The Ekeland Variational Principle, the Bishop-Phelps Theorem, and The

Total Page:16

File Type:pdf, Size:1020Kb

The Ekeland Variational Principle, the Bishop-Phelps Theorem, and The The Ekeland Variational Principle, the Bishop-Phelps Theorem, and the Brøndsted-Rockafellar Theorem Our aim is to prove the Ekeland Variational Principle which is an abstract result that found numerous applications in various fields of Mathematics. As its application to Convex Analysis, we provide a proof of the famous Bishop- Phelps Theorem and some related results. Let us recall that the epigraph and the hypograph of a function f : X ! [−∞; +1] (where X is a set) are the following subsets of X × R: epi f = f(x; t) 2 X × R : t ≥ f(x)g ; hyp f = f(x; t) 2 X × R : t ≤ f(x)g : Notice that, if f > −∞ on X then epi f 6= ; if and only if f is proper, that is, f is finite at at least one point. 0.1. The Ekeland Variational Principle. In what follows, (X; d) is a metric space. Given λ > 0 and (x; t) 2 X × R, we define the set Kλ(x; t) = f(y; s): s ≤ t − λd(x; y)g = f(y; s): λd(x; y) ≤ t − sg ⊂ X × R: Notice that Kλ(x; t) = hyp[t − λ(x; ·)]. Let us state some useful properties of these sets. Lemma 0.1. (a) Kλ(x; t) is closed and contains (x; t). (b) If (¯x; t¯) 2 Kλ(x; t) then Kλ(¯x; t¯) ⊂ Kλ(x; t). (c) If (xn; tn) ! (x; t) and (xn+1; tn+1) 2 Kλ(xn; tn) (n 2 N), then \ Kλ(x; t) = Kλ(xn; tn) : n2N Proof. (a) is quite easy. To show (b), let (¯x; t¯) 2 Kλ(x; t) and (y; s) 2 Kλ(¯x; t¯). Then λd(x; y) ≤ λd(x; x¯) + λd(¯x; y) ≤ (t − t¯) + (t¯− s) = t − s, and hence (y; s) 2 Kλ(x; t). To prove (c), we shall use (a) and (b). Notice that Kλ(xn+1; tn+1) ⊂ Kλ(xn; tn) for each n. Since (xm; tm) 2 Kλ(xn; tn) whenever m ≥ n, we can pass to the limit for m ! +1 to obtain (x; t) 2 Kλ(xn; tn) for each n. Thus T Kλ(x; t) ⊂ n Kλ(xn; tn). To show the equality, take (y; s) 2= Kλ(x; t), that is, λd(x; y) − t + s > 0. But then λd(xn; y) − tn + s > 0 for each sufficiently large T n. It follows that (y; s) 2= n Kλ(xn; tn). The proof of the following proposition, basic for Ekeland's principle, is some- how \geometric" and its basic idea is quite simple. The reader is invited to draw a picture to grasp this geometric idea of the proof. 1 2 Proposition 0.2. Let (X; d) be a complete metric space, and let PR denote the canonical projection of X × R onto R, that is, PR(x; t) = t. Let F ⊂ X × R be a closed set such that inf PR(F ) > −∞. Then for every λ > 0 and (x0; t0) 2 F there exists (¯x; t¯) 2 F \ Kλ(x0; t0) such that F \ Kλ(¯x; t¯) = f(¯x; t¯)g: P1 Proof. Fix a sequence f"ngn≥1 ⊂ (0; +1) such that 1 "n < +1. Let us inductively define sets Fn and points (xn; tn) in X × R for n ≥ 0. Case n = 0. Define F0 = F , and recall that we already have (x0; t0) 2 F0. Now let n ≥ 0, and assume that we have already defined Fk and (xk; tk) 2 Fk for 0 ≤ k ≤ n. Put Fn+1 = Fn \Kλ(xn; tn), and fix a point (xn+1; tn+1) 2 Fn+1 such that tn+1 ≤ inf PR(Fn+1) + "n+1. Notice that Fn+1 ⊂ Fn and Fn+1 = F \ Kλ(xn; tn)(n ≥ 0) by Lemma 0.1. Moreover, since (xn+1; tn+1) 2 Fn \ Kλ(xn; tn), we have that λd(xn; xn+1) ≤ tn − tn+1 ≤ inf PR(Fn) + "n − inf PR(Fn) = "n (n ≥ 1). Our choice of f"ng easily implies that the sequences fxng and ftng are Cauchy and hence ¯ T T convergent: (xn; tn) ! (¯x; t) 2 F1. Now, n Fn = n[F \ Kλ(xn; tn)] = T ¯ ¯ F \ n Kλ(xn; tn) = F \ Kλ(¯x; t) by Lemma 0.1. To show that F \ Kλ(¯x; t) is a singleton, it suffices to show that diam Fn ! 