Skip to main content
Taylor & Francis Open Select logoLink to Taylor & Francis Open Select
. 2018 Apr 10;34(3):489–514. doi: 10.1080/10556788.2018.1457151

Inducing strong convergence into the asymptotic behaviour of proximal splitting algorithms in Hilbert spaces

Radu Ioan Boţ 1,*, Ernö Robert Csetnek 1, Dennis Meier 1
PMCID: PMC6474735  PMID: 31057305

Abstract

Proximal splitting algorithms for monotone inclusions (and convex optimization problems) in Hilbert spaces share the common feature to guarantee for the generated sequences in general weak convergence to a solution. In order to achieve strong convergence, one usually needs to impose more restrictive properties for the involved operators, like strong monotonicity (respectively, strong convexity for optimization problems). In this paper, we propose a modified Krasnosel'skiĭ–Mann algorithm in connection with the determination of a fixed point of a nonexpansive mapping and show strong convergence of the iteratively generated sequence to the minimal norm solution of the problem. Relying on this, we derive a forward–backward and a Douglas–Rachford algorithm, both endowed with Tikhonov regularization terms, which generate iterates that strongly converge to the minimal norm solution of the set of zeros of the sum of two maximally monotone operators. Furthermore, we formulate strong convergent primal–dual algorithms of forward–backward and Douglas–Rachford-type for highly structured monotone inclusion problems involving parallel-sums and compositions with linear operators. The resulting iterative schemes are particularized to the solving of convex minimization problems. The theoretical results are illustrated by numerical experiments on the split feasibility problem in infinite dimensional spaces.

Keywords: fixed points of nonexpansive mappings, Tikhonov regularization, splitting methods, forward–backward algorithm, Douglas–Rachford algorithm, primal–dual algorithm

AMS Subject Classification: 47J25, 47H09, 47H05, 90C25

1. Introduction and preliminaries

Let H be a real Hilbert space endowed with inner product , and associated norm =,. Let T:HH be a nonexpansive mapping, that is TxTyxy for all x,yH. One of the most popular iterative methods for finding a fixed point of the operator T is the Krasnosel'skiĭ–Mann algorithm

xn+1=xn+λnTxnxnn0, (1)

where x0H is arbitrary and (λn)n0 is a sequence of nonnegative real numbers. Provided FixT={xH:Tx=x}, one can show under mild conditions imposed on (λn)n0, that the sequence (xn)n0 converges weakly to an element in FixT (see for instance [6]).

The applications and the impact of this fundamental result go beyond the usual fixed point theory, representing in fact the starting point for the derivation of algorithms and related convergence statements in connection with the solving of monotone inclusions. In this context, we mention the classical forward–backward algorithm for determining a zero of the sum of a set-valued maximally monotone operator and a single-valued and cocoercive one and the Douglas–Rachford algorithm for determining a zero of the sum of two set-valued maximally monotone operators. The paradigms behind these classical methods can be transferred to the solving of convex optimization problems, too (see [6]).

The iterative algorithms mentioned above share the common property that the generated sequences converge weakly to a solution of the problem under investigation. However, for applications where infinite dimensional functional spaces are involved, weak convergence is not satisfactory. In order to achieve strong convergence, one usually needs to impose more restrictive properties for the involved operators, like strong monotonicity when considering monotone inclusions and strong convexity when solving optimization problems. Since there evidently are applications for which these stronger properties are not fulfilled, the interest of the applied mathematics community in developing algorithms which generate iterates that strongly convergence is justified.

We mention in this sense the Halpern algorithm and its numerous variants designed for finding a fixed point of a nonexpansive mapping (see for instance [13]). In the context of solving monotone inclusions, we mention the proximal-Tikhonov algorithm

xn+1=Id+λn(A+μnId)1(xn)n0,

where A:H\rightrightarrowsH is a maximally monotone operator, Id is the identity operator on H and (λn)n0 and (μn)n0 are sequences of nonnegative real numbers. Under mild conditions imposed on (λn)n0 and (μn)n0, one can prove strong convergence of (xn)n0 to the minimal norm solution of the set of zeros of A (see [19,27]). It is important to emphasize the Tikhonov regularization terms (μnId)n0 in the above scheme, which actually enforces the strong convergence property. In the absence of the regularization term, the above numerical scheme becomes the classical proximal algorithm for determining a zero of the operator A, for which in general only weak convergence can be proved (see [23]). For more theoretical results concerning Tikhonov regularization and more motivational arguments for using such techniques, especially for optimization problems, we refer the reader to Attouch's paper [2]. For other techniques and tools in order to achieve strong convergence we mention also the works of Haugazeau [18] and [5].

In this article, we will first introduce and investigate a modified Krasnosel'skiĭ–Mann algorithm with relaxation parameters, having the outstanding property that it generates a sequence of iterates which converges strongly to the minimal norm solution of the fixed points set of a nonexpansive mapping. In contrast to [13,20,27] (see also the references therein), where the techniques and tools used have their roots in fixed point theory results for contractions, our convergence statements follow more directly. Relying on this, we derive a forward–backward and a Douglas–Rachford algorithm, both endowed with Tikhonov regularization terms, which generate iterates that strongly converge to the minimal norm solution of the set of zeros of the sum of two maximally monotone operators. The resulting iterative schemes are particularized to the minimization of the sum of two convex functions.

Furthermore, we deal with complexly structured monotone inclusions where parallel-sums and compositions with linear operators are involved. By making use of modern primal–dual techniques (see [10,12,15,17,28] and also [8,9]), we derive strongly convergent numerical schemes of forward–backward and Douglas–Rachford-type, both involving Tikhonov regularization terms and having the remarkable property that all the operators are evaluated separately. Moreover, the designed algorithms solve both the structured monotone inclusion problem and its dual monotone inclusion problem in the sense of Attouch-Théra (see [3]). When particularized to convex optimization problems, this means the concomitantly solving of a primal problem and its Fenchel dual one. For other types of primal–dual algorithms with strong convergence properties and their applications we refer the reader to [1,4]. Finally, in the last section we carry out numerical experiments on the split feasibility problem in infinite dimensional Hilbert spaces which illustrate the potential of the algorithm endowed with Tikhonov regularization terms.

In the remaining of this section, we recall some results which will play a decisive role in the convergence analysis of the proposed algorithms. The following result is related to the convergence of a sequence satisfying a sharp quasi-Fejér monotonicity property and follows as a direct consequence of [27, Lemma 2.5].

Lemma 1

Let (an)n0 be a sequence of nonnegative real numbers satisfying the inequality

an+1(1θn)an+θnbn+ϵnn0,

where

  1. 0θn1 for all n0 and n0θn=+;

  2. lim supn+bn0;

  3. ϵn0 for all n0 and n0ϵn<+.

Then the sequence (an)n0 converges to 0.

We close this section with a result that is a consequence of the demiclosedness principle (see [6, Corollary 4.18]) and it will be used in the proof of Theorem 3, which is the main result of this paper.

Lemma 2

Let T:HH be a nonexpansive operator and let (xn)n0 be a sequence in H and xH be such that wlimn+xn=x and (Txnxn)n0 converges strongly to 0 as n+. Then xFixT.

2. A strongly convergent Krasnosel'skiĭ–Mann algorithm

In order to induce strong convergence into the asymptotic behaviour of the Krasnosel'skiĭ–Mann algorithm for determining a fixed point of a nonexpansive mapping T:HH, we propose the following modified version of it:

xn+1=βnxn+λnT(βnxn)βnxnn0, (2)

where x0H is the starting point and (λn)n0 and (βn)n0 suitably chosen sequences of positive numbers. In the proof of the theorem below, we denote by projC:HC,projC(x)=argmincCxc, the projection operator onto the nonempty closed convex set CH. We notice that for a nonexpansive mapping T:HH, its set of fixed points FixT is closed and convex (see [6, Corollary 4.15]).

