Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and...

23
Non-linear conjugate gradient methods for vector optimization * L. R. Lucambio P´ erez L. F. Prudente April 20, 2017 Abstract In this work, we propose non-linear conjugate gradient methods for finding critical points of vector-valued functions with respect to the partial order induced by a closed, convex, and pointed cone with non-empty interior. No convexity assumption is made on the ob- jectives. The concepts of Wolfe and Zoutendjik conditions are extended for the vector- valued optimization. In particular, we show that there exist intervals of stepsizes satisfying the Wolfe-type conditions. The convergence analysis covers the vector extensions of the Fletcher-Reeves, Conjugate Descent, Dai-Yuan, Polak-Ribi` ere-Polyak, and Hestenes-Stiefel parameters that retrieve the classical ones in the scalar minimization case. Under inexact line searches and without regular restarts, we prove that the sequences generated by the proposed methods find points that satisfy the first-order necessary condition for Pareto- optimality. Keywords: Vector optimization; Pareto-optimality; conjugate gradient method; uncon- strained optimization; line search algorithm; Wolfe conditions. Mathematical Subject Classification(2010): 90C29 90C52 1 Introduction In vector optimization, one leads with a vector-valued function F : R n R m , and the partial order defined by a closed, convex, and pointed cone K R m with non-empty interior, denoted by int(K). The partial order K (K ) is given by u K v (u K v) if and only if v - u K (v - u int(K)). We are interested in the unconstrained vector problem Minimize K F (x), x R n , (1) where F is continuously differentiable. The concept of optimality has to be replaced by the concept of Pareto-optimality or efficiency. A point x is called Pareto-optimal if there is no y such that y 6= x and F (y) K F (x). A particular case, very important in the practical applications, is when K = R m + . This case corresponds to the multicriteria or multiobjective optimization. In multicriteria optimization, several objective functions have to be minimized simultaneously. In proper problems of this class, no single point minimizes all objectives at once. When K = R m + , a point is Pareto-optimal if there does not exist a different point with smaller or equal objective function values, such that there is a decrease in at least one objective. In other words, it is impossible to go better in one objective without going worse in another. Some applications of multicriteria and vectorial optimization can be found in engineering design [34], space exploration [45], antenna design [33–35], management science [22, 29, 46, 48], * This work was partially supported by CAPES and FAPEG. Instituto de Matem´ atica e Estat´ ıstica, Universidade Federal de Goi´ as, Avenida Esperan¸ ca, s/n, Campus Samambaia, Goiˆ ania, GO - 74690-900, Brazil (E-mails: [email protected], [email protected]). 1

Transcript of Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and...

Page 1: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

Non-linear conjugate gradient methods for vector optimization∗

L. R. Lucambio Perez† L. F. Prudente†

April 20, 2017

Abstract

In this work, we propose non-linear conjugate gradient methods for finding critical pointsof vector-valued functions with respect to the partial order induced by a closed, convex,and pointed cone with non-empty interior. No convexity assumption is made on the ob-jectives. The concepts of Wolfe and Zoutendjik conditions are extended for the vector-valued optimization. In particular, we show that there exist intervals of stepsizes satisfyingthe Wolfe-type conditions. The convergence analysis covers the vector extensions of theFletcher-Reeves, Conjugate Descent, Dai-Yuan, Polak-Ribiere-Polyak, and Hestenes-Stiefelparameters that retrieve the classical ones in the scalar minimization case. Under inexactline searches and without regular restarts, we prove that the sequences generated by theproposed methods find points that satisfy the first-order necessary condition for Pareto-optimality.

Keywords: Vector optimization; Pareto-optimality; conjugate gradient method; uncon-strained optimization; line search algorithm; Wolfe conditions.

Mathematical Subject Classification(2010): 90C29 90C52

1 Introduction

In vector optimization, one leads with a vector-valued function F : Rn → Rm, and the partialorder defined by a closed, convex, and pointed cone K ⊂ Rm with non-empty interior, denotedby int(K). The partial order �K (≺K) is given by u �K v (u ≺K v) if and only if v − u ∈ K(v − u ∈ int(K)). We are interested in the unconstrained vector problem

MinimizeK F (x), x ∈ Rn, (1)

where F is continuously differentiable. The concept of optimality has to be replaced by theconcept of Pareto-optimality or efficiency. A point x is called Pareto-optimal if there is noy such that y 6= x and F (y) ≺K F (x). A particular case, very important in the practicalapplications, is when K = Rm+ . This case corresponds to the multicriteria or multiobjectiveoptimization. In multicriteria optimization, several objective functions have to be minimizedsimultaneously. In proper problems of this class, no single point minimizes all objectives atonce. When K = Rm+ , a point is Pareto-optimal if there does not exist a different point withsmaller or equal objective function values, such that there is a decrease in at least one objective.In other words, it is impossible to go better in one objective without going worse in another.

Some applications of multicriteria and vectorial optimization can be found in engineeringdesign [34], space exploration [45], antenna design [33–35], management science [22, 29, 46, 48],

∗This work was partially supported by CAPES and FAPEG.†Instituto de Matematica e Estatıstica, Universidade Federal de Goias, Avenida Esperanca, s/n, Campus

Samambaia, Goiania, GO - 74690-900, Brazil (E-mails: [email protected], [email protected]).

1

Page 2: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

environmental analysis, cancer treatment planning, bilevel programming [21], location science[36], statistics [16], etc.

In recent papers, descent, steepest descent, Newton, Quasi-Newton, projected gradient,subgradient, interior point, and proximal methods were proposed for vector optimization, see[2, 3, 8, 19, 20, 23, 24, 26–28, 37, 44, 47]. The vector optimization problem in infinite dimensionwas considered in [4–7,9,10]. In this work, we propose a conjugate gradient method for the un-constrained vector optimization problem (1). Conjugate gradient methods, originally proposedin [18], constitute an important class of first-order algorithms for solving the unconstrainedoptimization problem

Minimize f(x), x ∈ Rn, (2)

where f : Rn → R is continuous differentiable. Owing to the efficiency of the algorithms,particularly in large dimensions, the extension to the vector case appears naturally.

The conjugate gradient algorithm for the scalar problem (2) generates a sequence of iteratesaccording to

xk+1 = xk + αkdk, k ≥ 0, (3)

where dk ∈ Rn is the line search direction, and αk > 0 is the step-size. The direction is definedby

dk =

{−∇f(xk), if k = 0,−∇f(xk) + βkd

k−1, if k ≥ 1,

where βk is a scalar algorithmic parameter. When f is strictly convex quadratic with Hessian Q,parameters βk can be chosen such that the dk are Q-conjugate, which means that (dk)TQdk+1 =0 for all k ≥ 0, and the stepsizes can be defined by

αk = arg min{f(xk + αdk) | α > 0

},

i.e., by exact minimizations in the line searches. In such a case the algorithm finds the minimizerof f over Rn after no more than n iterations, and it is called the linear conjugate gradient algo-rithm. For non-quadratic functions, different formulae for the parameters βk result in differentalgorithms, known as non-linear conjugate gradient methods. Some notable choices for βk areas follows:

Fletcher-Reeves (FR) [18]: βk = 〈gk, gk〉/〈gk−1, gk−1〉;

Conjugate Descent (CD) [17]: βk = −〈gk, gk〉/〈dk−1, gk−1〉;

Dai-Yuan (DY) [15]: βk = 〈gk, gk〉/〈dk−1, gk − gk−1〉;

Polak-Ribiere-Polyak (PRP) [41,42]: βk = 〈gk, gk − gk−1〉/〈gk−1, gk−1〉;

Hestenes-Stiefel (HS) [32]: βk = 〈gk, gk − gk−1〉/〈dk−1, gk − gk−1〉;

where gk = ∇f(xk) and 〈·, ·〉 denotes the usual inner product. A desirable property that theparameters βk should have is that the directions dk should be descent directions in the sensethat 〈∇f(xk), dk〉 < 0 for all k ≥ 0. In some convergence analyses, the more stringent sufficientdescent condition is required, namely,

〈∇f(xk), dk〉 ≤ −c‖∇f(xk)‖2 for some c > 0 and k ≥ 0,

where ‖ · ‖ denotes the Euclidian norm. For general functions, the exact line search means thatthe step-size αk is a stationary point of the problem of minimizing f at xk along direction dk,implying that

〈∇f(xk + αkdk), dk〉 = 0.

