888
Views
9
CrossRef citations to date
0
Altmetric
Article

Dualization and Automatic Distributed Parameter Selection of Total Generalized Variation via Bilevel Optimization

ORCID Icon, ORCID Icon, &
Pages 887-932 | Received 17 Feb 2021, Accepted 19 Apr 2022, Published online: 05 May 2022

Abstract

Total Generalized Variation (TGV) regularization in image reconstruction relies on an infimal convolution type combination of generalized first- and second-order derivatives. This helps to avoid the staircasing effect of Total Variation (TV) regularization, while still preserving sharp contrasts in images. The associated regularization effect crucially hinges on two parameters whose proper adjustment represents a challenging task. In this work, a bilevel optimization framework with a suitable statistics-based upper level objective is proposed in order to automatically select these parameters. The framework allows for spatially varying parameters, thus enabling better recovery in high-detail image areas. A rigorous dualization framework is established, and for the numerical solution, a Newton type method for the solution of the lower level problem, i.e. the image reconstruction problem, and a bilevel TGV algorithm are introduced. Denoising tests confirm that automatically selected distributed regularization parameters lead in general to improved reconstructions when compared to results for scalar parameters.

1. Introduction

In this work we analyze and implement a bilevel optimization framework for automatically selecting spatially varying regularization parameters α:=(α0,α1)C(Ω¯)2,α>0, in the following image reconstruction problem: (1.1) minimize12Ω(Tuf)2dx+TGVα2(u)over uBV(Ω),(1.1) where the second-order Total Generalized Variation (TGV) regularization is given by (1.2) TGVα2(u)=sup{Ωu div2ϕ dx:ϕCc(Ω,Sd×d),|ϕ(x)|α0(x),|divϕ(x)|α1(x), for all xΩ}.(1.2)

Here, ΩRd is a bounded, open image domain with Lipschitz boundary, Sd×d denotes the space of d × d symmetric matrices, T:L2(Ω)L2(Ω) is a bounded linear (output) operator, and f denotes given data which satisfies (1.3) f=Tutrue+η.(1.3)

In this context, η models a highly oscillatory (random) component with zero mean and known quadratic deviation (variance) σ2 from the mean. Further, L2(Ω) denotes the standard Lebesgue space [Citation1], and |·|, represents the Euclidean vector norm or its associated matrix norm. The space of infinitely differentiable functions with compact support in Ω and values in Sd×d is denoted by Cc(Ω,Sd×d). Further, we refer to Section 2 for the definition of the first- and second-order divergences div and div2, respectively.

Originally, the TGV functional was introduced for scalar parameters α0,α1>0 only; see [Citation2]. It serves as a higher order extension of the well-known Total Variation (TV) regularizer [Citation3,Citation4], preserves edges (i.e., sharp contrast) [Citation5,Citation6], and promotes piecewise affine reconstructions while avoiding the often adverse staircasing effect (i.e., piecewise constant structures) of TV [Citation7–9]; see for an illustration. These properties of TGV have made it a successful regularizer in variational image restoration for a variety of applications [Citation2, Citation10–16]. Extensions to manifold-valued data, multimodal and dynamic problems [Citation17–22] have been proposed as well. In all of these works, the choice of the scalar parameters α0,α1 is made “manually” via a direct grid search. Alternatively, selection schemes relying on a known ground truth utrue have been studied; see [Citation23–26]. The latter approach, however, is primarily of interest when investigating the mere capabilities of TGV regularization.

Figure 1. Gaussian denoising: Typical difference between TV (piecewise constant) and TGV reconstructions (piecewise affine).

Figure 1. Gaussian denoising: Typical difference between TV (piecewise constant) and TGV reconstructions (piecewise affine).

While there exist automated parameter choice rules for TV regularization, see for instance [Citation27] and the references therein, analogous techniques and results for the TGV parameters are very scarce. One of the very few contributions is [Citation28] where, however, a spatially varying fidelity weight rather then regularization parameter is computed. Compared to the choice of the regularization weight in TV-based models, the infimal convolution type regularization incorporated into the TGV functional significantly complicates the selection; compare the equivalent definition (2.1) below. Further difficulties arise when these parameters are spatially varying as in (1.2). In that case, by appropriately choosing α=(α0,α1), one wishes to smoothen homogeneous areas in the image while preserving fine scale details. The overall target is then to not only select the parameters in order to reduce noise while avoiding oversmoothing, as in the TV case, but also to ensure that the interplay of α0 and α1 will not produce any staircasing.

For this delicate selection task and inspired by [Citation27, Citation29] for TV, in this work we propose a bilevel minimization framework for an automated selection of α in the TGV case. Formally, the setting can be characterized as follows: (1.4) {minimize a statisticsbased (upper level) objective over (u,α)subjecttousolving(1.1)for a regularization weightα=(α0,α1).(1.4)

Note here that the optimization variable α enters the lower level minimization problem (1.1) as a parameter, thus giving rise to u=u(α). We also mention that this optimization format falls into the general framework which is discussed in our review paper [Citation30] where the general opportunities and mathematical as well as algorithmic aspects of bilevel optimization in generating structured non-smooth regularization functionals are discussed in detail.

As our statistical set-up parallels the one in [Citation27, Citation29], here we resort to the upper level objective proposed in that work. It is based on localized residuals R:L2(Ω)L(Ω) with (1.5) Ru(x)=Ωw(x,y)(Tuf)2(y) dy,(1.5) where wL(Ω×Ω) with ΩΩw(x,y)dxdy=1. Note that Ru(x) can be interpreted as a local variance keeping in mind that, assuming Gaussian noise of variance σ2, we have that Ω(Tutruef)2 dx=Ωη2 dx=σ2|Ω|. Consequently, if a reconstructed image u is close to utrue then it is expected that for every xΩ the value of Ru(x) will be close to σ2. Hence it is natural to consider an upper level objective which aims to approximately keep Ru within a corridor σ¯2σ2σ¯2 with positive bounds σ¯2,σ¯2. This can be achieved by utilizing the function F:L2(Ω)R with (1.6) F(v):=12Ωmax(vσ¯2,0)2dx+12Ωmin(vσ¯2,0)2dx.(1.6)

The function F(R·) is then indeed suitable as an upper level objective. This is demonstrated in , where we show (in the middle and right plots) the objective values for a series of scalar TGV denoising results and for a variety of parameters (α0,α1) for the image depicted on the left. Regarding the choices of σ¯,σ¯,w we refer to Section 5. Upon inspection of we find that the functional F(R·) is minimized for a pair of scalar parameters (α0,α1) that is close to the one maximizing the peak-signal-to-noise-ratio (PSNR). Note, however, that in order to truly optimize the PSNR, one would need the ground truth image utrue, which is of course typically not available. In contrast to this, we emphasize that F(R·) does not involve any ground truth information. Rather, it only relies on statistical properties of the noise.

Figure 2. Suitability of the functional F(R·) as an upper level objective. Evaluation of F(Ru) where u solves the TGV denoising problem (1.1) (T = Id), for a variety of scalar parameters (α0,α1).

Figure 2. Suitability of the functional F(R·) as an upper level objective. Evaluation of F(Ru) where u solves the TGV denoising problem (1.1) (T = Id), for a variety of scalar parameters (α0,α1).

For analytical and numerical reasons, rather than having (1.1) as the lower level problem for the bilevel minimization framework (1.4), one can use alternative formulations, as was done for instance in [Citation27, Citation29] for TV models, where the Fenchel predual problem was used instead, see also [Citation30] for a thorough discussion on the corresponding TGV model. This yields a bilevel problem which is expressed in terms of dual variables and is equivalent to the one stated in terms of the primal variable u. In this way, one has to treat a more amenable variational inequality of the first kind rather than one of second kind in the primal setting in the constraint system of the resulting bilevel optimization problem. Numerically, one may then utilize very efficient and resolution independent, function space based solution algorithms, like (inexact) semismooth Newton methods [Citation31]. The other option that we will consider in this work, is to minimize the upper level objective subject to the primal-dual optimality conditions, for which Newton methods can also be applied for their solution, see for instance [Citation32] for an inexact semismooth Newton solver which operates on the primal-dual optimality conditions for TV regularization. We should also mention that this approach requires TT to be invertible, with T being the adjoint of T, which is true when T is injective with closed range. We note that this does not exclude the use of our bilevel scheme to inverse problems whose forward operator does not satisfy this condition. A noninvertibility of TT can for instance be treated by adding a small regularization term of the form κ2Ωu2 dx, in combination with a Levenberg-Marquardt algorithmic scheme that sends the parameter κ to zero along the iterations. Alternatively, instead of Newton, one can resort to first-order methods for solving the lower level problem for which invertibility of TT is not required. Such approaches are definitely interesting and necessary for many inverse problems however since they would deviate from the main focus of our paper which is the automatic computation of spatially distributed regularization weights, we will keep this invertibility assumption in the following.

2. The structure of the paper

Basic facts on the TGV functional with spatially varying parameters along with functional analytic foundations needed for (pre)dualization are the subjects of Section 2. Section 2.4 is concerned with the derivation of the predual problem of (1.1) and the corresponding primal-dual optimality conditions. Regularized versions of these conditions are the focus of Section 3. Besides respective primal-dual optimality conditions, we study the asymptotic behavior of these problems and their associated solutions under vanishing regularization. Section 4 introduces the bilevel TGV problem for which the primal-dual optimality conditions serve as constraints. The numerical solution of the proposed bilevel problem is the subject of Section 5. It is also argued that every regularized instance of the lower level problem can be solved efficiently by employing an (inexact) semismooth Newton method. The paper ends by a report on extensive numerical tests along with conclusions drawn from these computational results.

Summarizing, this work provides not only a user-friendly and novel hierarchical variational framework for automatic selection of the TGV regularization parameters, but by making these parameters spatially dependent it leads to an overall performance improvement; compare, e.g., the results in Section 5.

3. The dual form of the weighted TGV functional

3.1. Total generalized variation

We recall here some basic facts about the TGV functional (1.2) with constant parameters α0,α1 and assume throughout that the reader is familiar with the basic concepts of functions of bounded variation (BV); see [Citation33] for a detailed account. For a function ϕCc(Ω,Sd×d) the first- and second-order divergences are, respectively, given by (divϕ)i=j=1dϕijxj,i=1,,d,anddiv2ϕ=i=1d2ϕiixi2+2i<j2ϕijxixj.

In [Citation14] it was shown that a function uL1(Ω) has finite TGV value if and only if it belongs to BV(Ω). Here BV(Ω) denotes the Banach space of function of bounded variation over Ω with associated norm ·BV(Ω). Moreover, the bounded generalized variation norm ·BGV:=·L1(Ω)+TGVα2(·) is equivalent to ·BV(Ω). Similarly to TV, TGV is a convex functional which is lower semicontinuous with respect to the strong L1 convergence. In [Citation14, Citation34] it is demonstrated that the TGV functional can be equivalently written as (2.1) TGVα2(u)=minwBD(Ω)α1|Duw|(Ω)+α0|Ew|(Ω),(2.1) where BD(Ω) is the Banach space of functions of bounded deformation, with E denoting the distributional symmetrized gradient [Citation35,Citation36]. The asymptotical behavior of the TGV model in image restoration with respect to scalars α0,α1 was studied in [Citation37]; see also [Citation6]. For instance, when T = Id and either α0 or α1 converges to zero, then the corresponding solutions of (1.1) converge (weakly in BV(Ω)) to f. When both of the parameters are sent to infinity, then the solutions converge weakly to the L2-linear regression solution for f. We further note that the set of affine functions constitutes the kernel of the TGV functional.

For specific symmetric functions u, there exist combinations of α0,α1 such that TGVα(u)=α1TV(u). In general one can show that there exists a constant C > 0 such that if α0/α1>C, then the TGV value does not depend on α0 and, up to an affine correction, it is equivalent to TV. In that case the reconstructed images still suffer from a kind of (affine) staircasing effect [Citation37].

The fine structure of TGV reconstructions has been studied analytically mainly in dimension one in [Citation5, Citation38–41]. Under some additional regularity assumptions (compare [Citation6]) it can be shown that for TGV denoising the jump set of the solution is essentially contained in the jump set of the data; see [Citation42] for the TV case.

3.2. The space W0q(div2;Ω)

Next we introduce several function spaces which will be useful in our subsequent development. For this purpose, let 1q and pLq(Ω,Rd). Recall that divpLq(Ω) if there exists wLq(Ω) such that Ωϕ·p dx=Ωϕw dx,for all ϕCc(Ω).