Theorem 3

Let (λn)n0 and (βn)n0 be real sequences satisfying the conditions:

  1. 0<βn1 for any n0,limn+βn=1,n0(1βn)=+ and n1|βnβn1|<+;

  2. 0<λn1 for any n0,lim infn+λn>0 and n1|λnλn1|<+.

Consider the iterative scheme (2) with and arbitrary starting point x0H and a nonexpansive mapping T:HH fulfilling FixT. Then (xn)n0 converges strongly to projFixT(0).

Proof.

For the beginning, we prove that (xn)n0 is bounded. Let xFixT. Due to the nonexpansiveness of T, we have for any n0:

xn+1x=(1λn)(βnxnx)+λnT(βnxn)Tx(1λn)βnxnx+λnT(βnxn)Txβnxnx=βn(xnx)+(βn1)xβnxnx+(1βn)x.

A simple induction leads to the inequality

xnxmax{x0x,x}n0,

hence (xn)n0 is bounded.

We claim that

xn+1xn0asn+. (3)

Indeed, by taking into account that T is nonexpansive and that (xn)n0 is bounded, we obtain for any n1 the following estimates:

xn+1xn=(1λn)βnxn(1λn1)βn1xn1+λnT(βnxn)λn1T(βn1xn1)(1λn)(βnxnβn1xn1)+(λn1λn)βn1xn1+λnT(βnxn)T(βn1xn1)+(λnλn1)T(βn1xn1)βnxnβn1xn1+|λnλn1|C1,

where C1>0.

Further, we derive for any n1:

xn+1xnβn(xnxn1)+(βnβn1)xn1+|λnλn1|C1βnxnxn1+|βnβn1|C2+|λnλn1|C1,

where C2>0. Statement (3) is a consequence of Lemma 1, for an:=xnxn1,bn:=0,ϵn:=|βnβn1|C2+|λnλn1|C1 and θn:=1βn,n1.

In the following we prove that

xnTxn0asn+. (4)

For any n0 we have the following inequalities:

xnTxnxn+1xn+xn+1Txn=xn+1xn+(1λn)(βnxnTxn)+λnT(βnxn)Txnxn+1xn+(1λn)βnxnTxn+λnβnxnxnxn+1xn+(1λn)βnxnβnTxn+(1λn)βnTxnTxn+λn(1βn)xnxn+1xn+(1λn)xnTxn+(1λn)(1βn)Txn+λn(1βn)xn.

From here we deduce that for any n0:

λnxnTxnxn+1xn+(1λn)(1βn)Txn+λn(1βn)xn.

Taking into account that (xn)n0 is bounded, (3) and the properties of the sequences involved, we derive from the last inequality that (4) holds.

In what follows we show that (xn)n0 actually converges strongly to projFixT(0):=x¯. Since T is nonexpansive, we have for any n0:

xn+1x¯=(1λn)(βnxnx¯)+λnT(βnxn)Tx¯(1λn)βnxnx¯+λnT(βnxn)Tx¯βnxnx¯.

Hence,

xn+1x¯2βnxnx¯2=βn(xnx¯)+(βn1)x¯2=βn2xnx¯2+2βn(1βn)x¯,xnx¯+(1βn)2x¯2βnxnx¯2+(1βn)2βnx¯,xnx¯+(1βn)x¯2n0. (5)

Next we show that

lim supn+x¯,xnx¯0. (6)

Assuming the contrary, there would exist a positive real number l and a subsequence (xkj)j0 such that

x¯,xkjx¯l>0j0.

Due to the boundedness of the sequence (xn)n0, we can assume without losing the generality that (xkj)j0 weakly converges to an element yH. According to Lemma 2, by taking into consideration (4), it follows that yFixT. From this and the variational characterization of the projection we easily derive

limj+x¯,xkjx¯=x¯,yx¯0,

which leads to a contradiction. This shows that (6) holds. Thus

lim supn+2βnx¯,xnx¯+(1βn)x¯20.

A direct application of Lemma 1 to (5), for an:=xnx¯2,bn:=2βnx¯,xnx¯+(1βn)x¯2,ϵn:=0 and θn:=1βn, n0, delivers the desired conclusion.

Remark 4

Condition (ii) in the previous theorem is satisfied by every monotonically increasing (and, in consequence, convergent) sequence (λn)n0(0,1] and also by every monotonically decreasing (and, in consequence, convergent) sequence (λn)n0(0,1] having as limit a positive number.

Condition (i) in the previous theorem is satisfied by every monotonically increasing sequence (βn)n0(0,1] which fulfils limn+βn=1 and n0(1βn)=+, as it is for instance the sequence with β0(0,12) and βn=11/(n+1) for any n1.

An immediate consequence of Theorem 3 is the following corollary, which proposes an iterative scheme that finds a minimal norm solution of the set of fixed points of an averaged operator. Let α(0,1) be fixed. We say that R:HH is an α-averaged operator if there exists a nonexpansive operator T:HH such that R=(1α)Id+αT. It is obvious that α-averaged operators are also nonexpansive. The 12-averaged operators are nothing else than the firmly nonexpansive ones and form the most important representatives of this class. For properties and insights into these families of operators we refer the reader to [6]. The following result will play in the next section a determinant role in the converge analysis of the forward–backward method endowed with Tikhonov regularization term.

Corollary 5

Consider the iterative scheme

xn+1=βnxn+λnR(βnxn)βnxnn0, (7)

with x0H as starting point, R:HH an α-averaged operator, for α(0,1), such that FixR and (λn)n0 and (βn)n0 real sequences satisfying the conditions:

  1. 0<βn1 for any n0,limn+βn=1,n0(1βn)=+ and n1|βnβn1|<+;

  2. 0<λn1/α for any n0,lim infn+λn>0 and n1|λnλn1|<+. Then (xn)n0 converges strongly to projFixR(0).

Proof.

Since R is α-averaged, there exists a nonexpansive operator T:HH such that R=(1α)Id+αT. The conclusion follows from Theorem 3, by taking into account that (7) is equivalent to

xn+1=βnxn+αλnT(βnxn)βnxnn0

and that FixR=FixT.

3. A forward–backward algorithm with Tikhonov regularization term

This section is dedicated to the formulation and convergence analysis of a forward–backward algorithm with Tikhonov regularization terms, which generates a sequence of iterates that converges strongly to the minimal norm solution of the set of zeros of the sum of two maximally monotone operators, one of them being single-valued.

For readers' convenience, we recall some standard notions and results in monotone operator theory which will be used in the following (see also [6,7,26]). For an arbitrary set-valued operator A:H\rightrightarrowsH we denote by GrA={(x,u)H×H:uAx} its graph. Then A1:H\rightrightarrowsH, which is the operator with GrA1={(x,u)H×H:xAu}, denotes the inverse operator of A. We use also the notation zerA={xH:0Ax} for the set of zeros of A. We say that A is monotone, if xy,uv0 for all (x,u),(y,v)GrA. A monotone operator A is said to be maximally monotone, if there exists no proper monotone extension of the graph of A on H×H. The resolvent of A, JA:H\rightrightarrowsH, is defined by

JA=(Id+A)1,

where Id:HH,Id(x)=x for all xH, is the identity operator on H. Moreover, if A is maximally monotone, then JA:HH is single-valued and maximally monotone (see [6, Proposition 23.7 and Corollary 23.10]).

Let γ>0. We say that B:HH is γ-cocoercive, if xy,BxByγBxBy2 for all x,yH.

The following technical result (see [21, Theorem 3(b)] and [16, Proposition 2.4]) gives an expression for the averaged parameter of the composition of two averaged operators. We refer also to [6, Proposition 4.32] for other results of this type.

Proposition 6

Let Ti:HH be αi-averaged, where αi(0,1),i=1,2. Then the composition T1T2 is α-averaged, where

α=α1+α22α1α21α1α2(0,1).

