BASICS OF CONVEX ANALYSIS

Similar documents
Convex Optimization Notes

Chapter 1. Optimality Conditions: Unconstrained Optimization. 1.1 Differentiable Problems

Lecture 4: Convex Functions, Part I February 1

Chapter 2 Convex Analysis

Local strong convexity and local Lipschitz continuity of the gradient of convex functions

Division of the Humanities and Social Sciences. Supergradients. KC Border Fall 2001 v ::15.45

Convex Functions and Optimization

Optimization and Optimal Control in Banach Spaces

A SET OF LECTURE NOTES ON CONVEX OPTIMIZATION WITH SOME APPLICATIONS TO PROBABILITY THEORY INCOMPLETE DRAFT. MAY 06

The Subdifferential of Convex Deviation Measures and Risk Functions

In Progress: Summary of Notation and Basic Results Convex Analysis C&O 663, Fall 2009

IE 521 Convex Optimization

Strongly convex functions, Moreau envelopes and the generic nature of convex functions with strong minimizers

Convex Analysis and Economic Theory AY Elementary properties of convex functions

Convex Analysis Background

Convex Optimization Conjugate, Subdifferential, Proximation

Almost Convex Functions: Conjugacy and Duality

Convex envelopes, cardinality constrained optimization and LASSO. An application in supervised learning: support vector machines (SVMs)

Convex Optimization Theory. Chapter 5 Exercises and Solutions: Extended Version

Convex Functions. Pontus Giselsson

Subdifferential representation of convex functions: refinements and applications

STABLE AND TOTAL FENCHEL DUALITY FOR CONVEX OPTIMIZATION PROBLEMS IN LOCALLY CONVEX SPACES

Epiconvergence and ε-subgradients of Convex Functions

GEOMETRIC APPROACH TO CONVEX SUBDIFFERENTIAL CALCULUS October 10, Dedicated to Franco Giannessi and Diethard Pallaschke with great respect

MAXIMALITY OF SUMS OF TWO MAXIMAL MONOTONE OPERATORS

Optimality Conditions for Nonsmooth Convex Optimization

Helly's Theorem and its Equivalences via Convex Analysis

SYMBOLIC CONVEX ANALYSIS

Convexity in R n. The following lemma will be needed in a while. Lemma 1 Let x E, u R n. If τ I(x, u), τ 0, define. f(x + τu) f(x). τ.

Math 273a: Optimization Subgradients of convex functions

8. Conjugate functions

Iterative Convex Optimization Algorithms; Part One: Using the Baillon Haddad Theorem

On Gap Functions for Equilibrium Problems via Fenchel Duality

Convex analysis and profit/cost/support functions

Convex Analysis Notes. Lecturer: Adrian Lewis, Cornell ORIE Scribe: Kevin Kircher, Cornell MAE

Optimality Conditions for Constrained Optimization

THE UNIQUE MINIMAL DUAL REPRESENTATION OF A CONVEX FUNCTION

Subgradients. subgradients and quasigradients. subgradient calculus. optimality conditions via subgradients. directional derivatives

A Dual Condition for the Convex Subdifferential Sum Formula with Applications

FIXED POINT ITERATIONS

Lecture 1: Background on Convex Analysis

Some Properties of the Augmented Lagrangian in Cone Constrained Optimization

Robust Duality in Parametric Convex Optimization

3.1 Convexity notions for functions and basic properties

Maximal Monotone Inclusions and Fitzpatrick Functions

Dedicated to Michel Théra in honor of his 70th birthday

STRONG AND CONVERSE FENCHEL DUALITY FOR VECTOR OPTIMIZATION PROBLEMS IN LOCALLY CONVEX SPACES

Math 273a: Optimization Convex Conjugacy

Subgradients. subgradients. strong and weak subgradient calculus. optimality conditions via subgradients. directional derivatives

