Alexandrov's Theorem on the Second Derivatives of Convex Functions Via

Total Page:16

File Type:pdf, Size:1020Kb

Alexandrov's Theorem on the Second Derivatives of Convex Functions Via ALEXANDROV'S THEOREM ON THE SECOND DERIVATIVES OF CONVEX FUNCTIONS VIA RADEMACHER'S THEOREM ON THE FIRST DERIVATIVES OF LIPSCHITZ FUNCTIONS RALPH HOWARD DEPARTMENT OF MATHEMATICS UNIVERSITY OF SOUTH CAROLINA COLUMBIA, S.C. 29208, USA [email protected] These are notes from lectures given the the functional analysis seminar at the University of South Carolina. Contents 1. Introduction 1 2. Rademacher's Theorem 2 3. A General Sard Type Theorem for Maps Between Spaces of the Same Dimension 6 3.1. A more general result. 11 4. An Inverse Function Theorem for Continuous Functions Differentiable at a Single Point 15 5. Derivatives of Set-Valued Functions and Inverses of Lipschitz Functions 17 6. Alexandrov's Theorem 19 7. Symmetry of the Second Derivatives 22 References 22 1. Introduction A basic result in the regularity theory of convex sets and functions is the theorem of Alexandrov that a convex function has second derivatives almost everywhere. The notes here are a proof of this following the ideas in the appendix of the article [4] of Crandall, Ishii, and Lions and they attribute the main idea of the proof to F. Mignot [5]. To make the notes more self contained I have included a proof of Rademacher's theorem on the differen- tiable almost everywhere of Lipschitz functions following the presentation in the book [8] of Ziemer (which I warmly recommend to anyone wanting Date: May 1998. 1 2 RALPH HOWARD to learn about the pointwise behavior of functions in Sobolev spaces or of bounded variation). Actually a slight generalization of Alexandrov's theo- rem is given in Theorem 5.3 which shows that set-valued functions that are inverses to Lipschitz functions are differentiable almost everywhere. To simplify notation I have assumed that functions have domains all of Rn. It is straightforward to adapt these proofs to locally Lipschitz functions or convex function defined on convex open subsets of Rn. As to notation. If x; y Rn then the inner product is denoted as usual 2 by either x y or x; y . Explicitly if x = (x1; : : : ; xn) and y = (y1; : : : ; yn) · h i x y = x; y = x1y1 + + xnyn: · h i ··· The norm of x is x = px x. Lebesgue measure in Rn will be denoted by n k k · and integrals with respect to this measure will be written as Rn f(x) dx. L n R If x0 R and r > 0 then the open and closed balls about x0 will be denoted by 2 n n B(x0; r) := x R : x x0 < r B(x0; r) x R : x x0 r : f 2 k − k g f 2 k − k ≤ g 2. Rademacher's Theorem We first review a little about Lipschitz functions in one variable. The following is a special case of a theorem of Lebesgue. 2.1. Theorem. Let f : R R satisfy f(x1) f(x0) M x1 x0 . Then ! j − j ≤ j − j the derivative f 0(t) exists for almost all t and f 0(t) M j j ≤ holds at all points where it does exist. Also for a < b b Z f 0(t) dt = f(b) f(a): a − Note if f : R R is Lipschitz and ' C01(R) then the product '(t)f(t) is also Lipschitz! and so the last result implies2 Z f 0(t)'(t) dt = Z f(t)'0(t) dt: R − R n Now if f : R R then denote by Dif the partial derivative ! @f Dif(x) = (x) @xi at points where this partial derivative exists. Let Df(x) denote Df(x) = (D1f(x);:::;Dnf(x)): n 2.2. Proposition. If f : R R is Lipschitz, say f(x1) f(x0) ! nk − k ≤ M x1 x0 , then Df(x) exists for almost all x R . Moreover all the k − k 2 partial derivative Dif satisfy Dif(x) M j j ≤ THEOREMS OF RADEMACHER AND ALEXANDROV 3 at points where they exist. Thus (2.1) Df(x) pnM: k k ≤ Finally if ' C (Rn) then 2 01 (2.2) Z Dif(x)'(x) dx = Z f(x)Di'(x) dx: Rn − Rn Proof. We show that D1f(x) exists almost