• Aucun résultat trouvé

The Geometry of Sparse Analysis Regularization

N/A
N/A
Protected

Academic year: 2021

Partager "The Geometry of Sparse Analysis Regularization"

Copied!
22
0
0

Texte intégral

(1)

HAL Id: hal-02169356

https://hal.archives-ouvertes.fr/hal-02169356

Preprint submitted on 1 Jul 2019

HAL

is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire

HAL, est

destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

The Geometry of Sparse Analysis Regularization

Xavier Dupuis, Samuel Vaiter

To cite this version:

Xavier Dupuis, Samuel Vaiter. The Geometry of Sparse Analysis Regularization. 2019. �hal-02169356�

(2)

THE GEOMETRY OF SPARSE ANALYSIS REGULARIZATION

XAVIER DUPUIS AND SAMUEL VAITER

Abstract. Analysis sparsity is a common prior in inverse problem or machine learning including special cases such as Total Variation regularization, Edge Lasso and Fused Lasso. We study the geometry of the solution set (a polyhedron) of the analysis`1 regularization (with`2 data fidelity term) when it is not reduced to a singleton without any assumption of the analysis dictionary nor the degradation operator. In contrast with most theoretical work, we do not focus on giving uniqueness and/or stability results, but rather describe a worst-case scenario where the solution set can be big in terms of dimension. Leveraging a fine analysis of the sub-level set of the regularizer itself, we draw a connection between support of a solution and the minimal face containing it, and in particular prove that extremal points can be recovered thanks to an algebraic test. Moreover, we draw a connection between the sign pattern of a solution and the ambient dimension of the smallest face containing it.

Finally, we show that any arbitrary sub-polyhedra of the level set can be seen as a solution set of sparse analysis regularization with explicit parameters.

1. Introduction. We focus on a convex regularization promoting sparsity in an analysis dictionary in the context of a linear inverse problem/regression problem where the regularization reads:

(1.1) min

x∈Rn

1

2||y−Φx||22+λ||Dx||1

wherey∈Rq is a observation/response vector,Φ :RnRq is the sensing/acquisition linear operator,D: RpRnis a dictionary andλ >0the hyper-parameter used as a trade-off between fidelity and regularization. Note that at this point, we do not make any assumption on the dictionaryDor the acquisition operatorΦ.

This convex regularization is known as analysis`1-regularization [8] in the inverse problem community or generalized Lasso [22] in statistics. Let us mention that it includes several popular regularizers as special cases such that (anisotropic) total variation [17] whenDis a discrete difference operator, wavelet coefficient analysis [19]

using a wavelet transform as an analysis dictionary or fused Lasso [20] when using the concatenation of the identity matrix and a discrete difference operator, i.e., using a Lasso regularization with an additional constraint on the (discrete) gradient. In the noiseless context, when y ∈ Im Φ, the following constrained formulation is used instead of the Tikhonov formulation(1.1)as

(1.2) min

x∈Rn||Dx||1 subject to Φx=y.

We focus here on the noisy version of the regularization in order to keep our discussion concise. The purpose of this paper is to answer the following question:

When the solution set of (1.1)is not reduced to a singleton, what is its “geometry”?

1.1. Previous works.

Uniqueness certificate of analysis regularization. Among several theoretical issues, sufficient condition for the uniqueness of the solution set of(1.1)have been extensively studied, see for instance [25,15,1,26]. Several uniqueness conditions can be proposed,

Fundings: This work was partly supported by ANR GraVa ANR-18-CE40-0005 and Projet ANER RAGA G048CVCRB-2018ZZ.

UniversitÃľ de Bourgogne, Dijon, France (xavier.dupuis@u-bourgogne.fr).

CNRS & UniversitÃľ de Bourgogne, Dijon, France (samuel.vaiter@u-bourgogne.fr).

(3)

where the simplest is for instance requiringn6qandΦhaving full rank: the`2-loss term is strictly convex, and uniqueness follows from it. The task of studying the case when the solution set is not reduced to a singleton can be seen as rather formal since most of the time the solution set is reduced to a singleton [22,25], but nevertheless, it exhibits interesting properties of sparse analysis regularization.

Solution set of generic convex program. Describing the geometry of the solution set in convex optimization has been a subject of intense study starting from the work of [13] and its generalization to non-smooth convex program [7]. Several extensions have proposed such as [10] for pseudo-linear programs or in a different setting (min- imization of concave function), [12] shows that one can describe one solution with minimal sparsity level.

Representer theorems. We shall also remark that coming from the statistics com- munity, [11] (and popularized in [18]) initiates a line of work coined as representer theorems, culminating recently in [6] and [23]. The basic idea of these kinds of results is to show that under some assumption, one can write every element of the solution set of a convex program as a sum of elementary atoms.

Description of polytopes. Convex polytopes and polyhedrons are central objects in geometry [27] and convex analysis. A part of our results provides a connection between faces and signs of vector living in the analysis domain. We can draw a connection with the study of oriented matroid [4] and zonotopes [5] (analysis `1-ball are zonotopes) as described in [27, Lecture 7], in particular in section 7.3.

1.2. Contributions. In contrast to these last lines of work, we take here a more direct and specific approach. We give below an overview of our contributions.

Geometry of the analysis`1-ball. The first part of our work (section 2) is dedicated to studying the geometry of the analysis `1-ball. We study across several results the direction and the relative interior of the the intersection between the sub-level set of the regularizer and another set. We refine our analysis progressively starting from any convex component of the level set, then looking to sub-polyhedra of the sub-level set ending by the faces itself of the level set. We show several specific results:

• The sign pattern defines a bijection between the set of exposed faces of the analysis `1-ball and the set of feasible signs in the dictionary D as proved inProposition 2.24.

• The extremal points of the analysis `1-ball can be recovered with a purely algebraic result thanks toCorollary 2.25. We draw a link between our result and a remark in [6] which is a topological argument

• We show that from a theoretical perspective (but unfeasible from a numeri- cal point of view in high dimension), the geometry of the analysis `1-ball is summarized by the Hasse diagram of the feasible signs.

Geometry of the solution set. Thanks to the study of the analysis`1-ball, we give in a second part (section 3) consequences on the solution set of (1.1). We show that:

• Using Theorem 2.16and Proposition 2.17, we describe the geometry of the solution set inProposition 3.6.

• The solution set of (1.1) admits extreme points if and only if, the condition denoted by (H0) and assumed to hold all throughout [25] or [24], namely Ker Φ∩KerD ={0}, holds. In this case, the extreme points are precisely those which satisfy the condition denoted by (HJ) in [24] at a given solution to perform a sensitivity analysis, seeProposition 3.8.

• For any affine space which intersect non-trivially the unit-sphere, one can find Φ, y such that the solution set is exactly this intersection, see Theorem 3.9

2

(4)

andProposition 3.10.