Shiqian Ma, MAT-258A: Numerical Optimization 1. Chapter 4. Subgradient

Math 5051 Measure Theory and Functional Analysis I Homework Assignment 3

1 Overview. 2 A Characterization of Convex Functions. 2.1 First-order Taylor approximation. AM 221: Advanced Optimization Spring 2016

Convex Analysis and Optimization Chapter 2 Solutions

SWFR ENG 4TE3 (6TE3) COMP SCI 4TE3 (6TE3) Continuous Optimization Algorithm. Convex Optimization. Computing and Software McMaster University

Math 273a: Optimization Subgradients of convex functions

Extreme Abridgment of Boyd and Vandenberghe s Convex Optimization

1 Lesson 0: Intro and motivation (Brunn-Minkowski)

LECTURE 12 LECTURE OUTLINE. Subgradients Fenchel inequality Sensitivity in constrained optimization Subdifferential calculus Optimality conditions

DEFINABLE VERSIONS OF THEOREMS BY KIRSZBRAUN AND HELLY

A Unified Analysis of Nonconvex Optimization Duality and Penalty Methods with General Augmenting Functions

Brøndsted-Rockafellar property of subdifferentials of prox-bounded functions. Marc Lassonde Université des Antilles et de la Guyane

Abstract. In this paper, the concepts and the properties of conjugate maps and directional derivatives of convex vector functions from a subset of

arxiv: v2 [math.fa] 21 Jul 2013

A Geometric Framework for Nonconvex Optimization Duality using Augmented Lagrangian Functions

Duality. for The New Palgrave Dictionary of Economics, 2nd ed. Lawrence E. Blume

Lecture 8 Plus properties, merit functions and gap functions. September 28, 2008

Constrained Optimization and Lagrangian Duality

Math 341: Convex Geometry. Xi Chen

UNDERGROUND LECTURE NOTES 1: Optimality Conditions for Constrained Optimization Problems

Chapter 1 Preliminaries

Continuity. Chapter 4

Convexity and constructive infima

A New Fenchel Dual Problem in Vector Optimization

Semicontinuous functions and convexity

Symmetric Matrices and Eigendecomposition

Static Problem Set 2 Solutions

Convex and Nonconvex Optimization Techniques for the Constrained Fermat-Torricelli Problem

Extended Monotropic Programming and Duality 1

arxiv:math/ v1 [math.fa] 4 Feb 1993

ZERO DUALITY GAP FOR CONVEX PROGRAMS: A GENERAL RESULT

Continuity. Chapter 4

Course 212: Academic Year Section 1: Metric Spaces

NEW MAXIMUM THEOREMS WITH STRICT QUASI-CONCAVITY. Won Kyu Kim and Ju Han Yoon. 1. Introduction

Translative Sets and Functions and their Applications to Risk Measure Theory and Nonlinear Separation

2019 Spring MATH2060A Mathematical Analysis II 1

MATHEMATICAL ECONOMICS: OPTIMIZATION. Contents

Optimization for Communications and Networks. Poompat Saengudomlert. Session 4 Duality and Lagrange Multipliers

7 Duality and Convex Programming

Chapter 2: Preliminaries and elements of convex analysis

Lecture: Duality of LP, SOCP and SDP

g 2 (x) (1/3)M 1 = (1/3)(2/3)M.

Radial Subgradient Descent

Review of Multi-Calculus (Study Guide for Spivak s CHAPTER ONE TO THREE)

Linear and non-linear programming

On the acceleration of the double smoothing technique for unconstrained convex optimization problems

Duality and Utility Maximization

Duality for almost convex optimization problems via the perturbation approach

Vector Spaces. Addition : R n R n R n Scalar multiplication : R R n R n.

