p1k+1/4(i,j))2+(p2k+2/4(i,j))2}𝐩k+1/4(i,j),\displaystyle\mathbf{p}^{k+2/4}(i,j)=\max\left\{0,1-\frac{\tau\alpha_{w}\left|\text{div}^{-}\boldsymbol{\lambda}^{k+1/4}(i,j)\right|^{2}}{\sqrt{(p_{1}^{k+1/4}(i,j))^{2}+(p_{2}^{k+2/4}(i,j))^{2}}}\right\}\mathbf{p}^{k+1/4}(i,j)\;, (46)

whose computational cost is 𝒪(MN)\mathcal{O}(MN).

To compute 𝝀k+2/4=(λ1k+2/4,λ2k+2/4)\boldsymbol{\lambda}^{k+2/4}=(\lambda_{1}^{k+2/4},\lambda_{2}^{k+2/4}), the updating formula (41) can be rewritten as

γ1𝝀k+2/42ταw(|𝐩k+2/4|𝝀k+2/4)=γ1𝝀k+1/4.\displaystyle\gamma_{1}\boldsymbol{\lambda}^{k+2/4}-2\tau\alpha_{w}\nabla(|\mathbf{p}^{k+2/4}|\nabla\cdot\boldsymbol{\lambda}^{k+2/4})=\gamma_{1}\boldsymbol{\lambda}^{k+1/4}\;. (47)

Instead of directly solving (47), we use the frozen coefficient method [50, 27, 36] and consider the following approximation

γ1𝝀k+2/4c2𝝀k+2/4=γ1𝝀k+1/4c2𝝀k+1/4+2ταw(|𝐩k+2/4|𝝀k+1/4)\displaystyle\gamma_{1}\boldsymbol{\lambda}^{k+2/4}-c\nabla^{2}\boldsymbol{\lambda}^{k+2/4}=\gamma_{1}\boldsymbol{\lambda}^{k+1/4}-c\nabla^{2}\boldsymbol{\lambda}^{k+1/4}+2\tau\alpha_{w}\nabla(|\mathbf{p}^{k+2/4}|\nabla\cdot\boldsymbol{\lambda}^{k+1/4}) (48)

for some small coefficient c>0c>0. In our experiment, we set c=1×109c=1\times 10^{-9}. We discretize (48) as

γ1𝝀k+2/4c+(div𝝀k+2/4)=γ1𝝀k+1/4c+(div𝝀k+1/4)+2ταw+(|𝐩k+2/4|(div𝝀k+1/4)).\displaystyle\gamma_{1}\boldsymbol{\lambda}^{k+2/4}-c\nabla^{+}(\text{div}^{-}\boldsymbol{\lambda}^{k+2/4})=\gamma_{1}\boldsymbol{\lambda}^{k+1/4}-c\nabla^{+}(\text{div}^{-}\boldsymbol{\lambda}^{k+1/4})+2\tau\alpha_{w}\nabla^{+}(|\mathbf{p}^{k+2/4}|(\text{div}^{-}\boldsymbol{\lambda}^{k+1/4})). (49)

Equation (49) can be written in a matrix form as

(γ1c1+1c1+2c2+1γ1c2+2)(λ1k+2/4λ2k+2/4)=(b1b2)\displaystyle\begin{pmatrix}\gamma_{1}-c\partial_{1}^{+}\partial_{1}^{-}&-c\partial_{1}^{+}\partial_{2}^{-}\\ -c\partial_{2}^{+}\partial_{1}^{-}&\gamma_{1}-c\partial_{2}^{+}\partial_{2}^{-}\end{pmatrix}\begin{pmatrix}\lambda^{k+2/4}_{1}\\ \lambda^{k+2/4}_{2}\end{pmatrix}=\begin{pmatrix}b_{1}\\ b_{2}\end{pmatrix} (50)

with

{b1=γ1λ1k+1/4c1+(div𝝀k+1/4)+2ταw1+(|𝐩k+2/4|(div𝝀k+1/4)),b2=γ1λ2k+1/4c2+(div𝝀k+1/4)+2ταw2+(|𝐩k+2/4|(div𝝀k+1/4)).\displaystyle\begin{cases}&b_{1}=\gamma_{1}\lambda_{1}^{k+1/4}-c\partial_{1}^{+}(\text{div}^{-}\boldsymbol{\lambda}^{k+1/4})+2\tau\alpha_{w}\partial_{1}^{+}(|\mathbf{p}^{k+2/4}|(\text{div}^{-}\boldsymbol{\lambda}^{k+1/4})),\\ &b_{2}=\gamma_{1}\lambda_{2}^{k+1/4}-c\partial_{2}^{+}(\text{div}^{-}\boldsymbol{\lambda}^{k+1/4})+2\tau\alpha_{w}\partial_{2}^{+}(|\mathbf{p}^{k+2/4}|(\text{div}^{-}\boldsymbol{\lambda}^{k+1/4})).\end{cases}

System (50) is equivalent to

(γ1c(𝒮1+)(𝒮1)/h2c(𝒮1+)(𝒮2)/h2c(𝒮2+)(𝒮1)/h2γ1c(𝒮2+)(𝒮2)/h2)(λ1k+2/4λ2k+2/4)=(b1b2).\displaystyle\begin{pmatrix}\gamma_{1}-c(\mathcal{S}_{1}^{+}-\mathcal{I})(\mathcal{I}-\mathcal{S}_{1}^{-})/h^{2}&-c(\mathcal{S}_{1}^{+}-\mathcal{I})(\mathcal{I}-\mathcal{S}_{2}^{-})/h^{2}\\ -c(\mathcal{S}_{2}^{+}-\mathcal{I})(\mathcal{I}-\mathcal{S}_{1}^{-})/h^{2}&\gamma_{1}-c(\mathcal{S}_{2}^{+}-\mathcal{I})(\mathcal{I}-\mathcal{S}_{2}^{-})/h^{2}\end{pmatrix}\begin{pmatrix}\lambda^{k+2/4}_{1}\\ \lambda^{k+2/4}_{2}\end{pmatrix}=\begin{pmatrix}b_{1}\\ b_{2}\end{pmatrix}. (51)

Applying discrete Fourier transform on both sides of (51) gives

(a11a12a21a22)(μ1k+2/4μ2k+2/4)=(b1b2),\displaystyle\begin{pmatrix}a_{11}&a_{12}\\ a_{21}&a_{22}\end{pmatrix}\mathcal{F}\begin{pmatrix}\mu^{k+2/4}_{1}\\ \mu^{k+2/4}_{2}\end{pmatrix}=\mathcal{F}\begin{pmatrix}b_{1}\\ b_{2}\end{pmatrix}, (52)

where

a11(i,j)=γ12c(cosζi1)/h2,a22(i,j)=γ12c(cosηj1)/h2,\displaystyle a_{11}(i,j)=\gamma_{1}-2c(\cos\zeta_{i}-1)/h^{2},\ a_{22}(i,j)=\gamma_{1}-2c(\cos\eta_{j}-1)/h^{2},
a12(i,j)=c(cosζi+1sinζi1)(cosηj1sinηj1)/h2,\displaystyle a_{12}(i,j)=c(\cos\zeta_{i}+\sqrt{-1}\sin\zeta_{i}-1)(\cos\eta_{j}-\sqrt{-1}\sin\eta_{j}-1)/h^{2},
a21(i,j)=c(cosηj+1sinηj1)(cosζi1sinζi1)/h2,\displaystyle a_{21}(i,j)=c(\cos\eta_{j}+\sqrt{-1}\sin\eta_{j}-1)(\cos\zeta_{i}-\sqrt{-1}\sin\zeta_{i}-1)/h^{2},

with ζi=2π(i1)/M,ηj=2π(j1)/N\zeta_{i}=2\pi(i-1)/M,\ \eta_{j}=2\pi(j-1)/N for i=1,,Mi=1,...,M and j=1,,Nj=1,...,N. We thus update 𝝀k+2/4\boldsymbol{\lambda}^{k+2/4} via

(λ1k+2/4λ2k+2/4)=Real(1[1a11a22a12a21(a11(b1)a12(b2)a21(b1)+a22(b2))]).\displaystyle\begin{pmatrix}\lambda_{1}^{k+2/4}\\ \lambda_{2}^{k+2/4}\end{pmatrix}=\text{Real}\left(\mathcal{F}^{-1}\left[\frac{1}{a_{11}a_{22}-a_{12}a_{21}}\begin{pmatrix}a_{11}\mathcal{F}(b_{1})-a_{12}\mathcal{F}(b_{2})\\ -a_{21}\mathcal{F}(b_{1})+a_{22}\mathcal{F}(b_{2})\end{pmatrix}\right]\right)\;. (53)

Since the linear system is solved explicitly under the forward and inverse FFT, the computational cost is 𝒪(MN(logM+logN))\mathcal{O}(MN(\log M+\log N)).

4.3 Discrete analogue of 𝐩k+3/4\mathbf{p}^{k+3/4}, 𝝀k+3/4\boldsymbol{\lambda}^{k+3/4}

According to Section 3.3, 𝐩k+3/4\mathbf{p}^{k+3/4} and 𝝀k+3/4\boldsymbol{\lambda}^{k+3/4} can be computed pointwisely. On χ1\chi_{1}, for each i,ji,j, we set 𝐩^k+3/4(i,j)=𝟎\widehat{\mathbf{p}}^{k+3/4}(i,j)=\mathbf{0} and discretize (25) as

𝝀^k+3/4(i,j)=𝝀k+2/4(i,j)max{1,|𝝀k+2/4(i,j)|2}.\displaystyle\widehat{\boldsymbol{\lambda}}^{k+3/4}(i,j)=\frac{\boldsymbol{\lambda}^{k+2/4}(i,j)}{\max\left\{1,\sqrt{|\boldsymbol{\lambda}^{k+2/4}(i,j)|^{2}}\right\}}. (54)

On χ2\chi_{2}, we discretize (31) as

𝝀~k+3/4(i,j)=θ𝐩k+2/4(i,j)+γ1𝝀k+2/4(i,j)|θ𝐩k+2/4(i,j)+γ1𝝀k+2/4(i,j)|,𝐩~k+3/4(i,j)=θ𝝀~k+3/4(i,j),\displaystyle\widetilde{\boldsymbol{\lambda}}^{k+3/4}(i,j)=\frac{\theta^{*}\mathbf{p}^{k+2/4}(i,j)+\gamma_{1}\boldsymbol{\lambda}^{k+2/4}(i,j)}{|\theta^{*}\mathbf{p}^{k+2/4}(i,j)+\gamma_{1}\boldsymbol{\lambda}^{k+2/4}(i,j)|},\quad\widetilde{\mathbf{p}}^{k+3/4}(i,j)=\theta^{*}\widetilde{\boldsymbol{\lambda}}^{k+3/4}(i,j)\;,

where θ\theta^{*} is computed by (30) with 𝐲(𝐱)=𝐩k+2/4(i,j),𝐰(𝐱)=𝝀k+2/4(i,j)\mathbf{y}(\mathbf{x})=\mathbf{p}^{k+2/4}(i,j),\mathbf{w}(\mathbf{x})=\boldsymbol{\lambda}^{k+2/4}(i,j). Functions 𝐩k+3/4\mathbf{p}^{k+3/4} and 𝝀k+3/4\boldsymbol{\lambda}^{k+3/4} are then updated using (32).