Theorem 7

Let A:H\rightrightarrowsH be a maximally monotone operator and B:HH a β-cocoercive operator, for β>0, such that zer(A+B). Let γ(0,2β]. Consider the iterative scheme

xn+1=(1λn)βnxn+λnJγAβnxnγB(βnxn)n0, (8)

with x0H as starting point and (λn)n0 and (βn)n0 real sequences satisfying the conditions:

  1. 0<βn1 for any n0,limn+βn=1,n0(1βn)=+ and n1|βnβn1|<+;

  2. 0<λn(4βγ)/2β for any n0,lim infn+λn>0 and n1|λnλn1|<+.

Then (xn)n0 converges strongly to projzer(A+B)(0).

Proof.

It is immediate that the iterative scheme (8) can be written in the form

xn+1=βnxn+λnT(βnxn)βnxnn0,

where T=JγA(IdγB).

We consider two cases. The first one is when γ(0,2β).

According to [6, Corollary 23.8 and Remark 4.24(iii)], JγA is 12-cocoercive. Moreover, by [6, Proposition 4.33], IdγB is γ/2β-averaged. Combining this with Proposition 6, we derive that T is 2β/(4βγ)-averaged. The statement follows now from Corollary 5, by noticing that FixT=zer(A+B) (see [6, Proposition 25.1(iv)]).

The second case is when γ=2β. The cocoercivity of B implies that IdγB is nonexpansive, hence the operator T=JγA(IdγB) is nonexpansive, too, the conclusion follows in this situation from Theorem 3.

Remark 8

The choice λn=1 for any n0 in the previous theorem leads to the iterative scheme

xn+1=JγAβnxnγB(βnxn)n0,

which further becomes in case B=0

xn+1=JγAβnxnn0.

This last relation can be equivalently written as

xn1βnxn+1+γβnAxn+1=Id+ϵnId+γβnA(xn+1),

where ϵnId (with ϵn:=1/βn1>0 and limn+ϵn=0) represents the Tikhonov regularization term, which enforces the strong convergence of the sequence (xn)n0 to the minimal norm solution. For other types of Tikhonov-like methods for monotone inclusion problems we refer the reader to [19,25,27] and the references therein.

In the remaining of this section we turn our attention to the solving of optimization problems of the form

minxH{f(x)+g(x)}, (9)

where f:HR{+} is a proper, convex and lower semicontinuous function and g:HR is a convex and Fréchet differentiable function with 1/β-Lipschitz continuous gradient, for β>0.

For a proper, convex and lower semicontinuous function f:HR{+}, its (convex) subdifferential at xH is defined as

f(x)={uH:f(y)f(x)+u,yx yH},

for xH with f(x)=+ and as f(x)=, otherwise. When seen as a set-valued mapping, the convex subdifferential is a maximally monotone operator (see [22]) and its resolvent is given by Jf=proxf (see [6]), where proxf:HH,

proxf(x)=argminyHf(y)+12yx2, (10)

denotes the proximal operator of f.

Corollary 9

Let f:HR{+} be a proper, convex and lower semicontinuous function and g:HR a convex and Fréchet differentiable function with 1/β-Lipschitz continuous gradient, for β>0, such that argminxH{f(x)+g(x)}. Let γ(0,2β]. Consider the iterative scheme

xn+1=(1λn)βnxn+λnproxγfβnxnγg(βnxn)n0, (11)

with x0H as starting point and (λn)n0 and (βn)n0 real sequences satisfying the conditions:

  1. 0<βn1 for any n0,limn+βn=1,n0(1βn)=+ and n1|βnβn1|<+;

  2. 0<λn(4βγ)/2β for any n0,lim infn+λn>0 and n1|λnλn1|<+.

Then (xn)n0 converges strongly to the minimal norm solution of (9).

Proof.

The statement is a direct consequence of Theorem 7, by choosing A:=f and B:=g and by taking into account that

zer(f+g)=argminxH{f(x)+g(x)}

and the fact that g is β-cocoercive due to the Baillon–Haddad Theorem (see [6, Corollary 18.16]).

4. A Douglas–Rachford algorithm with Tikhonov regularization term

In this section we derive from the Krasnosel'skiĭ–Mann algorithm formulated in Section 2 an iterative scheme of Douglas–Rachford-type, which generates sequences that strongly converge to a zero of of the sum of two set-valued maximally monotone operators.

In what follows, we denote by RA=2JAId the reflected resolvent of a maximally monotone operator A:H\rightrightarrowsH.

Theorem 10

Let A,B:H\rightrightarrowsH be two maximally monotone operators such that zer(A+B) and γ>0. Consider the following iterative scheme:

(n0)yn=JγB(βnxn)zn=JγA(2ynβnxn)xn+1=βnxn+λn(znyn)

with x0H as starting point and (λn)n0 and (βn)n0 real sequences satisfying the conditions:

  1. 0<βn1 for any n0,limn+βn=1,n0(1βn)=+ and n1|βnβn1|<+;

  2. 0<λn2 for any n0,lim infn+λn>0 and n1|λnλn1|<+.

Then the following statements are true:

  1. (xn)n0 converges strongly to x¯:=projFixRγARγB(0) as n+;

  2. (yn)n0 and (zn)n0 converge strongly to JγB(x¯)zer(A+B) as n+.

Proof.

Taking into account the iteration rules and the definition of the reflected resolvent, the iterative scheme in the enunciation of the theorem can be equivalently written as

xn+1=βnxn+λnJγA(2JγBId)(βnxn)JγB(βnxn)=βnxn+λnId+RγA2RγB(βnxn)Id+RγB2(βnxn)=βnxn+λn2T(βnxn)βnxnn0, (12)

where T:=RγARγB:HH is a nonexpansive operator (see [6, Corollary 23.10(ii)]). From [6, Proposition 25.1(ii)] we have zer(A+B)=JγB(FixT), hence FixT. By applying Theorem 3, we obtain that (xn)n0 converges strongly to x¯:=projFixT(0) as n+, hence (a) holds.

Further, by taking into account the definition of the sequence (yn)n0 and the continuity of the resolvent operator, we obtain that (yn)n0 converges strongly to JγBx¯zer(A+B) as n+. Finally, by taking the limit in the recursive formula of the sequence (xn)n0, we obtain that znyn converges strongly to 0 as n+, thus (b) holds, too.

Remark 11

The classical Douglas–Rachford method, which reads

(k0)yn=JγB(xn)zn=JγA(2ynxn)xn+1=xn+λn(znyn),

produces sequences for which in general only weak convergence to a zero of the A+B can be proved (see e.g. [6, Theorem 25.6]). In order to ensure strong convergence, one usually needs to impose restrictive conditions on the monotone operators involved, like uniform monotonicity (which is a generalization of strong monotonicity). This is not the case for the iterative scheme stated in Theorem 10, where we are able to guarantee strong convergence in the very general situation of maximally monotone operators.

Further, we look at optimization problems of the form

minxH{f(x)+g(x)}, (13)

where f,g:HR{+} are proper, convex and lower semicontinuous functions. We denote by

dom f={xH:f(x)<+}

the effective domain of the function f.

In order to proceed, we need the following notion. For SH a convex set, we denote by

sqri S:={xS:λ>0λ(Sx) is a closed linear subspace of H}

its strong quasi-relative interior. Notice that we always have intSsqriS (in general this inclusion may be strict). If H is finite-dimensional, then sqriS coincides with riS, the relative interior of S, which is the interior of S with respect to its affine hull. The notion of strong quasi-relative interior belongs to the class of generalized interiority notions and plays an important role in the formulation of regularity conditions, which are needed in convex optimization in order to guarantee duality results and also subdifferential sum formulas. The one considered in the next result is the so-called Attouch-Brézis regularity condition. We refer to [6,7,26,29] for more interiority notions and their impact on the duality theory.