everywhere, the argument for Dif being identical. Write x = (x1; x2; : : : ; xn) as x = (x1; x0) where x0 = n 1 (x2; : : : ; xn). Then for any x R let 0 2 − Nx := x1 R : D1f(x1; x0) does not exist. : 0 f 2 g Then by the one variable result N is a set of measure zero in R for all x0 x Rn 1. Therefore by Fubini's the set 0 2 − n N = Nx = x R : D1f(x) does not exist [ 0 f 2 g x Rn 1 02 − is a set of measure zero. That Dif(x) M at points where it exists is clear (or follows from the one dimensionalj j ≤ result). At points where Df(x) exists 2 2 2 2 Df(x) = D1f(x) + + Dnf(x) M + + M = pnM: k k p ··· ≤ p ··· Finally we show (2.2) in the case of i = 1. Using the notation above, Fubini's theorem, and the one variable integration by parts formula. Z D1f(x)'(x) dx = Z Z D1f(x1; x0)'(x1; x0) dx1 dx0 n n 1 R R − R = Z Z f(x1; x0)D1'(x1; x0) dx1 dx0 n 1 − R − R = Z f(x)D1'(x) dx − Rn This completes the proof. 2.3. Definition. Let f : Rn R, then for a fixed vector v Rn define ! 2 f(x + tv) f(x) df(x; v) := lim − : t 0 t ! When this limit exists it is directional derivative of f in the direction of v at the point x. 2.4. Proposition. Let f : Rn R be Lipschitz and let v Rn be a fixed vector. Then df(x; v) exists for almost! all x Rn and is given2 by the formula 2 (2.3) df(x; v) = Df(x) v · for almost all x. 4 RALPH HOWARD n Proof. Note if v = e1 where e1; : : : ; en is the standard coordinate basis of R then df(x; v) = df(x; e1) = D1f(x) and the fact that df(x; v) exists almost everywhere follows from Proposition 2.2. In the general case if v = 0 (and 6 n the case v = 0 is trivial) there is a linear coordinate system ξ1; : : : ; ξn on R so that df(x; v) = @f . But again Proposition 2.2 can be used to see that @ξ1 df(x; v) exits for almost all x Rn. 2 n To see that the formula (2.3) holds let ' C01(R ). Then as ' is smooth the usual form of the chain rule implies d'(2x; v) = D'(x) v. Let M be the Lipschitz constant of f. Then · f(x + tv) f(x) M tv − '(x) k k '(x) M v ' L : t ≤ t j j ≤ k kk k 1 j j f(x+tv) f(x) Therefore for 0 < t 1 the function x t − '(x) is uniformly j j ≤ 7! bounded and has compact support. Thus by the dominated convergence theorem and the version of integration by parts given in Proposition 2.2 f(x + tv) f(x) Z df(x; v)'(x) dx = lim Z − '(x) dx n t 0 n t R ! R 1 = lim Z f(x + tv)'(x) dx Z f(x)'(x) dx t 0 t n − n ! R R 1 = lim Z f(x)'(x tv) dx Z f(x)'(x) dx t 0 t n − − n ! R R '(x tv) '(x) = lim Z f(x) − − dx t 0 n t ! R = Z f(x)d'(x; v) dx Rn − = Z f(x)D'(x) v dx − Rn · n = vi Z f(x)Di'(x) dx − X n i=1 R n = vi Z Dif(x)'(x) dx X n i=1 R = Z Df(x) v'(x) dx: Rn · n Thus Rn df(x; v)'(x) dx = Rn Df(x) v'(x) dx for all ' C1(R ). ThereforeR d(x; v) = Df(x) v Rfor almost all· x Rn. 2 · 2 n m n 2.5. Definition. Let f : R R . Then f is differentiable at x0 R iff there is a linear map L: Rn! Rm so that 2 ! f(x) f(x0) = L(x x0) + o( x x0 ): − − k − k THEOREMS OF RADEMACHER AND ALEXANDROV 5 In this case the linear map L is easily seen to be unique and will be denoted by f 0(x0). By o( x x0 ) we mean a function of the form x x0 g(x; x0) where k − k k − k limx x0 g(x; x0) = 0. This definition could be given a little more formally by letting! Sn 1 = u Rn : u be the unit sphere in Rn. Then f : Rn Rm − f 2 k kg ! is differentiable at x with f 0(x) = L iff for all " > 0 there is a δ > 0 so that for all u Sn 1 2 − f(x0 + tu) f(x0) (2.4) 0 < t δ implies − Lu < ": j j ≤ t − 2.6. Theorem (Rademacher [6] 1919). If f : Rn Rm is Lipschitz, say ! f(x1) f(x0) M x1 x0 , then the derivative f 0(x) exists for almost allk x −Rn. Ink the ≤ casek m−= 1,k so that f : Rn R, then f (x) is given by 2 ! 0 f 0(x)v = Df(x) v · for almost all x Rn. 2 Proof. We first consider the case when m = 1 so that f is scalar valued.