4.4 Discrete analogue of 𝐩k+1,rk+1\mathbf{p}^{k+1},r^{k+1} and 𝐬k+1\mathbf{s}^{k+1}

Reorganizing the system (44), we get a linear system in vk+1,rk+1v^{k+1},r^{k+1} and 𝐬k+1=(s1k+1,s2k+1)\mathbf{s}^{k+1}=(s_{1}^{k+1},s_{2}^{k+1})

{(2+τI)vk+1+τrk+1+τ𝐬k+1=𝐩k+3/4+τf,τvk+1+[(γ2+τ)I+2ταw4]rk+1+τ𝐬k+1=γ2rk+3/4+τf,τvk+1τrk+1+[(γ3+2ταn)Iτdiv]𝐬k+1=γ3𝐬k+3/4τf.\displaystyle\begin{cases}(-\nabla^{2}+\tau I)v^{k+1}+\tau r^{k+1}+\tau\nabla\cdot\mathbf{s}^{k+1}=-\nabla\cdot\mathbf{p}^{k+3/4}+\tau f,\\ \tau v^{k+1}+[(\gamma_{2}+\tau)I+2\tau\alpha_{w}\nabla^{4}]r^{k+1}+\tau\nabla\cdot\mathbf{s}^{k+1}=\gamma_{2}r^{k+3/4}+\tau f,\\ -\tau\nabla v^{k+1}-\tau\nabla r^{k+1}+[(\gamma_{3}+2\tau\alpha_{n})I-\tau\nabla\text{div}]\mathbf{s}^{k+1}=\gamma_{3}\mathbf{s}^{k+3/4}-\tau\nabla f.\end{cases} (55)

Upon discretization, we obtain a linear system

𝐀[vk+1rk+1s1k+1s2k+1]=[b1kb2kb3kb4k]with{b1k=div𝐩k+3/4+τf,b2k=γ2rk+3/4+τf,b3k=γ3s1k+3/4τ1+f,b4k=γ3s2k+3/4τ2+f,\displaystyle\mathbf{A}\begin{bmatrix}v^{k+1}\\ r^{k+1}\\ s_{1}^{k+1}\\ s_{2}^{k+1}\end{bmatrix}=\begin{bmatrix}b^{k}_{1}\\ b^{k}_{2}\\ b^{k}_{3}\\ b^{k}_{4}\end{bmatrix}\leavevmode\nobreak\ \text{with}\leavevmode\nobreak\ \begin{cases}&b^{k}_{1}=-\text{div}^{-}\mathbf{p}^{k+3/4}+\tau f,\\ &b^{k}_{2}=\gamma_{2}r^{k+3/4}+\tau f,\\ &b^{k}_{3}=\gamma_{3}s_{1}^{k+3/4}-\tau\partial_{1}^{+}f,\\ &b^{k}_{4}=\gamma_{3}s_{2}^{k+3/4}-\tau\partial_{2}^{+}f,\end{cases}

where

𝐀=\displaystyle\mathbf{A}=
[[(𝒮1)(𝒮1+)/h2+(𝒮2)(𝒮2+)/h2]+τττ(𝒮1)/hτ(𝒮2)/hτ(γ2+τ)+2ταw𝒫τ(𝒮1)/hτ(𝒮2)/hτ(𝒮1+)/hτ(𝒮1+)/hγ3+2ταnτ(𝒮1+)(𝒮1)/h2τ(𝒮1+)(𝒮2)/h2τ(𝒮2+)/hτ(𝒮2+)/hτ(𝒮2+)(𝒮1)/h2γ3+2ταnτ(𝒮2+)(𝒮2)/h2]\displaystyle\begin{bmatrix}\begin{aligned} &-[(\mathcal{I}-\mathcal{S}_{1}^{-})(\mathcal{S}_{1}^{+}-\mathcal{I})/h^{2}\\ &+(\mathcal{I}-\mathcal{S}_{2}^{-})(\mathcal{S}_{2}^{+}-\mathcal{I})/h^{2}]\\ &+\tau\mathcal{I}\end{aligned}&\tau&\tau(\mathcal{I}-\mathcal{S}_{1}^{-})/h&\tau(\mathcal{I}-\mathcal{S}_{2}^{-})/h\\ \tau&(\gamma_{2}+\tau)\mathcal{I}+2\tau\alpha_{w}\mathcal{P}&\tau(\mathcal{I}-\mathcal{S}_{1}^{-})/h&\tau(\mathcal{I}-\mathcal{S}_{2}^{-})/h\\ -\tau(\mathcal{S}_{1}^{+}-\mathcal{I})/h&-\tau(\mathcal{S}_{1}^{+}-\mathcal{I})/h&\begin{aligned} &\gamma_{3}+2\tau\alpha_{n}\\ &-\tau(\mathcal{S}_{1}^{+}-\mathcal{I})(\mathcal{I}-\mathcal{S}_{1}^{-})/h^{2}\end{aligned}&-\tau(\mathcal{S}_{1}^{+}-\mathcal{I})(\mathcal{I}-\mathcal{S}_{2}^{-})/h^{2}\\ -\tau(\mathcal{S}_{2}^{+}-\mathcal{I})/h&-\tau(\mathcal{S}_{2}^{+}-\mathcal{I})/h&-\tau(\mathcal{S}_{2}^{+}-\mathcal{I})(\mathcal{I}-\mathcal{S}_{1}^{-})/h^{2}&\begin{aligned} &\gamma_{3}+2\tau\alpha_{n}\\ &-\tau(\mathcal{S}_{2}^{+}-\mathcal{I})(\mathcal{I}-\mathcal{S}_{2}^{-})/h^{2}\end{aligned}\end{bmatrix}

with

𝒫=\displaystyle\mathcal{P}= 1h4[((𝒮1)(𝒮1+))2+(𝒮2)(𝒮2+)(𝒮1)(𝒮1+)\displaystyle\frac{1}{h^{4}}\Big{[}((\mathcal{I}-\mathcal{S}_{1}^{-})(\mathcal{S}_{1}^{+}-\mathcal{I}))^{2}+(\mathcal{I}-\mathcal{S}_{2}^{-})(\mathcal{S}_{2}^{+}-\mathcal{I})(\mathcal{I}-\mathcal{S}_{1}^{-})(\mathcal{S}_{1}^{+}-\mathcal{I})
+(𝒮1)(𝒮1+)(𝒮2)(𝒮2+)+((𝒮2)(𝒮2+))2].\displaystyle\quad+(\mathcal{I}-\mathcal{S}_{1}^{-})(\mathcal{S}_{1}^{+}-\mathcal{I})(\mathcal{I}-\mathcal{S}_{2}^{-})(\mathcal{S}_{2}^{+}-\mathcal{I})+((\mathcal{I}-\mathcal{S}_{2}^{-})(\mathcal{S}_{2}^{+}-\mathcal{I}))^{2}\Big{]}.

Note that 𝒜\mathcal{A} does not depend on the iteration, thus it can be constructed beforehand and applied repeatedly in each iteration. Applying the discrete Fourier transform \mathcal{F}, we get

𝐃([vk+1rk+1s1k+1s2k+1])=([b1kb2kb3kb4k])\displaystyle\mathbf{D}\ \mathcal{F}\left(\begin{bmatrix}v^{k+1}\\ r^{k+1}\\ s_{1}^{k+1}\\ s_{2}^{k+1}\end{bmatrix}\right)=\mathcal{F}\left(\begin{bmatrix}b^{k}_{1}\\ b^{k}_{2}\\ b^{k}_{3}\\ b^{k}_{4}\end{bmatrix}\right) (56)

with 𝐃4×4\mathbf{D}\in\mathbb{R}^{4\times 4} whose (l,m)(l,m)-th entries dlmd_{lm}, 1l,m41\leq l,m\leq 4 are listed below

d11(i,j)=(τ2(cosζi1)/h22(cosηj1)/h2),d12=τ,\displaystyle d_{11}(i,j)=(\tau-2(\cos\zeta_{i}-1)/h^{2}-2(\cos\eta_{j}-1)/h^{2}),\quad d_{12}=\tau,
d13(i,j)=τ[1(cosζi1sinζi)]/h,d14(i,j)=τ[1(cosηj1sinηj)]/h,\displaystyle d_{13}(i,j)=\tau[1-(\cos\zeta_{i}-\sqrt{-1}\sin\zeta_{i})]/h,\quad d_{14}(i,j)=\tau[1-(\cos\eta_{j}-\sqrt{-1}\sin\eta_{j})]/h,
d21=τ,d22(i,j)=γ2+τ+2ταw4h4[(cosζi1)2+2(cosζi1)(cosηj1)+(cosηj1)2],\displaystyle d_{21}=\tau,\quad d_{22}(i,j)=\gamma_{2}+\tau+2\tau\alpha_{w}\frac{4}{h^{4}}\left[(\cos\zeta_{i}-1)^{2}+2(\cos\zeta_{i}-1)(\cos\eta_{j}-1)+(\cos\eta_{j}-1)^{2}\right],
d23=d13,d24=d14,\displaystyle d_{23}=d_{13},\quad d_{24}=d_{14},
d31(i,j)=d32(i,j)=τ(cosζi+1sinζi1)/h,\displaystyle d_{31}(i,j)=d_{32}(i,j)=-\tau(\cos\zeta_{i}+\sqrt{-1}\sin\zeta_{i}-1)/h,
d33(i,j)=γ3+2ταn2τ(cosζi1)/h2,\displaystyle d_{33}(i,j)=\gamma_{3}+2\tau\alpha_{n}-2\tau(\cos\zeta_{i}-1)/h^{2},
d34(i,j)=τ(cosζi+1sinζi1)(cosηj1sinηj1)/h2,\displaystyle d_{34}(i,j)=\tau(\cos\zeta_{i}+\sqrt{-1}\sin\zeta_{i}-1)(\cos\eta_{j}-\sqrt{-1}\sin\eta_{j}-1)/h^{2},
d41(i,j)=d42(i,j)=τ(cosηj+1sinηj1)/h,\displaystyle d_{41}(i,j)=d_{42}(i,j)=-\tau(\cos\eta_{j}+\sqrt{-1}\sin\eta_{j}-1)/h,
d43(i,j)=τ(cosηj+1sinηj1)(cosζi1sinζi1)/h2,\displaystyle d_{43}(i,j)=\tau(\cos\eta_{j}+\sqrt{-1}\sin\eta_{j}-1)(\cos\zeta_{i}-\sqrt{-1}\sin\zeta_{i}-1)/h^{2},
d44(i,j)=γ3+2ταn2τ(cosηj1)/h2.\displaystyle d_{44}(i,j)=\gamma_{3}+2\tau\alpha_{n}-2\tau(\cos\eta_{j}-1)/h^{2}.