Corollary 12

Let f,g:HR{+} be proper, convex and lower semicontinuous functions such that argminxH{f(x)+g(x)} and 0sqri(domfdomg) and γ>0. Consider the following iterative scheme:

(n0)yn=proxγg(βnxn)zn=proxγf(2ynβnxn)xn+1=βnxn+λn(znyn)

with x0H as starting point and (λn)n0 and (βn)n0 real sequences satisfying the conditions:

  1. 0<βn1 for any n0,limn+βn=1,n0(1βn)=+ and n1|βnβn1|<+;

  2. 0<λn2 for any n0,lim infn+λn>0 and n1|λnλn1|<+.

Then the following statements are true:

  1. (xn)n0 converges strongly to x¯:=projFixT(0) as n+, where T=(2proxγfId)(2proxγgId);

  2. (yn)n0 and (zn)n0 converge strongly to proxγg(x¯)argminxH{f(x)+g(x)} as n+.

Proof.

The result is a direct consequence of Theorem 10 for A=f and B=g and by noticing that the regularity condition 0sqri(dom fdom g) ensures the relation (see [6, Proposition 7.2])

zer(f+g)=argminxH{f(x)+g(x)}.

5. Strongly convergent primal–dual algorithms

The aim of this section is to induce strong convergence in the nowadays so popular primal–dual algorithms designed for solving highly structured monotone inclusions involving parallel-sums and compositions with linear operators.

5.1. A primal–dual algorithm of forward–backward-type with Tikhonov regularization terms

In this subsection, the following monotone inclusion problem will be in the focus of our investigations.

Problem 13

Let A:H\rightrightarrowsH be a maximally monotone operator and C:HH a μ-cocoercive operator, for some μ>0. Let m be a strictly positive integer and for any i=1,,m, let Gi be a real Hilbert space, Bi,Di:Gi\rightrightarrowsGi be maximally monotone operators such that Di are νi-strongly monotone, for some νi>0, and Li:HGi be a nonzero linear continuous operator. The problem is to solve the primal inclusion

find x¯Hsuch that 0Ax¯+i=1mLi(BiDi)(Lix¯)+Cx¯, (14)

together with the dual inclusion of Attouch-Théra type (see [3,15,28])

 find v¯1G1,,v¯mGmsuch that xH:i=1mLiv¯iAx+Cxv¯i(BiDi)(Lix),i=1,,m. (15)

Some of the notations used above are to be specified. The operator Li:GiH, defined via Lix,y=x,Liy for all xH and all yGi, denotes the adjoint of the linear continuous operator Li:HGi, for i=1,,m. We say that Di:Gi\rightrightarrowsGi is νi-strongly monotone, for some νi>0, if xy,uvνixy2 for all (x,u),(y,v)GrDi, i=1,,m. The parallel sum of the set-valued operators Bi, Di:Gi\rightrightarrowsGi is defined as BiDi:Gi\rightrightarrowsGi, BiDi=(Bi1+Di1)1, for i=1,,m.

We say that (x¯,v¯1,,v¯m)H×G1××Gm is a primal–dual solution to Problem 13, if

i=1mLiv¯iAx¯+Cx¯andv¯i(BiDi)(Lix¯),i=1,,m. (16)

It is easy to see that, if (x¯,v¯1,,v¯m)H×G1××Gm is a primal–dual solution to Problem 13, then x¯ is a solution to (14) and (v¯1,,v¯m)G1××Gm is a solution to (15). Moreover, if x¯H is a solution to (14), then there exists (v¯1,,v¯m)G1××Gm such that (x¯,v¯1,,v¯m) is a primal–dual solution to Problem 13 and, if (v¯1,,v¯m)G1××Gm is a solution to (15), then there exists x¯H such that (x¯,v¯1,,v¯m) is a primal–dual solution to Problem 13.

Theorem 14

In Problem 13, suppose that

0ranA+i=1mLi(BiDi)Li+C. (17)

Let τ and σi,i=1,,m, be strictly positive numbers such that

2min{τ1,σ11,,σm1}min{μ,ν1,,νm}1τi=1mσiLi21.

Let be the starting point (x0,v1,0,,vm,0)H×G1××Gm and set:

n0pn=JτAβnxnτβni=1mLivi,n+C(βnxn)xn+1=βnxn+λn(pnβnxn)For i=1,,mqi,n=JσiBi1βnvi,n+σiLi(2pnβnxn)Di1(βnvi,n)vi,n+1=βnvi,n+λn(qi,nβnvi,n) (18)

where (λn)n0 and (βn)n0 are real sequences satisfying the conditions:

  1. 0<βn1 for any n0,limn+βn=1,n0(1βn)=+ and n1|βnβn1|<+;

  2. 0<λn(4βρ1)/2βρ for any n0,lim infn+λn>0 and n1|λnλn1|<+,

for

β=min{μ,ν1,,νm}

and

ρ=min{τ1,σ11,,σm1}1τi=1mσiLi2.

Then there exists a primal–dual solution (x¯,v¯1,,v¯m) to Problem 13 such that the sequence of primal–dual iterates (xn,v1,n,,vm,n) converges strongly to (x¯,v¯1,,v¯m) as n+.

Remark 15

  1. Since Di:Gi\rightrightarrowsGi is νi-strongly monotone, we have that Di1:GiGi is νi-cocoercive, for i=1,,m.

  2. The resolvent of the inverse operator of a maximally monotone operator M:H\rightrightarrowsH can be computed as follows (see [6]):

Id=JγM+γJγ1M1γ1Id. (19)

Proof.

The idea is to apply Theorem 7 in an appropriate product space under the use of appropriate renorming techniques (see [28]). We consider the Hilbert space K=H×G1××Gm endowed with inner product and associated norm defined, for (x,v1,,vm), (y,q1,,qm)K, via

(x,v1,,vm),(y,q1,,qm)K=x,yH+i=1mvi,qiGiand(x,v1,,vm)K=xH2+i=1mviGi2, (20)

respectively. Furthermore, we consider the set-valued operator

M:K\rightrightarrowsK,(x,v1,,vm)(Ax,B11v1,,Bm1vm),

which is maximally monotone, since A and Bi, i=1,,m, are maximally monotone (see [6, Propositions 20.22 and 20.23]), and the linear continuous operator

S:KK,(x,v1,,vm)i=1mLivi,L1x,,Lmx,

which is skew-symmetric (i.e. S=S) and hence maximally monotone (see [6, Example 20.30]). We also consider the single-valued operator

Q:KK,(x,v1,,vm)Cx,D11v1,,Dm1vm,

which is once again maximally monotone, since C and Di are maximally monotone for i=1,,m. Therefore, since dom S=K, M+S is maximally monotone (see [6, Corollary 24.4(i)]). According to [28, page 672]

Q is βcocoercive.

Further, one can easily verify that (17) is equivalent to zer(M+S+Q) and (see also [15, page 317])

(x,v1,,vm)zer M+S+Q(x,v1,,vm) is a primal--dual solution to Problem 13. (21)

We also introduce the linear continuous operator

V:KK,(x,v1,,vm)xτi=1mLivi,v1σ1L1x,,vmσmLmx,

which is self-adjoint and ρ-strongly positive (see [28]), namely, the following inequality holds

x,VxKρxK2xK.

Therefore, its inverse operator V1 exists and it fulfils V11/ρ.

