499

Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures
Page 2: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

AN EXPLANATION OF CONSTRAINED OPTIMIZATIONFOR ECONOMISTS

In a constrained optimization problem, the decisionmaker wants toselect the “optimal” choice – the one most valuable to him or her – thatmeets all of the constraints imposed by the problem. Such problemsare at the heart of modern economics, where the typical behavioralpostulate is that a decisionmaker behaves “rationally”; that is, choosesoptimally from a set of constrained choices.

Most books on constrained optimization are technical and full ofjargon that makes it hard for the inexperienced reader to gain an holis-tic understanding of the topic. Peter B. Morgan’s Explanation of Con-strained Optimization for Economists solves this problem by emphasiz-ing explanations, both written and visual, of the manner in which manyconstrained optimization problems can be solved. Suitable as a text-book or a reference for advanced undergraduate and graduate studentsfamiliar with the basics of one-variable calculus and linear algebra, thisbook is an accessible, user-friendly guide to this key concept.

PETER B. MORGAN is an associate professor in the Department ofEconomics at the University at Buffalo.

Page 3: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

This page intentionally left blank

Page 4: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

An Explanation of ConstrainedOptimization for Economists

PETER B. MORGAN

UNIVERSITY OF TORONTO PRESSToronto Buffalo London

Page 5: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

c© University of Toronto Press 2015Toronto Buffalo London

www.utpublishing.com

Printed in the U.S.A.

ISBN 978-1-4426-4278-2 (cloth)ISBN 978-1-4426-7777-4 (paper)

Printed on acid-free paper.

Library and Archives Canada Cataloguing in Publication

Morgan, Peter B., 1949–, authorAn Explanation of Constrained Optimization for Economists / Peter B. Morgan

Includes bibliographical references and index.ISBN 978-1-4426-4653-7 (bound). ISBN 978-1-4426-1446-8 (pbk.)

1. Constrained optimization. 2. Economics, Mathematical. I. Title.

QA323.M67 2015 519.6 C2015-900428-4

University of Toronto acknowledges the financial assistance to its publishing programof the Canada Council for the Arts and the Ontario Arts Council, an agency of theGovernment of Ontario.

an Ontario government agencyun organisme du gouvernement de l’Ontario

University of Toronto Press acknowledges the financial support of the Government ofCanada through the Canada Book Fund for its publishing activities.

Page 6: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

To my grandmother, Phoebe Powley, for giving up so much to care for me.

To my mother, June Powley, for fiercely emphasizing education.

To my wife, Bea, with my love.

Page 7: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

ACKNOWLEDGMENTS

Thanks are due to Professor Carl Simon for allowing my classes at the Universityat Buffalo to use a mathematical economics coursepack that he authored. The genesisof this book is in part a result of questions asked by students who sought to completetheir understanding of Professor Simon’s fine coursepack.

I sincerely thank Song Wei for pointing out various errors in an early draft ofthis manuscript. Thanks are also due to the many classes that have allowed me topractice upon them and, especially, for their insistence upon having clear explanationsprovided to them. Special thanks are due to Nicole Hunter for her tireless checkingof this manuscript and her discovery of many potential embarrassments.

Several reviewers provided helpful suggestions and criticisms. I thank each fortheir time and efforts, and for their perspectives.

Full responsibility for any remaining errors, omissions, ambiguities, eccentricities,and the like remain with the author.

Page 8: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

Contents

List of Figures xii

List of Tables xvii

1 Introduction 1

2 Basics 6

2.1 Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2.2 Real Vector Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2.3 Some Linear Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.4 Direct Products of Sets . . . . . . . . . . . . . . . . . . . . . . . . . . 20

2.5 Addition of Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

2.6 Convex Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

2.7 Partial Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

2.8 Total Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

2.9 Continuous Differentiability . . . . . . . . . . . . . . . . . . . . . . . 44

2.10 Contour Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

2.11 Marginal Rates of Substitution . . . . . . . . . . . . . . . . . . . . . 47

2.12 Gradient Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

2.13 Inner Products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

2.14 Gradients, Rates of Substitution, and Inner Products . . . . . . . . . 55

2.15 Quadratic Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

2.16 Critical Points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

2.17 Hessian Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

2.18 Quadratic Approximations of Functions . . . . . . . . . . . . . . . . . 71

2.19 Saddle-Points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

2.20 Metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

2.21 Norms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

vii

Page 9: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

viii CONTENTS

2.22 How to Proceed from Here . . . . . . . . . . . . . . . . . . . . . . . . 83

2.23 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

2.24 Answers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

3 Basics of Topology 104

3.1 Topology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

3.2 Topological Bases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

3.3 Open and Closed Sets . . . . . . . . . . . . . . . . . . . . . . . . . . 109

3.4 Metric Space Topologies . . . . . . . . . . . . . . . . . . . . . . . . . 111

3.5 Bounded Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

3.6 Compact Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

3.7 How Do Our Topological Ideas Get Used? . . . . . . . . . . . . . . . 116

3.8 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

3.9 Answers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

4 Sequences and Convergence 123

4.1 Sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

4.2 Bounded Sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

4.3 Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

4.4 Subsequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

4.5 Cauchy Sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

4.6 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

4.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

4.8 Answers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

5 Continuity 145

5.1 Basic Ideas about Mappings . . . . . . . . . . . . . . . . . . . . . . . 146

5.2 Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

5.3 Semi-Continuity and Continuity . . . . . . . . . . . . . . . . . . . . . 155

5.4 Hemi-Continuity and Continuity . . . . . . . . . . . . . . . . . . . . . 158

5.5 Upper-Hemi-Continuity by Itself . . . . . . . . . . . . . . . . . . . . . 169

5.6 Joint Continuity of a Function . . . . . . . . . . . . . . . . . . . . . . 175

5.7 Using Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176

5.8 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177

5.9 Answers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

Page 10: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

CONTENTS ix

6 Hyperplanes and Separating Sets 191

6.1 Hyperplanes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192

6.2 Separations of Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193

6.3 Separating a Point from a Set . . . . . . . . . . . . . . . . . . . . . . 197

6.4 Separating a Set from a Set . . . . . . . . . . . . . . . . . . . . . . . 203

6.5 How Do We Use What We Have Learned? . . . . . . . . . . . . . . . 203

6.6 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205

6.7 Answers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206

7 Cones 216

7.1 Cones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216

7.2 Convex Cones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218

7.3 Farkas’s Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220

7.4 Gordan’s Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224

7.5 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231

7.6 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232

7.7 Answers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233

8 Constrained Optimization, Part I 238

8.1 The Constrained Optimization Problem . . . . . . . . . . . . . . . . . 238

8.2 The Typical Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . 241

8.3 First-Order Necessary Conditions . . . . . . . . . . . . . . . . . . . . 242

8.4 What Is a Constraint Qualification? . . . . . . . . . . . . . . . . . . . 247

8.5 Fritz John’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 249

8.6 Constraint Qualifications and Farkas’s Lemma . . . . . . . . . . . . . 255

8.7 Particular Constraint Qualifications . . . . . . . . . . . . . . . . . . . 257

8.8 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265

8.9 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267

8.10 Answers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269

9 Lagrange Functions 283

9.1 What Is a Lagrange Function? . . . . . . . . . . . . . . . . . . . . . . 283

9.2 Revisiting theKarush-Kuhn-TuckerCondition . . . . . . . . . . . . . . 286

9.3 Saddle-Points for Lagrange Functions . . . . . . . . . . . . . . . . . . 287

9.4 Saddle-Points and Optimality . . . . . . . . . . . . . . . . . . . . . . 291

9.5 Concave Programming Problems . . . . . . . . . . . . . . . . . . . . 293

9.6 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297

Page 11: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

x CONTENTS

9.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 298

9.8 Answers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299

10 Constrained Optimization, Part II 302

10.1 Necessary vs. Sufficient Conditions . . . . . . . . . . . . . . . . . . . 302

10.2 Second-Order Necessary Conditions . . . . . . . . . . . . . . . . . . . 304

10.3 Geometry of the Second-Order Necessary Condition . . . . . . . . . . 311

10.4 Second-Order Necessary Condition Testing . . . . . . . . . . . . . . . 320

10.5 Sufficient Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . 326

10.6 Second-Order Sufficient Condition Testing . . . . . . . . . . . . . . . 329

10.7 Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331

10.8 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332

10.9 Answers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 334

11 Optimal Solutions and Maximum Values 341

11.1 Questions to Be Answered . . . . . . . . . . . . . . . . . . . . . . . . 341

11.2 Solution Correspondences, Maximum Values . . . . . . . . . . . . . . 342

11.3 Existence of an Optimal Solution . . . . . . . . . . . . . . . . . . . . 347

11.4 Theorem of the Maximum . . . . . . . . . . . . . . . . . . . . . . . . 351

11.5 Comparative Statics Analysis . . . . . . . . . . . . . . . . . . . . . . 358

11.6 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360

11.7 Answers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364

12 Comparative Statics Analysis, Part I 373

12.1 What Is Comparative Statics Analysis? . . . . . . . . . . . . . . . . . 373

12.2 Quantitative Comparative Statics Analysis . . . . . . . . . . . . . . . 374

12.3 Implicit Function Theorems . . . . . . . . . . . . . . . . . . . . . . . 377

12.4 Starting Points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381

12.5 Assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381

12.6 Using the First-Order Conditions . . . . . . . . . . . . . . . . . . . . 382

12.7 Using the Maximum-Value Function . . . . . . . . . . . . . . . . . . . 391

12.8 Qualitative vs. Quantitative Analyses . . . . . . . . . . . . . . . . . . 405

12.9 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 407

12.10 Answers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 413

13 Comparative Statics Analysis, Part II 431

13.1 Qualitative Comparative Statics Analysis . . . . . . . . . . . . . . . . 431

13.2 Primal and Primal-Dual Problems . . . . . . . . . . . . . . . . . . . . 432

Page 12: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

CONTENTS xi

13.3 Primal-Dual First-Order Conditions . . . . . . . . . . . . . . . . . . . 43413.4 The Primal-Dual Second-Order Condition . . . . . . . . . . . . . . . 43713.5 Qualitative Comparative Statics Results . . . . . . . . . . . . . . . . 44013.6 An Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44213.7 The Second-Order Envelope Theorem . . . . . . . . . . . . . . . . . . 44613.8 Le Chatelier’s Principle . . . . . . . . . . . . . . . . . . . . . . . . . . 45613.9 More Comparisons of Comparative Statics . . . . . . . . . . . . . . . 45913.10 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46013.11 Answers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 461

Bibliography 474

Index 475

Page 13: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

List of Figures

2.1 Vectors are points, not arrows. . . . . . . . . . . . . . . . . . . . . . . 8

2.2 Linear combinations of e1 =(10

)and e2 =

(01

)generate all of �2. . . . 11

2.3 Linear combinations of v1 =(11

)and v2 =

(−1−1

)generate only vectors(

xx

). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.4 Various linear combinations of the basis vectors v1 =(21

)and v2 =

(13

). 14

2.5 The direct products of the sets S1 = [0, 1] and S2 = [1, 3]. . . . . . . . 21

2.6 S1 + S2 = ([0, 1]× [−1, 1]) + ([1, 3]× [3, 5]) = [1, 4]× [2, 6]. . . . . . . 23

2.7 Constructing S1 +S2 = ([0, 1]× [−1, 1]) + ([1, 3]× [3, 5]) = [1, 4]× [2, 6]. 24

2.8 S1 + S2 = ([1, 3]× [3, 5]) + ([−3,−1]× [−5,−3]) = [−2, 2]× [−2, 2]. . 25

2.9 The sets S1 and S2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

2.10 Some convex and nonconvex subsets of �2. . . . . . . . . . . . . . . . 27

2.11 A weakly convex budget set, a strictly convex set of weakly preferredconsumption bundles, and a weakly convex technology set. . . . . . . 30

2.12 Six intersections of convex sets. . . . . . . . . . . . . . . . . . . . . . 31

2.13 Three direct products of convex sets. . . . . . . . . . . . . . . . . . . 32

2.14 Two plots of f(x1, x2) = x1/21 x

1/22 . . . . . . . . . . . . . . . . . . . . . 34

2.15 Plots of f(x1; x2) = x1/21 x

1/22 with x2 fixed at values of 4 and 9. . . . . 35

2.16 The slopes at x1 = 3 of f(x1, x2) = x1/21 x

1/22 in the x1-direction, when

x2 = 4, and when x2 = 9. . . . . . . . . . . . . . . . . . . . . . . . . . 37

2.17 f(x1; x2=4), f(x1; x2=4 +Δx2), f(x2; x1=3), and f(x2; x1=3 +Δx1). 39

2.18 The true difference f(3 + Δx1; 4)− f(3; 4) and its first-order approxi-mation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

2.19 Decomposing Δtruef into its two components, Δtruef1 and Δtruef2. . . 42

2.20 The first-order approximations of Δtruef1 and Δtruef2. . . . . . . . . . 43

2.21 The graph of cosine(θ) for 0◦ ≤ θ ≤ 360◦. . . . . . . . . . . . . . . . . 53

2.22 The level-36 contour set Cf (36) and the gradient vectors of f evaluatedat the points

(1, 3−

√3)and

(1, 3 +

√3). . . . . . . . . . . . . . . . 57

xii

Page 14: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

LIST OF FIGURES xiii

2.23 Gradients and marginal rates of substitution for f(x1, x2) = x1/21 x

1/22 . 60

2.24 The graph of the positive-definite quadratic form Q(x1, x2) = x21 −x1x2 + 2x22. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

2.25 The graph of the positive-semi-definite quadratic form Q(x1, x2) =x21 − 2x1x2 + x22 = (x1 − x2)

2. . . . . . . . . . . . . . . . . . . . . . . 66

2.26 The graph of the indefinite quadratic form Q(x1, x2) = x21 − x22. . . . 67

2.27 The graph of f(x) = −50x+ x3 + 100 ln (1 + x) for 0 ≤ x ≤ 4. . . . . 72

2.28 The function f and the approximating quadratic g near to x = 1·1835. 73

2.29 f(x1, x2) = x21 − x22 has a saddle-point at (x1, x2) = (0, 0). . . . . . . . 77

2.30 f(x1, x2) = (x2− y2)/ex2+y2 has a saddle-point at (x1, x2) = (0, 0), and

also two local minima and two local maxima in the set [−2, 2]× [−2, 2]. 78

2.31 Two metrics that measure the distance between the functions f and g. 81

2.32 The region containing the vectors that are linear combinations x1v1 +

x2v2 with x1 ≥ 0 and x2 ≥ 0, and the region containing the vectors

that are linear combinations y1v1 + y2v

2 with y1 ≤ 0 and y2 ≤ 0. . . . 92

2.33 {(x1, x2) | 1 ≤ x1 ≤ 3, x2 = 6}+ {(2, 3), (4, 1)}. . . . . . . . . . . . . . 94

2.34 Summing a rectangle and a triangle. . . . . . . . . . . . . . . . . . . 95

2.35 The direct products S1 × S2 and S2 × S1 for problem 2.12. . . . . . . 96

2.36 A direct product of strictly convex sets that is a strictly convex set. . 99

2.37 Diagram for Problem 2.21. . . . . . . . . . . . . . . . . . . . . . . . . 100

3.1 Some sets that are bounded with respect to E2. . . . . . . . . . . . . 114

3.2 Some sets that are not bounded with respect to E2. . . . . . . . . . . 115

4.1 The limit superior of a sequence. . . . . . . . . . . . . . . . . . . . . 131

5.1 The graphs of f and F . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

5.2 In the left-side panel, f is continuous at x = x0. In the right-side panel,f is discontinuous at x = x0. . . . . . . . . . . . . . . . . . . . . . . . 153

5.3 A graph of an upper-semi-continuous function. . . . . . . . . . . . . . 156

5.4 A graph of a lower-semi-continuous function. . . . . . . . . . . . . . . 157

5.5 f−1(2, 4) = (1, 3) = {x ∈ [0, 4] | {x + 1} ⊂ (2, 4)} = {x ∈ [0, 4] |{x+ 1} ∩ (2, 4) = ∅}. . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

5.6 F−1((2, 4))+ = {x ∈ [0, 4] | [x + 1, x + 2] ⊂ (2, 4)} = (1, 2), butF−1((2, 4))− = {x ∈ [0, 4] | [x+ 1, x+ 2] ∩ (2, 4) = ∅} = (0, 3). . . . . 161

5.7 Upper-hemi-continuity at x = x0. . . . . . . . . . . . . . . . . . . . . 162

5.8 Lower-hemi-continuity at x = x0. . . . . . . . . . . . . . . . . . . . . 164

5.9 Upper-hemi-continuity at the point x = x0. . . . . . . . . . . . . . . . 171

Page 15: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

xiv LIST OF FIGURES

5.10 The graph for Problem 5.6. . . . . . . . . . . . . . . . . . . . . . . . 178

6.1 A hyperplane separating two sets. . . . . . . . . . . . . . . . . . . . . 192

6.2 Separated sets. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194

6.3 Strictly separated sets. . . . . . . . . . . . . . . . . . . . . . . . . . . 195

6.4 Part (i) of Theorem 6.1. . . . . . . . . . . . . . . . . . . . . . . . . . 198

6.5 Part (ii) of Theorem 6.1. . . . . . . . . . . . . . . . . . . . . . . . . . 199

6.6 Case (ii), with x0 a boundary point of the open set S. . . . . . . . . . 201

6.7 The budget constraint is a hyperplane that separates the budget setS1 from the set S2 of consumption bundles that are weakly preferredto the bundle (20/3, 10). . . . . . . . . . . . . . . . . . . . . . . . . . 207

6.8 The hyperplane’s normal p and the net trade vector z are orthogonal. 208

6.9 The budget constraint is a hyperplane that separates the budget setfrom the set of consumption bundles that are weakly preferred to therationally chosen bundle (16/3, 8, 8). . . . . . . . . . . . . . . . . . . 209

6.10 The net trade vector z and the normal p of the separating (budgetconstraint) hyperplane are orthogonal. . . . . . . . . . . . . . . . . . 210

6.11 Finding the point (x′1, x′2) in X that is closest to the point x0 = (8, 2). 211

7.1 A cone and a convex cone. . . . . . . . . . . . . . . . . . . . . . . . . 217

7.2 A cone and an affine cone. . . . . . . . . . . . . . . . . . . . . . . . . 218

7.3 Farkas’s two alternatives. . . . . . . . . . . . . . . . . . . . . . . . . . 221

7.4 Farkas’s second alternative. . . . . . . . . . . . . . . . . . . . . . . . 222

7.5 Farkas’s second alternative is that a hyperplane through the originseparates the singleton set {b} from the convex cone generated by a1

and a2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223

7.6 Gordan’s two alternatives. . . . . . . . . . . . . . . . . . . . . . . . . 225

7.7 A different perspective of Gordan’s two alternatives. . . . . . . . . . . 226

7.8 Where is the set V ? . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228

7.9 The graphs for problem 7.1. . . . . . . . . . . . . . . . . . . . . . . . 234

7.10 Diagrams for problem 7.5. . . . . . . . . . . . . . . . . . . . . . . . . 237

8.1 x∗ is a local maximum for f on C. . . . . . . . . . . . . . . . . . . . . 243

8.2 x∗ is not a local maximum for f on C. . . . . . . . . . . . . . . . . . . 244

8.3 The feasible set C has a cusp at x = x∗. . . . . . . . . . . . . . . . . . 248

8.4 The optimal solution x∗ = (2, 0) is a regular point. . . . . . . . . . . . 251

8.5 The feasible set has a cusp at the optimal solution x∗ = (2, 0). . . . . 253

Page 16: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

LIST OF FIGURES xv

8.6 The feasible set has a cusp at the optimal solution. Farkas’s secondalternative is true. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256

8.7 The Slater/Karlin Constraint Qualification. . . . . . . . . . . . . . . 258

8.8 The affine closed convex cone x′ + CD(x′) for the regular point x = x′. 260

8.9 The affine closed convex cone x′ + CD(x′) for the cusp x = x′. . . . . 261

8.10 h1 and h2 are constrained paths from x = x′. h3 is not. . . . . . . . . 262

8.11 The Kuhn-Tucker Constraint Qualification is satisfied at x = x′. . . . 263

8.12 The Kuhn-Tucker Constraint Qualification is not satisfied at x = x′. . 264

8.13 The feasible set, the optimal solution, and the gradient vectors forproblem 8.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270

8.14 The feasible set, the optimal solution, and the gradient vectors forproblem 8.2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272

8.15 The feasible set, the optimal solution, and the gradient vectors forproblem 8.3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274

8.16 Problem 8.5’s feasible set and two objective function contour sets. . . 277

8.17 The gradients of f , g1 and g2 at (x1, x2) = (1·177, 3·823) and at(x1, x2) = (3·823, 1·177). . . . . . . . . . . . . . . . . . . . . . . . . . 278

8.18 The singleton feasible set C and the gradient vectors for problem 8.6. 279

8.19 The feasible set, the optimal solution, and the gradient vectors forproblem 8.7. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281

9.1 The Lagrange function L(x, λ2) = 4√x+λ2(2−x) has a global x-max,

λ-min saddle-point at (x, λ2) = (2,√2 ). . . . . . . . . . . . . . . . . . 289

9.2 The Lagrange function L(x, λ2) = x2 + λ2(2 − x) has a global x-min,λ-max saddle-point at (x, λ2) = (2, 4). . . . . . . . . . . . . . . . . . . 290

10.1 The first-order and second-order necessary conditions that are impliedby f being maximized at x = (1, 1). . . . . . . . . . . . . . . . . . . . 306

10.2 f satisfies the first-order necessary condition but not the second-ordernecessary condition at x = (1, 1), which is therefore not a local maxi-mum for f . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307

10.3 The second-order necessary condition for problem (10.28). . . . . . . 312

10.4 The second-order necessary condition for problem (10.34). . . . . . . 314

10.5 The second-order necessary condition for problem (10.39). . . . . . . 317

10.6 The only perturbation that keeps both constraints binding is (0, 0). . 319

Page 17: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

xvi LIST OF FIGURES

11.1 On the compact interval [a, b], the upper-semi-continuous function fattains a global maximum at x = x∗, attains a local minimum, anddoes not attain a global minimum. . . . . . . . . . . . . . . . . . . . 348

11.2 On the compact interval [a, b], the lower-semi-continuous function fattains a global minimum at x = x∗, attains a local maximum, anddoes not attain a global maximum. . . . . . . . . . . . . . . . . . . . 349

11.3 On the compact set [a, b], the continuous function f(x) attains a globalmaximum at x = x∗, and attains a global minimum at both x = x∗and x = x∗∗. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 350

11.4 The ordinary demand correspondenceX∗(p1, y) is upper-hemi-continuouswith respect to p1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 352

11.5 X∗(γ) is upper-hemi-continuous at γ = γ0. . . . . . . . . . . . . . . . 35611.6 The unit simplices of dimensions one, two, and three, and some exam-

ples of mixed actions. . . . . . . . . . . . . . . . . . . . . . . . . . . . 36211.7 Graph of f(x) for problem 11.1. . . . . . . . . . . . . . . . . . . . . . 36511.8 The entrant’s and incumbent’s best-response correspondences. . . . . 37011.9 The entry game’s best-response correspondence. . . . . . . . . . . . . 372

12.1 The graph of the implicit optimal solution x∗(γ) to problem (12.4). . 37612.2 The restricted functions vr(γ; x′), vr(γ; x′′), and vr(γ; x′′′). . . . . . . 39912.3 The upper envelope v(γ) generated by the restricted functions vr(γ; x′),

vr(γ; x′′), and vr(γ; x′′′). . . . . . . . . . . . . . . . . . . . . . . . . . 40112.4 The upper envelope v(γ) generated by the family of functions vr(γ; x). 40212.5 Hotelling’s result that ∂π∗(p, w)/∂w1 = −x∗1(p, w). . . . . . . . . . . . 40512.6 The upper envelope for problem 12.4. . . . . . . . . . . . . . . . . . . 41812.7 The lower envelope for problem 12.5. . . . . . . . . . . . . . . . . . . 41912.8 Hotelling’s result that ∂π∗(p, ·)/∂p = y∗(p, ·). . . . . . . . . . . . . . . 42512.9 Shephard’s result that ∂cl(w1, w2, y)/∂w1 = xl1(w1, w2, y). . . . . . . . 427

13.1 The Second-Order Envelope Theorem. . . . . . . . . . . . . . . . . . 452

Page 18: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

List of Tables

2.1 True and first-order approximate values of f(3 + Δx1; x2 = 4)− f(3, 4). 412.2 Various evaluations of f and Q about (x1, x2) = (1, 1). . . . . . . . . 762.3 The values of f and Q at the requested points. . . . . . . . . . . . . . 103

12.1 x∗(γ) and its order 2 Taylor’s series approximation. . . . . . . . . . . 377

xvii

Page 19: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

This page intentionally left blank

Page 20: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

Chapter 1

Introduction

Hi there. Let’s start by stating what this book is not. It is not a book aboutmathematics. It is instead a book about economics. Particularly, it is a book abouteconomics spoken using the formal language of mathematics. What is the point? Whynot just speak about economics in a language such as English or Spanish and maybedraw some pictures to help with a verbal analysis of the economic issue that interestsyou? The problem is the lack of precision inherent in a verbal argument. Mathematicsis a precise language. Of course this precision comes with the cost that every littledetail must be treated precisely. This is usually what drives economics students crazyand sometimes scares them. Then, on top of this agony there is the language andsymbolism of mathematics itself. Who wants to worry about the meanings of squigglessuch as ∇, ⊗, �, and �? Actually, you should. These squiggles are nothing morethan convenient short forms of longer phrases. ∇, for example, denotes an importantsomething called a gradient vector. Surely it is easier to write down ∇ than it is towrite the words “gradient vector,” especially if you have to do this repeatedly. Sohere is the good news – the language is easy to learn and the ideas expressed by itare simple, practical, and usually rather obvious. There is nothing to fear. You andI have a lot to talk over and I think you will find it both valuable and fun – yes,fun! More than that, you will be able to protect yourself from those who will seekto challenge your arguments. Oftentimes an economist will have a strong intuitivehunch about the solution to a problem, but that is not enough. I should know – toooften I have had a “strong intuition” about the answer to a problem only to discovereventually that I was wrong. Intuition is helpful, but nothing more than that. Aprecise answer is the outcome of a logical analysis, and there are few better ways tocommunicate logically than to use the language we call mathematics.

1

Page 21: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2 INTRODUCTION

You may be one of the walking wounded who have studied mathematics or mathe-matical economics and have found it to be confusing. Perhaps you have found yourselfwondering if you are not as clever as you hoped. Perhaps you have been reduced tomindless memorization of mathematical arguments, having given up on ever properlyunderstanding them. I know what that is like. I also know that memorization isuseless since it teaches no understanding, and, in the end, understanding is all thatmatters. So I want you to know at the outset that I am sympathetic to your needsand that we are going to work together to get it all straight.

Where is the right place to start? I’m sure that you already have quite a substan-tial knowledge of the basic ideas, such as sets, derivatives, vectors, and things likemaxima and minima. The trouble is that quite often students are taught the meaningof important ideas in only special contexts. Since this is all they know, trouble ariseswhenever the student is asked to use an idea outside the special context in which itwas “taught.” Here is an example. Let me ask you, “What is an open set?” NowI’m serious about this. Stop reading. Pick up a pencil and on a piece of paper writedown your answer(s) to this question. Go on. Go no further until you have done this.

Done it?

Maybe you wrote something like, “A set is open if and only if its complement isa closed set.” If you did so then you are correct, but now I must ask you, “What isa closed set?” Don’t tell me that a set is closed if and only if its complement is anopen set, because then you have just gone around in a circle and said nothing useful.

Perhaps instead you wrote words like, “A set is open if and only if it consistsonly of interior points” or “A set is open if and only if it does not contain any of itsboundary points.” Of course, I now have to ask you what “interior” or “boundary”means. More important, even if I did, you would still be thinking about an openset in a rather special way. Now that’s OK, in that special context. The trouble is,economists sometimes have to think about problems outside that context, and thenyou won’t have a clue about the true, general meaning of “open set.”

I don’t want to worry you. My point is that your understanding of basic ideasmay have been distorted by presentations of special cases of them that suggested toyou that these special instances were more general. This can lead to a lot of confusionand distress. I would like my sometimes unpleasant experiences to help you to avoidthat suffering. We are going to be alert to these possibilities as needed, so pleasedon’t fret. These issues will get sorted out one at a time as and when we need to sortthem out.

So here’s the plan. We economists constantly talk about a “rational choice”or an “optimal solution.” When we do so we have in mind a problem in which a

Page 22: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

INTRODUCTION 3

decisionmaker wants as much payoff (utility, profit, or whatever) as is possible. Oftenthe decisionmaker’s choices are restricted by one or more constraints (a budget ora technology, for instance). This book is aimed at making sense of this rationalchoice idea that is at the very heart of modern economics. By making sense I meanconstructing a completely logical and precise understanding of the idea. If you haveless than this, then how can you call yourself a properly educated economist? I like tovisualize ideas. I spend a lot of time drawing pictures that display to me the essence ofan economics problem. I use the picture to help me to write down the mathematicaldescription of the problem. I work out the solution and then (this is important) I goback to the picture and see what the mathematics has told me. I need to convincemyself that the information given to me by the mathematics is completely consistentwith what the picture shows me. I need to see the solution and to understand it in avery basic and elementary way that lets me explain it with complete clarity to myselfand to anyone else interested in my discovery. You should have the same goals. Halfan understanding is a dangerous outcome.

I hope that you can now see why I say that this is not a book about mathematics.We will talk about economics in this book, and we will do so with coherent logic andprecision. We will use mathematics, draw lots of pictures, and make sense of all ofit as a cohesive whole, not as a bunch of disjointed bits and pieces. Trust me, this isgoing to be a good experience. I will need something from you, however, and that isa commitment from you to think. Understanding comes from patient contemplationand logical reasoning – these are your tools.

How should you use this book? That depends upon who you are. The corechapters are Chapters 8 and 10, which explain in some detail the ways in which wecan solve rational choice problems, also known as constrained optimization problems.These two chapters make extensive use of ideas that are developed in Chapters 5, 6,and 7. These chapters in turn rely upon ideas developed in Chapters 2, 3, and 4. Soif you already have a good understanding of the contents of Chapters 2, 3, 4, 5, 6,and 7, then you should proceed directly to Chapters 8 and 10. Otherwise, use thelisting of chapter contents to decide when and how to use parts of the earlier chaptersso that you can understand Chapters 8 and 10. No one book can hope to presenteverything you need to know about mathematics as it is applied to economics, soyou might need also to have accessible to you another book that presents more basicmaterials. Three, of many, widely used such books are Simon and Blume 1994, Hoyet al. 2011, and Chiang and Wainwright 2005.

I did not intend to write this book. It “just happened.” The blame lies with themany students who pressured me for clear explanations of mathematical ideas that

Page 23: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

4 INTRODUCTION

their studies forced them to examine. My first reaction was to look for accessible, nottoo technical, accurate explanatory references that would meet their needs. I soonfound that for the most part these references did not exist. The best available wasa fine coursepack prepared by Professor Carl Simon at the University of Michigan.Professor Simon graciously allowed me to use his coursepack to instruct studentstrying to complete their Ph.D. in Economics degrees at the University at Buffalo.Even so, the requests did not cease. I found myself spending many long eveningspreparing handouts on one topic after another. After several years of doing this Ihad unintentionally written about two-thirds of a book. For several years I resistedcompleting the task. I noticed that as time passed ever more blurry photocopies of myhandouts were being passed from older to newer cohorts of students. There appearedrequests for more materials, updates, and so on. So I finally decided to finish the job.I suspect that many books get written this way. In the end, the deciding factor isthat the materials are wasted if they are not made accessible. Here it is, at last. Ihope it is of value to you.

This book is not overly concerned with statements of theorems and their detailedproofs. Yes, there are theorems and there are proofs. But the emphasis of thisbook is on explanation. The question “Why?” is not a request for a proof. It is arequest for understanding, and conveying understanding often requires a less thancompletely rigorous story so that small details do not distract from the essentials ofan argument. I have tried always to remember this when writing. Where detailedproofs are provided, I have tried to explain their arguments.

I have found over the years that I can usually divide a group of students intotwo distinct sets. One set I usually label as the “techies,” meaning the studentswho concentrate upon mathematics and find satisfaction in deriving formulae andmemorizing proofs. But ask a typical techie to draw a picture illustrating the economiccontent of his or her work, or ask for an economist’s explanation of what he or shehas done, and you usually get a look back that suggests the student is wonderingif you are insane. “Why would I ever want to do that?” seems to be the impliedquestion. The answer is that economics merely uses mathematics, where appropriate,to understand more about economics. Mathematics itself is never the goal. The otherset I think of as the “hand wavers.” These students will happily draw a diagram,usually a little vaguely, and mutter in an imprecise way about what the solution to aproblem “should look like.” Asking for precise logic seems for many of these studentsto be a cruel and unreasonable request. So it will not surprise you to learn that mostof my classes start out disgruntled when they learn that I insist both upon logic andupon an economist’s understanding of what the logic yields. Usually, by about the

Page 24: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

INTRODUCTION 5

middle of the semester, both groups have improved upon the deficient skill and theclass becomes more homogeneous and better humored.

Let’s get started!

Page 25: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

Chapter 2

Basics

This book is not intended to be a substitute for the many others that already explainwell the basics of the calculus and of linear algebra. But for your convenience thischapter provides a review of the mathematical ideas that are used in various placesin the following chapters. If you are familiar with all of these ideas, then skip thischapter. Otherwise, select from within this chapter the appropriate reading. Thediscussions herein are less complete than in books that explain in detail the mathe-matical methods typically used by economists because the discussions provided hereare tailored only to their uses in later parts of this book.

2.1 Space

Mathematicians are always talking about “spaces.” They talk of metric spaces, vectorspaces, and all sorts of such stuff, and we too will often use the word “space,” so whatdoes it mean? A space is a set. Not just any old set though. A space is a set thatcontains elements that all satisfy the same collection of properties. For example, wemight be considering a set of elements all of which have the property that we canmeasure the distance between any pair of these elements. Such a set is called a metricspace (the word metric means that distances between points are measurable). Or itmight be that each element is a real number, and so we call such a set a real numberspace (a name that usually is abbreviated to real space). You will get the idea as wego along, so don’t be alarmed by the word “space.”

6

Page 26: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.2. REAL VECTOR SPACE 7

2.2 Real Vector Space

A real vector space is a set that is a space for two reasons. First, the set containsonly real numbers (hence, a real space). Second, the real numbers in a real vectorspace form ordered lists of numbers that are called vectors because they possess aparticular common collection of arithmetic properties. What are these properties?What is a vector?

First, think of an integer n ≥ 1 that is the common length of every list of realnumbers that we will consider. Symbols like xi and yi will denote the real numbers inthe ith positions in these lists, for i = 1, . . . , n. Consider, for example, the (vertically)ordered lists of real numbers

x =

⎛⎜⎜⎝

1−70

6·55

⎞⎟⎟⎠ and y =

⎛⎝ −316/79/2

⎞⎠ .

The number of positions in a list is the dimension n of the list. So the list x hasdimension n = 4 and the elements of this list are x1 = 1, x2 = −7, x3 = 0, andx4 = 6·55 (real numbers don’t have to be integers, right?). The list y has dimensionn = 3 and the elements of this list are y1 = −3, y2 = 16/7, and y3 = 9/2.

Now let’s insist that every ordered list of real numbers has the same dimension n,so that typical lists are written as

x =

⎛⎜⎜⎜⎜⎜⎝

x1x2...

xn−1

xn

⎞⎟⎟⎟⎟⎟⎠ , y =

⎛⎜⎜⎜⎜⎜⎝

y1y2...

yn−1

yn

⎞⎟⎟⎟⎟⎟⎠ , or z =

⎛⎜⎜⎜⎜⎜⎝

z1z2...

zn−1

zn

⎞⎟⎟⎟⎟⎟⎠ .

A set X of such n-dimensional ordered lists of real numbers is a set of n-dimensionalreal vectors, an n-dimensional real vector space, when all of the lists in X possess avariety of obvious arithmetic properties.

Definition 2.1 (Real Vector Space). A set X of n-dimensional ordered lists of realnumbers is an n-dimensional real vector space if and only if:

1. Addition. x, y ∈ X implies that z = x + y ∈ X. So, for example, if n = 3 and(1−46

)∈ X and

(20−2

)∈ X, then

(3−44

)∈ X also.

Page 27: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8 CHAPTER 2. BASICS

2. Commutativity. For every x, y ∈ X, x + y = y + x; i.e. the order of additiondoes not matter.

3. Associativity. x, y, z ∈ X implies that (x + y) + z = x + (y + z); i.e. how youpair up vectors when adding them does not matter.

4. Zero Vector. There is an ordered list 0 ∈ X such that for every x ∈ X, x+0 =0 + x = x. The zero vector is the n-dimensional list in which every element is

the number zero; e.g. 0 =(

000

)is the three-dimensional zero vector.

5. Inverse Vector. For each x ∈ X, there is −x ∈ X such that x + (−x) =

(−x) + x = 0; e.g. if x =(

1−46

)∈ X, then −x =

( −14−6

)∈ X and x + (−x) =(

1−46

)+( −1

4−6

)=(

000

). −x is the inverse of x, and x is the inverse of −x.

6. Scalar Multiplication. Given all of the above properties, if λ and μ are scalarreal numbers, then X also possesses the property of scalar multiplication if

(a) for every x, y ∈ X, λ(x+ y) = λx+ λy,

(b) for every x ∈ X, (λ+ μ)x = λx+ μx, and

(c) for every x ∈ X, (λμ)x = λ(μx).

Figure 2.1: Vectors are points, not arrows.

The left-side panel of Figure 2.1 displays the two-dimensional vectors x = ( 41 ),

y = ( −11 ), and z =

( −2−3

). The points displayed in the panel are the vectors. In the

right-side panel of the figure you again see the three vectors along with an arrow from

Page 28: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.3. SOME LINEAR ALGEBRA 9

the origin to each point. A common error for beginners is to believe that these arrowsare the vectors, but they are not. The arrows depict how distant each vector (point)is from the origin, and the directions of the vectors from the origin. Sometimes, then,it is visually more informative to use an arrow to indicate a vector than it is to drawthe (dot) vector itself. This will often be done in later chapters.

2.3 Some Linear Algebra

Several important results to do with constrained optimization amount to asking if aparticular set of linear equations has a particular type of solution. So let us think ofa simple pair of linear equations;

2x1 + x2 = 5

x2 + 4x2 = 13.(2.1)

The unique solution is x1 = 1 and x2 = 3. Now look at the vector equation

x1

(21

)+ x2

(14

)=

(513

). (2.2)

(2.2) and (2.1) are equivalent and have the same solution. They are equivalent because

x1

(21

)+ x2

(14

)=

(2x1x1

)+

(x24x2

)(scalar multiplication property of vectors)

=

(2x1 + x2x1 + 4x2

)(addition property of vectors)

=

(513

).

Any set of linear simultaneous equations like (2.1) is the same as a vector equationlike (2.2). That is, a system of m linear equations in n unknowns x1, . . . , xn

a11x1

a21x1...

am1x1

+

+

+

a12x2

a22x2...

am2x2

+

+

+

· · ·· · ·

· · ·

+

+

+

a1nxn

a2nxn...

amnxn

=

=

=

b1

b2...

bm

(2.3)

Page 29: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10 CHAPTER 2. BASICS

is equivalent to the vector equation

x1

⎛⎜⎜⎜⎝a11a21...am1

⎞⎟⎟⎟⎠+ x2

⎛⎜⎜⎜⎝a12a22...am2

⎞⎟⎟⎟⎠+ · · ·+ xn

⎛⎜⎜⎜⎝a1na2n...

amn

⎞⎟⎟⎟⎠ =

⎛⎜⎜⎜⎝b1b2...bm

⎞⎟⎟⎟⎠ . (2.4)

The left side of (2.2) is a weighted sum of the vectors ( 21 ) and ( 1

4 ). The weights arex1 and x2. A weighted sum of vectors is called a linear combination of the vectors.Thus the left side of (2.2) is a linear combination of ( 2

1 ) and ( 14 ), and this linear

combination creates, or generates, the vector ( 513 ) when the weights used in the linear

combination are x1 = 1 and x2 = 3. This means that the equations (2.1) have asolution if and only if there are values for the weights x1 and x2 that create a linearcombination of ( 2

1 ) and ( 14 ) that generates (

513 ). This is true in general. That is,

the simultaneous linear equation system (2.3) has a solution if and only ifthere exist weights x1, . . . , xn satisfying (2.4).

This is why we should now spend some time thinking about when vectors can be, orcannot be, generated as linear combinations of other vectors.

Take a given set V of vectors v and let L(V ) be the set of all of the vectors thatcan be created (generated) by linear combinations of the vectors v ∈ V . A basicexample is the set

V = {e1, e2}, where e1 =(10

)and e2 =

(01

).

Think of any two-dimensional vector at all, say, ( x1x2 ). This vector is generated from

e1 and e2 by the linear combination

x1e1 + x2e

2 = x1

(10

)+ x2

(01

)=

(x1x2

),

so {e1, e2} is a set of vectors that, by linear combination, can generate every vectorin the two-dimensional real vector space L({e1, e2}) = �2 (see Figure 2.2).

For another example consider the set

V = {v1, v2}, where v1 =(11

)and v2 =

(−1−1

).

Page 30: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.3. SOME LINEAR ALGEBRA 11

Figure 2.2: Linear combinations of e1 =(10

)and e2 =

(01

)generate all of �2.

A linear combination of v1 and v2 is a vector

v = x1v1 + x2v

2 = x1

(11

)+ x2

(−1−1

)=

(x1 − x2x1 − x2

)consisting of the same element in both positions. All such vectors are contained in the45◦ straight line L(V ) through the origin that contains both v1 and v2; see Figure 2.3.This straight line is only a tiny part of �2. A vector not in the line L(V ), such as ( 3

2 )for example, cannot be generated by v1 and v2. Equivalently, then, the simultaneouslinear equation system

1× x1 + (−1)× x2 = 3

1× x1 + (−1)× x2 = 2(2.5)

does not have a solution.The set L(V ) of vectors generated by linear combinations of the vectors in V is

called the set of vectors spanned by V and the set V is called a basis for L(V ). In ourexamples, {( 1

0 ) , (01 )} spans all of �2 and is a basis for �2, and

{( 11 ) ,( −1−1

)}spans

the 45◦ line {(x, x) | −∞ < x <∞} and is a basis for this line.Why does the set {e1, e2} generate all of �2 while the set {v1, v2} generates only

the 45◦ line in Figure 2.3? The answer is that neither of the vectors e1 and e2 can bewritten as a multiple of the other; i.e. there is no number λ such that ( 1

0 ) = λ ( 01 ).

Page 31: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12 CHAPTER 2. BASICS

Figure 2.3: Linear combinations of v1 =(11

)and v2 =

(−1−1

)generate only vectors(

xx

).

We say that the vectors e1 and e2 are linearly independent. In contrast, the vectorsv1 and v2 are multiples of each other; i.e. ( 1

1 ) = (−1)( −1−1

). Such vectors v1 and v2

are linearly dependent. To make further progress we need to formally define the ideasof linear independence and linear dependence.

Definition 2.2 (Linear Dependence, Linear Independence). A set of n-dimensionalvectors is linearly independent if and only if none of the vectors can be generated as alinear combination of other vectors in the set. The set of vectors is linearly dependentif and only if the set is not linearly independent.

A spanned vector space often has more than one basis. For example, {( 11 ) , (

−10 )}

is another basis for �2. The result that tells us this is called the Spanning Theorem.It is one of the most basic results in linear algebra.

Theorem 2.1 (Spanning Theorem). Any set of n linearly independent real vectorsof dimension n is a basis for the real vector space �n.

An Exercise. Show that the set {( 11 ) , (

−10 )} is a basis for �2.

Answer. We need to show that any vector in �2, say,(b1b2

), can be generated as a

linear combination of the vectors ( 11 ) and ( −1

0 ). That is, we need to show that, for

Page 32: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.3. SOME LINEAR ALGEBRA 13

any arbitrary but given values b1 and b2, there exist values x1 and x2 such that

x1

(11

)+ x2

(−10

)=

(b1b2

).

It is easy to see that this is true when

x1 = b2 and x2 = b2−b1.

An immediate and important implication of the Spanning Theorem is

Theorem 2.2. Any set of m real vectors of dimension n is a linearly dependent setif m > n.

Another Exercise. Use the Spanning Theorem to prove Theorem 2.2.

Answer. Consider any subset V of n of the m vectors. Either this subset is a linearlydependent set or it is not. If the subset V is linearly dependent, then so is the set of allm vectors. If the subset V is linearly independent, then, by Theorem 2.1, it is a basisfor �n and so each of the other m− n vectors may be written as linear combinationsof the vectors in V , making the set of all m vectors linearly dependent.

One of the most famous of the theorems to do with solving constrained opti-mization problems asks if there exists a nonzero and nonnegative solution to someparticular linear equations. For now we don’t need to worry about why this matters,but this is the right place to see what such a solution looks like. So have a lookat Figure 2.4. The figure displays two linearly independent vectors: v1 = ( 2

1 ) andv2 = ( 1

3 ). The set consisting of v1 and v2 is a basis for all of �2, so we can write anyvector

(b1b2

)as a unique linear combination of v1 and v2. That is, for any given values

of b1 and b2, there always exist unique values for x1 and x2, solving

x1

(21

)+ x2

(13

)=

(b1b2

).