In that case w is unique and we set divp=w. Based on this first-order divergence, we define the Banach space Wq(div;Ω):={pLq(Ω,Rd):divpLq(Ω)}, endowed with the norm pWq(div;Ω)q:=pLq(Ω,Rd)q+divpLq(Ω)q. Similarly one obtains the Banach space Wq(div2;Ω) as the space of all functions pLq(Ω,Sd×d) whose first- and second-order divergences, divp and div2p, respectively, belong to Lq(Ω). We note that for a pLq(Ω,Sd×d), we have that divpLq(Ω,Rd) if there exists an ωLq(Ω,Rd) such that ΩEϕ·p dx=Ωϕ·ω dx,for all ϕCc(Ω,Rd), with Eϕ denoting the L1 function representing the absolutely continuous part of Eϕ, with respect to the Lebesgue measure. Note that since ϕ is smooth, we have Eϕ=Eϕ. As before ω is unique and we set divp=ω. Finally div2pLq(Ω) if there exists a function vLq(Ω) such that Ωϕ·divp dx=Ωϕv dx,for all ϕCc(Ω).

This space is equipped with the norm pWq(div2;Ω)q:=pLq(Ω)q+divpLq(Ω,Rd)q+div2pLq(Ω)q. We refer to [Citation12] for a more general definition of these spaces. Note that when q = 2 these spaces are Hilbertian and then the standard notation is H(div;Ω) and H(div2;Ω); see [Citation43]. The Banach spaces W0q(div;Ω) and W0q(div2;Ω) are defined as W0q(div;Ω)=Cc(Ω,Rd)¯·Wq(div;Ω),W0q(div2;Ω)=Cc(Ω,Sd×d)¯·Wq(div2;Ω), with the analogous notation H0(div;Ω) and H0(div2;Ω) for q = 2. Using the definitions above, the following integration by parts formulae hold true: (2.2) Ωϕ·p dx=Ωϕ divp dx,for all pW0q(div;Ω),ϕC(Ω¯,R),(2.2) (2.3) ΩEϕ·p dx=Ωϕ· divp dx,for all pW0q(div2;Ω),ϕC(Ω¯,Rd),(2.3) (2.4) Ωϕ·divp dx=Ωϕ div2p dx,for all pW0q(div2;Ω),ϕC(Ω¯,R).(2.4)

3.3. Weighted TGV

Throughout the remainder of this work we use the weighted TGV functional (1.2) with α0,α1C(Ω¯) and α0(x),α1(x)>α¯>0, α¯R,xΩ. We denote by |·| the finite dimensional Euclidean norm. We note that for an R-valued finite Radon measure μ=(m1,,m), we denote by |μ| its total variation measure, where for every Borel EΩ |μ|(E)=sup{n=0|μ(En)|:En Borel pairwise disjoint, E=n=0En}.

Note also that it can be shown |μ|(Ω)=sup{i=1Ωϕi dμi:ϕCc(Ω,R),|ϕ(x)|1, for all xΩ}.

We will show that for uL2(Ω) the space Cc(Ω,Sd×d) in (1.2) can be substituted by H0(div2;Ω) (and by W0d(div2;Ω) for uLd/d1(Ω)). This fact will be instrumental when deriving the predual of the TGV minimization problem. For this we need the following result:

Proposition 2.1.

The weighted TGVα2 functional (1.2) admits the equivalent expression (2.5) TGVα2(u)=minwBD(Ω)Ωα1 d|Duw|+Ωα0 d|Ew|.(2.5)

Proof.

The proof is analogous to the one for the scalar TGV functional; see for instance [Citation34, Theorem 3.5] or [Citation12, Proposition 2.8]. Here, we highlight only the significant steps. Indeed, given uL1(Ω), the idea is to define U=C01(Ω,Rd)×C02(Ω;Sd×d),V=C01(Ω,Rd),Λ:UV,Λ(u1,u2)=u1divu2,F1:UR¯,F1(u1,u2)=Ωu divu1+I{|·(x)|α1(x)}(u1)+I{|·(x)|α0(x)}(u2),F2:VR¯,F2(v)=I{0}(v).

Here, IS(·) denotes the indicator function of a set S. Now, after realizing that (2.6) TGVα2(u)=sup(u1,u2)UF1(u1,u2)F2(Λ(u1,u2)),(2.6) the proof proceeds by showing that the dual problem of (2.6) is equivalent to (2.5) and then applying the Fenchel duality result [Citation44]. We note that in order to achieve zero duality gap between the primal and the dual problem the so-called Attouch-Brezis condition needs to be satisfied see [Citation45], that is one needs to show that the set λ0λ(dom(F2)Λ(dom(F1))) is a closed subspace of V. Indeed one can easily see that λ0λ(dom(F2)Λ(dom(F1)))={λ(u1+divu2):(u1,u2)U,|u1(x)|α1(x),|u2(x)|α0(x),xΩ}=V.

Note that for the above equality we crucially used the fact that α1 is bounded away from zero.

The other subtle point is the following density result which is required in order to show that (2.6) is indeed equal to (1.2): (2.7) {ϕCc(Ω,Sd×d):|ϕ(x)|α0(x),|divϕ(x)|α1(x), for all xΩ}¯·C02={ψC02(Ω,Sd×d):|ψ(x)|α0(x),|divψ(x)|α1(x), for all xΩ}.(2.7)

Indeed let ψ belong to the second set in (2.7), and let ϵ>0. Choose 0<λϵ<1 such that (2.8) ψλϵψC02<ϵ/2.(2.8)

Since α0 and α1 are continuous and bounded away from zero there exists αϵ>0, smaller than the minimum of α0,α1, such that |λϵψ(x)|α0(x)αϵ,|divλϵψ(x)|α1(x)αϵ, for all xΩ.

From standard density properties there exists a function ϕϵCc(Ω,Sd×d) such that the following conditions hold for all xΩ: (2.9) ϕϵλϵψC02<ϵ/2,|ϕϵ(x)λϵψ(x)|αϵ/2,|divϕϵ(x)divλϵψ(x)|αϵ/2,(2.9) which implies (2.10) |ϕϵ(x)|α0(x)αϵ/2,|divϕϵ(x)|α1(x)αϵ/2, for all xΩ.(2.10)

Then, from (2.10) it follows that ϕϵ belongs to the first set in (2.7) and from (2.8) and (2.9) we get that ψϕϵC02<ϵ.

Now we are ready to establish the density result needed for dualization. For the sake of the flow of presentation we defer the proof, which parallels the one of [Citation12, Proposition 3.3], to Appendix A. Below “a.e.” stands for “almost every” with respect to the Lebesgue measure.

Proposition 2.2.

For uL2(Ω), the weighted TGV functional (1.2) can be equivalently written as (2.11) TGVα2(u)=sup{Ωu div2p dx:pH0(div2;Ω),|p(x)|α0(x),|divp(x)|α1(x),fora.e.xΩ}.(2.11)

Remark:

By slightly amending the proof of Proposition 2.2, one can also show that if uLd/d1(Ω)BV(Ω) then the dual variables p can be taken to belong to W0d(div2;Ω) rather than H0(div2;Ω).

3.4. The predual weighted TGV problem

Now we study the predual problem for the weighted TGV model with continuous weights, i.e., we use the regularization functional (1.2) or equivalently (2.11). For TL(L2(Ω),L2(Ω)) we assume that B:=TT is invertible and define vB2=ΩvB1v, which induces a norm in L2(Ω) equivalent to the standard one; compare [Citation31]. When TT is not invertible one can add an extra regularization term of the form κ2Ωu2 dx, for some small κ>0, to the objective in (1.1). Then B=κId+TT will be invertible, see a similar approach in [Citation27, Citation29, Citation31]. We mention that a different set-up, as adopted for instance in [Citation46], assumes TL(Ld/d1,L2(Ω)), since BV(Ω) continuously embeds to Ld/d1(Ω). Then one readily finds that for d > 2, B cannot be invertible, in general. Even though invertibility of TT is not necessary when dealing only with the TGV regularization problem (1.1), additional challenges would arise in our overall bilevel optimization context, such as, e.g., a lack of uniqueness of the solution for the lower level problem, i.e., genuine set-valuedness of the regularization-weight-to-reconstruction map cf. the constraint set in (1.4). In the next proposition we prove zero duality gap for (1.1) and its predual. We note however that the invertibility assumption on TT is not needed to establish zero duality gap but rather to get an explicit formulation of the predual problem. An adaptation of the following proposition without this invertibility assumption can be done for instance by simply adjusting the corresponding proof of [Citation46, Theorem 5.4] where analogous dualization results for a class of structural TV regularizers were considered.

Proposition 2.3.

Let fL2(Ω), then there exists a solution to the primal problem (2.12) minimize12TufL2(Ω)2+TGVα2(u)over uBV(Ω),(2.12) as well as to its predual problem (2.13) minimize12Tfdiv2pB212fL22over pH0(div2;Ω)subject to|p(x)|α0(x),|divp(x)|α1(x),fora.e.xΩ,(2.13) and there is no duality gap, i.e., the primal and predual optimal objective values are equal. Moreover, the solutions u and p of these problems satisfy (2.14) Bu=Tfdiv2p.(2.14)

Proof.

We set U=H0(div2;Ω),V=L2(Ω), Λ:UV with Λp=div2p, and also F1:UR¯ and F2:VR¯ with (2.15) F1(p)=I{|·(x)|α0(x),fora.e.x}(p)+I{|div·(x)|α1(x),fora.e.x}(p),(2.15) (2.16) F2(ψ)=12TfψB212fL2(Ω)2.(2.16)

Immediately one gets that (2.17) infpUF1(p)+F2(Λp)=minpH0(div2;Ω)|p(x)|α0(x)|divp(x)|α1(x)12Tfdiv2pB212fL2(Ω)2.(2.17)

The problem in (2.17) admits a solution. Indeed, first observe that the objective is bounded from below. Then note that since 12T·fL2(Ω)2 is continuous at 0L2(Ω), its convex conjugate (see [Citation44] for a general definition) which is equal to 12Tf+·B212fL2(Ω)2 is coercive in L2(Ω); see [Citation47, Theorem 4.4.10]. Hence, any infimizing sequence (pn)nN is bounded in H0(div2;Ω), and thus there exist an (unrelabeled) subsequence and pH(div2;Ω) such that pnp,divpndivp and div2pndiv2p weakly in L2. We also have that p is a feasible point since the set {(h,divh,div2h):hH0(div2;Ω),|h(x)|α0(x),|divh(x)|α1(x),fora.e.xΩ}, is weakly closed. Then p is a minimizer of (2.17) as 12Tf·B2 is weakly lower semicontinuous in L2(Ω).

We now calculate the expression F1(Λu)+F2(u) for uV=L2(Ω). As before one verifies by direct computation that F2(u)=12TufL2(Ω)2. Moreover, F1(Λu)=suppU{Λu,pU,UF1(p)}=suppU{u,ΛpL2(Ω),L2(Ω)F1(p)}=sup|p(x)|α0(x)|divp(x)|α1(x)pH0(div2;Ω)Ωdiv2p dx=TGVα2(u), where for the last equality we used Proposition 2.2 and its underlying density result. In order to prove that there is no duality gap, it suffices again to verify the Attouch-Brezis condition and show that the set λ0λ(dom(F2)Λ(dom(F1))) is a closed subspace of V. It is immediate to see that dom(F2)=L2(Ω), and hence the condition holds true. Thus, we also get existence of a solution for the primal problem (2.12). Finally (2.14) follows from the optimality condition (Euler-Lagrange system) that corresponds to ΛpF2(u).