CONTINUOUS CONVEX SETS AND ZERO DUALITY GAP FOR CONVEX PROGRAMS

Lecture 3: Lagrangian duality and algorithms for the Lagrangian dual problem

Transcription:

BASICS OF CONVEX ANALYSIS MARKUS GRASMAIR 1. Main Definitions We start with providing the central definitions of convex functions and convex sets. Definition 1. A function f : R n R + } is called convex, if (1) f(λx + (1 λ)y) λf(x) + (1 λ)f(y) for all x, y R n and all 0 < λ < 1. It is called strictly convex, if it is convex and f(λx + (1 λ)y) < λf(x) + (1 λ)f(y) whenever x y, 0 < λ < 1, and f(x) <, f(y) <. It is important to note here that we allow the functions we consider to take the value +. This leads sometimes to complications in the notation, but will be very helpful later on. Indeed, convex functions taking the value + appear naturally as soon as we start looking at duality, which is possibly the central topic in convex analysis. Note, however, that the functions may not take the value. Definition 2. A set K R n is convex, if whenever x, y K, and 0 < λ < 1. λx + (1 λ)y K We note that there is a very close connection between the convexity of functions and the convexity of sets. Given a set K R n, we denote by χ K : R n R + } its characteristic function, given by 0 if x K, χ K (x) = + if x K. Lemma 3. A set K is convex, if and only if its characteristic function χ K is convex. Proof. Assume that K is convex, and let x, y R n, and 0 < λ < 1. If either x K or y K, then λχ K (x) + (1 λ)χ K (y) =, and therefore (1) (with f = χ K ) is automatically satisfied. On the other hand, if x, y K, then also λx+(1 λ)y K, and therefore χ K (λx + (1 λ)y) = 0 = λχ K (x) + (1 λ)χ K (y). Thus χ K is convex. Now assume that χ K is convex, and let x, y K, and 0 < λ < 1. The convexity of χ K implies that χ K (λx + (1 λ)y) λχ K (x) + (1 λ)χ K (y) = 0. Since χ K only takes the values 0 and +, this shows that, actually, χ K (λx + (1 λ)y) = 0, and therefore λx + (1 λ)y K. Thus K is convex. Date: March 2015. 1

2 MARKUS GRASMAIR Conversely, given a function f : R n R + }, define its epigraph 1 as epi(f) := (x, t) R n R : t f(x) }. Lemma 4. The function f : R n R + } is convex, if and only if its epigraph epi(f) R n+1 is convex. Proof. Assume first that the epigraph of f is convex and let x, y R n be such that f(x), f(y) <, and let 0 < λ < 1. Since (x, f(x)), (x, f(y)) epi(f) it follows that λ(x, f(x)) + (1 λ)(y, f(y)) epi(f), which means that λf(x) + (1 λ)f(y) f(λx + (1 λ)y). Since this holds for every x, y and 0 < λ < 1, the function f is convex. Now assume that the function f is convex and let (x, t), (y, s) epi(f), and 0 < λ < 1. Then λt + (1 λ)s λf(x) + (1 λ)f(y) f(λx + (1 λ)y), which implies that λ(x, t) + (1 λ)(y, s) epi(f). Thus epi(f) is convex. 2. Continuity A convex function need not be continuous as can be easily seen from the example of the characteristic function of a convex set, which is only ever continuous in the trivial situations K = and K = R n. An even more worrying situation (at least from an optimization point of view) occurs for instance in the situation of the function f : R R + } defined by +, if x < 0, f(x) = 1, if x = 0, x 2, if x > 0. This function can easily seen to be convex, and it is obviously not continuous at 0. The problem is that it is not even lower semi-continuous: If you approach the value 0 from the right, the function values tend to 0, but f(0) = 1. Thus this is a convex (and obviously coercive) function, which does not attain its minimum. Fortunately, problems with continuity and lower semi-continuity only occur near points where a convex function takes the value +, as the following result shows. In the following we will always denote by dom(f) := x R n : f(x) < } the domain of the function f. We will say that a function is proper, if its domain is non-empty; that is, the function is not identically equal to +. Proposition 5. Assume that f : R n R + } is convex and that x is contained in the interior of dom(f). Then f is locally Lipschitz continuous at x. In other words, discontinuities of convex functions can only appear at the boundary of their domain. Proof. See [3, Thm. 10.4]. 1 The word epigraph is composed of the greek preposition epi meaning above, over and the word graph. Thus it simply means above the graph.