The question is “Are the values of x1 and x2 positive or negative?” The answerdepends upon the values of b1 and b2. For example, if (b1, b2) = (3, 4), then (x1, x2) =(1, 1); if (b1, b2) = (3,−1), then (x1, x2) = (2,−1, ); and if (b1, b2) = (−5,−5), then(x1, x2) = (−2,−1). The dotted line in Figure 2.4 that contains both the origin andv1 is the line containing all multiples (positive, zero, and negative) x1v

1 of v1. Thedotted line containing both the origin and v2 is the line containing all multiples x2v

2

of v2. These two dotted lines partition �2 into four parts, labeled Regions 1, 2, 3,

Page 33: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

14 CHAPTER 2. BASICS

Figure 2.4: Various linear combinations of the basis vectors v1 =(21

)and v2 =

(13

).

and 4 in Figure 2.4. The figure displays a vector in Region 2 that is created by a linearcombination x1v

1+x2v2 in which x1 < 0 and x2 > 0. It will turn out that Region 1 is

of great interest to us. This is the region consisting of vectors that are generated bypositive (both x1 > 0 and x2 > 0) linear combinations of the two linearly independentvectors v1 and v2. Such regions are called convex cones and are crucial in computingsolutions to constrained optimization problems.

It is time to introduce a new, more brief notation. Let us write the list of vectorsused in (2.4) as the matrix A, the list of linear combination weights as the columnvector x, and the list of right-side values as the column vector b:

A ≡

⎛⎜⎜⎜⎝a11 a12 · · · a1na21 a22 · · · a2n...

......

...am1 am2 · · · amn

⎞⎟⎟⎟⎠ , x ≡

⎛⎜⎜⎜⎝x1x2...xn

⎞⎟⎟⎟⎠ , and b ≡

⎛⎜⎜⎜⎝b1b2...bn

⎞⎟⎟⎟⎠ .

The matrix A is just a space-saving way of writing down an ordered list of columnvectors all of the same dimension. Because A has m rows and n columns, we call Aan m× n matrix (always write the number of rows first and the number of columns

Page 34: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.3. SOME LINEAR ALGEBRA 15

second). Next, again for brevity, we write

Ax ≡

⎛⎜⎜⎜⎝a11 a12 · · · a1na21 a22 · · · a2n...

......

...am1 am2 · · · amn

⎞⎟⎟⎟⎠⎛⎜⎜⎜⎝x1x2...xn

⎞⎟⎟⎟⎠

≡ x1

⎛⎜⎜⎜⎝a11a21...am1

⎞⎟⎟⎟⎠+ x2

⎛⎜⎜⎜⎝a12a22...am2

⎞⎟⎟⎟⎠+ · · ·+ xn

⎛⎜⎜⎜⎝a1na2n...

amn

⎞⎟⎟⎟⎠ =

⎛⎜⎜⎜⎝b1b2...bm

⎞⎟⎟⎟⎠ = b.

(2.6)

Statement (2.6) defines what is meant by the symbol Ax, called the multiplicationof an m × n matrix A and an n-dimensional vector x. Such a multiplication is aconvenient brief way of writing down a linear combination of the column vectors thatmake up the matrix A. Ax = b is the same statement as (2.3) and (2.4). It says thatthe vector b can be written as a linear combination of the vectors that make up thecolumns of the matrix A and that, therefore, b belongs to the set of vectors that arespanned by the set of vectors that are the columns of the matrix A. We have arrivedat another important result in linear algebra.

Theorem 2.3. The simultaneous linear equation system Ax = b has at least onesolution if and only if b is a member of the set of vectors that is spanned by thecolumn vectors of the matrix A.

An Exercise. If the simultaneous linear equation system Ax = b has a solution,then must the set of vectors⎛

⎜⎜⎜⎝a11a21...am1

⎞⎟⎟⎟⎠ ,

⎛⎜⎜⎜⎝a12a22...am2

⎞⎟⎟⎟⎠ , · · · ,

⎛⎜⎜⎜⎝a1na2n...

amn

⎞⎟⎟⎟⎠ ,

⎛⎜⎜⎜⎝b1b2...bm

⎞⎟⎟⎟⎠ (2.7)

(i) be linearly dependent?(ii) If this set of vectors is linearly dependent, then must Ax = b have a solution?

Page 35: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

16 CHAPTER 2. BASICS

Answer.

(i) Yes. If Ax = b, then b is linearly dependent upon the column vectors making upthe matrix A.

(ii) No. (2.5) is a counterexample. Ax = b has a solution if and only if the vectorb is linearly dependent upon the column vectors of A. The set (2.7) can be linearlydependent because of linear dependence among the column vectors of A irrespectiveof whether or not b is linearly dependent upon the column vectors of A.

Now let’s think about how many solutions there can be to an equation systemsuch as (2.1). Reconsider our earlier examples of

2x1 + x2 = 5

x2 + 4x2 = 13and x1 − x2 = 3.

The first of these possesses just one solution, (x1, x2) = (1, 3). The second possessesinfinitely many solutions (x1, x2) = (x1, x1 − 3), one for each value of x1 ∈ �. Whatcauses these differences? Probably you are saying to yourself something like “The firstexample has two independent equations in two unknowns while the second examplehas only one equation in two unknowns.” I agree. But think about this from anotherpoint of view. In the first example, the left sides of the two equations are the linearcombination

Ax =

(2 11 4

)(x1x2

)= x1

(21

)+ x2

(14

)of two linearly independent vectors of dimension 2; neither ( 2

1 ) nor (14 ) is a multiple

of the other. This is actually what you mean when you say “independent equations.”In the second example, the left side of the equation is the linear combination

Bx =(1 −1

)(x1x2

)= x1(1) + x2(−1)

of two linearly dependent vectors of dimension 1; (1) and (−1) are multiples of eachother. Observe that the row vectors (2, 1) and (1, 4) in the matrix A are linearlyindependent of each other since neither is a multiple of the other, and that there isjust a single row vector (1,−1) in the matrix B. The number of linearly independentrow vectors in a matrix is called the rank of the matrix, so the rank of A is r(A) = 2and the rank of B is r(B) = 1. You can think of the rank of a matrix as the numberof “linearly independent equations” implied by the matrix.

Page 36: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.3. SOME LINEAR ALGEBRA 17

An Exercise. What are the ranks of the matrices

A =

⎛⎝2 1 10 −1 25 3 0

⎞⎠ , B =

⎛⎝2 10 −15 3

⎞⎠ , and C =

(2 1 51 4 13

)?

Answer. r(A) = 3 because none of the row vectors (2, 1, 1), (0,−1, 2) or (5, 3, 0) canbe generated as a linear combination of the other two row vectors. r(B) = 2 becauseany one of the row vectors (2, 1), (0,−1), or (5, 3) can be generated as a linearcombination of the other two. For example, (5, 3) = 5

2(2, 1) − 1

2(0,−1). r(C) = 2

because neither of the row vectors (2, 1, 5) and (1, 4, 13) is a multiple of the other.

Simply put, if the linear simultaneous equation system Ax = b displayed in (2.6)has a solution, then it has exactly one solution if the rank of the matrix A is thesame as the number of unknowns, n. This amounts to saying that the equationsystem has the same number of linearly independent equations as there are unknowns.Fortunately for us, in most applications of these ideas to constrained optimizationproblems the number of equations, m, and the number of unknowns, n, is the same.When m = n we can make simpler statements about when a linear simultaneousequation system Ax = b has a solution, has a unique solution, or has more than onesolution. In such cases the number of rows and the number of columns of the matrixA are both n, making the matrix A an n× n matrix. We call such a matrix a squarematrix of order n.

Theorem 2.4. If the linear simultaneous equation system (2.6) Ax = b has at leastone solution and m = n, then the system has a unique solution if the rank of A is n,and infinitely many solutions if the rank of A is less than n.

So now we need a way of figuring out if the rank of an order n square matrixA is n, or is less than n. Fortunately there is a relatively easy way to do this. Letme introduce you to the determinant of a square matrix (nonsquare matrices, wherem = n, do not have determinants). We will use det (A) to denote the determinantof a matrix A. The determinant of a matrix is just a number. It can be zero, orpositive, or negative. Let’s start with an order 1 matrix and an order 2 matrix. Thedeterminant of a 1× 1 matrix A = (a11) is just a11; det (a11) = a11. The determinantof a 2× 2 matrix A = ( a b

c d ) is

det

(a bc d

)= ad− bc. (2.8)

Page 37: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

18 CHAPTER 2. BASICS

For example, det ( 1 62 3 ) = −9 and det

( −1 4−3 −7

)= 19. Once we get to square matrices

of orders greater than 2 we have to work harder. There is more than one way tocompute the determinant of a matrix. I will explain only one, the “expansion bycofactors” method. Consider the 3× 3 matrix

A =

⎛⎝−3 1 −2

2 −5 17 2 3

⎞⎠ . (2.9)

For any i = 1, 2, 3 and for any j = 1, 2, 3, the ijth minor of the matrix A, denoted byMij, is the determinant of the matrix obtained from A after we exclude both the ith

row and the jth column of A. Thus, the minors of the matrix A above are

M11 = det

(−5 12 3

)= −17,

M21 = det

(1 −22 3

)= 7,

M31 = det

(1 −2−5 1

)= −9,

M12 = det

(2 17 3

)= −1,

M22 = det

(−3 −27 3

)= 5,

M32 = det

(−3 −22 1

)= 1,

M13 = det

(2 −57 2

)= 39

M23 = det

(−3 17 2

)= −13

M33 = det

(−3 12 −5

)= 13.

When we multiply the ijth minor of the matrix A by (−1)i+j, we obtain the ijth co-factor of the matrix A, denoted by Cij; i.e. Cij = (−1)i+jMij. So for our matrix thecofactors are

C11 = (−1)1+1M11 = −17,

C21 = (−1)2+1M21 = −7,

C31 = (−1)3+1M31 = −9,

C12 = (−1)1+2M12 = 1,

C22 = (−1)2+2M22 = 5,

C32 = (−1)3+2M32 = −1,

C13 = (−1)1+3M13 = 39,

C23 = (−1)2+3M23 = 13,

C33 = (−1)3+3M33 = 13.

The “expansion by cofactors” method of evaluating a determinant is simple. Allyou do is choose any row or any column and then, for each element in that row orcolumn, add up the values of the element multiplied by its cofactor. Let’s pick thethird column, so j = 3. The sum of the elements in the third column of (2.9), eachmultiplied by its cofactor, is

a13C13 + a23C23 + a33C33 = −2× 39 + 1× 13 + 3× 13 = −26.

Now let’s instead pick the second row, so i = 2. The sum of the elements in thesecond row, each multiplied by its cofactor, is

a21C21 + a22C22 + a23C23 = 2× (−7) + (−5)× 5 + 1× 13 = −26.

Page 38: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.3. SOME LINEAR ALGEBRA 19

Give it a try. Pick a row or a column and use the expansion by cofactors. No matterwhich row or column you choose, you will compute the same number, −26. This isthe value of the determinant of the matrix. So, if you like shortcuts, you will lookat a matrix to see which row or column contains the most zero elements (if any) andthen expand along that row or column to minimize the amount of boring arithmetic.

An Exercise. Compute the determinants of the matrices below.

A =

(4 −1−8 6

), B =

⎛⎝3 0 −45 1 −21 −1 6

⎞⎠ , and C =

⎛⎝ 4 −3 5−2 1 −3−3 4 −2

⎞⎠ . (2.10)

Answer. det (A) = 24 − 8 = 16. To minimize the work needed to compute thedeterminant of B, expand along either the first row or the second column. Let’schoose the second column. Then,

det (B) = (−1)2+2(1)M22 + (−1)3+2(−1)M32 =M22 +M32

= det

(3 −41 6

)+ det

(3 −45 −2

)= (18 + 4) + (−6 + 20) = 36.

Let’s expand along the second row of C. Then,

det (C) = (−1)2+1(−2)M21 + (−1)2+2(1)M22 + (−1)2+3(−3)M23

= 2M21 +M22 + 3M23 = 2

(−3 54 −2

)+

(4 5−3 −2

)+ 3

(4 −3−3 4

)= 2(6− 20) + (−8 + 15) + 3(16− 9) = 0.

For square matrices of orders n ≥ 4, the expansion by cofactors method of com-puting a determinant is similar to the computations we have already performed, butthe method quickly becomes very tedious as n increases. Once n is 5 or higher youprobably should use a computer to do the job for you. To evaluate the determinantof an order n matrix A, you again start by choosing any row i or any column j andthen expand along that row or column, giving

det (A) =n∑

j=1

(−1)i+jaijMij︸ ︷︷ ︸row expansion

=n∑

i=1

(−1)i+jaijMij︸ ︷︷ ︸column expansion

. (2.11)

Each of the minorsMij is a determinant of an order n−1 square matrix. Each of thesedeterminants must be evaluated, so, for each in turn, write the determinant as a sum

Page 39: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

20 CHAPTER 2. BASICS

like (2.11) of minors that are determinants of order n− 2 matrices. Keep going untilyou end up with only order 2 minors and use (2.8) to evaluate them. Problem 2.1asks you to evaluate the determinant of a 4 × 4 matrix to give you an idea of whatall of this involves.

Why are we computing determinants? Here is why.

Theorem 2.5. Let A be an order n square matrix. Then det (A) = 0 if and only ifr(A) = n.

Combine this statement with Theorem 2.4 and we learn that a system Ax = b ofn linear simultaneous equations in n unknowns with at least one solution has onlya unique solution if the determinant of A is not zero, since this ensures that thecolumns of A are a linearly independent set of vectors. In other words, the rank ofthe order n matrix A is n. We say that such a matrix has full rank. Let’s look againat the matrices in (2.10). There we see that det (A) = 16 = 0, so A has full rank,r(A) = 2, telling us that the columns of A are a linearly independent set of vectors.Similarly, det (B) = 36 = 0, so B is a full rank matrix, r(B) = 3, again meaning thatthe columns of B are a linearly independent set of vectors. det (C) = 0, so C is nota full rank matrix. r(C) < 3 tells us that at least one of the columns of C is a linear

combination of one or both of the other two columns. In fact,(

5−3−2

)= 2(

4−2−3

)+( −3

14

).

This section presents only a few topics in linear algebra, each of which is used laterin this book. Indeed, several of the important results in constrained optimizationtheory are applications of topics introduced here. Comprehensive introductions tolinear algebra are available in many other books.

2.4 Direct Products of Sets

Suppose we have two sets, possibly containing quite different sorts of elements. Let’ssay that the sets are

S1 = {1, 2, 3} and S2 = {frog, toad, newt, salamander}.We want to construct all possible ordered pairings of the elements of the first set withelements of the second set. Particularly, we want the first element in each pair to befrom the first set and the second element in each pair to be from the second set. Thecollection of all such pairs is called the direct product of S1 with S2 and is denoted byS1 × S2. Thus,

S1 × S2 = {(1, frog), (1, toad), (1, newt), (1, salamander), (2, frog), (2, toad), (2, newt),

(2, salamander), (3, frog), (3, toad), (3, newt), (3, salamander)}.

Page 40: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.4. DIRECT PRODUCTS OF SETS 21

By writing S1 first and S2 second in the symbol S1 × S2, we state that in each pairthe first element is from S1 and the second element is from S2.

So, what is the direct product S2 ×S1? Go ahead. Write it down before you readfurther.

The answer is the collection of pairs

S2 × S1 = {(frog, 1), (frog, 2), (frog, 3), (toad, 1), (toad, 2), (toad, 3), (newt, 1),(newt, 2), (newt, 3), (salamander, 1), (salamander, 2), (salamander, 3)}.

Notice that S1 × S2 = S2 × S1; the order of the direct product matters.

Briefly put, then, the direct product

S1 × S2 = {(x, y) | x ∈ S1, y ∈ S2}

and the direct product

S2 × S1 = {(y, x) | x ∈ S1, y ∈ S2}.

For a second example, suppose that S1 = [0, 1] and that S2 = [1, 3]. Then thedirect products S1 × S2 and S2 × S1 are as displayed in Figure 2.5.

Figure 2.5: The direct products of the sets S1 = [0, 1] and S2 = [1, 3].

Ordered lists, such as the above pairs, can be of any length. For example, we canthink of a triple (x, y, z), where x is an element of some set X, y is an element ofsome set Y , and z is an element of some set Z; i.e. (x, y, z) ∈ X × Y × Z. Or we

Page 41: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

22 CHAPTER 2. BASICS

can think of a list (x1, . . . , xn) of some arbitrary length n in which each element xibelongs to some set Xi, for i = 1, . . . , n. We write that

(x1, . . . , xn) ∈ X1 × · · · ×Xn = ×ni=1Xi.

For economists the usual such lists are n-dimensional vectors of real numbers.

Definition 2.3 (Direct Product of Sets). The direct product of the sets S1, . . . , Sn

is×n

i=1Si = S1 × · · · × Sn = {(x1, . . . , xn) | x1 ∈ S1, . . . , xn ∈ Sn}.When the sets S1, . . . , Sn are the same set, S, the symbol S × · · · × S︸ ︷︷ ︸

n times

is often

abbreviated to Sn. For most of us, the most common example of this is to write

�n ≡ �× · · · × �︸ ︷︷ ︸n times

= {(x1, . . . , xn) | x1 ∈ �, . . . , xn ∈ �}.

2.5 Addition of Sets

Here are two simple sets, S1 = {1 toaster, 3 toasters} and S2 = {4 toasters, 6 toasters}.Notice that all of the elements in both sets are measured in the same unit. This isimportant because we are about to add together elements from the two sets and forthis to have meaning we require that the elements being added are measured in thesame unit. The addition of these two sets is simply the collection of the sums of allof the possible pairings of the elements of the two sets. This is

S1 + S2 = {1 toaster + 4 toasters, 1 toaster + 6 toasters, 3 toasters + 4 toasters,

3 toasters + 6 toasters, }= {5 toasters, 7 toasters, 9 toasters}.

The order in which we add the elements of the sets does not matter so it follows thatS1 + S2 = S2 + S1.

It should be obvious that we can add together as many sets as we want, so long asthe elements in all of the sets have the same unit of measurement. Convince yourselfthat, if S3 = {0 toasters, 2 toasters}, then

S1 + S2 + S3 = {5 toasters, 7 toasters, 9 toasters, 11 toasters}

and

S1+S2+S3 = S1+S3+S2 = S2+S1+S3 = S2+S3+S1 = S3+S1+S2 = S3+S2+S1.

Page 42: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.5. ADDITION OF SETS 23

Definition 2.4 (Addition of Sets). Let X be a set for which the operation of addition,+, is defined. Let Si ⊆ X for i = 1, . . . , n. Then,

n∑i=1

Si = S1 + · · ·+ Sn = {x | x =n∑

i=1

xi, xi ∈ Si for i = 1, . . . , n}.

An alternative, equivalent, way of writing this sum of sets is

S1 + · · ·+ Sn = ∪x1∈S1

...xn∈Sn

{n∑

i=1

xi

}.

This too is the statement that the set S1+· · ·+Sn consists of all of the sums x1+· · ·+xnthat can be constructed from the elements of the sets S1, . . . , Sn.

An Exercise. S1 = {−4, 2, 5}. S2 = {1, 2}. What is the set S1 + S2?

Answer. S1 + S2 = {−3,−2, 3, 4, 6, 7}.

Figure 2.6: S1 + S2 = ([0, 1]× [−1, 1]) + ([1, 3]× [3, 5]) = [1, 4]× [2, 6].

An Exercise. The negative of a set S is denoted by −S and is the set that containsthe negatives of all of the elements of S. For example, if S = {−3, 0, 2}, then−S = {−2, 0, 3}. Is it true that adding a set S to its negative −S creates thesingleton set {0}; i.e. must S + (−S) = {0}?

Page 43: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

24 CHAPTER 2. BASICS

Answer. No, in general. If the set S is a singleton, say S = {5}, then the set −S ={−5} and then S + (−S) = {0}. But if, say, S = {−4,−1, 5}, then −S = {−5, 1, 4}and so S + (−S) = {−9,−6,−3, 0, 3, 6, 9} = {0}.An Exercise. What is the set S = S1 + S2 = S2 + S1 when S1 = {(x1, x2) | 0 ≤x1 ≤ 1, −1 ≤ x2 ≤ 1} and S2 = {(x1, x2) | 1 ≤ x1 ≤ 3, 3 ≤ x2 ≤ 5}?Answer. S = {(x1, x2) | 1 ≤ x1 ≤ 4, 2 ≤ x2 ≤ 6}; see Figure 2.6.

If you had some trouble getting this answer, then here is an easy way to doit. Think of adding the set S1 to each point in turn of the set S2. Let’s start bytaking the left-bottom corner point (1, 3) of S2 and then adding the entire set S1 tothis point. Then take the next point just a tiny bit directly above (1, 3) and againadd the entire set S1 to this point. Keep doing this until you reach the left-upperpoint (1, 5) of S2. You now have the “ribbon” displayed in the left-side panel ofFigure 2.7 of vertically aligned, overlapping rectangles that in union are the rectangle{(x1, x2) | 1 ≤ x1 ≤ 2, 2 ≤ x2 ≤ 6}.

Figure 2.7: Constructing S1 + S2 = ([0, 1]× [−1, 1]) + ([1, 3]× [3, 5]) = [1, 4]× [2, 6].

If we do the same thing but start, say, at the point (2, 3) in S2 and then continueup to the point (2, 5) in S2, then we will create another ribbon of vertically aligned,overlapping rectangles that in union are the rectangle {(x1, x2) | 2 ≤ x1 ≤ 3, 2 ≤x2 ≤ 6} (see the right-side panel of Figure 2.7). We do this for every vertical linein S2, finally creating the ribbon of vertically aligned, overlapping rectangles that in

Page 44: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.5. ADDITION OF SETS 25

union are the rectangle {(x1, x2) | 3 ≤ x1 ≤ 4, 2 ≤ x2 ≤ 6}; see the right-side panelof Figure 2.7. Now we take the union of all of these rectangles and we see that thesum of the sets S1 and S2 is the rectangle {(x1, x2) | 1 ≤ x1 ≤ 4, 2 ≤ x2 ≤ 6} that isdisplayed in Figure 2.6.

Another Exercise. Let S1 = {(x1, x2) | 1 ≤ x1 ≤ 3, 3 ≤ x2 ≤ 5} and let−S1 = {(x1, x2) | −3 ≤ x1 ≤ −1, −5 ≤ x2 ≤ −3}. What is the set S = S1 + (−S1)?

Answer. S = {(x1, x2) | −2 ≤ x1 ≤ 2, −2 ≤ x2 ≤ 2}; see Figure 2.8.

Figure 2.8: S1 + S2 = ([1, 3]× [3, 5]) + ([−3,−1]× [−5,−3]) = [−2, 2]× [−2, 2].

One More Exercise. Let S1, S2 ⊂ �2 with S1 = {(x1, x2) | x1 − x2 = 0} andS2 = {(x1, x2) | x1 + x2 = 0}. What is the set S1 + S2?

Answer. All of �2.

Now let’s see how good you have become. Try this next exercise. Let S1 ={(x1, x2) | 0 ≤ x1 ≤ 2, 0 ≤ x2 ≤ 1} and S2 be the triangle with vertices at (0, 0),(1, 0), and (1

2, 1) along with all of the points inside the triangle. These two sets

are displayed in Figure 2.9. What is the sum S1 + S2 of these two sets? This isproblem 2.9. Try it and then check your answer.

Before we leave this discussion of adding sets, there is one caution that I needto emphasize to you. When we say that some element x ∈ X, we often mean that

Page 45: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

26 CHAPTER 2. BASICS

Figure 2.9: The sets S1 and S2.

X is a reference set; i.e. X is the set of all of the elements x that we are allowedto consider. So suppose that we are allowed to consider only values for x satisfying0 ≤ x ≤ 4, meaning that the reference set is X = [0, 4]. Now consider the subsetsS1 and S2 of X that are S1 = [0, 2] and S2 = [1, 3]. What do you mean by S1 + S2?Is S1 + S2 = [1, 5]? If you intend that S1 + S2 must be a subset of the referenceset X = [0, 4], then the answer must be that S1 + S2 = [1, 4] only, since we are notpermitted to consider any values outside of [0, 4]. If instead you intend that S1 + S2

is a subset of X +X = [0, 8], then indeed S1 + S2 = [1, 5]. So be careful. When youadd a set S1 ⊂ X to another subset S2 ⊂ X, you need to consider whether or notyou intend that S1 + S2 ⊂ X also. Sometimes there is no need to pay attention tothis detail. For example, if X = � and S1, S2 ⊂ �, then S1 + S2 ⊂ � always. Whenadding sets, always remember what is the reference set.

2.6 Convex Sets

Simply put, a set is convex if, for every pair of elements in the set, every point in the“straight line” joining the two elements is a member of the set. Consider the subsetsof �2 that are displayed in Figure 2.10. The subsets S1, S2, and S3 are all convexsets because any straight line that runs from any arbitrarily chosen point in the setto any other arbitrarily chosen point in the same set is entirely contained within theset. The subsets S4, S5, and S6 are not convex since for each set there is at leastone pair of points in the set such that the straight line connecting the points is noteverywhere within the set.

The “straight line” between two members x′ and x′′ of some set is the set of convexcombinations of x′ and x′′. Each of these convex combinations is just a weightedaverage of x′ and x′′ in which the weights are θ and 1 − θ with 0 ≤ θ ≤ 1. More

Page 46: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.6. CONVEX SETS 27

Figure 2.10: Some convex and nonconvex subsets of �2.

formally, for a given value of θ ∈ [0, 1], the θ-weighted convex combination of x′ andx′′ is

x(θ) = (1− θ)x′ + θx′′.

If θ = 0, then the value of the convex combination is x(0) = x′. x(1) = x′′. x(12

)is

the point halfway along the straight line connecting x′ and x′′ or, if you prefer, it isthe simple arithmetic average of x′ and x′′.

Definition 2.5 (Convex Set). A set S is convex if it is empty or if, for every x′, x′′ ∈S and for every θ ∈ [0, 1], x(θ) = (1− θ)x′ + θx′′ ∈ S.

The above discussion and Figure 2.10 use the term “straight line” to describewhen a set is convex or not. This is well enough for sets in which straight linesare well-defined geometric objects, such as in the real spaces �n. But the ideasof convex combinations and convex sets are much more general than Figure 2.10suggests. Consider the set of all of the real-valued and continuous functions f : X �→� that are defined on some set X. Any convex combination of such functions is alsoa real-valued and continuous function defined on X, so this set of functions is convexeven though it cannot be depicted using a diagram as simple as Figure 2.10. Considerany two quadratic functions a1x

2 + b1x + c1 and a2x2 + b2x + c2 mapping from the

Page 47: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

28 CHAPTER 2. BASICS

real line �. The convex combination of these quadratics is

((1− θ)a1 + θa2)︸ ︷︷ ︸a

x2 + ((1− θ)b1 + θb2)︸ ︷︷ ︸b

x+ (1− θ)c1 + θc2︸ ︷︷ ︸c

,

which is another quadratic. Thus the set of all quadratics mapping from the realline is a convex set. There are endless other examples, but the main point here isthat convexity of sets is an idea that is used for much more than just subsets of realnumber vector spaces.

It is often necessary to distinguish between sets such as S1 and S2 in Figure 2.10.Both sets are convex. But notice that the boundary of S1 is everywhere “curved”while the boundary of S2 has a “straight” segment. Because the boundary of S1 iseverywhere curved, any line connecting two points in S1, excluding its end points,lies entirely in the interior of S1; i.e. with its end points removed, such a linenever intersects the boundary of S1. In contrast, in set S2 it is possible to find twopoints such that the straight line connecting them, with its end points removed, doesintersect with the set’s boundary – see the right-most of the three dashed lines in S2.A set such as S1 is called strictly convex. A set such as S2 is just convex, althoughwe should call it weakly convex whenever we wish to emphasize that it is not strictlyconvex. By the way, the interior of a set S is usually denoted by the symbol Int S.

Definition 2.6 (Strictly Convex Set). A set S is strictly convex if, for all x′, x′′ ∈ Swith x′ = x′′, x(θ) = (1− θ)x′ + θx′′ ∈ Int S for all θ ∈ (0, 1).

Notice that we consider only 0 < θ < 1. By not considering either θ = 0 or θ = 1,we are “chopping off” the endpoints x′ and x′′ from the line connecting x′ and x′′.Notice also that for a set to be strictly convex it must contain at least two differentpoints; we are not allowed to consider x′ = x′′.

It is important to note that a set must have a nonempty interior to be strictlyconvex. Consider, for example, the set S3 in Figure 2.10. S3 is (weakly) convex;because the set does not have an interior (Int S3 = ∅), it cannot be strictly convex.

Is a singleton set, such as, say, S = {3}, a convex set? Is a singleton set a strictlyconvex set? Consider your answers before proceeding to the next paragraph.

Look again at Definition 2.6 and recollect that a strictly convex set S must haveat least two distinct points, x′ and x′′ with x′ = x′′, in it. Obviously this is not so fora singleton set, which is, therefore, not a strictly convex set. Why so? For a set tohave a nonempty interior, it must contain more than one point, right? A singletonset has no interior and so cannot be strictly convex. Even so, is a singleton set aconvex set? Look again at Definition 2.5 and notice that there is no requirement that

Page 48: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.6. CONVEX SETS 29

the points x′ and x′′ be distinct; i.e. x′ = x′′ is allowed. So suppose S is the singletonset {x′}. Take x′′ = x′ and the convex combination x(θ) = (1− θ)x′+ θx′′ = x′ for allθ ∈ [0, 1], so x(θ) = x′ ∈ S for all θ ∈ [0, 1]. Thus a singleton set is always a weaklyconvex set and is never a strictly convex set.

Is the empty set ∅ a convex set? Definition 2.5 simply asserts that the empty set isconvex. Is the empty set a strictly convex set? Definition 2.6 requires that a strictlyconvex set contain at least two distinct points. Obviously this is not so for the emptyset, so the empty set is only weakly convex by definition.

For economists, common examples of convex sets are the budget set B(p1, p2,m) ={(x1, x2) | x1 ≥ 0, x2 ≥ 0, p1x1 + p2x2 ≤ m} with p1 > 0, p2 > 0 and m ≥ 0, the setof bundles WP(x′1, x

′2) = {(x1, x2) | (x1, x2) � (x′1, x

′2)} that a consumer with convex

preferences prefers at least as much as a given bundle (x′1, x′2), and the technology

set T = {(x, y) | x ≥ 0, 0 ≤ y ≤ f(x)} of a firm with a production function f(x)that exhibits decreasing returns to scale and uses a quantity x of a single input toproduce a quantity y of a single product. All three sets are displayed in Figure 2.11.As drawn, the budget set B(p1, p2,m) and the technology set T are weakly convexwhile the weakly preferred set WP(x′1, x

′2) is strictly convex.

Convex sets have a variety of useful properties. Included in these are that theintersection of convex sets is a convex set, the addition of convex sets is a convex setand the direct product of convex sets is a convex set.

Theorem 2.6 (Intersection of Convex Sets). Let S1, . . . , Sn be convex subsets of aset X. Then the set S = S1 ∩ · · · ∩ Sn is a convex subset of X.

Figure 2.12 illustrates the result by displaying six intersections of convex sets. S1

and S2 are both strictly convex sets. Their intersection, S1 ∩ S2, is also a strictlyconvex set. So is it always the case that the intersection of two strictly convex sets isa strictly convex set? Have a good look at Figure 2.12 before you answer.

The answer is “No.” For example, S7 and S8 are both strictly convex sets, buttheir intersection is only a singleton set, and a singleton set is always weakly convex.Again, both S9 and S10 are strictly convex sets, but their intersection is the empty set,which is a weakly convex set. But there is an important case in which the intersectionof two strictly convex sets is a strictly convex set. What is it? Hint: Look again atFigure 2.12 and note what is different about S1 ∩S2 and either of S7 ∩S8 or S9 ∩S10.

The crucial difference is that S1 ∩ S2 has a nonempty interior, while both S7 ∩ S8

and S9∩S10 have empty interiors. The theorem is that, “if S1 and S2 are both strictlyconvex subsets of X, then S1 ∩ S2 is a strictly convex subset of X if and only if theinterior of S1 ∩ S2 is not empty.”

Page 49: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

30 CHAPTER 2. BASICS

Figure 2.11: A weakly convex budget set, a strictly convex set of weakly preferredconsumption bundles, and a weakly convex technology set.

Can the intersection of weakly convex sets be a strictly convex set? Once again,have a good look at Figure 2.12 before you answer.

The answer is “Yes.” In Figure 2.12 the sets S5 and S6 are only weakly convex.Yet their intersection S5 ∩ S6 is a strictly convex set. Even so, as is illustrated byS11∩S12, the intersection of two weakly convex sets can be a weakly convex set. AndS3 ∩ S4 = S4 illustrates that the intersection of a strictly convex set with a weaklyconvex set can be a weakly convex set. It should take you only a moment to think ofsome examples in which the intersection of a strictly convex set and a weakly convexset is a strictly convex set.

Theorem 2.7 (Addition of Convex Sets). Let S1, . . . , Sn be convex subsets of a set X.Then the set S = S1 + · · ·+ Sn is a convex subset of X + · · ·+X︸ ︷︷ ︸

n times

.

Figure 2.7, Figure 2.8, and, together, Figures 2.9 and Figure 2.34 all illustrate theresult. Problems 2.16 and 2.17 ask you for proofs.

Page 50: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.6. CONVEX SETS 31

Figure 2.12: Six intersections of convex sets.

Theorem 2.8 (Direct Product of Convex Sets). For i = 1, . . . , n let Si be a convexsubset of a set Xi. Then the set S = S1 × · · · × Sn is a convex subset of the setX = X1 × · · · ×Xn.

Figure 2.13 provides some examples of the result. In the top panel, S1 = {x | 1 ≤x ≤ 3} and S2 = {x | 3 ≤ x ≤ 4} are both intervals in the real line �1. Their directproduct is the rectangle in �1 × �1 = �2 that is [1, 3] × [3, 4] = {(x1, x2) | 1 ≤ x1 ≤3, 3 ≤ x2 ≤ 4}. Notice that both S1 and S2 are strictly convex subsets of �1 and yettheir direct product is a weakly convex subset of �2.

In the middle panel of Figure 2.13, S1 = {(x1, x2) | (x1−2)2+(x2−2)2 = 4} is thesubset of �2 that is the circle of radius 2 centered at the point (x1, x2) = (2, 2). Theset S2 = {2} is the singleton set of �1 that contains only the number 2. Their directproduct is the circle of radius 2 centered at the point (x1, x2, x3) = (2, 2, 2) that liesonly in the “horizontal” plane x3 ≡ 2. The set S1 is not a convex set since it consistsonly of the circumference of the circle. As a consequence the set S1×S2 is not convex.However, if we change S1 by also including all of the points inside the circle, thenwe have changed S1 into the disk with radius 2 centered at (x1, x2) = (2, 2) in �2;i.e. now S1 = {(x1, x2) | (x1 − 2)2 + (x2 − 2)2 ≤ 4}. S1 is now a strictly convexsubset of �2. S2 is a weakly convex subset of R1. The direct product S1 × S2 then

Page 51: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

32 CHAPTER 2. BASICS

Figure 2.13: Three direct products of convex sets.

Page 52: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.7. PARTIAL DERIVATIVES 33

becomes the disk of radius 2 centered at (x1, x2, x3) = (2, 2, 2) that lies only in theplane x3 ≡ 2; i.e. now S1 × S2 = {(x1, x2, 2) | (x1 − 2)2 + (x2 − 2)2 ≤ 4}. This is aweakly convex subset of �3.

In the bottom panel of Figure 2.13 S1 is again the subset of �2 that is the circleof radius 2 centered at (x1, x2) = (2, 2). S2 is the interval [2, 3] in �1. The directproduct S1 × S2 is the “hollow vertical cylinder” made up of all of the circles ofradius 2 and centered at (x1, x2, x3) = (2, 2, x3) with “heights” 2 ≤ x3 ≤ 3; i.e.S1 × S2 = {(x1, x2, x3) | (x1 − 2)2 + (x2 − 2)2 = 4, 2 ≤ x3 ≤ 4}. The circularset S1 is not a convex set, so the hollow cylindrical direct product S1 × S2 is not aconvex set either. If, again, we alter S1 to the disk in �2 that has radius 2 and center(x1, x2) = (2, 2), then the direct product S1 × S2 becomes the “solid” version of thecylinder seen in the bottom panel of Figure 2.13. That is, S1 × S2 is composed ofall of the disks of radius 2 that are centered at (x1, x2, x3) = (2, 2, x3) with “heights”2 ≤ x3 ≤ 3; S1 × S2 = {(x1, x2, x3) | (x1 − 2)2 + (x2 − 2)2 ≤ 4, 2 ≤ x3 ≤ 3}. Thissolid cylinder is only a weakly convex set of �3 even though S1 is a strictly convexsubset of �2 and S2 is a strictly convex subset of �1.

2.7 Partial Derivatives

We all know that a derivative of a function f(x) at, say, x = 6, is a slope. Put inother words, it is, at the point x = 6, the rate at which the value of f changes as xincreases. What I want to emphasize is that any derivative is a directional quantity.To repeat myself with only a slight rewording, at x = 6 the value of the derivative ofthe function f(x) with respect to x is the rate of change at x = 6 of the value of fin the direction of increasing x. We scarcely take notice of this when thinking aboutfunctions of a single variable, f(x), because there is only one direction in which xcan be increased. But we have to be more careful when we think about the slope ofa function of more than a single variable. So suppose we have a function f(x1, x2) oftwo variables, x1 and x2. What do we mean now by the “slope of f”? Do we meanthe slope of f in the direction of increasing x1? Do we mean the slope of f in thedirection of increasing x2? Or do we mean the slope of f in some other direction?Unless we say exactly which direction interests us, the phrase “slope of f” has nomeaning. An example will help us to think about this.

Consider the function f(x1, x2) = x1/21 x

1/22 . Two graphs of this function are pro-

vided in Figure 2.14, both plotted only for 0 ≤ x1 ≤ 9 and 0 ≤ x2 ≤ 9. In the left-sideof Figure 2.14 the graph shows lines drawn on the surface that form a rectangular

Page 53: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

34 CHAPTER 2. BASICS

grid. Each one of these lines corresponds either to a fixed x1 value or to a fixed x2value. The graph in the right-side panel of Figure 2.14 shows various constant heightlines (contours) drawn on the surface. Each depiction of the graph has its distinctuses, as we shall see.

Figure 2.14: Two plots of f(x1, x2) = x1/21 x

1/22 .

Before we go further, I wish to introduce a useful notational device. When I writef(x1, x2) with a comma between x1 and x2, I mean that both x1 and x2 are variables;both can change in value. But when I write f(x1; x2) with a semi-colon between x1and x2, I mean that x1 is a variable but the value of x2 is fixed. Similarly, writingf(x2; x1) means that x2 is variable but the value of x1 is fixed. Again, for example,if we had a function g of x1, x2, x3, x4, and x5 and I write g(x1, x2, x3, x4, x5), then Imean that all of x1, . . . , x5 are variable, but if I write g(x2, x4; x1, x3, x5), then I meanthat only the values of x2 and x4 are variable and that the values of x1, x3, and x5are all fixed. Simply put, anything to the left of a semi-colon is variable and anythingto the right of a semi-colon is fixed.

The graph displayed in the left-side panel of Figure 2.15 is the same as the graphin the left-side panel of Figure 2.14 except that all but two of the grid lines have beenremoved. These are the lines, parallel to the x1-axis, along which the values of x2are fixed at x2 = 4 and x2 = 9. The diagram in the right-side panel of Figure 2.15 isthe same as in the left-side panel except that because we view the graph by lookingdirectly along the x2-axis we see only the x1-axis and the vertical axis. This givesthe graph a two-dimensional perspective. It appears that we are looking only at thegraphs of the functions f(x1; x2=4) = 2x

1/21 and f(x1; x2=9) = 3x

1/21 .

Page 54: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.7. PARTIAL DERIVATIVES 35

Figure 2.15: Plots of f(x1; x2) = x1/21 x

1/22 with x2 fixed at values of 4 and 9.

The derivative of f with respect to x1 evaluated at some point (x′1, x′2) is the slope

of f at that point in the direction of x1. This is the value of the slope of the curvethat is f(x1; x2 = x′2) at the point x1 = x′1 when x2 is fixed at the value x′2. Forexample, suppose we choose x′2 = 4 and ask what is the slope at the point x1 = 3 of

the curve that is f(x1; x2 = 4) = 2x1/21 ? The slope of the curve (in the x1-direction

since this is the curve’s only direction) is

df(x1; x2=4)

dx1=

d

dx1

(2x

1/21

)= x

−1/21 =

1√x1

.

Thus the value at the point (x1, x2) = (3, 4) of the slope in the x1-direction of thecurve f(x1, x2) is

df(x1; x2=4)

dx1

∣∣x1=3

=1√x1

∣∣x1=3

=1√3.

To make it clear that we are evaluating the slope of a function of more than onevariable, the notation changes from using a “d” to using a “∂ ” symbol (pronounced as“dell”) to denote directional differentiation of functions of more than a single variable,usually known as partial differentiation. Thus we may write

df(x1; x2=4)

dx1=

d

dx1

(2x

1/21

)= x

−1/21 =

1√x1

Page 55: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

36 CHAPTER 2. BASICS

to denote the rate of change of f(x1, x2) with respect to x1 when x2 is fixed at x2 = 4or, alternatively, write the slope of f(x1, x2) in the x1-direction when x2 is fixed atx2 = 4 as

∂f(x1, x2)

∂x1

∣∣x2=4

=1

2√x1

× x1/22

∣∣x2=4

=1

2√x1

× 41/2 =1√x1

.

Thus the value at the point (x1, x2) = (3, 4) of the slope in the x1-direction of thecurve f(x1, x2) is

∂f(x1, x2)

∂x1

∣∣x1=3x2=4

=1

2√x1

× x1/22

∣∣x1=3x2=4

=1

2√3× 41/2 =

1√3.

Similarly, if x2 is instead fixed at x2 = 9, then we are considering only the curvealong which x2 = 9. This is the curve f(x1; x2=9) = x

1/21 × 91/2 = 3x

1/21 . The slope

of this curve in the x1-direction is

df(x1; x2=9)

dx1=

d

dx1

(3x

1/21

)=

3

2√x1

.

Thus the value at the point (x1, x2) = (3, 9) of the slope in the x1-direction of thecurve f(x1, x2) is

df(x1; x2=9)

dx1

∣∣x1=3

=3

2√x1

∣∣x1=3

=3

2√3=

√3

2.

Alternatively, we may write the value at the point (x1, x2) = (3, 9) of the slope off(x1, x2) in the x1-direction when x2 is fixed at x2 = 9 as

∂f(x1, x2)

∂x1

∣∣x1=3x2=9

=1

2√x1

× x1/22

∣∣x1=3x2=9

=1

2√3× 91/2 =

√3

2.

I hope it is now clear that a partial derivative is just a derivative with respect toa single variable. A partial derivative is the rate of change, or slope of the graph, ofa function in the direction of only that single variable, so that, in effect, the values ofany other variables are constant.

An Exercise.

(i) Discover the equations that are the partial derivatives of f(x1, x2) = x1/21 x

1/22

with respect to x2, when x1 = 3, and then when x1 = 9. What are the values of these

Page 56: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.7. PARTIAL DERIVATIVES 37

Figure 2.16: The slopes at x1 = 3 of f(x1, x2) = x1/21 x

1/22 in the x1-direction, when

x2 = 4, and when x2 = 9.

derivatives when x2 = 4? Use the above diagrams to visualize what you have justcomputed.

(ii) Discover the equation that is the partial derivative of f(x1, x2) = x1/21 x

1/22

with respect to x2 for any arbitrary value of x1.

Answer to (i). When x1 is fixed at x1 = 3, the function is f(x2; x1=3) =√3 x

1/22 .

The derivative of this function (i.e. the slope of f(x2; x1=3) in the x2-direction) is

df(x2; x1=3)

dx2=

√3

2√x2

.

The value of this derivative for x2 = 4 is

df(x2; x1=3)

dx2

∣∣x2=4

=

√3

2√x2

∣∣x2=4

=

√3

4.

Page 57: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

38 CHAPTER 2. BASICS

When x1 is fixed at x1 = 9, the function is f(x2; x1=9) = 3x1/22 . The derivative

of this function (i.e. the slope of f(x2; x1=9) in the x2-direction) is

df(x2; x1=9)

dx2=

3

2√x2.

The value of this derivative for x2 = 4 is

df(x2; x1=9)

dx2

∣∣x2=4

=3

2√x2

∣∣x2=4

=3

4.

Answer to (ii). The partial derivative of f(x1, x2) = x1/21 x

1/22 with respect to x2 for

any arbitrary value of x1 is

∂f(x1, x2)

∂x2=

∂x2

(x1/21 x

1/22

)= x

1/21 × 1

2x1/22

=1

2

(x1x2

)1/2

.

2.8 Total Derivatives

Again consider the function f(x1, x2) = x1/21 x

1/22 . The graph of this function that is

displayed in the left-side panel of Figure 2.17 is the same as that displayed in theleft-side panel of Figure 2.14 except that all but four of the “grid” lines have beenremoved.

Two of these four “grid” lines are paths along the graph for fixed values of x1.Specifically, for one line x1 is fixed at the value x1 = 3 and for the other line x1 isfixed at the larger value x1 = 3+Δx1. The other two “grid” lines are paths along thegraph for fixed values of x2. Specifically, for one line x2 is fixed at the value x2 = 4and for the other line x2 is fixed at the larger value x2 = 4 +Δx2.

