• Non ci sono risultati.

Modified Fejér sequences and applications

N/A
N/A
Protected

Academic year: 2021

Condividi "Modified Fejér sequences and applications"

Copied!
26
0
0

Testo completo

(1)

Modified Fejér sequences and applications

The MIT Faculty has made this article openly available. Please share

how this access benefits you. Your story matters.

Citation Lin, Junhong, et al. “Modified Fejér Sequences and Applications.” Computational Optimization and Applications, vol. 71, no. 1, Sept. 2018, pp. 95–113.

As Published https://doi.org/10.1007/s10589-017-9962-1

Publisher Springer US

Version Author's final manuscript

Citable link http://hdl.handle.net/1721.1/117359

Terms of Use Article is made available in accordance with the publisher's policy and may be subject to US copyright law. Please refer to the publisher's site for terms of use.

(2)

Modified Fej´er Sequences and Applications

Junhong Lin†, Lorenzo Rosasco†, ◦, Silvia Villa, and Ding-Xuan Zhou*

LCSL, Istituto Italiano di Tecnologia and Massachusetts Institute of Technology, Cambridge,

MA 02139, USA

Dipartimento di Matematica, Politecnico di Milano, Milano 20133, Italy

DIBRIS, Universit´a degli Studi di Genova, Genova 16146, Italy

*Department of Mathematics, City University of Hong Kong, Kowloon, Hong Kong, China

October 30, 2017

Abstract

In this note, we propose and study the notion of modified Fej´er sequences. Within a Hilbert space setting, this property has been used to prove ergodic convergence of proximal incremental subgra-dient methods. Here we show that indeed it provides a unifying framework to prove convergence rates for objective function values of several optimization algorithms. In particular, our results apply to forward-backward splitting algorithm, incremental subgradient proxi-mal algorithm, and the Douglas-Rachford splitting method including and generalizing known results.

This material is based upon work supported by the Center for Brains, Minds and

Machines (CBMM), funded by NSF STC award CCF-1231216. The work by D. X. Zhou described in this paper is supported by a grant from the NSFC/RGC Joint Research Scheme [RGC Project No. N CityU120/14 and NSFC Project No. 11461161006]. L. R. acknowledges the financial support of the Italian Ministry of Education, University and Research FIRB project RBFR12M3AC, and S. V. the support of INDAM-GNAMPA, project 2017: “Algoritmi di ottimizzazione ad equazioni di evoluzione ereditarie”.

(3)

1

Introduction

We are interested in the study of convergence properties of optimization algorithms to solve the problem

min

x∈Hf (x),

where H is a Hilbert space and f : H →] − ∞, +∞] is a proper function. Let (xt)t∈N be the sequence generated by a chosen algorithm. The sequence is

said to be Fej´er monotone if, for every x∗ minimizer of f , kxt+1−x∗k2 ≤ kxt−

x∗k2. The notion of Fej´er monotonicity captures essential properties of (xt)t∈N

generated by a wide range of optimization methods and provides a common framework to analyze their convergence [11]. Quasi-Fej´er monotonicity is a relaxation of the above notion that allows for an additional error term [13, 22]. Generalizations of the above notion have been proposed, that allow to deal with variable metric algorithms [18, 40], and stochastic perturbations [15, 16, 22, 37, 38, 39].

In this paper, we propose and study a novel, related notion, to analyze the convergence of the objective function values f (xt), in addition to that of the

iterates. More precisely, we modify the notion of quasi-Fej´er monotonicity, by adding a term involving the objective function and say that a sequence satisfying the new requirement is modified Fej´er monotone (modified Fej´er for short). This property is the key step to derive ergodic convergence of the iterates generated by the proximal incremental gradient algorithm in [7]. In this paper, we show the wider usefulness of this new notion of monotonicity by deriving convergence rates for several optimization algorithms in a unified way. Based on this approach, we not only recover known results, such as the sublinear convergence rate for the proximal forward-backward splitting algorithm, but also derive new results. Interestingly, our results show that for projected subgradient, incremental proximal subgradient, and Douglas-Rachford algorithms, considering the last iterate leads to essentially the same convergence rate as considering the best iterate selection rule [36, 41], or ergodic means [8, 42], as typically done.

(4)

2

Modified Fej´

er Sequences

Throughout this paper, we assume that H is a Hilbert space, and f : H → ]−∞, ∞] is a proper function. We assume that the set of minimizers of f

X = {z ∈ H | f (z) = min

x∈Hf (x)}

is nonempty. We are interested in solving the following optimization problem

f∗ = min

x∈Hf (x). (1)

Given x ∈ H and a subset S ⊂ H, d(x, S) denotes the distance between x and S, i.e., d(x, S) = infx′∈Skx − x′k. R+ is the set of all non-negative real

numbers and N∗ = N \ {0}.

The following definition introduces the key notion we propose in this paper.

Definition 1. A sequence (xt)t∈N ∈ HN is modified Fej´er monotone with

respect to the objective function f and the sequence ((ηt, ξt))t∈N in R2+, if

(∀x ∈ domf ) kxt+1− xk2 ≤ kxt− xk2− ηt(f (xt+1) − f (x)) + ξt. (2)

Remark 1.

(i) Choosing x ∈ X in (2), we get

ηtf (xt+1) ≤ ξt+ ηtf∗ + kxt− xk2 < ∞.

This implies that xt ∈ domf for every t ∈ N.

(ii) All the subsequent results hold if condition (2) is replaced by the follow-ing weaker condition

(∀x ∈ X ∪ {xt}t∈N) kxt+1− xk2 ≤ kxt− xk2− ηt(f (xt+1) − f (x)) + ξt.

(3) However, in the proposed applications, condition (2) is always satisfied for every x ∈ domf .

(5)

(iii) Inequality (2) has been proved to be satisfied and implicitly used to de-rive convergence rate for many algorithms, considering the best iterate selection rule, e.g., [36, 41], or ergodic means [8, 42, 7]. More precisely, for not descending methods, common approaches keep track of the best point found so far, i.e. they study the following quantity,

(∀T ∈ N∗) bT = min 1≤t≤Tf (xt) − f∗, or this one: (∀T ∈ N∗) f T X t=1 xt/T ! − f∗.