0. (On X × R we can consider, e.g., the metric d^(x; t); (y; s) := maxd(x; y); jt − sj , or d^(x; t); (y; s) := d(x; y) + jt − sj which is clearly equivalent to the previous one.) For n ≥ 2 and (y; s) 2 Fn = Fn−1 \ Kλ(xn−1; tn−1), we have as above: λd(xn−1; y) ≤ tn−1 − s ≤ inf PR(Fn−1) + "n−1 − inf PR(Fn−1) = "n−1 ! 0. It follows that diam(Fn) ! 0, and we are done. We are ready for the famous \variational principle" due to I. Ekeland [J. Math. Anal. Appl. 47 (1974), 324{353]. It asserts that, under apropriate assumptions, if a point x0 is an \almost minimum point" for a function f then there exists a small Lipschitz perturbation of f attaining its (strict global) minimum at a point \near" to x0. As usual, \l.s.c." stands for \lower semicontinuous". Theorem 0.3 (Ekeland Variational Principle). Let (X; d) be a complete metric space, and f : X ! (−∞; +1] a proper l.s.c. function which is bounded below. Let " > 0, λ > 0, x0 2 X, and f(x0) ≤ inf f(X) + ". Then there exists x¯ 2 X such that (a) f(¯x) ≤ f(x0); (b) d(x0; x¯) ≤ ε/λ; (c) f(¯x) < f(x) + λd(¯x; x) for all x 2 X n fx¯g (that is, the function f + λ(¯x; ·) attains its strict global minimum at x¯.) Proof. The set F := epi f is a nonempty closed set in X × R,(x0; f(x0)) 2 F , f(x0) ≤ inf PR(F ) + " and the infimum is finite. By Proposition 0.2, there is (¯x; t¯) 2 F \ Kλ(x0; f(x0)) such that F \ Kλ(¯x; t¯) = f(¯x; t¯)g. Since (¯x; f(¯x)) 2 F \ Kλ(¯x; t¯), we necessarily have t¯ = f(¯x). Moreover, since (¯x; f(¯x)) 2 3 Kλ(x0; t0), we have λd(x0; x¯) ≤ f(x0) − f(¯x) ≤ inf f(X) + " − inf f(X) ≤ " which implies (a) and (b). Finally, if x 6=x ¯ then (x; f(x)) 2= Kλ(¯x; f(¯x)), that is, λd(¯x; x) > f(¯x) − f(x) which gives (c). p Remark 0.4. Notice that for λ = " , Theorem 0.3(b,c) become p p d(x0; x¯) ≤ " and f(¯x) < f(x) + " d(¯x; x) whenever x 6=x ¯. 0.2. The Bishop-Phelps Theorem. The classical Bishop-Phelps Theorem asserts that if X is a (real) Banach space then the elements of X∗ that attain the norm are dense in X∗. We shall prove a more general version of this theorem, where the unit ball of X is replaced by an arbitrary closed, convex, proper subset of X. We shall use the following standard terminology. If C is a nonempty set in a normed space X, x 2 C, x∗ 2 X∗ n f0g and x∗(x) = sup x∗(C), then we say that x is a support point of C and x∗ is a support functional for C. Notice that each support point belongs to the boundary @C of C, and all strictly positive multiples of a support functional are again support functionals. Moreover, the Hahn-Banach Theorem immediately gives that if C is a closed convex set with nonempty interior (in some topological vector space) then each point of @C is a support point of C. The main tool for the Bishop-Phelps Theorem will be the following auxiliary theorem. In the proof we shall use the easy and well-known fact that a closed hyperplane in X ×R that strictly separates two distinct points having the same first component coincides with the graph of a continuous affine function on X. Theorem 0.5. Let X be a Banach space, C ⊂ X a nonempty, closed, convex ∗ ∗ set. Let " > 0, x0 2 C and x0 2 X n f0g be such that ∗ ∗ ∗ inf x0(C) > −∞ and x0(x0) ≤ inf x0(C) + ": ∗ ∗ ∗ Then for each 0 < λ < kx0k there exist x1 2 C and x1 2 X n f0g such that " kx − x k ≤ ; kx∗ − x∗k ≤ λ and x∗(x ) = inf x∗(C) : 1 0 λ 1 0 1 1 1 Proof. The function ( x∗(x) if x 2 C, f(x) = 0 +1 otherwise, ∗ is a proper, convex, l.s.c. function on X (since epi f = (epi x0) \ C × R), and it satisfies f(x0) ≤ inf f(X) + ". By the Ekeland Variational Principle, there exists x1 2 X such that f(x1) ≤ f(x0), kx1 − x0k ≤ ε/λ, and f(x1) ≤ f(x) + λkx − x1k for all x 2 X. The first inequality gives that x1 2 C, while the last one can be rewritten in the form g(x) := f(x) − f(x1) ≥ −λkx − x1k =: h(x); x 2 X: 4 Since g is proper, convex and l.s.c., h is finite, concave and continuous, and g(x1) = h(x1) = 0, we can separate the epigraph epi g and the hypograph hyp h (which has nonempty interior!) by a closed hyperplane in X × R that ∗ ∗ necessarily passes through the point (x1; 0). Thus there exists y 2 X such that ∗ f(x) − f(x1) ≥ y (x − x1) ≥ −λkx − x1k; x 2 X: ∗ ∗ ∗ ∗ The first inequality can be written as x0(x) − x0(x1) ≥ y (x) − y (x1), x 2 C, that is, ∗ ∗ ∗ ∗ (x0 − y )(x) ≥ (x0 − y )(x1); x 2 C; while the second inequality gives (by putting x = x1 + v) y∗(v) ≤ λkvk; v 2 X; ∗ ∗ ∗ ∗ that is, ky k ≤ λ.
Recommended publications
  • On Quasi Norm Attaining Operators Between Banach Spaces
    ON QUASI NORM ATTAINING OPERATORS BETWEEN BANACH SPACES GEUNSU CHOI, YUN SUNG CHOI, MINGU JUNG, AND MIGUEL MART´IN Abstract. We provide a characterization of the Radon-Nikod´ymproperty in terms of the denseness of bounded linear operators which attain their norm in a weak sense, which complement the one given by Bourgain and Huff in the 1970's. To this end, we introduce the following notion: an operator T : X ÝÑ Y between the Banach spaces X and Y is quasi norm attaining if there is a sequence pxnq of norm one elements in X such that pT xnq converges to some u P Y with }u}“}T }. Norm attaining operators in the usual (or strong) sense (i.e. operators for which there is a point in the unit ball where the norm of its image equals the norm of the operator) and also compact operators satisfy this definition. We prove that strong Radon-Nikod´ymoperators can be approximated by quasi norm attaining operators, a result which does not hold for norm attaining operators in the strong sense. This shows that this new notion of quasi norm attainment allows to characterize the Radon-Nikod´ymproperty in terms of denseness of quasi norm attaining operators for both domain and range spaces, completing thus a characterization by Bourgain and Huff in terms of norm attaining operators which is only valid for domain spaces and it is actually false for range spaces (due to a celebrated example by Gowers of 1990). A number of other related results are also included in the paper: we give some positive results on the denseness of norm attaining Lipschitz maps, norm attaining multilinear maps and norm attaining polynomials, characterize both finite dimensionality and reflexivity in terms of quasi norm attaining operators, discuss conditions to obtain that quasi norm attaining operators are actually norm attaining, study the relationship with the norm attainment of the adjoint operator and, finally, present some stability results.