BASICS OF CONVEX ANALYSIS 3 3. Differentiation Assume now that f : R n R + } is any function and that x dom(f). Given d R n, we define the directional derivative of f at x in direction d as f(x + td) f(x) Df(x; d) := lim R ± } t 0 t t>0 provided that the limit exists (in the extended real line R ± }. For general functions f, the directional derivative need not necessarily exist (take for instance the function x x sin(1/x), which can be shown to have no directional derivatives at 0 in directions d 0). For convex functions, however, the situation is different. Lemma 6. Assume that f is convex, x dom(f), and d R n. Then Df(x; d) exists. If x is an element of the interior of dom(f), then Df(x; d) <. Proof. Let 0 t 1 < t 2. Then x + t 1 d = t 1 t 2 (x + t 2 d) + ( 1 t 1 t 2 ) x, and thus the convexity of f implies that ( f(x + t 1 d) 1 t ) 1 f(x) + t 1 f(x + t 2 d), t 2 t 2 which can be rewritten as f(x + t 1 d) f(x) f(x + t 2d) f(x). t 1 t 2 Thus the function t f(x + td) f(x) t is increasing. As a consequence, its limit as t 0 from above which is exactly the directional derivative Df(x; d) exists. In addition, if x is in the interior of the domain of f, then f(x + td) is finite for t sufficiently small, and therefore Df(x; d) <. Using the same argumenation as in the proof of the previous statement, one sees that f(x + td) f(x) + tdf(x; d) for all t 0. Thus a sufficient (and also necessary) condition for x dom(f) to be a global minimizer of f is that Df(x; d) 0 for all d R n. Next, we will introduce another notion of derivatives that is tailored to the needs of convex functions. Definition 7. Let f : R n R + } and x dom(f). An element ξ R n with the property that f(y) f(x) + ξ T (y x) for all y R n is called a subgradient of f at x. The set of all subgradients of f at x is called the subdifferential of f at x, and is denoted by f(x). An immediate consequence of the definition of the subdifferential is the following result:

4 MARKUS GRASMAIR Lemma 8. Assume that f : R n R + } is convex and proper. Then x R n is a global minimizer of f, if and only if Proof. Exercise. 0 f(x). The subdifferential is connected to the usual derivative and the directional derivative in the following way: Proposition 9. Assume that f : R n R + } is convex and that x dom(f). If f is differentiable at x, then f(x) = f(x)}. If f(x) contains a single element, e.g. f(x) = ξ}, then f is differentiable at x and ξ = f(x). Proof. See [3, Thm. 25.1]. Proposition 10. Assume that f : R n R + } is convex and that x dom(f). Then ξ f(x), if and only if Proof. See [3, Thm. 23.2]. ξ T d Df(x; d) for all d R n. Next we will compute the subdifferential in some particular (somehow important) cases: We consider first the function f : R n R, ( n ) 1/2. f(x) = x 2 = x 2 i If x 0, then f is differentiable at x with deriviative f(x) = x/ x 2, and therefore f(x) = x/ x 2 } for x 0. Now consider the case x = 0. We have ξ f(0), if and only if ξ T y f(y) f(0) = y 2 for all y R n, which is equivalent to the condition that ξ 2 1. Thus we have x } if x 0, f(x) = x 2 ξ R n : ξ 2 1 } if x = 0. Consider now the function f : R n R + }, 0, if x 2 1, f(x) = +, if x 2 > 1. If x R n satisfies x < 1, then f is differentiable at x with gradient equal to 0, and therefore f(x) = 0}. If x > 1, then f(x) = +, and therefore f(x) =. Finally, if x = 1, then ξ f(x), if and only if ξ T (y x) f(y) f(x) for all y R n, which is equivalent to the condition ξ T (y x) 0 for all y R n with y 2 1. The latter condition can in turn be shown to be equivalent to stating that ξ = λx for some λ 0.