The main novelty of this paper is to show that considering the last iterate, i.e. f (xT) − f∗, leads to essentially the same convergence rate

as that for considering the best iterate selection rule, or ergodic means. See Theorem 2.

(iv) A similar condition has been considered in [1] to study convergence properties of several optimization methods, and in [40] to study a vari-able metric forward-backward algorithm under relaxed differentiability assumptions.

(v) Extensions of the proposed notion could be considered, e.g. resem-bling variable metric and stochastic quasi-Fej´er monotonicity proper-ties, which have been recently proposed and investigated in [18, 15, 37, 39].

In the following remark we discuss the relation with classical Fej´er se-quences.

Remark 2 (Comparison with quasi-Fej´er sequences). IfP

t∈Nξt < +∞, Definition 1 implies that the sequence (xt)t∈Nis quasi-Fej´er

monotone with respect to X [13, 22]. Indeed, (2) implies

(∀x ∈ X ) kxt+1− xk2 ≤ kxt− xk2+ ξt.

Note that, in the study of convergence properties of quasi-Fej´er sequences corresponding to a minimization problem, the property is considered with

(6)

respect to the set of solutions X , while here we will consider modified Fej´er monotonicity for the entire space H.

We next present two main results to show how modified Fej´er sequences are useful to study the convergence of optimization algorithms. The first result shows that if a sequence is modified Fej´er monotone, one can bound its corresponding excess function values in terms of ((ηt, ξt))t∈N explicitly.

Theorem 1. Let (xt)t∈N ⊂ HN be a modified Fej´er sequence with respect to f

and ((ηt, ξt))t∈Nin R2+. Let (ηt)t∈N be a non-increasing sequence. Let T ∈ N∗.

Then ηT(f (xT +1) − f∗) ≤ d(x1, X ) 2 T + T −1 X t=1 1 T − t + 1ξt. (4) Proof. Let (ut)t∈N be a sequence in R. For every k ∈ {1, · · · , T − 1}, let

sk=PTj=T −kuj. Then, since sk = sk−1+ uT −k, 1 ksk−1− 1 k + 1sk = 1 k(k + 1)((k + 1)sk−1− ksk) = 1 k(k + 1)(sk−1− kuT −k).

Summing over k = 1, · · · , T − 1, and rearranging terms, we get

uT = 1 TsT −1+ T −1 X k=1 1 k(k + 1)(sk−1− kuT −k). (5) Let x ∈ domf and choose (∀t ∈ N) ut= ηt(f (xt+1) − f (x)). Then, we derive

the following error decomposition [28]:

ηT(f (xT +1) − f (x)) = 1 T T X t=1 ηt(f (xt+1) − f (x)) + T −1 X k=1 1 k(k + 1) T X t=T −k+1 ηt(f (xt+1) − f (xT −k+1)) + T −1 X k=1 1 k + 1 " 1 k T X t=T −k+1 ηt  − ηT −k # (f (xT −k+1) − f (x)) .

(7)

Let x = x∗ ∈ X . Since (ηt)t∈N is non-increasing and f (xT −k+1) − f∗ ≥ 0, the

last term of the above inequality is less than or equal to 0. Thus, we derive that ηT(f (xT +1) − f∗) ≤ 1 T T X t=1 ηt(f (xt+1) − f (x∗)) + T −1 X k=1 1 k(k + 1) T X t=T −k+1 ηt(f (xt+1) − f (xT −k+1)). (6)

For every j ∈ {1, . . . , T }, and for every x ∈ domf , summing up (2) over t = j, · · · , T , we get T X t=j ηt(f (xt+1) − f (x)) ≤ kxj− xk2+ T X t=j ξt. (7)

The above inequality with x = x∗ and j = 1 implies

1 T T X t=1 ηt(f (xt+1) − f (x∗)) ≤ 1 Tkx1− x∗k 2 + 1 T T X t=1 ξt. (8)

Inequality (7) with x = xT −k+1 and j = T − k + 1 yields

T X t=T −k+1 ηt(f (xt+1) − f (xT −k+1)) ≤ T X t=T −k+1 ξt, and thus T −1 X k=1 1 k(k + 1) T X t=T −k+1 ηt(f (xt+1) − f (xT −k+1)) ≤ T −1 X k=1 1 k(k + 1) T X t=T −k+1 ξt. (9)

(8)

Exchanging the order in the sum, we obtain T −1 X k=1 1 k(k + 1) T X t=T −k+1 ξt= T X t=2 T −1 X k=T −t+1 1 k(k + 1)ξt = T X t=2  1 T − t + 1 − 1 T  ξt = T X t=2 1 T − t + 1ξt− 1 T T X t=2 ξt. (10)

The result follows by plugging (8) and (10) into (6).

Remark 3. F´ejer monotonicity of a sequence is often useful not only for obtaining convergence rate estimates, but also for proving the convergence of the iterates to a minimizer, see e.g. [2, Proposition 2]. Here we mainly focus directly on convergence rates for the objective function values.

In the special case when, for every t ∈ N, ξt= 0, we derive the following

result.

Corollary 1. Let (xt)t∈N ∈ HN be a modified Fej´er sequence with respect to

f and a sequence ((ηt, 0))t∈N in R2+. Suppose that (ηt)t∈N is non-increasing.

Then for any T ∈ N∗,

f (xT +1) − f∗ ≤ 1 ηTT d(x1, X ) 2 .

In the case that ξtis non-increasing, we can get the following result, which

simplifies the upper bound in (4) from Theorem 1.