We have already discovered that the x1-directional rate of change of f(x1, x2) =

x1/21 x

1/22 evaluated at some specific point (x1, x2) = (x′1, x

′2), known as the value at

(x′1, x′2) of the partial derivative of f with respect to x1, is

∂f(x1, x2)

∂x1

∣∣x1=x′

1x2=x′

2

=1

2

(x2x1

)1/2 ∣∣x1=x′

1x2=x′

2

=1

2

(x′2x′1

)1/2

.

We also discovered that the x2-directional rate of change of f(x1, x2) = x1/21 x

1/22

evaluated at some specific point (x1, x2) = (x′1, x′2), known as the value at (x′1, x

′2) of

the partial derivative of f with respect to x2, is

∂f(x1, x2)

∂x2

∣∣x1=x′

1x2=x′

2

=1

2

(x1x2

)1/2 ∣∣x1=x′

1x2=x′

2

=1

2

(x′1x′2

)1/2

.

Page 58: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.8. TOTAL DERIVATIVES 39

Figure 2.17: f(x1; x2=4), f(x1; x2=4 +Δx2), f(x2; x1=3), and f(x2; x1=3 +Δx1).

At the particular point (x1, x2) = (3, 4), these directional rates of change have thevalues

∂f(x1, x2)

∂x1

∣∣x1=3x2=4

=1

2

(4

3

)1/2

≈ 0·577 and∂f(x1, x2)

∂x2

∣∣x1=3x2=4

=1

2

(3

4

)1/2

≈ 0·433.

Given that the value of x2 is fixed at x2 = 4, the change in the value of f that iscaused by changing the value of x1 from x1 = 3 to x1 = 3 +Δx1 is

Δtruef ≡ f(3 + Δx1, 4)− f(3, 4)

= (3 + Δx1)1/2 × 41/2 − 31/2 × 41/2

= 2((3 + Δx1)

1/2 − 31/2).

We can construct a first-order, or “linear,” approximation to this change by supposing,incorrectly, that the function f(x1; x2 = 4) is linear over the interval 3 ≤ x1 ≤ 3 +Δx1 with a constant slope equal to that of f at the point (x1, x2) = (3, 4). This

Page 59: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

40 CHAPTER 2. BASICS

approximation is

Δapproxf ≡ f(3, 4) +∂f(x1, x2)

∂x1

∣∣x1=3x2=4

×Δx1 − f(3, 4)

=1

2

(4

3

)1/2

×Δx1

=Δx1√

3.

Figure 2.18 displays the graph of f(x1; x2 = 4) = 2x1/21 and the graph of the linear

approximation that is

f(3, 4) +∂f(x1, x2)

∂x1

∣∣x1=3x2=4

×Δx1 = 2√3 +

Δx1√3.

Figure 2.18: The true difference f(3+Δx1; 4)− f(3; 4) and its first-order approxima-tion.

Clearly the approximate difference Δapproxf is not the same as the true differenceΔtruef . But are they “almost the same”? Table 2.1 presents values for Δtruef andΔapproxf for some values of Δx1 between 0 and 6. For this example, so long as Δx1is not “too big” the approximate difference Δapproxf has values that are quite closeto the true difference Δtruef , meaning that the first-order approximation for f works

Page 60: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.8. TOTAL DERIVATIVES 41

Table 2.1: True and first-order approximate values of f(3 + Δx1; x2 = 4)− f(3, 4).

Δx1 0 0·1 0·5 1 2 4 6

Δtruef 0 0·006 0·029 0·536 1·008 1·827 2·536Δapproxf 0 0·006 0·029 0·577 1·155 2·309 3·464

well for small enough values of Δx1. Particularly, then, for extremely small values ofΔx1 (i.e. for Δx1 ≈ 0) the first-order approximate difference and the true differenceare indistinguishable from each other. We say that such extremely small values ofΔx1 are infinitesimally small, and denote them by dx1 rather than by Δx1. The trueinfinitesimal change to the value of f that is caused by an infinitesimal change from(x1, x2) = (3, 4) to (x1, x2) = (3 + dx1, 4) is therefore

df =∂f(x1, x2)

∂x1

∣∣x1=3x2=4

× dx1 =1

2

(4

3

)1/2

× dx1 =dx1√3.

Now consider a change to the values of both x1 and x2. Particularly, considera change from (x1, x2) = (3, 4) to (x1, x2) = (3 + Δx1, 4 + Δx2). The difference socaused to the value of f is

Δtruef = f(3 + Δx1, 4 + Δx2)− f(3, 4).

Look at Figure 2.19. The change from (x1, x2) = (3, 4) to (x1, x2) = (3+Δx1, 4+Δx2)can be thought of as a “two-step” change, being a change first from (x1, x2) = (3, 4)to (x1, x2) = (3 + Δx1, 4), followed by a second change from (x1, x2) = (3 + Δx1, 4)to (x1, x2) = (3 + Δx1, 4 + Δx2). This “path” is indicated in Figure 2.19 by thedark arrows on the base plane of the figure. With each step, the value of f changes.Particularly, the change from (x1, x2) = (3, 4) to (x1, x2) = (3 + Δx1, 4) causes thevalue of f to change by

Δtruef1 = f(3 + Δx1, 4)− f(3, 4).

Then, the change from (x1, x2) = (3+Δx1, 4) to (x1, x2) = (3+Δx1, 4+Δx2) causesa further change of

Δtruef2 = f(3 + Δx1, 4 + Δx2)− f(3 + Δx1, 4)

to the value of f . The overall change to the value of f caused by the change from(x1, x2) = (3, 4) to (x1, x2) = (3 + Δx1, 4 + Δx2) is thus

Δtruef = Δtruef1 +Δtruef2.

Page 61: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

42 CHAPTER 2. BASICS

This decomposition of the overall change Δtruef into its two components Δtruef1 andΔtruef2 is displayed in Figure 2.19.

Figure 2.19: Decomposing Δtruef into its two components, Δtruef1 and Δtruef2.

Just as was done earlier for a change to only a single variable, the value of the truechange caused to the value of f can be approximated using first-order (i.e. linear)approximations. We do this by separately approximating each of Δtruef1 and Δtruef2.Figure 2.20 illustrates how this is done by magnifying the part of Figure 2.19 for3 ≤ x1 ≤ 3 + Δx1 and 4 ≤ x2 ≤ 4 + Δx2 and by adding two tangent lines. Thechange Δtruef1 caused by the change from (x1, x2) = (3, 4) to (x1, x2) = (3 + Δx1, 4)is approximated by using the tangent to f at (x1, x2) = (3, 4) in the x1 direction.This is the straight line starting at f(3, 4) with the constant slope ∂f(x1, x2)/∂x1evaluated at (x1, x2) = (3, 4). The change in the height of this straight line due tothe change from (x1, x2) = (3, 4) to (x1, x2) = (3 + Δx1, 4) is

Δapproxf1 =∂f(x1, x2)

∂x1

∣∣x1=3x2=4

×Δx1 =1

2

(4

3

)1/2

×Δx1 =Δx1√

3.

Similarly, the change Δtruef2 caused by the change from (x1, x2) = (3 + Δx1, 4) to(x1, x2) = (3+Δx1, 4+Δx2) is approximated by using the tangent to f at (x1, x2) =

Page 62: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.8. TOTAL DERIVATIVES 43

Figure 2.20: The first-order approximations of Δtruef1 and Δtruef2.

(3+Δx1, 4) in the x2 direction. This is the straight line starting at f(3+Δx1, 4) withthe constant slope ∂f(x1, x2)/∂x2 evaluated at (x1, x2) = (3 + Δx1, 4). The changein the height of this straight line due to the change from (x1, x2) = (3 + Δx1, 4) to(x1, x2) = (3 + Δx1, 4 + Δx2) is

Δapproxf2 =∂f(x1, x2)

∂x2

∣∣x1=3+Δx1x2=4

×Δx2 =1

2

(3 + Δx1

4

)1/2

×Δx2.

The sum of these two first-order approximations is our approximation for the truevalue of the overall change to the value of f ; i.e. our approximation for Δtruef is

Δapproxf = Δapproxf1 +Δapproxf2

=∂f(x1, x2)

∂x1

∣∣x1=3x2=4

×Δx1 +∂f(x1, x2)

∂x2

∣∣x1=3+Δx1x2=4

×Δx2.

Just as we saw above, the approximation Δapproxf1 becomes very close to thetrue value Δtruef1 as Δx1 becomes small, and the approximation Δapproxf2 becomes

Page 63: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

44 CHAPTER 2. BASICS

very close to the true value Δtruef2 as Δx2 becomes small. Therefore, the approxi-mation Δapproxf becomes very close to the true value Δtruef as both Δx1 and Δx2become small. When both Δx1 and Δx2 become infinitesimally small, we write theinfinitesimal change to the value of f that is caused by the infinitesimal change from(x1, x2) = (3, 4) to (x1, x2) = (3 + dx1, 4 + dx2) as

df =∂f(x1, x2)

∂x1

∣∣x1=3x2=4

× dx1 +∂f(x1, x2)

∂x2

∣∣x1=3x2=4

× dx2

=1

2

[(4

3

)1/2

dx1 +

(3

4

)1/2

dx2

].

This expression is called the total differential of f evaluated at (x1, x2) = (3, 4).The idea extends naturally to functions f(x1, . . . , xn) of any finite number n ≥ 1 ofvariables.

Definition 2.7 (Total Differential). Let f : �n �→ � be a differentiable function. Thetotal differential of f evaluated at (x1, . . . , xn) = (x′1, . . . , x

′n) is

df =∂f(x1, . . . , xn)

∂x1

∣∣x1=x′

1

...xn=x′

n

× dx1 + · · ·+ ∂f(x1, . . . , xn)

∂xn

∣∣x1=x′

1

...xn=x′

n

× dxn.

2.9 Continuous Differentiability

Sometimes we will consider functions that are continuously differentiable to someorder. What does this mean? Let’s start with just differentiability to some order.For example, a function is differentiable to order 2 if and only if all of the function’sfirst-order and second-order derivatives exist. Thus a function f(x1, x2) is differen-tiable to order 2 if and only if all of the derivatives ∂f(x1, x2)/∂x1, ∂f(x1, x2)/∂x2,∂2f(x1, x2)/∂x

21, ∂

2f(x1, x2)/∂x1∂x2, ∂2f(x1, x2)/∂x2∂x1, and ∂

2f(x1, x2)/∂x22 are all

well-defined functions. More generally we speak of functions that are differentiableto order r, where r is some integer; r = 1, 2, 3, . . ..

Definition 2.8 (Differentiability to Order r). Let S be a nonempty subset of �n thatis open with respect to the Euclidean topology on �n. A function f : �n → � isdifferentiable to order r at a point in S if and only if the function possesses everyderivative of all orders 1 to r at that point. If the function is differentiable to order rat every point in S, then the function is differentiable to order r on S.

Page 64: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.9. CONTINUOUS DIFFERENTIABILITY 45

Suppose that S is the interval (0, 1). Then, a function f is differentiable to order 3on (0, 1) if and only if at each point in (0, 1) all of the first-, second-, and third-orderderivatives of the function exist. If, as well, every one of these derivative functions isitself a continuous function at every point in (0, 1), then we say that the function f iscontinuously differentiable to order 3 on (0, 1). Thus adding the word “continuously”means that every derivative function to order 3 is a continuous function.

Definition 2.9 (Continuous Differentiability to Order r). Let f be a function that isdifferentiable to order r on a set S. If every one of the derivatives of f from order 1to order r is a continuous function at a particular point in S, then f is continuouslydifferentiable to order r at that point. If f is continuously differentiable to order r atevery point in S, then f is continuously differentiable on S.

Usually the symbol Cr(S) denotes the set of all functions that are continuouslydifferentiable to order r on a set S.

It is easy to think of functions that are continuously differentiable to any order.For example, the quadratic f(x) = ax2 + bx+ c is continuously differentiable on � toany order; i.e. to infinite order. In such a case, we write f ∈ C∞(�).

Now, think of the function g : � �→ � that is defined by

g(x) =

{x+ 1

2, −∞ < x < 0

12(x+ 1)2 , 0 ≤ x <∞.

The first-order derivative of this function is the continuous function

dg(x)

dx=

{1 , −∞ < x < 0

x+ 1 , 0 ≤ x <∞.

The second-order derivative is the function

d2g(x)

dx2=

{0 , −∞ < x < 0

1 , 0 ≤ x <∞

that is discontinuous at x = 0 and so is not everywhere differentiable on �. g isthus a function that is continuously differentiable to order 1 on �, differentiable butnot continuously differentiable to order 2 on �, and is not differentiable to any orderr ≥ 3.

Page 65: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

46 CHAPTER 2. BASICS

2.10 Contour Sets

Think again of the function f(x1, x2) = x1/21 x

1/22 defined for x1 ≥ 0 and x2 ≥ 0. This

function has the same value, 6, at each of the points (x1, x2) = (1, 36), (6, 6), (12, 3),and many others. The set of all such points is

Cf (6) = {(x1, x2) | x1 ≥ 0, x2 ≥ 0, f(x1, x2) = 6}.This is called the contour set of f with level 6 or the level set of f with isovalue 6.Both names, level set and contour set, are in common usage. To avoid confusion, Iwill use only the name “contour set.” Notice that the above contour set is a collectionof pairs (x1, x2) and is, therefore, a subset of the domain of the function. I mentionthis because it is common to use the label “contours” for horizontal lines on the graphof a function, such as those in the right-side panel of Figure 2.14. Such contours arenot contour sets since they are sets of points of the form (x1, x2, y), where y is theconstant “height” of the graph for these points (i.e. points in a contour line on agraph have a “height” dimension y that points in contour sets do not possess). For

example, some of the points in the contour of the graph of f(x1, x2) = x1/21 x

1/22 with

height or level 6 are (3, 12, 6), (36, 1, 6), and (6, 6, 6). The corresponding points in thecontour set of f with level 6 are (3, 12), (36, 1), and (6, 6).

An Exercise. Consider the function f : �2 �→ � that is defined by f(x, y) =40 − (x − 2)2 − (y − 3)2. What is the contour set with level 36 for f? What is thecontour set with level 40 for f? What is the contour set with level 41 for f?

Answer. The contour set with level 36 for f is

Cf (36) = {(x, y) | 40− (x− 2)2 − (y − 3)2 = 36} = {(x, y) | (x− 2)2 + (y − 3)2 = 4}.This set is the circle of radius 2 centered at the point (x, y) = (2, 3).

The contour set with level 40 for f is

Cf (40) = {(x, y) | 40− (x− 2)2 − (y − 3)2 = 40} = {(x, y) | (x− 2)2 + (y − 3)2 = 0}= {(2, 3)}.

This is the singleton set containing only the point (x, y) = (2, 3). A contour set doesnot have to be a curve.

The contour set with level 41 for f is

Cf (41) = {(x, y) | 41−(x−2)2−(y−3)2 = 40} = {(x, y) | (x−2)2+(y−3)2 = −1} = ∅.A contour set for a function f is empty whenever the level for the set is not containedin the image of f .

Page 66: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.11. MARGINAL RATES OF SUBSTITUTION 47

2.11 Marginal Rates of Substitution

Consider the function g(x1, x2) = x1/31 x

2/32 . The contour set with level 4 for g contains

the point (x1, x2) = (1, 8). Now consider another point (1+Δx1, 8+Δx2) that belongsto the same contour set as (1, 8). Then,

g(1, 8) = 11/382/3 = 4 = g(1 + Δx1, 8 + Δx2) = (1 + Δx1)1/3(8 + Δx2)

2/3. (2.12)

What is the relationship between Δx1 and Δx2? That is, if we change the value ofx2, from 8 to 8 + Δx2, then by what amount Δx1 must we change the value of x1,from 1 to 1 + Δx1, so as to keep the value of g constant at 4?

One way to answer this question is to try to solve for Δx1 in terms of Δx2. Oftenthis is not possible, but our example is simple enough that we can obtain such anequation. From (2.12),

4 = (1 + Δx1)1/3(8 + Δx2)

2/3 ⇒ 43 = 64 = (1 + Δx1)(8 + Δx2)2.

Therefore,

Δx1 = −1 +64

(8 + Δx2)2(2.13)

is the equation that tells us by how much, Δx1, to change the value of x1 when thevalue of x2 changes by Δx2 in order to remain in the level 4 contour set of g.

If, as happens in most cases, it is not possible to obtain an equation that explicitlystates Δx1 in terms of Δx2, then what can we do? The answer is that, so long as weconsider only infinitesimally small changes dx1 and dx2 to the values of x1 and x2,then we can obtain expressions for the rate at which x1 must change as x2 changes inorder to keep the value of the function g unchanged. These rate-of-change expressionsbetween x1 and x2 while keeping the value of g constant are called marginal rates ofsubstitution between x1 and x2.

Remember that the total derivative of a function g(x1, x2) is

dg =∂g(x1, x2)

∂x1dx1 +

∂g(x1, x2)

∂x2dx2.

The partial derivatives of g(x1, x2) = x1/31 x

2/32 are

∂g(x1, x2)

∂x1=

1

3

(x2x1

)2/3

and∂g(x1, x2)

∂x2=

2

3

(x1x2

)1/3

.

Page 67: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

48 CHAPTER 2. BASICS

The values of these partial derivatives at (x1, x2) = (1, 8) are

∂g(x1, x2)

∂x1

∣∣x1=1x2=8

=1

3

(8

1

)2/3

=4

3and

∂g(x1, x2)

∂x2

∣∣x1=1x2=8

=2

3

(1

8

)1/3

=1

3.

Thus, evaluated at (x1, x2) = (1, 8), the total derivative of g is

dg =4

3dx1 +

1

3dx2.

But dg = 0 because the point (x1, x2) = (1 + dx1, 8 + dx2), like the point (x1, x2) =(1, 8), is a member of the level 4 contour of g. Accordingly we write

dg = 0 =4

3dx1 +

1

3dx2, (2.14)

and then write thatdx1dx2

∣∣∣∣ g≡4x1=1x2=8

= − 1

4. (2.15)

This is a statement that a very small change in the value of x2 from x2 = 8 tox2 = 8 + dx2 must be accompanied by a very small change in the value of x1 fromx1 = 1 to x1 + dx1, where the change dx1 must be of the opposite sign to dx2 andone-quarter as large as dx2. (Note: It looks like (2.15) was obtained by rearranging(2.14) but this is not so – it is obtained by using the Implicit Function Theorem.)

Let’s check (2.15) by using (2.13) to evaluate at the point (x1, x2) = (1, 8) the rateat which Δx1 must change as Δx2 changes if we are to keep the value of g constantat 4. This rate is the derivative of Δx1 with respect to Δx2 which, from (2.13), is

d(Δx1)

d(Δx2)= − 2× 64

(8 + Δx2)3.

If we make Δx2 extremely small (i.e. Δx2 ≈ 0), then this rate-of-change is indistin-guishable from

d(Δx1)

d(Δx2)= − 2× 64

83= − 1

4,

which is the marginal rate of substitution value that we obtained in (2.15).What we have done so far for functions of two variables generalizes without any

additional work to functions of as many variables as you like. For differentiablefunctions f(x1, . . . , xn), the total derivative of f is

df =∂f(x1, . . . , xn)

∂x1dx1 +

∂f(x1, . . . , xn)

∂x2dx2 + · · ·+ ∂f(x1, . . . , xn)

∂xndxn.

Page 68: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.11. MARGINAL RATES OF SUBSTITUTION 49

Suppose that a point (x′1, . . . , x′n) and a very nearby point (x′1 + dx1, . . . , x

′n + dxn)

are both members of the function’s level α contour set; i.e. f(x′1, . . . , x′n) = α and

f(x′1 + dx1, . . . , x′n + dxn) = α. Then

df = 0 =∂f(x1, . . . , xn)

∂x1

∣∣x1=x′

1

...xn=x′

n

dx1 +∂f(x1, . . . , xn)

∂x2

∣∣x1=x′

1

...xn=x′

n

dx2

+ · · ·+ ∂f(x1, . . . , xn)

∂xn

∣∣x1=x′

1

...xn=x′

n

dxn.

(2.16)

This makes it easy to obtain the marginal rate of substitution between any pairof the variables x1, . . . , xn at some point in the function’s level-α contour set. Forexample, suppose you want the marginal rate of substitution between x1 and x2 atthe point (x1, . . . , xn) = (x′1, . . . , x

′n). Only x1 and x2 are being substituted for each

other, meaning that the values of all of the other variables x3, . . . , xn are being heldfixed at x′3, . . . , x

′n, so dx3 = · · · = dxn = 0 and (2.16) becomes only

df = 0 =∂f(x1, . . . , xn)

∂x1

∣∣x1=x′

1

...xn=x′

n

dx1 +∂f(x1, . . . , xn)

∂x2

∣∣x1=x′

1

...xn=x′

n

dx2.

Thus the marginal rate of substitution between x1 and x2 at the point (x1, . . . , xn) =(x′1, . . . , x

′n) is

dx2dx1

∣∣∣∣ df=0x1=x′

1

...xn=x′

n

= − ∂f(x1, . . . , xn)/∂x1∂f(x1, . . . , xn)/∂x2

∣∣∣∣x1=x′1

...xn=x′

n

.

Similarly, suppose you want the marginal rate of substitution between x3 and x5 atthe point (x1, . . . , xn) = (x′1, . . . , x

′n). Only x3 and x5 are being substituted for each

other, meaning that the values of all of the other variables x1, x2, x4, x6, . . . , xn arebeing held fixed at x′1, x

′2, x

′4, x

′6, . . . , x

′n, so dx1 = dx2 = dx4 = dx6 = · · · = dxn = 0

and (2.16) becomes only

df = 0 =∂f(x1, . . . , xn)

∂x3

∣∣x1=x′

1

...xn=x′

n

dx3 +∂f(x1, . . . , xn)

∂x5

∣∣x1=x′

1

...xn=x′

n

dx5.

Page 69: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

50 CHAPTER 2. BASICS

The marginal rate of substitution between x3 and x5 at (x1, . . . , xn) = (x′1, . . . , x′n) is

dx3dx5

∣∣∣∣ df=0x1=x′

1

...xn=x′

n

= − ∂f(x1, . . . , xn)/∂x5∂f(x1, . . . , xn)/∂x3

∣∣∣∣x1=x′1

...xn=x′

n

.

2.12 Gradient Vectors

In our review of partial derivatives, we emphasized that any derivative is a slope ina specified direction. With a function f(x1, x2) of just two variables, the directionsthat typically are emphasized are the x1 direction and the x2 direction. This is quitenatural since the rate of change of f in the x1 direction is the rate of change of thevalue of f as the value of only x1 is altered; i.e. the partial derivative of f with respectto x1. Similarly, the rate of change of f in the x2 direction is the rate of change ofthe value of f as the value of only x2 is altered; i.e. the partial derivative of f withrespect to x2. The ordered list of these two partial derivatives is called the gradientvector of f , and is denoted by

∇f(x1, x2) ≡(∂f(x1, x2)

∂x1,∂f(x1, x2)

∂x2

).

Notice that by default a gradient vector is a row vector; i.e. a vector in which theelements are written as a horizontal list. I mention this because most other vectors areby default column vectors; i.e. vectors in which the elements are written as verticallists.

The idea of a gradient vector extends naturally to functions f(x1, . . . , xn) of n ≥ 1variables. For such a function, the gradient vector is

∇f(x1, . . . , xn) =(∂f(x1, . . . , xn)

∂x1, . . . ,

∂f(x1, . . . , xn)

∂xn

)and, for brevity, is often denoted by just ∇f(x).An Exercise.(i) What is the gradient vector of the function f(x, y) = 40− (x− 2)2 − (y − 3)2?

(ii) What is the gradient vector of the function g(x1, x2) = x1/21 x

1/22 ?

Answer to (i). The gradient vector of f is the row vector

∇f(x, y) =(∂f(x, y)

∂x,∂f(x, y)

∂y

)= (−2(x− 2), −2(y − 3)).

Page 70: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.13. INNER PRODUCTS 51

Answer to (ii). The gradient vector of g is the row vector

∇g(x1, x2) =(∂g(x1, x2)

∂x1,∂g(x1, x2)

∂x2

)=

(1

2

(x2x1

)1/2

,1

2

(x1x2

)1/2).

2.13 Inner Products

There are all sorts of inner products, but the particular inner product that economistsmost frequently encounter is the Euclidean inner product. Because of its common use,the Euclidean inner product of two vectors x and x′ has been given a special, briefnotation and is usually written as x·x′. This is why the Euclidean inner product of xand x′ is often called the dot product of x and x′ or the scalar product of x and x′.

Definition 2.10 (Euclidean Inner Product). The Euclidean inner product on �n isthe function f : �n ×�n → � that, for x, x′ ∈ �n, is the multiplicative product of thevectors x and x′:

f(x, x′) ≡ x·x′ ≡ (x1, . . . , xn)

⎛⎜⎝x

′1...x′n

⎞⎟⎠ ≡ x1x

′1 + · · ·+ xnx

′n.

Notice that the Euclidean inner product of the vector x and the vector x′ is thesame as the Euclidean inner product of the vector x′ and the vector x; i.e. the orderof multiplication does not matter.

x·x′ = (x1, . . . , xn)

⎛⎜⎝x

′1...x′n

⎞⎟⎠ = x1x

′1 + · · ·+ xnx

′n

= x′1x1 + · · ·+ x′nxn = (x′1, . . . , x′n)

⎛⎜⎝x1...xn

⎞⎟⎠ = x′·x.

An Exercise. What is the value of the Euclidean inner product for each of thefollowing pairs of vectors?

a = (6, 2) and a′ = (−3, 3), b = (3, 24,−18) and b′ = (−1,−8, 6), c = (6, 12) andc′ = (−4, 2), and d = (6, 2, 0) and d′ = (3, 0, 1).

Page 71: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

52 CHAPTER 2. BASICS

Answer.

a·a′ = (6, 2)

(−33

)= −18 + 6 = −12.

b·b′ = (3, 24,−18)

⎛⎝−1−86

⎞⎠ = −3− 192− 108 = −303.

c·c′ = (6, 12)

(−42

)= −24 + 24 = 0.

d·d′ = (6, 2, 0)

⎛⎝301

⎞⎠ = 18.

Euclidean inner products can thus be zero, positive, or negative numbers.

The Euclidean inner product of any vector x with itself is the sum of the squaresof the vector’s elements;

x·x = (x1, . . . , xn)

⎛⎜⎝x1...xn

⎞⎟⎠ = x21 + · · ·+ x2n ≥ 0.

Thus the Euclidean inner product of any vector x with itself is strictly positive unlessthe vector is the zero vector, in which case the inner product is zero.

The Euclidean norm of a vector x ∈ �n is a common way of measuring thedistance of the vector x from the origin. The usual notation for the Euclidean normof x is ‖x‖E.Definition 2.11 (Euclidean Norm). Let x ∈ �n. The Euclidean norm of x is

‖x‖E ≡√x21 + · · ·+ x2n.

It is easy to see that the square of the Euclidean norm of x is the same as theEuclidean inner product of x with itself:

x·x = (x1, . . . , xn)

⎛⎜⎝x1...xn

⎞⎟⎠ = x21 + · · ·+ x2n =

(√x21 + · · ·+ x2n

)2

= ‖x‖2E.

Less obvious is the following very useful result. For a proof, see Simon and Blume1994, pp. 215–16.

Page 72: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.13. INNER PRODUCTS 53

Theorem 2.9. Let x, x′ ∈ �n and let θ be the angle between x and x′. Then,

x·x′ = ‖x‖E × ‖x′‖E × cos θ. (2.17)

You may have forgotten the properties of the cosine function. If so, then considerFigure 2.21, where you will observe that

cos θ

⎧⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎩

= 1 , if θ = 0◦ or θ = 360◦

= −1 , if θ = 180◦

= 0 , if θ = 90◦ or θ = 270◦

∈ (0, 1) , if 0◦ < θ < 90◦ or 270◦ < θ < 360◦

∈ (−1, 0) , if 90◦ < θ < 270◦.

Figure 2.21: The graph of cosine(θ) for 0◦ ≤ θ ≤ 360◦.

It follows, then, that if two vectors x and x′

• are positively collinear (the vectors point in exactly the same direction; θ = 0◦

or θ = 360◦), then their Euclidean inner product is the product of their distancesfrom the origin (norms) because cos 0◦ = cos 360◦ = 1; x′·x′′ = ‖x′‖E‖x′′‖E;

• are negatively collinear (the vectors point in exactly opposite directions; θ =180◦), then their Euclidean inner product is the negative of the product of theirdistances from the origin because cos 180◦ = −1; x′·x′′ = −‖x′‖E‖x′′‖E;

Page 73: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

54 CHAPTER 2. BASICS

• are orthogonal (θ = 90◦ or θ = 270◦; the vectors point in directions that areright-angled with respect to each other), then their Euclidean inner product iszero because cos 90◦ = cos 270◦ = 0; x′·x′′ = 0;

• have an acute angle between them (0◦ < θ < 90◦), then their Euclidean innerproduct is positive and is smaller in size than the product of their distances fromthe origin because 0 < cos θ < 1 for 0◦ < θ < 90◦; ‖x′‖E‖x′′‖E > x′·x′′ > 0;

• have an obtuse angle between them (90◦ < θ < 180◦), then their Euclidean innerproduct is negative and is smaller in size than the product of their distancesfrom the origin because −1 < cos θ < 0 for 90◦ < θ < 180◦; −‖x′‖E‖x′′‖E <x′·x′′ < 0.

An Exercise. For each of the four pairs of vectors of the previous exercise, determineif the two vectors are collinear, negatively collinear, separated by an acute angle, orseparated by an obtuse angle.

Answer. a·a′ = −12 < 0, so the angle θ between a and a′ is obtuse. From (2.17),

a·a′ = −12 = ‖a‖E‖a′‖E cos θ.

The Euclidean norms of a and a′ are

‖a‖E =√62 + 22 =

√40 and ‖a′‖E =

√(−3)2 + 32 =

√18 ,

so

cos θ = − 12√40×

√18

= − 1√5

⇒ θ ≈ cos−1(−0·447) ≈ 116·6◦.

Draw a and a′ to confirm visually that they are separated by an obtuse angle.The vectors b and b′ are negatively collinear. Notice that the Euclidean norm (i.e.

the Euclidean distance from the origin) of b is ‖b‖E =√

32 + 242 + (−18)2 =√909 =

3√101 and that the Euclidean norm of b′ is ‖b′‖E =

√(−1)2 + (−8)2 + 62 =

√101.

Let θ be the angle between b and b′. Then,

b·b′ = −303 = ‖b‖E‖b′‖E cos θ = 3√101×

√101 cos θ = 303 cos θ.

Thus cos θ = −1, which implies that θ = 180◦ and, therefore, that the vectors b andb′ are collinear and point in exactly opposite directions. Draw b and b′ to confirmvisually that they are negatively collinear.

c·c′ = 0, so c and c′ are orthogonal. Draw c and c′ to confirm visually that theyare orthogonal.

Page 74: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.14. GRADIENTS, RATES OF SUBSTITUTION, AND INNER PRODUCTS 55

d·d′ = 18 > 0, so the angle θ between d and d′ is acute. From (2.17),

d·d′ = 18 = ‖d‖E‖d′‖E cos θ.

The Euclidean norms of d and d′ are

‖d‖E =√62 + 22 + 02 =

√40 and ‖d′‖E =

√32 + 02 + 12 =

√10 ,

so

cos θ =18√

40×√10

=9

10⇒ θ = cos−1(0·9) ≈ 25·8◦.

Draw d and d′ to confirm visually that they are separated by an acute angle.

2.14 Gradients, Rates of Substitution, and Inner

Products

These three ideas are related to each other in ways that are crucial to understandingrational choice making, the main topic of this book. So let us take a little time to besure that these connections are recognized and understood. An example will help, soagain consider the function f(x, y) = 40 − (x − 2)2 − (y − 3)2. The gradient vectorfor this function is

∇f(x, y) =(∂f(x, y)

∂x,∂f(x, y)

∂y

)= (−2(x− 2),−2(y − 3)).

The marginal rate of substitution at any point (x, y) other than those with y = 3 is

MRSf (x, y) =dy

dx

∣∣∣∣df=0

= − ∂f(x, y)/∂x

∂f(x, y)/∂y

∣∣∣∣df=0

= − x− 2

y − 3.

Now let’s consider a contour set for f . The level-36 contour set for f , consideredearlier, is as good as any, and is

Cf (36) = {(x, y) | 40− (x− 2)2 − (y − 3)2 = 36} = {(x, y) | (x− 2)2 + (y − 3)2 = 4}.

This set is the circle of radius 2 centered at the point (x, y) = (2, 3), and is displayedin Figure 2.22. (x, y) = (1, 3 −

√3 ) ≈ (1, 1·27) and (x, y) =

(1, 3 +

√3)≈ (1, 4·73)

are two of the points in the contour set Cf (36).

Page 75: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

56 CHAPTER 2. BASICS

At the point (x, y) =(1, 3−

√3), the values of the function’s gradient vector and

marginal rate of substitution are

∇f(1, 3−

√3)=(2, 2

√3)

and MRSf

(1, 3−

√3)=

dy

dx

∣∣∣∣f≡36x=1y=3−√

3

= − 1√3.

Consider a small shift from the point(1, 3−

√3)to a point

(1 + dx, 3−

√3 + dy

)that too is contained in Cf (36). The changes dx and dy to the values of x and y areperturbations from the initial point

(1, 3−

√3). The value of the marginal rate of

substitution, −1/√3, says that the point

(1 + dx, 3−

√3 + dy

)that is very near to(

1, 3−√3)lies in the Cf (36) contour set only if

dy

dx

∣∣∣∣f≡36x=1y=3−√

3

= − 1√3

⇒ dy = − dx√3. (2.18)

In Figure 2.22 the short bold, dashed line through the point(1, 3−

√3)displays

the points (1 + dx, 3 −√3 + dy) that are very near to the point

(1, 3−

√3), and

obey (2.18). The line containing the small perturbations (dx, dy) is displayed asanother short bold, dashed line drawn through the origin in the right-side region ofFigure 2.22 that is surrounded by a finely dotted line. Also displayed is the value of f ’sgradient vector at

(1, 3−

√3). The arrow depicting the direction of the vector and

the short bold, dashed line depicting the perturbation(dx, dy) satisfying (2.18) areorthogonal to each another. Why? Well, the marginal rate of substitution equation(2.18) is

0 =1√3dx+ dy ⇒ 0 = 2

√3

(1√3dx+ dy

)⇒ 0 =

(2dx, 2

√3 dy

)⇒ 0 =

(2, 2

√3)(dx

dy

)

⇒ 0 = ∇f(1, 3−

√3)·(dxdy

).

(2.19)

The Euclidean inner product of two vectors is zero if the vectors are orthogonalto each other, so at the point

(1, 3−

√3)the function’s gradient vector and the

small perturbation vector (dx, dy) are orthogonal. Re-expressed, (2.19) is just the

Page 76: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.14. GRADIENTS, RATES OF SUBSTITUTION, AND INNER PRODUCTS 57

Figure 2.22: The level-36 contour set Cf (36) and the gradient vectors of f evaluatedat the points

(1, 3−

√3)and

(1, 3 +

√3).

statement made earlier that at any point in a contour set a small movement (dx, dy)to another very nearby point in the same contour set must give a total differentialvalue of zero.

Figure 2.22 displays the function’s marginal rate of substitution and its gradientvector at the point

(1, 3 +

√3)also. To check how well you have understood the

reasoning above, try to repeat the reasoning at this other point. You should concludethat at the point

(1, 3 +

√3)the gradient vector ∇f

(1, 3 +

√3)is orthogonal to the

vector (dx, dy) of small perturbations between the point(1, 3 +

√3)and any other

very nearby point(1 + dx, 3 +

√3 + dy

)that is also a member of the function’s level-

36 contour set.

Is it true in general that at any point in a contour set of a differentiable functionthe gradient vector of the function and the vector of small perturbations that moveonly to other nearby points within the contour set are orthogonal to each other? Yes,it is. Moreover, this simple idea is at the heart of the Karush-Kuhn-Tucker NecessityTheorem, one of the most useful results in all of constrained optimization theory.

Page 77: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

58 CHAPTER 2. BASICS

An Exercise. The points (x, y) =(3, 3−

√3)

≈ (3, 1·27) and(3, 3 +

√3)

≈(3, 4·73) are also members of the Cf (36) contour set described above. For each pointin turn, evaluate the function’s gradient vector and the short line containing the smallperturbations (dx, dy) that satisfy the marginal rate of substitution condition at thepoint. Plot the short line and the gradient vector on Figure 2.22, and then confirmthat the gradient vector and the vector of perturbations (dx, dy) are orthogonal toeach other by showing that the value of their Euclidean inner product is zero.

Answer. For the point (x, y) =(3, 3−

√3), the values of the function’s gradient

vector and marginal rate of substitution are

∇f(3, 3−

√3)=(−2, 2

√3)

and MRSf

(3, 3−

√3)=

dy

dx

∣∣∣∣f≡36x=3y=3−√

3

= +1√3.

The value of the marginal rate of substitution asserts that the vector of small pertur-bations (dx, dy) is

(dx, dy) withdy

dx

∣∣∣∣f≡36x=3y=3−√

3

= +1√3

⇒ dy = +1√3dx.

The Euclidean inner product of the gradient vector ∇f and the perturbation vector(dx, dy) is

∇f(3, 3−

√3)·(dxdy

)=(−2, 2

√3)(dx

dy

)=(−2, 2

√3)( dx

dx/√3

)= 0.

For the point (x, y) =(3, 3 +

√3), the values of the function’s gradient vector

and marginal rate of substitution are

∇f(3, 3 +

√3)=(−2,−2

√3)

and MRSf

(3, 3 +

√3)= − 1√

3.

The value of the marginal rate of substitution asserts that the vector of small pertur-bations (dx, dy) is

(dx, dy) withdy

dx

∣∣∣∣f≡36x=3y=3+

√3

= − 1√3

⇒ dy = − 1√3dx.

The Euclidean inner product of the gradient vector ∇f and the perturbation vector(dx, dy) is

∇f(3, 3 +

√3)·(dxdy

)=(−2,−2

√3)(dx

dy

)=(−2,−2

√3)( dx

− dx/√3

)= 0.

Page 78: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.14. GRADIENTS, RATES OF SUBSTITUTION, AND INNER PRODUCTS 59

Another Exercise. Consider the function f(x1, x2) = x1/21 x

1/22 .

(i) What is the gradient vector for this function?

(ii) The points (x1, x2) = (1, 16), (x1, x2) = (4, 4), and (x1, x2) = (16, 1) all belongto the level-4 contour set of f . For each point in turn, compute the value of thefunction’s gradient vector and the marginal rate of substitution between x1 and x2.Then compute the line containing the small perturbations dx = (dx1, dx2) that satisfythe marginal rate of substitution condition at the point. Finally, show that thegradient vector and the perturbation vector dx are orthogonal at the point.

Answer to (i). The gradient vector for f is the row vector

∇f(x1, x2) =(∂f(x1, x2)

∂x1,∂f(x1, x2)

∂x2

)=

(1

2

(x2x1

)1/2

,1

2

(x1x2

)1/2).

Answer to (ii). Along the f ≡ 4 contour set, the marginal rate of substitutionbetween x1 and x2 is

dx2dx1

∣∣f≡4

= − ∂f(x1, x2)/∂x1∂f(x1, x2)/∂x2

= − (1/2)(x2/x1)1/2

(1/2)(x1/x2)1/2= − x2

x1

for x1 and x2 satisfying x1/21 x

1/22 = 4.

See Figure 2.23. At the point (x1, x2) = (1, 16), the value of the function’s gradientvector is

∇f(1, 16) =(1

2

(16

1

)1/2

,1

2

(1

16

)1/2)

=

(2,

1

8

).

This vector is displayed in Bubble 1 in Figure 2.23. The value of the function’smarginal rate of substitution is

dx2dx1

∣∣∣∣df=0x1=1x2=16

= − 16

1= −16. (2.20)

Therefore the perturbation vector is dx = (dx1, dx2), where dx2 = −16dx1. Suchperturbation vectors are displayed in Bubble 1 as a short thick line containing theorigin. The Euclidean inner product of any such perturbation vector and the gradientvector at (x1, x2) = (1, 16) is

∇f(1, 16)·(dx1, dx2) =(2,

1

8

)(dx1

−16 dx1

)= 0.

Page 79: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

60 CHAPTER 2. BASICS

Figure 2.23: Gradients and marginal rates of substitution for f(x1, x2) = x1/21 x

1/22 .

The two vectors are therefore orthogonal.

At the point (x1, x2) = (4, 4), the value of the function’s gradient vector is

∇f(4, 4) =(1

2

(4

4

)1/2

,1

2

(4

4

)1/2)

=

(1

2,1

2

).

The value of the function’s marginal rate of substitution is

dx2dx1

∣∣∣∣df=0x1=4x2=4

= − 4

4= −1. (2.21)

Therefore the perturbation vector is dx = (dx1, dx2), where dx2 = −dx1. Suchperturbation vectors are displayed in Bubble 2 as a short thick line containing theorigin. The Euclidean inner product of any such perturbation vector and the gradientvector at (x1, x2) = (4, 4) is

∇f(1, 16)·(dx1, dx2) =(1

2,1

2

)(dx1− dx1

)= 0.

Page 80: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.15. QUADRATIC FORMS 61

The two vectors are therefore orthogonal.At the point (x1, x2) = (16, 1), the value of the function’s gradient vector is

∇f(16, 1) =(1

2

(1

16

)1/2

,1

2

(16

1

)1/2)

=

(1

8, 2

).

The value of the function’s marginal rate of substitution is

dx2dx1

∣∣∣∣df=0x1=16x2=1

= − 1

16. (2.22)

Therefore the perturbation vector is dx = (dx1, dx2), where −16dx2 = dx1. Suchperturbation vectors are displayed in Bubble 3 as a short thick line containing theorigin. The Euclidean inner product of any such perturbation vector and the gradientvector at (x1, x2) = (16, 1) is

∇f(16, 1)·(dx1, dx2) =(1

8, 2

)(−16 dx2dx2

)= 0.

The two vectors are therefore orthogonal.

2.15 Quadratic Forms

A quadratic form is just a special type of order 2 polynomial in n variables x1, . . . , xn.If n = 1, then the only possible quadratic form is

Q(x) = ax2 where a ∈ �.Examples are Q(x) = x2 (i.e. a = 1), Q(x) = −7x2 (i.e. a = −7) and Q(x) ≡ 0 (i.e.a = 0).

If n = 2, then the quadratic form is

Q(x1, x2) = a11x21 + a12x1x2 + a21x2x1 + a22x

22

= a11x21 + (a12 + a21)x1x2 + a22x

22,

where all of a11, a12, a21 and a22 are real numbers.The general quadratic form is

Q(x1, . . . , xn) = a11x21 + a12x1x2 + · · ·+ a1nx1xn +

a21x2x1 + a22x22 + · · ·+ a2nx2xn +

· · · +

an1xnx1 + an2xnx2 + · · ·+ annx2n,

(2.23)

Page 81: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

62 CHAPTER 2. BASICS

where each coefficient aij is a real number, for i, j = 1, . . . , n.

All of the terms in a quadratic form are of order 2 in x1, . . . , xn. That is, every termis either a square, x2i for i = 1, . . . , n, or it is a cross-product, xixj for i, j = 1, . . . , nwith i = j. There are no linear terms and there are no constants.

(2.23) is the product

xTAx = (x1, . . . , xn)

⎛⎜⎜⎜⎝a11 a12 · · · a1na21 a22 · · · a2n...

.... . .

...an1 an2 · · · ann

⎞⎟⎟⎟⎠⎛⎜⎝x1...xn

⎞⎟⎠ , (2.24)

and so the two identical statements (2.23) and (2.24) are called the quadratic form ofthe matrix A. Hopefully it is clear that only square matrices possess quadratic forms.

Notice that the quadratic form (2.23) contains two cross-product terms xixj foreach i, j = 1, . . . , n with i = j. For example, there is a term a12x1x2 and a terma21x2x1, which can be collected together to write (a12 + a21)x1x2. Similarly, there isa term a1nx1xn and a term an1xnx1. Their sum is (a1n + an1)x1xn. Halving each sumlets us rewrite the quadratic form (2.23) as

Q(x1, . . . , xn) = a11x21 +

a12 + a212

x1x2 + · · ·+ a1n + an12

x1xn +

a12 + a212

x2x1 + a22x22 + · · ·+ a2n + an2

2x2xn +

· · · +

a1n + an12

xnx1 +a2n + an2

2xnx2 + · · ·+ annx

2n.

(2.25)

(2.25) is the same as the product

xTBx = (x1, . . . , xn)

⎛⎜⎜⎜⎝

a11 (a12 + a21)/2 · · · (a1n + an1)/2(a12 + a21)/2 a22 · · · (a2n + an2)/2

......

. . ....

(a1n + an1)/2 (a2n + an2)/2 · · · ann

⎞⎟⎟⎟⎠⎛⎜⎝x1...xn

⎞⎟⎠ .

The matrix A is typically asymmetric while the matrix B, constructed from theelements of the matrix A, is symmetric. Thus the quadratic form of any asymmetricmatrix is also the quadratic form of a symmetric matrix. This fact can be a greatalgebraic convenience.