1.3. Notations. For a given integern, the set of all integers between1andnis denoted by[n] ={1, . . . , n}.

Vectors and support. Given u ∈ Rm, the support supp(u) and the sign vector sign(u)are defined by

supp(u) ={i∈[m] : ui6= 0} and sign(u) = (sign(ui))i∈[m],

and its cardinal is coined the `0-norm ||u||0 = |supp(u)|. The cosupportcosupp(u) is the set cosupp(u) = [m]\supp(u). Given u, v∈Rm, the inner product is written hu, vi=Pm

i=1uivi and the associated norm is written ||u||2=p

hu, ui. We will also use the`1-norm ||u||1=Pm

i=1|ui|and `-norm||u||= maxi∈[m]|ui|.

Linear operators. Given a linear operatorD∈Rn×m,DRm×nis the transpose operator, D+Rn×mits Moore–Penrose pseudo-inverse, KerD ⊆Rm its null-space and ImD ∈Rn its column-space. Given I ⊆[m], DIRn×|I| is the matrix formed by the column ofDindexed byI. The identity operator is denotedIdmorId. Given a vector u ∈ Rm, xI is the vector of components indexed by I. Given a subspace F ⊆Rm, we denote byΠF the orthogonal projection onF. Given a vectoru∈Rn, its diagonalized matrixdiag(u)∈Rn×n is the diagonal matrix such thatdiag(u)ii =ui

for everyi∈[n].

Convex analysis. Given a convex, lower semicontinuous, proper function f : RmR, its sub-differential∂f is given

∂f(u) ={η∈Rm : f(u)>f(v) +hη, u−vi}.

Given a convex set C, the affine hull aff(C) is the smallest affine set containing C, the directiondir(C)ofCis the direction ofaff(C)and its relative interiorri(C)is the interior of C relative to its affine hull aff(C). The relative boundary rbd(C)ofC is the boundary ofC relative toaff(C). The dimensiondim(C)ofCis the dimension of aff(C). We say thatx∈Cis an extreme point if there are no two differentx1, x2∈C such thatx= x1+x2 2. The set of all extreme points of C is denoted byext(C). For instance, given two points x1 6=x2Rn, the segment C = [x1, x2] is such that its affine hull is aff(C) = {x1+tx2 : t∈R}, its relative interior is the open segment ri(C) = (x1, x2), its relative boundary and extremal setext(C) = rbd(C) ={x1, x2}, its dimension is 1 and its direction isdir(C) =R(x1−x2).

2. The unit ball of the sparse analysis regularizer. This section contains the core of our results. After giving preliminary results on sign vectors in subsec- tion 2.1, we show that the unit ball is a convex polyhedron by giving its half-space representation in subsection 2.2. Then, we study properties of convex subset of the unit-sphere insubsection 2.3which lead us toLemma 2.8which turns to be the foun- dation of latter results. Subsection 2.4contains a sequence of results which represent our main contribution: Theorem 2.16which describes in details the affine components of the unit-ball,Proposition 2.17which instantiates this result to setting of an affine components included in the unit sphere, Proposition 2.18 which extends this result to any exposed faces and finally Proposition 2.19which gives a necessary and suffi- cient condition of extremality. Finally, insubsection 2.5, we reformulate our previous results in order to describe the exposed faces of the unit-ball, and to show that there exists a bijection between the set of exposed faces and feasible signs. We also draw a connection to the work of [6].

(5)

2.1. Preliminary results on sign vectors. We first define an order on the set of all possible signs{−1,0,+1}palong with a notion of consistency of signs which can be related to the idea of “sub-signs”.

Definition 2.1. Let s, s0∈ {−1,0,+1}p. We say that

• ss0 if for all i∈[p],si 6= 0 ⇒s0i=si;

• sands0 are consistent if for all i∈[p],si6= 0ands0i 6= 0 ⇒s0i =si. The following remarks connect the notion of support/cosupport to this sign pat- tern. In

Remark 2.2. 1. ss0⇒supp(s)⊂supp(s0)⇔cosupp(s0)⊂cosupp(s);

2. Ifsands0 are consistent, then

ss0 ⇔supp(s)⊂supp(s0)⇔cosupp(s0)⊂cosupp(s);

3. Ifss00and s0s00, thensand s0 are consistent;

4. The set{−1,0,+1}p endowed with the order relationis a poset.

The following lemma gives a characterization of the`1-norm which will be used intensively in latter results.

Lemma 2.3. Let θ∈ Rp. Then for any s∈ {−1,0,+1}p, hs, θi ≤ kθk1, and the equality holds if and only ifsign(θ)s.

Proof. We prove the result component-wisely. Let α ∈ R. Then for any s ∈ {−1,0,+1},sα≤ |α|. Suppose now thatsign(α)s. Ifα= 0, then sα=|α|, and if α6= 0, thens= sign(α)and sα=|α|. Conversely, suppose thatsign(α)s. Then α6= 0ands6= sign(α), i.e. s= 0or s=−sign(α). In both cases,sα <|α|.

2.2. Half-space representation of the unit ball. We denote by B1 (resp.

∂B1) the unit ball (resp. the unit sphere), or sub-level set (resp. level set) for the value1, of the sparse analysis regularizerR:x7→ kDxk1:

B1={x∈Rn :kDxk1≤1},

∂B1={x∈Rn :kDxk1= 1}.

SinceR is one-homogeneous, the results of this section apply to all sub-level sets for positive values.

Proposition 2.4. The unit ball B1 is a full-dimensional convex polyhedron, a half-space representation of which is given by

B1= \

s∈{−1,0,1}p

{x∈Rn:hDs, xi ≤1}.

Proof. First note thatB1has a nonempty interior (in particular0∈B1), namely {x∈Rn:kDxk1<1}, which is equivalent for a convex set to be of full dimension.

Second denoteA=T

s∈{−1,0,1}p{x : hDs, xi61}. Letx∈B1. ByLemma 2.3, hDs, xi=hs, Dxi ≤ kDxk1 ≤1 for any s∈ {−1,0,1}p so x∈A. Conversely, let x ∈ A and s = sign(Dx). By Lemma 2.3, kDxk1 = hs, Dxi = hDs, xi ≤ 1 so x∈B1. ThenB1=Ais a convex polyhedron.

Note that this half-space representation is redundant. The general question of the minimal representation of H-polyhedron is known to be hard, we shall leave it to future work. However, we can use this proposition to derive a way to construct exposed face ofB1 as claimed in the following lemma.

4

(6)

Lemma 2.5. Let s¯∈ {−1,0,1}p. Then

B1∩ {x∈Rn:hDs, xi¯ = 1}=∂B1∩ {x∈Rn: sign(Dx)s};¯

it is either empty or an exposed face ofB1.

Proof. Letx∈B1. Then byLemma 2.3,hD¯s, xi= 1if and only if1 =h¯s, Dxi ≤ kDxk1 ≤ 1, if and only if kDxk1 = 1 and sign(Dx) s. If the intersection is¯ nonempty, then{x∈Rn:hDs, xi¯ = 1}is a supporting hyperplane ofB1 byProposi- tion 2.4, and thus its intersection withB1 is an exposed face of the polyhedron.