2

Page 3: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

Since the exact line search is usually expensive and impractical, one may be satisfied withstepsizes satisfying, for example, the standard Wolfe conditions

f(xk + αkdk) ≤ f(xk) + ραk〈∇f(xk), dk〉,

〈∇f(xk + αkdk), dk〉 ≥ σ〈∇f(xk), dk〉

or the strong Wolfe conditions,

f(xk + αkdk) ≤ f(xk) + ραk〈∇f(xk), dk〉,∣∣∣〈∇f(xk + αkd

k), dk〉∣∣∣ ≤ σ∣∣∣〈∇f(xk), dk〉

∣∣∣where 0 < ρ < σ < 1. Under mild particularities the FR, CD, and DY methods with a linesearch satisfying the Wolfe conditions generate descent directions. This is not the case of PRPand HS methods. Furthermore, Powell showed in [43] that the PRP and HS methods with exactline searches can cycle without approaching a solution point. Throughout this paper, a methodis called globally convergent if

lim infk→∞

‖∇f(xk)‖ = 0. (4)

In the convergence analysis, an important step is to prove that the method satisfies the Zou-tendjik condition

∞∑k=0

〈∇f(xk), dk〉2

‖dk‖2<∞.

The concepts of Wolfe and Zoutendjik conditions are widely used to derive global convergenceresults of several line search algorithms for the scalar problem (2), particularly for non-linearconjugate gradient methods, see for example [40]. The convergence, in the sense of (4), ofthe FR and DY methods under the Wolfe conditions were established by Al-Baali [1] and Daiand Yuan [15], respectively. Dai and Yuan also studied the behavior of CD method in [14].In [25], Gilbert and Nocedal established the convergence of the conjugate gradient methodwith βk = max{βPRPk , 0}, and with βk = max{βHSk , 0}. Other choices for βk and some hybridmethods can be found in the literature. We suggest [11–13, 25, 31] for interesting reviews ofthese subjects.

In the context of vector optimization, we introduce the standard and strong Wolfe conditions.We say that d is a K-descent direction for F at x if there exists T > 0 such that 0 < t < Timplies that F (x + td) ≺K F (x). When d is a K-descent direction, we show that there existintervals of stepsizes satisfying the Wolfe conditions. This new theoretical result sheds light onalgorithmic properties and suggests the implementation of a Wolfe-type line search procedure.We also introduce the Zoutendjik condition for vector optimization and prove that it is satisfiedfor a general descent line search method. The considered assumptions are natural extensions ofthose made for the scalar case (2).

We present the general scheme of a non-linear conjugate gradient method for vector opti-mization, and study its convergence for different choices of parameter βk. The analysis covers thevector extensions of all the aforementioned parameters: FR, CD, DY, PRP, and HS. We pointout that the proposed vector extensions retrieve the classical parameters in the scalar minimiza-tion case (2). Under quite a reasonable hypothesis, we show that the sequences produced by themethods find points that satisfy the first-order necessary condition for Pareto-optimality. Weemphasize that the Wolfe and Zoutendjik conditions are essential tolls to prove the convergenceresults.

This paper is organized as follows. In Section 2, we summarize notions, notation, andresults related to unconstrained vector optimization that we use. In Section 3, we introducethe Wolfe and Zoutendjik conditions. We prove that there exist intervals of stepsizes satisfyingWolfe conditions, and that the Zoutendjik condition is satisfied for a general descent line search

3

Page 4: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

method. In Section 4, we present the general schemes for non-linear conjugate gradient methodsfor vector optimization, and study its convergence without imposing explicit restriction on βk.In Section 5, we analyze the convergence of the conjugate gradient algorithm related to thevector extensions of FR, CD, DY, PRP, and HS parameters. Finally, we make some finalremarks in Section 6.

2 General concepts in Vector Optimization

In this section we summarize the definitions, results, and notation of vector optimization. Werefer the reader to [20,26,28,38] and references therein. In [28], Grana Drummond and Svaiterhave developed quite useful tools for the vector optimization. In particular, they characterizedthe directional derivative and the steepest descent direction, which are concepts widely used inthe present work.

Let K ⊂ Rm be a closed, convex, and pointed cone with non-empty interior. The partialorder in Rm induced by K, �K , is defined by

u �K v ⇔ v − u ∈ K,

and the partial order defined by int(K), the interior of K, ≺K , is defined by

u ≺K v ⇔ v − u ∈ int(K).

Given a continuously differentiable function F : Rn → Rm, we are concerned with the prob-lem of finding an unconstrained K-minimizer, K-optimum, K-Pareto-optimal or K-efficientpoint of F , i.e., a point x∗ ∈ Rn such that there exists no other x ∈ Rn with F (x) �K F (x∗)and F (x) 6= F (x∗). In other words, we are seeking unconstrained minimizers for F with respectto the partial order induced by K. We denoted this problem by

MinimizeK F (x), x ∈ Rn.

The first-order derivative of F at x, the Jacobian of F at x, will be denoted by JF (x), andthe image on Rm by JF (x) will be denoted by Image(JF (x)). A necessary condition for K-optimality of x∗ is

−int(K) ∩ Image(JF (x∗)) = ∅.A point of Rn is called K-critical for F when it satisfies such condition. Therefore, if x is notK-critical, there exists v ∈ Rn such that JF (x)v ∈ −int(K). Every such vector v is a K-descentdirection for F at x, i.e., there exists T > 0 such that 0 < t < T implies that F (x+tv) ≺K F (x),see [38].

The positive polar cone of K is the set K∗ ={w ∈ Rm | 〈w, y〉 ≥ 0, ∀y ∈ K

}. Since K

is closed and convex, K = K∗∗, −K ={y ∈ Rm | 〈y, w〉 ≤ 0, ∀w ∈ K∗

}and −int(K) ={

y ∈ Rm | 〈y, w〉 < 0, ∀w ∈ K∗ − {0}}

. Let C ⊂ K∗ − {0} be compact and such that

K∗ = cone(conv(C)),

i.e., K∗ is the conic hull of the convex hull of C. In classical optimization K = R+, thenK∗ = R+ and we can take C = {1}. For multiobjective optimization K = Rm+ , then K∗ = Kand we may take C as the canonical basis of Rm. If K is a polyhedral cone, K∗ is also polyhedraland C can be taken as the finite set of extremal rays of K∗. For generic K, the set

C ={w ∈ K∗ | ‖w‖ = 1

},

has the desired properties, and, in this paper, C will denote exactly that.Define the function ϕ : Rm → R by

ϕ(y) = sup{〈y, w〉 | w ∈ C

}.

In view of the compactness of C, ϕ is well-defined. Function ϕ has some useful properties.

4

Page 5: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

Lemma 1. Let y and y′ ∈ Rm. Then:

(a) ϕ(y + y′) ≤ ϕ(y) + ϕ(y′) and ϕ(y)− ϕ(y′) ≤ ϕ(y − y′);

(b) If y �K y′, then ϕ(y) ≤ ϕ(y′); if y ≺K y′, then ϕ(y) < ϕ(y′);

(c) ϕ is Lipschitz continuous with constant 1.

Proof. See Lemma 3.1 in [28].

Function ϕ gives characterizations of −K and −int(K):

−K ={y ∈ Rm | ϕ(y) ≤ 0

}and − int(K) =

{y ∈ Rm | ϕ(y) < 0

}.

Note that ϕ(x) > 0 does not imply that x ∈ K, but x ∈ K implies that ϕ(x) ≥ 0 and x ∈ int(K)implies that ϕ(x) > 0.

Now define the function f : Rn × Rn → R by

f(x, d) = ϕ(JF (x)d) = sup{〈JF (x)d,w〉 | w ∈ C

}.

Function f gives a characterization of K-descent directions and of K-critical points:

• d is K-descent direction for F at x if and only if f(x, d) < 0;

• x is K-critical if and only if f(x, d) ≥ 0 for all d.

The next function allows us to extend the notion of steepest descent direction to the vectorcase. Define v : Rn → Rn by

v(x) = arg min

{f(x, d) +

‖d‖2

2| d ∈ Rn

},

and θ : Rn → R by θ(x) = f(x, v(x)) + ‖v(x)‖2/2. Since f(x, ·) is a real closed convex function,v(x) exists and is unique. Observe that in the scalar minimization case, where F : Rn → Rand K = R+, taking C = {1}, we obtain f(x, d) = 〈∇F (x), d〉, v(x) = −∇F (x) and θ(x) =−‖∇F (x)‖2/2. The following Lemma shows that v(x) can be considered the vector extensionof the steepest descent direction of the scalar case.

Lemma 2. (a) If x is K-critical, then v(x) = 0 and θ(x) = 0.

(b) If x is not K-critical, then v(x) 6= 0, θ(x) < 0, f(x, v(x)) < −‖v(x)‖22 < 0, and v(x) is a

K-descent direction for F at x.

(c) The mappings v and θ are continuous.

Proof. See Lemma 3.3 in [28].