Corollary 2. Let (xt)t∈N ∈ HN be a modified Fej´er sequence with respect

to an objective function f and a sequence ((ηt, ξt))t∈N in R2+. Suppose that

(ξt)t∈N and (ηt)t∈N are non-increasing. Let T ∈ N∗. Then

f (xT +1) − f∗ ≤ d(x1, X ) 2 + 2 T X t=1 ξt ! (T ηT)−1+ ξT 2+1⌋η −1 T log(T /2) (11)

(9)

Proof. Since ξt is non-increasing, T −1 X t=1 ξt T − t + 1 ≤ ξ⌊T2+1⌋ X T /2+1≤t≤T −1 1 T − t + 1 + 2T −1 X 1≤t<T /2+1 ξt ≤ ξT 2+1⌋log(T /2) + 2T −1 T X t=1 ξt.

The result follows directly from Theorem 1.

The next main result shows how to derive explicit rates for the objec-tive function values corresponding to a modified Fej´er sequence with respect to a polynomially decaying sequence ((ηt, ξt))t∈N in R2+. Interestingly, the

following result (as well as the previous ones) does not require convexity of f .

Theorem 2. Let (xt)t∈N ∈ HN be a modified Fej´er sequence with respect to

an objective function f and a sequence ((ηt, ξt))t∈N in R2+. Let η ∈ ]0, +∞[,

let θ1 ∈ [0, 1[, and set ηt= ηt−θ1. Let (θ2, ξ) ∈ R2+ and suppose that ξt= ξt−θ2

for all t ∈ N. Let T ∈ N∗. Then, setting c = 2θ2 + 2/(1 − θ

2): f (xT +1) − f∗ ≤                  d(x1, X ) 2 η T θ1−1+ ξ η  2θ2 + 2 1−θ2  Tθ1−θ2log T if θ 2 < 1 d(x1, X )2 η T θ1−1+4ξ ηT θ1−1log T if θ2 = 1  d(x1, X )2 η + ξ (θ2− 1)η  Tθ1−1 otherwise. (12) Proof. Let q ∈ ]0, +∞[. For n ≥ 2 we have (see [25, Theorem 3.3.3])

n X t=2 t−q ≤ Z n 1 u−qdu ≤    n1−q/(1 − q), when q < 1, log n, when q = 1, 1/(q − 1), when q > 1. (13)

We derive from Corollary 2 that

f (xT +1) −f∗ ≤ d(x1, X ) 2 + 2ξ T X t=1 t−θ2 ! η−1Tθ1−1+ 2θ2ξη−1Tθ1−θ2log(T /2). (14)

(10)

If θ2 < 1, equation (13) yields f (xT +1) − f∗ ≤ η−1d(x1, X ) 2 Tθ1−1+ ξη−1  2 1 − θ2 + 2θ2  Tθ1−θ2log T. (15)

The case θ2 = 1 is analogous. If θ2 > 1, from (13), it follows that

f (xT +1) − f∗ ≤ d(x1, X ) 2

+ 2ξ(1 − θ2)−1 η−1Tθ1−1+ 2θ2ξη−1Tθ1−θ2log T.

(16) Since 2θ2T−θ2log T ≤ T−1/(θ

2− 1), the result follows.

Remark 4. If a sequence (yt)t∈N∈ H N

satisfies

(∀x ∈ domf )(∀t ∈ N) kyt+1− xk2 ≤ kyt− xk2− ηt(f (yt) − f (x)) + ξt,

under the same assumptions of Corollary 2, it is possible to derive an in-equality analogous to (12).

3

Applications in Convex Optimization

In this section, we apply previous results to some convex optimization al-gorithms, including forward-backward splitting, projected subgradient, in-cremental proximal subgradient, and Douglas-Rachford splitting method. Convergence rates for the objective function values are obtained by using Theorem 2. The key observation is that the sequences generated by these algorithms are modified Fej´er monotone.

Throughout this section, we assume that f : H →] − ∞, ∞] is a proper, lower semicontinuous convex function. Recall that the subdifferential of f at x ∈ H is

∂f (x) = {u ∈ H : (∀y ∈ H) f (x) + hu, y − xi ≤ f (y)}. (17) The elements of the subdifferential of f at x are called subgradients of f at x. More generally, for ǫ ∈ ]0, +∞[, the ǫ-subdifferential of f at x is the set ∂ǫf (x) defined by

∂ǫf (x) = {u ∈ H : (∀y ∈ H) f (x) + hu, y − xi − ǫ ≤ f (y)}. (18)

The proximity operator of f [30] is

proxf(x) = argmin y∈H  f (y) + 1 2ky − xk 2  . (19)

(11)

3.1

Forward-Backward Splitting

In this subsection, we consider a forward-backward splitting algorithm for solving Problem (1), with objective function

f = ℓ + r (20)

where r : H → ]−∞, ∞] and ℓ : H → R are proper, lower semicontinuous, and convex. Since ℓ is real-valued, we have dom ∂ℓ = H [3, Proposition 16.14].

Algorithm 1. Given x1 ∈ H, a sequence of stepsizes (αt)t∈N⊂ ]0, +∞[, and

a sequence (ǫt)t∈N⊂ [0, +∞[ set, for every t ∈ N,

xt+1 = proxαtr(xt− αtgt) (21)

with gt∈ ∂ǫtℓ(xt).

The forward-backward splitting algorithm has been well studied [43, 10, 12, 9] and a review of this algorithm can be found in [14] under the assumption that ℓ is differentiable with a Lipschitz continuous gradient. Convergence is proved using arguments based on Fej´er monotonicity of the generated sequences [13]. Under the assumption that ℓ is a differentiable function with Lipschitz continuous gradient, the algorithm exhibits a sublinear convergence rate O(T−1) on the objective f [4]. If ℓ is not smooth, the algorithm has been

studied first in [34], and has a convergence rate O(T−1/2), considering the

best point selection rule [42]. A comprehensive study of proximal subgradient methods can be found in [5]. The use of ǫ-subgradients in a scaled version of algorithm (21) has been investigated in [6], in the special case where r is the indicator function of a convex and closed set, without focusing on convergence rates. Our objective here is to provide a convergence rate for the algorithm considering the last iterate, which shares the same rate (up-to logarithmic fac(up-tors) and (up-to allow the use of ǫ-subgradients, instead of subgradients. Before stating our main results, we introduce the following novel lemma for the forward-backward splitting for a (possibly) non-smooth ℓ. It recovers previous result (e.g. [4]) when ℓ is smooth.