To stabilize the computation, we add a small constant κ=1×109\kappa=1\times 10^{-9} on the diagonal of 𝐃\mathbf{D}. The inverse 𝐃1\mathbf{D}^{-1} can be explicitly computed using the Cramer’s rule. We can efficiently get vk+1,rk+1v^{k+1},r^{k+1} and 𝐬k+1\mathbf{s}^{k+1} by inverse FFT

[vk+1rk+1s1k+1s2k+1]=Real(1[𝐃1([b1kb2kb3kb4k])]).\displaystyle\begin{bmatrix}v^{k+1}\\ r^{k+1}\\ s_{1}^{k+1}\\ s_{2}^{k+1}\end{bmatrix}=\text{Real}\left(\mathcal{F}^{-1}\left[\mathbf{D}^{-1}\mathcal{F}\left(\begin{bmatrix}b^{k}_{1}\\ b^{k}_{2}\\ b^{k}_{3}\\ b^{k}_{4}\end{bmatrix}\right)\right]\right)\;. (57)

Similarly to (53), as we obtain the updated variables by explicit formula, the computational cost is still 𝒪(MN(logM+logN))\mathcal{O}(MN(\log M+\log N)). Overall, the computational cost of a single iteration of our proposed algorithm that updates (𝐩k,𝝀k,rk,𝐬k)(\mathbf{p}^{k},\boldsymbol{\lambda}^{k},r^{k},\mathbf{s}^{k}) to (𝐩k+1,𝝀k+1,rk+1,𝐬k+1)(\mathbf{p}^{k+1},\boldsymbol{\lambda}^{k+1},r^{k+1},\mathbf{s}^{k+1}) is in the order of 𝒪(MN(logM+logN))\mathcal{O}(MN(\log M+\log N)).

5 Numerical Experiments

In this section, we present various experiments to show the behaviors of the proposed model, analyze the effects of model parameters, and justify its effectiveness in the task of decomposition and denoising by ablation and comparison studies. To evaluate the image quality, we employ the PSNR metric

PSNR(uref,u^)=10log10(1u^uref2),\displaystyle\text{PSNR}(u_{\text{ref}},\widehat{u})=10\log_{10}\left(\frac{1}{\|\widehat{u}-u_{\text{ref}}\|_{2}}\right)\;,

where urefu_{\text{ref}} is the reference image, u^\widehat{u} is the corresponding approximation, and their intensity values are scaled to [0,1][0,1]. We test our method on various synthetic images as well as grayscale photos. For model parameters α0,αcurv,αw\alpha_{0},\alpha_{\text{curv}},\alpha_{w} and αn\alpha_{n}, we discuss their selections in Section 5.5 and fine-tune them according to the image contents. For algorithmic parameters, we fix τ=0.1,γ1=0.01,γ2=20\tau=0.1,\gamma_{1}=0.01,\gamma_{2}=20, and γ3=1\gamma_{3}=1. We set the global constants for numerical stability c=1×109c=1\times 10^{-9} and κ=1×109\kappa=1\times 10^{-9}; here cc is the frozen coefficient in (48), and κ\kappa is added to the diagonal of 𝐃\mathbf{D} in (56) for stability. We take the maximal number of iterations Itermax=1000\text{Iter}_{\max}=1000, and fix the threshold for the terminating condition ρ=1×106\rho=1\times 10^{-6}, i.e., the algorithm is terminated if max{rk+1rk2rk2,vk+1vn2vk2}<ρ\max\left\{\frac{\|r^{k+1}-r^{k}\|_{2}}{\|r^{k}\|_{2}},\frac{\|v^{k+1}-v^{n}\|_{2}}{\|v^{k}\|_{2}}\right\}<\rho. For the default initial conditions, we fix v0=0.001×f~+0.999×0.5v_{0}=0.001\times\widetilde{f}+0.999\times 0.5 where f~\widetilde{f} is the input ff convolved with a standard Gaussian kernel, 𝐩0=+v0\mathbf{p}_{0}=\nabla^{+}v_{0}, r0=fv0r_{0}=f-v_{0}, and 𝝀0=𝐬0=𝟎\boldsymbol{\lambda}_{0}=\mathbf{s}_{0}=\mathbf{0}. To reduce the boundary artifacts, we pad the input with 30 pixels on each side in a symmetric manner. The algorithm is implemented with MATLAB 2022a in the environment of Windows 11, and all the experiments are conducted in a laptop equipped with 12th Gen Intel(R) Core(TM) i7-12800H, 2400Mhz, and 14 cores.

5.1 General performances

fcleanf_{\text{clean}} fnoisef_{\text{noise}} uu^{*} vscaledv^{*}_{\text{scaled}} wscaledw^{*}_{\text{scaled}} nscaledn^{*}_{\text{scaled}}
Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
PSNR=22.10=22.10 PSNR=43.60=43.60 STD=19.69/255=19.69/255
Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
PSNR=22.10=22.10 PSNR=35.87=35.87 STD=19.97/255=19.97/255
Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
PSNR=22.12=22.12 PSNR=36.44=36.44 STD=19.82/255=19.82/255
Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
PSNR=22.12=22.12 PSNR=25.69=25.69 STD=23.19/255=23.19/255
Figure 3: Decomposition of synthetic images with noise (σ=20/255\sigma=20/255). The first column (fcleanf_{\text{clean}}) shows the clean images; the second column (fnoisef_{\text{noise}}) shows the noisy inputs; the third column (uu^{*}) shows summations of the identified vv^{*} and ww^{*} components; the subsequent columns show respectively vv^{*}, ww^{*} and nn^{*}, and for visualization, they are linearly scaled between 0 and 1. Model parameters for these examples are: α0=1×102\alpha_{0}=1\times 10^{-2}, αcurv=0.1\alpha_{\text{curv}}=0.1, αw=80\alpha_{w}=80, and αn=1×105\alpha_{n}=1\times 10^{-5}. PSNR values for fnoisef_{\text{noise}} and uu^{*} are computed, and the standard deviations (STD) of respective nn^{*} are reported.
fcleanf_{\text{clean}} fnoisef_{\text{noise}} uu^{*} vscaledv^{*}_{\text{scaled}} wscaledw^{*}_{\text{scaled}} nscaledn^{*}_{\text{scaled}}
Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
PSNR=22.11=22.11 PSNR=28.50=28.50 STD=19.87/255=19.87/255
Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
PSNR=22.11=22.11 PSNR=30.57=30.57 STD=19.24/255=19.24/255
Figure 4: Decomposition of photographic images with noise (σ=20/255\sigma=20/255). The first column (fcleanf_{\text{clean}}) shows the clean images; the second column (fnoisef_{\text{noise}}) shows the noisy inputs; the third column (uu^{*}) shows summations of the identified vv^{*} and ww^{*} components; the subsequent columns show respectively vv^{*}, ww^{*} and nn^{*}, and for visualization, they are linearly scaled between 0 and 1. Model parameters for these examples are: α0=1×103\alpha_{0}=1\times 10^{-3}, αcurv=0.5\alpha_{\text{curv}}=0.5, αw=50\alpha_{w}=50, and αn=0.1\alpha_{n}=0.1. PSNR values for fnoisef_{\text{noise}} and uu^{*} are computed, and the standard deviations (STD) of respective nn^{*} are reported.

We illustrate the general performances of the proposed model using two sets of experiments. In Figure 3, we show the decomposition results for synthetic images generated by convex combinations of piecewise constant components and smoothly varying shades. In Figure 4, we show the results for grayscale photos with richer details. For both sets of examples, Gaussian noise with intensity σ=20/255\sigma=20/255 is added.

In the first row of Figure 3, a white cross in the dark is shed by a soft light at the center. The proposed model successfully extracts the cross vv^{*} where the intensity variation due to the lighting is removed, and the important geometric features such as sharp corners as well as straight boundaries are accurately reconstructed. The lighting is captured by the ww^{*} component, and it is well separated from vv^{*}. The nn^{*} component corresponds to the noise whose standard deviation is close to the true noise level. In the second row, we test the proposed model on a combination of silhouettes of ellipses and a circular wave of lighting. Under the non-uniform lighting, the ellipses are correctly identified in vv^{*} and filled with nearly constant intensity. The ww^{*} component shows the correct wave pattern of the lighting. In the third row, we have a shaded globe, where the source of the light is fixed at the top left corner. The proposed method correctly assigns in the vv^{*} component the silhouettes of the globe where boundaries of continents are sharp. Noticeably, the proposed model keeps the soft shadow characterizing the volumetric feature of a ball in the ww^{*} component with the correct direction of lighting. In the last row, we have a challenging case where a binarized portrait of Einstein is blended with multiple blurry shades of the mass-energy equation. The ww^{*} component is more complicated than those in the previous examples. Our method successfully extracts the soft contents with complex shapes and leaves a relatively clean configuration for the vv^{*} component.

In Figure 4, we apply the proposed model to photo-realistic grayscale images. In all examples, we observe that contents with sharp boundaries are captured by vv^{*}; soft shadows and illumination variation are modeled by ww^{*}; and the nn^{*} keeps fine textures plus noise. Noticeably, in the first row, the rectangular reflectances on the dome are separated from the soft light and shadow defining the volume of the dome. In the second row, the foreground with crisp boundaries is distinguished from the background with blurry shapes.

The examples in Figure 3 and 4 justify the effectiveness of the proposed method. Thanks to the curvature regularization, the extracted vv^{*} component preserves the defining geometric features of the homogeneous part in the given input image, such as sharp boundaries and corners. The extracted ww^{*} component consistently matches the soft intensity variation caused by lighting and shadows. The extracted nn^{*} component holds the oscillatory residuals including noise and fine-scale textures. We note that when the input images do not contain rich textures, the nn^{*} components rendered by the proposed model do not contain structural details such as boundaries. It allows us to use the summation of vv^{*} and ww^{*}, which is denoted by uu^{*}, to reconstruct the clean image. This is justified by the PSNR values reported in Figure 3 and 4.