Page 82: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.15. QUADRATIC FORMS 63

An Exercise. Write each of the quadratic forms listed below in the matrix formxTAx and then in the form xTBx, where A is an asymmetric matrix and B is asymmetric matrix.(i) x21 − 8x1x2 + x22;(ii) 3x21 − 8x1x2 − 5x22; and(iii) 2x21 + x22 + 5x23 + 3x1x2 − 6x1x3 + 10x2x3.

Answer.

(i) The quadratic form is Q(x1, x2) = a11x21 + (a12 + a21)x1x2 + a22x

22, where a11 = 1,

a12 + a21 = −8 and a22 = 1. A is therefore any asymmetric matrix in which a11 = 1,a12 + a21 = −8 and a22 = 1. One of the infinitely many such matrices is

A =

(1 −462454 1

).

The only symmetric matrix, B, has elements b11 = 1, b12 = b21 = −4 and b22 = 1; i.e.

B =

(1 −4−4 1

).

(ii) The quadratic form is Q(x1, x2) = a11x21+(a12+ a21)x1x2+ a22x

22, where a11 = 3,

a12+a21 = −8 and a22 = −5. A is therefore any asymmetric matrix in which a11 = 3,a12 + a21 = −8 and a22 = −5. One of the infinitely many such matrices is

A =

(3 34

−42 −5

).

The only symmetric matrix, B, has elements b11 = 3, b12 = b21 = −4 and b22 = −5;i.e.

B =

(3 −4−4 −5

).

(iii) The quadratic form is Q(x1, x2, x3) = a11x21+(a12+ a21)x1x2+(a13+ a31)x1x3+

a22x22 + (a23 + a32)x2x3 + a33x

23, where a11 = 2, a22 = 1, a33 = 5, a12 + a21 = 3,

a13 + a31 = −6 and a23 + a32 = 10. A is therefore any asymmetric matrix in whicha11 = 2, a22 = 1, a33 = 5, a12 + a21 = 3, a13 + a31 = −6 and a23 + a32 = 10. One ofthe infinitely many such matrices is

A =

⎛⎝ 2 7 −3−4 1 17−3 −7 5

⎞⎠ .

Page 83: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

64 CHAPTER 2. BASICS

The only symmetric matrix, B, has elements b11 = 2, b22 = 1, b33 = 5, b12 = b21 = 3/2,b13 = b31 = −3 and b23 = b32 = 5; i.e.

B =

⎛⎝ 2 3/2 −33/2 1 5−3 5 5

⎞⎠ .

All quadratic forms fall into at least one of five categories. These are as follows.

Definition 2.12 (Definiteness of Quadratic Forms). A quadratic form Q(x1, . . . , xn)is(i) positive-definite if and only if Q(x1, . . . , xn) > 0 for all x ∈ �n, x = 0;(ii) positive-semi-definite if and only if Q(x1, . . . , xn) ≥ 0 for all x ∈ �n;(iii) negative-semi-definite if and only if Q(x1, . . . , xn) ≤ 0 for all x ∈ �n;(iv) negative-definite if and only if Q(x1, . . . , xn) < 0 for all x ∈ �n, x = 0;(v) indefinite if and only if there is at least one point (x′1, . . . , x

′n) ∈ �n and at least

one other point (x′′1, . . . , x′′n) ∈ �n for which Q(x′1, . . . , x

′n) > 0 and Q(x′′1, . . . , x

′′n) < 0.

Figure 2.24: The graph of the positive-definite quadratic form Q(x1, x2) = x21−x1x2+2x22.

If x = 0, then Q(x) = 0, so every quadratic form, no matter its definiteness type,has the origin as a point in its graph.

Page 84: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.15. QUADRATIC FORMS 65

Let’s take n = 2 and ask “What does the graph of a positive-definite quadraticform look like?” By its definition (see (i) above), for any (x1, x2) other than (0, 0),

a11x21 + (a12 + a21)x1x2 + a22x

22 > 0.

We are allowed to choose x2 = 0 (so long as we do not choose x1 = 0 at the sametime). Then we have a11x

21 > 0, so, necessarily, a11 > 0. Similarly, we are allowed to

choose x1 = 0 (so long as we do not choose x2 = 0 at the same time), in which casewe have a22x

22 > 0, so, necessarily, a22 > 0. So when the quadratic form is positive-

definite, it is a sum of two squares a11x21 and a22x

22 in which the coefficients a11 and

a22 are both strictly positive. The graphs of such functions are “positive parabolas”like the one displayed in Figure 2.24 in which a11 = 1, a12 + a21 = −1 and a22 = 2.This quadratic form is

Q(x1, x2) = x21 − x1x2 + 2x22 = x21 − x1x2 +x224

+7x224

=(x1 −

x22

)2+

7x224

> 0 for all (x1, x2) = (0, 0).

Notice that a positive-definite quadratic form has value zero only at the origin.Now let’s ask “What does a positive-semi-definite quadratic form look like?” By

its definition (see (ii) above), for any (x1, x2),

a11x21 + (a12 + a21)x1x2 + a22x

22 ≥ 0.

Choosing x2 = 0 gives a11x21 ≥ 0, so, necessarily, a11 ≥ 0. Similarly, necessarily,

a22 ≥ 0. So when the quadratic form is positive-semi-definite, it is a sum of twosquares a11x

21 and a22x

22 in which the coefficients a11 and a22 are both nonnegative

(either one or both could be zero). The graphs of such functions typically are again“positive parabolas” but may also include a line at a height of zero, as is displayedin Figure 2.25 for a11 = 1, a12 + a21 = −2 and a22 = 1. In this case the quadraticform is Q(x1, x2) = x21 − 2x1x2 + x22 = (x1 − x2)

2 and so has value zero at any point(x1, x2) with x1 = x2, not just at (0, 0).

It should be clear by now that if a quadratic form is negative-definite, then,necessarily, a11 < 0 and a22 < 0, and that if a quadratic form is negative-semi-definite,then, necessarily, a11 ≤ 0 and a22 ≤ 0. The typical graph of a negative-definitequadratic form is like the graph in Figure 2.24 turned upside down. The typicalgraph of a negative-semi-definite quadratic form looks like the graph in Figure 2.25turned upside down.

Figure 2.26 displays the graph of the quadratic form Q(x1, x2) = x21 − x22. Thisquadratic form takes both positive and negative values, so it is indefinite.

Page 85: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

66 CHAPTER 2. BASICS

Figure 2.25: The graph of the positive-semi-definite quadratic form Q(x1, x2) = x21 −2x1x2 + x22 = (x1 − x2)

2.

Theorem 2.10. The quadratic form of an n×n matrix A is

⎧⎪⎪⎨⎪⎪⎩

positive-definitepositive-semi-definitenegative-semi-definite

negative-definite

⎫⎪⎪⎬⎪⎪⎭

only if aii

⎧⎪⎪⎨⎪⎪⎩> 0≥ 0≤ 0< 0

⎫⎪⎪⎬⎪⎪⎭ for all i = 1, . . . , n.

Be clear that this theorem applies to asymmetric as well as to symmetric matrices.

An Exercise. A and B are both n× n matrices. If A is negative-semi-definite andB is negative-definite, then must the matrix A+B be negative-definite?

Answer. For every vector x ∈ �n, x = 0, xTAx ≤ 0 and xTBx < 0. Therefore,

xT (A+ B)x = xTAx+ xTBx < 0 ∀ x ∈ �n, x = 0.

That is, A+B is a negative-definite matrix.So long as the order n of a square matrix is not large, then it may be possible

without too much effort to determine algebraically to which of the five above definite-ness categories a matrix belongs. But for many cases such an approach is much too

Page 86: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.15. QUADRATIC FORMS 67

Figure 2.26: The graph of the indefinite quadratic form Q(x1, x2) = x21 − x22.

tedious. Is there an easier way to make this determination? There are two sensiblealternative approaches. One involves the computation of the eigenroots (also calledcharacteristic roots) of the matrix. I will not discuss this approach, but the interest-ed reader will find it explained in many other texts. The second approach uses thesuccessive principal minors of a square matrix.

Consider the 3× 3 matrix

A =

⎛⎝a11 a12 a13a21 a22 a23a31 a32 a33

⎞⎠ .

Now consider the square submatrices contained in A that all contain the top-leftelement a11 and that expand down along the main, or principal, diagonal of A. Theseare the submatrices

A1 =(a11), A2 =

(a11 a12a21 a22

)and A3 = A =

⎛⎝a11 a12 a13a21 a22 a23a31 a32 a33

⎞⎠ .

These submatrices are the leading principal submatrices of A. The determinants of

Page 87: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

68 CHAPTER 2. BASICS

these leading principal submatrices of A are called the leading principal minors of A:

detA1 = det(a11), detA2 = det

(a11 a12a21 a22

)

and detA3 = detA = det

⎛⎝a11 a12 a13a21 a22 a23a31 a32 a33

⎞⎠. (2.26)

Definition 2.13 (Leading Principal Submatrix). Let A be an n × n matrix. Fork = 1, . . . , n, the kth leading principal submatrix of A is the matrix formed by deletingfrom A the last n− k rows and columns.

Definition 2.14 (Leading Principal Minor). Let A be an n × n matrix. For k =1, . . . , n, the kth leading principal minor of A is the determinant of the kth leadingprincipal submatrix of A.

Odd though it may seem, for a symmetric matrix the strict definiteness (or not)of the matrix can be deduced by the signs of the matrix’s leading principal minors.Theorem 2.11 states how this is done. Notice that the theorem applies only to matricesthat are, first, symmetric and, second, have quadratic forms that are either positive-definite, negative-definite, or indefinite. Particularly, the theorem does not apply tosemi-definite matrices. The theorem also does not apply to asymmetric matrices,but I earlier noted that the quadratic form of any asymmetric matrix is the same asthe quadratic form of a symmetric matrix that is constructed from the asymmetricmatrix. The theorem can, therefore, be applied to any square asymmetric matrixby first constructing the symmetric matrix with the same quadratic form, and thenapplying the theorem to the symmetric matrix.

Theorem 2.11 (Leading Principal Minors and Definiteness). Let A be a symmetricn× n matrix. Then the quadratic form of A is(i) positive-definite if and only if all of the leading principal minors of A are positive;(ii) negative-definite if and only if all of the leading principal minors of A are nonzeroand alternate in sign, with the sign of the kth leading principal minor being (−1)k;and(iii) if any one of the leading principal minors of A is not zero and has a sign incon-sistent with either (i) or (ii) then the quadratic form of A is indefinite.

Statements (i) and (ii) are equivalency statements. Statement (iii) is only a sufficiencystatement.

Page 88: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.15. QUADRATIC FORMS 69

To see more clearly how Theorem 2.11 is used, let’s rewrite the above 3×3 matrixA as a symmetric matrix:

A =

⎛⎝a11 a12 a13a12 a22 a23a13 a23 a33

⎞⎠ .

The leading principal submatrices of A are now

A1 =(a11), A2 =

(a11 a12a12 a22

), and A3 = A =

⎛⎝a11 a12 a13a12 a22 a23a13 a23 a33

⎞⎠ ,

and the leading principal minors of A are

detA1 = det(a11), detA2 = det

(a11 a12a12 a22

), detA3 = detA = det

⎛⎝a11 a12 a13a12 a22 a23a13 a23 a33

⎞⎠.

Theorem 2.11 states that the quadratic form of A is positive-definite if and only if

detA1 = det(a11)= a11 > 0, (2.27)

detA2 = det

(a11 a12a12 a22

)= a11a22 − a212 > 0, (2.28)

and detA3 = detA = det

⎛⎝a11 a12 a13a12 a22 a23a13 a23 a33

⎞⎠ > 0. (2.29)

These conditions restrict the values that the matrix’s elements aij may have. Forexample, (2.27) requires that a11 > 0. Next, (2.28) requires that a11a22 > a212 ≥ 0,which, with a11 > 0, requires that a22 > 0. Together, the leading principal minorstatements (2.27), (2.28), and (2.29) provide all of the restrictions on the elements ofA that ensure that the quadratic form of A is positive-definite.

Theorem 2.11 also states that the quadratic form of the symmetric matrix A is

Page 89: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

70 CHAPTER 2. BASICS

negative-definite if and only if

detA1 = det(a11)= a11 < 0 (nonzero and has the sign (−1)1), (2.30)

detA2 = det

(a11 a12a12 a22

)= a11a22 − a212 > 0 (nonzero and has the sign (−1)2),

(2.31)

and detA3 = detA = det

⎛⎝a11 a12 a13a12 a22 a23a13 a23 a33

⎞⎠ < 0 (nonzero and has the sign (−1)3).

(2.32)

These conditions restrict the values that the matrix’s elements aij may have. (2.30)requires that a11 < 0. (2.31) requires that a11a22 > a212 ≥ 0, which, with a11 < 0,requires that a22 < 0. Together, the leading principal minor statements (2.30), (2.31)and (2.32) provide all of the restrictions on the elements of A that ensure that thequadratic form of A is negative-definite.

2.16 Critical Points

A critical point for a real-valued and differentiable function f(x1, . . . , xn) is a pointat which the function’s slope is zero in all of the x1, . . . , xn directions.

Definition 2.15 (Critical Point). Let O be a nonempty open subset of �n and letf : �n �→ � be a function that is differentiable on O. Then x∗ ∈ O is a criticalpoint of f if the gradient of f is the n-dimensional zero vector at x = x∗; i.e. if∂f(x)/∂xi = 0 at x = x∗ for all i = 1, . . . , n.

Any (interior to the domain) local maximum for f is a critical point, as is anyinterior local minimum or point of inflexion at which the gradient of f is the zerovector. Another type of critical point that will turn out to be of great interest to usis a saddle-point.

2.17 Hessian Matrices

The Hessian matrix of a twice-differentiable function is a matrix array of all of thefunction’s second-order derivatives. Specifically, if f(x1, . . . , xn) is a function that is

Page 90: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.18. QUADRATIC APPROXIMATIONS OF FUNCTIONS 71

differentiable to order 2, then the function’s Hessian matrix is the array of functions

Hf (x) =

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

∂2f

∂x21

∂2f

∂x1∂x2· · · ∂2f

∂x1∂xn∂2f

∂x2∂x1

∂2f

∂x22· · · ∂2f

∂x2∂xn...

.... . .

...

∂2f

∂xn∂x1

∂2f

∂xn∂x2· · · ∂2f

∂x2n

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠.

If the function f is continuously differentiable to order 2, then the order of differenti-ation does not matter: ∂2f/∂xi∂xj = ∂2f/∂xj∂xi for all i, j = 1, . . . , n. For order 2continuously differentiable functions, then, the Hessian matrices are symmetric.

2.18 Quadratic Approximations of Functions

Why would you want to use a quadratic function to approximate another function?Once you achieve a clear understanding of how to solve constrained optimizationproblems in which the constraint and objective functions are all differentiable, youwill realize that it is all about using quadratic functions as local approximations ofdifferentiable functions, so it is important that you understand this section clearly.Whether you realize it or not, you have already done this many times before. Alarge part of the chapter on constrained optimization is concerned with helping youto realize this and then to understand why it makes sense to do it. The central idea,used repeatedly in later chapters, is a special case of the famous “Taylor’s Theorem”published in 1715 by Brook Taylor. Oddly enough, this important result was ignoredfor another 57 years until the French mathematician Joseph Lagrange drew attentionto it in 1772. Taylor did not give a complete proof of his result; that was provided in1841 by another famous French mathematician, Augustin Cauchy. Another point ofinterest is that various special cases of Taylor’s result were known well before Taylor’spublication of his more general statement.

I will give you an example that I hope will make clear why quadratic approxima-tions are so useful for solving differentiable constrained optimization problems. Hereis a simple function:

f(x) = −50x+ x3 + 100 ln (1 + x) for x ∈ [0, 4]. (2.33)

Page 91: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

72 CHAPTER 2. BASICS

Figure 2.27: The graph of f(x) = −50x+ x3 + 100 ln (1 + x) for 0 ≤ x ≤ 4.

This function is not even close to being a quadratic, which is why I chose it. Fig-ure 2.27 displays the graph of the function for 0 ≤ x ≤ 4. The function has alocal maximum at x ≈ 1·18350342. Taylor’s Theorem tells us that a particularquadratic function is a good approximation to f(1·18350342 + Δx) for values of Δxthat are close to zero; i.e. for points close to x = 1·18350342. The quadratic isg(Δx) = a(Δx)2+ bΔx+ c, where c is the value of f at x = 1·18350342, b is the valueof the first-order derivative of f at x = 1·18350342, and a is one-half of the value ofthe second-order derivative of f at x = 1·18350342. That is,

c = f(x = 1·18350342) ≈ 20·575605,

b =df(x)

dx

∣∣x=1·18350342 =

(−50 + 3x2 +

100

1 + x

) ∣∣∣∣x=1·183503419

≈ 0,

and a =1

2

d2f(x)

dx2∣∣x=1·18350342 =

1

2

(6x− 100

(1 + x)2

) ∣∣∣∣x=1·183503419

≈ −6·936754975.

So the approximating quadratic is

g(Δx) = 20·575605− 6·936754975(Δx)2

for small values, both positive and negative, of Δx that are close to zero.

Page 92: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.18. QUADRATIC APPROXIMATIONS OF FUNCTIONS 73

Figure 2.28: The function f and the approximating quadratic g near to x = 1·1835.

Figure 2.28 shows the graphs of f and g for values of x in a small neighborhoodabout x = 1·18350342. For any small enough neighborhood about x = 1·18350342,such as that indicated by the double-ended arrow, the two graphs are almost thesame; the quadratic g is a close approximation to the function f . Following thesame procedure will give us a close quadratic approximation to f at any value ofx ∈ [0, 4], not just at a local maximum. So if, for example, we selected x = 2, apoint that is neither a local maximum nor a local minimum for f (see Figure 2.27),then we can use the above procedure to construct a new quadratic function that isa close approximation to the function f for values of x that are close to x = 2. Thepoint to grasp, then, is that at any point in the domain of any twice-continuouslydifferentiable function f there exists a quadratic function that is almost exactly thesame as the differentiable function f over a small enough neighborhood of domainvalues around the chosen point. This is the result that is the second-order version ofTaylor’s Theorem.

Page 93: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

74 CHAPTER 2. BASICS

So let’s see the second-order version of Taylor’s Theorem. I will introduce it intwo steps. The first time around, we will consider only functions that map from thereal line. The second pass will consider functions that map from �n. (There are manyother books in which you can read the general version of the theorem and see it inits full glory.)

Theorem 2.12 (Second-Order Taylor’s Theorem for Functions Mapping from �to �). Let S ⊆ � be a nonempty set that is open with respect to the Euclideantopology on �. Let f : S �→ � be a C2(S) function. Let x′, x′ +Δx ∈ S. Then thereexists a function R2(x

′,Δx) ∈ C2(S) such that

f(x′ +Δx) = f(x′) +df(x)

dx|x=x′Δx+

1

2

d2f(x)

dx2|x=x′ (Δx)2 +R2(x

′,Δx), (2.34)

whereR2(x

′,Δx)(Δx)2

→ 0 as Δx→ 0. (2.35)

The first three terms on the right-hand side of (2.34) are the Taylor quadratic

Q(x′,Δx) = f(x′) +df(x)

dx|x=x′Δx+

1

2

d2f(x)

dx2|x=x′ (Δx)2 (2.36)

that is the quadratic approximation to the function f at and nearby to x = x′. Theapproximation error at a point x = x′+Δx is the difference in the values at x = x′+Δxof f and the approximation Q(x′,Δx). This error is the second-order Taylor’s seriesremainder term R2(x

′,Δx). (2.35) states that as Δx → 0 the remainder term’scontribution to the right-hand side of (2.34) becomes negligibly small, making theTaylor quadratic an evermore accurate approximation to the function f at valuesx = x′ +Δx that are ever closer to x = x′. This is what we discovered in our aboveexample.

Now let’s see Taylor’s Theorem for functions that map from �n to �.

Theorem 2.13 (Second-Order Taylor’s Theorem for Functions Mapping from �n

to �). Let S ⊆ �n be a nonempty set that is open with respect to the Euclideanmetric topology on �n. Let f : S �→ � be a C2(S) function. Let x′ ∈ S. Let B(x′, ε)be an open ball centered at x′ with radius ε > 0 in S. Let x′ + Δx ∈ B(x′, ε). Thenthere exists a function R2(x

′,Δx) ∈ C2(S) such that, for every x′ +Δx ∈ B(x′, ε),

f(x′ +Δx) = f(x′) +∇f(x)|x=x′ ·Δx+ 1

2ΔxTHf (x)|x=x′Δx+R2(x

′,Δx), (2.37)

Page 94: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.18. QUADRATIC APPROXIMATIONS OF FUNCTIONS 75

whereR2(x

′,Δx)‖Δx‖2E

→ 0 as ‖Δx‖E → 0. (2.38)

The sum of the first three terms of (2.37) is again the Taylor quadratic functionof Δx = (Δx1, . . . ,Δxn) that more and more accurately approximates the functionf on the ball centered at x = x′ with radius ε as Δx approaches the zero vector.

Let’s have a look at (2.37) for the function f(x1, x2) = x1/21 +x1x2+ln (1 + x1 + x2).

The gradient of f is the row vector

∇f(x1, x2) =(∂f(x1, x2)

∂x1,∂f(x1, x2)

∂x2

)

=

(1

2x−1/21 + x2 +

1

1 + x1 + x2, x1 +

1

1 + x1 + x2

),

and the Hessian matrix of f is

Hf (x1, x2) =

⎛⎜⎜⎜⎝− 1

4x−3/21 − 1

(1 + x1 + x2)21− 1

(1 + x1 + x2)2

1− 1

(1 + x1 + x2)2− 1

(1 + x1 + x2)2

⎞⎟⎟⎟⎠ .

Now let’s pick a particular point (x1, x2), say (x1, x2) = (1, 1). At this point thevalues of the function, its gradient vector, and its Hessian matrix are

f(1, 1) = 2 + ln 3 ≈ 3·099, ∇f(1, 1) =(11

6,4

3

)and Hf (1, 1) =

⎛⎜⎜⎜⎝− 13

36

8

9

8

9− 1

9

⎞⎟⎟⎟⎠ .

So, evaluated at the point (x1, x2) = (1, 1), the Taylor quadratic is

Q(Δx1,Δx2) = f(1, 1) +∇f(1, 1)·(Δx1,Δx2) +1

2(Δx1,Δx2)Hf (1, 1)

(Δx1Δx2

)

= 2 + ln 3 +

(11

6,4

3

)(Δx1Δx2

)+

1

2(Δx1,Δx2)

⎛⎜⎜⎜⎝− 13

36

8

9

8

9− 1

9

⎞⎟⎟⎟⎠(Δx1Δx2

)

= 2 + ln 3 +11

6Δx1 +

4

3Δx2 −

13

72(Δx1)

2 +8

9Δx1Δx2 −

1

18(Δx2)

2

Page 95: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

76 CHAPTER 2. BASICS

Table 2.2: Various evaluations of f and Q about (x1, x2) = (1, 1).

(x1, x2) (Δx1,Δx2) f(x1, x2) Q(Δx1,Δx2) |f(x1, x2)−Q(Δx1,Δx2)|(1·00, 1·00) (0·00, 0·00) 3·09861 3·09861 0·00000(1·00, 0·98) (0·00,−0·02) 3·07192 3·07192 0·00000(0·98, 0·98) (−0·02,−0·02) 3·03554 3·03554 0·00000(1·02, 1·02) (0·02, 0·02) 3·16221 3·16221 0·00000(1·05, 0·95) (0·05,−0·05) 3·12081 3·12080 0·00001(1·05, 1·05) (0·05, 0·05) 3·25860 3·25858 0·00002(0·90, 0·90) (−0·10,−0·10) 2·78830 2·78847 0·00017(1·10, 1·10) (0·10, 0·10) 3·42196 3·42181 0·00015(0·50, 1·50) (−0·50, 0·50) 2·55572 2·56736 0·01164(1·50, 1·50) (0·50, 0·50) 4·86104 4·84514 0·01590(0·00, 3·00) (−1·00, 2·00) 1·38629 1·75139 0·36510(3·00, 3·00) (2·00, 2·00) 12·67796 12·04306 0·63490

where Δx1 and Δx2 are deviations from the values x1 = 1 and x2 = 1, respectively.Table 2.2 displays the values of the function f and the Taylor quadratic for variouspoints (x1, x2) = (1+Δx1, 1+Δx2) around the point (x1, x2) = (1, 1). The quadraticapproximation is accurate for values of (x1, x2) that are close to (x1, x2) = (1, 1).

2.19 Saddle-Points

A saddle-point is a particular type of critical point for a function of more than onevariable. A saddle-point is a point in the domain of the function that is a blend ofa local maximum and a local minimum in that, in some xi direction(s), the functionis locally maximized at the saddle-point, while in all of the other xj direction(s) thefunction is locally minimized at the saddle-point. It is easiest to get the idea of asaddle-point if we first consider a function f(x1, x2) of just two variables.

Let S ⊂ �2 be some nonempty neighborhood. A point (x1, x2) = (x′1, x′2) is a

saddle-point in S for a function f(x1, x2) if

f(x′1; x′2) = max

(x1, x′2)∈S

f(x1; x′2) and f(x

′2; x

′1) = min

(x′1, x2)∈S

f(x2; x′1).

Page 96: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.19. SADDLE-POINTS 77

That is,

f(x1; x′2) ≤ f(x′1; x

′2) for all (x1, x

′2) ∈ S and f(x′2; x

′1) ≤ f(x2; x

′1) for all (x

′1, x2) ∈ S.

Let’s read this carefully. The most important thing to notice is that two differentrestricted versions of the function f(x1, x2) are being considered. One of these isthe function of x1 only that is f(x1; x

′2); i.e. the function f(x1, x2) in which the

value of x2 is fixed at x2 = x′2. The other restricted function is f(x2; x′1); i.e. the

function f(x1, x2) in which the value of x1 is fixed at x1 = x′1. The point (x′1, x

′2) is a

saddle-point for the unrestricted function f(x1, x2) if the restricted function f(x1; x′2)

is maximized at x1 = x′1 and, also, the restricted function f(x2; x′1) is minimized at

x2 = x′2. The usual example of a function with a saddle-point is f(x1, x2) = x21 − x22.This function has only one critical point, at (x1, x2) = (0, 0), and it is a saddle-point, as may be seen in Figure 2.29. The function f(x1, x2) = x21 − x22 restricted byx2 = 0 is f(x1; x2 = 0) = x21. This function is minimized at x1 = 0. The functionf(x1, x2) = x21 − x22 restricted by x1 = 0 is f(x2; x1 = 0) = −x22. This function ismaximized at x2 = 0. Thus,

f(x2; x1=0) = −x22 ≤ f(x1=0, x2=0) = 0 ≤ f(x1; x2=0) = x21 ∀ (x1, x2) ∈ �2

and so (x1, x2) = (0, 0) is a saddle-point for f(x1, x2) = x21 − x22.

Figure 2.29: f(x1, x2) = x21 − x22 has a saddle-point at (x1, x2) = (0, 0).

Once we consider functions of more than two variables, the number of possibletypes of saddle-points increases rather rapidly. Nevertheless, the essential charac-teristic of any saddle-point possessed by some function is that at the saddle-point

Page 97: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

78 CHAPTER 2. BASICS

the function is locally maximized with respect to some of its variables and is locallyminimized with respect to the rest of its variables.

Definition 2.16 (Saddle-point). Let X ⊂ �n and Y ⊂ �m. Let the function f :X × Y �→ �. Then (x′, y′) ∈ X × Y is a saddle-point for f on X × Y if

f(x; y′) ≤ f(x′, y′) ∀ x ∈ X and f(x′, y′) ≤ f(y; x′) ∀ y ∈ Y.

Figure 2.30: f(x1, x2) = (x2 − y2)/ex2+y2 has a saddle-point at (x1, x2) = (0, 0), and

also two local minima and two local maxima in the set [−2, 2]× [−2, 2].

If the function is differentiable and if the saddle-point is in the interior of thefunction’s domain, then, necessarily, the value at the saddle-point of the gradientvector of the function is a zero vector, making the saddle-point a critical point. Wealready know that a function can possess more than one critical point, so it shouldnot surprise you that a function may have one or more saddle-points along with othertypes of critical points, such as local minima or local maxima. Figure 2.30 displaysthe graph of a function with five critical points, one of which is a saddle-point, two ofwhich are local minima, and two of which are local maxima. As the right-side panelsof Figures 2.29 and 2.30 show, it is sometimes easier graphically to locate a function’ssaddle-point by examining a contour plot of the function. A local maximum or localminimum displays as a point in the “middle” of concentric contours. A saddle-pointdisplays as a point towards which the contours “bulge.”

Page 98: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.20. METRICS 79

2.20 Metrics

How often have you measured the distance between two points? Have you ever neededto cut a cookie exactly in half? Have you ever wanted to know how high something isabove the ground? The thing common to these activities is distance. You want yourknife to cut at half of the distance across the cookie. You want to know the distanceof the something from the ground. Metric spaces are sets of points with the propertythat you can measure how distant any point in the set is from any other point in theset. You have used this idea each day since you were a child. In economics, we usemetric spaces all the time. Most times the space is the familiar set of real points withdistances between points measured by Pythagoras’s Theorem. But there are othermetric spaces that are much less familiar and yet are important to economists, andothers. So let us explore the ideas of a metric function and a metric space. You willquickly see their practical values.

It’s very simple. A metric is a function that measures distances between points.For example, the straight-line distance, measured in, say, kilometers, between Town Aand Town B might be called the “as the crow flies” distance between the towns. Thedistance is positive, and it is the same whether we measure it from Town A to Town Bor from Town B to Town A. The number of kilometers covered when driving a carbetween the same two towns might be called the “driving distance” between thetowns. Once again the distance is positive, and is the same whether we measure itfrom Town A to Town B or in the reverse direction. But the “as the crow flies”distance and the “driving distance” between the towns will not be the same unlessthe road between them is exactly straight. This is a simple example of the fact thatthere is often more than one reasonable way to measure a distance.

Although these are familiar examples for all of us, the idea of measuring distancesbetween points is an idea of wide applicability. We can, for example, measure dis-tances between colors, between sounds, or between quadratics. Let’s start by definingthe term metric. The usual symbol for a metric function is d, short for d istance.

Definition 2.17 (Metric). A metric defined on a set X is a function d : X×X �→ �+

with the following properties:

(i) nonnegativity: for all x′, x′′ ∈ X, d(x′, x′′) ≥ 0 with d(x′, x′′) = 0 if and only ifx′ = x′′;

(ii) symmetry: for all x′, x′′ ∈ X, d(x′, x′′) = d(x′′, x′);(iii) triangular inequality: for all x′, x′′, x′′′ ∈ X, d(x′, x′′′) ≤ d(x′, x′′) + d(x′′, x′′′).

Page 99: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

80 CHAPTER 2. BASICS

(i) says that distance is never negative, that the distance of any point from itself iszero, and that the distance between any two different points is strictly positive. (ii)says that the direction in which distance is measured does not matter; the distancefrom x′ to x′′ is the same as the distance from x′′ to x′. (iii) says that the “direct”distance between any two points x′ and x′′′ is never larger than the distance betweenx′ and any other point x′′ added to the distance between x′′ and x′′′. For example,it is never a shorter trip from Town A to Town B via Town C than it is to traveldirectly from Town A to Town B.

Let’s look at ways of measuring the distances between points in the real line, �.The obvious, and certainly the most commonly used, metric is the Euclidean distancefunction dE defined on �× �:

∀ x′, x′′ ∈ �, dE(x′, x′′) = |x′ − x′′|. (2.39)

Now let’s think about measuring distances between points in the real plane, �2. Onceagain, the most commonly used metric is the Euclidean distance function dE, which,defined on �2 ×�2, is

∀ x′, x′′ ∈ �2, dE(x′, x′′) = +

√(x′1 − x′′1)2 + (x′2 − x′′2)2. (2.40)

This is the famous distance measure devised by Pythagoras. But there are reason-able alternative ways of measuring distances between points in �2. One of thesealternatives is the metric function

d0(x′, x′′) = |x′1 − x′′1|+ |x′2 − x′′2|. (2.41)

Yet another is the metric function

d1(x′, x′′) = max {|x′1 − x′′1|, |x′2 − x′′2|}. (2.42)

All of the metrics (2.40), (2.41), and (2.42) may be adapted to measure distancesbetween points in �n for any n ≥ 1. Notice that for n = 1 all three metrics are thesame function.

How can we use metrics to measure distances between sounds and colors? That’seasy. Just measure the distance between the wave frequencies of two sounds or of twocolors.

How might we measure the distance between two continuous real-valued functionsdefined on an interval [a, b]? Two possibilities are displayed in Figure 2.31. One is

Page 100: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.20. METRICS 81

Figure 2.31: Two metrics that measure the distance between the functions f and g.

the function known as the sup metric:

d∞(f, g) = supx∈[a,b]

|f(x)− g(x)|. (2.43)

This, the least upper bound of the differences |f(x)−g(x)| over all of [a, b], is displayedon the left-hand axis of Figure 2.31. The other distance measure is the (shaded) areabetween the graphs of the functions f and g:

d2(f, g) =

∫ b

a

|f(x)− g(x)| dx. (2.44)

In both cases the distance between f and g is zero if and only if the two continuousfunctions are identical on [a, b]. If f and g are continuous but not identical, thenboth metrics show a strictly positive distance between the functions. Both metricshave the symmetry property. Both also have the triangular property (prove this asan exercise – it’s simple to do).

The primary uses that we shall make of metrics are, first, in determining whensequences converge or not, and, second, to define small neighborhoods around pointsof interest to us. These neighborhoods are called balls. You have often used them.

Page 101: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

82 CHAPTER 2. BASICS

For example, whenever you have said something like “consider a small open interval(x− ε, x+ ε) centered at a point x” you have in fact been talking of the ball of radiusε that is centered at the point x.

The pair (X, d) that consists of a set X, together with a metric function d definedon X ×X, is called a metric space. I’m sure that you already comprehend that thereare lots of metrics and lots of metric spaces.

For economists the most useful metric spaces are the Euclidean metric spaces.The pair (�, dE), consisting of the real line � together with the Euclidean metricfunction (2.39) defined on �×�, is the one-dimensional Euclidean metric space andis denoted by E1 ≡ (�, dE). The pair (�2, dE), consisting of the two-dimensional realspace �2 together with the Euclidean metric function (2.40) defined on �2 × �2, isthe two-dimensional Euclidean metric space and is denoted by E2 ≡ (�2, dE). Then-dimensional Euclidean space is En ≡ (�n, dE), where the Euclidean metric functiondefined on �n ×�n is

dE(x′, x′′) = +

√(x′1 − x′′1)2 + · · ·+ (x′n − x′′n)2 ∀ x′, x′′ ∈ �n.

2.21 Norms

A norm is a function that measures the distances of points in a space from theparticular point that is the “origin” of the space. A norm is usually denoted by thesymbol ‖·‖. The origin on the real line is the number zero, so the Euclidean norm onthe real line is, from (2.39),

‖x‖E = dE(x, 0) = |x− 0| = |x|. (2.45)

The origin in �2 is the point (0, 0), so the Euclidean norm in �2 is, from (2.40),

‖x‖E = dE(x, 0) =√(x1 − 0)2 + (x2 − 0)2 =

√x21 + x22 . (2.46)

The origin in the space of all real-valued continuous functions defined on [a, b] is thefunction o(x) ≡ 0, so we can use (2.44) to define the distance of a function f definedon [a, b] from the origin function o as

‖f‖ = d2(f, o) =

∫ b

a

|f(x)− o(x)| dx =

∫ b

a

|f(x)| dx.

I hope it is clear that, if we have a metric space (X, d) and the space X has an“origin” element, then we can define a norm function ‖·‖ on this space just by using

Page 102: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.22. HOW TO PROCEED FROM HERE 83

the metric function d to measure distances from only the origin element in X. Thepair (X, ‖·‖) that consists of the set X together with the norm ‖·‖ defined on X iscalled a normed space. For example, the n-dimensional Euclidean normed space is(�n, ‖·‖E), where

‖x‖E = +√x21 + · · ·+ x2n ∀ x ∈ �n.

2.22 How to Proceed from Here

Chapter 8 is the first chapter that directly deals with solving constrained optimizationproblems. Its explanations rely upon your having a reasonable understanding ofthe contents of Chapters 6 and 7. Chapter 6 introduces hyperplanes and how theymay be used to separate certain types of sets from one another. This material isgeometrically quite obvious and you may skip to this chapter directly if you arewilling to forgo the ability to follow some of the more detailed arguments and proofs.Chapter 7 introduces cones and convex cones. These are simple but very useful typesof sets. Sometimes we can use a hyperplane to separate a convex cone from anotherconvex cone, and sometimes not. When we can, and cannot, is at the very heart ofconstrained optimization theory. Understanding Chapter 7 thus requires a reasonableprior understanding of Chapter 6. So, in short, one reasonable strategy for proceedingfrom this point is to study just Chapters 6 and 7 prior to studying Chapter 8. Youwill get a useful, but somewhat incomplete, understanding of the basics of constrainedoptimization theory. You will also be able to proceed through Chapter 9 and a lot,but not all, of Chapter 10. For many readers, this may be a satisfactory outcome.

A second reasonable strategy for proceeding from this point is to jump to Chap-ter 6 and read until you encounter an idea that is new to you and that you wish tounderstand. You can then go to the appropriate part of an earlier chapter, gain theknowledge you need, and then return to where you were originally reading.

If you want to understand everything in detail, then simply read on. Enjoy.

Page 103: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

84 CHAPTER 2. BASICS

2.23 Problems

Problem 2.1.

A =

⎛⎜⎜⎝

2 0 −2 3−5 3 −1 51 0 7 −36 −2 3 3

⎞⎟⎟⎠ .

(i) What is the value of the determinant of the matrix A?(ii)

S =

⎧⎪⎪⎨⎪⎪⎩⎛⎜⎜⎝

2−516

⎞⎟⎟⎠ ,

⎛⎜⎜⎝

030−2

⎞⎟⎟⎠ ,

⎛⎜⎜⎝−2−173

⎞⎟⎟⎠ ,

⎛⎜⎜⎝

35−33

⎞⎟⎟⎠⎫⎪⎪⎬⎪⎪⎭ .

Is S a set of linearly independent vectors?

(iii) Can the vector

(307−651999−712

)be generated as a linear combination of the vectors in S?

(iv) Does the equation system

2x1 − 2x3 + 3x4 = b1

−5x1 + 3x2 − x3 + 5x4 = b2

x1 + 7x3 − 3x4 = b3

6x1 − 2x2 + 3x3 + 3x4 = b4

have a solution for any values for b1, b2, b3, and b4? If not, then why not? If so, thenhow many solutions are there?

Problem 2.2.

B =

⎛⎝ 2 0 −2−10 6 −176 −2 3

⎞⎠ .

(i) What is the value of the determinant of the matrix B?(ii)

V =

⎧⎨⎩⎛⎝ 2−106

⎞⎠ ,

⎛⎝ 0

6−2

⎞⎠ ,

⎛⎝ −2−173

⎞⎠⎫⎬⎭ .

Is V a set of linearly independent vectors?(iii) What is the set L(V ) of vectors in �3 that is spanned by V ?

Page 104: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.23. PROBLEMS 85

(iv) Does either of the equation systems

2x1 − 2x3 = 2

−10x1 + 6x2 − 17x3 = 2

6x1 − 2x2 + 3x3 = 2

and

2x1 − 2x3 = 2

−10x1 + 6x2 − 17x3 = −2

6x1 − 2x2 + 3x3 = 2

have a solution? If not, why not? If so, how many solutions are there?

Problem 2.3. Draw a diagram that displays the vectors v1 = ( 41 ) and v2 = ( −4

3 ).Then, for any x1 ≥ 0 and for any x2 ≥ 0, display the vectors x1v

1 and x2v2. Next,

display the vector x1v1 + x2v

2. On your diagram, where is the region that containsall of the vectors that are linear combinations x1v

1 + x2v2 with x1 ≥ 0 and x2 ≥ 0?

Use a similar method to discover the region that contains all of the vectors thatare linear combinations y1v

1 + y2v2 with y1 ≤ 0 and y2 ≤ 0.

Problem 2.4. Review Section 2.2 to remind yourself of what is meant by the term“vector space.” You know already that the real space �n is a vector space, but whatabout some other spaces?

(i) Let P be the space of all polynomial functions defined on �. For example, allquadratic and cubic functions defined on � belong to this space. Is P a vector space?

(ii) LetW be the space of all functions defined on � that are bounded by the constantM > 0; i.e. a function f ∈ W only if −M ≤ f(x) ≤ +M for every x ∈ �. Is W avector space?

Problem 2.5. A function f : �n → � of the form f(x) = a1x1 + · · · + anxn + b,where a1, . . . , an, b ∈ �, is an affine function defined on �n. A function g : �n → �of the form g(x) = a1x1 + · · · anxn, where a1, . . . , an ∈ �, is a linear function definedon �n. Let A(�n) be the space of all affine functions defined on �n. Let L(�n) bethe space of all linear functions defined on �n.

(i) Is A(�n) a vector space?

(ii) Is L(�n) a vector space?

(iii) Consider the singleton subset S = {0} ⊂ �n. Is S a vector space?

Definition 2.18 (Subspace). A set V ′ is a subspace of a vector space V if and onlyif

(a) V ′ ⊆ V , and

(b) x, y ∈ V ′ implies that ax+ by ∈ V ′ for all a, b ∈ �.

Page 105: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

86 CHAPTER 2. BASICS

A subspace of a vector space V is often called a linear manifold of V . If V ′ is asubspace of V and V ′ is strictly contained in V , then V ′ is a proper subspace of V .

(iv) Is the vector space V a subspace of itself? Is it a proper subspace of itself?(v) Is L(�n) a proper subspace of A(�n)?

Problem 2.6. S is a proper subspace of �n. What type of set must S be, other thanmerely the singleton set consisting of only the origin?

Problem 2.7. X = {−3,−2,−1, 0, 1, 2, 3}. S1 = {1, 2, 3}. S2 = {−3,−2,−1}. Trueor false: S1 + S2 = {0}?Problem 2.8. X = �2. S1 = {(x1, x2) ∈ �2 | 1 ≤ x1 ≤ 3, x2 = 6}. S2 ={(2, 3), (4, 1)}. What is the set S1 + S2?

Problem 2.9. Sum the two sets displayed in Figure 2.9.

Problem 2.10. X = �2. S1 = {(x1, x2) ∈ �2 | x2 = −x1, −1 ≤ x1 ≤ 0}.S2 = {(x1, x2) ∈ �2 | x2 = −x1, 0 ≤ x1 ≤ 1}. What is the set S1 + S2?

Problem 2.11. X = �2. S1 = {(x1, x2) ∈ �2 | x2 = −x1}. S2 = {(x1, x2) ∈ �2 |x2 = x1}. What is the set S1 + S2?

Problem 2.12. S1 = {(x1, x2) ∈ �2 | x2 = −x1, −1 ≤ x1 ≤ 0}. S2 = {x ∈ � | 0 ≤x ≤ 1}. What are the sets S1 × S2 and S2 × S1?

Problem 2.13. Prove that if S1 and S2 are convex subsets of a set X, then S1 ∩ S2

is a convex subset of X.

Problem 2.14. Use the result proved in problem 2.13, together with forward induc-tion, to prove that, if, for any finite k ≥ 2, S1, . . . , Sk are all convex subsets of X,then S1 ∩ · · · ∩ Sk is a convex subset of X.

Problem 2.15. Suppose that S1, . . . , Sn are all strictly convex subsets of a set X.Then as you have seen in the above theorem the intersection of these subsets is a setthat must be convex in X. Give two examples to show that this intersection neednot be strictly convex in X.

Problem 2.16. Let S1 and S2 be convex subsets of �n. Prove that S1 + S2 is aconvex subset of �n.

Problem 2.17. Let S1, . . . , Sk be convex subsets of �n, for finite k ≥ 2. Use the resultobtained in problem 2.16 together with forward induction to prove that S1 + · · ·+Sk

is a convex subset of �n.

Page 106: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.23. PROBLEMS 87

Problem 2.18. Prove the following theorem.

Theorem. Let S1, . . . , Sn be any finite collection of convex subsets of a set X.Then S1 × · · · × Sn is a convex subset of Xn = X × · · · ×X︸ ︷︷ ︸

n times

.

Problem 2.19. Let S1 and S2 be strictly convex subsets of �n. Prove that S1+S2 isa strictly convex subset of �n. Use this result to prove the more general result that,if S1, . . . , Sk for k ≥ 2 are all strictly convex subsets of �n, then S1 + · · · + Sk is astrictly convex subset of �n.

Problem 2.20. In Section 2.6 we encountered examples in which the direct productsof two strictly convex sets are sets that are only weakly convex. This may haveinstilled the impression that the direct product of two convex sets is never a strictlyconvex set. But, in fact, the direct product of two convex sets can be (does not haveto be) a strictly convex set. Provide an example in which the direct product of twostrictly convex sets is a strictly convex set.

Problem 2.21. Consider the function f(x1, x2) = x1(1 + x2) defined on �2.

(i) What is the value of the gradient vector of f at the point (1, 9)?

(ii) What is the contour set for f ≡ 10?

(iii) Consider the f ≡ 10 contour set at the point (1, 9). Define the marginal rateof substitution of x1 for x2, MRS(x1, x2), as the differential rate of change of x2 withrespect to x1 along the contour set at the point (x1, x2). What is the value of thismarginal rate of substitution at the point (1, 9)?

(iv) Construct the tangent to the f ≡ 10 contour set at the point (1, 9).

(v) What is the gradient vector of this tangent?