The primal-dual optimality conditions for the problems (2.12) and (2.13) read (2.18) pF1(Λu),(2.18) (2.19) ΛpF2(u),(2.19) and we note once again that (2.14) corresponds to (2.19) with F2 and Λ as in the proof of Proposition 2.12. Instead of making the optimality condition that corresponds to (2.18) explicit, we are interested in the analogous optimality conditions written in the variables u and w of the equivalent primal weighted TGV problem (2.20) minwBD(Ω)uBV(Ω)12TufL2(Ω)2+Ωα1 d|Duw|+Ωα0 d|Ew|.(2.20)

For this purpose note first that the predual problem (2.13) can be equivalently written as (2.21) {minimize12Tf+divqB212fL2(Ω)2 over (p,q)H0(div2;Ω)×H0(div;Ω),subject todivp=q, |p(x)|α0(x), |q(x)|α1(x), fora.e.xΩ.(2.21)

The next proposition characterizes the solutions (w, u) and (p, q) of the problems (2.20), and (2.21) respectively. Before we state it, we note that in its proof we will make use of the following density results: (2.22) Cα0¯L2(Ω)=Kα0,Cα1¯H0(div;Ω)=Kα1,(2.22) where (2.23) Cα0:={(div2ϕ,divϕ):ϕCc(Ω,Sd×d),|ϕ(x)|α0(x), for all xΩ},(2.23) (2.24) Kα0:={(div2p,divp):pH0(div2;Ω),|p(x)|α0(x),for a.e.xΩ},(2.24) (2.25) Cα1:={ψ:ψCc(Ω,Rd),|ψ(x)|α1(x), for all xΩ},(2.25) (2.26) Kα1:={q:qH0(div;Ω),|q(x)|α1(x),for a.e.xΩ}.(2.26)

These results can be proven by using the duality arguments of the proof of Proposition 2.2, which originate from [Citation12], or with the use of mollification techniques; see [Citation48–50].

Proposition 2.4.

The pair (p,q)H0(div2;Ω)×H0(div;Ω) is a solution to (2.21), and (w,u)BD(Ω)×BV(Ω) is a solution to (2.20) if and only if the following optimality conditions are satisfied: (2.27) Bu=Tf+divq,(2.27) (2.28) q=divp,(2.28) (2.29) |q(x)|α1(x)for a.e.xΩ(2.29) (2.30) and Duw,q˜q0 for every q˜H0(div;Ω) with |q˜(x)|α1(x)for a.e.xΩ,|p(x)|α0(x)for a.e.xΩ(2.30)  and Ew,p˜p0 for every p˜H0(div2;Ω) with |p˜(x)|α0(x)fora.e.xΩ.

Proof.

Define X=(X1,X2)=H0(div2,Ω)×H0(div,Ω), Y=(Y1,Y2)=H0(div;Ω)×L2(Ω),Λ:XY with Λ(p,q)=(q+divp,divq), and F1:XR¯,F2:YR¯ with (2.31) F1(p,q)=I{|·(x)|α0(x),for a.e.x}(p)+I{|·(x)|α1(x),for a.e.x}(q),(2.31) (2.32) F2(ϕ,ψ)=I{0}(ϕ)+12Tf+ψB212fL2(Ω)2.(2.32)

One checks immediately that min(p,q)XF1(p,q)+F2(Λ(p,q)) corresponds to (2.21) with the dual problem reading min(w,u)YF1(Λ(w,u))+F2(w,u). Observe that since Λ(w,u),(p,q)X,X=(w,u),Λ(p,q)Y,Y=w,divpY1,Y1w,qY1,Y1u,divqY2,Y2, we have (2.33) F1(Λ(w,u))=sup|p(x)|α0(x)pH0(div2;Ω)w,divpY1,Y1+sup|q(x)|α1(x)qH0(div;Ω)w,qY1,Y1u,divqY2,Y2.(2.33)

Note that the suprema above are always greater or equal to the corresponding suprema over Cc(Ω,Sd×d)H0(div2;Ω) and Cc(Ω,Rd)H0(div;Ω). One can easily check that there is no duality gap between these primal and dual problems – the Attouch-Brezis condition is satisfied – and thus since the primal problem is finite so is the dual. Using the fact that α0 is bounded from below, this implies in particular that w, seen as distribution through its action on Cc(Ω,Rd), has a distributional symmetrized gradient Ew with bounded Radon norm, and hence it is a Radon measure. It follows that wL1(Ω,Rd) yielding wBD(Ω); see [Citation34], which means that for ψCc(Ω,Rd) we have w,ψY1,Y1=Ωw·ψ dx. Using density of Cc(Ω,Rd) in H0(div;Ω) this also implies w,qY1,Y1=Ωw·q dx and similarly w,divpY1,Y1=Ωw·divp dx in (2.33). Using now also the density results (2.22) we have F1(Λ(w,u))=sup|p(x)|α0(x)pH0(div2;Ω)Ωw·divp dx+sup|q(x)|α1(x)qH0(div;Ω)Ωw·q dxΩu divq dx=sup|ϕ(x)|α0(x)ϕCc(Ω,Sd×d)Ωw·divϕ dx+sup|ψ(x)|α1(x)ψCc(Ω,Rd)Ωw·ψ dxΩu divψ dx=sup|ϕ(x)|α0(x)ϕCc(Ω,Sd×d)Ew,ϕ+sup|ψ(x)|α1(x)ψCc(Ω,Rd)Duw,ψ,=Ωα0 d|Ew|+Ωα1 d|Duw|.

Here we used the fact that since the distribution Duw has a finite Radon norm, due to α1 being bounded away from zero, it can be represented by an Rd-valued finite Radon measure and in particular uBV(Ω). Furthermore, as in the proof of Proposition 2.3 we have F2(w,u)=12TufL2(Ω)2.

The fact that there is no duality gap is ensured by Propositions 2.1, 2.2 and 2.3. We now turn our attention to the optimality conditions (2.34) (p,q)F1(Λ(w,u)),(2.34) (2.35) Λ(p,q)F2((w,u)).(2.35)

It can be checked again that (2.35) gives (2.27) and (2.28). We now expand on (2.34). We have that (p,q)F1(Λ(w,u)) which is equivalent to Λ(w,u)F1(p,q), that is F1(p,q)=0 and Λ(w,u),(p˜p,q˜q)X,XF1(p˜,q˜)w,div(p˜p)w,q˜qu,divq˜divqF1(p˜,q˜)Ew,p˜pI{|·(x)|α0(x),f.a.e.x}(p˜)Duw,q˜qI{|·(x)|α1(x),f.a.e.x}(q˜)Ew,p˜p0Duw,q˜q0, with the last two inequalities holding for any p˜H0(div2;Ω) with |p˜(x)|α0(x) for a.e. xΩ and for any q˜H0(div;Ω) with |q˜(x)|α1(x) for a.e. xΩ. Hence we obtain (2.29) and (2.30). □

4. A Series of regularized problems

4.1. Regularization of the primal problem

With the aim of lifting the regularity of u and w to avoid measure-valued derivatives, we next consider the following regularized version of the primal weighted TGV problem (2.20): (3.1) minimize12TufL2(Ω)2+Ωα1|uw|dx+Ωα0|Ew|dx+μ2uL2(Ω)2+ν2wH1(Ω,Rd)2over (w,u)H1(Ω,Rd)×H1(Ω),(3.1) for some constants 0<μ,ν1. Existence of solutions for (3.1) follows from standard arguments.

Observe that (3.1) is equivalent to min(w,u)X̂Q1(w,u)+Q2(R(w,u)) where X̂=H1(Ω,Rd)×H1(Ω), Ŷ=L2(Ω,Sd×d)×L2(Ω,Rd),R:X̂Ŷ with R(w,u)=(Ew,uw),Q1:X̂R,Q2:ŶR with Q(w,u)=12TufL2(Ω)2+μ2uL2(Ω,Rd)2+ν2wH1(Ω,Rd)2 and Q2(ψ,ϕ)=Ωα1|ϕ|dx+Ωα0|ψ|dx. Note that the Attouch-Brezis condition is satisfied since dom(Q2)=Y.

Proposition 3.1.

The pairs (w,u)H1(Ω,Rd)×H1(Ω) and (p,q)L2(Ω,Rd×d)×L2(Ω,Rd) are solutions to (3.1) and its predual problem, respectively, if and only if the following optimality conditions are satisfied: (3.2) BuμΔu+qTf=0 in H1(Ω),(3.2) (3.3) νwνΔwq+Ep=0 in H1(Ω,Rd),(3.3) (3.4) {|q|α1,α1(uw)q|uw|=0 if|q(x)|=α1(x),uw=0 if|q(x)|<α1(x),(3.4) (3.5) {|p|α0,α0Ewp|Ew|=0 if|p(x)|=α0(x),Ew=0 if|p(x)|<α0(x),(3.5) where the multiplications are regarded component-wise.

Proof.

The proof follows again easily by calculating the corresponding primal-dual optimality conditions. □

Next we study the relationship between the solutions of (2.20) and (3.1) as the parameters μ, ν tend to zero.

Proposition 3.2.

Let μn,νn0 and let (wn,un)nN be a sequence of solution pairs of the problem (3.1). Then unu and wnw in BV(Ω) and BD(Ω) respectively, where (w,u) is a solution pair for (2.20). The convergence of wn is up to a subsequence.

Proof.

For convenience of notation, define the energies En(w,u)=12TufL2(Ω)2+Ωα1|uw|dx+Ωα0|Ew|dx+μn2uL2(Ω)2+νn2wH1(Ω,Rd)2,E(w,u)=12TufL2(Ω)2+Ωα1d|Duw|+Ωα0d|Ew|.

We have (3.6) 12TunfL2(Ω)2+Ωα1|unwn|dx+Ωα0|Ewn|dxEn(wn,un)En(0,0)12fL2(Ω)2.(3.6)

Thus, the sequences (un)nN and (wn)nN are bounded in BV(Ω) and BD(Ω), respectively. In order to see this, note that by setting α¯i:=minxΩ¯αi(x), i = 0, 1, we get TGVα¯0,α¯12(un)=minwBD(Ω)α¯1unwM+α¯0EwMΩα1|unwn|dx+Ωα0|Ewn|dx12fL2(Ω)2.

Hence, (un)nN is bounded in the sense of second-order TGV. Again using [Citation47, Theorem 4.4.10], we get that 12T·fL2(Ω)2 is coercive, since it is the convex conjugate of 12Tf+·B212fL22 which is continuous at 0L2(Ω). This implies further that this sequence is bounded both on L2(Ω) and BV(Ω). The bound on (wn)nN in BD(Ω) then follows from (3.6).

From compactness theorems in those spaces (for BD(Ω) see for instance [Citation36, Remark 2.4]) we have that there exist uBV(Ω)L2(Ω) and wBD(Ω) such that unku in BV(Ω) and weakly in L2(Ω), and wnkw in BD(Ω) along suitable subsequences. Due to the lower semicontinuity of the functional E with respect to these convergences, we have for any pair (w˜,u˜)H1(Ω,Rd)×H1(Ω) (3.7) E(w,u)liminfkE(wnk,unk)liminfkEnk(wnk,unk)liminfkEnk(w˜,u˜)=E(w˜,u˜).(3.7)

Recall now that LD(Ω)={wL1(Ω,Rd):EwL1(Ω,Rd×d)} is a Banach space endowed with the norm wLD(Ω)=wL1(Ω,Rd)+EwL1(Ω,Rd×d) and that C(Ω¯,Rd) is dense in that space; see [Citation35, Proposition 1.3]. From this, in combination with the fact that C(Ω¯) is dense in W1,1(Ω)L2(Ω) we have that for every (ŵ,û)LD(Ω)×(W1,1(Ω)L2(Ω)) there exists a sequence (ŵh,ûh)hNC(Ω¯,Rd)×C(Ω¯)H1(Ω,Rd)×H1(Ω), such that E(ŵh,ûh)E(ŵ,û). Hence, since (3.7) holds we have that (3.8) E(w,u)E(ŵ,û), for all (ŵ,û)LD(Ω)×(W1,1(Ω)L2(Ω)).(3.8)

Finally, by following similar steps as in the proof of [Citation6, Theorem 3], we can show that for every (w,u)BD(Ω)×(BV(Ω)L2(Ω)) there exists a sequence (wh,uh)hNLD(Ω)×(W1,1(Ω)L2(Ω)) such that uhuL2(Ω)0,Ωα1|uhwh|dxΩα1d|Duw|,Ωα0|Ewh|dxΩα0d|Ew|, which implies again that E(wh,uh)E(w,u). This, together with (3.8) yields E(w,u)E(w,u), for all (w,u)BD(Ω)×BV(Ω).

This yields that (w,u) is a solution pair for (2.20). Finally, from the uniqueness of the solution u for (2.20) we get that the whole initial sequence (un)nN converges to u weakly in BV(Ω). The uniqueness follows from the fact that T is injective and hence the functional 12T·fL2(Ω)2 is strictly convex. □

We now proceed to the second level of regularization of the problem (3.1), which, in addition to lifting the regularity of u and w, respectively, also smoothes the non-differentiable constituents. For this purpose, we define the following primal problem, which will also be treated numerically, below: (pγ)minimize12TufL2(Ω)2+Ωα1φγ(uw)dx+Ωα0φγ(Ew)dx+μ2uL2(Ω)2+ν2wH1(Ω,Rd)2 over (w,u)H1(Ω,Rd)×H1(Ω).