For multiobjective optimization, where K = Rm+ , with C given by the canonical basis of Rm,v(x) can be computed by solving

Minimize α+1

2‖d‖2

subject to [JF (x)d]i ≤ α, i = 1, . . . ,m,

which is a convex quadratic problem with linear inequality constraints, see [20]. For generalproblems, we refer the reader to [28] for a detailed discussion regarding this issue.

The next result is a direct consequence of Lemma 1-(c).

5

Page 6: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

Lemma 3. Let be F : Rn → Rm differentiable. If the Jacobian of F is Lipschitz continuouswith constant L, then f(·, d) is Lipschitz continuous with constant L‖d‖.

We ended this section with the following observations.

Lemma 4. For any scalars a, b, and α 6= 0, we have:

(a) ab ≤ a2/2 + b2/2;

(b) 2ab ≤ 2α2a2 + b2/(2α2);

(c) (a+ b)2 ≤ 2a2 + 2b2;

(d) (a+ b)2 ≤ (1 + 2α2)a2 + [1 + 1/(2α2)]b2.

Proof. (a) We have 0 < (a− b)2 = a2− 2ab+ b2 which implies ab ≤ a2/2 + b2/2. (b) The resultfollows from (a) replacing a by 2αa, and b by b/α. Items (c)-(d) follow directly from (a) and(b), respectively.

3 On Line Search in Vector Optimization methods

Line search procedures are crucial tools in methods for optimization. Concepts such as Armijoand Wolfe conditions are widely used in algorithms for minimization. In classical optimiza-tion (2), for an iterative method such as (3), the Armijo condition requires the step-size αk tosatisfy

f(xk + αkdk) ≤ f(xk) + ραk〈∇f(xk), dk〉.

In [28], the authors extended the Armijo condition for vector optimization, and used it to studythe convergence of the steepest descent method. If d is a K-descent direction for F at x, it issaid that α > 0 satisfies the Armijo condition for 0 < ρ < 1 if

F (x+ αd) �K F (x) + ραJF (x)d.

Other vector optimization methods, including Newton, Quasi-Newton, and projected gradientmethods, that use the Armijo rule can be found in [8,19,20,23,24,26,27,37,44]. In connectionwith the scalar case, we say that α > 0 is obtained by means of an exact line search if

f(x+ αd, d) = 0.

Now we introduce the Wolfe-like conditions for vector optimization.

Definition. Let d ∈ Rn be a K-descent direction for F at x, and e ∈ K a vector such that

0 < 〈w, e〉 ≤ 1 for all w ∈ C.

Consider 0 < ρ < σ < 1. We say that α > 0 satisfies the standard Wolfe conditions if

F (x+ αd) �K F (x) + ραf(x, d)e, (5a)

f(x+ αd, d) ≥ σf(x, d), (5b)

and we say that α > 0 satisfies the strong Wolfe conditions if

F (x+ αd) �K F (x) + ραf(x, d)e, (6a)∣∣f(x+ αd, d)∣∣ ≤ σ∣∣f(x, d)

∣∣ . (6b)

6

Page 7: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

Note that vector e actually exists. Indeed, take any e ∈ int(K) and let 0 < γ− < γ+ bethe minimum and maximum values of the continuous linear function 〈w, e〉 over the compactset C. Defining e = e/γ+, we have 0 < γ−/γ+ ≤ 〈w, e〉 ≤ 1 for all w ∈ C. In multiobjectiveoptimization where K = Rm+ and C is the canonical basis of Rm, we may take e = [1, . . . , 1]T ∈Rm.

Next, we show that given F , K, x and d, where F is bounded below and continuous differen-tiable, K is a finitely generated cone, x is in the domain of F , and d is a K-descent direction forF at x, there exist intervals of stepsizes satisfying the standard and the strong Wolfe conditions.

Proposition 1. Assume that F is of class C1, d is a K-descent direction for F at x, and thereexists F ∈ Rm such that

F (x+ αd) �K F , (7)

for all α > 0. If C, the generator of K, is finite, then there exist intervals of positive stepsizessatisfying the standard Wolfe conditions (5) and the strong Wolfe conditions (6).

Proof. Given w ∈ C, define φw and lw : R→ R by

φw(α) = 〈w,F (x+ αd)〉,

andlw(α) = 〈w,F (x)〉+ αρf(x, d).

Observe that φw(0) = lw(0). Considering (7), we have that φw(α) is bounded below for allα > 0. By the definition of f , we have f(x, d) < 0, because d is a K-descent direction for F atx. Since ρ ∈ (0, 1), as in [40, Lemma 3.1], the line lw is unbounded below and must intersectthe graph of φw at least once for a positive α. Function φw is continuous differentiable becauseF is so. Hence, there exists Tw > 0 such that

〈w,F (x+ Twd)〉 = 〈w,F (x)〉+ Twρf(x, d),

and〈w,F (x+ αd)〉 < 〈w,F (x)〉+ αρf(x, d)

for α ∈ (0, Tw). Define T = min{Tw | w ∈ C}, and let w ∈ C be such that Tw = T . SinceT ≤ Tw for all w ∈ C, we obtain

〈w,F (x+ αd)〉 ≤ 〈w,F (x)〉+ αρf(x, d)

≤ 〈w,F (x)〉+ αρf(x, d)〈w, e〉

for all w ∈ C and α ∈ [0, T ], because f(x, d) < 0 and 0 < 〈w, e〉 ≤ 1. Therefore, (5a) and (6a)hold in [0, T ].

By the mean value theorem, there is tw ∈ (0 , Tw) such that

φ′w(tw) =φw(Tw)− φw(0)

Tw=〈w,F (x+ Twd)〉 − 〈w,F (x)〉

Tw= ρf(x, d). (8)

Define α∗ = min{tw | w ∈ C}, and let w∗ ∈ C be such that tw∗ = α∗. Note that α∗ ≤ tw <Tw = T . Assume that there exists w ∈ C such that

φ′w(α∗) = 〈w, JF (x+ α∗d)d〉 > 〈w∗, JF (x+ α∗d)d〉 = φ′w∗(α∗).

Then, by (8), and taking into account that f(x, d) < 0 and ρ ∈ (0, 1), we obtain

φ′w(α∗) > ρf(x, d) > f(x, d) ≥ 〈w, JF (x)d〉 = φ′w(0).

7

Page 8: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

By the intermediate value theorem, there exists α ∈ (0, α∗) such that φ′w(α) = ρf(x, d), that is,

〈w, JF (x+ αd)d〉 = ρf(x, d),

which is impossible by the definition of α∗. We conclude that

〈w, JF (x+ α∗d)d〉 ≤ 〈w∗, JF (x+ α∗d)d〉, ∀w ∈ C.

Thenf(x+ α∗d, d) = ρf(x, d),

because 〈w∗, JF (x+ α∗d)d〉 = ρf(x, d). Since 0 < ρ < σ, and f(x, d) < 0 we have

σf(x, d) < f(x+ α∗d, d) < 0.

By the continuity of f(x + αd, d) as a function of α, there is a neighborhood of α∗ containedin [0, T ], for which inequalities (5b) and (6b) hold. Therefore, the standard and strong Wolfeconditions hold in this neighborhood, completing the proof.

Now consider a general line search method for vector optimization

xk+1 = xk + αkdk, k ≥ 0 (9)

where dk is a K-descent direction for F at xk, and αk is the step-size. Let us suppose that thefollowing assumptions are satisfied.

Assumptions.

(A1) The cone K is finitely generated and there exists an open set N such that L ={x ∈ Rn |

F (x) �K F (x0)}⊂ N and the Jacobian JF is Lipschitz continuous on N with constant

L > 0, i.e., x and y ∈ N implies that ‖JF (x)− JF (y)‖ ≤ L‖x− y‖.

(A2) All monotonically non-increasing sequences in F (L) are bounded below, i.e., if the sequence{Gk}k∈N ⊂ F (L) and Gk+1 �K Gk, for all k, then there exists G ∈ Rm such that G �K Gkfor all k.

Under Assumptions (A1) and (A2), we establish that the general method (9) satisfies acondition of Zoutendijk’s type. We emphasize that the considered assumptions are naturalextensions of those made for the scalar case.

Proposition 2. Assume that Assumptions (A1) and (A2) hold. Consider an iteration of theform (9), where dk is a K-descent direction for F at xk and αk satisfies the standard Wolfeconditions (5). Then ∑

k≥0

f2(xk, dk)

‖dk‖2<∞. (10)

Proof. By (5b), Assumption (A1) and Lemma 3, we obtain

(σ − 1)f(xk, dk) ≤ f(xk+1, dk)− f(xk, dk) ≤ Lαk‖dk‖2.