5.2 Ablation study

(a) (b) (c) (d) (e) (f)
Input Model I Model II Model III Proposed True
Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
PSNR(u)(u^{*}) 32.3132.31 34.6434.64 35.0335.03 35.66\mathbf{35.66}
Figure 5: Ablation study. (a) Noisy input (size: 256×256256\times 256, σ=50/255\sigma=50/255), and regions in red boxes are zoomed-in in the subsequent rows. (b) Noise-free component uu^{*} by Model I (58), α0=6×102\alpha_{0}=6\times 10^{-2}. (c) Noise-free component uu^{*} by Model II (59), α0=2×102\alpha_{0}=2\times 10^{-2}, αw=80\alpha_{w}=80. (d) Noise-free component uu^{*} by Model III (60), α0=4×102,αcurv=2\alpha_{0}=4\times 10^{-2},\alpha_{\text{curv}}=2. (e) Noise-free component uu^{*} by the proposed model, α0=2.5×102,αw=80,αcurv=0.5\alpha_{0}=2.5\times 10^{-2},\alpha_{w}=80,\alpha_{\text{curv}}=0.5. (f) True image. For all scenarios, we fixed αn=1×104\alpha_{n}=1\times 10^{-4}. The noise-free components are obtained by summing the identified components except for the respective oscillatory parts.

To justify the combination of regularizations in our model (7), we compare it with the following settings:

  1. 1.

    Model I. An image ff is decomposed into two parts: cartoon part uu^{*} and oscillatory part nn^{*}

    {u,n}=argminu,n[α0u0+αnnH1(Ω)2+12Ω|f(u+n)|2𝑑𝐱].\displaystyle\{u^{*},n^{*}\}=\operatorname*{arg\,min}_{u,n}\left[\alpha_{0}\|\nabla u\|_{0}+\alpha_{n}\|n\|^{2}_{H^{-1}(\Omega)}+\frac{1}{2}\int_{\Omega}\left|f-(u+n)\right|^{2}\,d\mathbf{x}\right]\;. (58)
  2. 2.

    Model II. An image ff is decomposed into three parts: cartoon part vv^{*}, smooth part ww^{*}, and oscillatory part nn^{*}

    {v,w,n}=argminv,w,n[α0v0+αwΔwL2(Ω)2+αnnH1(Ω)2+12Ω|f(v+w+n)|2𝑑𝐱].\displaystyle\{v^{*},w^{*},n^{*}\}=\operatorname*{arg\,min}_{v,w,n}\left[\alpha_{0}\|\nabla v\|_{0}+\alpha_{w}\|\Delta w\|_{L^{2}(\Omega)}^{2}+\alpha_{n}\|n\|^{2}_{H^{-1}(\Omega)}+\frac{1}{2}\int_{\Omega}\left|f-(v+w+n)\right|^{2}\,d\mathbf{x}\right]\;. (59)
  3. 3.

    Model III. An image ff is decomposed into two parts: homogeneous structure uu^{*} and oscillatory part nn^{*}

    {u,n}=argminu,n[α0u0+αcurvΩ(u|u|)2|u|𝑑𝐱+αnnH1(Ω)2+12Ω|f(u+n)|2𝑑𝐱].\displaystyle\{u^{*},n^{*}\}=\operatorname*{arg\,min}_{u,n}\left[\alpha_{0}\|\nabla u\|_{0}+\alpha_{\text{curv}}\int_{\Omega}\left(\nabla\cdot\frac{\nabla u}{|\nabla u|}\right)^{2}\left|\nabla u\right|\,d\mathbf{x}+\alpha_{n}\|n\|^{2}_{H^{-1}(\Omega)}+\frac{1}{2}\int_{\Omega}\left|f-(u+n)\right|^{2}\,d\mathbf{x}\right]\;. (60)

Compared to the proposed model, there is no smoothly varying component in Model I, and the regularity of uu^{*} is mainly affected by the L0L^{0}-gradient. In Model II, the noise-less component uu^{*} is decomposed into a cartoon part vv^{*} and a smooth part ww^{*}, the same as in the proposed model; but there is no curvature regularization on vv^{*}. In Model III, we impose both L0L^{0}-gradient and curvature regularization on the noise-free approximation uu^{*}, but there is no component that models the smooth shading.

Figure 5 shows the uu^{*} components learned from the various settings above. In column (a), we present a noisy grayscale image of a squared ring, and the regions inside red boxes are zoomed-in and shown in the subsequent rows. We added Gaussian noise with intensity σ=50/255\sigma=50/255 on the input. In column (b), we show the uu^{*} component reconstructed by Model I (58). As was expected, using the regularization u0\|\nabla u\|_{0} enforces strong sparsity on the gradient of reconstructed uu^{*} component and yields sharp boundaries. However, analogous to the TV-norm regularization, using only L0L^{0}-gradient as regularization creates strong staircase effects. Moreover, L0L^{0}-gradient is local, yet larger-scale geometric information is necessary to render better results. The boundary in the third row shows undesirable irregularities. The smooth region is better reconstructed in Model II, where an extra component ww^{*} is introduced. However, due to the lack of geometric information in L0L^{0}-gradient, the boundary lines still present zig-zags. In Model III, the curvature regularization on uu^{*} is used, which imposes geometric priors on the level lines of vv^{*}. The straight boundary lines are better preserved compared to Model I and Model II. Without the smooth part ww^{*}, the intensity transition in the last row is not as smooth as in Model II. Combining the benefits of geometric regularization on the level lines and adding an extra smoothly varying component, our proposed model outperforms the others. Observe that the transition is smooth in the second row indicated by the red contour curves, and the boundary line is straight in the third row. These reconstructed features match the underlying image shown in the last column.

From this set of experiments, we conclude that L0L^{0}-gradient regularization induces strong sparsity on the gradient which causes significant staircase artifacts and irregular boundaries. These undesirable phenomenons are effectively mitigated by the curvature regularization on the level lines. In order to recover smoothly varying regions while preserving sharp boundaries, our experiments show that adding an extra smooth part in the decomposition is important.

5.3 Effect of parameter αw\alpha_{w} and αcurv\alpha_{\text{curv}}

In our proposed model, both vv^{*} and ww^{*} favor regularity. Particularly, vv^{*} prefers shapes with definite and regular boundaries; whereas ww^{*} tends to be blurry. We demonstrate that image content can be distributed and converted between these two components, yielding different decomposition results.

First, we study the effect of the model parameter αw\alpha_{w} associated with the term Ω2w22𝑑𝐱\int_{\Omega}\|\nabla^{2}w\|_{2}^{2}\,d\mathbf{x} that encourages isotropic smoothness of the ww^{*} component. Intuitively, larger values of αw\alpha_{w} induce smoother ww^{*} where image details are smeared. If αw\alpha_{w} is too large, regions with smoothly varying intensity and sufficiently regular level lines can be captured by the vv^{*} component.

We illustrate this conversion between ww^{*} and vv^{*} affected by αw\alpha_{w} in Figure 6. Given a noisy image (σ=10/255\sigma=10/255), we apply the proposed model with αw\alpha_{w} increasing from 1010 to 10001000 while fixing the other parameters. For each value of αw\alpha_{w}, we show the scaled vv^{*} and ww^{*} components in the first and second row, respectively, and in the third row, we show the respective sum u=v+wu^{*}=v^{*}+w^{*}. Observe that with a wide range of values of αw\alpha_{w}, the noise-free reconstruction uu^{*} remains visually similar, and the PSNR values are stable. When αw\alpha_{w} is small, most of the intensity variation is captured by ww^{*}, while vv^{*} only captures objects’ large-scale contours. Although hair, face, eyes, and many recognizable shapes are visible in ww^{*}, these distinct parts are blended with each other smoothly without sharp boundaries. As αw\alpha_{w} increases, we see that ww^{*} becomes more blurry, and vv^{*} shows richer intensity variations such as those in the background. Note that these smoothly varying components in vv^{*} have definite boundaries. For example, the zoom-ins of the red boxed region in the fourth row show that intensity transitions become sharper when αw\alpha_{w} increases. Other softer smooth parts without distinctive boundaries, such as the shades on the cheeks and chin, are captured by the ww^{*} component.

αw\alpha_{w} 10 50 100 500 1000
vscaledv_{\text{scaled}}^{*} Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
wscaledw_{\text{scaled}}^{*} Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
uu^{*} Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
Zoom-in uu^{*} Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
PSNR(uu^{*}) 34.5834.58 34.59 34.5334.53 34.3434.34 34.2834.28
Figure 6: Effect of αw\alpha_{w}. When αw\alpha_{w} increases, smooth intensity variations in ww^{*} (second row) become structural shapes in vv^{*} (first row): vv^{*} contains richer details, e.g., the shades in the background, and ww^{*} is more blurry. Overall, intensity transitions in uu^{*} (third row) become sharper, which is further illustrated by the zoom-ins of the red box (fourth row). The input image Zelda is added with Gaussian noise (σ=10/255\sigma=10/255). Other parameters are fixed: α0=1×103\alpha_{0}=1\times 10^{-3}, αcurv=0.5\alpha_{\text{curv}}=0.5, αn=10\alpha_{n}=10.

Second, we show the effect of the parameter αcurv\alpha_{\text{curv}} which is closely related to the regularity of the level lines in the vv^{*} component. Similar to the previous study on the effect of αw\alpha_{w}, by changing αcurv\alpha_{\text{curv}}, smoothly varying components can transfer between ww^{*} and vv^{*}. In particular, if αcurv\alpha_{\text{curv}} is large, certain regions with smooth intensity variation can be pushed into ww^{*} for lower total loss.

Figure 7 presents results of the proposed model on a noisy image (σ=10/255)(\sigma=10/255) with varying values of αcurv\alpha_{\text{curv}} when the other regularization parameters are fixed. When αcurv\alpha_{\text{curv}} increases from 0.010.01 to 1010, we observe that soft shades on peppers in vv^{*} are suppressed while the boundaries are preserved. Meanwhile, the ww^{*} component gains more variation in the intensity. If we keep increasing αcurv\alpha_{\text{curv}} above 0.10.1, the intensity variation near the boundaries and the highlights due to reflection slowly fade away in vv^{*}. In contrast, shapes of peppers are more apparent in ww^{*}. Since the sharp gradients in vv^{*} are reduced and the intensity variation in ww^{*} is intrinsically smooth, the reconstructed uu^{*} component becomes slightly blurry. This transition is also reflected by the decreasing PSNR. In Figure 7, we also show the level lines of uu^{*} in a zoomed-in region. The effect of αcurv\alpha_{\text{curv}} on uu^{*} is obvious. Increasing αcurv\alpha_{\text{curv}} yields more regular level lines, and the staircase effect due to L0L^{0}-gradient regularization is thus ameliorated.