BASICS OF CONVEX ANALYSIS 5 Thus we obtain that, if x > 1, } f(x) = λx : λ 0 if x = 1, 0}, if x < 1. Now assume that f is convex and that x, y R n, and that ξ f(x) and η f(y). Then the definition of the subdifferential implies that f(y) f(x) ξ T (y x) 0, f(x) f(y) η T (x y) 0. Summing up these inequalities, we see that (η ξ) T (y x) 0 whenever ξ f(x) and η f(y). In case f is differentiable, this inequality becomes ( f(y) f(x)) T (y x) 0 for all x, y R n. Now consider the case of a one-dimensional function. Then this further simplifies to (f (y) f (x))(y x) 0 for all x, y R, which can be restated as f (y) f (x) whenever y x. In other word, the derivative of a one-dimensional function is monotoneously increasing. Even more, the converse is also true: Proposition 11. Assume that f : R R is convex and differentiable. Then f is monotoneously increasing. Conversely, if g : R R is monotoneously increasing and continuous, then there exists a convex function f such that f = g. Proof. The necessity of the monotonicity of f has already been shown above. Conversely, if g is monotoneously increasing, we can define f(x) := x Then, we have for y > x and 0 < λ < 1 that f(λx + (1 λ)y) = λf(x) + (1 λ)f(y) showing that f is convex. + λ 0 λx+(1 λ)y x g(y) dy. g(z) dz (1 λ) y λx+(1 λ)y g(z) dz λf(x) + (1 λ)f(y) + λ(1 λ)(y x)g (λx + (1 λ)y) (1 λ)λ(y x)g (λx + (1 λ)y) = λf(x) + (1 λ)f(y), In higher dimensions, the situation is slightly more complicated. Still, one can show the following result: Proposition 12. Assume that f : R n R is differentiable. Then f is convex, if and only if f is increasing in the sense that ( f(y) f(x)) T (y x) 0 for all x, y R n. Proof. Apply Proposition 11 to the one-dimensional functions t f(x + t(y x)).

6 MARKUS GRASMAIR Note the difference between the two different statements: In the first proposition (in one dimension), the monotonicity of g implies that it is the derivative of some convex function. In the second proposition (in higher dimensions), we already know that the object we are dealing with ( f) is the derivative of some function; monotonocity of the gradient implies that f is convex, but we do not have to care about the existence of f. Finally, it is important to note that there is a close connection between properties of the Hessian of a function and its convexity: Proposition 13. Assume that f : R n R + } is twice differentiable on its domain, and that dom(f) is convex. Then f is convex, if and only if 2 f(x) is positive semi-definite for all x dom(f). In addition, if 2 f(x) is positive definite for all but a countable number of points x dom(f), then f is strictly convex. 4. Transformations of convex functions Next we will discuss, which operations preserve the convexity of functions, and how these operations affect the subdifferential. The first thing to note is that convex functions behave well under multiplication with positive scalars: If f : R n R + } is convex and λ > 0, then also λf is convex and (λf)(x) = λ f(x) for all x R n. Note that this result does not hold for negative scalars: If f is convex and λ < 0, then usually λf will not convex. Also, for λ = 0 some problems occur if f takes the value +. First, it is not clear how to define the product 0 (+ ) in a meaningful way, and, second, the equality (0 f)(x) = 0 f(x) need not necessarily be satisfied for all x. Next we consider a function that is defined as a supremum of convex functions: Lemma 14. Assume that I is any index set (not necessarily finite or even countable) and that f i : R n R + }, i I, are convex. Define g(x) := sup f i (x). i I Then g is convex. In addition, if x R n and g(x) = f j (x) for some j I, then Proof. If x, y R n and 0 < λ < 1, then f j (x) g(x). g(λx + (1 λ)y) = sup f i (λx + (1 λ)y) i I ( ) sup λf i (x) + (1 λ)f i (y) i I λ sup f i (x) + (1 λ) sup f i (y) i I i I = λg(x) + (1 λ)g(y). Thus g is convex. In addition, if g(x) = f j (x) and ξ f j (x), then g(y) f j (y) f j (x) + ξ T (y x) = g(x) + ξ T (y x), showing that ξ f j (x).