For a given ¯s, the set{x∈Rn : sign(Dx)s}¯ looks hard to describe. In fact, there exists a linear representation as told in the following lemma.

Lemma 2.6. Let s¯∈ {−1,0,1}p. Then

{x∈Rn : sign(Dx)s}¯ ={x∈Rn:DJ¯x= 0and diag(¯sI¯)DI¯x>0}

whereJ¯= cosupp(¯s)andI¯= supp(¯s).

Proof. It is a straightforward rewriting ofsign(Dx)s. Indeed,¯ sign(Dx)s¯⇔DJ¯x= 0and¯si(Dx)i>0,∀i∈I¯

⇔DJ¯x= 0and diag(¯sI¯)DI¯x>0.

Note that we can exchange the role of¯sandDx, and we also obtain that {x∈Rn : sign(Dx)¯s}={x∈Rn:DJ¯x= 0and diag(DI¯x)¯sI >0}.

2.3. Convex components of the unit sphere. In this section we consider nonempty convex subsetsC⊂∂B1. All the results will hold in particular for exposed faces ofB1.

We begin with a lemma on general convex sets.

Lemma 2.7. Let X be a nonempty convex set andC⊂X be a nonempty convex subset. Suppose that there exists an exposed face G of X such that ri(C)∩G 6=∅.

ThenC⊂G. Moreover, if Gis exposed inaff(X), thenG⊂rbd(X).

Proof. Recall that an exposed face ofXis defined asG=X∩{x:hα, xi=β}with {x:hα, xi=β} a supporting hyperplane ofX, i.e. such thatX ⊂ {x:hα, xi ≤ β}.

Suppose thatC 6⊂Gand let x∈C\G⊂X andx¯ ∈ri(C)∩G(nonempty). Then hα, xi< β andhα,xi¯ =β. Letd= ¯x−x∈dir(C). Sincex¯∈ri(C), x¯+εd∈C⊂X for|ε|small. Buthα,x+εdi¯ =β+ε(β−hα, xi)> βforε >0, which is a contradiction.

ThenC⊂G.

It is a classical result that G⊂ bd(X), see e.g. [9, Part III, Section 2.4]. If G is exposed inaff(X), we get thatG⊂rbd(X)by considering aff(X)as the ambient space.

The following lemma is the first result of a long number of consequences which study the direction and relative interior of the intersection of the unit ball with another set.

Lemma 2.8. Let x¯∈∂B1,¯s= sign(Dx),¯ J¯= cosupp(Dx), and¯ F¯=B1∩ {x∈Rn:hD¯s, xi= 1}.

(i) F¯ =∂B1∩ {x∈Rn: sign(Dx)¯s} and it is an exposed face;

(7)

(ii) C⊂F¯ for any nonempty convex subset C⊂B1 such that x¯∈ri(C);

(iii) dir( ¯F) = (D¯s)∩KerDJ¯;

(iv) ri(C)⊂ri( ¯F)for any nonempty convex subsetC⊂B1 such that x¯∈ri(C);

(v) ri( ¯F) =∂B1∩ {x∈Rn: sign(Dx) = ¯s}.

Proof. (i) The expression forF¯is given byLemma 2.5. It follows thatx¯∈F¯and thusF¯ is an exposed face ofB1.

(ii) LetC be a nonempty convex subset ofB1such thatx¯∈ri(C). Thenri(C)∩ F¯ 6=∅, and byLemma 2.7, C⊂F¯.

(iii) The inclusiondir( ¯F)⊂(Ds)¯follows from the definition ofF¯. Moreover, for anyx∈F¯,sign(Dx)s, which implies that¯ J¯⊂cosupp(Dx), i.e. DJ¯x= 0. Then F¯ ⊂KerDJ¯, anddir( ¯F)⊂KerDJ¯. Conversely, let d∈(Ds)¯∩KerDJ¯. Since d∈ KerDJ¯andsign(Dx) = ¯¯ s,cosupp(Dx)¯ ⊂cosupp(Dd)andsign(D(¯x+εd))¯sfor

|ε|small. Then byLemma 2.3,kx+εdk¯ 1=h¯s, D(¯x+εd)i=hDs,¯ x+εdi¯ =hDs,¯ xi¯ = 1 sinced∈(Ds)¯. Thenx¯+εd∈F¯ for|ε| small, andd∈dir( ¯F).

(iv)-(v) First, we prove that ∂B1∩ {x∈ Rn : sign(Dx) = ¯s} ⊂ri( ¯F)(thus in particular x¯ ∈ ri( ¯F)). Let x ∈ ∂B1 be such that sign(Dx) = ¯s. By (iii) and its proof, for any d∈dir( ¯F), x+εd∈F¯ for|ε| small. Thus x∈ri( ¯F). Second, let C be a nonempty convex subset ofB1such that x¯∈ri(C). By (ii),C ⊂F¯ ⊂∂B1. Let ˆ

x∈ri(C)and ˆs= sign(Dx). Note thatˆ ˆss¯sinceˆx∈F¯. LetFˆ =B1∩ {x∈Rn: hDˆs, xi= 1}. Applying (ii) toxˆ andC = ¯F, we get that F¯ ⊂Fˆ, which implies that

¯

ss. Thusˆ sign(Dx) = ¯ˆ sandri(C)⊂∂B1∩ {x∈Rn: sign(Dx) = ¯s}. This proves (iv) as well as the missing inclusion of (v) by settingC= ¯F (recall thatx¯∈ri( ¯F)).

The following proposition is a direct consequence ofLemma 2.8which allows to characterize faces ofB1 by an arbitrary convex subset of it.

Proposition 2.9. Let C be a nonempty convex subset of ∂B1. Let ¯x∈ ri(C),

¯

s= sign(Dx), and¯

F¯=B1∩ {x∈Rn:hD¯s, xi= 1}.

Then C ⊂F¯ and ri(C) ⊂ri( ¯F). Moreover, F¯ is the smallest face of B1 such that ri(C)∩F¯ 6=∅and the unique face ofB1 such thatri(C)∩ri( ¯F)6=∅.

Proof. ByLemma 2.8(i) and (iii),C⊂F¯ andri(C)⊂ri( ¯F). LetF be a face such thatri(C)∩F 6=∅. IfF =B1, thenF¯ ⊂F; otherwiseFis an exposed face sinceB1is a polyhedron, and byLemma 2.7,C⊂F. It follows that∅ 6= ri(C)∩ri( ¯F)⊂F∩ri( ¯F), and byLemma 2.7again,F¯⊂F. Suppose now thatri(C)∩ri(F)6=∅. Then permuting F andF, we get that¯ F ⊂F¯, thusF = ¯F.

Remark 2.10. The uniqueness actually holds with the same proof for a general nonempty convex set X: given a nonempty convex subsetC ⊂rbd(X), there exists at most one exposed face Gof X such thatri(C)∩ri(G)6=∅. The existence reduces to the existence, for anyx∈rbd(X), of an exposed faceGofX such thatx∈ri(G).