(12)

Lemma 1. Let (xt)t∈N∗ be the sequence generated by Algorithm 1. Then for

all t ∈ N∗, there holds

2αt[f (xt+1) − f (x)] ≤ kxt− xk2− kxt+1− xk2− kxt− xt+1k2

+ 2αt[hxt+1− xt, gt+1− gti + ǫt+1+ ǫt]. (22)

Proof. Let t ∈ N∗. By Fermat’s rule (see e.g. [3, Theorem 16.2]),

0 ∈ xt+1− xt+ αtgt+ αt∂r(xt+1).

Thus, there exists qt+1∈ ∂r(xt+1), such that xt+1 in (21) can be written as

xt+1 = xt− αtgt− αtqt+1. (23)

Let x ∈ domf . The convexity of r implies

r(xt+1) − r(x) ≤ hxt+1− x, qt+1i.

Multiplying both sides by 2αt, and combining with (23), we get

2αt[r(xt+1) − r(x)] ≤ 2αthxt+1− x, qt+1i

= 2hxt+1− x, xt− xt+1− αtgti

= 2hxt+1− x, xt− xt+1i + 2αthx − xt+1, gti.

A direct computation yields

2hxt+1− x, xt− xt+1i = 2hxt+1− x, xt− xi − 2kxt+1− xk2 = kxt− xk2− kxt+1− xk2− kxt− xt+1k2. (24) Therefore, 2αt[r(xt+1) − r(x)] ≤ kxt− xk2− kxt+1− xk2− kxt− xt+1k2+ 2αthx − xt+1, gti. Moreover, by (18), we have hx − xt, gti ≤ ℓ(x) − ℓ(xt) + ǫt,

(13)

and hxt− xt+1, gt+1i ≤ ℓ(xt) − ℓ(xt+1) + ǫt+1, and thus hx − xt+1, gti = hx − xt, gti + hxt− xt+1, gt+1i + hxt− xt+1, gt− gt+1i ≤ ℓ(x) − ℓ(xt) + ǫt+ ℓ(xt) − ℓ(xt+1) + ǫt+1+ hxt− xt+1, gt− gt+1i = ℓ(x) − ℓ(xt+1) + hxt+1− xt, gt+1− gti + ǫt+1+ ǫt. Consequently, we get 2αt[r(xt+1) − r(x)] ≤ kxt− xk2− kxt+1− xk2− kxt− xt+1k2 +2αt[ℓ(x) − ℓ(xt+1) + hxt+1− xt, gt+1− gti + ǫt+1+ ǫt].

Rearranging terms and recalling that f = ℓ+r, the desired result thus follows.

Theorem 3. Let α ∈ ]0, +∞[, let θ ∈ [0, 1[, and let, for every t ∈ N∗,

αt = αt−θ. Let ǫ ∈ ]0, +∞[, (ǫt)t∈N∗ ⊂ [0, +∞], and assume that ǫt ≤ ǫαt.

Let (xt)t∈N∗ be the sequence generated by Algorithm 1. Let T ∈ N, and assume

that there exists B ∈ ]0, +∞[ such that

(∀T ∈ {1, . . . , T }) kgtk ≤ B, (25)

Then, there exists c ∈ [0, +∞[ such that

f (xT +1) − f∗ ≤      d(x1, X ) 2 2α T θ−1+ 2αc(B2 + ǫ)T−θlog T if θ ≤ 1/2  d(x1, X )2 2α + 2αc(B 2 + ǫ)  Tθ−1 otherwise.

Proof. Let t ∈ N∗. By (22) and Cauchy-Schwartz inequality

kxt+1− xk2− kxt− xk2 ≤ − kxt− xt+1k2+ 2αt[hxt+1− xt, gt+1− gti + ǫt+1+ ǫt] − 2αt[f (xt+1) − f (x)] ≤ − kxt− xt+1k2+ 2αt[kxt+1− xtkkgt+1− gtk + ǫt+1+ ǫt] − 2αt[f (xt+1) − f (x)] ≤α2 tkgt+1− gtk 2 + 2αt[ǫt+1+ ǫt] − 2αt[f (xt+1) − f (x)].

(14)

Using the assumptions kgtk ≤ B and ǫt≤ ǫαt, kxt+1− xk2− kxt− xk2 ≤4B2 α2 t + 2ǫαt[αt+ αt+1] − 2αt[f (xt+1) − f (x)] ≤4(B2 + ǫ)α2 t − 2αt[f (xt+1) − f (x)]. (26)

Thus, (xt)t∈N∗ is a modified Fej´er sequence with respect to the objective

func-tion f and (2αt, 4(B2+ ǫ)αt2)



t∈N∗. The statement follows from Theorem 2,

applied with θ1 = θ, θ2 = 2θ, η = 2α and ξ = 4(B 2

+ ǫ)α2

.

The following remark collects some comments on the previous result.

Remark 5.

(i) Theorem 3 suggests the optimal choice for the stepsize decay rate θ = 1/2. In such a way, we get a convergence rate O(T−1/2log T ) for

forward-backward algorithm applied to a sum of nonsmooth functions with nonsummable diminishing stepsizes, considering the last iterate. As usual in this setting, no convergence is obtained using a fixed step-size.

(ii) In Theorem 3, the assumption on bounded approximate subgradients, which implies Lipschitz continuity of ℓ, is satisfied for several practical optimization problems. For example, when r is the indicator function of a closed, bounded, and convex set D ⊂ RN, it follows that (x

t)t∈N is

bounded. If ℓ is bounded on bounded sets, this implies that ℓ is Lipschitz continuous on bounded sets [44, Corollary 2.2.12] and thus that (gt)t∈N

is bounded as well [35, Proposition 1.11]. Similar results to those proved here may be obtained by imposing a less restrictive growth condition on ∂f , using a similar approach to that in [28] to bound the sequence of subgradients.