(vi) Show that the tangent and the gradient vector of the f ≡ 10 contour set at thepoint (1, 9) are orthogonal.

Problem 2.22. Consider the following matrices:

A =

(4 12

−16 1

), B =

(7 −2−4 1

), C =

(0 aa 0

), D =

⎛⎝ 1 −a ba 1 −c−b c 1

⎞⎠ ,

where a, b, and c are arbitrary real numbers.

(i) Write out the quadratic form of the matrix A as a quadratic equation in x1 and x2.Then use the method of “completion of squares” to determine from this equation ifthe matrix A has a particular definiteness property.

Page 107: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

88 CHAPTER 2. BASICS

(ii) Write out the quadratic form of the matrix B as a quadratic equation in x1and x2. Then use the method of “completion of squares” to determine from thisequation if the matrix B has a particular definiteness property.(iii) Does the matrix C have a particular definiteness property?(iv) Does the matrix D have a particular definiteness property?

Problem 2.23. Consider the function f : �× � → � that is

f(x, y) =

{0 , if x = y

1 , if x = yfor all x, y ∈ �.

Is f a metric on �× �?

Problem 2.24. Let the function d : X ×X → � be a metric on X ×X.(i) Let the function f : X × X → � be defined by f(x, y) = ad(x, y) + b, wherea, b ∈ � with a > 0. Is f a metric on X ×X?(ii) Let g : X ×X → � be defined by g(x, y) = d(x, y)2. Is g a metric on X ×X?(iii) Let h : X×X → � be defined by h(x, y) = +

√d(x, y). Is h a metric on X×X?

Problem 2.25. Consider the function f(x1, x2) = 2+x1−x2 defined on [0, 1]2. Doesthis function have a saddle-point? If so, is the saddle-point a critical point for thefunction?

Problem 2.26. Consider the function f(x1, x2) = (x21 − x22)e−(x2

1+x22) defined on �2.

Use the restricted functions f(x1; x2) and f(x2; x1) to prove that (x1, x2) = (0, 0) is asaddle-point for f . Is this saddle-point also a critical point for f?

Problem 2.27. Let f(x1, x2) = x31 − x21x2 + x32. Construct the Taylor quadraticQ(2 + Δx1, 1 + Δx2) approximation to the function f at the point (x1, x2) = (2, 1).Then evaluate the function f and the quadratic for the points (x1, x2) = (2, 1),(x1, x2) = (2·1, 1·1), (x1, x2) = (1·9, 0·9) (x1, x2) = (2·5, 1·5), (x1, x2) = (3, 2), and(x1, x2) = (1, 3) to get an idea of how well, or poorly, the quadratic approximates f .

Page 108: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.24. ANSWERS 89

2.24 Answers

Answer to Problem 2.1.(i) It is easiest if we expand down the second column of A. Doing so gives

det (A) = (−1)2+2(3)M22 + (−1)4+2(−2)M42 = 3M22 − 2M42.

M22 = det

⎛⎝2 −2 31 7 −36 3 3

⎞⎠

= (−1)1+1(2) det

(7 −33 3

)+ (−1)2+1(1) det

(−2 33 3

)+ (−1)3+1(6) det

(−2 37 −3

)= 2(21 + 9)− (−6− 9) + 6(6− 21) = −15.

M42 = det

⎛⎝ 2 −2 3−5 −1 51 7 −3

⎞⎠

= (−1)1+1(2) det

(−1 57 −3

)+ (−1)2+1(−5) det

(−2 37 −3

)+ (−1)3+1(1) det

(−2 3−1 5

)= 2(3− 35) + 5(6− 21) + (−10 + 3) = −146.

So,det (A) = 3× (−15)− 2× (−146) = 247.

(ii) Yes. det (A) = 0 only if S is a set of linearly independent vectors.(iii) Yes. S is a set of four linearly independent vectors of dimension 4, so S is a basisfor all of �4. That is, any vector in �4 can be generated as a linear combination ofthe vectors in S.(iv) Asking if the equation system has a solution is equivalent to asking if the vector(

b1b2b3b4

)can be generated as a linear combination of the vectors in S. We already know

this to be true, so the equation system has at least one solution. Since the rank ofthe matrix A is 4 (thus, A has full rank), there is only a unique solution.

Answer to Problem 2.2.(i) The arithmetic required is least if we expand across the first row of B. Doing sogives

det (B) = (−1)1+1(2)M11 + (−1)1+3(−2)M13 = 2M11 − 2M13

= 2det

(6 −17−2 3

)− 2 det

(−10 66 −2

)= 2(18− 34)− 2(20− 36) = 0.

Page 109: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

90 CHAPTER 2. BASICS

(ii) No. det (B) = 0 only if V is a set of linearly dependent vectors. For example,

−(

2−106

)− 9

2

(06−2

)=

( −2−173

).

(iii) V is a set of three linearly dependent vectors of dimension 3, so V is not a

basis for all of �3. Any vector

(b1b2b3

)contained in L(V ) can be generated as a

linear combination of any two of the vectors in V , so let us consider all of the linear

combinations of the vectors

(2

−106

)and

(06−2

). Thus a vector

(b1b2b3

)is contained in

L(V ) if and only if there are weights x1 and x2 for which

x1

⎛⎝ 2−106

⎞⎠+ x2

⎛⎝ 0

6−2

⎞⎠ =

⎛⎝ 2x1−10x1 + 6x26x1 − 2x2

⎞⎠ =

⎛⎝b1b2b3

⎞⎠ .

This requires

x1 =b12, −10x1 + 6x2 = −5b1 + 6x2 = b2 ⇒ x2 =

1

6(5b1 + b2)

and 6x1 − 2x2 = 3b1 −1

3(5b1 + b2) =

1

3(4b1 − b2) = b3.

Thus the vectors in L(V ), the vectors

(b1b2b3

)that can be generated as linear combi-

nations of

(2

−106

)and

(06−2

), are the vectors in which the elements b1, b2, and b3

satisfy the restriction 4b1 − b2 − 3b3 = 0. This is the equation of only a plane in �3.

For example, the vector

(222

)∈ L(V ), while the vector

(2−22

)/∈ L(V ).

(iv) The equation system

2x1 − 2x3 = 2

−10x1 + 6x2 − 17x3 = 2

6x1 − 2x2 + 3x3 = 2

has a solution if and only if the vector

(222

)can be generated as a linear combination

of the vectors

(2

−106

)and

(06−2

). We showed this to be true in part (iii), so the

equation system has at least one solution. We also know, from part (i), that the

Page 110: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.24. ANSWERS 91

matrix B has less than full rank, r(B) = 2 < 3, so the equation system has infinitelymany solutions. Check for yourself that the complete set of solutions is {(x1, x2, x3) |x1 = 1 + x3, x2 = 2 + 9x3/2, −∞ < x3 <∞}.

The equation system

2x1 − 2x3 = 2

−10x1 + 6x2 − 17x3 = −2

6x1 − 2x2 + 3x3 = 2

has a solution if and only if the vector

(2−22

)can be generated as a linear combination

of the vectors

(2

−106

)and

(06−2

). In part (iii) we showed this is not so. Trying to

solve the system produces two inconsistent equations.

Answer to Problem 2.3. Look at Figure 2.32. The figure displays the vectors v1 =( 41 ), v2 = ( −4

3 ), the finely dotted line from the origin (a ray) containing nonnegativemultiples x1v

1 of v1 and the finely dotted ray containing nonnegative multiples x2v2

of v2. The sum of a nonnegative multiple x1v1 of v1 and a nonnegative multiple x2v

2

of v2 necessarily lies within the region in Figure 2.32 that is bounded by the raycontaining v1 and the ray containing v2. If x1 = x2 = 0, then the sum x1v

1 + x2v2 is

the origin vector ( 00 ). If x1 = 0 and x2 > 0, then the sum x1v1+x2v2 = x2v

2, a vectorin the ray containing v2. If x1 > 0 and x2 = 0, then the sum x1v

1 + x2v2 = x1v

1, avector in the ray containing v1. If x1 > 0 and x2 > 0, then the sum x1v

1 + x2v2 is a

vector that lies strictly inside the region bounded by the rays containing v1 and v2.Thus the region containing all of the sums x1v

1 + x2v2 with x1 ≥ 0 and x2 ≥ 0 is the

upper shaded region in Figure 2.32, including its edges.The ray containing nonpositive multiples y1v

1 of v1 is the dot-dashed line fromthe origin that is opposite the ray containing v1. The ray containing nonpositivemultiples y2v

2 of v2 is the dot-dashed line from the origin that is opposite the raycontaining v2. The region containing all sums y1v

1 + y2v2 with y1 ≤ 0 and y2 ≤ 0 is

the region consisting of these two dot-dashed lines from the origin together with all ofthe points between them. This is the lower dot-filled region in Figure 2.32, includingits edges.

Answer to Problem 2.4.(i) P is a vector space. Since the sum of polynomials on � is a polynomial on �and since the scalar multiple of any polynomial on � is a polynomial on �, it iseasy to see that P possesses properties 1, 2, 3, and 6 of Definition 2.1. The zero

Page 111: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

92 CHAPTER 2. BASICS

Figure 2.32: The region containing the vectors that are linear combinations x1v1+x2v

2

with x1 ≥ 0 and x2 ≥ 0, and the region containing the vectors that are linearcombinations y1v

1 + y2v2 with y1 ≤ 0 and y2 ≤ 0.

polynomial, o(x) ≡ 0, is the zero element of P , since g(x) + o(x) = g(x) + 0 = g(x)for any g ∈ P . Thus P possesses property 4 of Definition 2.1. If g ∈ P , then(−g) ∈ P (i.e. the negative of a polynomial on � is a polynomial on �), and theng(x) + (−g(x)) = o(x) ≡ 0, so P possesses property 5 of Definition 2.1.

(ii) W is not a vector space. Consider the functions f, g ∈ W that are f(x) ≡ 3M/4and g(x) ≡ M/2 for all x ∈ �. Then f(x) + g(x) ≡ 5M/4 /∈ W , so W does nothave the addition property that af(x) + bg(x) ∈ W for any a, b ∈ �. As well,2f(x) ≡ 3M/2 /∈ W , so W also does not satisfy the scalar multiplication propertythat af(x) ∈ W for all a ∈ �.

Answer to Problem 2.5.

(i) Consider any f, g ∈ A(�n). Then, there are scalars a1, . . . , an, b ∈ � and scalarsa′1, . . . , a

′n, b

′ ∈ � for which f(x1, . . . , xn) = a1x1 + · · ·+ anxn + b and g(x1, . . . , xn) =a′1x1 + · · ·+ a′nxn + b′ for (x1, . . . , xn) ∈ �n.

Page 112: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.24. ANSWERS 93

For any α, β ∈ �, the function h = αf + βg is

h(x1, . . . , xn) = (αa1+βa′1)x1+ · · ·+(αan+βa

′n)xn+αb+βb

′ = a′′1x1+ · · ·+a′′nxn+b′′,

where a′′i = αai + βa′i for i = 1, . . . , n and b′′ = αb+ βb′. Thus αf + βg ∈ A(�n).f ∈ A(�n) implies −f ∈ A(�n), where (−f)(x1, . . . , xn) = (−a1)x1 + · · · +

(−an)xn + (−b).The zero element of A(�n) is the function o(x1, . . . xn) = 0× x1 + · · · 0× xn + 0.For any α ∈ � the function h = αf is

h(x1, . . . , xn) = αa1x1 + · · ·+ αanxn + αb = a′′1x1 + · · · a′′nxn + b′′,

where a′′i = αai for i = 1, . . . , n and b′′ = αb. Thus αf ∈ A(�n).(ii) Repeating the argument given in part (i) shows that L(�n) is a vector space –do it.(iii) Yes. It is called the trivial vector space.

Answer to Problem 2.6. S is any (hyper)plane in �n that includes the origin.That is,

S = {x ∈ �n | p·x = 0}is a proper subspace of �n for any vector p ∈ �n such that p = 0 and ‖p‖E < ∞.Why? First, S ⊂ �n (property (a) in Definition 2.18). Second, take any x′, x′′ ∈ S.Then p·x′ = 0 and p·x′′ = 0, so for any a, b ∈ �, p·ax′ = ap·x′ = 0 and p·bx′′ =bp·x′′ = 0. Thus p·(ax′ + bx′′) = 0, implying that ax′ + bx′′ ∈ S (property (b) inDefinition 2.18).

Answer to Problem 2.7. False. S1 + S2 = {−2,−1, 0, 1, 2}.Answer to Problem 2.8. See Figure 2.33.

S1 + S2 = {{(2, 3)}+ S1} ∪ {{(4, 1)}+ S1}= {(x1, x2) | 3 ≤ x1 ≤ 5, x2 = 9} ∪ {(x1, x2) | 5 ≤ x1 ≤ 7, x2 = 7}.

Answer to Problem 2.9. See Figure 2.34.

Answer to Problem 2.10. S1 + S2 = {(x1, x2) | x2 = −x1, −1 ≤ x1 ≤ 1}.Answer to Problem 2.11. S1 + S2 = �2.

Answer to Problem 2.12.

S1 × S2 = {(x1, x2, x3) ∈ �3 | x2 = −x1 for − 1 ≤ x1 ≤ 0, 0 ≤ x3 ≤ 1}.

Page 113: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

94 CHAPTER 2. BASICS

Figure 2.33: {(x1, x2) | 1 ≤ x1 ≤ 3, x2 = 6}+ {(2, 3), (4, 1)}.

See the top panel of Figure 2.35.

S2 × S1 = {(x1, x2, x3) ∈ �3 | 0 ≤ x1 ≤ 1, x3 = −x2 for − 1 ≤ x2 ≤ 0}.See the bottom panel of Figure 2.35.

Answer to Problem 2.13. There are three possibilities. First, S1 ∩ S2 is empty,in which case the result is true by definition (see Definition 2.5). Second, S1 ∩ S2 isa singleton, in which case the result is again (trivially) true. The third possibility isthat S1 ∩ S2 contains at least two elements, so suppose that this is so.

We start by arbitrarily choosing any two elements x′, x′′ ∈ S1 ∩ S2. What we nowhave to show is that the convex combination θx′+(1− θ)x′′ ∈ S1 ∩S2 for every valueof θ ∈ [0, 1].

x′ and x′′ belong to the intersection of S1 and S2, so we know that x′ and x′′

both belong to S1 and, also, that x′ and x′′ both belong to S2. S1 is a convex set, soθx′+(1−θ)x′′ ∈ S1 for all θ ∈ [0, 1]. Similarly, S2 is a convex set, so θx′+(1−θ)x′′ ∈ S2

also, for all θ ∈ [0, 1]. Since θx′+(1−θ)x′′ belongs to both S1 and S2 for all θ ∈ [0, 1],it belongs to their intersection S1 ∩ S2. This establishes the result.

Answer to Problem 2.14. The argument given in problem 2.13 establishes theresult for a pair of subsets of X. Thus we have already proved the result for k = 2.

The forward-induction argument works as follows. We assume that the result istrue for some value of k, and then prove that the result must be true for k+1. Then,since the result is proved already for k = 2, we know that the result is also true fork = 3, which then tells us that the result is true for k = 4 and so on.

Page 114: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.24. ANSWERS 95

Figure 2.34: Summing a rectangle and a triangle.

So, choose some value for k ≥ 2. We start with the convex subsets S1, . . . , Sk ofX and assume that their intersection S1 ∩ · · · ∩ Sk is a convex subset of X. Now weadd Sk+1, an additional convex subset of X. But the intersection S1 ∩ · · · ∩Sk ∩Sk+1

is the intersection of just two convex subsets of X, being the subset S1 ∩ · · · ∩Sk andthe subset Sk+1. We have already proved that the intersection of two convex sets isa convex set, so it follows that S1 ∩ · · · ∩ Sk ∩ Sk+1 is a convex subset of X.

Now we apply forward induction starting at k = 2, and the result is proved.

Answer to Problem 2.15. If the intersection is either the empty set or a singletonset then it is convex, but only weakly convex in X. If, however, the intersection is aset with a nonempty interior, then the intersection must be a strictly convex subsetof X. You should try to prove this since it is only a small extension of the proof ofproblem 2.13. Remember, a strictly convex set must have a nonempty interior.

Answer to Problem 2.16. We have to show that the convex combination of anytwo elements chosen from the subset S1+S2 is entirely contained in S1+S2. So, startby selecting arbitrarily any two elements x′, x′′ ∈ S1 + S2. Any element in S1 + S2 isthe sum of an element in S1 with an element in S2. So, there are elements y′ ∈ S1

and z′ ∈ S2 that sum to x′ (i.e. x′ = y′ + z′) and there are elements y′′ ∈ S1 and

Page 115: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

96 CHAPTER 2. BASICS

Figure 2.35: The direct products S1 × S2 and S2 × S1 for problem 2.12.

z′′ ∈ S2 that sum to x′′ (i.e. x′′ = y′′ + z′′).The next step is to note that, since S1 is a convex subset of �n and y′, y′′ ∈ S1,

any convex combination of y′ and y′′ is contained in S1; i.e.

θy′ + (1− θ)y′′ ∈ S1 for 0 ≤ θ ≤ 1.

Similarly, since S2 is a convex subset of �n and z′, z′′ ∈ S2, any convex combinationof z′ and z′′ is contained in S2; i.e.

θz′ + (1− θ)z′′ ∈ S2 for 0 ≤ θ ≤ 1.

Page 116: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.24. ANSWERS 97

Since θy′ + (1 − θ)y′′ ∈ S1 and θz′ + (1 − θ)z′′ ∈ S2, the sum of these elementsmust be an element of S1 + S2; i.e.

θy′+(1−θ)y′′+θz′+(1−θ)z′′ = θ(y′+z′)+(1−θ)(y′′+z′′) = θx′+(1−θ)x′′ ∈ S1+S2.

This establishes the result.

Answer to Problem 2.17. The argument given in problem 2.16 establishes theresult for a pair of subsets of �n. Thus we have already proved the result for k = 2.

So, assume, for some k ≥ 2, that, if S1, . . . , Sk are all convex subsets of �n, thenS1 + · · · + Sk is a convex subset of �n. Now, we introduce the set Sk+1 which, too,is a convex subset of �n. But in fact all we have here are two convex subsets of �n,being the subset S1 + · · ·+Sk and the subset Sk+1. We have proved already that thesum of two convex subsets of �n is a convex subset of �n, so applying this result tellsus that S1 + · · · + Sk + Sk+1 is also a convex subset of �n. Now, we apply forwardinduction starting from k = 2 and the result is proved.

Answer to Problem 2.18. Let S = S1 × · · · × Sn. Choose any two elementsx′0, x

′′0 ∈ S. Then, there are elements x′i, x

′′i ∈ Si for each i = 1, . . . , n such that

x′0 = (x′1, . . . , x′n) and x′′0 = (x′′1, . . . , x

′′n). Since each Si is a convex subset of X,

for any number θ where 0 ≤ θ ≤ 1 the element θx′i + (1 − θ)x′′i ∈ Si. But then(θx′1 + (1− θ)x′′1, . . . , θx

′n + (1− θ)x′′n ∈ S1 × · · · × Sn. That is, θx

′0 + (1− θ)x′′0 ∈ S,

so S is a convex subset of ×ni=1X = Xn.

Answer to Problem 2.19. We are given that both of the sets S1 and S2 are strictlyconvex sets. Therefore both sets have nonempty interiors (see Definition 2.6). Thesum of a point from the interior of S1 and a point from the interior of S2 must be apoint in the interior of S1 + S2, so the set S1 + S2 also has a nonempty interior.

Our purpose is to select arbitrarily any two points x′ and x′′ with x′ = x′′ fromthe interior of S1 + S2. We must then show that the strict convex combination of x′

and x′′, which is θx′ + (1 − θ)x′′ for 0 < θ < 1 (note: not θ = 0 and not θ = 1), liesentirely within the interior of S1 + S2.

Now that we have chosen the distinct points x′ and x′′ from S1+S2, we note thatboth x′ and x′′ are the sums of pairs of elements from S1 and S2. That is, there arepoints y′ and y′′ in S1 and points z′ and z′′ in S2 for which x

′ = y′+z′ and x′′ = y′′+z′′.S1 is a strictly convex set, so the strict convex combination that is θy′ + (1 − θ)y′′

for 0 < θ < 1 is everywhere contained in the interior of S1. Similarly, since S2 is astrictly convex set, the strict convex combination θz′ + (1 − θ)z′′ for 0 < θ < 1 iseverywhere contained in the interior of S2. Summing these two points in the interiors

Page 117: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

98 CHAPTER 2. BASICS

of S1 and S2 must give us a point in the interior of S1 + S2. That is, for 0 < θ < 1,

θy′ + (1− θ)y′′ + θz′ + (1− θ)z′′ = θ(y′ + z′) + (1− θ)(y′′ + z′′) = θx′ + (1− θ)x′′

is a point in the interior of S1 + S2. This establishes the result for summing a pair ofsets.

Now consider a sum S1 + · · · + Sk of k ≥ 2 strictly convex subsets. Assume thatthis sum is a strictly convex subset. Now introduce another strictly convex subsetSk+1. The sum S1 + · · ·+ Sk + Sk+1 is the sum of two strictly convex sets, being theset S1+ · · ·+Sk and the set Sk+1. Since we have already proved that the sum of a pairof strictly convex sets is a strictly convex set, we can assert that S1 + · · ·+ Sk + Sk+1

is a strictly convex set. Now we apply forward induction starting from k ≥ 2 and theresult is proved.

Answer to Problem 2.20. Consider the sets S1 ⊂ �2 and S2 ⊂ �1 that areS1 = {(x1, x2) | x21 + x22 ≤ r2} and S2 = {r | r ≥ 0}. If r = 0 then S1 is the singletonset {(0, 0)} and so is a weakly convex subset of �2. If r > 0 then S1 is a strictlyconvex subset of �2. S2 is a strictly convex subset of �1. S1 and S2 are displayedin the left-side panel of Figure 2.36. The direct product S1 × S2 of these two convexsets is S1 × S2 = {(x1, x2, r) ∈ �3 | x21 + x22 ≤ r2, r ≥ 0}. This strictly convex subsetof �3 is the solid version of the parabola that is displayed in the right-side panel ofFigure 2.36 and continued for all r ≥ 0.

Answer to Problem 2.21. See Figure 2.37.(i) The gradient vector function is

∇f(x1, x2) = (1 + x2, x1).

The value at the point (1, 9) of the gradient vector is ∇f(1, 9) = (10, 1).(ii) The f ≡ 10 contour set is the set S = {(x1, x2) ∈ �2 | x1(1 + x2) = 10}.(iii) The total derivative of f is

df =∂f(x1, x2)

∂x1dx1 +

∂f(x1, x2)

∂x2dx2.

When the perturbations dx1 and dx2 are constrained to the f ≡ 10 contour set, wehave df = 0. The total differential then gives the marginal rate of substitution atany point (x1, x2) on the f ≡ 10 contour set as

MRS(x1, x2) =dx2dx1

|f≡10 = − ∂f(x1, x2)/∂x1∂f(x1, x2)/∂x2

|f≡10 = − 1 + x2x1

|f≡10.

Page 118: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.24. ANSWERS 99

Figure 2.36: A direct product of strictly convex sets that is a strictly convex set.

The value at (1, 9) of the marginal rate of substitution is MRS(1, 9) = −10.(iv) The tangent is a straight line with slope −10, so the equation of the tangentis x2 = −10x1 + b for some intercept b. Since the point (1, 9) lies in this tangent,9 = −10 + b ⇒ b = 19. The equation of the tangent is therefore

x2 = −10x1 + 19 or g(x1, x2) = 10x1 + x2 ≡ 19.

(v) ∇g(x1, x2) = (10, 1).(vi) From part (i), the value at (1, 9) of the gradient vector of the f ≡ 10 contourset is ∇f(1, 9) = (10, 1). This is the same as the gradient vector of the tangent,∇g(x1, x2) = (10, 1). Since these two vectors are collinear, the tangent and thegradient vector of the f ≡ 10 contour set at the point (1, 9) must be orthogonal.

Answer to Problem 2.22.(i) The quadratic form of the matrix A is

QA(x1, x2) = (x1, x2)

(4 12

−16 1

)(x1x2

)= 4x21 − 4x1x2 + x22 = (2x1 − x2)

2 ≥ 0.

A is therefore a positive-semi-definite matrix.(ii) The quadratic form of the matrix B is

QB(x1, x2) = (x1, x2)

(7 −2−4 1

)(x1x2

)= 7x21 − 6x1x2 + x22 = −2x21 + (3x1 − x2)

2.

Page 119: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

100 CHAPTER 2. BASICS

Figure 2.37: Diagram for Problem 2.21.

This quadratic takes both positive and negative values, so B is an indefinite matrix.(iii) The quadratic form of the matrix C is

QC(x1, x2) = (x1, x2)

(0 aa 0

)(x1x2

)= 2ax1x2.

If a = 0, then this quadratic is identically zero, so C is then both a positive-semi-definite and a negative-semi-definite matrix. If a = 0, then the quadratic takes bothpositive and negative values, so C is then an indefinite matrix.(iv) The quadratic form of the matrix D is the same as the quadratic form of theorder-3 identity matrix; i.e.

QD(x1, x2) = (x1, x2, x3)

⎛⎝ 1 −a ba 1 −c−b c 1

⎞⎠⎛⎝x1x2x3

⎞⎠ = (x1, x2, x3)

⎛⎝1 0 00 1 00 0 1

⎞⎠⎛⎝x1x2x3

⎞⎠

= x21 + x22 + x23 > 0 ∀ (x1, x2, x3) = (0, 0, 0)

so D is a positive-definite matrix for all a, b and c in �.

Page 120: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.24. ANSWERS 101

Answer to Problem 2.23. Take a moment to review the definition of a metric (seeDefinition 2.17).

f possesses the nonnegativity property, since f(x, y) = 0 if and only if x = y andf(x, y) = 1 > 0 otherwise.

f also possesses the symmetry property, since, for any x, y ∈ � with x = y,f(x, y) = 1 = f(y, x).

Take any x, y, z ∈ �. If x = y = z then f(x, z) = 0 = 0 + 0 = f(x, y) + f(y, z).If x = y = z then f(x, z) = 0 + f(y, z) = f(x, y) + f(y, z). If x = y = z thenf(x, z) = 1 < f(x, y) + f(y, z) = 2. Thus f satisfies the triangular inequality and sois a metric on �2.

Answer to Problem 2.24.

(i) The function f is a metric only if f(x, x) = ad(x, x)+b = b = 0 for any x ∈ X. Sonow we need consider only f(x′, x′′) = ad(x′, x′′) for all x′, x′′ ∈ X. It is clear that fpossesses the nonnegativity property of a metric (since a > 0) and, also, the symmetryproperty. Does f satisfy the triangular inequality? Take any x′, x′′, x′′′ ∈ X. Since dis a metric, d(x′, x′′′) ≤ d(x′, x′′) + d(x′′, x′′′). Consequently,

f(x′, x′′′) = ad(x′, x′′′) ≤ a(d(x′, x′′) + d(x′′, x′′′)) = f(x′, x′′) + f(x′′, x′′′).

Thus f is a metric if b = 0.

(ii) g is not a metric. For example, let d be the Euclidean metric on the real line.Then d(1, 5) = 4 = 1+3 = d(1, 2)+d(2, 5) (thus satisfying the triangular inequality),but g(1, 5) = d(1, 5)2 = 16 ≤ g(1, 2) + g(2, 5) = d(1, 2)2 + d(2, 5)2 = 12 + 32 = 10.More generally, take x′ < x′′ < x′′′ ∈ �. Then,

g(x′, x′′′) = (x′ − x′′′)2 = (x′ − x′′ + x′′ − x′′′)2

= (x′ − x′′)2 + 2(x′ − x′′)(x′′ − x′′′) + (x′′ − x′′′)2

≤ (x′ − x′′)2 + (x′′ − x′′′)2

= g(x′, x′′) + g(x′′, x′′′).

(iii) h is a metric. First, for any x ∈ �, h(x, x) = +√d(x, x) = +

√0 = 0. Second, for

any x′, x′′ ∈ � with x′ = x′′, h(x′, x′′) = +√d(x′, x′′) > 0 since d(x′, x′′) > 0. Next, h

possesses the symmetry property because d possesses the symmetry property:

h(x′, x′′) = +√d(x′, x′′) = +

√d(x′′, x′) = h(x′′, x′).

Page 121: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

102 CHAPTER 2. BASICS

It remains to show that for any x′, x′′, x′′′ ∈ �, h(x′, x′′′) ≤ h(x′, x′′) + h(x′′, x′′′).

h(x′, x′′′) = +√d(x′, x′′′) ≤ +

√d(x′, x′′) + d(x′′, x′′′) (because d is a metric)

≤ +

√d(x′, x′′) + 2

√d(x′, x′′)×

√d(x′′, x′′′) + d(x′′, x′′′)

=

√(+√d(x′, x′′) +

√d(x′′, x′′′)

)2= +

√d(x′, x′′) +

√d(x′′, x′′′)

= h(x′, x′′) + h(x′′, x′′′).

Answer to Problem 2.25. When defined on [0, 1]× [0, 1], f(x1, x2) = 2 + x1 − x2has a saddle-point at (x1, x2) = (0, 0). The restricted function f(x1; x2=0) = 2 + x1is minimized at x1 = 0 and the restricted function f(x2; x1=0) = 2−x2 is maximizedat x2 = 0. However, the point (0, 0) is not a critical point for the function. Thisdemonstrates that a saddle-point for a function need not be a critical point for thefunction if the saddle-point is not in the interior of the function’s domain.

Answer to Problem 2.26. The function f(x1, x2) = (x21 − x22)e−(x2

1+x22) has five

critical points. (x1, x2) = (1, 0) and (x1, x2) = (−1, 0) are global maxima. (x1, x2) =(0, 1) and (x1, x2) = (0,−1) are global minima. (x1, x2) = (0, 0) is a saddle-point.

The gradient vector of f is

∇f(x1, x2) =(2x1(1− x21 + x22

)e−x2

1−x22 , −2x2

(1 + x21 − x22

)e−x2

1−x22

).

The solutions to ∇f(x1, x2) = (0, 0) are the critical points of f , listed above.The first-order derivative of the restricted function f(x1; x2=0) = x21e

−x21 is

df(x1; x2=0)

dx1= 2x1(1− x21)e

−x21 = 0

at x1 = −1, 0,+1. The restricted function’s second-order derivative is

d2f(x1; x2=0)

dx21= 2(1− 5x21 + 2x41)e

−x21 .

The values of this second-order derivative at x1 = −1, 0 + 1 are negative, positive,and negative, respectively. The function f(x1; x2=0) is locally minimized at x1 = 0.

The first-order derivative of the restricted function f(x2; x1=0) = −x22e−x22 is

df(x2; x1=0)

dx2= −2x2(1− x22)e

−x22 = 0

Page 122: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

2.24. ANSWERS 103

at x2 = −1, 0,+1. The restricted function’s second-order derivative is

d2f(x2; x1=0)

dx22= −2(1− 5x22 + 2x42)e

−x22 .

The values of this second-order derivative at x2 = −1, 0 + 1 are positive, negative,and positive, respectively. The function f(x2; x1=0) is locally maximized at x2 = 0.This and above the paragraph together show that (x1, x2) = (0, 0) is a saddle-pointfor the function f .

Table 2.3: The values of f and Q at the requested points.

(x1, x2) (Δx1,Δx2) f(x1, x2) Q(2 + Δx1, 1 + Δx2)

(2, 1) (0, 0) 5 5

(2·1, 1·1) (0·1, 0·1) 5·741 5·740(1·9, 0·9) (−0·1,−0·1) 4·339 4·340(2·5, 1·5) (+0·5,+0·5) 9·625 9·50(3, 2) (+1,+1) 17 16

(1, 3) (−1,+2) 25 20

Answer to Problem 2.27. The gradient vector for f(x1, x2) = x31 − x21x2 + x32 is

∇f(x1, x2) = (3x21 − 2x1x2,−x21 + 3x22) = (8,−1)

when evaluated at (x1, x2) = (2, 1). The function’s Hessian matrix is

Hf (x1, x2) =

(6x1 − 2x2 −2x1−2x1 6x2

)=

(10 −4−4 6

)

when evaluated at (x1, x2) = (2, 1). Last, f(2, 1) = 5, so the Taylor quadratic is

Q(2 + Δx1, 1 + Δx2) = 5 + (8,−1)

(Δx1Δx2

)+

1

2(Δx1,Δx2)

(10 −4−4 6

)(Δx1Δx2

)= 5 + 8Δx1 −Δx2 + 5(Δx1)

2 − 4Δx1Δx2 + 3(Δx2)2.

Page 123: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

Chapter 3

Basics of Topology

Obtaining a clear understanding of the essentials of constrained optimization theoryrequires repeated use of ideas that depend upon the properties of sets. Examplesare ideas as basic and important as an “open set” and “continuity of a mapping.”So we must start by examining basic properties of sets. The name of the branch ofmathematics that studies the properties of sets that are preserved by mappings istopology. I used to think that topology was a subject that mattered only for seriousmathematicians, and my heart would sink whenever a research paper said somethingominous like “Let T (X) denote the something-or-other topology on X . . . ” Help! Isurrender! The truth is that the basics of topology are easy. Yes, easy. Of coursesome sets and mappings can be rather complicated, and that is when topology getscomplicated. But the topological ideas that economists typically use are pretty simple.Do not be afraid.

3.1 Topology

The most basic of basics, the foundation upon which all will be built, is the referenceset. I will typically denote such a set by X. What the reference set contains willdepend upon the problem we are working on. For example, if it is a problem ofchoosing a color, then the reference set will be the set containing all possible colors.But if the problem is choosing which house to buy, then the reference set will be theset of all houses. The reference set X is the complete collection of elements relevantto the problem we are thinking about. The typical label we will use to denote anyelement in the set X is x, so you will often see the statement x ∈ X, meaning that xis an element of, or belongs to, or is contained in the set X.

104

Page 124: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

3.1. TOPOLOGY 105

You already know that the basic operations used on sets are union ∪ and inter-section ∩. We use these operations to construct new subsets of X from some originalcollection of subsets of X. A topology on a reference set X is merely a collectionof subsets of X that are formed by particular uses of the union and intersectionoperations. It is a simple idea.

Definition 3.1 (Topology). A topology T (X) on a reference set X is a collection ofsubsets of X such that

(i) ∅ ∈ T (X),(ii) X ∈ T (X),(iii) any union of any members of T (X) is a member of T (X), and(iv) any intersection of any finite collection of members of T (X) is a member of

T (X).

A few examples will help make you comfortable with the idea of a topology. Forthe first example, take any reference set X you like and then consider the collectionof sets C = {∅, X}. Is C a topology with respect to X? Think about it before youproceed to the next paragraph.

C is indeed a topology with respect to X. It is obvious that C possesses properties(i) and (ii). What about property (iii)? Let S1, S2, . . . be any sequence of subsets in C.Then either Si = ∅ for all i, or it does not. If Si = ∅ for all i, then S = ∪∞

i=1 = ∅ ∈ C.If not, then Si = X for at least one i so S = ∪∞

i=1Si = X ∈ C. C therefore possessesproperty (iii). What about property (iv)? This time we consider only a finite sequenceS1, . . . , Sn of subsets in C. Either Si = X for all i = 1, . . . , n or it does not. If Si = Xfor all i = 1, . . . , n, then S = ∩n

i=1 = X ∈ C. If not, then Si = ∅ for at least one i, soS = ∩n

i=1Si = ∅ ∈ C. Hence C possesses all of properties (i), (ii), (iii), and (iv) andis a topology with respect to the reference set X. What we have described here isin fact the simplest (the “coarsest,” in the parlance of topology) of all topologies inthat it is the smallest of all of the topologies that can be constructed from any givenreference set. This two-element topology is called the indiscrete topology on X.

Here is a second example. The reference set is X = {1, 2, 3}. The collection C ofsubsets of X is now

C = {∅, {1}, {1, 3}, {1, 2, 3}}.Work out for yourself if C is a topology with respect to X before continuing to thenext paragraph.

Yes, it is. For a third example, again take X = {1, 2, 3}. Consider the collection

C = {∅, {1}, {2}, {1, 3}, {1, 2, 3}}.

Page 125: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

106 CHAPTER 3. BASICS OF TOPOLOGY

Is C a topology with respect to X? Come to a conclusion before proceeding to thenext paragraph.

No, it is not. Why not? Because {1} ∪ {2} = {1, 2} /∈ C.For a fourth example, consider the collection

C = {∅, {1}, {1, 3}, {2, 3}, {1, 2, 3}}.

Is C a topology with respect to X = {1, 2, 3}?No, because {1, 3} ∩ {2, 3} = {3} /∈ C. By now you should have realized that it is

possible to form more than one topology from the same reference set X. Confirm foryourself that each of the following collections of subsets is a topology with respect toX = {1, 2, 3}:

C ′ = {∅, {1}, {1, 2, 3}}.C ′′ = {∅, {2}, {1, 3}, {1, 2, 3}}.C ′′′ = {∅, {1}, {2}, {3}, {1, 2}, {1, 3}, {2, 3}, {1, 2, 3}}.

If one topology on X contains all of the subsets contained in a second topology on X,then the first topology is said to be at least as refined as (or “no more coarse than”)the second topology.

Definition 3.2 (Refined Topology). Let T ′(X) and T ′′(X) be topologies on X. IfS ∈ T ′(X) for every S ∈ T ′′(X), but T ′(X) = T ′′(X) (i.e. T ′′(X) ⊂ T ′(X)),then the topology T ′(X) is strictly more refined, or strictly finer, or strictly lesscoarse than, or is a strict refinement of the topology T ′′(X), and the topology T ′′(X)is strictly less refined, or strictly coarser, or strictly more coarse than the topologyT ′′(X).

In our above examples, C ′ and C ′′ are not refinements of each other, becauseC ′ ⊂ C ′′ and C ′′ ⊂ C ′. C ′′′ is a strict refinement of C ′ because C ′′′ contains all of thesubsets in C ′ (C ′ ⊂ C ′′′) and has at least one more subset of X as well (C ′ = C ′′′). C ′′′ isalso a strict refinement of C ′′. For any reference set X, the coarsest, or least refined,topology on X is always the indiscrete topology.

What is the most refined of all of the possible topologies on a set X? It must bethe collection of all of the possible subsets of X, including ∅ and X. This is calledthe power set of X and is denoted by P(X). The power set of X is always a topologyon X. Why? Clearly ∅ ∈ P(X) and X ∈ P(X). Also, since P(X) contains all ofthe subsets of X, it must contain all of the subsets of X that can be created by allpossible unions and all possible finite intersections of the subsets of X. Hence P(X)

Page 126: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

3.2. TOPOLOGICAL BASES 107

must be a topology on X. It is easy to see that P(X) is the most refined of all of thetopologies on X. (Think of a quick, easy proof.) For this reason it is given a specialname. P(X) is called the discrete topology on X. Check for yourself that the aboveexample C ′′′ is the discrete topology on X = {1, 2, 3}.

3.2 Topological Bases

The next basic ideas in topology are a basis of a topology and generating a topologyfrom a basis. Simply put, here they are. From a given reference set X, we take acollection B of subsets of X. If this collection satisfies two particular properties, thenwe call it a basis in X. We then generate another collection C of subsets of X bytaking all of the possible unions of the subsets in B and then, if need be, addingthe empty set into C. C will be a topology on X, and we say that C is the topologygenerated by the basis B. The practical value of a basis is that it allows us a simpleway of constructing a topology. You will see a little later that you have already usedthis idea, even if you are not yet aware of having done so.

Definition 3.3 (Basis of a Topology). Let B be a collection of subsets of X. B is abasis for a topology on X if

(i) for every x ∈ X there is a B ∈ B with x ∈ B, and(ii) for any x ∈ B1∩B2, where B1, B2 ∈ B, there exists B3 ∈ B with B3 ⊆ B1∩B2

and x ∈ B3.

What does all this mean? Requirement (i) just says that every element of X mustbe contained in at least one of the subsets in B. If this is not so, then we cannotgenerate X by taking the union of all of the subsets in B and so we cannot generatea topology on X. But what is the purpose of requirement (ii) – it seems to be abit mysterious, doesn’t it? Remember that, for a collection of subsets of X to be atopology, any finite intersection of these subsets must be a subset contained in thetopology. Requirement (ii) ensures that this is so. Some examples will make thiseasier for you to see.

Again take X = {1, 2, 3}. Start with the collection of subsets of X that is

B1 = {{1}, {2, 3}}.

Is B1 a basis?Yes, it is. Each of the elements 1, 2, and 3 in X is contained in at least one of

the subsets contained in B1. Requirement (ii) does not apply since all of the subsets

Page 127: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

108 CHAPTER 3. BASICS OF TOPOLOGY

in B1 have only an empty intersection with each other. What, then, is the topologyon X that is generated by taking all possible unions of the subsets in B1 and thenincluding the empty set also? It is

T1(X) = {∅, {1}, {2, 3}, {1, 2, 3}}.

Take a moment to check that T1(X) really is a topology on X = {1, 2, 3}.Is

B2 = {∅, {1, 2}, {2, 3}}a basis?

No. Why not? Because 2 ∈ {1, 2} ∩ {2, 3} and there is no subset in B2 that is asubset of {1, 2} ∩ {2, 3} and contains 2. So what happens if we take the union of allof the subsets in B2? Do we generate a topology on X? The collection of subsets ofX that is generated is

C2 = {∅, {1, 2}, {2, 3}, {1, 2, 3}}.

But C2 is not a topology on X because {1, 2} ∩ {2, 3} = {2} /∈ C2. This would nothave happened had B2 satisfied requirement (ii).

IsB3 = {{1, 3}, {2}, {3}}

a basis?Yes, it is. Notice that 3 ∈ {3} ⊆ {1, 3} ∩ {3}. The topology generated on X by

taking all possible unions of the members of B3 and then also including the emptyset is

T3(X) = {∅, {2}, {3}, {1, 3}, {2, 3}, {1, 2, 3}}.Check for yourself that this really is a topology on X = {1, 2, 3}.

As a last example, consider

B4 = {∅, {1}, {2}, {3}}.

Confirm for yourself that this is a basis and that it generates the discrete topologyon X = {1, 2, 3}.

A moment of thought should make it clear to you that, if we have two differentbases, B′ and B′′, for X with B′ ⊂ B′′ and B′ = B′′ (i.e. the basis B′′ contains allof the subsets in the basis B′ and at least one more subset of X), then the topologygenerated by B′′ must be at least as refined as the topology generated by B′.

By the way, the term topological space means a reference set X paired up with atopology defined on X.

Page 128: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

3.3. OPEN AND CLOSED SETS 109

Definition 3.4 (Topological Space). If T (X) is a topology on a reference set X,then the ordered pair (X, T (X)) is called a topological space.

We have covered the basic ideas of topologies. The next thing to do is to examinethe meanings of two of the most important ideas in mathematics. These are opensets and closed sets.

3.3 Open and Closed Sets

You might be used to saying something like the following: A set is open if and onlyif every point in the set is interior to the set, and that a point is interior to a set ifand only if the point can be surrounded by an open ball of finite radius with thatball entirely contained in the interior of the set. Whew! What a mouthful. Not onlythat, such statements are special and are not to be used as a general definition of anopen set. As you will see in a moment, you should never say that a set is open, orclosed for that matter, unless you also say from which topology the set is taken. Yes,openness of a set is a topological idea. So for now I want you to forget entirely anynotions you may have that sets are open only if they are something like sets with onlyinterior points, or some such. Please, put them out of your mind. We will get back tothese ideas when we are ready to consider the special cases in which such statementsare true.

Given a reference set X and a subset S ⊆ X, the collection of all elements of Xthat are not contained in S is called the complement of S in X and will be denotedby Sc; Sc ≡ {x ∈ X | x /∈ S} .

The general definitions of open and closed sets are easy to understand. Here theyare.

Definition 3.5 (Open Set). Given a topology T (X) on a set X, a set S ⊆ X is openwith respect to T (X) if and only if S ∈ T (X).

All this says is that a subset S of X is open with respect to some topology T (X)on X if and only if S is one of the subsets contained in T (X). That’s all there is toit – really!

Definition 3.6 (Closed Set). Given a topology T (X) on a set X, a set S ⊆ X isclosed with respect to T (X) if and only if Sc ∈ T (X); i.e. if and only if Sc is openwith respect to T (X).

Page 129: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

110 CHAPTER 3. BASICS OF TOPOLOGY

So a subset S of X is closed with respect to some topology T (X) on X if andonly if the complement in X of S is one of the subsets contained in T (X). Simple,isn’t it?

Let’s play around with some examples. Start with the simplest of all topologies;i.e. the indiscrete topology T (X) = {∅, X}. The above definition of an open set saysthat both sets, ∅ and X, are open with respect to the indiscrete topology just becausethey are members of that topology. And since any topology on X contains both theempty set ∅ and the reference set X, these sets are always open sets. What is thecomplement of the open set X with respect to X? It is the empty set ∅, so the emptyset must be a closed set with respect to that topology. Similarly, the complementwith respect to X of the open set ∅ is the reference set X, so the reference set mustbe closed with respect to that topology. We have discovered that, no matter whattopology you are working with, the empty set ∅ and the reference set X are alwaysboth open and closed with respect to that topology.

Reconsider some of our earlier examples for the reference set X = {1, 2, 3}. Onetopology on X is C ′ = {∅, {1}, {2, 3}, {1, 2, 3}}. Therefore the sets ∅, {1}, {2, 3}, and{1, 2, 3} are all open with respect to the topology C ′. Hence their complements, whichare respectively the sets {1, 2, 3}, {2, 3}, {1}, and ∅ are all sets that are closed withrespect to the topology C ′. Thus all of the members of this topology happen to besets that are both open and closed with respect to the topology C ′.