Here φγ denotes the Huber-regularized version of the Euclidean norm. That is, for a vector vS,S=Rd or Rd×d and γ>0 we use (3.9) φγ(v)(x)={|v(x)|12γ if|v(x)|γ,12γ|v(x)|2 if|v(x)|<γ,(3.9) with |·| denoting either the Euclidean norm in Rd or the Frobenius norm in Rd×d. We mention that this type of Huber regularization of TV-type terms in the primal problem corresponds to an L2 regularization of the dual variables in the predual [Citation32, Citation51]. In order to illustrate this consider the following denoising problem Pγ without any H1 regularization: (3.10) minimize12ufL2(Ω)2+Ωα1dφγ1(Duw)+Ωα0dφγ2(Ew)over (w,u)BD(Ω)×BV(Ω),(3.10) where Ωα1dφγ1(Duw)=Ωα1φγ1(uw)dx+Ωα1d|Dsu|,Ωα0dφγ1(Ew)=Ωα0φγ2(Ew)dx+Ωα0d|Esw|.

Here Dsu and Esw denote the singular parts with respect to the Lebesgue measure of Du, and Ew respectively, following the Lebesgue decomposition Du=Ωu dx+Dsu and Ew=ΩEw dx+Esw. The corresponding predual problem of (3.10) is given by (3.11) maximize12f+divqL2(Ω)2γ02Ω1α0|p|2dxγ12Ω1α1|q|2dx+12fL2(Ω)2,over (p,q)W0d(div2;Ω)×W0d(div;Ω),subject to q=divp, |p(x)|α0(x), |q(x)|α1(x).(3.11) The proof is similar to the one of Proposition 2.4 with F1(p,q)=I{|·(x)|α0(x)}(p)+I{|·(x)|α1(x)}(q)γ02Ω1α0|p|2dxγ12Ω1α1|q|2dx, and in the dualization process we use the fact that for an S-valued measure μ we have, Ωαdφγ(μ)=sup{Ωϕ dμI{|·(x)|α(x)}(ϕ)γ2Ω1α|ϕ|2dx:ϕCc(Ω,S)}; see for instance [Citation52].

Returning to the (doubly) regularized primal problem Pγ, we are primarily interested in its associated first-order optimality conditions.

Proposition 3.3.

We have that the pairs (w,u)H1(Ω,Rd)×H1(Ω) and (p,q)L2(Ω,Rd×d)×L2(Ω,Rd) are solution to Pγ and its predual problem, respectively, if and only if the following optimality conditions are satisfied: (3.12) BuμΔu+qTf=0 in H1(Ω),(3.12) (3.13) νwνΔwq+Ep=0 in H1(Ω,Rd),(3.13) (3.14) max(|uw|,γ1)qα1(uw)=0 in L2(Ω,Rd),(3.14) (3.15) max(|Ew|,γ0)pα0Ew=0 in L2(Ω,Sd×d).(3.15)

(3.16)

The proof of Proposition 3.3 follows from calculating the corresponding primal-dual optimality conditions as in Proposition 3.1. Finally, in order to avoid constraint degeneracy and for the sake of differentiability for the bilevel scheme in the next section, we also employ here a smoothed version maxδ(·,γ) of max and its derivative, denoted by Xδ, defined as follows for r0 and for 0<δ2<γ: maxδ(r,γ)={γ12δ(r+δ2γ)2+γ,rXδ(r,γ)={0 if rγδ2,1δ(r+δ2γ) if γδ2<r<γ+δ2,1 if r>γ+δ2.

Thus we arrive at the following regularized optimality conditions: (Opt1) BuμΔu+qTf=0,(Opt1) (Opt2) νwνΔwq+Ep=0,(Opt2) (Opt3) maxδ(|uw|,γ1)qα1(uw)=0,(Opt3) (Opt4) maxδ(|Ew|,γ0)pα0Ew=0.(Opt4) Note that these would correspond to the optimality conditions of an analogue problem to Pγ where the Huber function φγ is substituted by an analogous C2 version φγ,δ, having the property φγ,δ(x)=x/maxδ(|x|,γ) pointwise, and which do not write down explicitly here. The relevant approximation result now follows, where we have set γ0=γ1=γ for simplicity.

Proposition 3.4.

Let (w,u,q,p) and (wγ,δ,uγ,δ,pγ,δ,qγ,δ) satisfy the optimality conditions (3.2)–(3.5) and (Opt1)(Opt4), respectively. Then, as γ,δ0, we have uγ,δu strongly in H1(Ω),wγ,δw strongly in H1(Ω,Rd) as well as divqγ,δdivq and qγ,δ+divpγ,δq+divp weakly in H1(Ω) and H1(Ω,Rd), respectively.

Proof.

By subtracting first two equations of the optimality system of Proposition 3.1 and 3.3, respectively, we get for all vH1(Ω),ωH1(Ω,Rd) (3.17) ΩB(uuγ,δ)v dx+μΩ(uuγ,δ)v dx=Ω(qγ,δq)v dx,(3.17) (3.18) νΩ(wwγ,δ)ω dx+νΩ(wwγ,δ)ω dx=Ω(qqγ,δ)ω dx+Ω(pγ,δp)Eω dx.(3.18)

When using v=uuγ,δ and ω=wwγ,δ in the equations above and adding them up we get (3.19) uuγ,δB12+μuuγ,δL2(Ω,Rd)2+νwwγ,δH1(Ω,Rd)2=R1+R2,(3.19) where R1:=Ω(qγ,δq)[uw(uγ,δwγ,δ)] dx,R2:=Ω(pγ,δp)E(wwγ,δ) dx.

We now estimate R1 and R2. Consider the partitions of Ω into disjoint sets (up to sets of measure zero) Ω=AI=Aγ,δIγ,δ, where A={xΩ:|uw|>0},I=ΩA,Aγ,δ={xΩ:|uγ,δwγ,δ|>γ+δ2},Iγ,δ=ΩAγ,δ.

We estimate R1 separately on the disjoint sets Aγ,δA, Aγ,δI, Iγ,δA and Iγ,δI. Recall that |q(x)|α1(x), as well as |qγ,δ(x)|α1(x) for almost every xΩ. Starting from Aγ,δA and noticing that q=α1uw|uw|,qγ,δ=α1uγ,δwγ,δ|uγ,δwγ,δ|, it follows that pointwise on Aγ,δA (with argument x left off for ease of notation) we have (qγ,δq)[uw(uγ,δwγ,δ)]=qγ,δ(uw)α1|uγ,δwγ,δ|α1|uw|+q(uγ,δwγ,δ)α1|uw|α1|uγ,δwγ,δ|α1|uw|+α1|uγ,δwγ,δ|=0.

Turning now to the set Aγ,δI and recalling uw=0 we have (qγ,δq)[uw(uγ,δwγ,δ)]α1|uγ,δwγ,δ|+|q||uγ,δwγ,δ|0.

For the set Iγ,δA, note that |uγ,δwγ,δ|γ+δ2,uγ,δwγ,δ=γα1qγ,δ.

Thus, we can estimate (qγ,δq)[uw(uγ,δwγ,δ)]qγ,δ(uw)α1|uw|qγ,δ(uγ,δwγ,δ)+q(uγ,δwγ,δ)α1|uw|α1|uw|qγ,δ(uγ,δwγ,δ)+q(uγ,δwγ,δ)(2γ+δ)α1

Similarly, for the set Iγ,δI we get (qγ,δq)[uw(uγ,δwγ,δ)]2(γ+δ)α1.

Combining the above estimates we have R1Ω(2γ+δ)α1 dx0 and for R2 we similarly get R2Ω(2γ+δ)α0 dx0.

Hence, from (3.19) and the fact that ·B1 is equivalent to ·L2(Ω), we obtain the desired convergences for uγ,δ and wγ,δ. From this result and using (3.17) and (3.18) we get that for every vH1(Ω) and for every ωH1(Ω,Rd) we have Ωvdivqγ,δ dxΩvdivq dx and Ωω(qγ,δ+divpγ,δ) dxΩω(q+divp) dx, as γ,δ0. This completes the proof. □

Finally, the following approximation result holds true, when all the parameters μ, ν, γ, δ tend to zero.

Proposition 3.5.

Let μn,νn,γn,δn0, and denote by uμn,νn,γn,δnH1(Ω) the solution of (Opt1)–(Opt1) with (μ,ν,γ,δ)=(μn,νn,γn,δn). Then uμn,νn,γn,δnu in BV(Ω), where u solves (2.20).

Proof.

It is easy to show that uμn,νn,γn,δnu in L1(Ω). Indeed, we have uμn,νn,γn,δnuL1(Ω)uμn,νn,0,0uL1(Ω)+uμn,νn,γn,δnuμn,νn,0,0L1(Ω).

According to Proposition 3.2 it holds that uμn,νn,0,0uL1(Ω)0. The other term tends to zero according to Equationequation (3.19) of Proposition 3.4. There, the estimates for R1, R2 are not affected if we substitute u and uγ,δ by uμn,νn,0,0 and uμn,νn,γn,δn, respectively. In other words, the estimate uμn,νn,0,0uμn,νn,γn,δnL2(Ω)2C(2γn+δn)|Ω|α0+α1 holds for some C > 0 and hence uμn,νn,γn,δnuμn,νn,0,0L1(Ω)0.

To finish the proof and show that the convergence is weak in BV(Ω), it suffices to establish that Ω|uμn,νn,γn,δn dx| is uniformly bounded in n; see [Citation33, Proposition 3.13]. Observe first that, if φγ,δ is the C2 regularized Huber function, that corresponds to the δ-smoothing of max, then as in the proof of Proposition 3.2 we get (3.20) Ωα1φγ,δ(uμn,νn,γn,δnwμn,νn,γn,δn)dx+Ωα0φγ,δ(Ewμn,νn,γn,δn)dx12fL2(Ω)2.(3.20)

As in (3.9) we also have that φγ,δ(·)|·|12γ, and hence we obtain (3.21) Ωα1|uμn,νn,γn,δnwμn,νn,γn,δn|dx+Ωα0|Ewμn,νn,γn,δn|dx12fL2(Ω)2+(α1+α0)|Ω|γn2K,(3.21) for some constant K > 0. Then, as in the proof of Proposition 3.2, we get that (uμn,νn,γn,δn)nN is bounded in TGV which, together with the L1 bound, gives the desired bound in TV. □

5. The bilevel optimization scheme

In this section we will adapt the bilevel optimization framework developed in [Citation27, Citation29] in order to automatically select the regularization functions α0 and α1. The main idea is to minimize a suitable upper level objective over both the image u and the regularization parameters α0, α1 subject to u being a solution to a (regularized) TGV-based reconstruction problem with these regularization weights.

It is useful to recall the definitions of the localized residual R and the function F as stated in the introduction: (4.1) Ru(x)=Ωw(x,y)(Tuf)2(y)dy,(4.1) where wL(Ω×Ω) with ΩΩw(x,y) dxdy=1 and (4.2) F(v):=12Ωmax(vσ¯2,0)2dx+12Ωmin(vσ¯2,0)2dx,(4.2) for some appropriately chosen σ¯2,σ¯2.

We thus end up to the following bilevel minimization problem: (4.3) {minJ(u,α0,α1):=F(R(u))+λ02α0H1(Ω)2+λ12α1H1(Ω)2,over (u,α0,α1)H1(Ω)×Aad0×Aad1, subject to (u,w)=argmin(u˜,w˜)12Tu˜fL2(Ω)2+Ωα1φγ,δ(u˜w˜)dx+Ωα0φγ,δ(Ew˜)dx+μ2u˜L2(Ω)2+ν2w˜H1(Ω,Rd)2.(4.3)

Utilizing the regularized primal-dual first-order optimality characterization (Opt1)–(Opt4) of the solution to the lower level problem of (4.3), we arrive at the following mathematical program with equilibrium constraints (MPEC, for short) which is equivalent to (4.3): (PTGV){minJ(u,α0,α1):=F(R(u))+λ02α0H1(Ω)2+λ12α1H1(Ω)2,over (u,α0,α1)H1(Ω)×Aad0×Aad1, subject to       BuμΔu+qTf=0,νwνΔwq+Ep=0,maxδ(|uw|,γ1)qα1(uw)=0,maxδ(|Ew|,γ0)pα0Ew=0.