Since f(xk, dk) < 0 and ‖dk‖ 6= 0, we obtain

f2(xk, dk)

‖dk‖2≤ Lαk

f(xk, dk)

σ − 1. (11)

By (5a), we have that {F (xk)}k≥0 is monotone decreasing and that

F (xk+1)− F (x0) �K ρ

k∑j=0

αjJF (xj)dj ,

8

Page 9: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

for all k ≥ 0. Then, by Assumption (A2), since {F (xk)}k≥0 ⊂ L, there exists F ∈ Rn such that

F − F (x0) �K ρk∑j=0

αjJF (xj)dj ,

for all k ≥ 0. Therefore,

〈F − F (x0), w〉 ≤ ρ〈k∑j=0

αjJF (xj)dj , w〉,

for all k ≥ 0 and w ∈ C. Denoting inf{〈F − F (x0), w〉 | w ∈ C

}by γ, we obtain

γ

ρ≤

k∑j=0

αj〈JF (xj)dj , w〉 ≤k∑j=0

αjf(xj , dj) < 0.

We conclude that ∑k≥0

αkf(xk, dk)

σ − 1<∞,

which together with (11) gives∑k≥0

f2(xk, dk)

‖dk‖2<∞.

Remark. The Zoutendijk condition (10) can be used to derive convergence results for severalline search algorithms. In particular, the method of steepest descent converges in the sense that

limk→∞

‖v(xk)‖ = 0,

provided that it uses a line search satisfying the standard Wolfe conditions. In fact, considerthe iterative method (9) with dk = v(xk), and assume that αk satisfies (5) for all k ≥ 0. Ifv(xk) = 0 for any k, the result trivially holds. Suppose now that v(xk) 6= 0 for all k ≥ 0. Sinceθ(xk) = f(xk, v(xk)) + ‖v(xk)‖2/2 < 0, it follows that

f2(xk, dk)

‖dk‖2=f2(xk, v(xk))

‖v(xk)‖2≥ 1

4‖v(xk)‖2.

Summing this inequality over all k, we obtain from the Zoutendijk condition that∑k≥0

‖v(xk)‖2 <∞,

which implies that ‖v(xk)‖ → 0.

4 The Non-linear Conjugate Gradient Algorithm

In the following we define the non-linear conjugate gradient algorithm (NLCG Algorithm) forthe vector optimization problem (1).

NLCG Algorithm.

Given x0 ∈ Rn. Compute v(x0), and initialize k ← 0.

Step 1 If v(xk) = 0, then STOP.

9

Page 10: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

Step 2 Define

dk =

{v(xk), if k = 0v(xk) + βkd

k−1, if k ≥ 1,(12)

where βk is an algorithmic parameter.

Step 3 Compute a step-size αk > 0 by a line search procedure, and set xk+1 = xk + αkdk.

Step 4 Compute v(xk+1), set k ← k + 1, and go to Step 1.

The choice for updating the parameter βk, and the adopted line search procedure remaindeliberately open. In the next section, we consider several choices of βk combined with an ap-propriate line search strategy that result in globally convergent methods. The NLCG algorithmsuccessfully stops if a K-critical point of F is found. Hence, from now on, let us consider thatv(xk) 6= 0 for all k ≥ 0.

Throughout the paper the line search procedure must be such that the step-size satisfiesthe standard or strong Wolfe conditions. Thus, for the well-definedness of the NLCG algo-rithm, Proposition 1 requires dk to be a K-descent direction of F at xk which is equivalent tof(xk, dk) < 0. In some convergence analysis, we will require the more stringent condition

f(xk, dk) ≤ cf(xk, v(xk)), (13)

for some c > 0 and for all k ≥ 0. In connection with the scalar case, we call (13) the suffi-cient descent condition. Condition (13) can be obtained for an adequate line search procedure,provided that dk−1 is a K-descent direction of F at xk−1. In this case, by Proposition 1 aline search along dk−1 can be performed enforcing the (standard or strong) Wolfe conditions toobtain xk. Moreover, the line search procedure can be implemented in such a way that it gives,in the limit, a point xk with f(xk, dk−1) = 0. Now, from (12), and the definition of f , if βk ≥ 0,we have

f(xk, dk) ≤ f(xk, v(xk)) + βkf(xk, dk−1).

Then, if βk is bounded, the line search procedure can be applied to reduce∣∣∣f(xk, dk−1)

∣∣∣ suffi-

ciently and obtain (13). A line search procedure with these characteristics may be coded basedon the work of More and Thuente [39]. We refer the reader to [25] for a careful discussion aboutthis issue in the classical optimization approach.

In the next lemma, we give a sufficient condition on βk for ensuring the descent property ondk.

Lemma 5. Assume that in the NLCG algorithm, the sequence βk is defined so that it has thefollowing property:

βk ∈

[0,∞), if f(xk, dk−1) ≤ 0[0,−f(xk, v(xk))/f(xk, dk−1)

), if f(xk, dk−1) > 0

(14)

or

βk ∈

[0,∞), if f(xk, dk−1) ≤ 0[0,−µf(xk, v(xk))/f(xk, dk−1)

], if f(xk, dk−1) > 0

(15)

for some µ ∈ [0, 1). If property (14) holds, then dk is a K-descent direction for all k. If property(15) holds, then dk satisfies the sufficient descent condition with c = 1− µ for all k.

10

Page 11: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

Proof. Let us show the second statement. Inequality (13) with c = 1−µ clearly holds for k = 0.For k ≥ 1, by the definition of dk, we have

〈JF (xk)dk, w〉 = 〈JF (xk)v(xk), w〉+ βk〈JF (xk)dk−1, w〉, ∀w ∈ Rm.

First suppose that f(xk, dk−1) ≤ 0. Since βk ≥ 0, for all w ∈ C, we obtain

〈JF (xk)dk, w〉 ≤ 〈JF (xk)v(xk), w〉 ≤ f(xk, v(xk)) ≤ (1− µ)f(xk, v(xk)).

Now assume that f(xk, dk−1) > 0. By the definition of f , and taking into account (15), weverify for all w ∈ C that

〈JF (xk)dk, w〉 ≤ f(xk, v(xk)) + βkf(xk, dk−1) ≤ (1− µ)f(xk, v(xk)).

In both cases f(xk, dk) = ϕ(JF (xk)dk) ≤ (1− µ)f(xk, v(xk)). The proof of the first statementcan be similarly obtained.

Without imposing an explicit restriction on βk, the main convergence result of the NLCGalgorithm is as follows. Note that Theorem 1-(ii) is the vector extension of [13, Theorem 2.3].

Theorem 1. Consider the NLCG algorithm. Assume that Assumptions (A1) and (A2) holdand that ∑

k≥0

1

‖dk‖2=∞. (16)

(i) If dk satisfies the sufficient descent condition (13), and αk satisfies the standard Wolfeconditions (5), then lim inf

k→∞‖v(xk)‖ = 0.

(ii) If βk ≥ 0, dk is K-descent direction of F at xk, and αk satisfies the strong Wolfe condi-tions (6), then lim inf

k→∞‖v(xk)‖ = 0.

Proof. Assume by contradiction that there exists a constant γ such that

‖v(xk)‖ ≥ γ for all k ≥ 0.

Let us show item (i). Observe that, by f(xk, v(xk)) + ‖v(xk)‖2/2 < 0 and (13), we obtain

c2γ4

4‖dk‖2≤ c2‖v(xk)‖4

4‖dk‖2≤ c2f2(xk, v(xk))

‖dk‖2≤ f2(xk, dk)

‖dk‖2.

Since, under the hypotheses of item (i), the Zoutendijk condition (10) holds, we are in contra-diction with (16). Then, item (i) is demonstrated.

Now consider item (ii). Remembering (12) we have −βkdk−1 = −dk + v(xk). Then,

0 ≤ β2k‖dk−1‖2 ≤

[‖dk‖+ ‖v(xk)‖

]2≤ 2‖dk‖2 + 2‖v(xk)‖2,

where the second inequality follows from Lemma 4-(c). Hence

‖dk‖2 ≥ −‖v(xk)‖2 +β2k

2‖dk−1‖2. (17)

On the other hand, by the definition of dk and the positiveness of βk, we obtain

f(xk, dk) ≤ f(xk, v(xk)) + βkf(xk, dk−1).

11

Page 12: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

Then

0 ≤ −f(xk, v(xk)) ≤ −f(xk, dk) + βkf(xk, dk−1) ≤ −f(xk, dk)− σβkf(xk−1, dk−1), (18)

because we are assuming that αk satisfies the strong Wolfe conditions, i.e., (6b) holds. From (18)and using Lemma 4-(b) with α = 1, we obtain

f2(xk, v(xk)) ≤ f2(xk, dk) + σ2β2kf

2(xk−1, dk−1) + 2σf(xk, dk)βkf(xk−1, dk−1)

≤ (1 + 2σ2)

[f2(xk, dk) +

β2k

2f2(xk−1, dk−1)

].