Another example we considered is the topology on X = {1, 2, 3} that is C ′′ ={∅, {1}, {1, 2, 3}}. Therefore the subset {1} is open with respect to C ′′, and so thesubset {1}c = {2, 3} is closed with respect to the topology C ′′. What about the subset{2}? It is not open with respect to C ′′ since it is not a member of C ′′. Nor is it closedwith respect to C ′′ since {2}c = {1, 3} is not open with respect to C ′′ because {1, 3} isnot a member of C ′′. Thus whether or not a set is open or closed depends not only onthe set but also upon the topology being used. Given a topology on X, some subsetsof X are open sets, some are closed sets, some subsets may be both open and closed,and some subsets may be neither open nor closed – it all depends upon the particulartopology you are using.

Here is a little teaser for you. Take any reference set X and construct its discretetopology T (X). Now take any subset S of X. By construction, S ∈ T (X) and soS is open with respect to the discrete topology. This establishes that any subset ofX is open with respect to the discrete topology on X. Must S also be closed withrespect to the discrete topology on X? In other words, must every possible subset ofX be a set that is both open and closed with respect to the discrete topology on X?Think it over. I’ll let you know in a little while.

Page 130: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

3.4. METRIC SPACE TOPOLOGIES 111

Typically a lot of topologies are possible for a given reference set X. In economicsit is usually the case that X is some subset of �n. This will be the case we examinemost in this book, but you should always remember that the ideas we have methere are very general. For instance, the reference set can be something like X ={green, butterfly, Fred, Africa, water}, and we can logically talk of topologies on thisset and of open and closed subsets of this X.

Did you decide if every member of the discrete topology on X must be both openand closed with respect to that topology, no matter what is the reference set X?The answer is “Yes” and it is easy to see why. The discrete topology on X containsevery possible subset of X, so every possible subset of X is open with respect to thediscrete topology on X. Pick any one of these open subsets. Its complement mustalso belong to the discrete topology because that topology contains every possiblesubset of X, and so this set must be open with respect to the topology, but it mustalso be closed since it is the complement of an open set. If this is confusing to you,look back at the discrete topology C ′′′ used as an example earlier. Every set in C ′′′ isa subset of the reference set X = {1, 2, 3} that is both open and closed with respectto the topology C ′′′. Try working out an example for yourself.

Now let’s move on to some special cases that will look more familiar to you.

3.4 Metric Space Topologies

If you have forgotten, or not yet examined, what is a metric space, then you need tostop reading this section and explore Section 2.20. Come back here once you haveunderstood it.

The idea we are going to use repeatedly in this section is that of a ball.

Definition 3.7 (Ball). Let (X, d) be a metric space. Let x ∈ X. Let ε ∈ �++. Theball centered at x with radius ε is the set

Bd(x, ε) = {x′ ∈ X | d(x, x′) < ε}.

In simple words, the ball Bd(x, ε) is the set of elements in X that are distant fromx by strictly less than ε. Such a ball is often called an ε-neighborhood about x. Noticethat every element in the ball belongs to X, so Bd(x, ε) ⊆ X. You have met sets likethis before. Simple examples are intervals in the real line such as

Bd(x, ε) ={x′ ∈ � | d(x, x′) =

∣∣x− x′∣∣ < ε

}= (x− ε, x+ ε),

Page 131: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

112 CHAPTER 3. BASICS OF TOPOLOGY

discs in �2 such as

Bd(x, ε) ={x′ ∈ �2 | d(x, x′) =

√(x1 − x′1)2 + (x2 − x′2)2 < ε

}, (3.1)

and balls in �n such as

Bd(x, ε) ={x′ ∈ �n | d(x, x′) =

√(x1 − x′1)2 + · · ·+ (xn − x′n)2 < ε

}. (3.2)

Some care has to be taken when constructing balls in bounded spaces since, remember,any ball is entirely contained in its reference space. For example, if X = [0, 4] anddistance is measured by the Euclidean metric, then

BdE(3, 1) = {x ∈ [0, 4] |∣∣x− 3

∣∣ < 1} = (2, 4),

but BdE(3, 2) = {x ∈ [0, 4] |∣∣x− 3

∣∣ < 2} = (1, 4].

Similarly, if X = [0, 4]× [0, 4] and distance is again measured by the Euclidean metric,then

BdE((4, 2), 2) = {(x1, x2) ∈ [0, 4]2 |√

(x1 − 4)2 + (x2 − 2)2 < 2}is the half -disc centered at (4, 2). Draw it.

Balls do not have to be intervals, discs, or spherical subsets of real numbers.Consider the metric space that is the set C(X) of all functions that are continuous,real-valued, and bounded on a set X, with the distance between any two such func-tions measured by the sup norm d∞ (see (2.43)). Then in the metric space (C(X), d∞)the ball centered at the function f and with radius ε > 0 is the set

Bd∞(f, ε) =

{g ∈ C(X) | d∞(f, g) = sup

x∈X|f(x)− g(x)| < ε

}.

Such a set looks nothing like a nice round Euclidean ball.Now let’s form the collection of all of the balls in a metric space (X, d). This is

the collectionB(X) = {Bd(x, ε) | ∀ ε > 0 and ∀ x ∈ X} .

For example, if the metric space is the two-dimensional Euclidean metric space E2 =(�2, dE), then B(�2) is the collection of every possible disc (3.1). The claim I wantto make here is that, no matter what is the metric space you are considering, B(X)is a topological basis in that space. Let’s see why.

Does B(X) satisfy the two requirements of a basis? Look again at Definition 3.3.It is easy to see that requirement (i) is satisfied since B(X) contains at least one ball

Page 132: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

3.4. METRIC SPACE TOPOLOGIES 113

centered at every x ∈ X. What about requirement (ii)? Choose any two balls youlike from B(X). Call them Bd(x

′, ε′) and Bd(x′′, ε′′). If Bd(x

′, ε′) ∩ Bd(x′′, ε′′) = ∅,

then requirement (ii) is irrelevant. Suppose Bd(x′, ε′) ∩ Bd(x

′′, ε′′) = ∅. Choose anyx ∈ Bd(x

′, ε′) ∩ Bd(x′′, ε′′). B(X) contains balls Bd(x, ε) that are centered at x. In

fact, B(X) contains such a ball for every possible radius ε > 0. Therefore there is aball Bd(x, ε) with a radius ε small enough that the ball is entirely contained in theintersection Bd(x

′, ε′) ∩ Bd(x′′, ε′′) and, obviously, x is contained in this ball. Thus

requirement (ii) is satisfied and so B(X) is indeed a basis in (X, d).Therefore, if we take all of the possible unions of the balls in B(X), we will generate

a topology on (X, d). Any set in this topology is thus the union of some collectionof balls and, by definition, every such set is open with respect to the topology sogenerated.

You have already met this construction, in the special case of the Euclidean spaces.In En = (�n, dE) the balls are perfectly round spheres (see (3.1) and (3.2)) with radiimeasured by the Euclidean metric, with every point in �n being the center of onesuch ball for every possible positive radius:

B(�n) = {BdE(x, ε) ∀ ε > 0 and ∀ x ∈ �n}.

The topology generated by this basis is called the Euclidean metric topology on �n

so any set that is open with respect to this topology is a set that is the union of somesubset of the ε-balls in B(�n). I hope it is now clear why we say that a set is open inEn (i.e the set belongs to the topology on �n that is generated by the basis B(�n)consisting of all possible Euclidean ε-balls in �n) if and only if every point in the setis surrounded by an ε-ball that is entirely contained within the set. This “definition”is just a restatement of the manner in which the sets in the Euclidean metric topologyon �n are generated by taking unions of ε-balls. I also hope that you now understandthat such a statement can be made for any set in any metric space topology thatis generated from the basis that consists of all of the possible ε-balls in that metricspace. It is not a statement that is restricted only to the Euclidean metric spaces.Do understand, though, that the statement is restricted to metric spaces.

Think once more about the topologies C ′, C ′′, and C ′′′ on X = {1, 2, 3} that wehave already repeatedly examined. In C ′′, for example, the subset {1, 3} is open withrespect to C ′′, but there are no balls here. Why not? Because there is no metricfunction defined on X = {1, 2, 3}. Openness of a set is an idea that does not relyupon being able to measure distances.

Well, that is about all we need to say about bases and about open and closed sets.The next important idea is that of a bounded set.

Page 133: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

114 CHAPTER 3. BASICS OF TOPOLOGY

3.5 Bounded Sets

This is an obvious idea. A set is bounded if and only if it is inside a ball with afinite radius. This guarantees that the set does not “go forever” in any direction. Forexample, think of a peanut inside of a basketball. The peanut is a subset containedinside the larger set that is the basketball. Because the basketball has only a finiteradius, the peanut is a bounded set. Notice, then, that to talk about whether a setis or is not bounded we need a distance measure (something that we do not need todecide if a set is open or closed), and so we talk about bounded sets in the contextof a metric space.

Definition 3.8 (Bounded Set). Let (X, d) be a metric space. A set S ⊆ X is boundedin (X, d) if there is a point x ∈ X and a finite radius r > 0 such that S ⊆ Bd(x, r).

Figure 3.1: Some sets that are bounded with respect to E2.

All of the sets in Figure 3.1 are bounded with respect to the metric space E2. Forexample, each of the sets S1 and S2 is contained in the ball BdE(y1, r1) with finiteradius r1, so both sets are bounded. So is the set S1∪S2 and the set S4. The set S3 isthe same as the ball of finite radius BdE(y3, r3) that contains it, so it too is bounded.

Page 134: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

3.6. COMPACT SETS 115

In contrast, each of the three sets displayed in Figure 3.2 is not bounded withrespect to E2. The half-space that is the set S7 = {(x1, x2) | x1 + x2 ≥ 2} is not“bounded above” and so cannot be contained inside any ball with only a finite radius.The set of points S8 = {(x1, x2) | x2 ≤ −x21} is not “bounded below” and so it toocannot be contained inside any ball with only a finite radius. An example of a discreteset that is not bounded is the set S9 of all of the integers.

Figure 3.2: Some sets that are not bounded with respect to E2.

The next important idea is that of a compact set.

3.6 Compact Sets

Just think of the word, compact. It seems to suggest something that is not too spreadout and with edges to it – isn’t that what you mean when you say, for example, thata car or an apartment is compact? With respect to the Euclidean metric topologies,how do we describe a set that does not extend all over the place; i.e. a set in whichany two points are not “too far” apart? We say that such a set is bounded, don’t we?And how do we say that a set contains its “edges”? We say that the set is closed.

Page 135: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

116 CHAPTER 3. BASICS OF TOPOLOGY

This probably is already familiar to you since you will often have heard the statementthat a set is compact if and only if it is both closed and bounded. If at this momentyou are nodding to yourself and saying that’s right, then stop! The statement is nottrue in general. It is true in a special case that is very important to economists, solet’s state the relevant theorem and have a close look at it.

Theorem 3.1 (Heine-Borel). Let S ⊂ �n. S is compact with respect to the Euclideanmetric space En = (Rn, dE) if and only if S is both closed and bounded with respectto this metric space.

The conditions imposed by this theorem are severe. First, it considers only subsetsof real spaces. Second, it considers only one topology on the real spaces – justone! So the Heine-Borel Theorem, while very important and useful, is rather special.Fortunately, the Euclidean metric topology is where we economists do most of ourthinking. But there are times when economists have to work in other topologicalspaces, and then compactness is not the same as closed and bounded. It is true thatin any metric-space topology a set is compact only if (note: just “only if,” not “ifand only if”) it is bounded and closed. But in general metric space topologies, theconverse is typically not true, so be careful.

3.7 How Do Our Topological Ideas Get Used?

By now you may be wondering why the ideas discussed in this chapter matter. Whocares if some set is compact? So what if a set is open, or not? Isn’t constrainedoptimization really about computing things? If so, then why aren’t we getting readyto crunch some numbers? I think we all become a little impatient when trying tounderstand clearly some body of knowledge. The details never seem to end, andseem to slow up our progress towards the heart of whatever it is we are trying tounderstand. But experience has taught me the hard way that patient and organizedstudy of a topic is the only way to master it. Nothing that has been presented in thischapter is less than essential to a detailed understanding of the types of constrainedoptimization problems that confront economists.

Why does it matter whether a set is open or closed? Most constrained optimizationproblems have two parts to them. The first part is the set of things from which adecisionmaker is allowed to choose. The second part is a function that states thedecisionmaker’s valuations of these available choices. For the typical economist’sproblem, if the set of available choices is not closed, then there may not be a solution

Page 136: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

3.7. HOW DO OUR TOPOLOGICAL IDEAS GET USED? 117

to the problem. As well, the function that states the decisionmaker’s valuationsis usually “continuous” (meaning “varies smoothly without sudden jumps”), but afunction is continuous only if it is defined over open sets. Continuity of functions is areally basic, important idea that we will use a lot, and for this we will need to knowwhen sets are open or not.

Why should economists care about bounded sets? Well, what is it that makeseconomics such an important and practical science? The answer is scarcity. What isscarcity? It is the fact that in our world resources are available only in finite quantities.Recognizing this is the same as saying that the set consisting of all available resourcesis bounded. In every economics problem I can think of to do with allocating someresource (a budget or a production input, perhaps), the set of available choices is abounded set and, typically, a compact (both closed and bounded) set.

So, have patience and proceed to the next necessary ideas, which are sequencesand limits. Sequences and, if they converge, their limits are used in Euclidean metricspace environments to verify if sets are open or closed, and to verify if functions arecontinuous or not, so you should view the contents of the next chapter as providing youa means of determining the properties of sets that we have discussed in this chapter.You will find that there are many places in the chapters that follow where we willuse sequences to establish results that are crucial for solving constrained optimizationproblems.

Page 137: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

118 CHAPTER 3. BASICS OF TOPOLOGY

3.8 Problems

Problem 3.1. Let X = {1, 2, 3, 4, 5}. Consider the following collection of subsets ofX:

C1 = {{1, 2}, {3, 4, 5}, ∅}.Generate the collection D1 of subsets of X by taking all possible unions of the sets inC1.(i) Show that C1 is a topological basis in X. Then show that D1 is a topology withrespect to X.

(ii) Now consider the following collection of subsets of X:

C2 = {{4}, {1, 2}, {3, 4, 5}, ∅}.

Show that C2 is a topological basis in X. Then generate the collection D2 of subsetsof X by taking all of the possible unions of the sets in C2 and show that D2 is thetopology in X that is generated by C2.(iii) Notice that {4} ∈ C2 is a subset of the basis element {3, 4, 5} ∈ C1. Also notice,by comparing how you generated the topologies D1 and D2, that this is why D1 isstrictly contained in D2; more formally, D2 is finer in X than is D1. It is a generalresult. Here is the theorem.

Theorem. Let B1 and B2 be bases in X. Let T1 be the topology in X that isgenerated by B1 and let T2 be the topology in X that is generated by B2. Then T2

is finer in X than is T1 if and only if, for every x ∈ X and for every basis elementB ∈ B1 that contains x, there is a basis element C ∈ B2 such that x ∈ C ⊆ B.

Problem 3.2. Let X = {1, 2, 3, 4, 5}. Let B = {∅, {1, 2}, {3, 4, 5}}. B is a basis fora topology on X.

(i) What is the topology on X that is generated by B?(ii) Consider the topology generated in part (i). Is it possible for all of the sets ina topology (other than the indiscrete and discrete topologies) to be both open andclosed with respect to that topology?

Problem 3.3. It is claimed in the text of this chapter that a particular collection ofsubsets of �n is a basis for the Euclidean metric topology on �n. The collection ofsets is

B(�n) = {BdE(x, ε) ∀ x ∈ �n, ∀ ε > 0}.Explain why this claim is true.

Page 138: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

3.9. ANSWERS 119

Problem 3.4. Prove the following theorem.Theorem. If S1, . . . , Sn is a finite collection of subsets that are bounded with

respect to a metric space (X, d), then S = ∪ni=1Si is bounded with respect to (X, d).

Problem 3.5. Present a counterexample to the assertion that, if S1, S2, . . . is anycountable collection of subsets that are bounded with respect to a metric space (X, d),then S = ∪∞

i=1Si must be bounded with respect to (X, d).

Problem 3.6. Prove the following theorem.Theorem. Let S1, S2, . . . be any countable collection of subsets that are bounded

with respect to a metric space (X, d). Then S = ∩∞i=1Si is bounded with respect to

(X, d).

Problem 3.7. Present a counterexample to the assertion that S = ∩∞i=1Si is bounded

with respect to (X, d) only if S1, S2, . . . is a countable collection of subsets that arebounded with respect to a metric space (X, d).

Problem 3.8. Prove the following theorem.Theorem. Let S1, . . . , Sn be any finite collection of subsets that are compact with

respect to the metric space En. Then S = ∪ni=1Si is compact with respect to En.

Problem 3.9. Present a counterexample to the assertion that, if S1, S2, . . . is a count-able collection of subsets that are compact with respect to the metric space En, thenS = ∪∞

i=1Si must be compact with respect to En.

Problem 3.10. Present a counterexample to the assertion that S = ∪ni=1Si is com-

pact with respect to En only if S1, . . . , Sn is a countable collection of subsets that arecompact with respect to En.

Problem 3.11. Prove the following theorem.Theorem. Let S1, . . . , Sn be any finite collection of subsets that are compact

with respect to En. Then S = ∩ni=1Si is compact with respect to En.

3.9 Answers

Answer to Problem 3.1.(i) C1 is a basis in X if and only if

(a) x ∈ X implies x ∈ B for at least one member B in the collection C1, and(b) ifB1, B2 ∈ C1 and x ∈ B1∩B2, then there existsB3 ∈ C1 such that B3 ⊆ B1∩B2

and x ∈ B3.

Page 139: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

120 CHAPTER 3. BASICS OF TOPOLOGY

It is easily seen that each of the individual elements 1, 2, 3, 4, and 5 belongs to atleast one of the members of C1. Thus (a) is true. (b) is trivially satisfied since thereis no x ∈ X that belongs to the intersection of elements of C1. Therefore C1 is a basis.Particularly, C1 is the basis for the topology that is generated by taking unions of theelements of C1. This is the collection

D1 = {∅, {1, 2}, {3, 4, 5}, {1, 2, 3, 4, 5}}.Inspection shows that unions of any members of D1 generate only members of D1 andthat finite intersections of any members of D1 generate only members of D1. Also,∅ ∈ D1 and X = {1, 2, 3, 4, 5} ∈ D1. Hence D1 is the topology with respect to X thatis generated by the basis C1. All of the members of D1 are thus sets that are openwith respect to the topology D1 in X = {1, 2, 3, 4, 5}.(ii) C2 is a basis in X if and only if

(a) x ∈ X implies x ∈ B for at least one member B in the collection C2, and(b) ifB1, B2 ∈ C2 and x ∈ B1∩B2, then there existsB3 ∈ C2 such that B3 ⊆ B1∩B2

and x ∈ B3.

Again it is easily seen that each of the individual elements 1, 2, 3, 4, and 5 belongto at least one of the members of C2. Thus (a) is true. The only element of X thatbelongs to an intersection of elements in C2 is 4, since 4 ∈ {4}∩ {3, 4, 5}. Notice that4 ∈ {4} ⊆ {4} ∩ {3, 4, 5}. Thus (b) is also true and so C2 is a basis. Particularly, C2is the basis for the topology that is generated by taking unions of the elements of C2.This is the collection

D2 = {∅, {4}, {1, 2}, {1, 2, 4}, {3, 4, 5}, {1, 2, 3, 4, 5}}.Inspection shows that unions of any members of D2 generate only members of D2 andthat finite intersections of any members of D2 generate only members of D2. Also,∅ ∈ D2 and X = {1, 2, 3, 4, 5} ∈ D2, so D2 is the topology with respect to X that isgenerated by the basis C2. All of the members of D2 are thus sets that are open withrespect to the topology D2 in X = {1, 2, 3, 4, 5}.Answer to Problem 3.2.(i) The topology is

T = {∅, {1, 2}, {3, 4, 5}, {1, 2, 3, 4, 5}}.

(ii) Each of the sets in T is open with respect to T . The complements of each ofthese sets are thus closed with respect to T . These closed sets are, respectively,

{1, 2, 3, 4, 5}, {3, 4, 5}, {1, 2} and ∅.

Page 140: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

3.9. ANSWERS 121

Therefore every set in T is both open and closed with respect to T . The answer tothe question “Is it possible for all of the sets in a topology (other than the indiscreteand discrete topologies) to be both open and closed with respect to that topology?”is “Yes”.

Answer to Problem 3.3. B(�n) is a basis for a topology on �n if and only if B(�n)satisfies properties (i) and (ii) in Definition 3.3. It is clear that, by construction,

B(�n) = {BdE(x, ε) ∀ x ∈ �n, ∀ ε > 0}

satisfies property (i). What about property (ii)? Take any two members B1 andB2 from B(�n) such that B1 ∩ B2 = ∅. Let x ∈ B1 ∩ B2. Notice that B1 ∩ B2

cannot be a singleton set. Why? Well, if it were, then x would be a boundary pointfor both the ball B1 and the ball B2. But, by construction, none of the balls inB(�n) contains any of its boundary points, so B1 ∩ B2 is not a singleton set. Thisimplies that, for any x ∈ B1 ∩ B2, there is a small enough ε′ > 0 such that the ballB3 = BdE(x, ε

′) = {y ∈ �n | dE(x, y) < ε′} ⊂ B1 ∩ B2. Again, by construction,B3 ∈ B(�n). Thus B(�n) also satisfies property (ii), establishing that B(�n) is abasis for a topology on �n.

Answer to Problem 3.4. The result is trivial for n = 1. Consider the case of n = 2.Since S1 is a bounded set, there exists a point x1 ∈ X and a radius ε1 > 0 such thatS1 ⊆ Bd(x1, ε1). Similarly, since S2 is a bounded set, there exists a point x2 ∈ X anda radius ε2 > 0 such that S2 ⊆ Bd(x2, ε2). Let x

′ ∈ S1 and let x′′ ∈ S2. Then, by thetriangular inequality property of a metric function,

d(x′, x′′) ≤ d(x′, x1) + d(x1, x2) + d(x2, x′′) ≤ ε1 + d(x1, x2) + ε2.

The right-hand-side quantity is a finite and positive number that is independent ofboth x′ and x′′. Hence

d(x′, x′′) ≤ ε1 + d(x1, x2) + ε2 ≡ δ ∀ x′ ∈ S1, ∀ x′′ ∈ S2.

Thus S1 ∪ S2 ⊆ Bd(x1, δ), which establishes that S1 ∪ S2 is a bounded subset of X.

The result is extended to the union of any finite number n ≥ 2 of sets by induction.For any n ≥ 2, write the union of the sets S1, . . . , Sn as ∪n−1

i=1 Si∪Sn. By the inductionhypothesis, ∪n−1

i=1 Si is a bounded set. Therefore the union of the two bounded subsets∪n−1

i=1 Si and Sn must be bounded in (X, d). Since the result is true for n = 2, it istrue for all n ≥ 2.

Page 141: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

122 CHAPTER 3. BASICS OF TOPOLOGY

Answer to Problem 3.5. Let X = � and let d be the Euclidean metric. Fori = 0, 1, 2, . . ., let Si = [i, i+ 1] ⊂ �. Each set Si is bounded, but

S = ∪∞i=0Si = ∪∞

i=0[i, i+ 1] = [0,∞)

is not bounded with respect to En.

Answer to Problem 3.6. S1 is bounded in (X, d), so there exists x ∈ X and ε1 > 0such that S1 ⊆ Bd(x, ε1). S = ∩∞

i=1Si ⊆ S1 so S ⊆ Bd(x, ε1). That is, S is boundedin (X, d).

Answer to Problem 3.7. Consider the countable collection S1 and S2, whereS1 = (−∞, 0] and S2 = [0,∞). Neither S1 nor S2 is bounded with respect to E1. YetS1 ∩ S2 = {0} is bounded with respect to E1.

Answer to Problem 3.8. Each Si is closed in En and the union of finitely manyclosed sets is a closed set. Thus S is a closed set. From Problem 3.4, S is also boundedand so is compact with respect to En.

Answer to Problem 3.9. The counterexample provided for Problem 3.5 is a coun-terexample here also.

Answer to Problem 3.10. Let S1 = [0, 1] and Si = (0, 1) for all i = 2, 3, . . ..Obviously all of S2, S3, . . . are open, and thus not compact, with respect to E1. ButS1 ∪ S2 ∪ S3 ∪ · · · = [0, 1], which is compact with respect to E1.

Answer to Problem 3.11. The intersection of any finite collection of subsets thatare closed with respect to (X, d) is a subset that is closed with respect to (X, d).The proof given in Problem 3.6 states that S is bounded in (X, d). Then S is bothclosed and bounded with respect to (X, d). If X = �n and d is the Euclidean metric,then (X, d) = En and so S is compact with respect to En (by the Heine-BorelTheorem).

Page 142: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

Chapter 4

Sequences and Convergence

4.1 Sequences

This chapter presents only a few of the many ideas and results to do with sequences.Its content is determined solely by what is required to understand discussions in laterchapters of this book. There is little value in duplicating the extensive discussionsof sequences that are presented in other places, so the interested reader is invited toadvance his or her knowledge of sequences beyond the contents of this chapter by pe-rusing any of the many mathematics books that provide more expansive expositions.

We are going to make a lot of use of the set of positive integers because theyare the numbers that we most often use as “counters” when constructing a sequence.The positive integers are called the natural numbers. This set {1, 2, 3, . . .} is oftendenoted by N , but N is also used to denote many other quantities so, to avoidambiguity we will follow mathematicians and use Z+ to denote the set of naturalnumbers; Z+ = {1, 2, 3, . . .}.

Definition 4.1 (Sequence). A sequence of elements from a set X is a surjectivefunction f : Z+ → X.

All this means is that we construct a mapping of the form {(n1, x1), (n2, x2), . . .}that more commonly we write as {xn1 , xn2 , . . .} by using the subscript ni on xni

todenote the natural number ni that is mapped only to the element xni

∈ X. That thefunction is required to be surjective (see Definition 5.7) means every element in thesequence is associated with at least one “counter” index element n. In most cases theindices are 1, 2, 3, . . ., so we usually see a sequence written as x1, x2, x3, . . .. Whenwe say that x1 is the first member of a sequence, we are saying that the number 1 is

123

Page 143: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

124 CHAPTER 4. SEQUENCES AND CONVERGENCE

associated with (i.e. the number 1 is mapped to) the first value in the sequence. And,more generally, when we say that xi is the i

th member of a sequence, we are sayingthat the number i is associated with the ith value in the sequence, for i = 1, 2, 3, . . ..A particular element in a sequence may be associated with more than one indexelement. For example, it might be that x3 = x15 = 9; i.e. the third and fifteenthelements of the sequence are both the number 9.

A sequence can be infinite, x1, x2, . . . , xi, . . ., or it can be finite, x1, . . . , xm forsome finite integer m ≥ 1. There are various notations used for sequences. Somepeople write a sequence simply as xn, but this leads to ambiguities so I will not useit. Others use (xn), 〈xn〉 or 〈xn〉∞n=1. Still others use {xn} or {xn}∞n=1. I will use thelast of these. {xn}∞n=1 will denote a sequence of infinitely many elements. {xn}mn=1

will denote a finite sequence of n elements.

Economics is the science of scarcity. Scarcity imposes bounds upon what is achiev-able. Consequently many of the sequences that concern economists are bounded se-quences.

4.2 Bounded Sequences

A sequence is bounded, meaning both bounded above and bounded below, if all ofthe elements in the sequence are not larger than some finite upper bound and alsoare not less than some finite lower bound.

Definition 4.2 (Bounded Sequence). Let X ⊂ � and let {xn}∞n=1 be a sequence ofelements from X. {xn}∞n=1 is bounded on X if there exists B ∈ X such that |xn| ≤ Bfor all n = 1, 2, . . ..

Definition 4.2 is the same as saying that the set {x1, x2, . . .} is contained in aclosed ball centered at zero and with a finite radius of at least B in the range of thefunction f : Z+ → X that defines the sequence {xn}∞n=1.

Bounds are not unique. Consider the sequence {2,−1, 2,−1, 2,−1, . . .}. Anynumber 2 or larger is an upper bound for this sequence. Any number −1 or less is alower bound. 2 is significant in that it is the least upper bound for the sequence. −1is significant in that it is the greatest lower bound.

Now consider the sequence{1− 1

n

}∞n=1

={0, 1

2, 23, 34, . . .}. This sequence is bound-

ed, above and below. The greatest lower bound is zero, and this is a member (thefirst element) of the sequence. The least upper bound is 1, which is not a memberof the sequence. So, a bounded sequence may or may not contain its least upper

Page 144: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

4.3. CONVERGENCE 125

bound or its greatest lower bound. The least upper bound of a sequence {xn}∞n=1 isthe supremum of the sequence and is denoted by sup xn. The greatest lower boundof the sequence is the infimum of the sequence and is denoted by inf xn .

Some sequences are only bounded above, meaning that there is a finite numberB such that xn ≤ B for every n ≥ 1. An example is S = {0,−1,−2,−3, . . .}. Forthis sequence the least upper bound of zero is the supremum: sup (S) = 0. Since thesequence is unbounded below, there is no numerical value for the infimum, and wewrite inf (S) = −∞. Other sequences are only bounded below, meaning that there is afinite number B such that B ≤ xn for every n ≥ 1. An example is S ′ = {0, 1, 2, 3, . . .}.The greatest lower bound for this sequence is zero: inf (S ′) = 0. There is no numericalvalue for the supremum of this sequence, so we write sup (S ′) = +∞.

Be careful with your terminology. If you say that a sequence is “bounded,” thenyou mean that the sequence is bounded above and bounded below. If you mean onlyone of these, then say so.

4.3 Convergence

Let’s refresh our memories about the idea of convergence of a sequence. We havesome set X and we construct a sequence of elements xn from X. Most sequences arenot convergent. But if, as n becomes very large, the elements of the sequence become“arbitrarily close” to some value x0 ∈ X, then we state that the sequence convergesand that x0 is the limit of the sequence. To state this exactly requires somethingthat gives a precise meaning of “close.” The usual tool is a metric function d thatallows us to measure the distance d(xk, x0) between any element xk in the sequenceand the limiting value x0. So the idea is that, eventually, when we get far enoughalong a convergent sequence, the distances d(xk, x0) all are very small and continueto get ever closer to zero as k increases; i.e. as we continue to go ever further outalong the sequence.

Definition 4.3 (Convergent Sequence). Let (X, d) be a metric space. Let {xn}∞n=1 bea sequence of elements of X. {xn}∞n=1 converges to x0 ∈ X with respect to the metricd if and only if, for every ε > 0, there exists an integer M(ε) such that d(xn, x0) < εfor every n ≥M(ε).

Think of the following simple example. Let the reference set X = [1, 2] and letthe metric function be the Euclidean metric. Consider the sequence

s∗ ={1, 1 +

1

2, 1 +

2

3, 1 +

3

4, . . . , 1 +

n− 1

n, . . .

}.

Page 145: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

126 CHAPTER 4. SEQUENCES AND CONVERGENCE

Is s∗ a convergent sequence? If so, then what is its limit? Can you prove it? Don’tread on until you have an answer.

s∗ is convergent on X = [1, 2] with respect to the Euclidean metric. The limitis 2. A proof is that the Euclidean distance of any element xn = 1 + (n− 1)/n from2 is

dE(xn, 2) =

√(1 +

n− 1

n− 2

)2

=1

n→ 0 as n→ ∞.

Thus, as n → ∞ (i.e. as we move further and further along the sequence), thedistances of elements xn from 2 become arbitrarily close to zero.

That was easy. Let’s modify the example a little by changing the reference set toX = [1, 2). Is s∗ still a convergent sequence? If so, then what is its limit? Again,don’t read on until you have thought about it and come to a reasoned response.

s∗ is not convergent. If you said otherwise, then you have made a common butserious error. A sequence cannot be convergent unless there exists a limit for it. So ifyou want to say that s∗ is still convergent then you have to say that its limit exists.But you cannot say that the limit of s∗ is 2 because 2 is not an element of the referenceset; as far as we are concerned, the number 2 does not exist. And if the limit doesnot exist, then the sequence is not convergent. Never think about whether a sequenceconverges or not without taking careful note of the reference set.

It is obvious that a sequence that converges cannot have elements in it that arepersistently wildly different from each other. In fact the opposite should occur, atleast eventually. So that must mean that all of the elements of a convergent sequenceof real numbers should lie between some upper bound and some lower bound; i.e. aconvergent sequence of real numbers should be bounded. This is indeed so, and it iseasy to prove. Before stating the theorem and its proof, however, have a look at thefollowing convergent sequence of real numbers. Let the reference set X = � and usethe Euclidean metric to measure distances.

s∗∗ = {−1000, 2000,−2500, 8200,−10, 8,−4, 3, 0, 2, 2, 2, . . . , 2, . . .}.It is easy to see that this is a convergent sequence with 2 as its limit. You can seethat any number at least as large as 8200 is an upper bound on the sequence and thatany number less than −2500 is a lower bound, so the sequence is bounded (meaningboth bounded above and bounded below). The fact that the early elements of thesequence vary greatly in value affects neither that the sequence converges nor thatthe sequence is bounded.

Now let’s prove our assertion that any sequence that converges is a boundedsequence.

Page 146: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

4.3. CONVERGENCE 127

Theorem 4.1 (A Convergent Sequence is Bounded). Let X ⊂ � and let {xn}∞n=1

be a sequence of elements from X. If {xn}∞n=1 is convergent on X, then {xn}∞n=1 isbounded on X.

Proof. We are given that the sequence {xn}∞n=1 converges to some limiting value x0:xn → x0 ∈ X. So pick any distance ε > 0 and there must exist an integer M(ε) suchthat, for the part of the sequence xM(ε), xM(ε)+1, xM(ε)+2, . . ., the distance between eachof these elements of the sequence and the limiting value x0 is less than ε; dE(xn, x0) =|xn − x0| < ε for every n ≥M(ε). Therefore

|xn| < |x0|+ ε ≡ L1 for every n ≥M(ε).

That is, the part of the sequence that is xM(ε), xM(ε)+1, xM(ε)+2, . . . is bounded by thenumber L1 = |x0|+ ε.

The rest, the first part, of the sequence is x1, x2, . . . , xM(ε)−1. The distancesof these sequence elements from the limiting value x0 form a finite set of valuesd(x1, x0) = |x1 − x0|, . . . , d(xM(ε)−1, x0) = |xM(ε)−1 − x0|. Thus

|x1| < |x0|+ d(x1, x0), . . . , |xM(ε)−1| < |x0|+ d(xM(ε)−1, x0),

and so

|x1| < |x0|+max {d(x1, x0), . . . , d(xM(ε)−1, x0)}...

...

|xM(ε)−1| < |x0|+max {d(x1, x0), . . . , d(xM(ε)−1, x0)}.

Thus the first part of the sequence is bounded by the number

L2 ≡ |x0|+max {d(x1, x0), . . . , d(xM(ε)−1, x0)}.

Hence the entire sequence is bounded by the larger of the two numbers L1 and L2:

|xn| ≤ max {L1, L2} for every n ≥ 1.

It is easy to see that the converse is not true. That is, a bounded sequence neednot be convergent. For example, the sequence {0, 1, 0, 1, 0, 1, . . .} is bounded but it isnot convergent. However, there are two simple and important types of sequences ofreal numbers that are always convergent.

Page 147: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

128 CHAPTER 4. SEQUENCES AND CONVERGENCE

Definition 4.4 (Monotonic Sequence). Let X ⊂ �. Let {xn}∞n=1 be a sequence ofelements from X. If xn ≤ xn+1 for every n ≥ 1, then {xn}∞n=1 is a monotonicincreasing sequence on X. If xn < xn+1 for every n ≥ 1, then {xn}∞n=1 is a strictlymonotonic increasing sequence on X. If xn ≥ xn+1 for every n ≥ 1, then {xn}∞n=1 isa monotonic decreasing sequence on X. If xn > xn+1 for every n ≥ 1, then {xn}∞n=1

is a strictly monotonic decreasing sequence on X.

When a sequence is increasing we write xn ↗, and when a sequence is decreasingwe write xn ↘. If a sequence is increasing and is convergent to a limit x0, then wewrite xn ↗ x0. If a sequence is decreasing and is convergent to a limit x0, then wewrite xn ↘ x0.

What if a sequence {xn}∞n=1 is increasing and is bounded above? Well, the elementsin the sequence cannot decrease as n increases and neither can they increase pasta certain value. Shouldn’t such a sequence be convergent? Similarly, shouldn’t adecreasing sequence that is bounded below be convergent? Both statements are trueand are very useful.

Theorem 4.2 (Monotonic Increasing Sequence Bounded Above Converges). Let{xn}∞n=1 be a monotonic increasing sequence of real numbers that is bounded above.Then {xn}∞n=1 is convergent.

Proof. xn ↗. Let B be an upper bound for {xn}∞n=1. Then x = sup xn ≤ B andso the least upper bound x exists. Since x is the least upper bound and since xn ↗for any ε > 0, there must exist an integer M(ε) such that |xn − x| < ε for everyn ≥M(ε). Hence the sequence converges to x.

Theorem 4.3 (Monotonic Decreasing Sequence Bounded Below Converges). Let{xn}∞n=1 be a monotonic decreasing sequence of real numbers that is bounded below.Then {xn}∞n=1 is convergent.

The proof is almost the same as for Theorem 4.2, so take a few minutes andconstruct it for yourself.

Here is one last small caution before we move along to the next section. Supposeyou have a sequence of real numbers xn with xn < b for every n ≥ 1. Suppose thatthe sequence converges to some limit x0. Must x0 < b? Think about it. Write downan example or two. When you are ready, move to the next paragraph.

x0 < b may not be true. x0 ≤ b is always true. Think of the sequence in whichxn = 2− 1/n. Clearly xn < 2 for every n ≥ 1. Even so, xn → 2. Similarly, if xn > bfor every n ≥ 1 and xn → x0, then it is always true that x0 ≥ b but the statementthat x0 > b may be false.

Page 148: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

4.4. SUBSEQUENCES 129

4.4 Subsequences

How common is it to work with monotonic sequences? For economists, the answeris “Quite often.” Perhaps more important still is that any sequence of real numbers(convergent or not) possesses at least one subsequence that is monotonic. In a lot ofthe cases that matter to economists, this is a really important piece of information.Let’s start by understanding what is a subsequence.

Definition 4.5 (Subsequence). Let {xn}∞n=1 be a sequence in X. Let {k1, k2, . . .} bea subset of {1, 2, . . .} with k1 < k2 < · · · . Then {xki}∞i=1 is a subsequence of {xn}∞n=1.

Put into simple words, a subsequence is a subset of the elements of the parentsequence. The order of the elements in the subsequence is the same as in the parentsequence. For example, the sequence {−2,−4, 15} is a subsequence of the sequence{3,−2, 7, 0,−4,−8, 9, 15, 2}, while the sequence {−4,−2, 15} is not.

As the following theorem states, monotonic subsequences of real numbers arecommonplace.

Theorem 4.4 (Any Sequence of Real Numbers Contains a Monotonic Subsequence).Any sequence {xn}∞n=1 of real numbers possesses at least one monotonic subsequence.

Proof. For each n ≥ 1 define the subsequence sn = {xn, xn+1, . . .}. There are twopossibilities. Either the parent sequence {xn}∞n=1 possesses a largest element, or itdoes not.

Suppose that {xn}∞n=1 does not possess a largest element. Then it possesses amonotonically increasing sequence {yk}∞k=1 constructed as follows. Set y1 = x1. Thenset y2 equal to the first element in s2 that is larger than y1 (such an element existsbecause the original sequence {xn}∞n=1 does not have a maximal element). Denote thiselement by xm2 . Then set y3 equal to the first element in sm2+1 that is larger than y2;call this element xm3 . Continuing in this way constructs a sequence {y1, y2, y3, . . .} ={x1, xm2 , xm3 , . . .} that is monotonically increasing.

Now suppose instead that {xn}∞n=1 does possess a largest element. Then the se-quence possesses a monotonically decreasing sequence {zk}∞k=1 constructed as follows.For each i ≥ 1 let max si be the largest element in si. Set z1 = max s1 and de-note this element by xm1 . Then set z2 = max sm1 and denote this element by xm2 .Clearly z1 ≥ z2. Continue in this manner. The sequence {zk}∞k=1 is monotonicallydecreasing.

In the above proof we could equally well have said that the two cases to beconsidered are, first, that {xn}∞n=1 does not possess a minimal element and, second,

Page 149: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

130 CHAPTER 4. SEQUENCES AND CONVERGENCE

that {xn}∞n=1 does possess a minimal element. You might like to “reprove” the theoremfor these two cases. The statement of the theorem will not alter.

If we put together what we have learned about sequences of real numbers fromTheorems 4.2, 4.3, and 4.4, then we immediately have a tremendously useful andfamous theorem. Here it is. Proving it wasn’t all that hard, was it?

Theorem 4.5 (Bolzano-Weierstrass). Every bounded sequence of real numbers pos-sesses at least one convergent subsequence.

Remember our earlier example of the nonconvergent sequence of real numbers{0, 1, 0, 1, 0, 1, . . .}? This is a bounded sequence, so it must possess at least oneconvergent subsequence. Actually it possesses lots of convergent subsequences. Hereare just a few of them:

{0, 0, 0, 0, 0, 0, 0, . . . , 0, . . .} → 0,

{1, 1, 1, 1, 1, 1, 1, . . . , 1, . . .} → 1,

{0, 1, 0, 0, 0, 0, 0, . . . , 0, . . .} → 0,

and {0, 0, 0, 1, 1, 1, 1, . . . , 1, . . .} → 1.

There are two types of subsequences that are especially useful when deciding if asequence is, or is not, convergent. They are the limit superior subsequence and thelimit inferior subsequence.

Definition 4.6 (Limit Superior and Limit Inferior). The limit superior of a sequence{xn}∞n=1 with xn ∈ � for all n ≥ 1 is

lim sup xn = limn→∞

supk≥n

xk. (4.1)

The limit inferior of the sequence {xn}∞n=1 is

lim inf xn = limn→∞

infk≥n

xk. (4.2)

If the limit superior does not exist, then lim sup xn = +∞. If the limit inferior doesnot exist, then lim inf xn = −∞.

These two limits seem a bit odd at a first glance, but there is nothing difficultabout either of them. Let’s consider the limit superior first. From (4.1) we see that,for any given n ≥ 1, we are asked to consider the subsequence {xn, xn+1, xn+2, . . .}and then, for this particular subsequence, to discover its supremum (i.e. its least

Page 150: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

4.4. SUBSEQUENCES 131

upper bound). So let’s call this supremum sn = sup {xn, xn+1, xn+2, . . .}. Then wedo the same thing for the shorter subsequence {xn+1, xn+2, . . .}; i.e. we discover thesupremum sn+1 = sup {xn+1, xn+2, . . .}. We keep doing this for each n ≥ 1. As nincreases we are evaluating the supremum of an ever smaller collection of elements ofthe original sequence {xn}∞n=1 so the sequence of these suprema is weakly decreasing:sn ≥ sn+1 for every n ≥ 1. Figure 4.1 will help you to understand the idea.

………………

……

Figure 4.1: The limit superior of a sequence.

In Figure 4.1 you see plotted some of the elements xn of the original sequence{xn}∞n=1. For n = 1 we evaluate s1, the supremum of every element of the originalsequence. For n = 2 we evaluate the supremum s2 of a new sequence that is theoriginal sequence without the element x1. In the figure it happens that s1 = s2.We keep doing this for each value of n. By the time we reach n = 20, we areevaluating s20, the least upper bound on the sequence {x20, x21, . . .}. In the figure s20is smaller than s1, . . . , s19 because the elements x1, . . . , x19 of the original sequenceare no longer being considered. As n continues to increase, the suprema sn neverincrease and, in the figure, gradually fall in value towards a limiting value x0. This

Page 151: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

132 CHAPTER 4. SEQUENCES AND CONVERGENCE

limit of the sequence of suprema is the limit superior of the original sequence {xn}∞n=1;i.e. x0 = limn→∞ sn = limn→∞ sup {xn, xn+1, . . .} = lim sup xn.

The limit inferior of a sequence is very similar in concept to the limit superior ofthe sequence. We start by considering the infimum (i.e. greatest lower bound) of thesubsequence {xn, xn+1, xn+2, . . .}. This is sn = inf {xn, xn+1, xn+2, . . .}. Then we con-sider the sequence of these infima as n→ ∞. The limit of this sequence of infima is thelimit inferior of the original sequence; limn→∞ sn = limn→∞ inf {xn, xn+1, xn+2, . . .} =lim inf xn.

It is possible, of course, that a particular sequence is not convergent. Even so,such a sequence might possess both a limit superior and a limit inferior. An obviousexample is once again the sequence {1,−1, 1,−1, 1,−1, . . .}. This is not a convergentsequence. Even so, the limit superior of the sequence is 1 and the limit inferior is −1.The fact that the limit superior and the limit inferior are not the same tells us thatthe sequence is not convergent.