For a singletonC={x}, the previous proposition becomes the following.¯ Corollary 2.11. Let x¯ ∈ ∂B1 ands¯= sign(Dx). Then¯ F¯ =B1∩ {x∈Rn : hD¯s, xi= 1} is the smallest face of B1 such that x¯ ∈ F¯ and the unique face of B1

such that ¯x∈ri( ¯F).

We can also derive fromProposition 2.9andLemma 2.8 the following properties about the mappingx7→sign(Dx).

Corollary 2.12. Let C be a nonempty convex subset of the unit-sphere ∂B1. Thenmaxx∈Csign(Dx) is well-defined and this maximum is attained everywhere in

6

(8)

ri(C). In particular sign(D·)is constant onri(C).

For anyx, x0∈C,[x, x0]is a nonempty convex subset of∂B1and thussign(D·) is constant on]x, x0[. Moreover, sinces= sign(Dx)ands0= sign(Dx0)are both maxx∈Csign(Dx), they are consistent (seeRemark 2.2). It follows that the constant values00ofsign(D·)on]x, x0[can be given explicitly:

s00i =





si ifsi6= 0, s0i ifs0i6= 0, 0 otherwise.

It is also the maximum ofsign(D·)over[x, x0].

Finally we get a general sufficient condition of extremality.

Corollary 2.13. Let C be a nonempty convex subset of ∂B1. Let x¯ ∈ C and

¯

s= sign(Dx). If¯ x¯ is the uniquex∈C such that sign(Dx)¯s, thenx¯∈ext(C).

Proof. Letx1, x2∈Csuchx¯= x1+x2 2. Then[x1, x2]is a nonempty convex subset of ∂B1. ByCorollary 2.12, sign(Dx) = ¯s for any x∈]x1, x2[. By uniqueness of x,¯ ]x1, x2[={x}, i.e.¯ x1=x2= ¯xand thusx¯ is an extreme point.

Observe that the uniqueness condition in theCorollary 2.13can be written asC∩F¯ = {¯x}withF¯=B1∩ {x∈Rn:hD¯s, xi= 1} the smallest face ofB1 containingx.¯

2.4. Sub-polyhedra of the unit ball. In this section we consider nonempty convex polyhedra of the formA∩B1withAan affine subspace. The results on convex components of the unit sphere apply to such sets if A ∩B1⊂∂B1, and in any case, as we will see, to exposed faces of such polyhedra. Again, the results of this section will hold in particular for exposed faces ofB1.

We begin with a useful lemma.

Lemma 2.14. Let Abe an affine subspace and C be a nonempty convex set such that A ∩ri(C)6=∅. Then

ri(A ∩C) =A ∩ri(C)and dir(A ∩C) = dir(A)∩dir(C).

Proof. Since ri(A) = A, we have ri(A)∩ri(C) 6= ∅. Then by [9, Part III, Proposition 2.1.10], ri(A ∩C) = ri(A)∩ri(C) = A ∩ri(C). Let us now prove that dir(A ∩C) = dir(A)∩dir(C). Letd∈dir(A ∩C)andx¯∈ri(A ∩C)(nonempty). Then

¯

x+εd∈ A ∩Cfor|ε|small, andd∈dir(A)∩dir(C). Similarly, letd∈dir(A)∩dir(F) and x¯ ∈ ri(A)∩ri(C) (nonempty). Then x¯+εd ∈ A ∩C for |ε| small, and d ∈ dir(A ∩C).

The following lemma will be used inTheorem 2.16.

Lemma 2.15. Let A be an affine subspace such that ∅ 6= A ∩B1 ⊂ ∂B1. Let

¯

x∈ A ∩∂B1,s¯= sign(Dx), and¯ J¯= cosupp(Dx). Then¯ (i) dir(A)⊂S

s{d∈Rn:hDs, di ≥0};

(ii) dir(A)∩ P

sR+Ds 6={0};

(iii) dir(A)∩KerDJ¯⊂(Ds)¯, i.e. dir(A)∩(D¯s)∩KerDJ¯= dir(A)∩KerDJ¯. (iv) A ∩ {x∈Rn : sign(Dx)s} ⊂¯ ∂B1, i.e. A ∩∂B1∩ {x∈Rn : sign(Dx)

¯s}=A ∩ {x∈Rn : sign(Dx)¯s};

(v) A ∩ {x∈Rn : sign(Dx) = ¯s} ⊂∂B1, i.e. A ∩∂B1∩ {x∈Rn : sign(Dx) =

¯s}=A ∩ {x∈Rn : sign(Dx) = ¯s}.

(9)

Proof. (i) Suppose that the inclusion does not hold and letd∈dir(A)such that hDs, di<0 for all s¯s. ThenhDs,x¯+εdi<1 for all sand ε >0 small. Indeed, byLemma 2.3, if ss, then¯ hDs,x¯+εdi<1 forε >0; otherwise, hDs,x¯+εdi<1 forεsmall. Still byLemma 2.3,kx¯+εdk1<1forε >0small, i.e. x¯+εd∈ A ∩B˚1, which is in contradiction withA ∩B1⊂∂B1.

(ii) Suppose that the intersection is reduced to{0} and consider the dual cone of both sides of the expression (we denote by C ={y ∈Rn :∀x∈ C,hy, xi ≥0} of a subsetC ⊂Rn). Sincedir(A) and

P

sR+Ds

6={0} are two polyhedral cones, we get that

dir(A)⊥∗+ X

s

R+Ds

=Rn,

where dir(A)⊥∗ = dir(A) and P

sR+Ds

=T

sDs. It follows from (i) that Rn⊂S

sDs, which is not true (note e.g. thatd=x−x¯ withx∈B˚1is such that hDs, di<0)for alls¯s).

(iii) First note that for d ∈ KerDJ¯, hDs, di = hD¯s, di for all s s. Indeed,¯ (Dd)i = 0 if ¯si = 0 and si = ¯si ifs¯i 6= 0, thus si(Dd)i = ¯si(Dd)i for all i and hs, Ddi= h¯s, Ddi. Let now d∈ dir(A)∩KerDJ¯. Since by (i) there exists s s¯ such thathDs, di ≥ 0, it follows that hD¯s, di ≥0. And since −d∈dir(A)∩KerDJ¯ too, we get thathDs, di¯ = 0, which proves the inclusion and the equivalent equality.

(iv) Letx∈ Asuch thatsign(Dx)s¯(in particular,J¯⊂cosupp(Dx)) and let d=x−x. Then¯ d∈dir(A)∩KerDJ¯(recall thatx¯∈ AandJ¯= cosupp(Dx)), and¯ by (ii), d∈ (D¯s). By Lemma 2.3, kDxk1 = h¯s, Dxi =hD¯s,x¯+di= hDs,¯ xi¯ = kDxk¯ 1= 1, which proves the inclusion and the equivalent equality.

(v) The proof is the same as for (iv).