Since dk is a K-descent direction for F at xk, we obtain θ(xk) = f(xk, v(xk)) + ‖v(xk)‖2/2 < 0.Hence,

f2(xk, dk) +β2k

2f2(xk−1, dk−1) ≥ c1‖v(xk)‖4, (19)

where c1 =[4(1 + 2σ2)

]−1. Note that

f2(xk, dk)

‖dk‖2+f2(xk−1, dk−1)

‖dk−1‖2=

1

‖dk‖2

[f2(xk, dk) +

‖dk‖2

‖dk−1‖2f2(xk−1, dk−1)

]is, by (17), greater than or equal to

1

‖dk‖2

f2(xk, dk) +

[β2k

2− ‖v(xk)‖2

‖dk−1‖2

]f2(xk−1, dk−1)

and, by (19), this last expression is greater than or equal to

‖v(xk)‖2

‖dk‖2

[c1‖v(xk)‖2 − f2(xk−1, dk−1)

‖dk−1‖2

].

Since, under the hypotheses of item (ii), the Zoutendijk condition (10) holds, and it impliesthat f(xk, dk)/‖dk‖ tends to zero, we have

f2(xk, dk)

‖dk‖2+f2(xk−1, dk−1)

‖dk−1‖2≥ c1

‖v(xk)‖4

‖dk‖2≥ c1γ

4∑k≥0

1

‖dk‖2

for all sufficiently large k. Hence, we are in contradiction with (16) and the proof is complete.

5 Convergence Analysis for Specific βk

In this section, we analyze the convergence properties of the NLCG algorithm related to thevector extensions of FR, CD, DY, PRP, and HS choices for parameter βk. In all cases, theproposed vector extensions retrieve the classical parameters in the scalar minimization case.

For some results, we need to replace the Assumption (A2) by the following stronger hypoth-esis.

Assumption.

(A3) The level set L = {x | F (x) �K F (x0)} is bounded.

If dk is a K-descent direction of F at xk in the NLCG algorithm, we claim that under

Assumption (A3) the sequence{f(xk, v(xk))

}is bounded. Indeed, in this case {xk} ⊂ L and,

by continuity arguments, there are constants γ > 0 and c > 0 such that ‖v(xk)‖ ≤ γ and‖JF (xk)‖ ≤ c for all k ≥ 0. Then, for all w ∈ C and k ≥ 0, it turns out that

0 > f(xk, v(xk)) ≥ 〈JF (xk)v(xk), w〉 ≥ −‖JF (xk)‖‖v(xk)‖ ≥ −cγ, (20)

because ‖w‖ = 1.

12

Page 13: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

5.1 Fletcher–Reeves

The vector extension of the FR parameter is given by

βFRk =f(xk, v(xk))

f(xk−1, v(xk−1)).

In the next theorem, under a suitable hypothesis, we show that it is possible to obtain globalconvergence if the parameter βk is bounded in magnitude by any fraction of βFRk .

Theorem 2. Consider the NLCG algorithm, and let 0 ≤ δ < 1.

(i) Assume that Assumptions (A1) and (A2) hold. If

|βk| ≤ δβFRk ,

dk satisfies the sufficient descent condition (13), and αk satisfies the standard Wolfe con-ditions (5), then lim inf

k→∞‖v(xk)‖ = 0.

(ii) Assume that Assumptions (A1) and (A3) hold. If

0 ≤ βk ≤ δβFRk ,

dk is a K-descent direction of F at xk, and αk satisfies the strong Wolfe conditions (6),then lim inf

k→∞‖v(xk)‖ = 0.

Proof. Assume by contradiction that there exists γ > 0 such that

‖v(xk)‖ ≥ γ for all k ≥ 0. (21)

By (12), and Lemma 4-(d) with a = ‖v(xk)‖, b = ‖dk−1‖ and α = δ/√

2(1− δ2), we have

‖dk‖2 ≤[‖v(xk)‖+|βk| ‖dk−1‖

]2≤ 1

1− δ2‖v(xk)‖2 +

1

δ2β2k‖dk−1‖2.

Thus,

‖dk‖2

f2(xk, v(xk))≤ 1

1− δ2

‖v(xk)‖2

f2(xk, v(xk))+

1

δ2

β2k‖dk−1‖2

f2(xk, v(xk))

≤ 1

1− δ2

‖v(xk)‖2

f2(xk, v(xk))+

‖dk−1‖2

f2(xk−1, v(xk−1)),

because|βk| ≤ δf(xk, v(xk))/f(xk−1, v(xk−1)). Remembering 0 < γ2 ≤ ‖v(xk)‖2 ≤ −2f(xk, v(xk)),we obtain

‖dk‖2

f2(xk, v(xk))≤ 4

(1− δ2)γ2+

‖dk−1‖2

f2(xk−1, v(xk−1)).

Applying this relation repeatedly, it follows that

‖dk‖2

f2(xk, v(xk))≤ 4

(1− δ2)γ2k +

‖d0‖2

f2(x0, v(x0))

≤ 4

(1− δ2)γ2k +

4

γ2=

4

γ2

(1

1− δ2k + 1

).

(22)

Hence,f(xk, v(xk))2

‖dk‖2≥ γ2(1− δ2)

4

1

k + 1− δ2≥ γ2(1− δ2)

4

1

k + 1. (23)

13

Page 14: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

Consider (i). By the sufficient descent condition (13) and (23) we have∑k≥0

f2(xk, dk)

‖dk‖2≥∑k≥0

c2 f2(xk, v(xk))

‖dk‖2≥ c2(1− δ2)γ2

4

∑k≥0

1

k + 1=∞.

We are in contradiction because, under the hypotheses of (i), the Zoutendijk condition (10)holds. Thus, item (i) is demonstrated.

Now let us show item (ii). By (20) we obtain f2(xk, v(xk)) ≤ c2γ2. Therefore, by (22), itfollows that

‖dk‖2 ≤ c1k + c2,

where c1 = 4c2γ2/[γ2(1− δ2)] > 0 and c2 = 4c2γ2/γ2 > 0. Then,∑k≥0

1

‖dk‖2=∞,

which implies by Theorem 1-(ii) that lim infk→∞

‖v(xk)‖ = 0, contradicting (21).

We clearly have the following corollary.

Corollary 1. Assume that Assumptions (A1) and (A3) hold. Consider a NLCG algorithm with

βk = δβFRk , where 0 ≤ δ < 1.

If αk satisfies the strong Wolfe conditions (6), and dk is a K-descent direction of F at xk, thenlim infk→∞

‖v(xk)‖ = 0.

We point out that the descent condition can be enforced while performing the line search.In the scalar minimization case, if |βk| ≤ βFRk and αk satisfies the strong Wolfe conditions (6)with σ < 1/2, it is possible to show that the sufficient descent condition (13) always holds,see [1, 25]. For vector optimization, this is an open question.

5.2 Conjugate Descent

Next, we derive the convergence result related to the CD parameter, which is given by

βCDk =f(xk, v(xk))

f(xk−1, dk−1).

In the following lemma, we show that the NLCG algorithm generates descent directions if0 ≤ βk ≤ βCDk and the line search satisfies the strong Wolfe conditions.

Lemma 6. Consider a NLCG algorithm with 0 ≤ βk ≤ βCDk and suppose that αk satisfiesthe strong Wolfe conditions (6). Then dk satisfies the sufficient descent condition (13) withc = 1− σ.

Proof. The proof is by induction. For k = 0, since d0 = v(x0), f(x0, v(x0)) < 0, and 0 < σ < 1,we see that inequality (13) with c = 1− σ holds. For some k ≥ 1, assume that

f(xk−1, dk−1) ≤ (1− σ)f(xk−1, v(xk−1)) < 0.

Hence, βCDk > 0 and βk is well defined. By the definition of f , and the strong Wolfe condi-tion (6b), we have

f(xk, dk) ≤ f(xk, v(xk)) + βkf(xk, dk−1) ≤ f(xk, v(xk))− σβkf(xk−1, dk−1)

≤ f(xk, v(xk))− σβCDk f(xk−1, dk−1) = (1− σ)f(xk, v(xk)),

concluding the proof.

14

Page 15: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

Now we show that the NLCG algorithm generates a globally convergent sequence if βk isnon-negative and bounded above by an appropriate fraction of βCDk and the line search satisfiesthe strong Wolfe conditions. It is worth mentioning that even for the scalar case the conjugategradient method with βk = βCDk and strong Wolfe line search may not converge to a stationarypoint, see [14].

Theorem 3. Assume that Assumptions (A1) and (A2) hold. Consider a NLCG algorithm with

βk = ηβCDk , where 0 ≤ η < 1− σ.

If αk satisfies the strong Wolfe conditions (6), then lim infk→∞

‖v(xk)‖ = 0.