Now consider again the convergent sequence {1,−1, 12,−1

2, 13,−1

3, . . .}. I hope it

is clear to you that the limit superior of this sequence is the limit of the sequenceof suprema that is {1, 1

2, 12, 13, 13, . . .} and that the limit inferior of the sequence is

the limit of the sequence of infima that is {−1,−1,−12,−1

2,−1

3. . .}. The limits of

both sequences are zero. Clearly the limit of the original sequence too is zero, so forthis convergent sequence the limit superior, the limit inferior, and the limit are allthe same value. This is, in fact, an alternative way of stating when a sequence isconvergent.

Theorem 4.6. Let (X, d) be a metric space and let xn ∈ X for all n ≥ 1. Then thesequence {xn}∞n=1 is convergent if and only if lim sup xn = lim inf xn. The value ofthe limit of the sequence is limn→∞ xn = lim sup xn = lim inf xn.

4.5 Cauchy Sequences

According to our earlier definition of convergence of a sequence, you need to know thevalue of the limit x0 of the sequence to establish if the sequence is convergent, for, ifnot, then how will you be able to measure the differences d(xk, x0) and decide if thesedifferences eventually shrink to zero as k increases? This is a Catch-22 situation. IfI know the limit of a sequence, then I already know it is convergent and to whatvalue. If I don’t know the limit of a sequence, then I cannot check that the sequenceconverges and so I will not be able to discover the limit.

Page 152: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

4.5. CAUCHY SEQUENCES 133

This seems to be most unsatisfactory, and I am sure that you are muttering toyourself that you don’t really need to know the limit of a sequence in order to checkthat it is a convergent sequence. Perhaps you think that all you really need to check isthat eventually the neighboring values in the sequence all become almost the same aseach other, and get closer and closer in value as we move further and further along thesequence? Unfortunately that’s not always true, as we will see, though I sympathizewith what you have in mind, but a modification of your idea has proved to be mostuseful. It was given to us by the great French mathematician Augustin-Louis Cauchy,and now is the time to work it out for ourselves. Let’s look at a few examples.

Sequences that are convergent can converge in various ways. Think about thefollowing sequences.

s′ ={1,

1

2,1

3,1

4,1

5, . . .

},

s′′ ={1,−1

2,1

3,−1

4,1

5, . . .

},

and s′′′ ={1, 0,

1

2, 0, 0,

1

3, 0, 0, 0,

1

4, 0, 0, 0, 0,

1

5, . . .

}.

All three of these sequences converge. All three converge to the same limit, zero. Butthe convergences occur in different ways. Of them all, the sequence s′ can be judged toconverge in the “smoothest” manner. As you travel along s′, the elements stay positiveand the distances |(1/n)−(1/(n+1))| between neighboring elements all monotonicallydecrease towards zero. The sequence s′′ also contains only elements that always getcloser to zero, but the differences |(1/n) + (1/(n+1))| between neighboring elementsare greater than for s′. Even so, these differences also monotonically shrink towardszero as n increases. The sequence s′′′ consists of segments commencing with 1/n andfollowed by n zeros. In this case, as n increases there are longer and longer stringsof zeros interspersed with smaller and smaller nonzero numbers 1/n. Notice thatthe distance between neighboring zeros is zero and that the distance between anynonzero element 1/n and its zero neighbors is greater than zero. Thus the distancesbetween neighboring elements of s′′′ do not monotonically decrease towards zero. ButI’m sure that you were able quickly to deduce that each of the sequences converged,and converged to zero, even though I did not tell you that the value of each limitis zero. What was your reasoning? I’m ready to bet that you reasoned that, asyou moved further and further along the sequence, the distance between any of theremaining elements of the sequence was generally getting smaller and eventually wouldbe arbitrarily close to zero for all of the remaining elements. If so, then accept my

Page 153: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

134 CHAPTER 4. SEQUENCES AND CONVERGENCE

congratulations because you have discovered the important idea of a Cauchy sequence.

Definition 4.7 (Cauchy Sequence). Let (X, d) be a metric space. Let {xn}∞n=1 be asequence of elements of X. If, for any positive ε > 0, there exists an integer M(ε) ≥ 1such that d(xi, xj) < ε for all i, j ≥ M(ε), then the sequence is a Cauchy sequencewith respect to d.

What does all this mean? Here’s the idea. Choose a value for ε > 0, say, ε = 12.

If a sequence is Cauchy, then there will be an element, the M(12

)thelement, in the

sequence such that, for every pairing of elements selected from the M(12

)thelement

and its successors, the distance between the pair of elements is strictly less than 12.

In sequence s′, M(12

)= 2. In sequence s′′, M

(12

)= 4. In sequence s′′′, M

(12

)= 4.

Now choose a smaller value for ε, say, ε = 110. For sequence s′, the distance between

any pair of elements after the third element is strictly less than 110

so M(

110

)= 3. For

sequence s′′, M(

110

)= 20. For sequence s′′′, M

(110

)= 56. The value that you select

for ε is an upper bound on the distances that you will accept from all possible pairsof elements in the sequence. As you command ε to become nearer to zero, you areinsisting that the elements of the “right-hand end” of the sequence all become closerto each other and thus closer to a common value. Typically this will happen only ifyou consider elements far enough along in the sequence, and so typically the value ofM(ε) rises as the value you impose on ε gets closer to zero. And since you can makethe value of ε as close to zero as you like, it should be clear that any Cauchy sequenceis a convergent sequence. Let’s try a few more examples.

Is the sequences = {1,−1, 1,−1, 1,−1, . . .}

Cauchy or not? Think about it and come to a definite conclusion before reading anyfurther.

For the sequence to be Cauchy, you must be able, for any ε > 0, to find anM(ε)th

element such that all pairs of elements further along the sequence are distant by lessthan the value you chose for ε. Let’s choose ε = 1

2. Is there in s any element such

that all later pairs of elements are distant from each other by less than 12? No, there

is not. No matter how far you look along the sequence, there will always be elementsdistant from each other by 2, which is not less than ε = 1

2. Therefore s is not a

Cauchy sequence.Is the sequence

s =

{1, 1 +

1

2, 1 +

1

2+

1

3, . . . ,

n∑i=1

1

i, . . .

}

Page 154: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

4.5. CAUCHY SEQUENCES 135

a Cauchy sequence? Don’t read on until you have a definite answer – this is animportant example.

s is not a Cauchy sequence. It is not even a convergent sequence. The trap (I hopeyou did not get caught in it) in this example is that the distance d(xn, xn+1) = 1/(n+1) between an element in the sequence and its immediate neighbor monotonicallydecreases towards zero as n increases, and this might seem to suggest that the sequenceis convergent. Not so. Looking at the distances between only neighboring elementsdoes not tell you if a sequence is convergent. You need to look at the distancesbetween all pairs of elements, not just neighbors. Suppose you choose ε = 1/100.If you look at, say, the 100th and 101st elements, then indeed the distance betweenthese neighboring elements is 1/101 < 1/100. But notice that the distance betweenthe 100th and 102nd element is 1/101 + 1/102 = 203/10302 > 1/100. In fact forany n the distance d(xn, xk) =

∑ki=n+1 1/i strictly increases as k > n increases and,

moreover, has no upper bound. A sequence is Cauchy if and only if, for any positivebound ε (no matter how close to zero) placed on the distances between all pairs ofsuccessive elements, there is always an element in the sequence beyond which all ofthese pairwise distances are strictly less than ε. s is a counterexample to the incorrectassertion that a sequence must converge if the distance between neighboring elementsxn and xn+1 goes to zero as n→ ∞.

What are the relationships between Cauchy sequences and convergent sequences?It should be reasonably easy to see that a convergent sequence must be a Cauchysequence, but must every Cauchy sequence be convergent? The following theoremgives the answers. Have a look at the proofs. They are instructive and quite easy tofollow.

Theorem 4.7. Let (X, d) be a metric space. Let {xn}∞n=1 be a sequence of elementsdrawn from X.

(i) {xn}∞n=1 is a convergent sequence only if it is a Cauchy sequence.

(ii) If {xn}∞n=1 is a Cauchy sequence, then it may not be convergent in X.

(iii) If {xn}∞n=1 is a Cauchy sequence, then it is a sequence that is bounded in X.

(iv) If {xn}∞n=1 is a Cauchy sequence that contains a subsequence that is convergentin X, then {xn}∞n=1 is also convergent in X and has the same limit as its subsequence.

Proof of (i). Since the sequence {xn}∞n=1 is convergent in X, there is an x0 ∈ X thatis the limit of the sequence; xn → x0. Therefore, for any ε > 0 there is an integerM(ε/2) such that

d(xn, x0) <ε

2for every n ≥M(ε/2).

Page 155: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

136 CHAPTER 4. SEQUENCES AND CONVERGENCE

Since d is a metric,

d(xn, xm) ≤ d(xn, x0) + d(x0, xm) for every n,m ≥ 1,

and so, for n,m > M(ε/2) in particular,

d(xn, xm) <ε

2+ε

2= ε.

That is, the sequence {xn}∞n=1 is Cauchy.Proof of (ii). An example suffices as a proof. Choose X = [1, 2) and consider thesequence {xn}∞n=1, where xn = 2− 1/n for n = 1, 2, . . .. It is easily shown that this isa Cauchy sequence, but it does not converge since 2 /∈ X.Proof of (iii). {xn}∞n=1 is Cauchy. The task is to show that there is a ball of finiteradius that contains every element of the sequence. Choose any ε > 0. Then thereexists an integer M(ε) such that d(xn, xm) < ε for every n,m ≥ M(ε). Hence, forn =M(ε) in particular,

d(xM(ε), xm) < ε for every m ≥M(ε),

and soxm ∈ Bd(xM(ε), ε) for every m ≥M(ε).

Defineρ ≡ max {d(x1, xM(ε)), d(x2, xM(ε)), . . . , d(xM(ε)−1, xM(ε))}+ μ,

where μ > 0 is any positive real number. Then

xm ∈ Bd(xM(ε), ρ) for every m = 1, . . . ,M(ε)− 1.

Therefore

xm ∈ Bd(xM(ε),max {ε, ρ}) for every m ≥ 1.

Proof of (iv). {xn}∞n=1 is Cauchy. {xmk}∞k=1 is a convergent subsequence of {xn}∞n=1.

Let x0 denote the limit of {xmk}∞k=1. Then, for any ε > 0, there exists an integer

Mxm(ε/2) such that

d(x0, xmk) <

ε

2for every k ≥Mxm(ε/2).

The integer Mxm(ε/2) is the number (position) of an element in {xmk}∞k=1. Because

{xmk}∞k=1 is a subsequence of {xn}∞n=1, this element is also a member of {xn}∞n=1. Let

Page 156: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

4.6. REMARKS 137

Mx(ε/2) be the position of this element in {xn}∞n=1. Mx(ε/2) ≥ Mxm(ε/2) because{xm1 , xm2 , . . .} is a subset of {x1, x2, . . .}.

Since {xn}∞n=1 is a Cauchy sequence, for the same choice of ε there exists an integerNx(ε/2) such that

d(xi, xj) <ε

2for every i, j ≥ Nx(ε/2).

Since d is a metric,d(x0, xi) ≤ d(x0, xmk

) + d(xmk, xi).

Define M(ε) = max {Mx(ε/2), Nx(ε/2)}. Then, for every i ≥M(ε),

d(x0, xi) <ε

2+ε

2= ε.

Therefore x0 is also the limit of the sequence {xn}∞n=1.

Combining (i) and (iv) of the above theorem gives the following more commonlyexpressed statement.

Corollary 4.1. Any subsequence of a convergent sequence is convergent and convergesto the same limit as the parent sequence.

In general not every Cauchy sequence in some space will converge to a point inthat space – this we have already seen and discussed. However, there are spaces inwhich every possible Cauchy sequence does converge (to a point in that space). Theseare called complete space.

Definition 4.8 (Complete Metric Space). Let (X, d) be a metric space. (X, d) is acomplete metric space if and only if every Cauchy sequence in X is convergent in Xwith respect to d.

You will no doubt be relieved to discover that the Euclidean metric spaces are allcomplete spaces.

4.6 Remarks

Most of the time economists work in Euclidean spaces where sequences and theirlimits, if they exist, often are used to establish whether sets are closed and whetherfunctions are continuous. So, in the chapters that follow, which almost always developideas in Euclidean spaces, this is how we will use the ideas presented in this chapter,and we will do this a lot. As remarked at the end of Chapter 3, when analyzing

Page 157: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

138 CHAPTER 4. SEQUENCES AND CONVERGENCE

constrained optimization problems, it is important to determine the properties of setsof available choices and whether or not the function that reveals a decisionmaker’svaluations of available choices is continuous. Convergence of sequences is how we willtry to determine when such properties are present.

The next chapter presents an explanation of various ideas of continuity of map-pings. What sort of mappings would interest an economist? All sorts, is the answer.For example, economists talk of demand functions. These are mappings that reportquantities demanded for given prices and incomes. How does quantity demandedchange as, say, a particular price changes? Does quantity demanded change smooth-ly, for example, or does it jump suddenly? This is a continuity question. Perhapsyou are familiar with determining the rate of change of quantity demanded as a pricechanges. Such rates of change are well defined if the demand function is differentiableand may not be defined otherwise. Differential rates of change do not exist for afunction unless the function is continuous. I could make similar remarks for all sortsof functions that are important to economists. The point, then, is that continuity isa really important topic for economists interested in solving constrained optimizationproblems, and to talk of continuity we will often need to use sequences and limits.

Page 158: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

4.7. PROBLEMS 139

4.7 Problems

Problem 4.1. Consider the sequence {xi}∞i=1 = {1, 1+ 12, 1+ 2

3, . . . , 1+ i−1

i, . . .}. For

each i = 1, 2, 3, . . ., let si = sup {1 + i−1i, 1 + i

i+1, 1 + i+1

i+2, . . .} be the supremum of the

elements xi, xi+1, . . . and let si = inf {1 + i−1i, 1 + i

i+1, 1 + i+1

i+2, . . .} be their infimum.

(i) Prove that, with respect to the Euclidean metric on �, the sequence {xi}∞i=1

converges to 2.(ii) What is the sequence {si}∞i=1? Does this sequence converge with respect to theEuclidean metric on �? If so, then to what limit? What is the value of the lim sup xi?(iii) What is the sequence {si}∞i=1? Does this sequence converge with respect to theEuclidean metric on �? If so, then to what limit? What is the value of the lim inf xi?(iv) Do the answers to parts (ii) and (iii) confirm that the sequence {xi}∞i=1 is con-vergent to 2? Explain.(v) Prove that {xi}∞i=1 is a Cauchy sequence.

Problem 4.2. Consider the sequence {xi}∞i=1, in which xi =∑i

j=11j. This sequence

is monotonically increasing and is unbounded above, since∑i

j=11j→ ∞ as i→ ∞.

(i) Prove that the sequence {xi}∞i=1 is not convergent with respect to the Euclideanmetric on �.(ii) Prove that the sequence {xi}∞i=1 is not a Cauchy sequence with respect to theEuclidean metric on �.Problem 4.3. Let f : � → � be a function with the property that,

for any x′, x′′ ∈ �, dE(f(x′), f(x′′)) ≤ θdE(x′, x′′), where 0 ≤ θ < 1.

Such a mapping is called a contraction with modulus θ on �, with respect to theEuclidean metric, because the Euclidean distance between any two image values,f(x′) and f(x′′), is never larger than only a fraction θ of the Euclidean distancebetween the domain values x′ and x′′. In this sense, then, f “contracts” or “shrinks”distances. A simple example is f(x) = 1

2x. The distance between any two domain

values x′ and x′′ is |x′ − x′′|. The distance |12x′ − 1

2x′′| = 1

2|x′ − x′′| between the image

values 12x′ and 1

2x′′ is only one-half of the distance between x′ and x′′. Thus f(x) = 1

2x

is a contraction on � with modulus θ = 12.

Start with any x1 ∈ �. Then construct the infinite sequence {x1, x2, x3, x4, . . .},in which xi+1 = f(xi) for i = 1, 2, 3, . . .. That is,

x2 = f(x1),

x3 = f(x2) = f(f(x1)) = f 2(x1),

x4 = f(x3) = f(f 2(x1)) = f 3(x1),

Page 159: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

140 CHAPTER 4. SEQUENCES AND CONVERGENCE

and so on. A useful example is the function f(x) = x1/2 + 2 defined for x ≥ 1. Thisis a contraction with modulus θ = 1

2on the interval [1,∞). So let us take x′1 = 1 as

a starting value for our sequence. Then the first ten elements in the sequence are

x′1 = 1,

x′2 = f(1) = 11/2 + 2 = 3,

x′3 = f(3) = 31/2 + 2 ≈ 3·732050808,x′4 = f(3·732050808) = 3·7320508081/2 + 2 ≈ 3·931851653,x′5 = f(3·931851653) = 3·9318516531/2 + 2 ≈ 3·982889723,x′6 = f(3·982889723) = 3·9828897231/2 + 2 ≈ 3·995717847,x′7 = f(3·995717847) = 3·9957178471/2 + 2 ≈ 3·998929175,x′8 = f(3·998929175) = 3·9989291751/2 + 2 ≈ 3·999732276,x′9 = f(3·999732276) = 3·9997322761/2 + 2 ≈ 3·999933068, and

x′10 = f(3·999933068) = 3·9999330681/2 + 2 ≈ 3·999983267.

This sequence converges to 4. Now let’s try generating a new sequence, using thesame function f , but from a quite different starting value x′′1 = 300. The first tenelements in this new sequence are

x′′1 = 300,

x′′2 = f(300) = 3001/2 + 2 ≈ 19·32050808,x′′3 = f(19·32050808) = 19·320508081/2 + 2 ≈ 6·395509991,x′′4 = f(6·395509991) = 6·3955099911/2 + 2 ≈ 4·528934556,x′′5 = f(4·528934556) = 4·5289345561/2 + 2 ≈ 4·128129356,x′′6 = f(4·128129356) = 4·1281293561/2 + 2 ≈ 4·031779849,x′′7 = f(4·031779849) = 4·0317798491/2 + 2 ≈ 4·007929244,x′′8 = f(4·007929244) = 4·0079292441/2 + 2 ≈ 4·001981330,x′′9 = f(4·001981330) = 4·0019813301/2 + 2 ≈ 4·000495271, and

x′′10 = f(4·000495271) = 4·0004952711/2 + 2 ≈ 4·000123814.

This new sequence also converges to 4. In fact, no matter what starting value youtake for x1 from the set [1,∞), you will, using the above procedure, always create asequence that converges to 4. Try it for yourself. We will explore why this is so a

Page 160: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

4.8. ANSWERS 141

little later. For now it is enough to observe that the two sequences we have createdare both convergent.

The examples we have just worked through are a special case of a much moregeneral and very useful result. In the following parts of this question, we will establishthat, if f : � �→ � is a function that is a contraction with modulus θ, then the sequence{xi}∞i=1 that is generated by the algorithm xi+1 = f(xi) for i ≥ 1 is a Cauchy sequencethat converges to a limit x0 ∈ �. This limit is the solution to the equation x = f(x).

(i) Let f : � → � be a function that is a contraction with modulus θ on � withrespect to the Euclidean metric. 0 ≤ θ < 1. Prove that the sequence {xi}∞i=1 is aCauchy sequence.

(ii) Review the statements of Theorem 4.5 and parts (iii) and (iv) of Theorem 4.7.Then prove that the sequence {xi}∞i=1, in which xi+1 = f(xi) for i ≥ 1, is convergent.Use x0 to denote the limit of the sequence.

(iii) Prove that the limit x0 of the sequence {xi}∞i=1 is the solution to the equationx = f(x).

4.8 Answers

Answer to Problem 4.1.

(i) The Euclidean distance between any element xi of the sequence and the number2 is

dE(xi, 2) = |xi − 2| = 2− xi = 1− i− 1

i=

1

i.

Choose any ε > 0. Then,

dE(xi, 2) =1

i< ε if i >

1

ε=M(ε).

Thus, for any value ε > 0, no matter how small, there is a finite number M(ε) = 1/εsuch that, for every i > M(ε), every element xi of the sequence is distant from 2 byless than ε. The sequence {xi}∞i=1 is therefore convergent to 2, with respect to theEuclidean metric on �; lim xi = 2.

(ii) For any i = 1, 2, 3, . . ., the least upper bound for the sequence {1+ i−1i, 1+ i

i+1, 1+

i+1i+2, . . .} is the number 2. Thus si = 2 for i = 1, 2, 3, . . .. Clearly this constant-valued

sequence converges to its limit of 2. Thus lim sup xi = 2.

(iii) The sequence {1, 1+ 12, 1+ 2

3, . . . , 1+ i−1

i, . . .} is strictly monotonically increasing,

so for each i = 1, 2, 3, . . ., the greatest lower bound of the subsequence {1 + i−1i, 1 +

Page 161: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

142 CHAPTER 4. SEQUENCES AND CONVERGENCE

ii+1, 1+ i+1

i+2, . . .} is the first, smallest element; i.e. si = 1+ i−1

ifor i = 1, 2, 3, . . .. The

sequence of infima {si}∞i=1 = {1+ i−1i}∞i=1 is identical to the original sequence {xi}∞i=1.

We can therefore repeat the proof for part (i) to establish that the sequence {si}∞i=1

converges to 2. Thus lim inf xi = 2.

(iv) The sequence {xi}∞i=1 is convergent if and only if lim inf xi = lim sup xi. This isso, since lim inf xi = lim sup xi = 2. This confirms that lim xi = 2.

(v) The Euclidean distance between any pair of elements xi and xj in the sequenceis

dE(xi, xj) = |xi − xj| =∣∣∣∣ i− 1

i− j − 1

j

∣∣∣∣ =∣∣∣∣1j − 1

i

∣∣∣∣ .Choose any ε > 0, no matter how small. Then, for i, j > 2/ε,∣∣∣∣1j − 1

i

∣∣∣∣ < 1

j+

1

i<ε

2+ε

2= ε,

so d(xi, xj) < ε for all i, j > M(ε) = 2/ε. {xi}∞i=1 is therefore a Cauchy sequence.

Answer to Problem 4.2.

(i) Suppose that the sequence {xi}∞i=1 does have a limit; call it x0. The Euclideandistance between the fixed number x0 and the ith element in the sequence is

dE(xi, x0) = |xi − x0| =∣∣∣∣∣

i∑j=1

1

j− x0

∣∣∣∣∣→ ∞ as i→ ∞

because∑i

j=11j→ ∞ as i → ∞. Therefore, for any ε > 0 there does not exist

any integer M(ε) such that dE(xi, x0) < ε for all i > M(ε). That is, {xi}∞i=1 is notconvergent.

(ii) Without loss of generality, suppose that i > k. The Euclidean distance betweenthe ith and kth elements in the sequence is

dE(xi, xk) = |xi − xk| =∣∣∣∣∣

i∑j=1

1

j−

k∑j=1

1

j

∣∣∣∣∣ =i∑

j=k+1

1

j→ ∞ as i→ ∞.

Therefore, for any ε > 0, there does not exist any integerM(ε) such that dE(xi, xk) <ε for all i, k > M(ε). That is, {xi}∞i=1 is not a Cauchy sequence.

Answer to Problem 4.3.

(i) Choose any initial element x1 ∈ � and then generate the sequence {xi}∞i=1 usingthe algorithm that xi+1 = f(xi) for i ≥ 1. The sequence {xi}∞i=1 is a Cauchy sequence

Page 162: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

4.8. ANSWERS 143

if and only if, for any ε > 0, there exists an integer M(ε) such that d(xi, xj) < ε forevery i, j ≥M(ε).

Choose any i ≥ 2. Then, since f is a contraction with modulus θ,

dE(xi, xi+1) = dE(f(xi−1), f(xi)) ≤ θdE(xi−1, xi)

and, similarly,

dE(xi−1, xi) = dE(f(xi−2), f(xi−1)) ≤ θdE(xi−2, xi−1),

and so on. Thus

dE(xi, xi+1) ≤ θdE(xi−1, xi) ≤ θ2dE(xi−2, xi−1) ≤ · · · ≤ θi−1dE(x1, x2). (4.3)

Choose any i and k with i = k. Without loss of generality, suppose that i < k, sothat we may write k = i+Δ with Δ ≥ 1. By repeated applications of the triangularproperty of the metric function dE,

dE(xi, xi+Δ) ≤ dE(xi, xi+1) + dE(xi+1, xi+2) + · · ·+ dE(xi+Δ−1, xi+Δ). (4.4)

Using (4.3) term by term in (4.4) gives

dE(xi, xi+Δ) ≤ dE(x1, x2)(θi−1 + θi + · · ·+ θi+Δ−2

)=

θi−1

1− θ(1− θΔ)dE(x1, x2).

0 ≤ θ < 1, so θi−1 → 0 as i → ∞. Hence, for any ε > 0, there is a value Δ(ε) for Δ,and thus a value M(ε) = i+Δ(ε), such that d(xi, xk) < ε for all i, k > M(ε). {xi}∞i=1

is therefore a Cauchy sequence.

(ii) First, note from part (iii) of Theorem 4.7 that any Cauchy sequence of real num-bers is a bounded sequence. Then, from part (i) of this answer, we know that {xi}∞i=1

is a bounded sequence. The Bolzano-Weierstrass Theorem (Theorem 4.5) then tellsus that our bounded sequence {xi}∞i=1 possesses a convergent subsequence. Finally,part (iv) of Theorem 4.7 tells us that our Cauchy sequence {xi}∞i=1 is convergentbecause it possesses a convergent subsequence.

(iii) The sequence is constructed using the algorithm that xi+1 = f(xi) for all i ≥ 1.(4.3) shows that dE(xi, xi+1) = dE(xi, f(xi)) → 0 as i → ∞. Therefore the limit x0has the property that dE(x0, f(x0)) = 0, which implies that x0 = f(x0).

Problem 4.3 is a special case of Banach’s Fixed-Point Theorem, a result of hugepractical importance that was given to us in 1922 by the famous Polish mathematician

Page 163: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

144 CHAPTER 4. SEQUENCES AND CONVERGENCE

Stefan Banach. It is worth spending a few moments to understand why this resultmatters. Then, if you want more information about the theorem and its many uses,you can look elsewhere. One of my favorite references is Chapter 3 of Franklin 2002.The solution to an equation x = f(x) is called a fixed-point for f . This might seemlike a rather special equation, but it is not. In fact, the solution to any equationg(x) = 0 can be converted easily into a fixed-point equation. All we do is definef(x) ≡ g(x) + x. Then the generic equation g(x) = 0 has the same solution asthe fixed-point equation f(x) = g(x) + x = x. So one way to think of solving anyequation at all is to convert the equation into a fixed-point equation, and then usesome fixed-point solution algorithm, like Banach’s theorem, to compute numericallythe value of the solution. Banach’s Theorem applies only to contraction mappings,but there are simple ways to convert almost any mapping into a local contraction,so the theorem is more widely applicable than might at first be thought. Also, thetheorem applies not just to mappings from and to the real line but to mappings in anyBanach space; i.e. a complete and normed space. The Euclidean metric spaces areall Banach spaces, but there are all sorts of other Banach spaces, including some veryimportant function spaces, so there is an enormous set of applications for Banach’swonderful theorem.

Page 164: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

Chapter 5

Continuity

This chapter presents and explains four notions of continuity: continuity, joint con-tinuity, hemi-continuity, and semi-continuity. By necessity it is a chapter that ex-tensively uses the ideas of open sets, sequences and limits. Therefore, if you need toreview these ideas, you should do so before proceeding into this chapter.

Most economists like to think that a small change to some economic policy vari-able, such as an interest rate or a tax rate, will result in only a small change tothings that matter in the economy, such as incomes, quantities traded, total produc-tion, exchange rates, and so on. If this is true, then the relationships (mappings)between policy variables such as interest rates and tax rates, and outcome variablessuch as incomes and consumption levels, are continuous. But must this be so? Is itnot possible, perhaps, that a small change to an interest rate could cause a suddenlarge change to incomes? Certainly an economist wants to know if this is so. Inother words, every economist wants to know if important economic relationships arecontinuous mappings. It’s a big idea.

How much of this chapter should you read? You can read the chapters deal-ing with constrained optimization without knowing anything about joint continuity,hemi-continuity, and semi-continuity. The only idea of continuity you need for thosechapters is the continuity of a real-valued function. If you also want to understandthe chapter on the properties of optimal solution correspondences, then you need tounderstand the ideas of joint continuity and hemi-continuity. That means readingjust about everything in this chapter. So, you choose. Read what you want or need.You can always come back to this chapter if later on you find it necessary to do so.

Let us begin by defining some basic ideas. The first is the idea of amap ormapping.These terms mean the same thing, so we will use only “mapping.” A mapping is a

145

Page 165: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

146 CHAPTER 5. CONTINUITY

“rule of association,” meaning that a mapping associates an element from one setX with an element, or a subset of elements, in another set Y . X, the set that ismapped from, is called the domain of the mapping. Y , the set that is mapped to, iscalled the codomain of the mapping. You have met these ideas many times beforenow. We need be careful with some terminology that we will use in this and laterchapters. Mathematicians have a variety of names for different types of mappings.Unfortunately the usage of some of these names seems to have changed over time,with some names replacing others and meanings changing. It gets confusing, so let’stake a little time to define what in this book is meant by the terms that will matterto us.

5.1 Basic Ideas about Mappings

We will talk of two types of mappings: functions and correspondences. Let’s startwith the simpler of the two.

Definition 5.1 (Function). A mapping from X to Y is a function if each elementx ∈ X is associated with (mapped to) a unique element y ∈ Y .

A function is thus a “single-valued” mapping from an element of X to an elementof Y . An example is the rule of association that maps x to y = 2x; a given value forx is mapped to only a single value for y. We will typically use lower-case letters suchas f and g to denote functions. The mathematical way of writing the statement thata function f maps from a set X to a set Y is f : X �→ Y .

The rule of association that maps x ≥ 0 to√x is not a function since, for example,

x = 4 is mapped to more than one value;√4 = ±2. It would be more accurate, then,

to say that the element x = 4 is mapped to the set {−2,+2}. This is an example ofa mapping from an element in one set X to a subset of another set Y . We will callsuch a mapping a correspondence.

Definition 5.2 (Correspondence). A mapping from X to Y is a correspondence ifeach element x ∈ X is associated with (mapped to) a unique subset of Y .

A correspondence is also a “single-valued” mapping in the sense that an elementx ∈ X is mapped to just one subset of Y . But this one subset of Y can contain manyelements of Y , so you can think of a correspondence as allowing a single elementin X to be mapped to one or more elements in Y . There are correspondences forwhich any element x ∈ X is mapped to just a singleton subset of Y . There is no

Page 166: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.1. BASIC IDEAS ABOUT MAPPINGS 147

significant difference between such a correspondence and a function. Mappings thatare correspondences typically will be denoted by capital letters such as F and G.The statement that a correspondence F maps from a set X to a set Y is just as for afunction: F : X �→ Y . An example of a correspondence F : � �→ � is F (x) = [x, x+1].A domain element x is mapped to (associated with) the interval from x to x + 1 inthe codomain. The set [x, x + 1] is unique for given x, but contains many codomainelements.

To define a mapping completely we must specify both which elements in X aremapped from and which elements in Y are mapped to. These are called the domainand the range of the mapping.

The domain of a function f is the collection of all of the elements x that areactually mapped from by f .

Definition 5.3 (Domain of a Function). The domain of a function f is

{x ∈ X | there exists y ∈ Y with y = f(x)}.

For example, if Y = �, then the domain of the function f(x) = +√x (the “+”

means take only the nonnegative square root) is the set of all of the nonnegativereal numbers, �+, and is not the entire real line (complex numbers are not allowedbecause Y = � contains only real numbers).

Before we can talk of the domain of a correspondence, we must consider whatsubsets can be created from the set Y . The collection of all of the subsets that can becreated from the elements of Y is the discrete topology on Y (also called the powerset of Y ), which we will denote by P(Y ). The domain of a correspondence mappingfrom X to Y is the collection of all of the elements in X that are mapped to somenonempty subset of Y ; i.e. to some member of P(Y ) other than the empty set.

Definition 5.4 (Domain of a Correspondence). The domain of a correspondence Fis

{x ∈ X | F (x) ∈ P(Y ) \ ∅}.

For example, if Y = � the domain of the correspondence F (x) = {−√x,+

√x }

is the set �+ of the nonnegative real numbers. The \∅ bit of the above statementmeans “excluding the empty set.”

The image or range of a function is the collection of all of the elements of Y thatare mapped to by the function for at least one element in its domain.

Page 167: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

148 CHAPTER 5. CONTINUITY

Definition 5.5 (Image/Range of a Function). The image or range of a functionf : X �→ Y is

f(X) = {y ∈ Y | y = f(x) for at least one x ∈ X} =⋃x∈X

{f(x)}.

Similarly, the range of a correspondence F is the collection of all of the elementsin all of the subsets in P(Y ) that are actually mapped to by F .

Definition 5.6 (Image/Range of a Correspondence). The image or range of a cor-respondence F : X �→ Y is

F (X) =⋃x∈X

{y ∈ Y | y ∈ F (x)}.

For example, the image or range of the correspondence F (x) = ±√x, where

the square roots must be real numbers, is the union of the entire collection of pairs{(−√

x,+√x ) | x ≥ 0}; i.e. the image or range is the entire real line �. We will use

the terms image and range of a mapping interchangeably throughout the rest of thisbook.

It is possible that the range of a mapping is a strict subset of the mapping’scodomain, but it is also possible that the range and the codomain are the same set.Special language distinguishes these two cases.

Definition 5.7 (Onto/Surjective Mapping). A function f : X �→ Y is called onto orsurjective if Y = f(X). A correspondence F : X �→ Y is called onto or surjective ifY = F (X).

When a mapping is onto/surjective, every element in the codomain is mapped toby at least one element in the domain. Stated loosely, there are no “surplus” elementsin the codomain. Consider, for example, the correspondence F : [0, 1] �→ [0, 10] thatis defined by

F (x) = [x+ 1, x+ 2] for 0 ≤ x ≤ 1. (5.1)

The image of F is F ([0, 1]) = [1, 3]. There is no x ∈ X = [0, 1] that maps to any ofthe codomain elements in [0, 1)∪(3, 10]; these are “surplus” elements in the codomainY = [0, 10]. F is not onto/surjective. If the correspondence (5.1) is instead definedas F : [0, 1] �→ [1, 3], then it is onto/surjective.

It is also helpful to understand the meaning of terms such as one-to-one andinjective.

Page 168: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.1. BASIC IDEAS ABOUT MAPPINGS 149

Definition 5.8 (One-to-One or Injective Mapping). A function f : X �→ Y is calledone-to-one or injective if

∀ x′, x′′ ∈ X with x′ = x′′, f(x′) = f(x′′).

A correspondence F : X �→ Y is called one-to-one or injective if

∀ x′, x′′ ∈ X with x′ = x′′, F (x′) ∩ F (x′′) = ∅.

That is, when a mapping is one-to-one, each element in its domain is uniquelymapped to a distinct element, or a distinct set of elements, in its range. In other words,any element y in the image of the mapping is reached from only one element in thedomain of the mapping. Is the correspondence (5.1) injective? For each x ∈ [0, 1],the image F (x) is uniquely the set [x+1, x+2]. But this does not make F injective.Why? Consider the range element y = 2. This is mapped to by every one of theelements in the domain [0, 1]; i.e. 2 ∈ [x+1, x+2] for every x ∈ [0, 1]. Thus F is notinjective. In contrast, the correspondence G : �+ �→ � where G(x) = {−√

x,+√x } is

injective because, for x′ = x′′, the images G(x′) and G(x′′) have no common element;{−

√x′,+

√x′ } ∩ {−

√x′′,+

√x′′ } = ∅.

Definition 5.9 (Bijective Mapping). A mapping is bijective if it is both surjectiveand injective.

Clearly bijective mappings are special one-to-one mappings. Consider the functionf : �+ �→ � that is defined by

f(x) =1

1 + x.

The image of f is the interval (0, 1]. The codomain of f is specified as � and obviously(0, 1] is a strict subset of �. Thus f is injective but it is not surjective, so f is notbijective. Now consider the function g : �+ �→ (0, 1] that is defined by the sameequation as for f . The image and the codomain of g are the same, (0, 1], so g issurjective. We already know that g is injective, so g is bijective.

The last idea that we need for now is that of the graph of a mapping.

Definition 5.10 (Graph of a Mapping). The graph of a function f : X �→ Y is

Gf = {(x, y) | x ∈ X and y = f(x)}.

The graph of a correspondence F : X �→ Y is

GF = {(x, y) | x ∈ X and y ∈ F (x)}.

Page 169: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

150 CHAPTER 5. CONTINUITY

The graph of a mapping f or F is a subset of the direct product X × Y of thedomain and the codomain of the mapping. Some examples will help to make the ideaclear. Start with the function f : [0, 2] �→ [1, 3] that is defined by f(x) = x + 1. Thegraph of f is the subset of [0, 2]× [1, 3] ⊂ �2 that is

Gf = {(x, y) | 0 ≤ x ≤ 2 and y = x+ 1}.

This set of pairs is the straight line in �2 with endpoints at (0, 1) and (2, 3); see theleft-side panel of Figure 5.1. Now consider the correspondence F : [1, 3] �→ [1, 4] thatis defined by F (x) = [x, x + 1]. The graph of F is the subset of [1, 3] × [1, 4] ⊂ �2

that is

GF = {(x, y) | 1 ≤ x ≤ 3 and x ≤ y ≤ x+ 1}.

This is the set of points in the parallelogram with vertices at (1, 1), (1, 2), (3, 3), and(3, 4); see the right-side panel of Figure 5.1.

Figure 5.1: The graphs of f and F .

Page 170: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.2. CONTINUITY 151

5.2 Continuity

Now let us turn to the idea of the continuity of a mapping. What does it mean to saythat a mapping F is continuous at some point x0 in its domain? We consider onlyslight movements from x0 to close by values, say, x′ and x′′. Continuity of F meansthat the sets F (x0), F (x

′) and F (x′′) must be almost the same or, if you prefer,because x′ and x′′ are close to x0 the sets F (x′) and F (x′′) cannot be “suddenlydifferent” from F (x0). This description is valid for mappings from real spaces to realspaces, and it is only these mappings that we will concentrate upon in this book.But before doing so, let’s realize that continuity is an idea that is much more widelyapplicable.

We must start with a definition of continuity. As you will see in a moment,continuity of a mapping is an idea that depends upon specifications of topologies onthe domain and the range of the mapping.

Definition 5.11 (Continuous Mapping). Let (X, TX) and (Y, TY ) be topological s-paces. Let the function f : X �→ Y . Let the correspondence F : X �→ Y .

Relative to (X, TX) and (Y, TY ), f is continuous on X if and only if, for every setO ⊂ Y that is open with respect to TY , the set f−1(O) is open with respect to TX .

Relative to (X, TX) and (Y, TY ), F is continuous on X if and only if, for everyset O ⊂ Y that is open with respect to TY , the set F−1(O) is open with respect to TX .

Let’s build a weird example to get a better idea of the wide applicability of theidea of continuity. Let X = {popcorn, blue, Rome}. Let Y = {shoes, radio, pansy}.

T (X) = {∅, {popcorn}, {blue}, {Rome }, {popcorn, blue}, {popcorn, Rome},{blue, Rome}, {popcorn, blue, Rome}}

is a topology on X, so all of the sets in the collection T (X) are open with respect tothis topology. Similarly,

T (Y ) = {∅, {radio}, {shoes, radio}, {radio, pansy}, {shoes, radio, pansy}}

is a topology on Y , so all of the sets in T (Y ) are open with respect to T (Y ). Nowlet’s define the correspondence F : X �→ Y as follows:

F (x) =

⎧⎪⎨⎪⎩{radio} , x = blue

{shoes, radio} , x = popcorn

{radio, pansy} , x = Rome.

Page 171: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

152 CHAPTER 5. CONTINUITY

Strange as it may seem at first glance, F is a correspondence that is continuouswith respect to the topology T (X) on the domain X and the topology T (Y ) onthe codomain Y . Why? The set {radio} is open with respect to the codomaintopology T (Y ). The inverse image of this set is F−1({radio}) = {blue}, which is a setthat is open with respect to the domain topology T (X). Next, the set {shoes, radio}is open with respect to the codomain topology T (Y ). The inverse image of this setis F−1({shoes, radio}) = {popcorn}, which is a set that is open with respect to thedomain topology T (X). Finally, the set {radio, pansy} is open with respect to thecodomain topology T (Y ). The inverse image of this set is F−1({radio, pansy}) ={Rome}, which is a set that is open with respect to the domain topology T (X). Fsatisfies Definition 5.11 and so is a continuous correspondence relative to the twogiven topologies. Fun though this example might be, we will spend most of our timethinking about continuity in cases where both the domain X and the codomain Y aremetrizable (meaning that a metric function can be defined upon X and a, possiblydifferent, metric function can be defined on Y ).

Most of us are introduced to continuity in a case that is special for two reasons.First, the mappings are functions and, second, the mappings are from some real spaceto another real space, so think of a function f : �n �→ �m. The Euclidean metricdE can be used to measure distances on both of these spaces. When we do so, wehave constructed the Euclidean metric spaces (�n, dE) ≡ En and (�m, dE) ≡ Em.The typical basis element of the Euclidean metric topology for En is a ball (open, ofcourse) B(x, ε) of radius ε > 0 centered at a point x ∈ X. This ball is the set of allpoints that are distant (as measured by dE) from x by strictly less than ε:

B(x, ε) = {x′ ∈ X | dE(x, x′) < ε}.

The topology on X that is induced by the metric dE is constructed by taking allpossible unions of all of the balls B(x, ε) for every x ∈ X and for every ε > 0. Thusevery point in a set that is open with respect to the Euclidean metric topology on Xis surrounded by at least one such ball, and that ball is itself entirely within the openset. Similar statements apply to any set that is open with respect to the Euclideanmetric topology on Y . This allows the following more familiar statement that isequivalent to Definition 5.11 when we restrict ourselves only to functions f mappingfrom one Euclidean metric space to another Euclidean metric space.

Definition 5.12 (Continuous Function in Euclidean Metric Spaces). Let X ⊆ �n

and Y ⊆ �m. Let f : X �→ Y be a function. With respect to the Euclidean metrictopologies on �n and �m, f is continuous at x0 ∈ X if and only if, for each ε > 0,

Page 172: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.2. CONTINUITY 153

there exists δ > 0 such that

dE(x0, x′) < δ ⇒ dE(f(x0), f(x

′)) < ε.

f is continuous on X if and only if f is continuous at every x ∈ X.

A lot of people find this statement hard to read, but really it is not. What it saysis that, if x′ is “close” to x0 (distant from x0 by less than δ), then the images f(x′) andf(x0) must also be close to each other (distant from each other by less than ε), andthis “closeness” may be made as severe as one likes by making ε ever closer to zero.Put another way, the statement says that there cannot be a “break” or a “jump” inthe graph of the function at the point (x0, f(x0)). Why? Because the statement saysthat a small distance between x′ and x0 cannot result in a big distance between f(x′)and f(x0). The idea is illustrated in Figure 5.2.

Figure 5.2: In the left-side panel, f is continuous at x = x0. In the right-side panel,f is discontinuous at x = x0.

We start by choosing any (possibly very small) strictly positive value for ε. Wethen look at all of the image values f(x) that are in the codomain of f and are distantfrom f(x0) by less than ε. In the left-side panel of the figure, all of these image values

Page 173: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

154 CHAPTER 5. CONTINUITY

are contained in the vertical axis interval (f(x0)−ε, f(x0)+ε). In the right-side panel,all of these image values are contained in the vertical axis interval (f(x0)− ε, f(x0)].Then we ask, “Is there a strictly positive number δ (not necessarily small) such thatall of the x-values in the interval (x0 − δ, x0 + δ) provide image values f(x) that arenot distant from f(x0) by ε or more?” For the function displayed in the left-sidepanel, the answer is “Yes.” You see there displayed on the horizontal axis an interval(x0 − δ, x0 + δ) with the property that, for any x′ ∈ (x0 − δ, x0 + δ), the distancebetween f(x′) and f(x0) is less than ε. No matter what value we choose for ε, theanswer to the question will always be “Yes.” In the right-side panel, there is at leastone strictly positive value for ε such that the answer is “No.” For the value chosen forε in the right-side panel, there does not exist any strictly positive value for δ such thatf(x) is distant from f(x0) by less than ε for every value x ∈ (x0 − ε, x0 + ε). Thereare values x′′ larger than x0 and distant from x0 by less than δ with image valuesf(x′′) that are distant from f(x0) by more than ε. The source of the problem is the“jump” in the value of the function at x = x0. Such a function is thus discontinuousat x = x0.

It is important to understand that Definition 5.12 makes exactly the same state-ment as Definition 5.11 when the mapping being considered is a function that mapsfrom one Euclidean metric space to another. Why is this so? Definition 5.12 says tochoose a point x0 and to choose a strictly positive distance ε. This allows us to definethe ball (in the codomain �m of f),

BY (f(x0), ε/2) = {y ∈ Y | dE(f(x0), y) < ε/2}.

This ball is a set that is open with respect to the Euclidean metric topology on �m.The ball contains all points in Y that are distant from f(x0) by less than the givenvalue ε/2. From Definition 5.11, f is continuous at x0 if and only if the inverse imageof the open set BY (f(x0), ε/2) is a subset of X that is open with respect to theEuclidean metric topology on the domain �n. Necessarily, x0 ∈ f−1(BY (f(x0), ε/2)),and so, because f−1(BY (f(x0), ε/2)) is an open set, there must exist a ball of somestrictly positive radius δ/2 containing all the points in X that are distant from x0 bystrictly less than δ/2. This is the ball (in the domain of f),

BX(x0, δ/2) = {x ∈ X | dE(x0, x) < δ/2}.