The following theorem is similar toLemma 2.8when we replace convex subset by sub-polyhedra (here of the unit sphere).

Theorem 2.16. Let Abe an affine subspace intersecting ∂B1. Let x¯∈ A ∩∂B1,

¯

s= sign(Dx),¯ J¯= cosupp(D¯x), andF¯ =B1∩ {x∈Rn:hDs, xi¯ = 1}. Then G¯ =A ∩F¯

(i) is the smallest face of A ∩B1 such that x¯ ∈ G¯ and the unique face of A ∩B1

such thatx¯∈ri( ¯G)(G¯ is possibly equal toA ∩B1 itself );

(ii) satisfies the following:

G¯=A ∩∂B1∩ {x∈Rn: sign(Dx)s},¯ ri( ¯G) =A ∩∂B1∩ {x∈Rn: sign(Dx) = ¯s}, dir( ¯G) = dir(A)∩(Ds)¯∩KerDJ¯;

(iii) satisfies the following, in the case where A ∩B1⊂∂B1: G¯=A ∩ {x∈Rn: sign(Dx)s},¯ ri( ¯G) =A ∩ {x∈Rn: sign(Dx) = ¯s}, dir( ¯G) = dir(A)∩KerDJ¯.

Proof. (i) First note thatG¯ = (A ∩B1)∩ {x:hD¯s, xi= 1} is an exposed face of A ∩B1 (in particular it is a convex subset of A ∩B1) and is such that x¯∈ri( ¯G) =

8

(10)

A ∩ri( ¯F)byLemma 2.14. LetGbe a face ofA ∩B1such thatx¯∈G. IfG=A ∩B1, thenG¯⊂G; otherwiseGis an exposed face sinceA ∩B1is a convex polyhedron, and byLemma 2.7, G¯ ⊂G. Suppose now thatx¯ ∈ri(G). Then permutingGandG, we¯ get thatG⊂G, thus¯ G= ¯G.

(ii) By definition,G¯=A∩F¯andx¯∈ A∩ri( ¯F). ByLemma 2.14,ri( ¯G) =A∩ri( ¯F) anddir( ¯G) = dir(A)∩dir( ¯F). The expression of these sets follows fromLemma 2.8.

(iii) We proved the strengthened expression of the previous sets inLemma 2.15.

We get the next result onA ∩B1 itself in the case where it is a subset of∂B1. Proposition 2.17. Let A be an affine subspace such that ∅ 6= A ∩B1 ⊂ ∂B1. Then

A ∩B1=A ∩F¯

withF¯=B1∩ {x∈Rn:hD¯s, xi= 1}ands¯= maxx∈A∩B1sign(Dx)(or equivalently

¯

s= sign(Dx)¯ for somex¯∈ri(A∩B1)). In particular, the results ofTheorem2.16(iii) hold forA ∩B1.

Proof. Since A ∩B1 is a nonempty convex subset of ∂B1, ¯s is well-defined by Corollary 2.12. Let x¯∈ri(A ∩B1)⊂ A ∩∂B1. Then byTheorem 2.16(i), A ∩F¯ is the unique face of A ∩B1 containingx¯in its relative interior; it is thus equal to the faceA ∩B1.

In the general case, we can describe all the exposed faces ofA ∩B1.

Proposition 2.18. Let Abe an affine subspace such that∅ 6=A ∩B16= aff(A ∩ B1). LetGbe a faceA ∩B1 exposed in aff(A ∩B1). Then

G=A ∩F¯

with F¯ =B1∩ {x ∈Rn :hDs, xi¯ = 1} and s¯= maxx∈Gsign(Dx) (or equivalently

¯

s = sign(Dx)¯ for some x¯ ∈ ri(G)). In particular, the results of Theorem 2.16 (ii) (or (iii) ifA ∩B1⊂∂B1) hold forG.

Proof. ByLemma 2.7,G⊂rbd(A∩B1). Let us show thatrbd(A∩B1)⊂ A∩∂B1. We distinguish two cases: if A ∩B1 ⊂ ∂B1, there is nothing to prove; otherwise, A ∩B˚1 6=∅. By Lemma 2.14applied the full-dimensional convexB1, ri(A ∩B1) = ri(A)∩ri(B1) =A∩B˚1. Thenrbd(A∩B1) =A∩B1\A∩B˚1=A∩∂B1. In particular, Gis a nonempty convex subset of∂B1, thus¯sis well-defined byCorollary 2.12.

Letx¯∈ri(G)⊂ A ∩∂B1. Then byTheorem 2.16(i),A ∩F¯ is the unique face of A ∩B1 containingx¯ in its relative interior; it is thus equal to the faceG.

In this setting, we get a necessary and sufficient condition of extremality. Note that the notion of extremality can be related to the topology of the set, here our condition only use an algebraic characterization.

Proposition 2.19. Let A be an affine subspace intersecting ∂B1. Let x¯ ∈ A ∩

∂B1,¯s= sign(Dx), and¯ J¯= cosupp(Dx). Then¯

¯

x∈ext(A ∩B1)⇔dir(A)∩(Ds)¯∩KerDJ¯={0}

⇔dir(A)∩KerDJ¯={0}in the case where A ∩B1⊂∂B1.

Moreover, A ∩B1 admits extreme points (that necessarily belong to A ∩∂B1) if and only if it is compact (i.e. is a convex polytope), if and only if dir(A)∩KerD={0}.

Proof. Since A ∩B1 is a convex polyhedron, x¯ ∈ ext(A ∩B1) ⇔ {¯x} is a face A ∩B1 ⇔ {¯x}= ¯GofTheorem 2.16⇔dir( ¯G) ={0}.

(11)

Recall that extreme points belong torbd(A∩B1), and that as inProposition 2.18, rbd(A ∩B1)⊂ A ∩∂B1. Note also that we always have dir(A)∩KerD⊂dir(A)∩ (D¯s)∩KerDJ¯. Thus ifA∩B1admits extreme points, thendir(A)∩KerD={0}by the beginning of the corollary, which implies thatA∩B1is bounded and thus compact, which in turn implies the existence of extreme points [9, Part III, Proposition 2.3.3].

Remark 2.20. It is possible to show directly (viaCorollary 2.13) that the condi- tion above is sufficient. Indeed, assume that dir(A)∩(D¯s)∩KerDJ¯= {0}. Let x ∈ A ∩∂B1 such that sign(Dx) ¯s; let us show that x = ¯x: x−¯x ∈ dir(A);

by Lemma 2.3, h¯s, Dxi = kDxk1 = kDxk¯ 1 =h¯s, Dxi, thus¯ x−x¯ ∈ (Ds)¯; by Remark 2.2, J¯ ⊂ cosupp(Dx), thus x−x¯ ∈ KerDJ¯. Then ¯s is minimal, and by Corollary 2.13,x¯∈ext(A ∩B1).

2.5. Consequence results on the unit ball. The previous results will be at the core of our study of the solution set of (1.1) insection 3. Nevertheless, we can also dive deeper into this analysis in order to fully characterize the faces of the unit-ball as a byproduct.