αcurv\alpha_{\text{curv}} 0.01 0.1 1 5 10
vscaledv_{\text{scaled}}^{*} Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
wscaledw_{\text{scaled}}^{*} Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
uu^{*} Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
Zoom-in uu^{*} Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
PSNR(uu^{*}) 33.0633.06 32.7732.77 32.0532.05 31.1931.19 30.6730.67
Figure 7: Effect of αcurv\alpha_{\text{curv}} (σ=10/255\sigma=10/255). When αcurv\alpha_{\text{curv}} increases, vv^{*} (first row) becomes smoother while objects’ boundaries are preserved. The ww^{*} component (second row) gains more geometric features such as shapes. Consequently, the sum uu^{*} (third row) is smoother. Thanks to the curvature regularization, stair-casing in the zoomed-in contours (last row) is effectively removed when increasing αcurv\alpha_{\text{curv}}. Other parameters are fixed: α0=1×103\alpha_{0}=1\times 10^{-3}, αw=50\alpha_{w}=50, and αn=10\alpha_{n}=10.

5.4 Comparison on cartoon+smooth+oscillation decomposition

CEP [13] (CPU time: 76.16 sec)
fnoisef_{\text{noise}} uu^{*} vscaledv^{*}_{\text{scaled}} wscaledw^{*}_{\text{scaled}} nscaledn^{*}_{\text{scaled}}
Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
PSNR=29.04=29.04
HKLM [29] (CPU time: 199.48 sec)
uu^{*} vscaledv^{*}_{\text{scaled}} wscaledw^{*}_{\text{scaled}} nscaledn^{*}_{\text{scaled}}
Refer to caption Refer to caption Refer to caption Refer to caption
PSNR=25.08=25.08
Proposed (CPU time: 14.83 sec)
uu^{*} vscaledv^{*}_{\text{scaled}} wscaledw^{*}_{\text{scaled}} nscaledn^{*}_{\text{scaled}}
Refer to caption Refer to caption Refer to caption Refer to caption
PSNR=32.06=\mathbf{32.06} STD=49.08=49.08
Figure 8: Comparison of CEP [13], HKLM [29] and the proposed model for decomposing a noisy synthetic image (size: 128×128128\times 128, noise: σ=50/255\sigma=50/255). For CEP (see also (5)), we use λ=0.5,α=10,μ=2\lambda=0.5,\alpha=10,\mu=2 and δ\delta determined by SURE [19]. For HKLM (see also (6)), we use β1=1×104\beta_{1}=1\times 10^{-4}, β2=0.5\beta_{2}=0.5, β3=5×106\beta_{3}=5\times 10^{-6}, and a=30a=30 for the MC penalty (an approximation of the L0L^{0}-norm, see [29] for details). For the proposed model, we use α0=2×102\alpha_{0}=2\times 10^{-2}, αcurv=0.15\alpha_{\text{curv}}=0.15, αw=50\alpha_{w}=50 and αn=1×103\alpha_{n}=1\times 10^{-3}.
CEP [13] (CPU time: 1655.30 sec)
fnoisef_{\text{noise}} uu^{*} vscaledv^{*}_{\text{scaled}} wscaledw^{*}_{\text{scaled}} nscaledn^{*}_{\text{scaled}}
Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
PSNR=29.11=29.11
HKLM [29] (CPU time: 4201.45 sec)
uu^{*} vscaledv^{*}_{\text{scaled}} wscaledw^{*}_{\text{scaled}} nscaledn^{*}_{\text{scaled}}
Refer to caption Refer to caption Refer to caption Refer to caption
PSNR=27.93=27.93
Proposed (CPU time: 179.20 sec)
uu^{*} vscaledv^{*}_{\text{scaled}} wscaledw^{*}_{\text{scaled}} nscaledn^{*}_{\text{scaled}}
Refer to caption Refer to caption Refer to caption Refer to caption
PSNR=31.46=\mathbf{31.46}
Figure 9: Comparison on cartoon+smooth+oscillation decomposition of CEP [13], HKLM [29] and the proposed model. fnoisef_{\text{noise}} is the input butterfly (size: 512×512512\times 512, noise: σ=10/255\sigma=10/255). For CEP (see also (5)), we set λ=5,α=10,μ=2\lambda=5,\alpha=10,\mu=2 and δ\delta determined from SURE [19]. For [29] (see also (6)), we use β1=1×104\beta_{1}=1\times 10^{-4}, β2=3\beta_{2}=3, β3=5×106\beta_{3}=5\times 10^{-6}, and a=30a=30 for the MC penalty. For the proposed model, we use α0=1×103\alpha_{0}=1\times 10^{-3}, αcurv=0.05\alpha_{\text{curv}}=0.05, αw=30\alpha_{w}=30 and αn=10\alpha_{n}=10.
(a) (b) (c) (d) (e)
Refer to caption Refer to caption Refer to caption Refer to caption Refer to caption
Figure 10: Zoom-ins of the boxed region in Figure 9. (a) Clean image. (b) Noisy input. (c) CEP [13]. (d) HKLM [29]. (e) Proposed.

We compare the proposed model with models proposed by Chan et al. [13] (CEP) and Huska et al. [29] (HKLM), which are stated in (5) and (6), respectively. In both the proposed model and HKLM, an input image is decomposed into three components corresponding to cartoon-like part vv^{*}, smooth part ww^{*}, and oscillatory part nn^{*}. For the four-component model (5), in addition to vv^{*} and ww^{*}, we take n=n1+n2n^{*}=n_{1}^{*}+n_{2}^{*}, i.e., the oscillatory part is the summation of the texture and noise.

In Figure 8, we show the decomposition results from CEP, HKLM, and DOSH when applied to a noisy synthetic image (σ=50/255\sigma=50/255), respectively. For CEP [13] (see also (5)), we set λ=0.5,α=10\lambda=0.5,\alpha=10, μ=2\mu=2 and the wavelet threshold δ\delta is determined using Stein’s Unbiased Risk Estimate (SURE) [19]. For HKLM [29] (see also (6)), the authors proposed an automatic scheme for tuning the regularization parameters. However, we find that they fail to yield good decomposition results in general for our test images. To reach good PSNR values in uu^{*}, after experimenting with different ranges of parameters, we manually set β1=1×104\beta_{1}=1\times 10^{-4}, β2=0.5\beta_{2}=0.5, β3=5×106\beta_{3}=5\times 10^{-6}, and the MC penalty (an approximation of the L0L^{0}-norm, see [29] for details) parameter a=30a=30. For the proposed model, we use α0=2×102\alpha_{0}=2\times 10^{-2}, αcurv=0.15\alpha_{\text{curv}}=0.15, αw=50\alpha_{w}=50 and αn=1×103\alpha_{n}=1\times 10^{-3}. We also present the summations of vv^{*} and ww^{*} as the noise-less approximations for fnoisef_{\text{noise}} and evaluate their respective PSNR values. In this example with heavy noise, both HKLM and our proposed model successfully separate the piecewise constant part from fnoisef_{\text{noise}} with strong sparsity of gradient thanks to the L0L^{0}-gradient regularization, and CEP with TV-norm fails to produce clean vv^{*} part. Moreover, we observe that, with a heavy noise, regularization using only L0L^{0}-gradient as in HKLM yields irregular boundaries. In contrast, with the additional regularization on the curvature of level-set curves of vv^{*}, our proposed model renders more satisfying results.

In Figure 9, we compare these methods using a grayscale photo of butterfly with noise (σ=10/255\sigma=10/255) and show the zoom-in comparisons in Figure 10. For CEP (see also (5)), we set λ=5,α=10,μ=2\lambda=5,\alpha=10,\mu=2 and δ\delta determined from SURE [19]. For HKLM (see also (6)), we use β1=1×104\beta_{1}=1\times 10^{-4}, β2=3\beta_{2}=3, β3=5×106\beta_{3}=5\times 10^{-6}, and a=30a=30 for the MC penalty. For the proposed method, we use α0=1×103\alpha_{0}=1\times 10^{-3}, αcurv=0.05\alpha_{\text{curv}}=0.05, αw=30\alpha_{w}=30 and αn=10\alpha_{n}=10. All methods separate the structural part including patterns on the wing in the front, the smooth part including the blurry leaves in the back, as well as the oscillatory component such as noise. Compared to CEP, both HKML and our proposed model produce better separations between the foreground details and background variations. Since HKLM uses only L0L^{0}-gradient for regularizing the vv^{*} component, we observe that uu^{*} of HKML has the sharpest boundaries compared to CEP and the proposed model in Figure 10. This also yields irregular curves and strong staircase effects in (d), which are effectively suppressed in CEP in (c) and the proposed model in (e). Since CEP uses TV-norm and the proposed model uses the stronger sparsity-inducing L0L^{0}-gradient, the vv^{*} component and consequently the uu^{*} component resulted from the proposed model are clearer than those from CEP. Moreover, for this example, the noise-less approximation uu^{*} produced by the proposed model has the highest PSNR value compared to the other two. We conclude that our proposed model produces more visually appealing decomposition results with better quality.

Our proposed model is also more efficient than CEP and HKLM. For the image of size 512×512512\times 512 in Figure 9, CEP takes around 1655 seconds, HKLM takes over 4000 seconds, whereas our model only requires around 180 seconds. We note that HKLM requires to solve a dense linear system of size (4NM)×(4NM)(4NM)\times(4NM), which costs around 𝒪((NM)3)\mathcal{O}((NM)^{3}). As discussed in Section 4, by leveraging FFT, the heaviest computation in our model requires solving a linear system of size 4×44\times 4 for each pixel in parallel, and the solution can be explicitly expressed. Overall, the computational cost of each iteration of our method is in the order or 𝒪(MN(logM+logN))\mathcal{O}(MN(\log M+\log N)). This is similar for CEP, which involves both FFT and the discrete wavelet transform.