(iii) An inequality similar to (26) has been obtained in [5, Lemma 2.1]. The-orem 3 improves [5, Corollary 2.4] in two aspects. First, the assumption (25) is weaker than the assumption kgt+ utk ≤ B for some ut∈ ∂r(xt),

(15)

in [5]. Second, [5] shows convergence rate only for the best point, i.e, the one with smallest function value:

(∀T ∈ N∗) bT = argmin 1≤t≤T

f (xt). (27)

whereas our result holds for any iterate.

If the function ℓ in (20) is differentiable, with a Lipschitz differentiable gradient, we recover the following well-known convergence result. We include the proof for completeness.

Proposition 1. [4, Theorem 3.1] Let β ∈ [0, +∞[ and assume that ∇ℓ is β-Lipschitz continuous. Consider Algorithm 1 with ǫt = 0 and (αt)t∈N

non-increasing, with αt∈ ]0, 1/β[ for all t ∈ N∗. Then, for every T ∈ N∗,

f (xT +1) − f∗ ≤

d(x1, X )2

αTT

(28)

Proof. Since ∇ℓ is β-Lipschitz continuous, it follows from the descent lemma [3, Theorem 18.15] and the definition of the forward-backward algorithm that

ℓ(xt+1) − ℓ(xt) ≤ h∇ℓ(xt), xt+1− xti +

β

2kxt+1− xtk

2

. (29)

Since (xt− xt+1)/αt− ∇f (xt) ∈ ∂r(xt+1), it follows from convexity of ℓ and

r, that (∀y ∈ H) ℓ(xt) − ℓ(y) ≤ h∇ℓ(xt), xt− yi (30) (∀y ∈ H) r(xt+1) − r(y) ≤ xt+1− xt αt + ∇ℓ(xt), y − xt+1 . (31) Summing up (29),(30),(31) we derive f (xt+1) − f (y) ≤ xt− xt+1 αt , xt+1− y + β 2kxt− xt+1k 2 . Therefore 2αt(f (xt+1) − f (y)) ≤ kxt− yk2− kxt+1− yk2+ (βαt− 1) kxt− xt+1k2,

which implies, since αt≤ 1/β, that (xt)t∈N is modified Fej´er monotone with

respect to f and the sequence (2αt, 0). The statement follows from

(16)

Remark 6. For the forward-backward algorithm, it has been proved in [17] that f (xT +1) − f∗ = o(1/T ) also for αt ∈ ]0, 2/β[, even in the presence

of errors, and with relaxation. The concept of modified Fej´er monotonicity allows to derive nonasymptotic bounds on the sequence of iterates, but only if αt ∈ ]0, 1/β[ for every t ∈ N∗.

3.2

Projected Approximate Subgradient Method

Let D be a convex and closed subset of H, and let ιD be the indicator function

of D. In this subsection, we consider Problem (1) with objective function given by

f = ℓ + ιD (32)

where ℓ : H → R is lower semicontinuous and convex. It is clear that (32) is a special case of (20) corresponding to a given choice of r. The forward-backward algorithm in this case reduces to the following projected subgradi-ent method (see e.g. [8, 26, 36, 41] and references therein), which allows to use ǫ-subgradients, see [2, 11].

Algorithm 2. Given x1 ∈ H, a sequence of stepsizes (αt)t∈N⊂ ]0, +∞[, and

a sequence (ǫt)t∈N⊂ [0, +∞[ , set, for every t ∈ N,

xt+1 = PD(xt− αtgt) (33)

with gt∈ ∂ǫtℓ(xt).

The algorithm has been studied using different rules for choosing the stepsizes. Here, as a corollary of Theorem 3, we derive the convergence rate for the objective function values, for a nonsummable diminishing stepsize. Theorem 4. For some α1 > 0, ǫ ≥ 0 and θ ∈ [0, 1), let αt = ηt−θ and

ǫt ≤ ǫαt for all t ∈ N∗. Let (xt)t∈N be a sequence generated by Algorithm 2.

Assume that for all t ∈ N∗, kg

tk ≤ B. Then, there exists c ∈ ]0, +∞[ such

that, for every T ∈ N∗

f (xT +1) − f∗ ≤        d(x1, X )2 2α1 Tθ−1+ α 1c(B2+ 2ǫ)T−θlog T if θ ≤ 1/2  d(x1, X )2 2α1 + α1c(B2+ 2ǫ)  Tθ−1 otherwise

(17)

Choosing θ = 1/2, we get a convergence rate of order O(T−1/2log T ) for

projected approximate subgradient method with nonsummable diminishing stepsizes, which is optimal up to a log factor without any further assump-tion on f [19, 33]. Since the subgradient method is not a descent method, a common approach keeps track of the best point found so far, see (27). The projected subgradient method with diminishing stepsizes of the form (αt−θ)

t∈N, with θ ∈ ]0, 1], satisfies f (xT +1) − f∗ = O(T−1/2). Our result

shows that considering the last iterate for projected approximate subgradi-ent method esssubgradi-entially leads to the same convergence rate, up to a logarithmic factor, as the one corresponding to the best iterate, even if the function value may not decrease at each iteration. To the best of our knowledge, our result is the first of this kind, without any assumption on strong convexity of f , or on a conditioning number with respect to subgradients (as in [23] using stepsizes {γt/kgtk}t). Note that, using nonsummable diminishing stepsizes,

convergence rate O(T−1/2) was shown, but only for a subsequence of (x t)t∈N∗

[2]. Finally, let us mention that using properties of quasi-Fej´er sequences, convergence properties were proved in [11].

3.3

Incremental Subgradient Proximal Algorithm

In this subsection, we consider an incremental subgradient proximal algo-rithm [7, 31] for solving (1), with objective function f given by, for some m ∈ N∗,

m

X

i=1

(ℓi+ ri),

where for each i, ℓi : H → R and ri : H → ]−∞, +∞] are convex, proper,

and lower semicontinuous. The algorithm is similar to the proximal subgra-dient method, the main difference being that at each iteration, xt is updated

incrementally, through a sequence of m steps.