Note that in view of the equivalence of (4.3) and (PTGV), we will still refer the latter as the bilevel TGV problem. A few words about (PTGV) are in order. Here, αi are forced to be contained in the box constraint sets (4.4) Aadi:={αiH1(Ω):α¯iαiα¯i},i=0,1,(4.4) with α¯i,α¯iL2(Ω) and 0<ϵ¯α¯i(x)<α¯i(x)ϵ¯ in Ω for some ϵ¯,ϵ¯>0, i = 0, 1. Note that the H1 regularity on the parameter functions α0,α1 facilitates the existence and differential sensitivity analysis as established in [Citation27, Citation29] for the TV case. Note, however, that this setting does not guarantee a priori that these functions belong to C(Ω¯), the regularity required for applying the dualization results of the previous sections. Nevertheless, under mild data assumptions, one can make use of the following regularity result of the H1–projection onto the sets Aad0 and Aad1; see [Citation29, Corollary 2.3] for a proof.

Proposition 4.1.

Let ΩR with =1,2,3 be a bounded convex set and let Aad:={αH1(Ω):α¯αα¯}, where α¯,α¯H2(Ω) such that α¯α¯ and α¯ν=α¯ν=0 in H1/2(Ω). Then if ω=PAad(ω):=argminαA12αωH1(Ω)2, it holds ωH2(Ω) and ων=0ωH2(Ω) and ων=0.

In particular, if α¯0,α¯0, α¯1,α¯1 as well as the initializations for α1 and α0 are constant functions, then along the projected gradient iterations of Algorithm Citation2, the weights are guaranteed to belong to H2(Ω) which (for dimension d2) embeds into C(Ω¯).

We briefly note that in the TV case it can be shown [Citation30, Citation46] that W1,1 regularity for the regularization parameter α suffices to establish a dualization framework. A corresponding result is not yet known for TGV, even though one expects that it could be shown by similar arguments. Hence, here we will also make use of the H1–projection regularity result as described above.

Regarding the box constraints (4.4) in [Citation25] it was shown that for a PSNR-optimizing upper level objective J˜(u,α)=u(α)fL2(Ω)2 subject to H1 and Huber regularized TV and TGV denoising problems, under some mild conditions on the data f, the optimal scalar solutions α and (α0,α1) are strictly positive. As depicted in the upper level objective discussed here appears close to optimizing the PSNR, keeping the parameters strictly positive via (4.4) seems, however, necessary for the time being.

We now briefly discuss how to treat the bilevel problem (PTGV). Let (α0,α1)u(α0,α1) denote the solution map for the lower level problem, equivalently of the optimality conditions (Opt1)–(Opt4). Then the problem (PTGV) admits the following reduced version (4.5) 4minĴ(α0,α1):=J(u(α0,α1),α0,α1)over α0Aad0,α1Aad1.(4.5)

Similarly to the TV case [Citation27], one can show that the reduced functional Ĵ:H1(Ω)×H1(Ω)R is differentiable. We can then apply the KKT framework in Banach spaces [Citation53]: (4.6) {minimizeJ(x)over xX,subject toxC and g(x)=0,(4.6) where V,A,Z are Banach spaces, X=V×A, J:XR and g:XZ are Fréchet differentiable and continuous differentiable functions, respectively, and CX is a non-empty, closed convex set. In the bilevel TGV problem (PTGV) we have V=H1(Ω)×H1(Ω,Rd)×L2(Ω,Rd),L2(Ω,Sd×d), A=H1(Ω)×H1(Ω), C=V×Aad0×Aad1, and Z=H1(Ω)×H1(Ω,Rd)×L2(Ω,Rd),L2(Ω,Sd×d). Here g:XZ is defined by the optimality conditions (Opt1)–(Opt4). Finally, for x=(u,w,q,p,α0,α1):=(x,α0,α1), we have J(x)=J(u,α0,α1). Note that the framework of (4.6) guarantees the existence of an adjoint variable xV with the help of which an optimal solution of (PTGV) can be characterized. This adjoint variable also allows the computation of the derivative of the reduced objective Ĵ in an amenable way, see next section.

We will skip here the proofs for the differentiability of the functions g and the reduced objective J as well as the existence proofs for (PTGV). These results can be shown similarly to the corresponding assertions for TV; see [Citation27, Theorem. 6.1, Proposition 6.2, Proposition 6.3].

6. Numerical implementation

In this section we will describe a Newton method for the lower level problem, a projected gradient algorithm for the solution of the discretized version of the bilevel problem (PTGV), as well as provide corresponding numerical examples in denoising.

6.1. Newton solver for the lower level problem

Before we proceed to devising of a projected gradient algorithm for the solution of the bilevel problem, we discuss first a primal-dual Newton algorithm for the solution of the first-order optimality conditions (Opt1)– (Opt4) re-written here for the sake of readability: (5.1) BuμΔudivqf=0,(5.1) (5.2) νwνΔwqdivp=0,(5.2) (5.3) maxδ(|uw|,γ1)qα1(uw)=0,(5.3) (5.4) maxδ(|Ew|,γ0)pα0Ew=0.(5.4)

A few words on the discrete involved quantities are in order. Images (d = 2) are considered as elements of Uh:={u | u:ΩhR} where Ωh={1,2,,n}×{1,2,,m} is a discrete cartesian grid that corresponds to the image pixels. The mesh size, defined as the distance between the grid points, is set to h=1/nm. We define the associated discrete function spaces Wh=Uh×Uh,Vh=Uh×Uh×Uh, so that pVh with p=(p11,p12,p22). For the discrete gradient and divergence we have, :WhVh and div:VhWh satisfying the adjoint relation =div, setting zero values at the ghost points. The discretized symmetrized gradient Ew is defined as 12(w+(w)). For the discretized versions of the Laplacian, we use the standard five-point stencils with zero Neumann boundary conditions, by setting the function values of ghost grid points to be the same with the function value of the nearest grid point in Ωh. Note that these act on the primal variables u and w, which satisfy natural boundary conditions in contrast to the dual variable.

The system of Equationequations (5.1)–(5.4) can be shortly written as gpd(x)=0, where x=(u,w,q,p). We compute the derivative of gpd at a point x=(u,w,q,p) as the following block-matrix: Dgpd(x)=Dgpd(u,w,q,p)=[ABCD], where (5.5) A=[BμΔ00ν(IΔ)],B=[div0Idiv],D=[maxδ(|uw|,γ1)00maxδ(|Ew|,γ0)],(5.5) (5.6) C=[α1+qXδ(|uw|,γ1)uw|uw|·α1I+qXδ(|uw|,γ1)uw|uw|·(I)0α0E+pXδ(|Ew|,γ0)Ew|Ew|·E].(5.6)

Given xk, the Newton iteration for solving the system of Equationequations (5.1)–(5.4), or gpd(x)=0 for short, reads xk+1=xkDF(xk)1F(xk), which can also be written as (5.7) Dgpd(xk)xk+1=Dgpd(xk)xkgpd(xk).(5.7)

Here it is convenient to introduce the notation Dgpd(xk)=Dgpd(uk,wk,qk,pk)=[ABCkDk] since only the submatrices C and D depend on k. Note that the right-hand side Dgpd(xk)xkgpd(xk) of the linear system (5.7) can be written as Dgpd(xk)xkF(xk)=(b1kb2k), where b1k=(f,0), andb2k=(qkXδ(|ukwk|,γ1)|ukwk|,pkXδ(|Ewk|,γ0)|Ewk|).

Notation-wise, the components that appear in b2k should be regarded as the diagonals of the corresponding diagonal matrices that we mentioned before, multiplied component-wise. By introducing the notation x1k=(uk,wk),x2k=(qk,pk), the Newton system (5.7) can be written as (5.8) [ABCkDk](x1k+1x2k+1)=(b1kb2k).(5.8)

The above system can be simplified utilizing the Schur complement: First solve for the primal variables x1k+1=(uk+1,wk+1) and then recover the dual ones x2k+1=(qk+1,pk+1). This yields (ABDk1Ck)x1k+1=b1kBDk1b2k,x2k+1=Dk1(b2kCkx1k+1).

The folllowing result then holds.

Lemma 5.1.

If (qk,pk) belong to the feasible set, i.e., |qk|α1 and |pk|α0 component-wise, then the matrix Sk:=(ABDk1Ck) is positive definite and for the minimum eigenvalues we have λmin(Sk)λmin(A)>0. Furthermore, Sk1 is bounded independently of k.

The proof of Lemma 5.1 follows the steps of the analogous proof in [Citation32] and is hence omitted. Summarizing, the Newton method for the solution of the (5.1)–(5.4) is outlined in Algorithm Citation1. Here we have followed [Citation32] and project in every iteration the variables q, p onto the feasible sets such that the result of Lemma 5.1 holds.

Algorithm 1 Newton algorithm for the solution of the regularized TGV primal problem Pγ

while some stopping criterion is not satisfied do

Solve the linear system for x1k+1=(uk+1,wk+1) (ABDk1Ck)x1k+1=b1kBDk1b2k

Update x˜2k+1=(q˜k+1,p˜k+1) as follows x˜2k+1=Dk1(b2kCkx1k+1)

Compute qk+1,pk+1 as projections of q˜k+1,p˜k+1 onto the feasible sets {q:|q|α1},{p:|p|α0}

end while

The projections onto the feasible sets are defined, respectively, as (5.9) q=q˜max{1,|q˜|α1},p=p˜max{1,|p˜|α0},(5.9) with the equalities above to be considered component-wise.

6.2. The numerical algorithm for (PTGV)

We now describe our strategy for solving the discretized version of the bilevel TGV problem (PTGV). We note that in most of the experiments we will keep α0 a scalar – this is justified by the numerical results; see the relevant discussion later on. We will always mention the small modifications on the algorithm when α0 is a scalar. We will also make use here of the discrete Laplacian with zero Neumann boundary conditions Δ:UhUh which is used to act on the weight function α1. These are the desired boundary conditions for α0,α1 as dictated by the regularity result for the H1–projection in [Citation29, Corollary 2.3]. For a function uUh we define the discrete 2 norm as u2(Ωh)2=h2(i,j)Ωh|ui,j|2.

For the discrete H1 norm applied to the weight function α1 we use α1H1(Ωh)=hα1(IΔ)α1, while the dual norm is defined as rH1(Ωh)=(IΔ)1rH1(Ωh)=hr(IΔ)1r based on the H1H1(Ω) Riesz map αr=(IΔ)α. For the discrete version of the averaging filter in the definition of the localized residuals (4.1) we use a filter of size nw×nw, with entries of equal value whose sum is equal to one. With these definitions the discrete version of the bilevel TGV (PTGV), is the following: (PTGVh){minimize 12(R(u)σ¯2)+2(Ωh)2+12(σ¯2R(u))+2(Ωh)2+λ02α0H1(Ωh)2+λ02α1H1(Ωh)2,over (u,α0,α1)Uh×(Aad0)h×(Aad1)h, subject to       BuμΔudivqf=0,νwνΔwqdivp=0,maxδ(|uw|,γ1)qα1(uw)=0,maxδ(|Ew|,γ0)pα0Ew=0.

Here the box constraint sets are defined as (Aad0)h={α0Uh:α¯0(α0)i,jα¯0,for all (i,j)Ωh},(Aad1)h={α1Uh:α¯1(α1)i,jα¯1,for all (i,j)Ωh}.

The discretized versions of (5.1)–(5.4) and the upper level objective are still denoted by gpd(x)=0, and J respectively.

Regarding the choice of the lower and upper bounds for the local variance σ¯2 and σ¯2, respectively, we follow here the following rules, where σ2 is the variance of the “Gaussian” noise contaminating the data: (5.10) σ¯2=σ2(1+2nw),σ¯2=σ2(12nw).(5.10)

The formulae (5.10) are based on the statistics of the extremes; see [Citation29, Section 4.2.1].

We now proceed by describing the algorithm for the numerical solution of (PTGVh). In essence, we employ a discretized projected gradient method with Armijo line search. We briefly describe how the discrete gradient of the reduced objective functional is computed with the help of the adjoint equation. The corresponding discretized version of the latter Dxgpd(x)=DxJ(x), where x:=(u,w,q,p) is the adjoint variable, reads (5.11) [ACBD](uwqp)=(2(uf)(w((R(u)σ¯2)+(σ¯2R(u))+))000):=(b1b2),(5.11) where the matrices above were defined in (5.5) and (5.6). The equation can be solved again for x1:=(u,w) first and then subsequently for x2:=(q,p) as follows (AC(D)1B)x1=b1,x2=(D)1(b2Bx1).