Recommended publications
  • The Fundamental Theorem of Calculus for Lebesgue Integral
    Divulgaciones Matem´aticasVol. 8 No. 1 (2000), pp. 75{85 The Fundamental Theorem of Calculus for Lebesgue Integral El Teorema Fundamental del C´alculo para la Integral de Lebesgue Di´omedesB´arcenas([email protected]) Departamento de Matem´aticas.Facultad de Ciencias. Universidad de los Andes. M´erida.Venezuela. Abstract In this paper we prove the Theorem announced in the title with- out using Vitali's Covering Lemma and have as a consequence of this approach the equivalence of this theorem with that which states that absolutely continuous functions with zero derivative almost everywhere are constant. We also prove that the decomposition of a bounded vari- ation function is unique up to a constant. Key words and phrases: Radon-Nikodym Theorem, Fundamental Theorem of Calculus, Vitali's covering Lemma. Resumen En este art´ıculose demuestra el Teorema Fundamental del C´alculo para la integral de Lebesgue sin usar el Lema del cubrimiento de Vi- tali, obteni´endosecomo consecuencia que dicho teorema es equivalente al que afirma que toda funci´onabsolutamente continua con derivada igual a cero en casi todo punto es constante. Tambi´ense prueba que la descomposici´onde una funci´onde variaci´onacotada es ´unicaa menos de una constante. Palabras y frases clave: Teorema de Radon-Nikodym, Teorema Fun- damental del C´alculo,Lema del cubrimiento de Vitali. Received: 1999/08/18. Revised: 2000/02/24. Accepted: 2000/03/01. MSC (1991): 26A24, 28A15. Supported by C.D.C.H.T-U.L.A under project C-840-97. 76 Di´omedesB´arcenas 1 Introduction The Fundamental Theorem of Calculus for Lebesgue Integral states that: A function f :[a; b] R is absolutely continuous if and only if it is ! 1 differentiable almost everywhere, its derivative f 0 L [a; b] and, for each t [a; b], 2 2 t f(t) = f(a) + f 0(s)ds: Za This theorem is extremely important in Lebesgue integration Theory and several ways of proving it are found in classical Real Analysis.
    [Show full text]
  • CORE View Metadata, Citation and Similar Papers at Core.Ac.Uk
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by Bulgarian Digital Mathematics Library at IMI-BAS Serdica Math. J. 27 (2001), 203-218 FIRST ORDER CHARACTERIZATIONS OF PSEUDOCONVEX FUNCTIONS Vsevolod Ivanov Ivanov Communicated by A. L. Dontchev Abstract. First order characterizations of pseudoconvex functions are investigated in terms of generalized directional derivatives. A connection with the invexity is analysed. Well-known first order characterizations of the solution sets of pseudolinear programs are generalized to the case of pseudoconvex programs. The concepts of pseudoconvexity and invexity do not depend on a single definition of the generalized directional derivative. 1. Introduction. Three characterizations of pseudoconvex functions are considered in this paper. The first is new. It is well-known that each pseudo- convex function is invex. Then the following question arises: what is the type of 2000 Mathematics Subject Classification: 26B25, 90C26, 26E15. Key words: Generalized convexity, nonsmooth function, generalized directional derivative, pseudoconvex function, quasiconvex function, invex function, nonsmooth optimization, solution sets, pseudomonotone generalized directional derivative. 204 Vsevolod Ivanov Ivanov the function η from the definition of invexity, when the invex function is pseudo- convex. This question is considered in Section 3, and a first order necessary and sufficient condition for pseudoconvexity of a function is given there. It is shown that the class of strongly pseudoconvex functions, considered by Weir [25], coin- cides with pseudoconvex ones. The main result of Section 3 is applied to characterize the solution set of a nonlinear programming problem in Section 4. The base results of Jeyakumar and Yang in the paper [13] are generalized there to the case, when the function is pseudoconvex.