A first consequence or reformulation of the previous results is that all the exposed faces ofB1are of the formB1∩ {x∈Rn:hD¯s, xi= 1}.

Proposition 2.21. Let F be an exposed face ofB1. Then F =B1∩ {x∈Rn:hDs, xi¯ = 1}

with s¯ = maxx∈Fsign(Dx) (or equivalently ¯s = sign(Dx)¯ for some x¯ ∈ ri(F)).

Moreover,

F =∂B1∩ {x∈Rn: sign(Dx)s}¯

={x∈Rn:hDs, xi¯ = 1} ∩ {x∈Rn : sign(Dx)s},¯ ri(F) =∂B1∩ {x∈Rn: sign(Dx) = ¯s}

={x∈Rn:hDs, xi¯ = 1} ∩ {x∈Rn : sign(Dx) = ¯s}, dir(F) = (D¯s)∩KerDJ¯(whereJ¯= cosupp(¯s)).

Proof. SinceF is a nonempty convex subset of ∂B1, ¯sis well-defined by Corol- lary 2.12. The first statement is a consequence ofCorollary 2.11. The next statements follow fromLemma 2.8 andProposition 2.17withA={x∈Rn:hD¯s, xi= 1} a sup- porting hyperplane of B1 defining the face F (recall or note by Lemma 2.7 that A ∩B1⊂∂B1for anyAsupporting hyperplane ofB1).

Remark 2.22. Any exposed face F of B1 satisfies KerD ⊂ dir(F) ⊂ (D¯s) with equality if and only if (Ds)¯∩KerDJ¯ = KerD (for the left inclusion) and (D¯s) ⊂KerDJ¯(for the right inclusion).

Together withLemma 2.6, the previous proposition gives the following half-space representations of the exposed faces ofB1.

Corollary 2.23. Let F be an exposed face ofB1. Then

F ={x∈Rn:hD¯s, xi= 1, DJ¯x= 0,diag(¯sI¯)DI¯x≥0}

withs¯= maxx∈Fsign(Dx),J¯= cosupp(¯s), andI¯= supp(¯s).

By Lemma 2.8 and Proposition 2.21, the mapping F 7→ maxx∈Fsign(Dx) is a bijection between the set of exposed faces of B1 and the set of feasible signs {sign(Dx) : x ∈ ∂B1}. The next observation is that this bijection preserves the partial orders (i.e. it is an order isomorphism).

10

(12)

Proposition 2.24. Let F1 and F2 be two exposed faces of B1 and si= maxx∈Fisign(Dx),i= 1,2. Then

F1⊂F2⇔s1s2. In this case, denoting byJ1= cosupp(s1),

F1=F2∩ {x∈Rn :J1⊂cosupp(Dx)}, ri(F1) =F2∩ {x∈Rn :J1= cosupp(Dx)}, dir(F1) = dir(F2)∩KerDJ1.

Proof. Suppose thatF1⊂F2 and letx∈ri(F1)⊂F2; thens1= sign(Dx)s2. Conversely, suppose that s1 s2 and let x ∈ F1; then sign(Dx) s1 s2, thus x∈F2.

Assume now that these two conditions hold. Then, we haveF1⊂F2∩ {x:J1⊂ cosupp(Dx)}sinceF1⊂ {x:J1⊂cosupp(Dx)}(recallRemark 2.2: sign(Dx)s1

implies that J1 ⊂cosupp(Dx)). Conversly, let x∈ F2∩ {x: J1 ⊂cosupp(Dx)}.

Since sign(Dx)and s1 are both s2, they are consistent and thus the converse is true: J1 ⊂ cosupp(Dx) implies that sign(Dx) s1. And since x ∈ F2 ⊂ ∂B1, x∈F1. The same proof holds for ri(F1) and{x:J1 = cosupp(Dx)}. For dir(F1), note thatKerDJ

1 ⊂KerDJ

2 (sinceJ2= cosupp(s2)⊂J1). Thendir(F2)∩KerDJ

1 = (Ds2)∩KerDJ1 (anddir(F1) = (Ds1)∩KerDJ1). By the proof ofLemma 2.15(ii), ford∈KerDJ

1,hDs, di=hDs1, difor allss1. In particular,(Ds1)∩KerDJ

1 = (Ds2)∩KerDJ

1, which concludes the proof.

In the spirit of [6], we consider the compact polyhedron(KerD)∩B1, which is isomorphic to the projection of B1 onto the quotient of the ambient space by the lineality space KerD. Proposition 2.19gives the following necessary and sufficient condition of extremality.

Corollary 2.25. The convex polyhedron (KerD)∩B1 is compact (i.e. is a convex polytope). It admits extreme points that belong to (KerD) ∩∂B1. Given

¯

x∈(KerD)∩∂B1,s¯= sign(Dx), and¯ J¯= cosupp(Dx),¯

¯

x∈ext((KerD)∩B1)⇔(KerD)∩(Ds)¯∩KerDJ¯={0}.

Remark 2.26. In [6, Section 4.1.3], the authors notice that since (KerD)∩B1

and ImD∩B`1 are in bijection through D and its pseudo-inverse (D)+, it holds that

¯

x∈ext((KerD)∩B1)⇔x¯= (D)+θ¯withθ¯∈ext(ImD∩B`1).

Thus we have two conditions of extremality, which are of different nature but of course equivalent, as it can be shown directly. We have already derived inRemark 2.20that if(KerD)∩(Ds)¯∩KerDJ¯={0}, thenx¯∈ext((KerD)∩B1); it follows that θ¯=Dx¯∈ext(ImD∩B`1)and is such that(D)+θ¯= ¯x(recall that(D)+Dis the orthogonal projection ontoImD= (KerD)). Conversely, letθ¯∈ext(ImD∩B`1) and x¯ = (D)+θ¯ (note that Dx¯ = ¯θ since D(D)+ is the orthogonal projection onto ImD). Let d ∈ (KerD)∩(D¯s)∩KerDJ¯. Note that cosupp(¯θ) = ¯J ⊂ cosupp(Dd). Thensign(¯θ+εDd) = ¯sfor|ε|small, and byLemma 2.3,kθ+εD¯ dk1= h¯s,θ¯+εDdi=h¯s,θi¯ +εhD¯s, di =kθk¯ 1, i.e. θ¯+εDd∈B`1 for for |ε| small (and θ¯+εDd∈ImD). By extremality ofθ,¯ Dd= 0, i.e. d∈KerD∩(KerD)={0}, which ends the proof.

(13)

2.6. Testing the extremality. In this subsection, we aim to show that the results of subsection 2.5 can be exploited to numerically test the extremality of a point.

Our first definition formalize the idea of feasible sign, i.e., signs which are attained by some vector in the ambiant space.

Definition 2.27. We say that a signs∈ {−1,0,1}p isfeasiblewith respect toD if there exists x∈Rn such that s= sign(Dx).