The derivatives of the reduced objective with respect to α0 and α1, respectively, are (5.12) Ĵα0(α0,α1)=(Dα0gpd)x+Dα0J(α0,α1)(5.12) (5.13) =[IdId2Id](000diag(Ew))(uwqp)+λ0(IdΔ)α0=[IdId2Id]diag(Ew)p+λ0(IdΔ)α0,(5.13) (5.14) Ĵα1(α0,α1)=(Dα1gpd)x+Dα1J(α0,α1)=[IdId](00diag(Duw)0)(uwqp)+λ1(IdΔ)α1,=[IdId]diag(Duw)q+λ1(IdΔ)α1,(5.14) where x=(u,w,q,p) solves gpd(x)=0 for α0,α1. The corresponding reduced gradients are (5.15) αiĴ(α0,α1)=(IΔ)1Ĵαi(α0,α1),i=0,1.(5.15)

We note that in the case of a scalar α0, we set λ0=0. Then, Ĵα0(α0,α1)=[1121]diag(Ew)p, and α0Ĵ(α0,α1)=Ĵα0(α0,α1). Here 1 denotes a matrix of size 1×nm with all entries equal to one. In summary, the projected gradient algorithm for the solutions of (PTGVh) is described in Algorithm Citation2

We lastly note that the projections P(Aad0)h,P(Aad1)h are computed as described in [Citation29, Algorithm Citation4], that is via the semismooth Newton method developed in [Citation54]. We only mention that the original discretized H1–projection problem P(Aad)h(α˜) given by (5.16) {min12αα˜H1(Ωh)2:=h22(αα˜)(IΔ)(αα˜),overα(Aad)h={αUh:α¯αi,jα¯},(5.16) is approximated by the following penalty version: (5.17) minαUh12αα˜H1(Ωh)2+1ϵα(12(αα¯)+2(Ωh)2+12(α¯α)+2(Ωh)2),(5.17) with some small ϵα>0. For the projection regarding a scalar α0, we simply set P(Aad0)h(α0)=max(min(α0,α¯0),α¯0).

Algorithm 2: Discretized projected gradient method for the bilevel TGV problem (PTGVh)

Input: f, α¯0, α¯0,α¯1, α¯1,σ¯,σ¯, λ0, λ1, μ, ν, γ0, γ1, δ, nwτ00,τ10, 0<c<1,0<θ<1θ+

Initialize: α00(Aad0)h, α10(Aad1)h and set k = 0.

repeat

 Use the Algorithm Citation1 to compute the solution xk=(uk,wk,qk,pk) of the lower level problem gpd(uk,wk,qk,pk)=0

 Solve the adjoint Equationequation (5.11) for (u,w,q,p)

 Compute the derivative of the reduced objective with respect to α0 and α1 as in (5.13) and (5.14)

 Compute the reduced gradients αiĴ(α0k,α1k)=(IΔ)1Ĵαi(α0k,α1k),i=0,1

 Compute the trial points αik+1=P(Aadi)h(αikτikαiĴ(α0k,α1k)),i=0,1

while Ĵ(α0k+1,α1k+1)>Ĵ(α0k,α1k)+c(Ĵα0(α0k,α1k)(α0k+1α0k)+Ĵα1(α0k,α1k)(α1k+1α1k))

do (Armijo line search)

Set τ0k:=θτ0k,τ1k:=θτ1k and re-compute αik+1=P(Aadi)h(αikτikαiĴ(α0k,α1k)),i=0,1

end while

 Update τ0k+1=θ+τ0k,τ1k+1=θ+τ1k and k:=k+1

until some stopping condition is satisfied

6.3. Numerical examples in denoising

We now discuss some weighted TGV numerical examples, with regularization weights produced automatically by Algorithm Citation2. We are particularly interested in the degree of improvement over the scalar TGV examples. We are also interested in whether the statistics-based upper level objective enforces an automatic choice of regularization parameters that ultimately leads to a reduction of the staircasing effect. Our TGV results are also compared with the bilevel weighted TV method of [Citation27, Citation29]. In order to have a fair comparison, we use a Huber TV regularization for the corresponding lower lever problem, with the latter substituted by the TV primal-dual optimality conditions, udivpf=0max(|u|,γ)pαu=0, where α is now the spatially dependent regularization parameter for TV. We use the same values for the Huber parameter γ both in bilevel TV and bilevel TGV. The associated test images are depicted in with resolution n=m=256. The first one is the well-known “Cameraman” image which essentially consists of a combination of piecewise constant parts and texture. The next two images, “Parrot” and “Turtle” contain large piecewise affine type areas, thus they are more suitable for the TGV prior. The final image “hatchling” is characterized by highly oscillatory patterns of various kinds, depicting sand in various degrees of focus.

Figure 3. Test images, resolution 256×256.

Figure 3. Test images, resolution 256×256.

Parameter values: For the lower level primal-dual TGV problem we used μ = 0, ν = 1, δ=105, γ0=γ1=103, and a mesh size h = 1. For the H1–projection, we set ϵα=106, and we also weighted the discrete Laplacian Δ with 6×104. For the lower and upper bounds of α0 and α1 we set here α¯0=102,α¯0=10 and α¯1=104,α¯1=10. We also set λ1=1011 and when we spatially varied α0 we also set λ0=1011. We used a normalized nw×nw filter for w (i.e., with entries 1/nw2), with nw = 7. The local variance barriers σ¯2 and σ¯2 were set according to (5.10). For our noisy images we have σ2=102, and thus the corresponding values for (σ¯,σ¯) are (0.00798,0.01202). For the Armijo line search the parameters were τ00=0.05, τ10=100,c=109,θ=0.25,θ+=2. We solved each lower level problem until the residual of each of the optimality conditions (5.1)–(5.4) had Euclidean norm less than 104. MATLAB’s backslash was used for the solution of the linear systems.

We note that the initialization of the algorithm needs some attention. As it was done in [Citation29] for the TV case, α00 and α01 must be large enough in order to produce cartoon-like images, providing the local variance estimator with useful information. However, if α0 is initially too large then there is a danger of falling into the regime, in which the TGV functional and hence the solution map of (at least the non-regularized) lower level problem does not depend on α0. In that case the derivative of the reduced functional with respect to α0 will be close to zero, thus making no or little progress with respect to its optimal choice. Indeed this was confirmed after some numerical experimentation. Note that an analogous phenomenon can occur also in the case where α0 is much smaller than α1. In that case it is the effect of α1 which vanishes. This has been shown theoretically in [Citation37, Proposition 2] for dimension one, but numerical experiments indicate that this phenomenon persists also in higher dimensions. In our examples we used α10=0.25 and α00=0.2. Regarding the termination of the projected gradient algorithm, we used a fixed number of iterations n = 40. Neither the upper level objective nor the argument changed significantly after running the algorithm for more iterations; see for instance . The same holds true for the corresponding PSNR and SSIM values. We also note that a termination criterion as in [Citation29] based on the proximity measures P(Aadi)h(αikαiĴ(α0k,α1k))αikH1(Ωh), i = 0, 1, is also possible here.

Figure 4. Upper level objective values vs projected gradient iterations for the problem (PTGVh) (right) of (scalar α0, spatial α1).

Figure 4. Upper level objective values vs projected gradient iterations for the problem (PTGVh) (right) of Figure 5 (scalar α0, spatial α1).

We note that due to the line search, the number of times that the lower level problem has to be solved is more than the number of projected gradient iterations. For instance for the four examples the lower level problem had to be solved 58, 57, 57, and 57 times, respectively, with typically 8-12 Newton iterations needed per each lower level problem, This resulted in each bilevel problem (40 projected gradient iterations) requiring approximately 45-50 times the CPU time needed to solve one instance of the lower level problem. We note however that in general after 8-9 projected gradient iterations the PSNR and SSIM values of the reconstruction were essentially reaching their top values, requiring about 6-8 times the CPU time needed for the lower level problem. Further improvement with respect to computational times can be achieved using faster descent methods than the one we used here, like for instance [Citation55,Citation56].

For the first series of examples we keep the parameter α0 scalar, whose value nevertheless is determined by the bilevel algorithms. We depict the examples in . The first row shows the noisy images, while the second contains the bilevel TV results. The third row depicts the best scalar TGV results with respect to SSIM, where we have computed the optimal scalars α0,α1 with a manual grid method. The fourth row shows the results of (PTGVh). Detailed sections of all the images of are highlighted in . The weight functions α1 for the bilevel TV and the bilevel TGV algorithms are shown in . In we report all PSNR and SSIM values of the best scalar methods (scalar TV, scalar TGV) with respect to both quality measures, the corresponding values of the bilevel TV and TGV algorithms, as well as the ones the correspond to solving the bilevel TGV with the statistics-based upper level objective (4.2) but using scalar parameters only (third row from the end in ). For the latter case, we used no additional regularization for the weights in the upper level objective. We also report the PSNR and SSIM values of the computed results when both α0, α1 are spatially varying (last row), which we discuss later on in this section. We next comment on the results for each image.

Figure 5. First row: noisy images. Second row: bilevel TV. Third row: Best scalar TGV (SSIM). Fourth row: bilevel TGV.

Figure 5. First row: noisy images. Second row: bilevel TV. Third row: Best scalar TGV (SSIM). Fourth row: bilevel TGV.

Figure 6. Details of the reconstructions shown in .

Figure 6. Details of the reconstructions shown in Figure 5.

Figure 7. First row: the computed regularization functions α for bilevel TV. Second row: the computed regularization functions α1 for bilevel TGV.

Figure 7. First row: the computed regularization functions α for bilevel TV. Second row: the computed regularization functions α1 for bilevel TGV.

Table 1. PSNR and SSIM comparisons for the images of .

Cameraman: Here both the best PSNR and SSIM are obtained by the bilevel TV algorithm. This is probably not surprising due to the piecewise constant nature of this image. However, the bilevel TGV algorithm improve upon its scalar version with respect to both measures.

Parrot: Here the best results with respect to both PSNR and SSIM are achieved by the bilevel TGV algorithm, (PTGVh). There is significant improvement over all TV methods, which is due to the parameters being chosen in a way such that the staircasing effect diminishes. Furthermore, we observe improvement over the scalar TGV result especially around the parrot’s eye, where the weights α1 drop significantly; see the second column of .

Turtle: We get analogous results here as well, with the bilevel TGV (PTGVh) producing the best results both with respect to PSNR and SSIM. There is a significant reduction of the staircasing effect, while the weight α1 drops in the detailed areas of the image (head and flipper of the turtle).

Hatchling: In this image, the best PSNR and best SSIM are achieved by the scalar TGV when its parameters are manually optimized with respect to each measure (using the ground truth). However, the (ground truth-free) bilevel TGV achieves a better SSIM and PSNR value compared to these two results, respectively, achieving a better balance with respect to these measures. Similarly the scalar TV results are generally better than the ones of bilevel TV. We attribute this to the fact that the natural oscillatory features of the image are interpreted as noise by the upper level objective. Nevertheless, all the bilevel methods are able to locate and preserve better the eyes area, i.e., sand in focus, with the weight α1 dropping there significantly.

We remark that in all four examples, the ground truth-free bilevel TGV with at least one spatially varying parameter always produces better results with respect to both PSNR and SSIM, than the corresponding ground truth-free bilevel TGV with both parameters being scalar, compare the third last with the second last row of .

Finally, we would like to experiment with the case were also the weight α0 varies spatially. We note that by spatially varying both TGV parameters, the reduced problem becomes highly non-convex with many combinations of these parameters leading to similar values for the upper level objective. In order to deal with this, we solve a slightly different problem by performing one step of alternating optimization. Here we optimize only with respect to a spatially varying α0 having fixed the spatial weight α1 as it has been computed by the previous experiments, see the last row of . According to our numerical experiments, this strategy produces satisfactory results. As initialization for α0, we set it constant, equal to 5.