    [Show full text]
  • [Math.FA] 3 Dec 1999 Rnfrneter for Theory Transference Introduction 1 Sas Ihnrah U Twl Etetdi Eaaepaper
    Transference in Spaces of Measures Nakhl´eH. Asmar,∗ Stephen J. Montgomery–Smith,† and Sadahiro Saeki‡ 1 Introduction Transference theory for Lp spaces is a powerful tool with many fruitful applications to sin- gular integrals, ergodic theory, and spectral theory of operators [4, 5]. These methods afford a unified approach to many problems in diverse areas, which before were proved by a variety of methods. The purpose of this paper is to bring about a similar approach to spaces of measures. Our main transference result is motivated by the extensions of the classical F.&M. Riesz Theorem due to Bochner [3], Helson-Lowdenslager [10, 11], de Leeuw-Glicksberg [6], Forelli [9], and others. It might seem that these extensions should all be obtainable via transference methods, and indeed, as we will show, these are exemplary illustrations of the scope of our main result. It is not straightforward to extend the classical transference methods of Calder´on, Coif- man and Weiss to spaces of measures. First, their methods make use of averaging techniques and the amenability of the group of representations. The averaging techniques simply do not work with measures, and do not preserve analyticity. Secondly, and most importantly, their techniques require that the representation is strongly continuous. For spaces of mea- sures, this last requirement would be prohibitive, even for the simplest representations such as translations. Instead, we will introduce a much weaker requirement, which we will call ‘sup path attaining’. By working with sup path attaining representations, we are able to prove a new transference principle with interesting applications. For example, we will show how to derive with ease generalizations of Bochner’s theorem and Forelli’s main result.
    [Show full text]
  • Convexity I: Sets and Functions
    Convexity I: Sets and Functions Ryan Tibshirani Convex Optimization 10-725/36-725 See supplements for reviews of basic real analysis • basic multivariate calculus • basic linear algebra • Last time: why convexity? Why convexity? Simply put: because we can broadly understand and solve convex optimization problems Nonconvex problems are mostly treated on a case by case basis Reminder: a convex optimization problem is of ● the form ● ● min f(x) x2D ● ● subject to gi(x) 0; i = 1; : : : m ≤ hj(x) = 0; j = 1; : : : r ● ● where f and gi, i = 1; : : : m are all convex, and ● hj, j = 1; : : : r are affine. Special property: any ● local minimizer is a global minimizer ● 2 Outline Today: Convex sets • Examples • Key properties • Operations preserving convexity • Same for convex functions • 3 Convex sets n Convex set: C R such that ⊆ x; y C = tx + (1 t)y C for all 0 t 1 2 ) − 2 ≤ ≤ In words, line segment joining any two elements lies entirely in set 24 2 Convex sets Figure 2.2 Some simple convexn and nonconvex sets. Left. The hexagon, Convex combinationwhich includesof x1; its : :boundary : xk (shownR : darker), any linear is convex. combinationMiddle. The kidney shaped set is not convex, since2 the line segment between the twopointsin the set shown as dots is not contained in the set. Right. The square contains some boundaryθ1x points1 + but::: not+ others,θkxk and is not convex. k with θi 0, i = 1; : : : k, and θi = 1. Convex hull of a set C, ≥ i=1 conv(C), is all convex combinations of elements. Always convex P 4 Figure 2.3 The convex hulls of two sets in R2.