The algorithmic scheme (18) in the statement of the theorem can be written by using this notations as

n0βnτ1xni=1mLivi,nτ1pn+i=1mLiqi,nC(βnxn)Apn+i=1mLiqi,nxn+1=βnxn+λn(pnβnxn)For i=1,,mβn(σi1vi,nLixn)σi1qi,n+LipnDi1(βnvi,n)Bi1(qi,n)Lipnvi,n+1=βnvi,n+λn(qi,nβnvi,n), (22)

By introducing the sequences

xn=(xn,v1,n,,vm,n),andyn=(pn,q1,n,,qm,n)n0,

the scheme (22) can equivalently be written in the form

n0βnV(xn)V(yn)Q(βnxn)M+S(yn)xn+1=βnxn+λnynβnxn. (23)

Furthermore, we have for any n0

βnV(xn)V(yn)Q(βnxn)M+S(yn)βnVQ(βnId)(xn)(M+S+V)(yn)yn=(M+S+V)1βnVQ(βnId)(xn)yn=Id+V1(M+S)1βnIdV1Q(βnId)(xn)yn=JAβnxnB(βnxn),

where

A:=V1M+SandB:=V1Q. (24)

Let KV be the Hilbert space with inner product and norm defined, for x,yK, by

x,yKV=x,VyKandxKV=x,VxK, (25)

respectively. As the set-valued operators M+S and Q are maximally monotone on K, the operators A and B are maximally monotone on KV (see also [28]). Furthermore, B is βρ-cocoercive on KV. Moreover, since V is self-adjoint and ρ-strongly positive, weak and strong convergence in KV are equivalent with weak and strong convergence in K, respectively.

Taking this into account, it follows that (23) becomes

n0 xn+1=βnxn+λnJAβnxnB(βnxn)βnxn,

which is the algorithm presented in Theorem 7 for determining the zeros of A+B in case γ=1. However, we have

zer(A+B)=zer(V1M+S+Q)=zer(M+S+Q).

According to Theorem 7, xn converges strongly to projzer(A+B)(0,0,,0) in the space KV as n+ and the conclusion follows from (21).

In the remaining of this subsection we investigate the convergence property of the algorithm (18) in the context of simultaneously solving complexly structured convex optimization problems and their Fenchel duals. The problem under investigation is the following one.

Problem 16

Let fΓ(H) and h:HR be a convex and differentiable function with a μ1-Lipschitz continuous gradient, for some μ>0. Let m be a strictly positive integer and for i=1,,m, let Gi be a real Hilbert space, gi,liΓ(Gi) such that li is νi-strongly convex, for some νi>0 and Li:HGi a nonzero linear continuous operator. Consider the convex optimization problem

infxHf(x)+i=1m(gili)(Lix)+h(x) (26)

and its Fenchel-type dual problem

supviGi, i=1,,mfhi=1mLivii=1mgi(vi)+li(vi). (27)

We denote by Γ(H) the set of proper, convex and lower semicontinuous functions defined on H with values in the extended real line R{+}. The conjugate of a function f is f:HR¯, f(p)=sup{p,xf(x):xH} for all pH. Moreover, if fΓ(H), then fΓ(H), as well, and (f)1=f. Finally, having two proper functions f,g:HR¯, their infimal convolution is defined by fg:HR¯, (fg)(x)=infyH{f(y)+g(xy)} for all xH.

Corollary 17

In Problem 16, suppose that

0ranf+i=1mLi(gili)Li+h. (28)

Let τ and σi,i=1,,m, be strictly positive numbers such that

2min{τ1,σ11,,σm1}min{μ,ν1,,νm}1τi=1mσiLi21.

Let be the starting point (x0,v1,0,,vm,0)H×G1××Gm and set:

n0pn=proxτfβnxnτβni=1mLivi,n+h(βnxn)xn+1=βnxn+λn(pnβnxn)For i=1,,mqi,n=proxσigiβnvi,n+σiLi(2pnβnxn)li(βnvi,n)vi,n+1=βnvi,n+λn(qi,nβnvi,n), (29)

where (λn)n0 and (βn)n0 are real sequences satisfying the conditions:

  1. 0<βn1 for any n0,limn+βn=1,n0(1βn)=+ and n1|βnβn1|<+;

  2. 0<λn(4βρ1)/2βρ for any n0,lim infn+λn>0 and n1|λnλn1|<+,

for

β:=min{μ,ν1,,νm}

and

ρ:=min{τ1,σ11,,σm1}1τi=1mσiLi2.

Then there exists (x¯,v¯1,,v¯m)H×G1××Gm such that (xn,v1,n,,vm,n) converges strongly to (x¯,v¯1,,v¯m) as n+ and x¯ is an optimal solution of the problem (26), (v¯1,,v¯m) is an optimal solution of (27) and the optimal objective values of the two optimization problems coincide.

Remark 18

The proximal-point operator of the conjugate function can be computed via the Moreau's decomposition formula

proxγf+γprox(1/γ)fγ1Id=Id, (30)

which is valid for γ>0 and fΓ(H) (see [6]).

Proof.

Consider the maximal monotone operators

A=f,C=h,Bi=giandDi=li,i=1,,m.

The Baillon–Haddad Theorem (see [6, Corollary 18.16]) ensures that C is μ-cocoercive. Since li is νi-strongly convex, Di is νi-strongly monotone, for i=1,,m. According to [6, Proposition 17.10, Theorem 18.15], Di1=li is a monotone and νi1-Lipschitz continuous operator for i=1,,m. The strong convexity of the functions li guarantees that giliΓ(Gi) (see [6, Corollary 11.16, Proposition 12.14]) and (gili)=gili, i=1,,m, (see [6, Proposition 24.27]).

Hence, the monotone inclusion problem (14) reads

find x¯Hsuch that 0f(x¯)+i=1mLi(gili)(Lix¯)+h(x¯), (31)

while the dual monotone inclusion problem (15) reads

 find v¯1G1,,v¯mGmsuch that xH: i=1mLiv¯if(x)+h(x)v¯i(gili)(Lix),i=1,,m. (32)

One can see that if (x¯,v¯1,,v¯m)H×G1××Gm is a primal–dual solution to (31)–(32), namely,

i=1mLiv¯if(x¯)+h(x¯)andv¯i(gili)(Lix¯),i=1,,m, (33)

then x¯ is an optimal solution of the problem (26), (v¯1,,v¯m) is an optimal solution of (27) and the optimal objective values of the two problems coincide. Notice that (33) is nothing else than the system of optimality conditions for the primal–dual pair of convex optimization problems (26)–(27).

The conclusion follows now from Theorem 14.

Remark 19

(i) The relation (28) in the above theorem is fulfilled if the primal problem (26) has an optimal solution x¯H and a suitable regularity condition holds. Under these auspices there exists an optimal solution to (27) (v¯1,,v¯m)G1××Gm, such that (x¯,v¯1,,v¯m) satisfies the optimality conditions (33) and, consequently, (28) holds.

(ii) Further, let us discuss some conditions ensuring the existence of a primal optimal solution. Suppose that the primal problem (26) is feasible, which means that its optimal objective value is not identical +. The existence of optimal solutions for (26) is guaranteed if, for instance, f+h is coercive (that is limx(f+h)(x)=+) and for all i=1,,m, gi is bounded from below. Indeed, under these circumstances, the objective function of (26) is coercive (use also [6, Corollary 11.16 and Proposition 12.14] to show that for all i=1,,m, gili is bounded from below and giliΓ(Gi)) and the statement follows via [6, Corollary 11.15]. On the other hand, if f+h is strongly convex, then the objective function of (26) is strongly convex, too, thus (26) has a unique optimal solution (see [6, Corollary 11.16]).

(iii) We discuss at this point a suitable regularity condition as mentioned at item (i) above. Since dom(gili)=domgi+domli, i=1,,m, one can use to this end the regularity condition of interiority-type (see also [15])

(0,,0)sqrii=1m(dom gi+dom li){(L1x,,Lmx):xdom f}. (34)

This is fulfilled provided that one of the following conditions is verified (see [15, Proposition 4.3]):

  1. domgi+domli=Gi, i=1,,m;

  2. H and Gi are finite-dimensional and there exists xri dom f such that Lixriri dom gi+ri dom li, i=1,,m.

5.2. A primal–dual algorithm of Douglas–Rachford-type with Tikhonov regularization terms

The problem that we investigate in this section reads as follows.