    [Show full text]
  • POLARS and DUAL CONES 1. Convex Sets
    POLARS AND DUAL CONES VERA ROSHCHINA Abstract. The goal of this note is to remind the basic definitions of convex sets and their polars. For more details see the classic references [1, 2] and [3] for polytopes. 1. Convex sets and convex hulls A convex set has no indentations, i.e. it contains all line segments connecting points of this set. Definition 1 (Convex set). A set C ⊆ Rn is called convex if for any x; y 2 C the point λx+(1−λ)y also belongs to C for any λ 2 [0; 1]. The sets shown in Fig. 1 are convex: the first set has a smooth boundary, the second set is a Figure 1. Convex sets convex polygon. Singletons and lines are also convex sets, and linear subspaces are convex. Some nonconvex sets are shown in Fig. 2. Figure 2. Nonconvex sets A line segment connecting two points x; y 2 Rn is denoted by [x; y], so we have [x; y] = f(1 − α)x + αy; α 2 [0; 1]g: Likewise, we can consider open line segments (x; y) = f(1 − α)x + αy; α 2 (0; 1)g: and half open segments such as [x; y) and (x; y]. Clearly any line segment is a convex set. The notion of a line segment connecting two points can be generalised to an arbitrary finite set n n by means of a convex combination. Given a finite set fx1; x2; : : : ; xpg ⊂ R , a point x 2 R is the convex combination of x1; x2; : : : ; xp if it can be represented as p p X X x = αixi; αi ≥ 0 8i = 1; : : : ; p; αi = 1: i=1 i=1 Given an arbitrary set S ⊆ Rn, we can consider all possible finite convex combinations of points from this set.
    [Show full text]
  • On the Ekeland Variational Principle with Applications and Detours
    Lectures on The Ekeland Variational Principle with Applications and Detours By D. G. De Figueiredo Tata Institute of Fundamental Research, Bombay 1989 Author D. G. De Figueiredo Departmento de Mathematica Universidade de Brasilia 70.910 – Brasilia-DF BRAZIL c Tata Institute of Fundamental Research, 1989 ISBN 3-540- 51179-2-Springer-Verlag, Berlin, Heidelberg. New York. Tokyo ISBN 0-387- 51179-2-Springer-Verlag, New York. Heidelberg. Berlin. Tokyo No part of this book may be reproduced in any form by print, microfilm or any other means with- out written permission from the Tata Institute of Fundamental Research, Colaba, Bombay 400 005 Printed by INSDOC Regional Centre, Indian Institute of Science Campus, Bangalore 560012 and published by H. Goetze, Springer-Verlag, Heidelberg, West Germany PRINTED IN INDIA Preface Since its appearance in 1972 the variational principle of Ekeland has found many applications in different fields in Analysis. The best refer- ences for those are by Ekeland himself: his survey article [23] and his book with J.-P. Aubin [2]. Not all material presented here appears in those places. Some are scattered around and there lies my motivation in writing these notes. Since they are intended to students I included a lot of related material. Those are the detours. A chapter on Nemyt- skii mappings may sound strange. However I believe it is useful, since their properties so often used are seldom proved. We always say to the students: go and look in Krasnoselskii or Vainberg! I think some of the proofs presented here are more straightforward. There are two chapters on applications to PDE.
    [Show full text]
  • ABOUT STATIONARITY and REGULARITY in VARIATIONAL ANALYSIS 1. Introduction the Paper Investigates Extremality, Stationarity and R
    1 ABOUT STATIONARITY AND REGULARITY IN VARIATIONAL ANALYSIS 2 ALEXANDER Y. KRUGER To Boris Mordukhovich on his 60th birthday Abstract. Stationarity and regularity concepts for the three typical for variational analysis classes of objects { real-valued functions, collections of sets, and multifunctions { are investi- gated. An attempt is maid to present a classification scheme for such concepts and to show that properties introduced for objects from different classes can be treated in a similar way. Furthermore, in many cases the corresponding properties appear to be in a sense equivalent. The properties are defined in terms of certain constants which in the case of regularity proper- ties provide also some quantitative characterizations of these properties. The relations between different constants and properties are discussed. An important feature of the new variational techniques is that they can handle nonsmooth 3 functions, sets and multifunctions equally well Borwein and Zhu [8] 4 1. Introduction 5 The paper investigates extremality, stationarity and regularity properties of real-valued func- 6 tions, collections of sets, and multifunctions attempting at developing a unifying scheme for defining 7 and using such properties. 8 Under different names this type of properties have been explored for centuries. A classical 9 example of a stationarity condition is given by the Fermat theorem on local minima and max- 10 ima of differentiable functions. In a sense, any necessary optimality (extremality) conditions de- 11 fine/characterize certain stationarity (singularity/irregularity) properties. The separation theorem 12 also characterizes a kind of extremal (stationary) behavior of convex sets. 13 Surjectivity of a linear continuous mapping in the Banach open mapping theorem (and its 14 extension to nonlinear mappings known as Lyusternik-Graves theorem) is an example of a regularity 15 condition.