    [Show full text]
  • Convex Functions
    Convex Functions Prof. Daniel P. Palomar ELEC5470/IEDA6100A - Convex Optimization The Hong Kong University of Science and Technology (HKUST) Fall 2020-21 Outline of Lecture • Definition convex function • Examples on R, Rn, and Rn×n • Restriction of a convex function to a line • First- and second-order conditions • Operations that preserve convexity • Quasi-convexity, log-convexity, and convexity w.r.t. generalized inequalities (Acknowledgement to Stephen Boyd for material for this lecture.) Daniel P. Palomar 1 Definition of Convex Function • A function f : Rn −! R is said to be convex if the domain, dom f, is convex and for any x; y 2 dom f and 0 ≤ θ ≤ 1, f (θx + (1 − θ) y) ≤ θf (x) + (1 − θ) f (y) : • f is strictly convex if the inequality is strict for 0 < θ < 1. • f is concave if −f is convex. Daniel P. Palomar 2 Examples on R Convex functions: • affine: ax + b on R • powers of absolute value: jxjp on R, for p ≥ 1 (e.g., jxj) p 2 • powers: x on R++, for p ≥ 1 or p ≤ 0 (e.g., x ) • exponential: eax on R • negative entropy: x log x on R++ Concave functions: • affine: ax + b on R p • powers: x on R++, for 0 ≤ p ≤ 1 • logarithm: log x on R++ Daniel P. Palomar 3 Examples on Rn • Affine functions f (x) = aT x + b are convex and concave on Rn. n • Norms kxk are convex on R (e.g., kxk1, kxk1, kxk2). • Quadratic functions f (x) = xT P x + 2qT x + r are convex Rn if and only if P 0.
    [Show full text]
  • Generalizations of the Riemann Integral: an Investigation of the Henstock Integral
    Generalizations of the Riemann Integral: An Investigation of the Henstock Integral Jonathan Wells May 15, 2011 Abstract The Henstock integral, a generalization of the Riemann integral that makes use of the δ-fine tagged partition, is studied. We first consider Lebesgue’s Criterion for Riemann Integrability, which states that a func- tion is Riemann integrable if and only if it is bounded and continuous almost everywhere, before investigating several theoretical shortcomings of the Riemann integral. Despite the inverse relationship between integra- tion and differentiation given by the Fundamental Theorem of Calculus, we find that not every derivative is Riemann integrable. We also find that the strong condition of uniform convergence must be applied to guarantee that the limit of a sequence of Riemann integrable functions remains in- tegrable. However, by slightly altering the way that tagged partitions are formed, we are able to construct a definition for the integral that allows for the integration of a much wider class of functions. We investigate sev- eral properties of this generalized Riemann integral. We also demonstrate that every derivative is Henstock integrable, and that the much looser requirements of the Monotone Convergence Theorem guarantee that the limit of a sequence of Henstock integrable functions is integrable. This paper is written without the use of Lebesgue measure theory. Acknowledgements I would like to thank Professor Patrick Keef and Professor Russell Gordon for their advice and guidance through this project. I would also like to acknowledge Kathryn Barich and Kailey Bolles for their assistance in the editing process. Introduction As the workhorse of modern analysis, the integral is without question one of the most familiar pieces of the calculus sequence.