5.5 Selection of model parameters

(a) (b) (c)
Refer to caption Refer to caption Refer to caption
Figure 11: Images used for studying parameters in subsection 5.5. (a) house, (b) pepper, (c) blonde.
(a) (b) (c)
Refer to caption Refer to caption Refer to caption
Figure 12: PSNR contour plots of uu^{*} components generated by the proposed model for different combinations of αcurv\alpha_{\text{curv}} and αw\alpha_{w}. The inputs are (a) house, (b) pepper, and (c) blonde, which have increasing levels of details and textures. Images have Gaussian noise (σ=10/255\sigma=10/255). Other parameters are fixed: α0=1×103\alpha_{0}=1\times 10^{-3}, αn=10\alpha_{n}=10. The parameter αw\alpha_{w} is relatively stable when αw1\alpha_{w}\geq 1. Choice of αcurv\alpha_{\text{curv}} can depend on image contents. Higher values of αcurv\alpha_{\text{curv}} are suitable for images with simple shapes and lower values for those with rich details.
(a) (b) (c)
Refer to caption Refer to caption Refer to caption
Figure 13: PSNR contour plots of uu^{*} components generated by the proposed model for different values of αn\alpha_{n} under various levels of noise σ\sigma. The inputs are (a) house, (b) pepper, and (c) blonde, which have increasing levels of details and textures. Other parameters are fixed: α0=1×103\alpha_{0}=1\times 10^{-3}, αcurv=0.1\alpha_{\text{curv}}=0.1, αw=50\alpha_{w}=50. For low noise regime (σ<20)(\sigma<20), αn1\alpha_{n}\approx 1 is appropriate; for middle levels of noise (20σ<60)(20\leq\sigma<60), we suggest using αn104\alpha_{n}\approx 10^{-4}; and for high noise regime, αn1×106\alpha_{n}\approx 1\times 10^{-6} is more suitable.
(a) (b) (c)
Refer to caption Refer to caption Refer to caption
Figure 14: PSNR values of uu^{*} components for noisy inputs with varying α0\alpha_{0}. (a) σ=10\sigma=10, (b) σ=40\sigma=40, (c) σ=80\sigma=80. In these examples, αcurv=0.1\alpha_{\text{curv}}=0.1, αw=50\alpha_{w}=50. The parameter αn\alpha_{n} is fixed at 1010, 1×1021\times 10^{-2} and 1×1041\times 10^{-4}, respectively.

There are four model parameters in the proposed model: α0,αcurv,αw\alpha_{0},\alpha_{\text{curv}},\alpha_{w} and αn\alpha_{n}, and it is generally difficult to determine their optimal values a priori. We demonstrate proper ranges for these model parameters that render good quality for uu^{*} in terms of PSNR metric.

We take house, pepper, and blonde in Figure 11 with increasing levels of details. In the first set of experiments, we add Gaussian noise (σ=10/255\sigma=10/255) and apply the proposed model to them with different combinations of αcurv\alpha_{\text{curv}} and αw\alpha_{w} while fixing α0=1×103\alpha_{0}=1\times 10^{-3} and αn=10\alpha_{n}=10 from heuristics. Figure 12 shows the PSNR contours of the uu^{*} components learned from the test images. We observe that, for αw<1\alpha_{w}<1, the PSNR values are relatively low due to lack of smoothness, and they increase as αw\alpha_{w} grows. When 1αw1031\leq\alpha_{w}\leq 10^{3}, PSNR values become relatively insensitive to the choice of αw\alpha_{w}. As for αcurv\alpha_{\text{curv}}, we observe that the optimal values depend on image contents. Specifically, if the images are relatively simple, such as house, αcurv\alpha_{\text{curv}} around 0.110.1\sim 1 yields slightly better quality. As the images contain richer details such as those in pepper and blonde, the maximal PSNR values are obtained when αcurv\alpha_{\text{curv}} takes smaller values around 10310210^{-3}\sim 10^{-2}. We suggest fixing αcurv\alpha_{\text{curv}} in the order of 0.10.1 and αw\alpha_{w} in the order of 1010. In case fine-tuning is needed, a proper range for αcurv\alpha_{\text{curv}} is 103110^{-3}\sim 1, where smaller values are suitable for complex images, and larger values for simple ones. As for αw\alpha_{w}, keeping it above 11 is a good practice. Investigations of their effects on the decomposition are collected in Section 5.3.

We conduct our second set of experiments to search for appropriate αn\alpha_{n} while fixing αcurv=0.1\alpha_{\text{curv}}=0.1 and αw=50\alpha_{w}=50. Since αn\alpha_{n} is directly associated with the noise level, we show in Figure 13 the PSNR values of uu^{*} components learned from test images by the proposed model, where both αn\alpha_{n} and the noise level σ\sigma vary. We observe that for noise levels below 2020, the optimal choices of αn\alpha_{n} stably remain around 0.1100.1\sim 10. When the noise level grows above 2020, we observe that the optimal choice of αn\alpha_{n} has a clear dependence on σ\sigma. In particular, smaller αn\alpha_{n} is needed for larger σ\sigma to capture oscillations due to noises. We could fit αn\alpha_{n} in log-scale with σ>20\sigma>20 using a quadratic polynomial; instead, we suggest a strategy for images with noise σ<100\sigma<100 with better generalizability. We roughly divide the levels of noise into three regimes: low noise (σ<20\sigma<20), medium noise (20σ<6020\leq\sigma<60), and high noise (60σ<10060\leq\sigma<100). For images with low noise, use αn=10\alpha_{n}=10; for images with medium noise, use αn=1×102\alpha_{n}=1\times 10^{-2}; and for those with heavy noise, use αn=1×104\alpha_{n}=1\times 10^{-4}.

With αcurv\alpha_{\text{curv}}, αw\alpha_{w}, and αn\alpha_{n} determined above, we proceed to tune α0\alpha_{0}. Using various values of α0\alpha_{0}, we record in Figure 14 the corresponding PSNR values of the uu^{*} components for the three test images with different levels of noise. Specifically, we add Gaussian noises with σ=10/255\sigma=10/255 in (a), σ=40/255\sigma=40/255 in (b), and σ=80/255\sigma=80/255 in (c). As discussed above, we use αn=10\alpha_{n}=10 for (a), αn=1×102\alpha_{n}=1\times 10^{-2} for (b), and αn=1×104\alpha_{n}=1\times 10^{-4} for (c); and we fix αcurv=0.1\alpha_{\text{curv}}=0.1 and αw=50\alpha_{w}=50. For the tested images, when the noise is mild as in (a) and (b), we observe two plateaus in PSNR values of uu^{*} components connected by decreasing phases between 1×103<α0<1×1021\times 10^{-3}<\alpha_{0}<1\times 10^{-2}. The transitions from high PSNR to low PSNR regime are smooth, and they correspond to removing rich image details such as textures. However, when the noise level is high as in (c), we observe that as α0\alpha_{0} increases, the PSNR values improve slightly. This shows that L0L^{0}-gradient can contribute to perturbation reduction in case of heavy noise. For general images, we suggest using α0=1×103\alpha_{0}=1\times 10^{-3} as the default.

6 Conclusion

In this paper, we propose a novel model for decomposing grayscale images into three components: the structural part, the smooth part, and the oscillatory part, each of which has a distinctive visual feature. To take advantage of the strong sparsity induced by the L0L^{0}-gradient regularization while avoiding the undesirable staircase effect, our model integrates the curvature regularization on the level lines of the structural part. As a result, the reconstructed structural part effectively preserves sharp boundaries enclosing homogeneous regions. To solve the associated energy minimization problem, we develop an efficient numerical algorithm based on operator splitting. The original non-convex non-smooth problem is separated into four sub-problems. Each of them either admits a closed-form solution or enjoys a form suitable for applying FFT. We present various experiments to justify the proposed model. In particular, we show that our model successfully separates underlying patterns from soft shadows even when both of them have complicated geometries with fine details. We also justify the regularization terms included in the proposed model and illustrate the effects of parameters. We compare our model with CEP [13] and HKLM [29] for three-component decomposition and show that our model achieves visually appealing results and outstanding computational efficiency. To amend the loss of fine-scale details, a natural extension is to model noise and textures separately. The proposed method can also be extended to color image decomposition when applied channel-wise. To avoid potential artifacts, we leave it as a future work to investigate appropriate regularizers that leverage correlations among color channels.

Acknowledgement

We would like to thank Dr. Martin Huska at University of Bologna for kindly sharing the code of [29]. We would also like to thank Prof. Sung Ha Kang at Georgia Institute of Technology for inspiring discussions on high-order regularizations.