Problem 20

Let A:H\rightrightarrowsH be a maximally monotone operator. Let m be a strictly positive integer and for any i=1,,m, let Gi be a real Hilbert space, Bi, Di:Gi\rightrightarrowsGi be maximally monotone operators and Li:HGi a nonzero linear continuous operator. The problem is to solve the primal inclusion

find x¯Hsuch that 0Ax¯+i=1mLi(BiDi)(Lix¯) (35)

together with the dual inclusion

find v¯1G1,,v¯mGmsuch that (xH)i=1mLiv¯iAxv¯i(BiDi)(Lix),i=1,,m. (36)

Different to Problem 13, the operators Di, i=1,,m are general maximally monotone operators, thus they will have to be addressed through their resolvents. This is why in this context a primal–dual algorithm relying on the Douglas–Rachford paradigm is more appropriate.

Theorem 21

In Problem 20, suppose that

0ranA+i=1mLi(BiDi)Li. (37)

Let τ,σi>0,i=1,,m, be strictly positive numbers such that

τi=1mσiLi2<4.

Let be the starting point (x0,v1,0,,vm,0)H×G1×Gm and set:

n0p1,n=JτAβnxnτ2βni=1mLivi,nw1,n=2p1,nβnxnFor i=1,,mp2,i,n=JσiBi1βnvi,n+σi2Liw1,nw2,i,n=2p2,i,nβnvi,nz1,n=w1,nτ2i=1mLiw2,i,nxn+1=βnxn+λn(z1,np1,n)For i=1,,mz2,i,n=JσiDi1w2,i,n+σi2Li(2z1,nw1,n)vi,n+1=βnvi,n+λn(z2,i,np2,i,n), (38)

where (λn)n0 and (βn)n0 real sequences satisfying the conditions:

  1. 0<βn1 for any n0,limn+βn=1,n0(1βn)=+ and n1|βnβn1|<+;

  2. 0<λn2 for any n0,lim infn+λn>0 and n1|λnλn1|<+.

Then there exists an element (x¯,v¯1,,v¯m)H×G1×Gm such that the following statements are true:

  1. by setting
    p¯1=JτAx¯τ2i=1mLiv¯i,p¯2,i=JσiBi1v¯i+σi2Li(2p¯1x¯),i=1,,m,
    the element (p¯1,p¯2,1,,p¯2,m)H×G1××Gm is a primal–dual solution to Problem 20;
  2. (xn,v1,n,,vm,n) converges strongly to (x¯,v¯1,,v¯m) as n+;

  3. (p1,n,p2,1,n,,p2,m,n) and (z1,n,z2,1,n,,z2,m,n) converge strongly to (p¯1,p¯2,1,,p¯2,m) as n+.

Proof.

For the proof we use Theorem 10 (see also [8]) in the same setting as in the proof of Theorem 14, namely, by considering K=H×G1××Gm endowed with inner product and associated norm defined in (20). Furthermore, we consider again the maximally monotone operator

M:K\rightrightarrowsK,(x,v1,,vm)(Ax,B11v1,,Bm1vm),

the linear continuous skew-symmetric operator

S:KK,(x,v1,,vm)i=1mLivi,L1x,,Lmx

and the (this time not necessarily single-valued) maximally monotone operator

Q:K\rightrightarrowsK,(x,v1,,vm)0,D11v1,,Dm1vm.

Since dom S=K, both 12S+Q and 12S+M are maximally monotone (see [6, Corollary 24.4(i)]). Furthermore,

(37)zerM+S+Q

and

(x,v1,,vm)zerM+S+Q(x,v1,,vm) is a primal--dual solution to Problem 20. (39)

We introduce the linear continuous operator

V:KK,(x,v1,,vm)xτ12i=1mLivi,v1σ112L1x,,vmσm12Lmx,

which is self-adjoint and ρ-strongly positive (see also [8]), for

ρ:=112τi=1mσiLi2min1τ,1σ1,,1σm>0,

namely, the following inequality holds

x,VxKρxK2xK.

Therefore, its inverse operator V1 exists and it fulfils V11/ρ.

The algorithmic scheme (38) is equivalent to

(n0)βnxnp1,nτ12i=1mLi(βnvi,np2,i,n)12i=1mLip2,i,n+Ap1,nw1,n=2p1,nβnxnFor i=1,,mβnvi,np2,i,nσi12Li(βnxnp1,n)12Lip1,n+Bi1p2,i,nw2,i,n=2p2,i,nβnvi,nw1,nz1,nτ12i=1mLiw2,i,n=0xn+1=βnxn+λn(z1,np1,n)For i=1,,mw2,i,nz2,i,nσi12Li(w1,nz1,n)12Liz1,n+Di1z2,i,nvi,n+1=βnvi,n+λn(z2,i,np2,i,n). (40)

By considering for any n0 the notations

xn=(xn,v1,n,,vm,n),yn=(p1,n,p2,1,n,,p2,m,n)andzn=(z1,n,z2,1,n,,z2,m,n),

the iterative scheme (40) can be written as

n0V(βnxnyn)12S+MynV(2ynβnxnzn)12S+Qznxn+1=βnxn+λnznyn, (41)

which is further equivalent to

n0yn=Id+V1(12S+M)1βnxnzn=Id+V1(12S+Q)12ynβnxnxn+1=βnxn+λnznyn. (42)

Let KV be the Hilbert space with inner product and norm defined, for x,yK, via

x,yKV=x,VyKandxKV=x,VxK, (43)

respectively. As the set-valued operators 12S+M and 12S+Q are maximally monotone on K, the operators

B:=V112S+MandA:=V112S+Q (44)

are maximally monotone on KV. Furthermore, since V is self-adjoint and ρ-strongly positive, strong convergence in KV is equivalent with strong convergence in K.

Taking this into account, (42) becomes

n0yn=JBβnxnzn=JA2ynβnxnxn+1=βnxn+λnznyn, (45)

which is the Douglas–Rachford algorithm formulated in Theorem 10 in case γ=1 for determining the zeros of A+B. It is easy to see that

zer(A+B)=zer(V1M+S+Q)=zer(M+S+Q).

By Theorem 10(a), there exists x¯=(x¯,v¯1,,v¯m)Fix(RARB), such that JBx¯zer(A+B)=zer(M+S+Q). The claim follows from Theorem 10, (39) and by writing JBx¯ in terms of the resolvents of the operators involved in the expression of B.

We close this section by considering the variational case.

Problem 22

Let fΓ(H),m be a strictly positive integer and for i=1,,m, let Gi be a real Hilbert space, gi,liΓ(Gi) and Li:HGi a nonzero linear continuous operator. Consider the convex optimization problem

infxHf(x)+i=1m(gili)(Lix) (46)

and its conjugate dual problem

sup(v1,,vm)G1××Gmfi=1mLivii=1mgi(vi)+li(vi). (47)

Corollary 23

In Problem 22, suppose that

0ranf+i=1mLi(gili)Li. (48)

Let τ,σi>0,i=1,,m, be strictly positive numbers such that

τi=1mσiLi2<4.

Let be the starting point (x0,v1,0,,vm,0)H×G1×Gm and set:

n0p1,n=proxτfβnxnτ2βni=1mLivi,nw1,n=2p1,nβnxnFor i=1,,mp2,i,n=proxσigiβnvi,n+σi2Liw1,nw2,i,n=2p2,i,nβnvi,nz1,n=w1,nτ2i=1mLiw2,i,nxn+1=βnxn+λn(z1,np1,n)For i=1,,mz2,i,n=proxσiliw2,i,n+σi2Li(2z1,nw1,n)vi,n+1=βnvi,n+λn(z2,i,np2,i,n), (49)

where (λn)n0 and (βn)n0 are real sequences satisfying the conditions:

  1. 0<βn1 for any n0,limn+βn=1,n0(1βn)=+ and n1|βnβn1|<+;

  2. 0<λn2 for any n0,lim infn+λn>0 and n1|λnλn1|<+.