In we depict the computed spatially varying parameters α0 as well as the corresponding PSNR and SSIM values. Observe that the shape of α0 is different to the one of α1, compare the last row of to the second row of . This implies that a non-constant ratio of α0/α1 is preferred throughout the image domain. Secondly, by spatially varying α0 we only get a slight improvement with respect to PSNR and SSIM in the first two images, compare the second last and the last row of . However, it is interesting to observe the spatial adaptation of α0 with respect to piecewise constant versus piecewise smooth areas. The values of α0 are high in large piecewise constant areas, like the background of cameraman, the left area of the parrot image, as well as the top-right and the bottom-left corner of the turtle image. This is not so surprising as large values of α0 imply a large ratio α0/α1 and a promotion of TV like behavior in those areas. We can observe this in more detail at the parrot image, see last row of . On the contrary, the values of α0 are kept small in piecewise smooth areas like the right part of the parrot image and the sun rays around the turtle’s body. This results in low ratio α0/α1 and thus to a more TGV like behavior, reducing the staircasing effect. This is another indication of the fact that by minimizing the statistics-based upper level objective one is able not only to better preserve detailed areas but also to finely adjust the TGV parameters such that the staircasing is reduced.

Figure 8. Experiments with optimizing over a spatially varying α0. Top row: the automatically computed scalar parameters α0, that correspond to the images of the last row of . Middle row: the automatically computed spatially varying parameters α0, where α1 has been kept fixed (last row of ). The weight α0 is adapted to piecewise constant parts having there large values and hence promoting TV like behavior, see for instance the parrot image at the last row. On the contrary α0 has low values in piecewise smooth parts promoting a TGV like behavior reducing the staircasing.

Figure 8. Experiments with optimizing over a spatially varying α0. Top row: the automatically computed scalar parameters α0, that correspond to the images of the last row of Figure 5. Middle row: the automatically computed spatially varying parameters α0, where α1 has been kept fixed (last row of Figure 7). The weight α0 is adapted to piecewise constant parts having there large values and hence promoting TV like behavior, see for instance the parrot image at the last row. On the contrary α0 has low values in piecewise smooth parts promoting a TGV like behavior reducing the staircasing.

Lastly, we would like to see the degree of improvement of bilevel TGV with spatial parameters over the other ground truth-free approaches, i.e. using the statistics-based upper level objective, over a larger set of images. We thus used the 24 images of the Kodak dataset http://r0k.us/graphics/kodak/, and selected two middle square parts from each photograph, resized to 256 × 256 pixels, see left part of . We added two more photographs (see bottom-right part in the depicted collection) to reach 50 photographs in total. In each photograph, we run the bilevel TV and bilevel TGV experiments after adding the same type of Gaussian noise (with only α1 being spatially varying), as well as the bilevel TGV with both parameters to be scalar and we compared the corresponding PSNR differences, PSNRspatialTGVPSNRspatialTV, PSNRspatialTGVPSNRscalarTGV, as well as the corresponding SSIM ones, for every one of the 50 images. The PSNR improvement for the bilevel TGV over bilevel TV (both spatial parameters) was 0.05 ± 0.14 (mean ± standard deviation) while the improvement with respect to SSIM was more moderate, 0.003 ± 0.008. In (top) we also show the corresponding histograms of this difference. The fact that for some images the bilevel TV algorithm produces images of higher PSNR and/or SSIM, as it was the case for the cameraman, is perhaps not so surprising as the same has been observed in similar tests in [Citation26] in the case of scalar parameters only and for a bilevel scheme that maximizes the PSNR of the reconstructed image, using a ground truth-based upper level objective. We stress that for large ratio α0/α1 TGV is equivalent to TV only up an affine correction [Citation37] and for images that have a piecewise constant structure, setting α0 in TGV to be large enough might not be enough to achieve the same reconstruction as TV. On the other hand the improvement of bilevel TGV with spatially varying α1 versus its scalar analogue was 0.08±0.17 for the PSNR and 0.005 ± 0.008 for the SSIM, see bottom histograms of . We note that the last two methods require a similar computational effort since the additional H1–projection for the spatially varying weight is typically quite fast and does not significantly contribute to the overall computational time.

Figure 9. Histograms of the PSNR and SSIM differences between the bilevel spatial TGV and bilevel spatial TV (top) and between the bilevel spatial TGV and bilevel scalar TGV (bottom), for reconstructions of noisy versions of the 50 images on the left. In all bilevel algorithms the ground truth-free statistics-based upper level objective was used.

Figure 9. Histograms of the PSNR and SSIM differences between the bilevel spatial TGV and bilevel spatial TV (top) and between the bilevel spatial TGV and bilevel scalar TGV (bottom), for reconstructions of noisy versions of the 50 images on the left. In all bilevel algorithms the ground truth-free statistics-based upper level objective was used.

7. Conclusion

In this work we have adapted the bilevel optimization framework of [Citation27, Citation29] for automatically computing spatially dependent regularization parameters for the TGV regularizer. We established a rigorous dualization framework for the lower level TGV minimization problem that formed the basis for its algorithmic treatment via a Newton method. We showed that the bilevel optimization framework with the statistics/localized residual based upper level objective is able to automatically produce spatially varying parameters that not only adapt to the level of detail in the image but also reduce the staircasing effect.

Future continuation of this work includes adaptation of the bilevel TGV framework for advanced inverse problems tasks, i.e., Magnetic Resonance Imaging (MRI) and Positron Emission Tomography (PET) reconstruction as well as in multimodal medical imaging problems where structural TV based regularizers (edge aligning) have been suggested. That will also require devising a bilevel scheme where the invertibility assumption on T is dropped, since the linear operators involved in these inverse problems do not satisfy this assumption. Adaptation of the framework for different noise distributions e.g. Poisson, Salt & Pepper as well as combination of those [Citation57,Citation58], should also be investigated. Finally, a fine structural analysis of the weighted TGV regularized solutions in the spirit of [Citation59,Citation60] would be also of interest.

Additional information

Funding

This work is funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy – The Berlin Mathematics Research Center MATH+ (EXC-2046/1, project ID: 390685689). It was further supported by the MATHEON Research Center project CH12 funded by the Einstein Center for Mathematics (ECMath) Berlin. M.H and K.P. acknowledge support of Institut Henri Poincaré (UMS 839 CNRS-Sorbonne Université), and LabEx CARMIN (ANR-10-LABX-59-01). C.N.R. was supported by NSF grant DMS-2012391. H.S. acknowledges the financial support of Alexander von Humboldt Foundation.