References

  • [1] L. Ambrosio and S. Masnou. A direct variational approach to a problem arising in image reconstruction. Interfaces and Free Boundaries, 5(1):63–81, 2003.
  • [2] G. Anzellotti. Pairings between measures and bounded functions and compensated compactness. Annali di Matematica pura ed applicata, 135:293–318, 1983.
  • [3] G. Aubert and J.-F. Aujol. Modeling very oscillating signals. Application to image processing. Applied Mathematics and Optimization, 51:163–182, 2005.
  • [4] J.-F. Aujol, G. Aubert, L. Blanc-Féraud, and A. Chambolle. Image decomposition into a bounded variation component and an oscillating component. Journal of Mathematical Imaging and Vision, 22:71–88, 2005.
  • [5] J.-F. Aujol and A. Chambolle. Dual norms and image decomposition models. International Journal of Computer Vision, 63:85–104, 2005.
  • [6] J.-F. Aujol, G. Gilboa, T. Chan, and S. Osher. Structure-texture image decomposition—modeling, algorithms, and parameter selection. International Journal of Computer Vision, 67:111–136, 2006.
  • [7] J.-F. Aujol and S. H. Kang. Color image decomposition and restoration. Journal of Visual Communication and Image Representation, 17(4):916–928, 2006.
  • [8] G. Bellettini et al. Variational approximation of functionals with curvatures and related properties. Journal of Convex Analysis, 4:91–108, 1997.
  • [9] P. Blomgren, T. F. Chan, P. Mulet, and C.-K. Wong. Total variation image restoration: numerical methods and extensions. In Proceedings of International Conference on Image Processing, volume 3, pages 384–387. IEEE, 1997.
  • [10] A. Buades, T. M. Le, J.-M. Morel, and L. A. Vese. Fast cartoon+texture image filters. IEEE Transactions on Image Processing, 19(8):1978–1986, 2010.
  • [11] A. Chambolle. An algorithm for total variation minimization and applications. Journal of Mathematical Imaging and Vision, 20:89–97, 2004.
  • [12] T. Chan, S. Esedoglu, F. Park, A. Yip, et al. Recent developments in total variation image restoration. Mathematical Models of Computer Vision, 17(2):17–31, 2005.
  • [13] T. F. Chan, S. Esedoglu, and F. E. Park. Image decomposition combining staircase reduction and texture extraction. Journal of Visual Communication and Image Representation, 18(6):464–486, 2007.
  • [14] K. Chen and H. Han. Three-stage approach for 2d/3d diffeomorphic multi-modality image registration with textural control: diffeomorphic multi-modality image registration. SIAM Journal on Imaging Sciences, 2024.
  • [15] Q. Chen and V. Koltun. A simple model for intrinsic image decomposition with depth cues. In Proceedings of the IEEE International Conference on Computer Vision, pages 241–248, 2013.
  • [16] E. De Giorgi and L. Ambrosio. New functionals in calculus of variations. In Nonsmooth Optimization and Related Topics, pages 49–59. Springer, 1989.
  • [17] L.-J. Deng, R. Glowinski, and X.-C. Tai. A new operator splitting method for the Euler elastica model for image smoothing. SIAM Journal on Imaging Sciences, 12(2):1190–1230, 2019.
  • [18] Q. Ding, T. Niu, X. Zhang, and Y. Long. Image-domain multimaterial decomposition for dual-energy CT based on prior information of material images. Medical physics, 45(8):3614–3626, 2018.
  • [19] D. L. Donoho and I. M. Johnstone. Adapting to unknown smoothness via wavelet shrinkage. Journal of the American Statistical Association, 90(432):1200–1224, 1995.
  • [20] V. Duval, J.-F. Aujol, and L. Vese. Projected gradient based color image decomposition. In Scale Space and Variational Methods in Computer Vision: Second International Conference, SSVM 2009, Voss, Norway, June 1-5, 2009. Proceedings 2, pages 295–306. Springer, 2009.
  • [21] E. Esser. Applications of lagrangian-based alternating direction methods and connections to split Bregman. CAM report, 9:31, 2009.
  • [22] J. B. Garnett, P. W. Jones, T. M. Le, and L. A. Vese. Modeling oscillatory components with the homogeneous spaces BMO˙α\dot{\text{BMO}}^{-\alpha} and W˙α,p\dot{W}^{-\alpha,p^{*}}. Pure and Applied Mathematics Quarterly, 7(2):275–318, 2011.
  • [23] G. Gilboa and S. Osher. Nonlocal operators with applications to image processing. Multiscale Modeling & Simulation, 7(3):1005–1028, 2009.
  • [24] R. Glowinski and P. Le Tallec. Augmented Lagrangian and operator-splitting methods in nonlinear mechanics. SIAM, 1989.
  • [25] R. Glowinski, S. J. Osher, and W. Yin. Splitting methods in communication, imaging, science, and engineering. Springer, 2017.
  • [26] R. Glowinski, T.-W. Pan, and X.-C. Tai. Some facts about operator-splitting and alternating direction methods. Splitting Methods in Communication, Imaging, Science, and Engineering, pages 19–94, 2016.
  • [27] Y. He, S. H. Kang, and H. Liu. Curvature regularized surface reconstruction from point clouds. SIAM Journal on Imaging Sciences, 13(4):1834–1859, 2020.
  • [28] Y. Hu and M. Jacob. Higher degree total variation (HDTV) regularization for image recovery. IEEE Transactions on Image Processing, 21(5):2559–2571, 2012.
  • [29] M. Huska, S. H. Kang, A. Lanza, and S. Morigi. A variational approach to additive image decomposition into structure, harmonic, and oscillatory components. SIAM Journal on Imaging Sciences, 14(4):1749–1789, 2021.
  • [30] Y. Lan, Z. Li, J. Sun, and Y. Xiang. Dosnet as a non-black-box pde solver: When deep learning meets operator splitting. Journal of Computational Physics, 491:112343, 2023.
  • [31] T. M. Le and L. A. Vese. Image decomposition using total variation and div(BMO). Multiscale Modeling & Simulation, 4(2):390–423, 2005.
  • [32] N. Le Bihan and S. Sangwine. Color image decomposition using quaternion singular value decomposition. In 2003 International Conference on Visual Information Engineering VIE 2003, pages 113–116. IET, 2003.
  • [33] L. H. Lieu and L. A. Vese. Image restoration and decomposition via bounded total variation and negative Hilbert-Sobolev spaces. Applied Mathematics and Optimization, 58:167–193, 2008.
  • [34] H. Liu, J. Liu, R. Chan, and X.-C. Tai. Double-well net for image segmentation. arXiv preprint arXiv:2401.00456, 2023.
  • [35] H. Liu, X.-C. Tai, and R. Glowinski. An operator-splitting method for the Gaussian curvature regularization model with applications to surface smoothing and imaging. SIAM Journal on Scientific Computing, 44(2):A935–A963, 2022.
  • [36] H. Liu, X.-C. Tai, R. Kimmel, and R. Glowinski. A color elastica model for vector-valued image regularization. SIAM Journal on Imaging Sciences, 14(2):717–748, 2021.
  • [37] H. Liu, X.-C. Tai, R. Kimmel, and R. Glowinski. Elastica models for color image regularization. SIAM Journal on Imaging Sciences, 16(1):461–500, 2023.
  • [38] M. Lysaker, A. Lundervold, and X.-C. Tai. Noise removal using fourth-order partial differential equation with applications to medical magnetic resonance images in space and time. IEEE Transactions on Image Processing, 12(12):1579–1590, 2003.
  • [39] S. Masnou and J.-M. Morel. Level lines based disocclusion. In Proceedings 1998 International Conference on Image Processing. ICIP98 (Cat. No. 98CB36269), pages 259–263. IEEE, 1998.
  • [40] S. Masnou and J.-M. Morel. On a variational theory of image amodal completion. Rendiconti del Seminario Matematico della Universita di Padova, 116:211–252, 2006.
  • [41] Y. Meyer. Oscillating patterns in image processing and nonlinear evolution equations: the fifteenth Dean Jacqueline B. Lewis memorial lectures, volume 22. American Mathematical Soc., 2001.
  • [42] D. Mumford. Elastica and computer vision. In Algebraic Geometry and its Applications: Collections of Papers from Shreeram S. Abhyankar’s 60th Birthday Conference, pages 491–506. Springer, 1994.
  • [43] R. M. Nguyen and M. S. Brown. Fast and effective L0L_{0} gradient minimization by region fusion. In Proceedings of the IEEE International Conference on Computer Vision, pages 208–216, 2015.
  • [44] S. Ono. L0L_{0} gradient projection. IEEE Transactions on Image Processing, 26(4):1554–1564, 2017.
  • [45] S. Osher, A. Solé, and L. Vese. Image decomposition and restoration using total variation minimization and the H1H^{-1} norm. Multiscale Modeling & Simulation, 1(3):349–370, 2003.
  • [46] M. T. Pimenta and M. Montenegro. Existence of a BV solution for a mean curvature equation. Journal of Differential Equations, 299:51–64, 2021.
  • [47] L. I. Rudin, S. Osher, and E. Fatemi. Nonlinear total variation based noise removal algorithms. Physica D: Nonlinear Phenomena, 60(1-4):259–268, 1992.
  • [48] J. Shen. Piecewise h1+h0+h1h^{-1}+h^{0}+h^{1} images and the mumford-shah-sobolevmodel for segmented image decomposition. Applied Mathematics Research Express, 2005(4):143–167, 2005.
  • [49] J. Shen, S. H. Kang, and T. F. Chan. Euler’s elastica and curvature-based inpainting. SIAM Journal on Applied Mathematics, 63(2):564–592, 2003.
  • [50] X.-C. Tai, J. Hahn, and G. J. Chung. A fast algorithm for Euler’s elastica model using augmented lagrangian method. SIAM Journal on Imaging Sciences, 4(1):313–344, 2011.
  • [51] X.-C. Tai, H. Liu, and R. Chan. PottsMGNet: A mathematical explanation of encoder-decoder based neural networks. SIAM Journal on Imaging Sciences, 17(1):540–594, 2024.
  • [52] L. A. Vese and S. J. Osher. Modeling textures with total variation minimization and oscillating patterns in image processing. Journal of Scientific Computing, 19:553–572, 2003.
  • [53] Y.-W. Wen, H.-W. Sun, and M. K. Ng. A primal-dual method for the meyer model of cartoon and texture decomposition. Numerical Linear Algebra with Applications, 26(2):e2224, 2019.
  • [54] Y.-W. Wen, M. Zhao, and M. Ng. Cartoon and texture decomposition for color image in opponent color space. Applied Mathematics and Computation, 414:126654, 2022.
  • [55] X. Wu, J. Zheng, C. Wu, and Y. Cai. Variational structure–texture image decomposition on manifolds. Signal Processing, 93(7):1773–1784, 2013.
  • [56] J. Xu, X. Feng, Y. Hao, and Y. Han. Image decomposition and staircase effect reduction based on total generalized variation. Journal of Systems Engineering and Electronics, 25(1):168–174, 2014.
  • [57] L. Xu, C. Lu, Y. Xu, and J. Jia. Image smoothing via L0L_{0} gradient minimization. In Proceedings of the 2011 SIGGRAPH Asia Conference, pages 1–12, 2011.
  • [58] M. Yashtini and S. H. Kang. Alternating direction method of multiplier for Euler’s elastica-based denoising. In Scale Space and Variational Methods in Computer Vision: 5th International Conference, SSVM 2015, Lège-Cap Ferret, France, May 31-June 4, 2015, Proceedings 5, pages 690–701. Springer, 2015.
  • [59] C.-H. Zhang. Nearly unbiased variable selection under minimax concave penalty. The Annals of Statistics, 38:894–942, 2010.
  • [60] H. Zhang, L. Tang, Z. Fang, C. Xiang, and C. Li. Nonconvex and nonsmooth total generalized variation model for image restoration. Signal Processing, 143:69–85, 2018.
  • [61] W. Zhu. A first-order image denoising model for staircase reduction. Advances in Computational Mathematics, 45(5):3217–3239, 2019.
  • [62] W. Zhu and T. Chan. Image denoising using mean curvature of image surface. SIAM Journal on Imaging Sciences, 5(1):1–32, 2012.

Appendix A Proof of Proposition 2.1

Proof.

The proof is analogous to [17]. Denote

E1(v,w,n)=α0v0+αcurvΩ(v|v|)2|Dv|𝑑𝐱+αwΔwL2(Ω)2+αnnH1(Ω)2,\displaystyle E_{1}(v,w,n)=\alpha_{0}\|\nabla v\|_{0}+\alpha_{\text{curv}}\int_{\Omega}\left(\nabla\cdot\frac{\nabla v}{|\nabla v|}\right)^{2}\,\left|Dv\right|\,d\mathbf{x}+\alpha_{w}\|\Delta w\|_{L^{2}(\Omega)}^{2}+\alpha_{n}\|n\|_{H^{-1}(\Omega)}^{2},
E2(v,w,n)=12Ω|f(v+w+n)|22𝑑𝐱.\displaystyle E_{2}(v,w,n)=\frac{1}{2}\int_{\Omega}|f-(v+w+n)|_{2}^{2}d\mathbf{x}.