BASICS OF CONVEX ANALYSIS 7 There are two important things to note here: First, the result states that the supremum of any number of convex functions is convex; there is no need to restrict oneself to only finitely many functions. Second, the minimum even of two convex functions is usually not convex any more. Next we discuss sums of convex functions: Proposition 15. Assume that f, g : R n R + } are convex. Then also the function f + g is convex and (f + g)(x) f(x) + g(x) for all x R n. If in addition f and g are lower semi-continuous and there exists some y dom(f) dom(g) such that either one of the functions f and g is continuous at y, then (f + g)(x) = f(x) + g(x) for all x R n. Proof. The first part of the proposition is straightforward: (f + g)(λx + (1 λ)y) = f(λx + (1 λ)y) + g(λx + (1 λ)y) Also, if ξ f(x) and η g(x), then λf(x) + (1 λ)f(y) + λg(x) + (1 λ)g(y) = λ(f + g)(x) + (1 λ)(f + g)(y). (f + g)(y) f(x) + ξ T (y x) + g(x) + η T (y x) = (f + g)(x) + (ξ + η) T (y x), showing that (f + g)(x) f(x) + g(x). The second part is tricky. A proof can for instance be found in [2, Chap. I, Prop. 5.6]. In the above proposition, the sum of subdifferentials is defined as f(x) + g(x) = ξ + η R n : ξ f(x) and η g(x) }. Example 16. Define the functions f, g : R 2 R + }, 0, if x 2 1, 0, if x (2, 0) 2 1, f(x) = and g(x) = +, if x 2 > 1, +, if x (2, 0) 2 > 1, Both of these functions are convex. Moreover 0, if x = (1, 0), (f + g)(x) = + else. Now consider the subdifferential of the different functions at (1, 0). We have f(1, 0) = λ(1, 0) : λ 0 } and Thus On the other hand, Thus in the case we have g(1, 0) = µ( 1, 0) : µ 0 }. f(1, 0) + g(1, 0) = R(1, 0). (f + g)(1, 0) = R 2. f(1, 0) + g(1, 0) (f + g)(1, 0). Finally, we study the composition of convex functions and linear transformations:

8 MARKUS GRASMAIR Proposition 17. Assume that f : R n R + } is convex and that A R n m. Then the function g : R m R + } defined by g(x) = f(ax) is convex. If in addition there exists y Ran A such that f is continuous at y, then for all x R n. g(x) = A T f(ax) Proof. Proving the convexity of g is straightforward. The proof of the representation of the subdifferential can for instance be found in [2, Chap. I, Prop. 5.7]. Note that the second part of the previous result is just a kind of chain rule for the composition of convex and linear functions. In the particular case where f is differentiable, the result reads as (f A)(x) = A T f(ax), which is exactly the usual chain rule. Differentiability, however, is not required in the convex setting. 5. Duality Assume that f : R n R + } is convex and lower semi-continuous and consider any linear function x ξ T x (for some ξ R n ). Then it is easy to convince oneself that there exists some β R such that f(x) ξ T x β for all x R n (this follows basically from the definition and the local Lipschitz continuity of convex functions). In other words, β ξ T x f(x) for all x R n. Now consider the smallest such number β, that is, β := inf β R : β ξ T x f(x) for all x R n}. It is easy to see that β > unless the function f is everywhere equal to +. Thus one can equivalently write β = sup x R n ξ T x f(x). The mapping that maps ξ to β plays a central role in covnex analysis. Definition 18. Let f : R n R + } be proper. Then its conjugate f : R n R + } is defined as f (ξ) := sup x R n ξ T x f(x). Lemma 19. Let f : R n R + } be proper. semi-continuous. Then f is convex and lower Proof. The function f is the supremum of the collection of all affine (and thus in particular convex and lower semi-continuous) functions h x (ξ) := ξ T x f(x). Thus also f is convex and lower semi-continuous.