Algorithm 3. Let t ∈ N∗. Given x

t ∈ H, an iteration of the incremental

proximal subgradient algorithm generates xt+1 according to the recursion,

(18)

where ψm

t is obtained at the end of a cycle, namely as the last step of the

recursion ψ0 t = xt, ψit = proxαtri(ψ i−1 t − αtgit), ∀gti ∈ ∂ℓi(ψi−1t ), i = 1, · · · , m (35) for a suitable sequence of stepsizes {αt}t∈N∗ ⊂ ]0, +∞[.

Several versions of incremental subgradient proximal algorithms have been studied in [7], where convergence results for various stepsizes rules and both for stochastic of cyclic selection of the components are given. Concern-ing the function values, the results are stated in terms of the best iterate, i.e., (27). See also [32] for the study of the special case of incremental subgradient methods under different stepsizes rules. The paper [24] provides convergence results using approximate subgradients instead of gradients.

In this section, we derive a sublinear convergence rate for the incremen-tal subgradient proximal algorithm in a straightforward way, relying on the properties of modified Fej´er sequences assuming a boundedness assumption on the subdifferentials, already used in [32].

Theorem 5. Let α ∈ ]0, +∞[, let θ ∈ [0, 1[, and let, for every t ∈ N∗, α t=

αt−θ. Let (x

t)t∈N∗ be the sequence generated by Algorithm 3. Let B ∈ ]0, +∞[

be such that

(∀t ∈ N∗)(∀g ∈ ∂ℓi(xt) ∪ ∂ri(xt)) kgk ≤ B.

Then, there exists c ∈ ]0, +∞[ such that, for every T ∈ N∗,

f (xT) − f∗ ≤      d(x1, X )2 2α T θ−1+cα(4m + 5)mB 2 2 T −θlog T if θ ≤ 1/2  d(x1, X ) 2 2α + cα(4m + 5)mB2 2  Tθ−1 otherwise. (36) Proof. It was shown in [7, Proposition 3 (Equation 27)] that,

kxt+1− xk 2 ≤ kxt− xk 2 − 2αt[f (xt) − f (x)] + α 2 t(4m + 5) mB 2 .

Thus, (xt)t∈N∗ is a modified Fej´er sequence with respect to the objective

function f , and ((2αt, α2t(4m + 5) mB 2

(19)

applying Remark 4 with θ1 = θ, θ2 = 2θ, η = 2α and ξ = α2(4m + 5) mB2.

Remark 7.

(i) The choice θ = 1/2 in Theorem 5, yields a convergence rate of order O(T−1/2log T ) for the objective function values.

(ii) In [7, Proposition 5] a bound similar to (36) is derived for the best it-erate (27) with a fixed stepsize. In contrast to this previous result, our result holds for any last iterate, considering both the fixed and diminish-ing stepsize settdiminish-ing. Note that, neither [7, Proposition 5] nor Theorem 5 imply convergence for the fixed stepsize.

As in Theorem 5, we can derive convergence rates for the projected in-cremental subgradient method. Analogously to what we have done for the forward-backward algorithm in Section 3.1, Theorem 5 can be extended to analyze convergence of the approximate and incremental subgradient method in [24].

3.4

Douglas-Rachford splitting method

In this subsection, we consider Douglas-Rachford splitting algorithm for solv-ing (1). Given ℓ : H → ]−∞, +∞] and r : H → ]−∞, +∞] convex and lower semicontinuous functions, we suppose that f = ℓ + r in (1).

Algorithm 4. Let (αt)t∈N∗ ∈ ]0, +∞[

N

. Let t ∈ N∗. Given x

t ∈ H, an

iteration of Douglas-Rachford algorithm generates xt+1 according to

   yt+1 = proxαtℓ(xt) zt+1= proxαtr(2yt+1− xt), xt+1= xt+ zt+1− yt+1. (37)

The algorithm has been introduced in [21] to solve matrix equations. Then it has been extended for solving the minimization problem of the sum of two convex functions [27], and then to monotone inclusions involving the sum of two nonlinear operators [29]. A review of this algorithm can be found in [14]. The convergence of the iterates is established using the theory of

(20)

Fej´er sequences [13]. Our objective here is to establish a new result, namely a convergence rate for the objective function values.

Theorem 6. Let α ∈ ]0, +∞[, and let θ ∈ [0, 1[. For every t ∈ N∗, let

αt = αt−θ. Let (yt, xt, zt)



t∈N∗ be the sequences generated by Algorithm 4.

Assume that there exists B ∈ ]0, +∞[ such that, for every t ∈ N∗:

(∀v ∈ ∂ℓ(yt))(∃ut∈ ∂ℓ(xt))(∃st∈ ∂r(xt))

kvk ≤ B, kutk ≤ B and kstk ≤ B. (38)

Then, there exists c ∈ ]0, +∞[, such that, for every T ∈ N∗,

f (xT +1) − f∗ ≤      d(x1, X )2 2α T θ−1+ 5cαB 2 2 T −θlog T if θ ≤ 1/2  d(x1, X ) 2 2α + 5cαB2 2  Tθ−1 otherwise.

Proof. Let t ∈ N∗, set v

t+1= (xt−yt+1)/αtand wt+1= (2yt+1−xt−zt+1)/αt.

By Fermat’s rule, vt+1 ∈ ∂ℓ(yt+1) and wt+1 ∈ ∂r(zt+1). (39) We can rewrite (37) as    yt+1 = xt− αtvt+1, zt+1= (2yt+1− xt) − αtwt+1, xt+1 = xt+ zt+1− yt+1, (40)

By (40), we have for any x ∈ domf,

ℓ(yt+1) − ℓ(x) ≤ hyt+1− x, vt+1i.

Multiplying both sides by 2αt,

2αt[ℓ(yt+1) − ℓ(x)] ≤ 2αthyt+1− x, vt+1i = 2hyt+1− x, xt− yt+1i.