References

  • Adams, R. A., Fournier, J. (2003). Sobolev Spaces. 2nd ed. Cambridge, MA: Academic Press,
  • Ambrosio, L., Fusco, N., Pallara, D. (2000). Functions of Bounded Variation and Free Discontinuity Problems. USA: Oxford University Press.
  • Attouch, H., Brezis, H. (1986). Duality for the Sum of Convex Functions in General Banach Spaces. Vol. 34, North-Holland Mathematical Library, pp. 125–133.
  • Barzilai, J., Borwein, J. M. (1988). Two-point step size gradient methods. IMA J Numer Anal. 8(1):141–148. DOI: 10.1093/imanum/8.1.141.
  • Benning, M., Brune, C., Burger, M., Müller, J. (2013). Higher-order TV methods: Enhancement via Bregman iteration. J Sci Comput. 54(2-3):269–310. DOI: 10.1007/s10915-012-9650-3.
  • Bergounioux, M., Papoutsellis, E., Stute, S., Tauber, C. (2018). Infimal convolution spatiotemporal PET reconstruction using total variation based priors, HAL preprint https://hal.archives-ouvertes.fr/hal-01694064.
  • Borwein, J. M., Vanderwerff, J. D. (2010). Convex Functions. Cambridge: Cambridge University Press.
  • Bredies, K., Dong, Y., Hintermüller, M. (2013). Spatially dependent regularization parameter selection in total generalized variation models for image restoration. Inter. J. Comput. Math. 90(1):109–123. DOI: 10.1080/00207160.2012.700400.
  • Bredies, K., Holler, M. (2012). Artifact-free JPEG decompression with total generalized variation. VISAP 2012: Proceedings of the International Conference on Computer Vision and Applications,
  • Bredies, K., Holler, M. (2013). A TGV regularized wavelet based zooming model. Scale Space and Variational Methods in Computer Vision. Springer, DOI: 10.1007/978-3-642-38267-3_13., pp. 149–160.
  • Bredies, K., Holler, M. (2014). Regularization of linear inverse problems with total generalized variation. J. Inverse Ill-Posed Prob. 22(6):871–913. DOI: 10.1515/jip-2013-0068.
  • Bredies, K., Holler, M. (2015). A TGV-based framework for variational image decompression, zooming, and reconstruction. Part I: Analytics. SIAM J. Imaging Sci. 8(4):2814–2850. DOI: 10.1137/15M1023865.
  • Bredies, K., Holler, M. (2015). A TGV-based framework for variational image decompression, zooming, and reconstruction. Part II: Numerics. SIAM J. Imaging Sci. 8(4):2851–2886. DOI: 10.1137/15M1023877.
  • Bredies, K., Holler, M., Storath, M., Weinmann, A. (2018). Total generalized variation for manifold-valued data. SIAM J. Imaging Sci. 11(3):1785–1848. DOI: 10.1137/17M1147597.
  • Bredies, K., Kunisch, K., Pock, T. (2010). Total generalized variation. SIAM J. Imaging Sci. 3(3):492–526. DOI: 10.1137/090769521.
  • Bredies, K., Kunisch, K., Valkonen, T. (2013). Properties of L1-TGV 2: The one-dimensional case. J. Math. Anal. Appl. 398(1):438–454. DOI: 10.1016/j.jmaa.2012.08.053.
  • Bredies, K., Valkonen, T. (2011). Inverse problems with second-order total generalized variation constraints. Proceedings of SampTA 2011 - 9th International Conference on Sampling Theory and Applications, Singapore,
  • Burger, M., Papafitsoros, K., Papoutsellis, E., Schönlieb, C. B. (2016). Infimal convolution regularisation functionals of BV and [Formula: see text] Spaces: Part I: The Finite [Formula: see text] Case. J. Math. Imaging Vis. 55(3):343–369. DOI: 10.1007/s10851-015-0624-6.
  • Calatroni, L., Chung, C., Los Reyes, J. D., Schönlieb, C. B., Valkonen, T. (2017). Bilevel approaches for learning of variational imaging models. RADON Book Series on Computational and Applied Mathematics. Vol. 18, Berlin, Boston: De Gruyter, https://www.degruyter.com/view/product/458544.
  • Calatroni, L., De Los Reyes, J. C., Schönlieb, C. B. (2017). Infimal convolution of data discrepancies for mixed noise removal. SIAM J. Imaging Sci. 10(3):1196–1233. DOI: 10.1137/16M1101684.
  • Calatroni, L., Papafitsoros, K. (2019). Analysis and automatic parameter selection of a variational model for mixed gaussian and salt-and-pepper noise removal. Inverse Prob. 35(11):114001. DOI: 10.1088/1361-6420/ab291a.
  • Caselles, V., Chambolle, A., Novaga, M. (2007). The discontinuity set of solutions of the TV denoising problem and some extensions. Multiscale Model. Simul. 6(3):879–894. DOI: 10.1137/070683003.
  • Chambolle, A., Duval, V., Peyré, G., Poon, C. (2017). Geometric properties of solutions to the total variation denoising problem. Inverse Prob. 33(1):015002. http://stacks.iop.org/0266-5611/33/i=1/a=015002. DOI: 10.1088/0266-5611/33/1/015002.
  • Chambolle, A., Lions, P. L. (1997). Image recovery via total variation minimization and related problems. Numerische Mathematik. 76(2):167–188. DOI: 10.1007/s002110050258.
  • Van Chung, C., De los Reyes, J. C., Schönlieb, C. B. (2017). Learning optimal spatially-dependent regularization parameters in total variation image denoising. Inverse Prob. 33(7):074005. DOI: 10.1088/1361-6420/33/7/074005.
  • De Los Reyes, J. C., Schönlieb, C. B., Valkonen, T. (2016). The structure of optimal parameters for image restoration problems. J. Math. Anal. Appl. 434(1):464–500. DOI: 10.1016/j.jmaa.2015.09.023.
  • De Los Reyes, J. C., Schönlieb, C. B., Valkonen, T. (2017). Bilevel parameter learning for higher-order Total Variation regularisation models. J. Math. Imaging Vis. 57(1):1–25. DOI: 10.1007/s10851-016-0662-8.
  • Demengel, F., Temam, R. (1984). Convex functions of a measure and applications. Indiana Univ. Math. J. 33(5):673–709. DOI: 10.1512/iumj.1984.33.33036.
  • Ekeland, I., Temam, R. (1999). Convex analysis and variational problems. Classics Appl. Mathem. Soci. Indust. Appl. Math.
  • Girault, V., Raviart, P. A. (1986). Finite Element Method for Navier-Stokes Equation. Berlin, Germany: Springer.
  • Hintermüller, M., Holler, M., Papafitsoros, K. (2018). A function space framework for structural total variation regularization with applications in inverse problems. Inverse Prob. 34(6):064002. http://stacks.iop.org/0266-5611/34/i=6/a=064002. DOI: 10.1088/1361-6420/aab586.
  • Hintermüller, M., Kunisch, K. (2006). Path-following methods for a class of constrained minimization problems in function space. SIAM J. Optim. 17(1):159–187. DOI: 10.1137/040611598.
  • Hintermüller, M., Papafitsoros, K. (2019). Generating structured nonsmooth priors and associated primal-dual methods. Processing, Analyzing and Learning of Images, Shapes, and Forms: Part 2. In: Ron K. and Xue-Cheng T., eds. Handbook of Numerical Analysis, Vol. 20, DOI: 10.1016/bs.hna.2019.08.001., pp. 437–502.
  • Hintermüller, M., Papafitsoros, K., Rautenberg, C. N. (2017). Analytical aspects of spatially adapted total variation regularisation. J. Math. Anal. Appl. 454(2):891–935. DOI: 10.1016/j.jmaa.2017.05.025.
  • Hintermüller, M., Papafitsoros, K., Rautenberg, C. N. (2020). Variable step mollifiers and applications. Integr. Equ. Oper. Theory. 92(6). DOI: 10.1007/s00020-020-02608-2.
  • Hintermüller, M., Rautenberg, C. N. (2015). On the density of classes of closed convex sets with pointwise constraints in Sobolev spaces. J. Math. Anal. Appl. 426(1):585–593. DOI: 10.1016/j.jmaa.2015.01.060.
  • Hintermüller, M., Rautenberg, C. N. (2017). Optimal selection of the regularization function in a weighted total variation model. Part I: Modelling and theory. J. Math. Imaging Vis. 59(3):498–514. DOI: 10.1007/s10851-017-0744-2.
  • Hintermüller, M., Rautenberg, C. N., Rösel, S. (2017). Density of convex intersections and applications. Proc. Royal Soci. London A Math. Phys. Engng Sci. 473(2205) DOI: 10.1098/rspa.2016.0919.
  • Hintermüller, M., Rautenberg, C. N., Wu, T., Langer, A. (2017). Optimal selection of the regularization function in a weighted total variation model. Part II: Algorithm, its analysis and numerical tests. J. Math. Imaging Vis. 59(3):515–533. DOI: 10.1007/s10851-017-0736-2.
  • Hintermüller, M., Stadler, G. (2006). An infeasible primal-dual algorithm for total bounded variation–based inf-convolution-type image restoration. SIAM J. Sci. Comput. 28(1):1–23. DOI: 10.1137/040613263.
  • Holler, M., Kunisch, K. (2014). On infimal convolution of TV-type functionals and applications to video and image reconstruction. SIAM J. Imaging Sci. 7(4):2258–2300. DOI: 10.1137/130948793.
  • Huber, R., Haberfehlner, G., Holler, M., Kothleitner, G., Bredies, K. (2019). Total generalized variation regularization for multi-modal electron tomography. Nanoscale. 11(12):5617–5632. DOI: 10.1039/c8nr09058k.
  • Jalalzai, K. Discontinuities of the minimizers of the weighted or anisotropic total variation for image reconstruction, arXiv preprint 1402.0026 (2014), http://arxiv.org/abs/1402.0026.
  • Jalalzai, K. (2016). Some remarks on the staircasing phenomenon in total variation-based image denoising. J. Math. Imaging Vis. 54(2):256–268. DOI: http://dx.doi.org/10.1007/s10851-015-0600-1.
  • Knoll, F., Bredies, K., Pock, T., Stollberger, R. (2011). Second order total generalized variation (TGV) for MRI. Magn. Reson. Med. 65(2):480–491. DOI: 10.1002/mrm.22595.
  • Knoll, F., Holler, M., Koesters, T., Otazo, R., Bredies, K., Sodickson, D. K. (2017). Joint MR-PET reconstruction using a multi-channel image regularizer. IEEE Trans. Med. Imaging. 36(1):1–16. DOI: 10.1109/TMI.2016.2564989.
  • Kunisch, K., Hintermüller, M. (2004). Total bounded variation regularization as a bilaterally constrained optimization problem. SIAM J. Appl. Math. 64(4):1311–1333. DOI: 10.1137/S0036139903422784.
  • Nesterov, Y. E. (1983). A method for solving the convex programming problem with convergence rate O(1/k2). Soviet Math. Dokl. 27:367–372.
  • Papafitsoros, K. (2014). Novel higher order regularisation methods for image reconstruction. Ph.D. thesis. University of Cambridge. https://www.repository.cam.ac.uk/handle/1810/246692.
  • Papafitsoros, K., Bredies, K. (2015). Department of Applied Mathematics and Theoretical Physics, University of Cambridge, Wilberforce Road, CB3 0WA, Cambridge A study of the one dimensional total generalised variation regularisation problem. Inverse Prob. Imaging. 9(2):511–550. DOI: 10.3934/ipi.2015.9.511.
  • Papafitsoros, K., Valkonen, T. (2015). Asymptotic behaviour of total generalised variation. Scale Space and Variational Methods in Computer Vision: 5th International Conference, SSVM 2015, Proceedings. In: Jean-François A., Mila N., and Nicolas P., eds. Springer International Publishing, pp. 702–714. DOI: 10.1007/978-3-319-18461-6_56.
  • Pöschl, C., Scherzer, O. (2015). Exact solutions of one-dimensional total generalized variation. Comm. Math. Sci. 13(1):171–202. DOI: 10.4310/CMS.2015.v13.n1.a9.
  • Ring, W. (2000). Structural properties of solutions to total variation regularization problems. Esaim: M2an. 34(4):799–810. DOI: 10.1051/m2an:2000104.
  • Rudin, L. I., Osher, S., Fatemi, E. (1992). Nonlinear total variation based noise removal algorithms. Physica D: Nonlinear Phenomena. 60(1-4):259–268. DOI: 10.1016/0167-2789(92)90242-F.
  • Schloegl, M., Holler, M., Schwarzl, A., Bredies, K., Stollberger, R. (2017). Infimal convolution of total generalized variation functionals for dynamic MRI. Magn Reson Med. 78(1):142–155. DOI: 10.1002/mrm.26352.
  • Temam, R. (1985). Mathematical Problems in Plasticity. Vol. 15, Gauthier-Villars Paris.
  • Temam, R., Strang, G. (1980). Functions of bounded deformation. Arch. Rational Mech. Anal. 75(1):7–21. DOI: 10.1007/BF00284617.
  • Valkonen, T. (2017). The jump set under geometric regularisation. Part 2: Higher-order approaches. J. Math. Anal. Appl. 453(2):1044–1085. DOI: 10.1016/j.jmaa.2017.04.037.
  • Valkonen, T., Bredies, K., Knoll, F. (2013). Total generalized variation in diffusion tensor imaging. SIAM J. Imaging Sci. 6(1):487–525. DOI: 10.1137/120867172.
  • Zowe, J., Kurcyusz, S. (1979). Regularity and stability for the mathematical programming problem in Banach spaces. Appl Math Optim. 5(1):49–62. DOI: 10.1007/BF01442543.

Appendix A

We provide here the proof of Proposition 2.2:

Proof of Proposition 2.2

. The proof follows [Citation12, Proposition 3.3]. Denote by Cα,Kα the following convex sets (A.1) Cα={div2ϕ:ϕCc(Ω,Sd×d),|ϕ(x)|α0(x),|divϕ(x)|α1(x), for all xΩ},(A.1) (A.2) Kα={div2p:pH0(div2;Ω),|p(x)|α0(x),|divp(x)|α1(x),fora.e.xΩ}.(A.2)

It suffices to show that (A.3) Cα¯L2(Ω)=Kα.(A.3)

We first show that Kα is closed in L2(Ω). Let gL2(Ω) and assume that there exists (pn)nNH0(div2;Ω) where every pn satisfies the convex constraints and div2png in L2(Ω). By boundedness of α0,α1 we have that there exist h0L2(Ω,Sd×d), h1L2(Ω,Rd) and a subsequence of (pnk)kN such that pnkh0anddivpnkh1, in L2(Ω) and L2(Ω,Rd), respectively. Using that, we have for every ϕCc(Ω,Rd) (A.4) Ωϕ·h0 dx=limkΩϕ·pnk dx=limkΩϕ·divpnk dx=Ωϕ·h1 dx,(A.4) thus h1=divh0. Similarly we derive that g=divh1=div2h0 and hence h0H(div2;Ω). Finally note that the set {(h,divh,div2h):hH0(div2;Ω),|h(x)|α0(x),|divh(x)|α1(x),fora.e.xΩ}, is a norm-closed and convex subset of L2(Ω,(Sd×d×Rd×R)) and hence weakly closed. Since (pnk,divpnk,div2pnk)kN belongs to that set, converging weakly to (h0,divh0,div2h0) we get that the latter also belongs there. Thus, Kα is closed in L2(Ω) and since CαKα, we get Cα¯L2(Ω)Kα.

It remains to show the other direction, i.e., KαCα¯L2(Ω). Toward that, note first that the functional TGVα2(Ω):L2(Ω)R¯, can also be written as TGVα2(u)=ICα(u).

Using the convexity of Cα one gets TGVα2  (v)=ICα(v)=ICα¯L2(Ω)(v).

Secondly, note that due to the lower bounds on α0,α1, for uL2(Ω), we have that TGVα2(u)< if and only if uBV(Ω). Indeed this holds from the equivalence of the (scalar) ·BGV with ·BV(Ω) and from the estimate TGVα¯,α¯2(u)TGVα2(u)α1TV(u), for every uL2(Ω). This means that if for div2pKα it holds (A.5) Ωu div2p dxTGVα2(u),for all uBV(Ω)L2(Ω),(A.5) then in fact the inequality (A.5) will hold for every uL2(Ω) and thus TGVα2  (div2p)=0 which implies div2pCα¯L2(Ω). Thus in order to finish the proof it suffices to show (A.5) for every div2pKα. In view of Proposition 2.1 it suffices to show (A.6) Ωu div2p dxminwBD(Ω)Ωα1 d|Duw|+Ωα0 d|Ew|(A.6) for all uBV(Ω)L2(Ω). The first step toward that is to show that for every wBD(Ω) and for every pH0(div2;Ω) with |p(x)|α0(x) and |divp(x)|α1(x) for a.e. xΩ, it holds (A.7) Ωw·divp dxΩα0 d|Ew|.(A.7)

Indeed, note first that from (2.3), we get for every ϕC(Ω¯,Rd) (A.8) |Ωϕ·divp dx|=|Ωp·Eϕ dx|Ω|p||Eϕ| dxΩ α0 d|Eϕ|.(A.8)

Recall now that every wBD(Ω) can be strictly approximated by a sequence (ϕn)nNC(Ω¯,Rd), that is ϕnw in L1(Ω,Rd) and |Eϕ|(Ω)|Ew|(Ω), see [Citation12, Proposition 2.10]. Furthermore, using that, along with Reshetnyak’s continuity theorem [Citation33, Theorem 2.39] we also get that Ωα0 d|Eϕn|Ωα0 d|Ew|,as n.

Using the above and the fact that divpL(Ω,Rd), by taking limits in (A.8) we obtain (A.7). Finally in order to obtain (A.5) let again pH0(div2;Ω) with |p(x)|α0(x) and |divp(x)|α1(x) for a.e. xΩ, and let ϕC(Ω¯,R). Then by using (2.4) and (A.7), we have for every wBD(Ω) Ωϕ div2p dx|Ωϕ·divp dx|=|Ω(ϕw)·divp dx+Ωw divp dx|Ω|ϕw||divp| dx+Ωα0 d|Ew|Ωα1|ϕw| dx+Ωα0 d|Ew|.

Similarly as before given uBV(Ω)L2(Ω) and wBD(Ω), there exists a sequence (ϕn)nNC(Ω¯,R) such that ϕnu in L2(Ω) and |ϕnw|(Ω)|Duw|(Ω). This follows from a modification of the proof of [Citation12, Proposition 2.10] where one takes advantage of the L2 integrability of u. Using again the Reshetnyak’s continuity theorem and taking limits we get that for every wBD(Ω) Ωu div2p dxΩα1 d|Duw|+Ωα0 d|Ew|.

By taking the minimum over wBD(Ω), we obtain (A.5). □