    [Show full text]
  • Monotonic Transformations: Cardinal Versus Ordinal Utility
    Natalia Lazzati Mathematics for Economics (Part I) Note 10: Quasiconcave and Pseudoconcave Functions Note 10 is based on Madden (1986, Ch. 13, 14) and Simon and Blume (1994, Ch. 21). Monotonic transformations: Cardinal Versus Ordinal Utility A utility function could be said to measure the level of satisfaction associated to each commodity bundle. Nowadays, no economist really believes that a real number can be assigned to each commodity bundle which expresses (in utils?) the consumer’slevel of satisfaction with this bundle. Economists believe that consumers have well-behaved preferences over bundles and that, given any two bundles, a consumer can indicate a preference of one over the other or the indi¤erence between the two. Although economists work with utility functions, they are concerned with the level sets of such functions, not with the number that the utility function assigns to any given level set. In consumer theory these level sets are called indi¤erence curves. A property of utility functions is called ordinal if it depends only on the shape and location of a consumer’sindi¤erence curves. It is alternatively called cardinal if it also depends on the actual amount of utility the utility function assigns to each indi¤erence set. In the modern approach, we say that two utility functions are equivalent if they have the same indi¤erence sets, although they may assign di¤erent numbers to each level set. For instance, let 2 2 u : R+ R where u (x) = x1x2 be a utility function, and let v : R+ R be the utility function ! ! v (x) = u (x) + 1: These two utility functions represent the same preferences and are therefore equivalent.
    [Show full text]
  • Chapter 5 Convex Optimization in Function Space 5.1 Foundations of Convex Analysis
    Chapter 5 Convex Optimization in Function Space 5.1 Foundations of Convex Analysis Let V be a vector space over lR and k ¢ k : V ! lR be a norm on V . We recall that (V; k ¢ k) is called a Banach space, if it is complete, i.e., if any Cauchy sequence fvkglN of elements vk 2 V; k 2 lN; converges to an element v 2 V (kvk ¡ vk ! 0 as k ! 1). Examples: Let ­ be a domain in lRd; d 2 lN. Then, the space C(­) of continuous functions on ­ is a Banach space with the norm kukC(­) := sup ju(x)j : x2­ The spaces Lp(­); 1 · p < 1; of (in the Lebesgue sense) p-integrable functions are Banach spaces with the norms Z ³ ´1=p p kukLp(­) := ju(x)j dx : ­ The space L1(­) of essentially bounded functions on ­ is a Banach space with the norm kukL1(­) := ess sup ju(x)j : x2­ The (topologically and algebraically) dual space V ¤ is the space of all bounded linear functionals ¹ : V ! lR. Given ¹ 2 V ¤, for ¹(v) we often write h¹; vi with h¢; ¢i denoting the dual product between V ¤ and V . We note that V ¤ is a Banach space equipped with the norm j h¹; vi j k¹k := sup : v2V nf0g kvk Examples: The dual of C(­) is the space M(­) of Radon measures ¹ with Z h¹; vi := v d¹ ; v 2 C(­) : ­ The dual of L1(­) is the space L1(­). The dual of Lp(­); 1 < p < 1; is the space Lq(­) with q being conjugate to p, i.e., 1=p + 1=q = 1.