Then there exists an element (x¯,v¯1,,v¯m)H×G1×Gm such that the following statements are true:

  1. by setting
    p¯1=proxτfx¯τ2i=1mLiv¯i,p¯2,i=proxσigiv¯i+σi2Li(2p¯1x¯),i=1,,m,
    the element (p¯1,p¯2,1,,p¯2,m)H×G1××Gm is a primal–dual solution to Problem 20, namely,
    i=1mLiv¯if(x¯)andv¯i(gili)(Lix¯),i=1,,m, (50)
    hence p¯1 is an optimal solution to (46) and (p¯2,1,,p¯2,m) is an optimal solution to (47);
  2. (xn,v1,n,,vm,n) converges strongly to (x¯,v¯1,,v¯m) as n+;

  3. (p1,n,p2,1,n,,p2,m,n) and (z1,n,z2,1,n,,z2,m,n) converge strongly to (p¯1,p¯2,1,,p¯2,m) as n+.

Remark 24

The hypothesis (48) in Theorem 23 is fulfilled, if the primal problem (46) has an optimal solution, the regularity condition (34) holds and

0sqri(dom gidom li)for i=1,,m.

According to [6, Proposition 15.7], the latter also guarantees that giliΓ(Gi), i=1,,m.

6. Numerical experiments: applications to the split feasibility problem

Let H and G be real Hilbert spaces and L:HG a bounded linear operator. Let C and Q be nonempty, closed and convex subsets of H and G, respectively. The split feasibility problem (SFP) searches a point xH with the property

xCandLxQ. (51)

The (SFP) was originally introduced by Censor and Elfving [11] for solving inverse problems in the context of phase retrieval, medical image reconstruction and intensity modulated radiation therapy.

We show that the strongly convergent primal–dual algorithms which we have investigated in Section 5 are excellently suited to solve the (SFP), especially in the case when infinite dimensional Hilbert spaces are involved. For this purpose, we note that problem (51) can be written equivalently in the form

minxHδC(x)+δQ(Lx), (52)

where

δS(x):=0,if xS+,else

denotes the indicator function of a subset S of a Hilbert space. Another alternative way to write problem (51) is as the minimization problem

minxH12dC2(x)+δQ(Lx), (53)

where dC(x)=infuCxu denotes the distance of the point xH to the set C. Both optimization problems are special instances of Problem 16, consequently, we will solve them by making use of the algorithm stated in Corollary 17.

The optimization problem (52) can be stated in the framework of Problem 16 by taking f=δC, m=1, g1=δQ, l1=δ{0}, L1=L and h=0. The iterative scheme stated in Corollary 17 reads:

n0pn=PC(βnxnτβnLvn)xn+1=βnxn+λn(pnβnxn)qn=βnvn+σL(2pnβnxn)σPQ(σ1βnvn+L(2pnβnxn))vn+1=βnvn+λn(qnβnvn). (54)

The optimization problem (52) can be stated in the framework of Problem 16 by taking f=0, m=1, g1=δQ, l1=δ{0}, L1=L and h=12dC2. Noticing that (12dC2)=IdPC, the iterative scheme stated in Corollary 17 reads:

n0pn=βnxnτ(βnLvn+βnxnPC(βnxn))xn+1=βnxn+λn(pnβnxn)qn=βnvn+σL(2pnβnxn)σPQ(σ1βnvn+L(2pnβnxn))vn+1=βnvn+λn(qnβnvn). (55)

For the numerical example we consider the following setup: let H=G=L2([0,2π]):={f:[0,2π]R:02π|f(t)|2dt<+} equipped with the scalar product f,g:=02πf(t)g(t)dt and the associated norm f:=(02π|f(t)|2dt)1/2 for all f,gL2([0,2π]). The sets

C:=xL2([0,2π]):02πx(t)dt1

and

Q:=xL2([0,2π]):02π|x(t)sin(t)|2dt16

are nonempty, closed and convex subsets of L2([0,2π]). Notice that C={xL2([0,2π]):x,u1} and Q={xL2([0,2π]):xf4}, where u:[0,2π]R,u(t)=1 for all t[0,2π], and f:[0,2π]R,f(t)=sint for all t[0,2π]. We define the linear continuous operator L:L2([0,2π])L2([0,2π]) as

(Lx)(t):=02πx(s)dst.

Let x,yL2([0,2π]). Then

Lx,y=02πLx(t)y(t)dt=02πx(s)02πty(t)dtds,

and therefore

(Ly)(t)=02πsy(s)ds.

Furthermore, by using the Cauchy-Schwarz inequality yields

L2=supx=102πTx(t)2dt=supx=102πt2dt02πx(s)ds2supx=183π302πx(s)2ds2π=163π4.

The projection onto the set C can be computed as (see [6, Example 28.16])

PC(x)=102πx(t)dt2π+x,if 02πx(t)dt>1x,else.

On the other hand PQ is given by [6, Example 28.10]

PQ(x)=sin+4(xsin)(02π|x(t)sin(t)|2dt)1/2,if 02π|x(t)sin(t)|2dt>16x,else.

We implemented the algorithms (54) and (55) in MATLAB used symbolic computation for generating the sequences of iterates. One can easily notice that, in this particular setting, the split feasibility problem and, consequently, both addressed optimization problems are solvable. The numerical experiments confirmed that the primal–dual algorithms involving Tikhonov regularization terms outperform the ones without Tikhonov regularization terms, which correspond to the case when βn=1 for every n0 and for which is known that they weakly convergence to a primal–dual solution of the corresponding KKT system of optimality conditions (Tables 1 and 2).

Table 1. Comparison of the variants without and with Tikhonov regularization terms of the primal–dual algorithm (54), for different starting values and step sizes τ=0.1 and σ=0.01.

    Number of iterations
x0 v0 βn=1 βn=11n+1
t210 t210 13 1
t210 12et 20 11
t210 et+t224 21 12
12et t210 > 150 11
12et 12et 20 12
12et et+t224 21 13
et+t224 t210 > 150 15
et+t224 12et 20 13
et+t224 et+t224 21 13

Table 2. Comparison of the variants without and with Tikhonov regularization terms of the primal–dual algorithm (55), for different starting values and step sizes τ=0.1 and σ=0.01.

    Number of iterations
x0 v0 βn=1 βn=11n+1
t210 t210 24 1
t210 12et 46 10
t210 et+t224 46 10
12et t210 30 6
12et 12et 24 11
12et et+t224 35 21
et+t224 t210 32 6
et+t224 12et 36 12
et+t224 et+t224 24 11

In the two tables above we present the numbers of iterations needed by the two algorithms to approach a solution of the split feasibility problem (SFP). We consider as stopping criterion

E(xn):=12PC(xn)xn2+12PQ(Lxn)Lxn2103,

while taking as relaxation variables λn=0.4 for every n0 and as Tikhonov regularization parameters βn:=11/(n+1) for every n0, respectively, βn=1 for every n0 for the variants without Tikhonov regularization terms.

7. Further work

We point out some directions of research related to proximal methods with Tikhonov regularization terms, which merit to be addressed starting from the investigations made in this paper:

  1. To consider in the numerical algorithms of type forward–backward and Douglas–Rachford proposed in this paper dynamic step sizes, which are known to increase the flexibility of the algorithms. This can be for instance done by formulating first a Krasnosel'skiĭ–Mann–type algorithm for determining an element in the intersection of the sets of fixed points of a family of nonexpansive operators (Tk)k0. Suitable choices of the operators in this family, in the spirit of the investigations we made in the Sections 3 and 4, can lead to iterative algorithms with dynamic step sizes (see [14]).

  2. To employ in the iterative schemes discussed in this article inertial and memory effects, which are known to contribute to the acceleration of the convergence behaviour of the algorithms.

  3. To translate the proposed numerical methods to a framework that goes beyond the Hilbert space setting considered in this article, in order to allow applications where functional spaces are involved. We refer the reader to [24,25,27] and the references therein for tools and techniques which allow to prove convergence statements in Banach spaces with some particular underlying geometric structures.