BASICS OF CONVEX ANALYSIS 9 Example 20. Let f(x) = 1 p x p p = 1 p n x i p with 1 < p < +. In order to compute f (ξ), we have to maximize the function x h ξ (x) := ξ f(x). Since h ξ is concave (that is, h ξ is convex) and differentiable, maximization is equivalent to solving the equation h ξ (x) = 0. Now note that h ξ (x) = ξ f(x), and Thus the maximum x satisfies f(x) = x 1 x 1 p 2. x n x n p 2 x i x i p 2 = ξ i for all i. Solving these equations for x i, we obtain Thus Defining this can be written as Example 21. Let Then f (ξ) = h ξ (x ) = f (ξ) = 1 p x i = ξ i ξ i 1 p 1 1.. n 1 ξ i p 1 +1 1 p p := n p p 1, f(x) = x 1 = n ξ i ξ i p = 1 p ξ p p. n x i. ( f (ξ) = sup ξ T x x R n n ) x i. p p 1. Now assume that ξ j > 1 for some 1 j n. Now consider only vectors x in the supremum above for which the j th entry is λ sgn ξ j and all the other entries of which are zero. Then we see that f (ξ) sup λ>0 ( ξj λ sgn ξ j λ ) = λ( ξ j 1) = +. On the other hand, if ξ j 1 for all j (or in other words ξ 1, then ( n ) n f (ξ) = sup ξ T ( x x i = sup ξi x i x i ) = 0, x R n x R n λ>0 since all the entries in the sum are non-positive. Thus we see that f 0, if ξ 1, (ξ) = +, if ξ > 1. A trivial consequence of the definition of the conjugate of a function is the so called Fenchel inequality:

10 MARKUS GRASMAIR Lemma 22. If f : R n R + } is proper, then f(x) + f (ξ) ξ T x for all x, ξ R n. Proof. This follows immediately from the definition of f. Proposition 23. Assume that f : R n R + } is proper, convex and lower semi-continuous and that x, ξ R n. Then the following are equivalent: ξ T x = f(x) + f (ξ), ξ f(x), x f (ξ). Proof. See [3, Thm. 23.5]. As a next step, it is also possible to compute the conjugate of the f (x) := (f ) (x) = sup ξ R n ( ξ T x f (ξ) ). Theorem 24. Assume that f : R n R + } is proper, convex, and lower semi-continuous. Then f = f. Sketch of Proof. First recall that f is by construction convex and lower semicontinuous. Moreover we have that f ( (x) = sup ξ T x f (ξ) ) ( = sup ξ T ( x sup ξ T y f(y) )) ξ R n ξ R n y R n ( = sup inf ξ T (x y) + f(y) ) f(x) y R n ξ R n for every x R n (for the last inequality, we simply choose y = x). Thus f is a convex and lower semicontinuous function that is smaller or equal to f. Now assume that there exists some x R n such that f (x) < f(x). One can show 2 that, because of the convexity and lower semicontinuity of f, in this case there exist δ > 0 and ξ R n such that f(y) f (x) + ξ T (y x) + δ for all y R n. In other words, the function f and the point f (x) can be separated strictly by the (shifted) hyperplane defined by ξ. Therefore f ( (x) = sup inf ζ T (x y) + f(y) ) ζ R n y R n ( inf ζ T (x y) + f(y) ) f (x) + δ, y R n which is obviously a contradiction. Thus f (x) f(x) for every x R n. Since we have already shown that f (x) f(x) for all x, this implies that, actually f = f. Remark 25. Note that we have actually shown in the proof that for all proper functions f (not necessarily convex or lower semicontinuous), the function f is the largest convex and lower semicontinuous function below f. That is, the function f is the convex and lower semicontinuous hull of f. 2 Basically, the proof of this statement relies on so called separation theorems, which state that it is possible to separate a closed convex set from a point outside said set by means of a hyperplane. Applying this result to the epigraph of f yields the claimed result. A large collection of such separation theorems can be found in [3, Section 11].