    [Show full text]
  • Non-Linear Inner Structure of Topological Vector Spaces
    mathematics Article Non-Linear Inner Structure of Topological Vector Spaces Francisco Javier García-Pacheco 1,*,† , Soledad Moreno-Pulido 1,† , Enrique Naranjo-Guerra 1,† and Alberto Sánchez-Alzola 2,† 1 Department of Mathematics, College of Engineering, University of Cadiz, 11519 Puerto Real, CA, Spain; [email protected] (S.M.-P.); [email protected] (E.N.-G.) 2 Department of Statistics and Operation Research, College of Engineering, University of Cadiz, 11519 Puerto Real (CA), Spain; [email protected] * Correspondence: [email protected] † These authors contributed equally to this work. Abstract: Inner structure appeared in the literature of topological vector spaces as a tool to charac- terize the extremal structure of convex sets. For instance, in recent years, inner structure has been used to provide a solution to The Faceless Problem and to characterize the finest locally convex vector topology on a real vector space. This manuscript goes one step further by settling the bases for studying the inner structure of non-convex sets. In first place, we observe that the well behaviour of the extremal structure of convex sets with respect to the inner structure does not transport to non-convex sets in the following sense: it has been already proved that if a face of a convex set intersects the inner points, then the face is the whole convex set; however, in the non-convex setting, we find an example of a non-convex set with a proper extremal subset that intersects the inner points. On the opposite, we prove that if a extremal subset of a non-necessarily convex set intersects the affine internal points, then the extremal subset coincides with the whole set.
    [Show full text]
  • Convex Sets and Convex Functions 1 Convex Sets
    Convex Sets and Convex Functions 1 Convex Sets, In this section, we introduce one of the most important ideas in economic modelling, in the theory of optimization and, indeed in much of modern analysis and computatyional mathematics: that of a convex set. Almost every situation we will meet will depend on this geometric idea. As an independent idea, the notion of convexity appeared at the end of the 19th century, particularly in the works of Minkowski who is supposed to have said: \Everything that is convex interests me." We discuss other ideas which stem from the basic definition, and in particular, the notion of a convex and concave functions which which are so prevalent in economic models. The geometric definition, as we will see, makes sense in any vector space. Since, for the most of our work we deal only with Rn, the definitions will be stated in that context. The interested student may, however, reformulate the definitions either, in an ab stract setting or in some concrete vector space as, for example, C([0; 1]; R)1. Intuitively if we think of R2 or R3, a convex set of vectors is a set that contains all the points of any line segment joining two points of the set (see the next figure). Here is the definition. Definition 1.1 Let u; v 2 V . Then the set of all convex combinations of u and v is the set of points fwλ 2 V : wλ = (1 − λ)u + λv; 0 ≤ λ ≤ 1g: (1.1) In, say, R2 or R3, this set is exactly the line segment joining the two points u and v.
    [Show full text]
  • The Krein–Milman Theorem a Project in Functional Analysis
    The Krein{Milman Theorem A Project in Functional Analysis Samuel Pettersson November 29, 2016 2. Extreme points 3. The Krein{Milman theorem 4. An application Outline 1. An informal example 3. The Krein{Milman theorem 4. An application Outline 1. An informal example 2. Extreme points 4. An application Outline 1. An informal example 2. Extreme points 3. The Krein{Milman theorem Outline 1. An informal example 2. Extreme points 3. The Krein{Milman theorem 4. An application Outline 1. An informal example 2. Extreme points 3. The Krein{Milman theorem 4. An application Convex sets and their \corners" Observation Some convex sets are the convex hulls of their \corners". Convex sets and their \corners" Observation Some convex sets are the convex hulls of their \corners". kxk1≤ 1 Convex sets and their \corners" Observation Some convex sets are the convex hulls of their \corners". kxk1≤ 1 Convex sets and their \corners" Observation Some convex sets are the convex hulls of their \corners". kxk1≤ 1 kxk2≤ 1 Convex sets and their \corners" Observation Some convex sets are the convex hulls of their \corners". kxk1≤ 1 kxk2≤ 1 Convex sets and their \corners" Observation Some convex sets are the convex hulls of their \corners". kxk1≤ 1 kxk2≤ 1 kxk1≤ 1 Convex sets and their \corners" Observation Some convex sets are the convex hulls of their \corners". kxk1≤ 1 kxk2≤ 1 kxk1≤ 1 x1; x2 ≥ 0 Convex sets and their \corners" Observation Some convex sets are not the convex hulls of their \corners". Convex sets and their \corners" Observation Some convex sets are not the convex hulls of their \corners".