Similarly, we have

(21)

Combining the above two estimates, we get

2αt[ℓ(yt+1) + r(zt+1) − ℓ(x) − r(x)]

≤ 2hyt+1− x, xt− yt+1i + 2hzt+1− x, 2yt+1− xt− zt+1i.

The third equality in (40) implies that that zt+1 = xt+1− xt+ yt+1, and thus

2αt[ℓ(yt+1) + r(zt+1) − ℓ(x) − r(x)]

≤ 2hyt+1− x, xt− yt+1i + 2hxt+1− xt+ yt+1− x, 2yt+1− xt+1− yt+1i

= 2hxt− xt+1, xt+1− xi.

= kxt− xk2− kxt+1− xk2− kxt− xt+1k2.

Adding 2αt[ℓ(xt+1) + r(xt+1) − ℓ(yt+1) − r(zt+1)] to both sides, and recalling

that f = ℓ + r,

2αt[f (xt+1) − f (x)] ≤ kxt− xk2− kxt+1− xk2− kxt− xt+1k2

+ 2αt[l(xt+1) + r(xt+1) − l(yt+1) − r(zt+1)]. (41)

Let ut+1 ∈ ∂ℓ(xt+1) and st+1 ∈ ∂r(xt+1) such that kut+1k ≤ B and kst+1k ≤

B. Convexity of ℓ and r, and (40) yield,

ℓ(xt+1) − ℓ(yt+1) ≤ hxt+1− yt+1, ut+1i = hxt+1− xt, ut+1i + hxt− yt+1, ut+1i = hxt+1− xt, ut+1i + αthvt+1, ut+1i ≤ kxt+1− xtkkuk + αtkvt+1kkut+1k ≤ kxt+1− xtkB + αtB2 ≤ kxt+1− xtk2/(2αt) + B2αt/2 + αtB2, and r(xt+1) − r(zt+1) ≤ hxt+1− zt+1, st+1i = αthvt+1, st+1i ≤ αtkvt+1kkst+1k ≤ αtB 2 .

Introducing the last two estimates into (41), and by a direct calculation,

(22)

Thus, (xt)t∈N∗ is a modified quasi-Fej´er sequence with respect to the objective

function f and (2αt, 5α2tB 2

)

t∈N∗. The statement follows from Theorem 2

with θ1 = θ and θ2 = 2θ.

Remark 8.

(i) Condition (38) is verified if ℓ and r are Lipschitz continuous on H. In this case the subdifferential is nonempty at every point and uniformly bounded, see [35, Proposition 1.11].

(ii) Choosing θ = 1/2, we get a convergence rate O(T−1/2log T ) for the

algorithm with nonsummable diminishing stepsizes. Convergence does not follow using a fixed stepsize.

(iii) In the case where ℓ is the indicator function of a linear subspace of H (thus taking also the value +∞) and r is Lipschitz continuous, noner-godic convergence rates for the objective function values corresponding to the Douglas-Rachford iteration can be derived by [20, Corollary 3.5].

4

Concluding remarks

We studied a modified notion of Fej´er monotonicity, providing various con-vergence results for different optimization algorithms. Possible generalization and extension of the proposed notion can be considered, for instance includ-ing the stochastic and the variable metric settinclud-ings.

References

[1] Attouch, H., Bolte, J., and Svaiter, B., Convergence of descent meth-ods for semi-algebraic and tame problems: proximal algorithms, forward-backward splitting, and regularized Gauss-Seidel methods, Math. Pro-gram. Ser. A 137, 91-129, 2011.

[2] Alber, Y. I., Iusem, A. N., and Solodov, M. V.: On the projected sub-gradient method for nonsmooth convex optimization in a Hilbert space. Math. Program. 81, 23-35 (1998).

(23)

[3] Bauschke, H. H., and Combettes, P. L.: Convex Analysis and Monotone Operator Theory in Hilbert spaces, Springer, New York, 2011.

[4] Beck, A., and Teboulle, M.: A fast iterative shrinkage-thresholding al-gorithm for linear inverse problems. SIAM J. Imaging Sciences 2, 183-202 (2009).

[5] Bello-Cruz, J.-Y., On proximal subgradient splitting method for mini-mizing the sum of two nonsmooth convex functions, Set-Valued Var. Anal. 25, 245-263 (2017).

[6] Bonettini, S., Benfenati, A., and Ruggiero, V.: Scaling techniques for ǫ-subgradient methods, SIAM J. Optim. 26, 891-921 (2016).

[7] Bertsekas, D. P.: Incremental proximal methods for large scale convex optimization. Math. Program., Ser. B 129, 163-195 (2009).

[8] Boyd, S., Xiao, L., and Mutapcic, A.: Subgradient methods. https://web.stanford.edu/class/ee392o/subgrad method.pdf. Accessed 14 October 2015.

[9] Bredies, K. and Lorenz, D. A.: Linear convergence of iterative soft-thresholding. J. Fourier Anal. Appl. 14, 813-837 (2008).

[10] Chen, G. H., and Rockafellar, R. T.: Convergence rates in forward– backward splitting. SIAM J. Optim. 7, 421-444 (1997).

[11] Combettes, P. L.: Quasi-Fej´erian analysis of some optimization algo-rithms. Stud. Comput. Math. 8, 115-152 (2001).

[12] Combettes, P.L. and Wajs, V.R.: Signal recovery by proximal forward-backward splitting. Multiscale Model. Simul. 4, 1168-1200 (2005).

[13] Combettes, P. L.: Fej´er monotonicity in convex optimization. In: C. A. Floudas and P. M. Pardalos (eds.) Encyclopedia of Optimization (pp. 1016-1024). Springer, New York (2009).

(24)

[14] Combettes, P. L., and Pesquet, J. C.: Proximal splitting methods in sig-nal processing. In: Fixed-point algorithms for inverse problems in science and engineering (pp. 185-212). Springer, New York, 2011.

[15] Combettes, P.L., and Pesquet, J.C.: Stochastic Quasi-Fej´er Block-Coordinate Fixed Point Iterations with Random Sweeping, SIAM J. Op-tim. 25 1221-1248 (2015).