Proof. It follows from Lemma 6 that dk satisfies the sufficient descent condition (13) withc = 1− σ for all k ≥ 0. Therefore,

0 ≤ βk =η

1− σ(1− σ)

f(xk, v(xk))

f(xk−1, dk−1)≤ η

1− σf(xk, v(xk))

f(xk−1, v(xk−1))=

η

1− σβFRk .

Since 0 ≤ η/(1− σ) < 1 the proof follows from Theorem 2.

5.3 Dai–Yuan

Consider the DY vector parameter

βDYk =−f(xk, v(xk))

f(xk, dk−1)− f(xk−1, dk−1).

Following the methodology of CD, let us prove that the NLCG algorithm with 0 ≤ βk ≤ βDYkand a line search satisfying the strong Wolfe conditions also generates descent directions.

Lemma 7. Consider a NLCG algorithm with 0 ≤ βk ≤ βDYk and suppose that αk satisfiesthe strong Wolfe conditions (6). Then dk satisfies the sufficient descent condition (13) withc = 1/(1 + σ).

Proof. We proceed by induction. For k = 0, since d0 = v(x0), f(x0, v(x0)) < 0, and 0 < σ < 1,we see that inequality (13) with c = 1/(1 + σ) is satisfied. For some k ≥ 1, assume that

f(xk−1, dk−1) ≤ 1

1 + σf(xk−1, v(xk−1)) < 0.

From the Wolfe condition (5b), we obtain

f(xk, dk−1) ≥ σf(xk−1, dk−1) > f(xk−1, dk−1),

because σ < 1 and f(xk−1, dk−1) < 0. Hence, βDYk > 0 and βk is well defined. By the definitionof dk and positiveness of βk we obtain

f(xk, dk) ≤ f(xk, v(xk)) + βkf(xk, dk−1).

If f(xk, dk−1) ≤ 0 the result holds trivially. Now suppose that f(xk, dk−1) > 0. Then,

f(xk, dk) ≤ f(xk, v(xk)) + βkf(xk, dk−1) ≤ f(xk, v(xk)) + βDYk f(xk, dk−1).

Define lk = f(xk, dk−1)/f(xk−1, dk−1). By the the strong Wolfe condition (6b) we have lk ∈[−σ, σ]. Using the above inequality, direct calculations show that

f(xk, dk) ≤ 1

1− lkf(xk, v(xk)) ≤ 1

1 + σf(xk, v(xk)),

which concludes the proof.

15

Page 16: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

We observe that, if αk satisfies the standard Wolfe conditions (5), it is possible to showthat DY method generates simple descent directions. In the following, we state the convergenceresult related to the DY method.

Theorem 4. Assume that Assumptions (A1) and (A2) hold. Consider a NLCG algorithm with

βk = ηβDYk , where 0 ≤ η < 1− σ1 + σ

.

If αk satisfies the strong Wolfe conditions (6), then lim infk→∞

‖v(xk)‖ = 0.

Proof. It follows from Lemma 7 that dk satisfies the sufficient descent condition (13) withc = 1/(1 + σ) for all k ≥ 0. Then, by the Wolfe condition (6b), we have

f(xk, dk−1)− f(xk−1, dk−1) ≥ (σ − 1)f(xk−1, dk−1) ≥ σ − 1

1 + σf(xk−1, v(xk−1)) > 0.

Then,−f(xk−1, v(xk−1))

f(xk, dk−1)− f(xk−1, dk−1)≤ 1 + σ

1− σ.

Define δ = η(1 + σ)/(1− σ). Thus, by the definition of βk, we obtain

βk = δ1− σ1 + σ

[−f(xk, v(xk))

]f(xk, dk−1)− f(xk−1, dk−1)

≤ δ1− σ1 + σ

[−f(xk, v(xk))

−f(xk−1, v(xk−1))

][−f(xk−1, v(xk−1))

f(xk, dk−1)− f(xk−1, dk−1)

]

≤ δ f(xk, v(xk))

f(xk−1, v(xk−1))= δβFRk .

Since 0 ≤ δ < 1 the proof follows from Theorem 2.

Theorem 4 state the global convergence of a NLCG algorithm with an appropriate fractionof DY parameter. We conclude this section by noting that it is possible to obtain globalconvergence with a slight modification in the DY parameter. Let us define the modified DYparameter by

βmDYk =−f(xk, v(xk))

f(xk, dk−1)− τf(xk−1, dk−1),

where τ > 1. Observe that, if the line searches are exact, we have βmDYk ≤ 1

τβFRk and the

convergence can be obtained by using Theorem 2.

Theorem 5. Assume that Assumptions (A1) and (A2) hold. Consider a NLCG algorithm with

βk = βmDYk .

If αk satisfies the strong Wolfe conditions (6), then lim infk→∞

‖v(xk)‖ = 0.

Proof. Similarly to Lemma 7, we may show that βmDYk > 0, and that dk satisfies the sufficientdescent condition (13) with c = τ/(τ + σ), i.e.,

f(xk, dk) ≤ τ

τ + σf(xk, v(xk)), (24)

for all k ≥ 0. Note that

f(xk, dk) ≤ f(xk, v(xk)) + βmDYk f(xk, dk−1) = τβmDYk f(xk−1, dk−1),

16

Page 17: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

which implies that

βmDYk ≤ f(xk, dk)

τf(xk−1, dk−1). (25)

Assume by contradiction that there exists γ > 0 such that

‖v(xk)‖ ≥ γ for all k ≥ 0.

By (12), and Lemma 4-(d) with a = ‖v(xk)‖, b = ‖dk−1‖ and α = 1/√

2(τ2 − 1), we have

‖dk‖2 ≤[‖v(xk)‖+ βmDYk ‖dk−1‖

]2≤ τ2

τ2 − 1‖v(xk)‖2 + τ2(βmDYk )2‖dk−1‖2.

Thus, from (24) and (25) and observing that f(xk, v(xk)) < 0, we obtain

‖dk‖2

f2(xk, dk)≤ τ2

τ2 − 1

‖v(xk)‖2

f2(xk, dk)+

‖dk−1‖2

f2(xk−1, dk−1)

≤ (τ + σ)2

τ2 − 1

‖v(xk)‖2

f2(xk, v(xk))+

‖dk−1‖2

f2(xk−1, dk−1).

Then, remembering that 0 < γ2 ≤ ‖v(xk)‖2 ≤ −2f(xk, v(xk)), we obtain

‖dk‖2

f2(xk, dk)≤ 4(τ + σ)2

(τ2 − 1)γ2+

‖dk−1‖2

f2(xk−1, dk−1).

Applying above relation repeatedly, it follows that

‖dk‖2

f2(xk, dk)≤ c1k + c2,

where c1 = 4(τ + σ)2/[(τ2 − 1)γ2] > 0 and c2 = ‖v(x0)‖2/f2(x0, v(x0)) > 0. Then,

∑k≥0

f2(xk, dk)

‖dk‖2≥∑k≥0

1

c1k + c2=∞,

contradicting the Zoutendijk condition (10).

5.4 Polak–Ribiere–Polyak and Hestenes–Stiefel

In this section, we analyze the convergence of the NLCG algorithm related to the PRP and HSparameters given by

βPRPk =−f(xk, v(xk)) + f(xk−1, v(xk))

−f(xk−1, v(xk−1))

and

βHSk =−f(xk, v(xk)) + f(xk−1, v(xk))

f(xk, dk−1)− f(xk−1, dk−1).

respectively. In the scalar minimization case, Powell showed in [43] that PRP and HS meth-ods with exact line searches can cycle without approaching a solution point. Gilbert andNocedal [25] proved that global convergence can be obtained for βk = max{βPRPk , 0} and forβk = max{βHSk , 0}. Dai et al. in [13] showed that the positiveness restriction on βk cannot berelaxed for the PRP method. We can retrieve these results for the vector optimization context.

Our results in this section are based on the work [25] of Gilbert and Nocedal. They studiedthe convergence of PRP and HS methods, for the scalar minimization case, introducing theso-called Property (∗). The vector extension of this property is as follows.

17

Page 18: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

Property (∗). Consider a NLCG algorithm, and suppose that

0 < γ ≤ ‖v(xk)‖ ≤ γ, (26)

for all k ≥ 0. Under this assumption, we say that the method has Property (∗) if there existconstants b > 1 and λ > 0 such that for all k:

|βk| ≤ b,

and

‖sk−1‖ ≤ λ⇒ |βk| ≤1

2b,

where sk−1 = xk − xk−1.

In [25], the authors established global convergence results assuming that the sufficient de-scent condition holds. On the other hand, Dai et al. showed in [13] that the convergence canbe obtained by assuming simple descent. In the present work, we retrieve the results of [13].