    [Show full text]
  • Stability in the Almost Everywhere Sense: a Linear Transfer Operator Approach ∗ R
    CORE Metadata, citation and similar papers at core.ac.uk Provided by Elsevier - Publisher Connector J. Math. Anal. Appl. 368 (2010) 144–156 Contents lists available at ScienceDirect Journal of Mathematical Analysis and Applications www.elsevier.com/locate/jmaa Stability in the almost everywhere sense: A linear transfer operator approach ∗ R. Rajaram a, ,U.Vaidyab,M.Fardadc, B. Ganapathysubramanian d a Dept. of Math. Sci., 3300, Lake Rd West, Kent State University, Ashtabula, OH 44004, United States b Dept. of Elec. and Comp. Engineering, Iowa State University, Ames, IA 50011, United States c Dept. of Elec. Engineering and Comp. Sci., Syracuse University, Syracuse, NY 13244, United States d Dept. of Mechanical Engineering, Iowa State University, Ames, IA 50011, United States article info abstract Article history: The problem of almost everywhere stability of a nonlinear autonomous ordinary differential Received 20 April 2009 equation is studied using a linear transfer operator framework. The infinitesimal generator Available online 20 February 2010 of a linear transfer operator (Perron–Frobenius) is used to provide stability conditions of Submitted by H. Zwart an autonomous ordinary differential equation. It is shown that almost everywhere uniform stability of a nonlinear differential equation, is equivalent to the existence of a non-negative Keywords: Almost everywhere stability solution for a steady state advection type linear partial differential equation. We refer Advection equation to this non-negative solution, verifying almost everywhere global stability, as Lyapunov Density function density. A numerical method using finite element techniques is used for the computation of Lyapunov density. © 2010 Elsevier Inc. All rights reserved. 1. Introduction Stability analysis of an ordinary differential equation is one of the most fundamental problems in the theory of dynami- cal systems.
    [Show full text]
  • Sum of Squares and Polynomial Convexity
    CONFIDENTIAL. Limited circulation. For review only. Sum of Squares and Polynomial Convexity Amir Ali Ahmadi and Pablo A. Parrilo Abstract— The notion of sos-convexity has recently been semidefiniteness of the Hessian matrix is replaced with proposed as a tractable sufficient condition for convexity of the existence of an appropriately defined sum of squares polynomials based on sum of squares decomposition. A multi- decomposition. As we will briefly review in this paper, by variate polynomial p(x) = p(x1; : : : ; xn) is said to be sos-convex if its Hessian H(x) can be factored as H(x) = M T (x) M (x) drawing some appealing connections between real algebra with a possibly nonsquare polynomial matrix M(x). It turns and numerical optimization, the latter problem can be re- out that one can reduce the problem of deciding sos-convexity duced to the feasibility of a semidefinite program. of a polynomial to the feasibility of a semidefinite program, Despite the relative recency of the concept of sos- which can be checked efficiently. Motivated by this computa- convexity, it has already appeared in a number of theo- tional tractability, it has been speculated whether every convex polynomial must necessarily be sos-convex. In this paper, we retical and practical settings. In [6], Helton and Nie use answer this question in the negative by presenting an explicit sos-convexity to give sufficient conditions for semidefinite example of a trivariate homogeneous polynomial of degree eight representability of semialgebraic sets. In [7], Lasserre uses that is convex but not sos-convex. sos-convexity to extend Jensen’s inequality in convex anal- ysis to linear functionals that are not necessarily probability I.
    [Show full text]
  • Subgradients