BASICS OF CONVEX ANALYSIS 11 The following list taken from [1, p. 50] gives a short overview of some important functions and their conjugates. Note that all these functions are lower semicontinuous and convex, and thus we have f = f; that is, the functions on the left hand side are also the conjugates of the functions on the right hand side: f(x) = 0, f 0, if ξ = 0, (ξ) = 0 if x 0, f(x) = + else, 0 if x 1, f(x) = + else, f (ξ) = f (ξ) = ξ, + else, 0 if ξ 0, + else, f(x) = x p /p, f (ξ) = ξ p /p, f(x) = 1 + x 2, f 1 ξ (ξ) = 2, if y 1, +, else, f(x) = e x, f ξ log ξ ξ, if ξ 0, (ξ) = + else. In addition, the next list provides a few basic rules for the computation of conjugates: f(x) = h(tx), f (ξ) = h (ξ/t) t 0, f(x) = h(x + y), f (ξ) = h (ξ) ξ T y, y R n, f(x) = λh(x), f (ξ) = λh (ξ/λ), λ > 0. 6. Primal and dual problem We now consider the special (but important) case of an optimization problem on R n of the form f(x) + g(ax) min, where f : R n R + } is convex, A R m n, and g : R m R + } is convex. Its dual problem is the optimization problem on R m given by f (A T ξ) + g ( ξ) min. One of the central theorems in convex analysis is the following one, which shows that, under certain non-restrictive assumptions, these two problems are equivalent in some sense. Theorem 26. Assume that f : R n R + } is convex, A R m n, and g : R m R + } is convex. Assume moreover that there exists x dom(f) such that g is continuous at Ax. Assume moreover that x R n and ξ R m. Then the following are equivalent: (1) The vector x minimizes the function x f(x) + g(ax) and ξ minimizes the function ξ f (A T ξ) + g ( ξ). (2) We have (3) We have (4) We have f(x ) + g(ax ) + f (A T ξ ) + g ( ξ ) = 0. x f (A T ξ ) and Ax g ( ξ ). A T ξ f(x ) and ξ g(ax ).

12 MARKUS GRASMAIR Proof. See [3, Thms. 31.2, 31.3]. Basically, this result allows one to switch freely between the primal (original) problem and the dual problem and to try to solve the one that appears to be easier. For instance, if n m, the primal problem may be much more complicated than the dual problem, because the number of variables is much larger. Thus, instead of solving the n-dimensional primal problem, one can try the solve the m- dimensional dual problem. If in addition the function f is differentiable (which is the case if f is strictly convex), then x can be easily recovered from the condition x f (A T ξ ), since the subdifferential of f contains only a single element. References [1] Jonathan M. Borwein and Adrian S. Lewis. Convex analysis and nonlinear optimization. CMS Books in Mathematics/Ouvrages de Mathématiques de la SMC, 3. Springer-Verlag, New York, 2000. Theory and examples. [2] I. Ekeland and R. Temam. Convex Analysis and Variational Problems. North-Holland, Amsterdam, 1976. [3] R. T. Rockafellar. Convex Analysis, volume 28 of Princeton Mathematical Series. Princeton University Press, Princeton, 1970. Department of Mathematics, Norwegian University of Science and Technology, 7491 Trondheim, Norway E-mail address: markus.grasmair@math.ntnu.no