We begin by presenting two lemmas. In the first we show that if there exists γ > 0 suchthat ‖v(xk)‖ ≥ γ for all k ≥ 0, then the search directions change slowly, asymptotically. In thesecond lemma, we demonstrate that, in this case, if the method has Property (∗), then a certainfraction of the stepsize cannot be too small.

Lemma 8. Suppose that Assumptions (A1) and (A3) hold. Consider a NLCG algorithm whereβk ≥ 0, dk is a K-descent direction of F at xk, and αk satisfies the strong Wolfe conditions (6).In addition, suppose that there exists γ > 0 such that ‖v(xk)‖ ≥ γ for all k ≥ 0. Then:

(i)∑k≥0

‖v(xk)‖4

‖dk‖2<∞;

(ii)∑k≥1

‖uk − uk−1‖2 <∞, where uk = dk/‖dk‖.

Proof. Since dk is aK-descent direction of F at xk, it turns out that dk 6= 0. Hence, ‖v(xk)‖4/‖dk‖2and uk are well defined. The proof of part (i) can be obtained from the proof of Theorem 1-(ii).

Now consider part (ii). Define rk = v(xk)/‖dk‖ and δk = βk‖dk−1‖/‖dk‖. Observe thatuk = rk + δku

k−1. Proceeding as in [25, Lemma 4.1] we obtain ‖uk−uk−1‖ ≤ 2‖rk‖. Therefore,using part (i), we have

γ2∑k≥1

‖uk − uk−1‖2 ≤ 4γ2∑k≥1

‖rk‖2 ≤ 4∑k≥1

‖rk‖2‖v(xk)‖2 = 4∑k≥1

‖v(xk)‖4

‖dk‖2<∞,

concluding the proof.

For λ > 0 and positive integer ∆, define

Kλk,∆ = {i ∈ N | k ≤ i ≤ k + ∆− 1, ‖sk−1‖ > λ}

and denote by |Kλk,∆| the number of elements of Kλk,∆.

Lemma 9. Suppose that Assumptions (A1) and (A3) hold. Consider a NLCG algorithm wheredk is a K-descent direction of F at xk, αk satisfies the strong Wolfe conditions (6), and assumethat the method has Property (∗). If there exists γ > 0 such that ‖v(xk)‖ ≥ γ, for all k ≥ 0,then there exists λ > 0 such that, for any ∆ ∈ N and any index k0, there is a greater indexk ≥ k0 such that

|Kλk,∆| >∆

2.

18

Page 19: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

Proof. Using Lemma 4-(c), we have

‖dl‖2 ≤ (‖v(xl)‖+ |βl|‖dl−1‖)2 ≤ 2‖v(xl)‖2 + 2β2l ‖dl−1‖2 ≤ 2γ2 + 2β2

l ‖dl−1‖2,

for all l ∈ N. Then, we can proceed by contradiction as in [25, Lemma 4.2] to show that thereexists k0 such that

‖dl‖2 ≤ c(l − k0 + 2) (27)

for any index l ≥ k0 + 1, where c is a certain positive constant independent of l. On the otherhand, from Lemma 8-(i), we have

γ4∑k≥0

1

‖dk‖2≤∑k≥0

‖v(xk)‖4

‖dk‖2<∞,

contradicting (27) and concluding the proof.

Theorem 6. Suppose that Assumptions (A1) and (A3) hold. Consider a NLCG algorithm whereβk ≥ 0, dk is a K-descent direction of F at xk, αk satisfies the strong Wolfe conditions (6),and assume that the method has Property (∗). Then lim inf

k→∞‖v(xk)‖ = 0.

Proof. Noting that {xk} belongs to the bounded set L, and using Lemmas 8 and 9, the proofis by contradiction exactly as in [25, Theorem 4.3].

Now we can prove the convergence result related to the PRP and HS parameters.

Theorem 7. Suppose that Assumptions (A1) and (A3) hold. Consider a NLCG algorithm with

βk = max{βPRPk , 0} or βk = max{βHSk , 0}.

If αk satisfies the strong Wolfe conditions (6), and dk is a K-descent direction of F at xk, thenlim infk→∞

‖v(xk)‖ = 0.

Proof. First, observe that if a NLCG algorithm with βk has Property (∗), so does a NLCGalgorithm with max{βk, 0}. Therefore, by Theorem 6, it is sufficient to prove that PRP and HSmethods have Property (∗). Assume that (26) holds for all k ≥ 0. Then

f(xk, v(xk)) < −1

2‖v(xk)‖2 ≤ −1

2γ2.

Since dk satisfies the sufficient descent condition (13), the relation (20) holds. Hence,

1

2γ2 ≤ −f(xk, v(xk)) ≤ cγ, (28)

where c is such that ‖JF (xk)‖ ≤ c for all k ≥ 0. Observe that there exists w ∈ C such that∣∣∣f(xk−1, v(xk))∣∣∣ =∣∣∣〈JF (xk−1)v(xk), w〉

∣∣∣ ≤ ‖JF (xk−1)‖‖v(xk)‖ ≤ cγ, (29)

because ‖w‖ = 1. Moreover, from Lemma 3 together with Assumption (A1), we have∣∣∣−f(xk, v(xk)) + f(xk−1, v(xk))∣∣∣ ≤ L‖v(xk)‖‖xk − xk−1‖ ≤ Lλγ, (30)

when ‖sk−1‖ ≤ λ.For the PRP method, define b = 4cγ/γ2 and λ = γ2/(4Lγb). By (28) and (29) we have

∣∣∣βPRPk

∣∣∣ ≤ −f(xk, v(xk)) +∣∣∣f(xk−1, v(xk))

∣∣∣−f(xk−1, vk−1)

≤ 4cγ

γ2= b,

19

Page 20: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

and when ‖sk−1‖ ≤ λ, it follows from (28) and (30) that∣∣∣βPRPk

∣∣∣ ≤ 2Lλγ

γ2=

1

2b.

Therefore, the PRP method has Property (∗).Now consider the HS method. By the sufficient descent condition (13) and the Wolfe con-

dition (6b) we obtain

f(xk, dk−1)− f(xk−1, dk−1) ≥ σf(xk−1, dk−1)− f(xk−1, dk−1)= −(1− σ)f(xk−1, dk−1)≥ −c(1− σ)f(xk−1, vk−1)

≥ c(1− σ)γ2

2> 0.

(31)

Define b = 4cγ/[c(1− σ)γ2] and λ = c(1− σ)γ2/(4Lγb). By (28), (29) and (31),

∣∣∣βHSk ∣∣∣ ≤ −f(xk, v(xk)) +∣∣∣f(xk−1, v(xk))

∣∣∣f(xk, dk−1)− f(xk−1, dk−1)

≤ 4cγ

c(1− σ)γ2= b.

Now, if ‖sk−1‖ ≤ λ, (30) and (31) imply∣∣∣βHSk ∣∣∣ ≤ 2Lλγ

c(1− σ)γ2=

1

2b,

concluding that the HS method has Property (∗).

6 Final Remarks

In this work, we have proposed non-linear conjugate gradient methods for solving unconstrainedvector optimization problems. Throughout the paper all the assumptions are natural extensionsof those made for the scalar minimization case. The vector extensions of the Fletcher–Reeves,Conjugate Descent, Dai–Yuan, Polak–Ribiere–Polyak, and Hestenes–Stiefel parameters wereconsidered. In particular, we showed that it is possible to obtain global convergence of the NLCGalgorithm with any fraction of the FR parameter. However, whether the FR method, withinexact line searches, automatically generates descent directions remains as an open question.We also showed that if parameter βk is non-negative and bounded above by an appropriatefraction of the CD parameter, the global convergence can be achieved. With respect to thePRP and HS parameters, all convergence results present in [25] for the classical optimizationwere retrieved. In addition, we proposed a slight modification in DY parameter for which it waspossible to show global convergence. The convergence analyses were made assuming inexactline searches and without regular restarts.

Regarding the line search procedure, we have introduced the standard and strong Wolfeconditions for vector optimization, and showed the existence of intervals of stepsizess satisfyingthem. Moreover, Proposition 1 sheds light on algorithmic properties and suggests the imple-mentation of a Wolfe-type line search procedure. We also introduced the Zoutendjik conditionfor vector optimization and proved that it is satisfied for a general descent line search method.As far as we know, this is the first paper in which this type of result has been presented in thevector optimization context. We expect that the proposed extensions of Wolfe and Zoutendjikconditions can be used in the convergence analysis of other methods.

It is worth mentioning that the positiveness of parameter βk seems to be essential for ob-taining methods that generate descent directions. Lemma 5 and the technique used in its proofcorroborate this intuition. In [30], Hager and Zhang proposed an efficient conjugate gradientmethod for scalar minimization for which the parameter βk can be negative. A challengingproblem is to extend their work for the vector-valued optimization.