Note that we can replace at no cost Rn by B1 or ∂B1 thanks to the homogenity of the the`1-norm. Testing if a sign is feasible has the complexity of a linear program onnvariables withpconstraints:

Lemma 2.28. Letc∈Rn,s∈ {−1,0,1}p,J= cosupp(x),I+={i : hdi, xi>0}

andI={i : hdi, xi<0}. The signs is feasible if, and only if, the solution set of

(2.1) min

x∈Rnhc, xi subject to





DJx = 0 DI+x >+1 DIx 6−1

is non-empty. Moreover, if s is feasible, then any solution x of (2.1) is such that s= sign(Dx).

We defer tosection 3 how the choice of c can leads to interesting properties. Here, we wrote the problem as a linear program to put an emphasis that existing solvers allow us to test this property. Note that finding all feasible signs is quite costly since it needs an exponential (inp) number of linear programs.

Thanks toCorollary 2.25, we have the definition

Definition 2.29. We say that a signs∈ {−1,0,1}p ispre-extremalif it satisfies (KerD)∩(D¯s)∩KerDJ¯={0},

whereJ = cosupp(s), and is extremalif it feasible and pre-extremal.

Checking if a sign is pre-extremal boils down to compute the null-space of the matrix

B = U D¯s DJ¯

where U is a basis of the null-space ofD. In order to find the dimension of KerB, one can use either QR reduction or SVD (Singular Value Decomposition). Here, we used an SVD approach.

We now illustrate these definitions in low dimension (it is known that the study of the number of faces is a very difficult task in general [14,3]), whenDcorrespond to the incidence matrix of a complete graph onn= 4vertices andp= 6 edges. Among 3p = 729possible signs, only 75 are feasible, and among them14 are extremal. We report in Figure 2.1 the pattern of such signs up to centrosymetry of the unit-ball, i.e., we only show 7 of the 14 extremal signs.

We can go further, and consider the sub-poset (of the poset of all signs) of feasi- ble signs and try to represent it with an Hasse diagram associated thanks toProposi- tion 2.24. Figure 2.2depicts such a diagram where purple nodes represent the minimal element, e.g., the extremal signs whereas yellow nodes are maximal element (every feasible sign is a sub-sign of a maximal sign). The construction of the diagram is done by constructing the directed acyclic graph associated to the poset. Note that if

12

(14)

0 1

2

3 [1 1 1 0 0 0]

0 1

2

3 [ 1 1 0 0 -1 -1]

0 1

2

3 [ 1 0 1 -1 0 1]

0 1

2

3 [ 1 0 0 -1 -1 0]

0 1

2

3 [0 1 1 1 1 0]

0 1

2

3 [ 0 1 0 1 0 -1]

0 1

2

3 [0 0 1 0 1 1]

Fig. 2.1: Extremal signs of a complete graph onn= 4vertices. Red edges correspond to positive sign, blue edges to negative sign and gray to 0.

such DAG was known, finding extremal points is reduced to the problem of finding terminal node of the DAG, which can be done by breadth-first search. Such construc- tion is obviously of a pure theoretical interest since it scales exponentially with the dimension.

3. The solution set of sparse analysis regularization. This section is an application of the previous one towards the solution set of (1.1). Remark that a similar analysis can be performed (in a less challenging way) for the noiseless problem (1.2).

In the firstsubsection 3.1, we show that the solution set can be seen as a particular sub-polyhedron of the unit-ball. Using this result, we derive several structural results in subsection 3.2 on the solution set thanks to section 2. Finally, we show that arbitrary sub-polyhedra of the unit ball can be seen as solution set of (1.1) with a specific choice of parameters.

3.1. The solution set as a sub-polyhedron of the unit ball. This section studies the structure of the solution set of (1.1), that we denote by X:

X = argmin

x∈Rn

L(x)def.= 1

2ky−Φxk22+λkDxk1, wherey∈Rq, Φ :RnRq,D:RpRn is linear andλ >0.

Theorem 3.1. The solution set of (1.1)is a nonempty convex polyhedron of the form

X=A ∩Br

withr≥0andAan affine subspace such that∅ 6=A ∩Br⊂∂Branddir(A) = Ker Φ.

Namely, ifx∈X, thenr=kDxk1 andA=x+ Ker Φ.

Proof. It is easy to see that the objective function x7→ L(x) is a nonnegative, convex, closed continuous function with full domain (in particular proper). However, it is not coercive, hence existence of minimizers and compactness of the solution set are not straightforward. Following [16, Chapter 8], the recession coneRLofLis given

(15)

(1, 1, 1, 1, 1, 1)(1, 1, 1, 1, 1, 0) (1, 1, 1, 1, 1, -1)

(1, 1, 1, 1, 0, -1) (1, 1, 1, 1, -1, -1)

(1, 1, 1, 0, 1, 1)

(1, 1, 1, 0, 0, 0) (1, 1, 1, 0, -1, -1) (1, 1, 1, -1, 1, 1)

(1, 1, 1, -1, 0, 1) (1, 1, 1, -1, -1, 1)

(1, 1, 1, -1, -1, 0) (1, 1, 1, -1, -1, -1)

(1, 1, 0, 1, -1, -1)

(1, 1, 0, 0, -1, -1) (1, 1, 0, -1, -1, -1)

(1, 1, -1, 1, -1, -1)

(1, 1, -1, 0, -1, -1) (1, 1, -1, -1, -1, -1) (1, 0, 1, -1, 1, 1)

(1, 0, 1, -1, 0, 1) (1, 0, 1, -1, -1, 1)

(1, 0, 0, -1, -1, 0)

(1, 0, -1, -1, -1, -1) (1, -1, 1, -1, 1, 1)

(1, -1, 1, -1, 0, 1) (1, -1, 1, -1, -1, 1)

(1, -1, 0, -1, -1, 1)

(1, -1, -1, -1, -1, 1) (1, -1, -1, -1, -1, 0)(1, -1, -1, -1, -1, -1) (0, 1, 1, 1, 1, 1)

(0, 1, 1, 1, 1, 0) (0, 1, 1, 1, 1, -1)

(0, 1, 0, 1, 0, -1) (0, 1, -1, 1, -1, -1)(0, 0, 1, 0, 1, 1)

Fig. 2.2: Hasse diagram of the poset of feasible signs of a complete graph on n= 4 vertices.

by

RLdef.= {z∈Rn : L(z)60},

whereLis the recession function ofLgiven by L(z)def.= lim

t→+∞

L(tz)

t ∈R∪ {+∞}.

It is clear that RL is non-negative, hence the recession cone RL is given by RL = {z∈Rn : L(z) = 0}. The lineality spaceLL is the subspace of Rn formed by ele- mentsdsuch thatd∈RLand−d∈RL, i.e,LL=RL∩(−RL). The following lemma characterizes the structure ofRL andLL.

Lemma 3.2. The recession cone RL and the lineality spaceLL ofL are given by RL=LL= KerD∩Ker Φ.