[16] Combettes, P.L., and Pesquet, J.C.: Stochastic Quasi-Fej´er Block-Coordinate Fixed Point Iterations with Random Sweeping II:Mean-square and linear convergence, https://arxiv.org/abs/1704.08083, (2017).

[17] Combettes, P.L., Salzo, S., and Villa, S.: Consistent learning by compos-ite proximal thresholding, Mathematical Programming, published online 2017-03-25.

[18] Combettes, P. L., and V˜u, B.C.: Variable metric quasi Fej´er monotonic-ity. Nonlinear Anal. 78, 17-31 (2013).

[19] Darzentas, J.: Problem complexity and method efficiency in optimiza-tion. J. Oper. Res. Soc., 35, 455-455 (1984).

[20] Davis, D.: Convergence rate analysis of the forward-Douglas-Rachford splitting scheme, SIAM J. Optim. 25, 1760-1786 (2015).

[21] Douglas, J., and Rachford, H. H.: On the numerical solution of heat conduction problems in two and three space variables. Trans. Amer. Math. Soc. 82, 421-439 (1956).

[22] Ermol’ev, Yu. M. and Tuniev, A. D.: Random Fej´er and quasi-Fej´er sequences, Theory of Optimal Solutions – Akademiya Nauk Ukrainsko˘ı SSR Kiev 2, 76–83 (1968) ; translated in: American Mathematical Society Selected Translations in Mathematical Statistics and Probability 13,143-148 (1973).

[23] Goffin, J. L.: On convergence rates of subgradient optimization methods. Math. Program. 13, 329-347 (1977).

(25)

[24] Kiwiel, K. C.: Convergence of approximate and incremental subgradient methods for convex optimization. SIAM J. Optim. 14, 807-840 (2004).

[25] Knopp, K. Infinite Sequences and Series, Dover Publications Inc., New York 1956.

[26] Larsson, T., Patriksson, M., Stromberg, A.-B.: On the convergence of conditional ǫ-subgradient methods for convex programs and convex-concave saddle-point problems, EJOR 151, 461-473, 2003.

[27] Lieutaud, J.: Approximation d’Op´erateurs par des M´ethodes de D´ecomposition. Th`ese, Universit´e de Paris (1969).

[28] Lin J., Rosasco L., and Zhou D. X.: Iterative regularization for learning with convex loss functions. Journal of Machine Learning Research 17, 1-38 (2016).

[29] Lions, P. L., and Mercier, B.: Splitting algorithms for the sum of two nonlinear operators. SIAM J. Numer. Anal. 16, 964-979, (1979).

[30] Moreau, J.J.: Fonctions convexes duales et points proximaux dans un espace hilbertien, C. R. Acad. Sci. Paris 255, 2897–2899, (1962).

[31] Nedic, A., and Bertsekas, D. P.: Incremental subgradient methods for nondifferentiable optimization. SIAM J. Optim. 12, 109-138, (2001).

[32] Nedic, A., and Bertsekas, D.: Convergence rate of incremental subgra-dient algorithms. In Stochastic optimization: algorithms and applications (pp. 223-264). Springer, New York 2001.

[33] Nesterov, Y.: Introductory Lectures on Convex Optimization. Springer, New York, 2004.

[34] Passty, G. B.: Ergodic convergence to a zero of the sum of monotone operators in Hilbert space. J. Math. Anal. Appl. 72, 383–390 (1979).

[35] R. R. Phelps, Convex Functions, Monotone Operators and Differentia-bility, 2nd ed. Springer, New York, 1993.

(26)

[36] Polyak, B. T.: Introduction to Optimization. Optimization Software, New York 1987.

[37] Rosasco, L., Villa, S., and V˜u, B.C., A stochastic forward-backward splitting method for solving monotone inclusions in Hilbert spaces, J. Op-tim. Theory Appl. 169, 388-406 (2016).

[38] Rosasco, L., Villa, S., and V˜u, B.C., A stochastic inertial forward-backward splitting algorithm for multivariate monotone inclusions, Optim. 65, 1293-1314 (2016).

[39] Rosasco, L., Villa, S., and V˜u, B.C., A first-order stochastic primal-dual algorithm with correction step, Numer. Funct. Anal. Optim. 38, 602-626 (2017).

[40] Salzo, S.: The variable metric forward-backward splitting algorithm un-der mild differentiability assumptions arXiv1605.00952v1, 2016

[41] Shor, N. Z.: Minimization Methods for Non-Differentiable Functions. Springer, New York, 1979.

[42] Singer, Y., and Duchi, J. C.: Efficient learning using forward-backward splitting. In Advances in Neural Information Processing Systems (pp. 495-503) (2009).

[43] Tseng, P.: Applications of a splitting algorithm to decomposition in convex programming and variational inequalities. SIAM J. Control Optim. 29, 119-138 (1991).

[44] C. Z˘alinescu, Convex Analysis in General Vector Spaces. World Scien-tific, River Edge, 2002.

Riferimenti

Documenti correlati

Cartesian reference system belonging to the triangle

We introduce a Gel’fand-Levitan-Marchenko theory, discuss some energy oscillation phenomena and provide some physical applications of an appropriate trace formula

If the ultimate end of representation using perspective – and all of the classic delineation derived from it – was founded on the attempt to (re)construct – or reproduce – space

In this work we discuss the prospects of using UAVs for these mentioned activities: in Section 2 we consider two concepts to measure the environmental parameters and the seeing;

Here, we employ the concept of ring minimal surface to quantify the extent of threadings in polymer solutions of the double-folded rings vs rings in equilibrated molecular

All’interno dell’unità di maglia urbana o dell’insediamento si trova un siste- ma di spazi vuoti privati, semiprivati o semipubblici, che danno aria, luce, assolvono a

The inverse scattering problem for the Schr¨ odinger equation on the line, first solved by Faddeev [45] and presented in [40, 43, 32, 75], consists of recovering the potential u(x)

Among the regulatory T cell population, two different subsets have been described on the basis of their distinct suppressive mechanisms: naturally occurring CD4 + CD25 +