    [Show full text]
  • On the Variational Principle
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by Elsevier - Publisher Connector JOURNAL OF MATHEMATICAL ANALYSIS AND APPLICATIONS 47, 324-353 (1974) On the Variational Principle I. EKELAND UER Mathe’matiques de la DC&ion, Waiver& Paris IX, 75775 Paris 16, France Submitted by J. L. Lions The variational principle states that if a differentiable functional F attains its minimum at some point zi, then F’(C) = 0; it has proved a valuable tool for studying partial differential equations. This paper shows that if a differentiable function F has a finite lower bound (although it need not attain it), then, for every E > 0, there exists some point u( where 11F’(uJj* < l , i.e., its derivative can be made arbitrarily small. Applications are given to Plateau’s problem, to partial differential equations, to nonlinear eigenvalues, to geodesics on infi- nite-dimensional manifolds, and to control theory. 1. A GENERAL RFNJLT Let V be a complete metric space, the distance of two points u, z, E V being denoted by d(u, v). Let F: I’ -+ 08u {+ co} be a lower semicontinuous function, not identically + 00. In other words, + oo is allowed as a value for F, but at some point w,, , F(Q) is finite. Suppose now F is bounded from below: infF > --co. (l-1) As no compacity assumptions are made, there need be no point where this infimum is attained. But of course, for every E > 0, there is some point u E V such that: infF <F(u) < infF + E.
    [Show full text]
  • Convex Sets, Functions, and Problems
    Convex Sets, Functions, and Problems Nick Henderson, AJ Friend (Stanford University) Kevin Carlberg (Sandia National Laboratories) August 13, 2019 1 Convex optimization Theory, methods, and software for problems exihibiting the characteristics below I Convexity: I convex : local solutions are global I non-convex: local solutions are not global I Optimization-variable type: I continuous : gradients facilitate computing the solution I discrete: cannot compute gradients, NP-hard I Constraints: I unconstrained : simpler algorithms I constrained : more complex algorithms; must consider feasibility I Number of optimization variables: I low-dimensional : can solve even without gradients I high-dimensional : requires gradients to be solvable in practice 2 Set Notation Set Notation 3 Outline Set Notation Convexity Why Convexity? Convex Sets Convex Functions Convex Optimization Problems Set Notation 4 Set Notation I Rn: set of n-dimensional real vectors I x ∈ C: the point x is an element of set C I C ⊆ Rn: C is a subset of Rn, i.e., elements of C are n-vectors I can describe set elements explicitly: 1 ∈ {3, "cat", 1} I set builder notation C = {x | P (x)} gives the points for which property P (x) is true n I R+ = {x | xi ≥ 0 for all i}: n-vectors with all nonnegative elements I set intersection N \ C = Ci i=1 is the set of points which are simultaneously present in each Ci Set Notation 5 Convexity Convexity 6 Outline Set Notation Convexity Why Convexity? Convex Sets Convex Functions Convex Optimization Problems Convexity 7 Convex Sets I C ⊆ Rn is
    [Show full text]
  • On Nonconvex Subdifferential Calculus in Banach Spaces1
    Journal of Convex Analysis Volume 2 (1995), No.1/2, 211{227 On Nonconvex Subdifferential Calculus in Banach Spaces1 Boris S. Mordukhovich, Yongheng Shao Department of Mathematics, Wayne State University, Detroit, Michigan 48202, U.S.A. e-mail: [email protected] Received July 1994 Revised manuscript received March 1995 Dedicated to R. T. Rockafellar on his 60th Birthday We study a concept of subdifferential for general extended-real-valued functions defined on arbitrary Banach spaces. At any point considered this basic subdifferential is a nonconvex set of subgradients formed by sequential weak-star limits of the so-called Fr´echet "-subgradients of the function at points nearby. It is known that such and related constructions possess full calculus under special geometric assumptions on Banach spaces where they are defined. In this paper we establish some useful calculus rules in the general Banach space setting. Keywords : Nonsmooth analysis, generalized differentiation, extended-real-valued functions, nonconvex calculus, Banach spaces, sequential limits 1991 Mathematics Subject Classification: Primary 49J52; Secondary 58C20 1. Introduction It is well known that for studying local behavior of nonsmooth functions one can suc- cessfully use an appropriate concept of subdifferential (set of subgradients) which replaces the classical gradient at points of nondifferentiability in the usual two-sided sense. Such a concept first appeared for convex functions in the context of convex analysis that has had many significant applications to the range of problems in optimization, economics, mechanics, etc. One of the most important branches of convex analysis is subdifferen- tial calculus for convex extended-real-valued functions that was developed in the 1960's, mainly by Moreau and Rockafellar; see [23] and [25, 27] for expositions and references.
    [Show full text]