14

(16)

Proof. Lett >0andz∈Rn. We have, 1

tL(tz) = 1

2tky−Φtzk22+λkDzk1

= 1

2tkyk22− hy,Φzi+tkΦzk22+λkDzk1. Hence,

L(z) =−hy,Φzi+λkDzk1Ker Φ(z).

In particular,L(z) = 0if, and only if,z∈KerD∩Ker Φ. SinceKerD∩Ker Φis a subspace, we haveRL=LL.

The polyhedral structure ofX, as we will see, relies on the following lemma.

Lemma 3.3. Let C ⊂Rn be a nonempty convex set. ThenL is constant on C if and only ifΦandkD· k1 are constant onC.

Proof. Assume thatLis constant onCand letx0∈C. Suppose that there exists x1∈ C such thatΦx1 6= Φx0 and let x= x0+x2 1 (note that x∈C). Then by strict convexity ofu7→ ky−uk22,

y−Φx

2 2=

y−(1

2Φx0+1 2Φx1)

2 2

<1 2

y−Φx0

2 2+1

2

y−Φx1

2 2. Together with the convexity inequality of the`1 norm:

Dx 1≤ 1

2 Dx0

1+1

2 Dx1

1,

we get thatL(x)< 12L(x0) + 12L(x1), which is in contradiction with L constant on C. ThenΦis constant onC and thuskD· k1 too. The converse is straightforward.

We add the following lemma, that gives locally the directions where kD· k1 is constant.

Lemma 3.4. Let A be an affine subspace andx¯ ∈ A. Then kD· k1 is constant in a neighborhood of x¯ inAif and only if

dir(A)⊂(D¯s)∩KerDJ¯

wheres¯= sign(Dx)¯ andJ¯= cosupp(Dx).¯

Proof. By definition,kD· k1 is constant in a neighborhood ofx¯inAif and only if there existsε > 0 such thatB(¯x, ε)∩ A ⊂∂Br with r =kDxk¯ 1. We denote by C =B(¯x, ε)∩ A(note that x¯ ∈ri(C)). By Proposition 2.9, C⊂∂Br if and only if C⊂F¯ withF¯ =Br∩ {x∈Rn :hD¯s, xi= 1}ifr >0andF¯= KerD ifr= 0. Since

¯

x∈ri( ¯F),C ⊂F¯ if and only ifdir(C)⊂dir( ¯F). The result follows by noticing that dir(C) = dir(A)(e.g. byLemma 2.14) anddir( ¯F) = (D¯s)∩KerDJ¯(byLemma 2.8 in the caser >0, obvious in the caser= 0).

Together, the previous two lemmas give the following.

Corollary 3.5. Let A be an affine subspace and ¯x∈ A. ThenL is constant in a neighborhood ofx¯ in Aif and only if

dir(A)⊂Ker Φ∩(D¯s)∩KerDJ¯ wheres¯= sign(Dx)¯ andJ¯= cosupp(Dx).¯

(17)

We now go back to the proof ofTheorem 3.1. ByLemma 3.2, the recession cone and the lineality space of Lcoincide. Then by [16, Theorem 27.1(a-b)], the solution set X is nonempty. Since L is convex (and closed), the solution set is also convex (and closed). Moreover,Lis constant onX. Then byLemma 3.3,ΦandkD· k1 are constant onX, i.e.

X ⊂(x+ Ker Φ)∩∂Br

withx∈X andr=kDxk1. But sinceL(x)is the minimum ofL, (x+ Ker Φ)∩Br⊂X.

It follows that

X = (x+ Ker Φ)∩Br

with(x+ Ker Φ)∩Br⊂∂Br, as it was to be proved.

3.2. Consequence results on the solution set. In this section, we apply the results on sub-polyhedra of the unit ball (subsection 2.4) to the solution setX of (1.1).

Proposition 3.6. Let x¯ ∈ ri(X), s¯ = sign(Dx), and¯ F¯ = Br ∩ {x ∈ Rn : hD¯s, xi=r}with r=kDxk¯ 1. Then

X = (¯x+ Ker Φ)∩F .¯

It follows that

X = (¯x+ Ker Φ)∩ {x∈Rn: sign(Dx)s},¯ ri(X) = (¯x+ Ker Φ)∩ {x∈Rn: sign(Dx) = ¯s}, dir(X) = Ker Φ∩KerDJ¯(whereJ¯= cosupp(¯s)).

Moreover, the faces ofX are exactly the sets of the form{x∈X:J ⊂cosupp(Dx)}

with J¯⊂J; their relative interior is given by {x∈X :J = cosupp(Dx)} and their direction byKer Φ∩KerDJ.

Proof. First note that the results are trivial in the case r = 0, as s¯ = 0 and F¯ =Br= KerD. Therefore we consider the case r >0. The first statement follows fromTheorem 3.1andProposition 2.17, and the second one fromTheorem 2.16(iii).

For the last statement, note thatGis a face ofX if and only ifG= (¯x+ Ker Φ)∩F withF a face ofBrsuch thatF ⊂F¯. Indeed, the direct implication holds forG=∅ withF =∅, forG=XwithF = ¯F, and forGexposed inaff(X)byProposition 2.18.

Conversely, letF =∅ or Br∩ H(with Ha supporting hyperplane) be a face ofBr. Then (¯x+ Ker Φ)∩F = ∅ or X∩ H is a face of X. The conclusion follows from Proposition 2.24.

Thanks to Proposition 3.6, we can draw several conclusions. In particular the role ofs¯andJ¯allows us to derive properties of the solution set.

The signs¯is shared by all the interior solutions of (1.1), which are also maximal solutions (¯s= maxx∈Xsign(Dx)). Such a solution can be obtained numerically by the algorithm described in [2]. Future work should include an analysis of the behavior of more common algorithms such as first-order proximal methods.

The knowledge of¯s(or ofJ¯which is the minimal cosupport) gives the dimension of the solution set (1.1) as dim(X) = dim(Ker Φ∩KerDJ¯) (up to determining the

16

Références

Documents relatifs

Representations of linear functionals In this section we prove the following main result of the paper.. Theorem

It is also possible to use a dictionary D of translation invariant wavelets, so that the corresponding regularization term R A can be viewed as a multiscale (higher order)

Local Behavior of Sparse Analysis Regularization: Applications to Risk Estimation... Local Behavior of Sparse

Our main contribution consists in providing a geometrical interpretation of a solution with a maximal D-support, namely the fact that such a solution lives in the relative interior

In this paper we have developed representer theorems for convex regularized inverse problems (1.1), based on fundamental properties of the geometry of convex sets: the solution set

Finally, while these compliance measures are designed to make sense with respect to known results in the area of sparse recovery, one might design other compliance measures tailored

Similar conver- gence results for bound-constrained and linearly constrained optimization algorithms assert that every limit point of the algorithm is stationary, but they do not

Namely, if {L,} is an interpolating sequences of hyperp lanes, then {aj\ is an interpolating sequence of points, so they must satisfy Varopoulos's necessary condition (cf.. the