Funding Statement

Research of Radu Ioan Boţ was partially supported by FWF (Austrian Science Fund), project I 2419-N32. Research of Ernö Robert Csetnek was supported by FWF (Austrian Science Fund), projects M 1682-N25 and P 29809-N32. Research of Dennis Meier was partially supported by FWF (Austrian Science Fund), project I 2419-N32, and by the Doctoral Programme Vienna Graduate School on Computational Optimization (VGSCO), project W1260-N35.

Disclosure statement

No potential conflict of interest was reported by the authors.

ORCID

Radu Ioan Boţ  http://orcid.org/0000-0002-4469-314X

References

  • [1].Alotaibi A., Combettes P.L., and Shahzad N., Best approximation from the Kuhn-Tucker set of composite monotone inclusions, Numer. Funct. Anal. Optim. 36(12) (2015), pp. 1513–1532. doi: 10.1080/01630563.2015.1077864 [DOI] [Google Scholar]
  • [2].Attouch H., Viscosity solutions of minimization problems, SIAM J. Optim. 6(3) (1996), pp. 769–806. doi: 10.1137/S1052623493259616 [DOI] [Google Scholar]
  • [3].Attouch H. and Théra M., A general duality principle for the sum of two operators, J. Convex Anal. 3(1) (1996), pp. 1–24. [Google Scholar]
  • [4].Attouch H., Briceño-Arias L.M., and Combettes P.L., A strongly convergent primal-dual method for nonoverlapping domain decomposition, Numer. Math. 133(3) (2016), pp. 443–470. doi: 10.1007/s00211-015-0751-4 [DOI] [Google Scholar]
  • [5].Bauschke H.H. and Combettes P.L., A weak-to-strong convergence principle for Fejér-monotone methods in Hilbert spaces, Math. Oper. Res. 26(2) (2001), pp. 248–264. doi: 10.1287/moor.26.2.248.10558 [DOI] [Google Scholar]
  • [6].Bauschke H.H. and Combettes P.L., Convex Analysis and Monotone Operator Theory in Hilbert Spaces, CMS Books in Mathematics, Springer, New York, 2011. [Google Scholar]
  • [7].Borwein J.M. and Vanderwerff J.D., Convex Functions: Constructions, Characterizations and Counterexamples, Cambridge University Press, Cambridge, 2010. [Google Scholar]
  • [8].Boţ R.I. and Hendrich C., A Douglas-Rachford type primal-dual method for solving inclusions with mixtures of composite and parallel-sum type monotone operators, SIAM J. Optim. 23(4) (2013), pp. 2541–2565. doi: 10.1137/120901106 [DOI] [Google Scholar]
  • [9].Boţ R.I., Csetnek E.R., Heinrich A., and Hendrich C., On the convergence rate improvement of a primal-dual splitting algorithm for solving monotone inclusion problems, Math. Program. 150(2) (2015), pp. 251–279. doi: 10.1007/s10107-014-0766-0 [DOI] [Google Scholar]
  • [10].Briceño-Arias L.M. and Combettes P.L., A monotone + skew splitting model for composite monotone inclusions in duality, SIAM J. Optim. 21(4) (2011), pp. 1230–1250. doi: 10.1137/10081602X [DOI] [Google Scholar]
  • [11].Censor Y. and Elfving T., A multiprojection algorithm using Bregman projections in a product space, Numer. Algorithms 8 (1994), pp. 221–239. doi: 10.1007/BF02142692 [DOI] [Google Scholar]
  • [12].Chambolle A. and Pock T., A first-order primal-dual algorithm for convex problems with applications to imaging, J. Math. Imaging Vis. 40(1) (2011), pp. 120–145. doi: 10.1007/s10851-010-0251-1 [DOI] [Google Scholar]
  • [13].Chidume C., Geometric Properties of Banach Spaces and Nonlinear Iterations, Lecture Notes in Mathematics 1965, Springer-Verlag, London, 2009. [Google Scholar]
  • [14].Combettes P.L., Solving monotone inclusions via compositions of nonexpansive averaged operators, Optimization 53(5–6) (2004), pp. 475–504. doi: 10.1080/02331930412331327157 [DOI] [Google Scholar]
  • [15].Combettes P.L. and Pesquet J.-C., Primal-dual splitting algorithm for solving inclusions with mixtures of composite, Lipschitzian, and parallel-sum type monotone operators, Set-Valued Var. Anal. 20(2) (2012), pp. 307–330. doi: 10.1007/s11228-011-0191-y [DOI] [Google Scholar]
  • [16].Combettes P.L. and Yamada I., Compositions and convex combinations of averaged nonexpansive operators, J. Math. Anal. Appl. 425(1) (2015), pp. 55–70. doi: 10.1016/j.jmaa.2014.11.044 [DOI] [Google Scholar]
  • [17].Condat L., A primal-dual splitting method for convex optimization involving Lipschitzian, proximable and linear, composite terms, J. Optim. Theory Appl. 158(2) (2013), pp. 460–479. doi: 10.1007/s10957-012-0245-9 [DOI] [Google Scholar]
  • [18].Haugazeau Y., Sur les inéquations variationnelles et la minimisation de fonctionnelles convexes, Thése, Université de Paris, France, 1968.
  • [19].Lehdili N. and Moudafi A., Combining the proximal algorithm and Tikhonov regularization, Optimization 37(3) (1996), pp. 239–252. doi: 10.1080/02331939608844217 [DOI] [Google Scholar]
  • [20].Moudafi A., Viscosity approximation methods for fixed-points problems, J. Math. Anal. Appl. 241(1) (2000), pp. 46–55. doi: 10.1006/jmaa.1999.6615 [DOI] [Google Scholar]
  • [21].Ogura N. and Yamada I., Non-strictly convex minimization over the fixed point set of an asymptotically shrinking nonexpansive mapping, Numer. Funct. Anal. Optim. 23(1–2) (2002), pp. 113–137. doi: 10.1081/NFA-120003674 [DOI] [Google Scholar]
  • [22].Rockafellar R.T., On the maximal monotonicity of subdifferential mappings, Pacific J. Math. 33(1) (1970), pp. 209–216. doi: 10.2140/pjm.1970.33.209 [DOI] [Google Scholar]
  • [23].Rockafellar R.T., Monotone operators and the proximal point algorithm, SIAM J. Control Optim. 14(5) (1976), pp. 877–898. doi: 10.1137/0314056 [DOI] [Google Scholar]
  • [24].Sahu D.R. and Yao Y.C., The prox-Tikhonov regularization method for the proximal point algorithm in Banach spaces, J. Global Optim. 51(4) (2011), pp. 641–655. doi: 10.1007/s10898-011-9647-8 [DOI] [Google Scholar]
  • [25].Sahu D.R., Ansari Q.H., and Yao J.C., The prox-Tikhonov-like forward-backward method and applications, Taiwanese J. Math. 19(2) (2015), pp. 481–503. doi: 10.11650/tjm.19.2015.4972 [DOI] [Google Scholar]
  • [26].Simons S., From Hahn-Banach to Monotonicity, Springer, Berlin, 2008. [Google Scholar]
  • [27].Xu H.-K., Iterative algorithms for nonlinear operators, J. London Math. Soc. 66(1) (2002), pp. 240–256. doi: 10.1112/S0024610702003332 [DOI] [Google Scholar]
  • [28].Vũ B.C., A splitting algorithm for dual monotone inclusions involving cocoercive operators, Adv. Comput. Math. 38(3) (2013), pp. 667–681. doi: 10.1007/s10444-011-9254-8 [DOI] [Google Scholar]
  • [29].Zălinescu C., Convex Analysis in General Vector Spaces, World Scientific, Singapore, 2002. [Google Scholar]

Articles from Optimization Methods & Software are provided here courtesy of Taylor & Francis

RESOURCES