Let (v,w,n)(v^{*},w^{*},n^{*}) be a minimizer of Model (8). Consider the tuple (v+c1,w+c2,n)(v^{*}+c_{1},w^{*}+c_{2},n^{*}), where c1,c2c_{1},c_{2}\in\mathbb{R} are scalar fields. Notice that

E1(v,w,n)=E1(v+c1,w+c2,n),\displaystyle E_{1}(v^{*},w^{*},n^{*})=E_{1}(v^{*}+c_{1},w^{*}+c_{2},n^{*})\;,

and

E2(v+c1,w+c2,n)=\displaystyle E_{2}(v^{*}+c_{1},w^{*}+c_{2},n^{*})= 12Ω|f(v+c1+w+c2+n)|2𝑑𝐱\displaystyle\frac{1}{2}\int_{\Omega}|f-(v^{*}+c_{1}+w^{*}+c_{2}+n^{*})|^{2}d\mathbf{x}
=\displaystyle= E2(v,w,n)+(c1+c2)Ω(v+w+nf)𝑑𝐱+|Ω|2(c1+c2)2,\displaystyle E_{2}(v^{*},w^{*},n^{*})+(c_{1}+c_{2})\int_{\Omega}(v^{*}+w^{*}+n^{*}-f)d\mathbf{x}+\frac{|\Omega|}{2}(c_{1}+c_{2})^{2}, (61)

where |Ω||\Omega| is the area of Ω\Omega. The function on the right-hand-side of (61) is quadratic in (c1+c2)(c_{1}+c_{2}), which takes its minimal value when

c1+c2=Ω(f(v+w+n))𝑑𝐱|Ω|.\displaystyle c_{1}+c_{2}=\frac{\int_{\Omega}(f-(v^{*}+w^{*}+n^{*}))d\mathbf{x}}{|\Omega|}. (62)

If Ω(f(v+w+n))𝑑𝐱0\int_{\Omega}(f-(v^{*}+w^{*}+n^{*}))d\mathbf{x}\neq 0, it would imply that

E1(v,w,n)+E2(v,w,n)>E1(u+c1,w+c2,n)+E2(u+c1,w+c2,n),\displaystyle E_{1}(v^{*},w^{*},n^{*})+E_{2}(v^{*},w^{*},n^{*})>E_{1}(u^{*}+c_{1},w^{*}+c_{2},n^{*})+E_{2}(u^{*}+c_{1},w^{*}+c_{2},n^{*}),

which contradicts to the assumption that (v,w,n)(v^{*},w^{*},n^{*}) is a minimizer of Model (8). We thus conclude that Ωv+w+nd𝐱=Ωf𝑑𝐱\int_{\Omega}v^{*}+w^{*}+n^{*}d\mathbf{x}=\int_{\Omega}fd\mathbf{x}. ∎

Appendix B Equivalence between Hessian regularizer and Laplacian regularizer

We compare the following two regularizers

EH(w)=122wL2(Ω)2 and EL(w)=12ΔwL2(Ω)2.\displaystyle E_{H}(w)=\frac{1}{2}\|\partial^{2}w\|_{L^{2}(\Omega)}^{2}\quad\mbox{ and }\quad E_{L}(w)=\frac{1}{2}\|\Delta w\|_{L^{2}(\Omega)}^{2}.

Functional EHE_{H} is used in [29] and ELE_{L} is used in this paper as a regularizer for the ww component. It was observed in [13] that EH=ELE_{H}=E_{L} for compactly supported smooth functions. We show that more general boundary conditions, EHE_{H} and ELE_{L} have the same set of critical points. To prove that, it is sufficient to show that their variations concerning ww have the same form. For simplicity, we assume Ω=[0,L1]×[0,L2]\Omega=[0,L_{1}]\times[0,L_{2}] for some L1,L2>0L_{1},L_{2}>0.

The functional EHE_{H} can be written as

EH(w)=12Ω(2wx12)2+2(2wx1x2)2+(2wx22)2d𝐱.\displaystyle E_{H}(w)=\frac{1}{2}\int_{\Omega}\left(\frac{\partial^{2}w}{\partial x_{1}^{2}}\right)^{2}+2\left(\frac{\partial^{2}w}{\partial x_{1}\partial x_{2}}\right)^{2}+\left(\frac{\partial^{2}w}{\partial x_{2}^{2}}\right)^{2}d\mathbf{x}.

The Fréchet derivative of EHE_{H} with respect to ww in the direction hh is

EHw,h=\displaystyle\left\langle\frac{\partial E_{H}}{\partial w},h\right\rangle= Ω(4wx14+24wx12x22+4wx24)h𝑑𝐱\displaystyle\int_{\Omega}\left(\frac{\partial^{4}w}{\partial x_{1}^{4}}+2\frac{\partial^{4}w}{\partial x_{1}^{2}\partial x_{2}^{2}}+\frac{\partial^{4}w}{\partial x_{2}^{4}}\right)h\ d\mathbf{x}
2wx12,hx1|x1=0x1=L1+3wx13,h|x1=0x1=L12wx22,hx2|x2=0x2=L2+3wx23,h|x2=0x2=L2\displaystyle-\left.\left\langle\frac{\partial^{2}w}{\partial x_{1}^{2}},\frac{\partial h}{\partial x_{1}}\right\rangle\right|_{x_{1}=0}^{x_{1}=L_{1}}+\left.\left\langle\frac{\partial^{3}w}{\partial x_{1}^{3}},h\right\rangle\right|_{x_{1}=0}^{x_{1}=L_{1}}-\left.\left\langle\frac{\partial^{2}w}{\partial x_{2}^{2}},\frac{\partial h}{\partial x_{2}}\right\rangle\right|_{x_{2}=0}^{x_{2}=L_{2}}+\left.\left\langle\frac{\partial^{3}w}{\partial x_{2}^{3}},h\right\rangle\right|_{x_{2}=0}^{x_{2}=L_{2}}
2wx1x2,hx2|x1=0x1=L1+3wx12x2,h|x2=0x2=L2\displaystyle-\left.\left\langle\frac{\partial^{2}w}{\partial x_{1}\partial x_{2}},\frac{\partial h}{\partial x_{2}}\right\rangle\right|_{x_{1}=0}^{x_{1}=L_{1}}+\left.\left\langle\frac{\partial^{3}w}{\partial x_{1}^{2}\partial x_{2}},h\right\rangle\right|_{x_{2}=0}^{x_{2}=L_{2}}
2wx1x2,hx1|x2=0x2=L2+3wx1x22,h|x1=0x1=L1.\displaystyle-\left.\left\langle\frac{\partial^{2}w}{\partial x_{1}\partial x_{2}},\frac{\partial h}{\partial x_{1}}\right\rangle\right|_{x_{2}=0}^{x_{2}=L_{2}}+\left.\left\langle\frac{\partial^{3}w}{\partial x_{1}\partial x_{2}^{2}},h\right\rangle\right|_{x_{1}=0}^{x_{1}=L_{1}}. (63)

The functional ELE_{L} can be written as

EL(w)=12Ω(2wx12)2+2(2wx122wx22)+(2wx22)2d𝐱.\displaystyle E_{L}(w)=\frac{1}{2}\int_{\Omega}\left(\frac{\partial^{2}w}{\partial x_{1}^{2}}\right)^{2}+2\left(\frac{\partial^{2}w}{\partial x_{1}^{2}}\frac{\partial^{2}w}{\partial x_{2}^{2}}\right)+\left(\frac{\partial^{2}w}{\partial x_{2}^{2}}\right)^{2}d\mathbf{x}.

The Fréchet derivative of ELE_{L} with respect to ww in the direction hh is

ELw,h=\displaystyle\left\langle\frac{\partial E_{L}}{\partial w},h\right\rangle= Ω(4wx14+24wx12x22+4wx24)h𝑑𝐱\displaystyle\int_{\Omega}\left(\frac{\partial^{4}w}{\partial x_{1}^{4}}+2\frac{\partial^{4}w}{\partial x_{1}^{2}\partial x_{2}^{2}}+\frac{\partial^{4}w}{\partial x_{2}^{4}}\right)h\ d\mathbf{x}
(2wx12+2wx22),hx1|x1=0x1=L1+(3wx13+3wx12x2),h|x1=0x1=L1\displaystyle-\left.\left\langle\left(\frac{\partial^{2}w}{\partial x_{1}^{2}}+\frac{\partial^{2}w}{\partial x_{2}^{2}}\right),\frac{\partial h}{\partial x_{1}}\right\rangle\right|_{x_{1}=0}^{x_{1}=L_{1}}+\left.\left\langle\left(\frac{\partial^{3}w}{\partial x_{1}^{3}}+\frac{\partial^{3}w}{\partial x_{1}^{2}\partial x_{2}}\right),h\right\rangle\right|_{x_{1}=0}^{x_{1}=L_{1}}
(2wx22+2wx22),hx2|x2=0x2=L2+(3wx23+3wx1x22),h|x2=0x2=L2.\displaystyle-\left.\left\langle\left(\frac{\partial^{2}w}{\partial x_{2}^{2}}+\frac{\partial^{2}w}{\partial x_{2}^{2}}\right),\frac{\partial h}{\partial x_{2}}\right\rangle\right|_{x_{2}=0}^{x_{2}=L_{2}}+\left.\left\langle\left(\frac{\partial^{3}w}{\partial x_{2}^{3}}+\frac{\partial^{3}w}{\partial x_{1}\partial x_{2}^{2}}\right),h\right\rangle\right|_{x_{2}=0}^{x_{2}=L_{2}}. (64)

By using the boundary conditon

2wx12=2wx22=3wx13=3wx23=3wx12x2=3wx1x22=0 on Ω,\displaystyle\frac{\partial^{2}w}{\partial x_{1}^{2}}=\frac{\partial^{2}w}{\partial x_{2}^{2}}=\frac{\partial^{3}w}{\partial x_{1}^{3}}=\frac{\partial^{3}w}{\partial x_{2}^{3}}=\frac{\partial^{3}w}{\partial x_{1}^{2}\partial x_{2}}=\frac{\partial^{3}w}{\partial x_{1}\partial x_{2}^{2}}=0\mbox{ on }\partial\Omega,

all boundary integrals in (63) and (64) vanish. We get the same Fréchet derivative for both EHE_{H} and ELE_{L}

EHw,h=ELw,h=Ω(4wx14+24wx12x22+4wx24)h𝑑𝐱.\displaystyle\left\langle\frac{\partial E_{H}}{\partial w},h\right\rangle=\left\langle\frac{\partial E_{L}}{\partial w},h\right\rangle=\int_{\Omega}\left(\frac{\partial^{4}w}{\partial x_{1}^{4}}+2\frac{\partial^{4}w}{\partial x_{1}^{2}\partial x_{2}^{2}}+\frac{\partial^{4}w}{\partial x_{2}^{4}}\right)h\ d\mathbf{x}.

Therefore, EHE_{H} and ELE_{L} have the same optimality condition and thus the same set of critical points.