    Subgradients Ryan Tibshirani Convex Optimization 10-725/36-725 Last time: gradient descent Consider the problem min f(x) x n for f convex and differentiable, dom(f) = R . Gradient descent: (0) n choose initial x 2 R , repeat (k) (k−1) (k−1) x = x − tk · rf(x ); k = 1; 2; 3;::: Step sizes tk chosen to be fixed and small, or by backtracking line search If rf Lipschitz, gradient descent has convergence rate O(1/) Downsides: • Requires f differentiable next lecture • Can be slow to converge two lectures from now 2 Outline Today: crucial mathematical underpinnings! • Subgradients • Examples • Subgradient rules • Optimality characterizations 3 Subgradients Remember that for convex and differentiable f, f(y) ≥ f(x) + rf(x)T (y − x) for all x; y I.e., linear approximation always underestimates f n A subgradient of a convex function f at x is any g 2 R such that f(y) ≥ f(x) + gT (y − x) for all y • Always exists • If f differentiable at x, then g = rf(x) uniquely • Actually, same definition works for nonconvex f (however, subgradients need not exist) 4 Examples of subgradients Consider f : R ! R, f(x) = jxj 2.0 1.5 1.0 f(x) 0.5 0.0 −0.5 −2 −1 0 1 2 x • For x 6= 0, unique subgradient g = sign(x) • For x = 0, subgradient g is any element of [−1; 1] 5 n Consider f : R ! R, f(x) = kxk2 f(x) x2 x1 • For x 6= 0, unique subgradient g = x=kxk2 • For x = 0, subgradient g is any element of fz : kzk2 ≤ 1g 6 n Consider f : R ! R, f(x) = kxk1 f(x) x2 x1 • For xi 6= 0, unique ith component gi = sign(xi) • For xi = 0, ith component gi is any element of [−1; 1] 7 n
    [Show full text]
  • NP-Hardness of Deciding Convexity of Quartic Polynomials and Related Problems
    NP-hardness of Deciding Convexity of Quartic Polynomials and Related Problems Amir Ali Ahmadi, Alex Olshevsky, Pablo A. Parrilo, and John N. Tsitsiklis ∗y Abstract We show that unless P=NP, there exists no polynomial time (or even pseudo-polynomial time) algorithm that can decide whether a multivariate polynomial of degree four (or higher even degree) is globally convex. This solves a problem that has been open since 1992 when N. Z. Shor asked for the complexity of deciding convexity for quartic polynomials. We also prove that deciding strict convexity, strong convexity, quasiconvexity, and pseudoconvexity of polynomials of even degree four or higher is strongly NP-hard. By contrast, we show that quasiconvexity and pseudoconvexity of odd degree polynomials can be decided in polynomial time. 1 Introduction The role of convexity in modern day mathematical programming has proven to be remarkably fundamental, to the point that tractability of an optimization problem is nowadays assessed, more often than not, by whether or not the problem benefits from some sort of underlying convexity. In the famous words of Rockafellar [39]: \In fact the great watershed in optimization isn't between linearity and nonlinearity, but convexity and nonconvexity." But how easy is it to distinguish between convexity and nonconvexity? Can we decide in an efficient manner if a given optimization problem is convex? A class of optimization problems that allow for a rigorous study of this question from a com- putational complexity viewpoint is the class of polynomial optimization problems. These are op- timization problems where the objective is given by a polynomial function and the feasible set is described by polynomial inequalities.
    [Show full text]
  • Chapter 6. Integration §1. Integrals of Nonnegative Functions Let (X, S, Μ
    Chapter 6. Integration §1. Integrals of Nonnegative Functions Let (X, S, µ) be a measure space. We denote by L+ the set of all measurable functions from X to [0, ∞]. + Let φ be a simple function in L . Suppose f(X) = {a1, . , am}. Then φ has the Pm −1 standard representation φ = j=1 ajχEj , where Ej := f ({aj}), j = 1, . , m. We define the integral of φ with respect to µ to be m Z X φ dµ := ajµ(Ej). X j=1 Pm For c ≥ 0, we have cφ = j=1(caj)χEj . It follows that m Z X Z cφ dµ = (caj)µ(Ej) = c φ dµ. X j=1 X Pm Suppose that φ = j=1 ajχEj , where aj ≥ 0 for j = 1, . , m and E1,...,Em are m Pn mutually disjoint measurable sets such that ∪j=1Ej = X. Suppose that ψ = k=1 bkχFk , where bk ≥ 0 for k = 1, . , n and F1,...,Fn are mutually disjoint measurable sets such n that ∪k=1Fk = X. Then we have m n n m X X X X φ = ajχEj ∩Fk and ψ = bkχFk∩Ej . j=1 k=1 k=1 j=1 If φ ≤ ψ, then aj ≤ bk, provided Ej ∩ Fk 6= ∅. Consequently, m m n m n n X X X X X X ajµ(Ej) = ajµ(Ej ∩ Fk) ≤ bkµ(Ej ∩ Fk) = bkµ(Fk). j=1 j=1 k=1 j=1 k=1 k=1 In particular, if φ = ψ, then m n X X ajµ(Ej) = bkµ(Fk). j=1 k=1 In general, if φ ≤ ψ, then m n Z X X Z φ dµ = ajµ(Ej) ≤ bkµ(Fk) = ψ dµ.
    [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]