20

Page 21: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

References

[1] M. Al-Baali, Descent Property and Global Convergence of the Fletcher-Reeves Methodwith Inexact Line Search, IMA Journal of Numerical Analysis, 5 (1985), pp. 121–124.

[2] M. A. Ansary and G. Panda, A modified Quasi-Newton method for vector optimizationproblem, Optimization, 64 (2015), pp. 2289–2306.

[3] J. Y. Bello Cruz, A Subgradient Method for Vector Optimization Problems, SIAM Jour-nal on Optimization, 23 (2013), pp. 2169–2182.

[4] H. Bonnel, A. N. Iusem, and B. F. Svaiter, Proximal methods in vector optimization,SIAM Journal on Optimization, 15 (2005), pp. 953–970.

[5] L. C. Ceng, B. S. Mordukhovich, and J. C. Yao, Hybrid approximate proximalmethod with auxiliary variational inequality for vector optimization, Journal of optimizationtheory and applications, 146 (2010), pp. 267–303.

[6] L. C. Ceng and J. C. Yao, Approximate proximal methods in vector optimization, Eu-ropean Journal of Operational Research, 183 (2007), pp. 1–19.

[7] T. D. Chuong, Generalized proximal method for efficient solutions in vector optimization,Numerical functional analysis and optimization, 32 (2011), pp. 843–857.

[8] , Newton-like methods for efficient solutions in vector optimization, ComputationalOptimization and Applications, 54 (2013), pp. 495–516.

[9] T. D. Chuong, B. S. Mordukhovich, and J. C. Yao, Hybrid approximate proximalalgorithms for efficient solutions in vector optimization, Journal of Nonlinear and ConvexAnalysis, 12 (2011), pp. 257–285.

[10] T. D. Chuong and J. C. Yao, Steepest descent methods for critical points in vectoroptimization problems, Applicable Analysis, 91 (2012), pp. 1811–1829.

[11] Y. H. Dai, Convergence analysis of nonlinear conjugate gradient methods, in Optimizationand Regularization for Computational Inverse Problems and Applications, Springer, 2010,pp. 157–181.

[12] , Nonlinear Conjugate Gradient Methods, John Wiley & Sons, Inc., 2010.

[13] Y. H. Dai, J. Han, G. Liu, D. Sun, H. Yin, and Y. X. Yuan, Convergence Propertiesof Nonlinear Conjugate Gradient Methods, SIAM Journal on Optimization, 10 (2000),pp. 345–358.

[14] Y. H. Dai and Y. .Yuan, Convergence properties of the conjugate descent method, Ad-vances in Mathematics, 25 (1996), pp. 552–562.

[15] Y. H. Dai and Y. Yuan, A Nonlinear Conjugate Gradient Method with a Strong GlobalConvergence Property, SIAM Journal on Optimization, 10 (1999), pp. 177–182.

[16] P. De, J. B. Ghosh, and C. E. Wells, On the minimization of completion time variancewith a bicriteria extension, Operations Research, 40 (1992), pp. 1148–1155.

[17] R. Fletcher, Practical Method of Optimization, Unconstrained Optimization, vol. 1, 1980.

[18] R. Fletcher and C. M. Reeves, Function minimization by conjugate gradients, TheComputer Journal, 7 (1964), pp. 149–154.

21

Page 22: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

[19] J. Fliege, L. M. Grana Drummond, and B. F. Svaiter, Newton’s Method for Mul-tiobjective Optimization, SIAM Journal on Optimization, 20 (2009), pp. 602–626.

[20] J. Fliege and B. F. Svaiter, Steepest descent methods for multicriteria optimization,Mathematical Methods of Operations Research, 51 (2000), pp. 479–494.

[21] J. Fliege and L. N. Vicente, Multicriteria approach to bilevel optimization, Journal ofoptimization theory and applications, 131 (2006), pp. 209–225.

[22] J. Fliege and R. Werner, Robust multiobjective optimization & applications in portfoliooptimization, European Journal of Operational Research, 234 (2014), pp. 422 – 433.

[23] E. H. Fukuda and L. M. Grana Drummond, On the convergence of the projectedgradient method for vector optimization, Optimization, 60 (2011), pp. 1009–1021.

[24] E. H. Fukuda and L. M. Grana Drummond, Inexact projected gradient method forvector optimization, Computational Optimization and Applications, 54 (2013), pp. 473–493.

[25] J. C. Gilbert and J. Nocedal, Global Convergence Properties of Conjugate GradientMethods for Optimization, SIAM Journal on Optimization, 2 (1992), pp. 21–42.

[26] L. M. Grana Drummond and A. N. Iusem, A Projected Gradient Method for VectorOptimization Problems, Computational Optimization and Applications, 28 (2004), pp. 5–29.

[27] L. M. Grana Drummond, F. M. P. Raupp, and B. F. Svaiter, A quadraticallyconvergent Newton method for vector optimization, Optimization, 63 (2014), pp. 661–677.

[28] L. M. Grana Drummond and B. F. Svaiter, A steepest descent method for vectoroptimization, Journal of Computational and Applied Mathematics, 175 (2005), pp. 395 –414.

[29] M. Gravel, J. M. Martel, R. Nadeau, W. Price, and R. Tremblay, A multi-criterion view of optimal resource allocation in job-shop production, European journal ofoperational research, 61 (1992), pp. 230–244.

[30] W. W. Hager and H. Zhang, A new conjugate gradient method with guaranteed descentand an efficient line search, SIAM Journal on Optimization, 16 (2005), pp. 170–192.

[31] W. W. Hager and H. Zhang, A survey of nonlinear conjugate gradient methods, Pacificjournal of Optimization, 2 (2006), pp. 35–58.

[32] M. R. Hestenes and E. Stiefel, Methods of conjugate gradients for solving linearsystems, vol. 49, Journal of Research of the National Bureau of Standards, 1952.

[33] A. Huttere and J. J., Optimization of the location of antennas for treatment planningin hyperthermia, 2000.

[34] J. Jahn, A. Kirsch, and C. Wagner, Optimization of rod antennas of mobile phones,Mathematical Methods of Operations Research, 59 (2004), pp. 37–51.

[35] A. Juschke, J. Jahn, and A. Kirsch, A bicriterial optimization problem of antennadesign, Computational Optimization and Applications, 7 (1997), pp. 261–276.

[36] T. M. Leschine, H. Wallenius, and W. A. Verdini, Interactive multiobjective analysisand assimilative capacity-based ocean disposal decisions, European Journal of OperationalResearch, 56 (1992), pp. 278–289.

22

Page 23: Non-linear conjugate gradient methods for vector optimization · subgradient, interior point, and proximal methods were proposed for vector optimization, see [2,3,8,19,20,23,24,26{28,37,44,47].

[37] F. Lu and C.-R. Chen, Newton-like methods for solving vector optimization problems,Applicable Analysis, 93 (2014), pp. 1567–1586.

[38] D. T. Luc, Theory of vector optimization, Lectures Notes in economics and mathematicalsystems, vol. 319, 1989.

[39] J. J. More and D. J. Thuente, Line Search Algorithms with Guaranteed SufficientDecrease, ACM Trans. Math. Softw., 20 (1994), pp. 286–307.

[40] J. Nocedal and S. Wright, Numerical optimization, Springer Science & Business Media,2006.

[41] E. Polak and G. Ribiere, Note sur la convergence de methodes de directions conjuguees,Revue francaise d’informatique et de recherche operationnelle, serie rouge, 3 (1969), pp. 35–43.

[42] B. T. Polyak, The conjugate gradient method in extremal problems, USSR ComputationalMathematics and Mathematical Physics, 9 (1969), pp. 94 – 112.

[43] M. J. D. Powell, Nonconvex minimization calculations and the conjugate gradientmethod, in Numerical analysis, Springer, 1984, pp. 122–141.

[44] S. Qu, M. Goh, and F. T. Chan, Quasi-Newton methods for solving multiobjectiveoptimization, Operations Research Letters, 39 (2011), pp. 397–399.

[45] M. Tavana, A subjective assessment of alternative mission architectures for the humanexploration of Mars at NASA using multicriteria decision making, Computers & OperationsResearch, 31 (2004), pp. 1147–1164.

[46] M. Tavana, M. A. Sodenkamp, and L. Suhl, A soft multi-criteria decision analysismodel with application to the European Union enlargement, Annals of Operations Research,181 (2010), pp. 393–421.

[47] K. D. Villacorta and P. R. Oliveira, An interior proximal method in vector opti-mization, European Journal of Operational Research, 214 (2011), pp. 485–492.

[48] D. White, Epsilon-dominating solutions in mean-variance portfolio analysis, EuropeanJournal of Operational Research, 105 (1998), pp. 457–466.

23