Now note that, if we take any two points x′, x′′ ∈ BX(x0, δ/2), then, by the triangularinequality property of a metric function, the distance between x′ and x′′ is

dE(x′, x′′) ≤ dE(x

′, x0) + dE(x0, x′′) <

δ

2+δ

2= δ.

Page 174: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.3. SEMI-CONTINUITY AND CONTINUITY 155

x′, x′′ ∈ BX(x0, δ/2) if and only if f(x′) and f(x′′) both are elements of BY (f(x0), ε/2),so, again by the triangular inequality property of a metric function, the distancebetween f(x′) and f(x′′) is

dE(f(x′), f(x′′)) ≤ dE(f(x

′), f(x0)) + dE(f(x0), f(x′′)) <

ε

2+ε

2= ε.

Hence,dE(x

′, x′′) < δ ⇒ dE(f(x′), f(x′′)) < ε.

This is the statement made in Definition 5.12. The two definitions coincide becausesets that are open with respect to Euclidean topologies are always unions of (open)balls. In other topologies this need not be so; in any such case, Definition 5.12 is notequivalent to Definition 5.11.

5.3 Semi-Continuity and Continuity

Like full continuity, the ideas of lower-semi-continuity and upper-semi-continuity aredefined with respect to topologies. However, they are much more restrictive thanis continuity. Upper-semi-continuity and lower-semi-continuity are ideas that applyonly to functions and, particularly, only to real-valued functions.

Definition 5.13 (Upper-Semi-Continuous Function). Let (X, TX) be a topologicalspace. Let Y ⊂ �. Let f : X �→ Y be a function. Relative to TX and the Euclideanmetric topology on the real line, f is upper-semi-continuous at x0 ∈ X if and only ifthe set {x ∈ X | f(x) < f(x0)} is open with respect to TX . f is upper-semi-continuouson X if and only if f is upper-semi-continuous at every x ∈ X.

Definition 5.14 (Lower-Semi-Continuous Function). Let (X, TX) be a topologicalspace. Let Y ⊂ �. Let f : X �→ Y be a function. Relative to TX and the Euclideanmetric topology on the real line, f is lower-semi-continuous at x0 ∈ X if and only ifthe set {x ∈ X | f(x) > f(x0)} is open with respect to TX . f is lower-semi-continuouson X if and only if f is lower-semi-continuous at every x ∈ X.

Figure 5.3 displays the graph of an upper-semi-continuous function. Notice that,at x = x0, the image f(x0) is at the upper end of the jump discontinuity. This causesthe set {x ∈ � | f(x) < f(x0)} not to include x = x0, thereby making this set openwith respect to E1. Figure 5.4 displays the graph of a lower-semi-continuous function.Notice that the image f(x0) is at the lower end of the discontinuity at x = x0. This

Page 175: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

156 CHAPTER 5. CONTINUITY

Figure 5.3: A graph of an upper-semi-continuous function.

causes the set {x ∈ � | f(x) > f(x0)} not to include x = x0, so that the set is openwith respect to E1.

Here is a little teaser for you. Look again at Figure 5.3 and think of a horizontalline drawn at the height that is the right-side limit of f(x) as x→ x0 from above; i.e.limx→x+

0f(x). Is the set {x | f(x) < limx→x+

0f(x)} open with respect to E1? Think

about it. When you are ready, look at Problem 5.6 and its answer.

For a function to be continuous at a point x = x0 there cannot be discontinuitiesof the type displayed in Figures 5.3 and 5.4, in which case the sets {x ∈ � | f(x) <f(x0)} and {x ∈ � | f(x) > f(x0)} will both be open with respect to E1. This is thecontent of the following theorem, which we could, if we wished, use as an alternativeway of defining continuity for a real-valued function.

Theorem 5.1. Let (X, TX) be a topological space. Let Y ⊂ �. Let f : X �→ Y bea function. Relative to TX and the Euclidean metric topology on the real line, f iscontinuous at x0 ∈ X if and only f is both lower-semi-continuous and upper-semi-continuous at x0. f is continuous on X if and only if f is both lower-semi-continuousand upper-semi-continuous on X.

The theorem offers an alternative way of defining continuity of a real-valued func-tion (but not of more general mappings). How? Look again at Definition 5.11. Once

Page 176: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.3. SEMI-CONTINUITY AND CONTINUITY 157

Figure 5.4: A graph of a lower-semi-continuous function.

we insist that the function f is real-valued, Definition 5.11 becomes the special casethat we will call

Definition 5.15 (Continuous Real-Valued Function). Let (X, TX) be a topologicalspace. Let f : X �→ � be a function. Relative to (X, TX) and the Euclidean metrictopology on the real line, f is continuous at x0 ∈ X if and only if, for every openneighborhood BdE(f(x0), ε) about f(x0), the set f

−1(BdE(f(x0), ε)) is open with respectto TX .

Let’s look closely at this statement and understand why it is equivalent to Defi-nition 5.11 for the special case of real-valued functions. BdE(f(x0), ε) is a subset ofthe codomain � that contains the point f(x0) and is open with respect to E1. Be-cause the topology on the codomain is the Euclidean metric topology, for every pointy ∈ BdE(f(x0), ε), there must exist a ball with a strictly positive radius ε(y) > 0 thatis centered on y and is entirely contained in BdE(f(x0), ε). In fact, the neighborhoodBdE(f(x0), ε) is the union of all such balls:

BdE(f(x0), ε) =⋃

y∈BdE(f(x0), ε)

BdE(y, ε(y)).

What is the inverse image of any one of these balls? It is f−1(BdE(y, ε(y))). Sincef is continuous and since BdE(y, ε(y)) is open with respect to the topology on the

Page 177: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

158 CHAPTER 5. CONTINUITY

codomain, f−1(BdE(y, ε(y))) must be open with respect to the topology on the do-main. The inverse image of any union of sets in the codomain of a function is theunion of the inverse images of these sets in the domain of the function, so,

f−1(BdE(f(x0), ε)) =⋃

y∈BdE(f(x0), ε)

f−1(BdE(y, ε(y))).

Because any union of sets open with respect to a topology is a set that is open withrespect to that topology, the set f−1(BdE(f(x0), ε)) must be open with respect to TX .

How does Definition 5.15 lead to Theorem 5.1? Think of any one of the ballsBdE(y, ε(y)). Since the codomain of f is the real line, the ball is simply an openinterval of the line:

BdE(y, ε(y)) = {y′ ∈ Y | dE(y, y′) < ε(y)}= (y − ε(y), y + ε(y))

= (−∞, y + ε(y)) ∩ (y − ε(y),∞).

Hence BdE(y, ε(y)) is open with respect to E1 if and only if both of the sets (y −ε(y),∞) and (−∞, y+ε(y)) are open with respect to E1. This is true if and only if f isupper-semi-continuous at f−1(y+ε(y)) and is lower-semi-continuous at f−1(y−ε(y)).

Now think of any set O ⊂ Y that is open with respect to E1. This subset consistsonly of points that are interior to the set, which, in the Euclidean metric topologyon �, means that each point y ∈ O is contained in an open ball BdE(y, ε(y)) that isitself entirely within O. Moreover, O must be the union of all such balls. Thus,

O =⋃y∈O

BdE(y, ε(y))

=⋃y∈O

{(y − ε(y),∞) ∩ (−∞, y + ε(y))}.

O is open with respect to E1 if and only if every one of the intervals (−∞, y + ε(y))and (y− ε(y),∞) is open with respect to E1, for every y ∈ O. For f to be continuouson X, we need any such subset O to be open with respect to E1. This is true if andonly if f is both upper-semi-continuous and lower-semi-continuous at every pointy ∈ Y . This concludes the argument that establishes Theorem 5.1.

5.4 Hemi-Continuity and Continuity

The ideas of lower-hemi-continuity and upper-hemi-continuity have little to do withupper-semi-continuity and lower-semi-continuity. It is therefore important not to

Page 178: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.4. HEMI-CONTINUITY AND CONTINUITY 159

confuse the hemi-continuities with the semi-continuities. Be warned that in someolder documents the term semi-continuity is used to mean what we here mean byhemi-continuity, so when you encounter the term semi-continuity, it is prudent totake a moment to check exactly what the author means by it.

Like continuity, upper-hemi-continuity and lower-hemi-continuity are ideas thatare defined with respect to topologies. They are rather general notions, but, as weshall see, they do not apply to all types of correspondences. Each hemi-continuity, byitself, is a more general idea than is full continuity since, for a correspondence to befully continuous, it must be both upper-hemi-continuous and lower-hemi-continuous.Neither an upper-hemi-continuous correspondence nor a lower-hemi-continuous cor-respondence need be (fully) continuous.

Why do we need to think about these hemi-continuity ideas at all? Do they haveany practical merit? When we think of only functions, the idea of continuity is quitesimple. We say that a function (remember, a function is a single-valued mapping) iscontinuous at some point x0 ∈ X if and only if, for every set O ⊂ Y that is openin the topology on the codomain with f(x0) ∈ O, the inverse image set f−1(O) isopen in the topology on the domain and x0 ∈ f−1(O). When f is a function, the setf−1(O) has just one meaning: it is the set of all of the domain elements that map tothe elements in O. Unfortunately, when the mapping is a correspondence F that isnot a function, there are two natural ways to think of the inverse image set F−1(O).A couple of examples will help you to understand the choice that is now before us.

Start by thinking of the function f : [0, 4] �→ [1, 5] that is defined by f(x) = x+1.Let O = (2, 4). O is open in the codomain with respect to the Euclidean metrictopology. The set f−1(O) = (1, 3), a set that is open in the domain with respect to theEuclidean metric topology. There is no question as to what is the set f−1(O), but thereare two ways to interpret its meaning. Have a look at Figure 5.5. One interpretationis that f−1(O) is the collection of all values of x ∈ [0, 4] for which the image set f(x)is contained in (2, 4); i.e. x ∈ f−1(O) if and only if {f(x)} ⊂ O. An alternativeinterpretation is that f−1(O) is the collection of all values of x ∈ [0, 4] for whichthe image set f(x) intersects (2, 4); i.e. x ∈ f−1(O) if and only if {f(x)} ∩ O = ∅.Because f is single-valued, these two different interpretations of the inverse image ofthe set O = (2, 4) give us the same value, the interval (1, 3), so we do not have totrouble ourselves over which of these two interpretations is “correct.” Make sure youunderstand the statements of this paragraph before you proceed further.

Now think of the correspondence F : [0, 4] �→ [1, 6] that is defined by F (x) =[x + 1, x + 2]. Again let O = (2, 4). Look at Figure 5.6. What is now meant byF−1(O)? Suppose we use the interpretation that F−1(O) is the collection of all values

Page 179: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

160 CHAPTER 5. CONTINUITY

Figure 5.5: f−1(2, 4) = (1, 3) = {x ∈ [0, 4] | {x + 1} ⊂ (2, 4)} = {x ∈ [0, 4] |{x+ 1} ∩ (2, 4) = ∅}.

of x ∈ [0, 4] for which the image set F (x) is contained in (2, 4); i.e. x ∈ F−1(O) if andonly if F (x) ⊂ O. Then F−1((2, 4)) = (1, 2). The alternative interpretation is thatF−1(O) is the collection of all values of x ∈ [0, 4] for which F (x) intersects (2, 4); i.e.x ∈ F−1(O) if and only if F (x) ∩O = ∅. Then F−1((2, 4)) = (0, 3). Obviously, (1, 2)and (0, 3) are different sets. Now it really does matter which way we interpret inverseimages. As you will see shortly, one interpretation leads to upper-hemi-continuitywhile the other leads to lower-hemi-continuity.

When the inverse image F−1(O) is taken to mean the set of all elements x in thedomain X such that the image of each x is completely contained in the set O, weobtain the set called the upper inverse of F at the set O. This we will denote byF−1(O)+:

F−1(O)+ = {x ∈ X | F (x) ⊂ O}. (5.2)

It is the set of all domain elements that are mapped to sets in the codomain thatcontain only elements in the set O. We could, therefore, choose to define a corre-spondence F to be “continuous” at a point x0 ∈ X if and only if, for every set O thatis open in the topology on the codomain of F and contains F (x0), the upper inverseF−1(O)+ is a set that is open in the topology on the domain of F . This idea of “con-tinuity” that arises from using the upper inverse of F is called upper-hemi-continuity.

Page 180: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.4. HEMI-CONTINUITY AND CONTINUITY 161

Figure 5.6: F−1((2, 4))+ = {x ∈ [0, 4] | [x + 1, x + 2] ⊂ (2, 4)} = (1, 2), butF−1((2, 4))− = {x ∈ [0, 4] | [x+ 1, x+ 2] ∩ (2, 4) = ∅} = (0, 3).

When the inverse image F−1(O) is taken to mean the set of every element x inthe domain X such that the image F (x) intersects O, we obtain the set called thelower inverse of F at the set O. This will be denoted by F−1(O)−:

F−1(O)− = {x ∈ X | F (x) ∩O = ∅}. (5.3)

It is the set of all domain elements that are mapped to sets in the codomain thatcontain at least one of the elements in the set O. Thus we could instead chooseto define a correspondence F to be “continuous” at a point x0 ∈ X if and only if,for every set O that is open in the topology on the codomain of F and containsF (x0), the lower inverse F

−1(O)− is a set that is open in the topology on the domainof F . This idea of “continuity” that arises from using the lower inverse of F is calledlower-hemi-continuity.

For any nonempty image set F (x), F (x) ⊂ O implies that F (x) ∩ O = ∅. Theconverse is not true. Hence the upper-inverse of F at O is always a subset of thelower-inverse of F at O; F−1(O)+ ⊆ F−1(O)−.

Page 181: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

162 CHAPTER 5. CONTINUITY

The term “continuous correspondence” is reserved only for correspondences thatare both upper-hemi-continuous and lower-hemi-continuous. Therefore, to under-stand clearly what is meant by the (full) continuity of a correspondence, it is nec-essary to understand both upper-hemi-continuity and lower-hemi-continuity. Let’sstart by thinking about upper-hemi-continuity.

Definition 5.16 (Upper-Hemi-Continuous Correspondence). Let (X, TX) and (Y, TY )be topological spaces. Let F : X �→ Y . Relative to (X, TX) and (Y, TY ), F is upper-hemi-continuous at x0 ∈ X if and only if, for every set O that is open in TY withF (x0) ⊂ O, there exists a set V ⊂ X that is open with respect to TX and x ∈ Vimplies that F (x) ⊂ O. F is upper-hemi-continuous on X if and only if F is upper-hemi-continuous at every x ∈ X.

Figure 5.7: Upper-hemi-continuity at x = x0.

This definition uses (5.2) as its interpretation of the inverse image F−1(O). Theidea is quite simple. Take a point x0 in the domain of F . Then look at its image F (x0).This is a set in the range Y ; see Figure 5.7. Now take any open set O that containsthe set F (x0). Think of O as setting boundaries on F (x0). Loosely stated, F is upper-hemi-continuous at x0 if and only if, for every x that is “close” to x0, the set F (x)

Page 182: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.4. HEMI-CONTINUITY AND CONTINUITY 163

is “bounded on the outside” (in the sense of complete containment) by that sameopen set O. Now recollect that O is any open set in Y that contains F (x0), so, inparticular, think of the open sets containing F (x0) that are almost the same as F (x0)(think of “shrink-wrapping” F (x0), as is displayed in Figure 5.7, where F (x) ⊂ O forevery x in the (open) ball of radius ε centered at x0. O encloses F (x0) more tightly asε → 0). That means that the set F (x0) sets the outer boundaries mentioned above,so F is upper-hemi-continuous at x0 if and only if, for all x ∈ X that are really closeto x0, any set F (x) that extends in any direction by more than does F (x0) does soby only very slightly more – see the sets F (x) in Figure 5.7 for x just above x0. Butfor x close to x0, the sets F (x) can suddenly be much smaller than the set F (x0) –see the sets F (x) in Figure 5.7 for x just below x0.

Lower-hemi-continuity has the opposite meaning to upper-hemi-continuity. As weshall see, when F is lower-hemi-continuous at x0, the set F (x0) sets inner boundariesfor all of the sets F (x) for x “close” to x0. Particularly, for every x close to x0, anyset F (x) that extends in some direction less than does F (x0) must do so only veryslightly – see the sets F (x) in Figure 5.8 for x just above x0. But for x close to x0,the sets F (x) can suddenly be much larger than the set F (x0) – see the sets F (x) inFigure 5.8 for x just below x0.

Definition 5.17 (Lower-Hemi-Continuous Correspondence). Let (X, TX) and (Y, TY )be topological spaces. Let F : X �→ Y . Relative to (X, TX) and (Y, TY ), F is lower-hemi-continuous at x0 ∈ X if and only if, for every set O that is open in TY withF (x0) ∩ O = ∅, there exists a set V ⊂ X that is open with respect to TX and x ∈ Vimplies that F (x) ∩ O = ∅. F is lower-hemi-continuous on X if and only if F islower-hemi-continuous at every x ∈ X.

This definition uses (5.3) as its interpretation of the inverse image F−1(O). Asan example to help to visualize the meaning of lower-hemi-continuity, think of F (x0)as being an interval in Y , as displayed in Figure 5.8. Then think of a long but finitesequence of very small open sets {Oj}nj=1 such that every one of these sets intersectsF (x0) and their union includes (i.e. covers) F (x0):

F (x0) ∩Oj = ∅ ∀ j = 1, . . . , n and F (x0) ⊂n⋃

j=1

Oj.

Lower-hemi-continuity insists that, for each one of the open sets Oj that intersectF (x0), for every x that is really close to x0 the set F (x) must also intersect each oneof the open sets Oj. Picture in your mind that F (x0) is an interval of finite length in

Page 183: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

164 CHAPTER 5. CONTINUITY

the vertical axis, that each open set Oj is a really small ball in the vertical axis thatis centered somewhere in F (x0), that each of these tiny open balls overlaps its twoneighbors, and that their union covers (i.e. contains) all of the set F (x0) like a stringof overlapping links in a chain. In Figure 5.8 the union of the overlapping intervals(balls) O1, O2, and O3 contains F (x0). The set of x-values for which F (x) ∩ O1 = ∅is the open interval V1. Similarly, the sets of x-values for which F (x) ∩ O2 = ∅and F (x) ∩ O3 = ∅ are the open intervals V2 and V3. If O = O1 ∪ O2 ∪ O3, thenV = V1 ∪ V2 ∪ V3.

Figure 5.8: Lower-hemi-continuity at x = x0.

Lower-hemi-continuity insists that, for every open set Oj that intersects F (x0),all of the sets F (x) must also intersect the same open set Oj for every x that is reallyclose to x0. Thus F (x) must intersect every single Oj and so must either be almostthe same as F (x0) or else entirely enclose F (x0). In this sense, F (x0) is like an “innerbound” for each one of the sets F (x). Look again at Figure 5.8. Notice that everyvalue of x in V1 that is “close” to x0 creates an image F (x) that intersects each of O1,O2, and O3. Similarly, for j = 2, 3, every value of x in Vj that is “close” to x0 createsan image F (x) that intersects each of O1, O2, and O3. We are allowed to choose theintervals O1, O2, and O3 so this is only just true, in which case the union of O1, O2,and O3 is an open set that essentially limits how small F (x) can be for values of xthat are really close to x0.

Page 184: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.4. HEMI-CONTINUITY AND CONTINUITY 165

What is the meaning of continuity of a correspondence? The basic idea of continu-ity is that a small change to x from an initial value x′ to a new value x′′ will result ina new image set F (x′′) that is only slightly different from the initial image set F (x′).In other words, a small change to x must cause only a small change to F (x). We haveseen that upper-hemi-continuity allows the change from x′ to x′′ to cause the imageset of F to explode suddenly, but not by too much, since any sudden jump in the sizeor position of F (x) will be bounded on the outside by an open set that just containsthe set F (x0). We have also seen that lower-hemi-continuity allows the small changefrom x′ to x′′ to cause the image set of F to implode suddenly, but not by too much,since any sudden jump in the size or position of F (x) will be bounded on the insideby F (x0). It seems reasonable to say that a correspondence F is fully continuous ifand only if neither such explosions nor implosions occur when a small change is madeto x. In other words, we define a correspondence to be continuous at a point x0 ifand only if it is both upper-hemi-continuous and lower-hemi-continuous at x0.

Definition 5.18 (Continuous Correspondence). Let (X, TX) and (Y, TY ) be topolog-ical spaces. Let F : X �→ Y . Relative to (X, TX) and (Y, TY ), F is continuous atx0 ∈ X if and only if F is both upper-hemi-continuous and lower-hemi-continuousat x0. F is continuous on X if and only if F is continuous for every x ∈ X.

Definition 5.18 might seem to have little to do with Definition 5.11, the correspon-dence part of which is restated below for your convenience. Yet both are definitions ofthe same thing – namely, the continuity of a correspondence at a point in its domain.Therefore the two statements must be equivalent.

Definition 5.11′ (Continuous Correspondence). Let (X, TX) and (Y, TY ) be topolog-ical spaces. Let the correspondence F : X �→ Y . Relative to (X, TX) and (Y, TY ), Fis continuous if and only if, for every set O ⊂ Y that is open with respect to TY theset F−1(O) is open with respect to TX .

According to Definition 5.18, a correspondence is continuous at a domain pointx0 if and only if both Definitions 5.16 and 5.17 are satisfied. That means that, forany set O ⊂ Y that is open in the codomain of F and contains F (x0), there mustexist a set V ⊂ X that is open in the domain of F and x ∈ V implies F (x) ⊂ Oand also that, for any set O′ ⊂ Y that is open in the codomain of F and intersectsF (x0), there must exist a set V ′ ⊂ X that is open in the domain of F and x ∈ V ′

implies F (x) ∩ O′ = ∅. These two requirements reduce to the statement made forcorrespondences in Definition 5.11. Why?

The requirement that F (x) ∩O = ∅ ensures that F (x) = ∅; i.e. x must be in thedomain of F . But why does it also imply that the image set F (x0) is not a sudden

Page 185: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

166 CHAPTER 5. CONTINUITY

change in size or position compared to F (x) for x very close to x0? Again think ofa long but finite sequence {Oj}nj=1 of sets open in the codomain topology with theproperties that

Oj ∩ F (x0) = ∅ ∀ j = 1, . . . , n andn⋃

j=1

Oj = O and F (x0) ⊂ O. (5.4)

The picture is again of a long line of finitely many small chain links Oj, with each linkoverlapping with F (x0) by a bit and with their union O completely covering F (x0),but only just. Think again of O as a “shrink-wrapping” of F (x0). Look again at (5.4)and convince yourself that F (x0) is such a “chain-link-covered” set – draw a pictureof it to help you follow what comes next.

Suppose that, for some x very close to x0, the image set F (x) is significantly smallerthan F (x0). Then some of the little chain links Oj do not intersect with F (x). Butif F is continuous at x0, then, from (5.4), this cannot happen.

Now suppose that, for some x very close to x0, the image set F (x) is significantlylarger than F (x0). Then F (x) ⊂ O. But if F is continuous at x0, then, from (5.4),this cannot happen.

Once you put the conclusions of the above two paragraphs together, it becomesclear that continuity of F at x0 means that, as neighboring values of x become veryclose to x0, the image sets F (x) must become very close to F (x0).

Some examples of correspondences will help to make these ideas clearer to you.Consider the correspondence

F1 : � �→ � that is F1(x) = [1, 2] ∀ x ∈ �. (5.5)

Draw the graph before reading any further. Is this correspondence upper-hemi-continuous at some point x0 ∈ �? The image F1(x0) = [1, 2]. Take any set Othat is open in the codomain with respect to E1 and contains the closed interval[1, 2]. Since O is open and [1, 2] is closed, O must strictly include [1, 2]; i.e. O mustbe a “bit vertically longer” than [1, 2] at both ends of [1, 2]. What is the upper-inverseof O? It is

F−11 (O)+ = {x ∈ � | F1(x) ⊂ O} = �. (5.6)

� is an open subset of the domain � with respect to E1, so F1 is upper-hemi-continuous at x0. Since x0 is an arbitrary choice from the domain �, F1 is upper-hemi-continuous on �. Is the correspondence lower-hemi-continuous at some pointx0 ∈ �? The image F1(x0) = [1, 2]. Take any set O′ that is open in the codomain

Page 186: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.4. HEMI-CONTINUITY AND CONTINUITY 167

with respect to E1 and intersects the closed interval [1, 2]; O′ ∩ [1, 2] = ∅. What isthe lower-inverse of O′? It is

F−11 (O′)− = {x ∈ � | F1(x) ∩O′ = ∅} = �. (5.7)

� is an open subset of the domain � with respect to E1, so F1 is lower-hemi-continuousat x0. Since x0 is an arbitrary choice from the domain �, F1 is lower-hemi-continuouson �. The correspondence F1 is therefore (fully) continuous on �. This is hardlya surprise, since you would expect, I hope, that any constant-valued correspondencewould be continuous.

Now consider the correspondence

F2 : � �→ � that is F2(x) = [x, x+ k], where k > 0, ∀ x ∈ �. (5.8)

Draw the graph before reading any further. Is this correspondence upper-hemi-continuous at some point x0 ∈ �? The image F2(x0) = [x0, x0 + k]. Take anyset O that is open in the range with respect to E1 and contains the closed inter-val [x0, x0 + k]. Since O is open and [x0, x0 + k] is closed, O must strictly include[x0, x0 + k]; i.e. O must be a bit vertically longer than [x0, x0 + k] at both ends of[x0, x0+k], so take O = (a, b), where a < x0 and x0+k < b. What is the upper-inverseof O? It is

F−12 (O)+ = {x ∈ � | F2(x) ⊂ (a, b)} = (a, b− k). (5.9)

(a, b − k) is an open subset of the domain � with respect to E1. We can make asimilar argument for any open subset O that contains [x0, x0 + k], so F2 is upper-hemi-continuous at x0. Since x0 is an arbitrary choice from the domain �, F2 isupper-hemi-continuous on �. Is the correspondence lower-hemi-continuous at somepoint x0 ∈ �? The image F2(x0) = [x0, x0 + k]. Take any set O′ that is openin the codomain with respect to E1 and intersects the closed interval [x0, x0 + k];O′ ∩ [x0, x0 + k] = ∅. Take O′ = (c, d), where c < x0 < d < x0 + k. What is thelower-inverse of O′? It is

F−12 (O′)− = {x ∈ � | [x, x+ k] ∩ (c, d) = ∅} = (c− k, d). (5.10)

(c − k, d) is an open subset of the domain � with respect to E1. We can make asimilar argument for any open subset O′ that intersects [x0, x0 + k], so F2 is lower-hemi-continuous at x0. Since x0 is an arbitrary choice from the domain �, F2 islower-hemi-continuous on �. The correspondence F2 is therefore (fully) continuouson �.

Page 187: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

168 CHAPTER 5. CONTINUITY

Notice that the two correspondences that we have considered so far are bothclosed-valued; i.e. F1(x) and F2(x) are both closed sets in their codomains for everyx in their domains. Does this matter? Consider

F3 : � �→ � that is F3(x) = (1, 2) ∀ x ∈ �. (5.11)

Draw the graph before reading any further. Is this correspondence upper-hemi-continuous at some point x0 ∈ �? The image F3(x0) = (1, 2). Take any set Othat is open in the codomain with respect to E1 and contains the open interval (1, 2).Since O is open and (1, 2) is open, the smallest such set O is (1, 2) itself, so takeO = (a, b), where a ≤ 1 and 2 ≤ b. What is the upper-inverse of O? It is

F−13 (O)+ = {x ∈ � | F3(x) = (1, 2) ⊂ (a, b)} = �. (5.12)

� is an open subset of the domain � with respect to E1, so F3 is upper-hemi-continuous at x0. Since x0 is an arbitrary choice from the domain �, F3 is upper-hemi-continuous on �. Is the correspondence lower-hemi-continuous at some pointx0 ∈ �? The image F3(x0) = (1, 2). Take any set O′ that is open in the codomainwith respect to E1 and intersects the open interval (1, 2); O′ ∩ (1, 2) = ∅. What isthe lower-inverse of O′? It is

F−13 (O′)− = {x ∈ � | F3(x) ∩O′ = (1, 2) ∩O′ = ∅} = �. (5.13)

� is an open subset of the domain � with respect to E1, so F3 is lower-hemi-continuousat x0. Since x0 is an arbitrary choice from the domain �, F3 is lower-hemi-continuouson �. The correspondence F3 is therefore (fully) continuous on �. This exampleestablishes that neither an upper-hemi-continuous correspondence, nor a lower-hemi-continuous correspondence, nor a continuous correspondence has to be closed-valued.

Another instructive example is the correspondence

F4 : � �→ � that is F4(x) = (x, x+ k) where k > 0, ∀ x ∈ �. (5.14)

Draw the graph before reading any further. Is this correspondence upper-hemi-continuous at some point x0 ∈ �? The image F4(x0) = (x0, x0 + k). Take anyset O that is open in the codomain with respect to E1 and contains the open interval(x0, x0 + k). Particularly, choose O = (x0, x0 + k). What is the upper-inverse of O?It is

F−14 (O)+ = {x ∈ � | F4(x) ⊂ (x0, x0 + k))} = {x0}. (5.15)

Page 188: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.5. UPPER-HEMI-CONTINUITY BY ITSELF 169

{x0} is a closed subset of the domain � with respect to E1, so F4 is not upper-hemi-continuous at x0 and therefore is not upper-hemi-continuous on �. The corre-spondence F4 is therefore not continuous on �. Is the correspondence lower-hemi-continuous at some point x0 ∈ �? The image F4(x0) = (x0, x0 + k). Take any setO′ that is open in the codomain with respect to E1 and intersects the open interval(x0, x0 + k); O′ ∩ (x0, x0 + k) = ∅. Suppose, for example, that O′ = (c, d), wherec < x0 < d < x0 + k. What is the lower-inverse of O′? It is

F−14 (O′)− = {x ∈ � | (x, x+ k) ∩ (c, d) = ∅} = (c− k, d). (5.16)

(c − k, d) is an open subset of the domain � with respect to E1. We can make asimilar argument for any open subset O′ that intersects (x0, x0 + k), so F4 is lower-hemi-continuous at x0. Since x0 is an arbitrary choice from the domain �, F4 islower-hemi-continuous on �.

Lastly, do you remember an example earlier in this section (see Figure 5.5) inwhich we concluded that, for a function, the upper-inverse and the lower-inverse setsare always the same? This is true for any function. With this in mind, construct aproof for the following theorem. It’s an easy task.

Theorem 5.2 (Hemi-Continuous Function is Continuous). Let (X, TX) and (Y, TY )be topological spaces. Let f : X �→ Y be a function.

(i) If f is upper-hemi-continuous on X with respect to TX and TY , then f iscontinuous on X with respect to TX and TY .

(ii) If f is lower-hemi-continuous on X with respect to TX and TY , then f iscontinuous on X with respect to TX and TY .

5.5 Upper-Hemi-Continuity by Itself

The properties possessed by any correspondence depend greatly upon the types oftopologies defined upon the domain and the range of the correspondence. In manyinstances an economist will be interested in mappings from one metric space to anoth-er. It turns out that upper-hemi-continuous correspondences have some rather niceproperties when a metric topology is defined on the domain and a metric topology isdefined on the range. Several of these are listed below.

Let’s start by understanding the statement that a correspondence F : X �→ Y iscompact-valued. It means that, for any domain element x ∈ X, the image F (x) is aset that is compact with respect to the topology on the codomain Y . We will often

Page 189: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

170 CHAPTER 5. CONTINUITY

make reference to such correspondences in this section. One important theorem thatunderlies many of the others that will be stated here is the following.

Theorem 5.3. Let (X, dX) and (Y, dY ) be metric spaces. Let F : X �→ Y be anupper-hemi-continuous correspondence. If K ⊂ X is compact with respect to themetric topology on X, then F (K) is compact with respect to the metric topology on Y .

Now we can proceed to examine one of the most commonly used characterizationsof upper-hemi-continuous correspondences that map from a metric space to a metricspace.

Theorem 5.4. Let (X, dX) and (Y, dY ) be metric spaces. Let F : X �→ Y be acorrespondence.

(a) F is upper-hemi-continuous at x0 ∈ X if, for any sequence {xn}∞n=1 withxn ∈ X for every n ≥ 1 and xn → x0, any sequence {yn}∞n=1 with yn ∈ F (xn) forevery n ≥ 1 possesses a subsequence {ynk

}∞k=1 such that ynk→ y0 ∈ F (x0).

(b) If F is compact-valued, then the converse of (a) is true; i.e. F is upper-hemi-continuous at x0 ∈ X only if, for any sequence {xn}∞n=1 with xn ∈ X for every n ≥ 1and xn → x0, any sequence {yn}∞n=1 with yn ∈ F (xn) for every n ≥ 1 possesses asubsequence {ynk

}∞k=1 such that ynk→ y0 ∈ F (x0).

Part (a) is a statement of sufficiency. Part (b) is a statement of necessity. We willsee why these two statements are true in a moment, but for now note that the abovetheorem does not state that every correspondence F mapping from one metric spaceto another is upper-hemi-continuous at x0 ∈ X if and only if, for any sequence {xn}∞n=1

with xn ∈ X for every n ≥ 1 and xn → x0, any sequence {yn}∞n=1 with yn ∈ F (xn) forevery n ≥ 1 possesses a subsequence {ynk

}∞k=1 such that ynk→ y0 ∈ F (x0). Believing

this to be true is a common error.Let’s have a closer look at part (a) of Theorem 5.4. The idea is illustrated in

Figure 5.9. Take as true that, for any sequence {xn}∞n=1 with xn ∈ X for every n ≥ 1and xn → x0, any sequence {yn}∞n=1 with yn ∈ F (xn) for every n ≥ 1 possesses asubsequence {ynk

}∞k=1 such that ynk→ y0 ∈ F (x0). Why is it that F must then be

upper-hemi-continuous at x0? The sequence xn → x0. This generates a sequenceof image sets F (xn). We are allowed to select any one yn from each set F (xn),and are assured that when we do so the sequence that we have created of valuesyn ∈ F (xn) has a convergent subsequence ynk

∈ F (xnk) that converges to a value

y0 ∈ F (x0). Therefore there is an integer K such that ynk∈ F (x0) for every k > K.

In Figure 5.9 the sequence x1, x2, . . .→ x0. The sequence y1, y2, . . . need not converge,but it contains a subsequence y′1, y

′2, . . . that converges to some element y′0 ∈ F (x0).

Page 190: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.5. UPPER-HEMI-CONTINUITY BY ITSELF 171

Because this is true for every sequence xn and any sequence yn ∈ F (xn), it followsthat every possible limit point y0 ∈ F (x0). This is possible only if every image F (xnk

)is either contained in F (x0) (see the sets F (xn) just to the left of F (x0) in Figure 5.9)or are almost the same as F (x0) (see the sets F (xn) just to the right of F (x0) in thefigure) for values xnk

that are very close to x0. This is exactly what is meant by Fbeing upper-hemi-continuous at x0.

Figure 5.9: Upper-hemi-continuity at the point x = x0.

Now let’s examine part (b) of the theorem. Pick some point x0 ∈ X and asequence {xn}∞n=1 in X with xn → x0. This generates a sequence of image sets F (xn).For each n select as you please one element yn from F (xn). Since F is compact-valued, the sequence {yn}∞n=1 must contain a subsequence {ynk

}∞k=1 that converges toa value we will call y0. The question is, “Must y0 belong to F (x0)?” Suppose thaty0 ∈ F (x0). Then is F upper-hemi-continuous at x0? No. Why not? We are toldthat the subsequence of range values ynk

converges to y0. For each ynkthere is an

associated domain value xnk. Each xnk

belongs to the xn sequence that converges tox0. Every subsequence of a convergent sequence is convergent, and to the same limit– see Corollary 4.1. Hence xnk

→ x0. Now y0 ∈ F (x0) and the sequence ynk→ y0,

so there is an integer K such that ynk∈ F (x0) for every k > K (draw the picture).

Page 191: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

172 CHAPTER 5. CONTINUITY

Hence F (xnk) ⊂ F (x0) for every k > K. The picture you should have in your head at

this moment is of a sequence of images F (xnk) that are not entirely contained in F (x0)

for values xnkthat are very close to x0. Hence F cannot be upper-hemi-continuous

at x0.

Theorem 5.5. Let (X, dX) and (Y, dY ) be metric spaces. If F : X �→ Y is a compact-valued and upper-hemi-continuous correspondence, then F is closed-valued.

The statement is that, if F is both compact-valued and upper-hemi-continuous,then, for any x ∈ X, the image F (x) is a set that is closed in Y with respectto the metric topology defined on Y . Consider, for example, the constant-valuedcorrespondence F : � �→ � that is defined by

F (x) = (1, 2) ∀ x ∈ �.

For any x ∈ �, the image set (1, 2) is not closed with respect to E1. Yet we havealready seen that this correspondence is (fully) continuous, not just upper-hemi-continuous. Does this example contradict Theorem 5.5? No, because F is not acompact-valued correspondence.

More common misconceptions arise when the graph of an upper-hemi-continuouscorrespondence is considered, so let us now turn to thinking about the graph of acorrespondence. Definition 5.10 states what is meant by the graph of a correspondenceand is repeated here for your convenience.

Definition 5.10′ (Graph of a Correspondence) The graph GF of a correspondenceF : X �→ Y is

GF = {(x, y) | x ∈ X and y ∈ F (x)}.Definition 5.19 (Closed Graph). Let (X, TX) and (Y, TY ) be topological spaces. Letthe correspondence F : X �→ Y . F has a closed graph if and only if the set GF ={(x, y) | x ∈ X and y ∈ F (x)} is closed with respect to the topology on the productspace X × Y .

If we restrict ourselves to correspondences that map only from a metric space toa metric space, then the statement of Definition 5.19 is equivalent to the followingstatement.

Definition 5.20 (Closed Graph for Metric Spaces). Let (X, dX) and (Y, dY ) be metricspaces. Let the correspondence F : X �→ Y . Let x0 ∈ X. F is closed-valued at x0 ifand only if, for every sequence {xn}∞n=1 with xn ∈ X for all n ≥ 1 and xn → x0 andevery sequence {yn}∞n=1 with yn ∈ F (xn) for every n ≥ 1 and yn → y0, y0 ∈ F (x0). Fis closed-valued if and only if F (x0) is closed with respect to (X, dX) at every x0 ∈ X.

Page 192: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.5. UPPER-HEMI-CONTINUITY BY ITSELF 173

Phrased in terms of the graph of a correspondence, the statement of Definition 5.20is exactly the same as the following statement.

Definition 5.20′ (Closed Graph for Metric Spaces). Let (X, dX) and (Y, dY ) bemetric spaces. Let the correspondence F : X �→ Y . F has a closed graph GF if andonly if, for every sequence {(xn, yn)}∞n=1 with xn ∈ X for all n ≥ 1 and xn → x0 andyn ∈ F (xn) for every n ≥ 1 and yn → y0, (x0, y0) ∈ GF .

For most people, accustomed to working and thinking in a Euclidean topologicalsetting, a set is closed if and only if it contains all of its limit points, both interiorlimit points and boundary limit points. This is all that the two above definitions say.In each case there is a domain sequence of elements xn that converges to some pointx0 in the domain. For each n there is another element yn in the range. The sequence{yn}∞n=1 may or may not converge, but if it does, then it converges to a point y0 thatbelongs to the image set F (x0). So any such convergent sequence yn has its limitpoint in the set F (x0). This is what we see in figures such as 5.9 that display graphsof correspondences that are upper-hemi-continuous at a point x0.

It is often asserted that a correspondence F : X �→ Y is upper-hemi-continuousif and only if the graph of F is a set that is closed in the topology defined on theproduct space X×Y . We have already seen that this cannot be true in general. Lookagain at the correspondence defined in (5.11). The graph of F3 is the set

GF3 = {(x, y) | x ∈ � and 1 < y < 2}.

This is a set that is open with respect to the Euclidean topology on � × � = �2.Yet F3 is upper-hemi-continuous. The truth of the matter is that the graph of anupper-hemi-continuous correspondence may be a closed set, but need not be, andthat a correspondence with a closed graph may be, but need not be, upper-hemi-continuous. We have just had an example of the first of these two assertions. For anexample of the second assertion, consider the correspondence F5 : � �→ �, where

F5(x) =

{{1, 4} ; x = 2

{2, 3} ; x = 2.

The graph of this correspondence is a set that is closed with respect to E2, butthe correspondence is not upper-hemi-continuous (F5 is not lower-hemi-continuouseither). The confusion typically arises from a misreading of the following theorem.

Theorem 5.6 (Closed Graph Theorem). Let (X, dX) and (Y, dY ) be metric spaces.Let the correspondence F : X �→ Y .

Page 193: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

174 CHAPTER 5. CONTINUITY

(a) If the graph of F is closed with respect to the topology on X × Y and if therange Y is compact with respect to the topology on Y , then F is upper-hemi-continuouson X.

(b) If F is upper-hemi-continuous on X and is closed-valued on X, then the graphof F is closed with respect to the topology on X × Y .

An often-cited example of (a) that demonstrates the importance of requiring thatthe range Y be compact is as follows. Let F6 : [0, 2] �→ �+, where

F6(x) =

⎧⎪⎪⎪⎨⎪⎪⎪⎩

{0} ; x = 0

{0,

1

x

}; 0 < x ≤ 2.

The graph of F6 is a closed (but not a bounded) subset of �2+, and F6 is not upper-

hemi-continuous at x = 0. To the contrary, F6 is lower-hemi-continuous at x = 0.Notice that the range of F6 is �+, which is a set that is closed but not bounded, andthus not compact, with respect to E1.

What if we modified F6 by insisting that its range Y be some bounded subset of�+? Let’s choose Y = [0, 10]. Now we must ask “what are the images of the domainelements x ∈ (0, 1/10).” Suppose we set all these images equal to the doubleton{0, 10}. Then our modified correspondence is F7 : [0, 2] �→ [0, 10], where

F7(x) =

⎧⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎩

{0} ; x = 0

{0, 10} ; 0 < x < 1/10

{0,

1

x

}; 1/10 ≤ x ≤ 2.

The graph of F7 is a bounded but not closed (hence not compact) subset of �2+ (why?)

and F7 is not upper-hemi-continuous at x = 0.Let’s try again. Let’s modify the image of F7 at x = 0 from the singleton {0} to

the doubleton {0, 10}. This gives us the correspondence F8 : [0, 2] �→ [0, 10], where

F8(x) =

⎧⎪⎪⎪⎨⎪⎪⎪⎩{0, 10} ; 0 ≤ x < 1/10

{0,

1

x

}; 1/10 ≤ x ≤ 2.

The graph of F8 is a closed and bounded (hence compact) subset of �2+ and F8 is

upper-hemi-continuous at every x ∈ [0, 2].

Page 194: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.6. JOINT CONTINUITY OF A FUNCTION 175

5.6 Joint Continuity of a Function

What does it mean to say that a function is jointly continuous?

Definition 5.21 (Jointly Continuous Function). Let f : X×Γ �→ Y be a function. fis jointly continuous with respect to (x, γ) at (x0, γ0) ∈ X×Γ if and only if, for everyset O ⊂ Y that is open with respect to the topology TY on Y , the set f−1(O) ⊂ X ×Γis open with respect to the topology TX×Γ on X × Γ.

When we restrict ourselves to functions that map from a metric space to a metricspace, Definition 5.21 becomes equivalent to the following statement.

Definition 5.22 (Jointly Continuous Function in Metric Spaces). Let (X, dX), (Γ, dΓ)and (Y, dY ) be metric spaces. Let f : X×Γ �→ Y be a function. f is jointly continuousin (x, γ) at (x0, γ0) ∈ X × Γ if and only if, for every sequence {(xn, γn)}∞n=1 with(xn, γn) ∈ X × Γ for all n ≥ 1 and (xn, γn) → (x0, γ0), limn→∞ f(xn, γn) = f(x0, γ0).

It should be clear that, if a function f(x, γ) is jointly continuous with respect to(x, γ), then the function is continuous with respect to x for a fixed value of γ and thefunction is continuous with respect to γ for a fixed value of x. (What are the simpleproofs of these two assertions?) It is tempting to infer that the reverse is true; i.e.that, if the function f(x, γ) is continuous with respect to x for any fixed value of γand the function is also continuous with respect to γ for any fixed value of x, thenthe function must be jointly continuous with respect to (x, γ). This is not true. Theusual counterexample is the function f : �2

+ �→ �+ that is

f(x, γ) =

⎧⎨⎩

x2 + γ2, for (x, γ) = (0, 0)

0 , for (x, γ) = (0, 0).

Fix γ at any value, and it is obvious that the function is continuous with respectto x ∈ �+. Fix x at any value, and it is obvious that the function is continuouswith respect to γ ∈ �+. But is the function jointly continuous in (x, γ)? Consider asequence {(xn, γn)}∞n=1 such that (xn, γn) → (0, 0) with xn = γn for all n ≥ 1. Then,

limn→∞

f(xn, γn) =1

2= f(0, 0) = 0.

Therefore f is not jointly continuous at (x, γ) = (0, 0) and so is not jointly continuouson �2

+.

Page 195: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

176 CHAPTER 5. CONTINUITY

5.7 Using Continuity

This chapter was hard work, so there had better be a good reason for all that effort. Inevery chapter after this, continuity is used a lot. Almost every argument presented asa part of developing understanding of solving constrained optimization problems re-lies heavily upon continuity. Sometimes continuity is assumed to be there; sometimesit must be shown to be there. The next chapter considers when we can “separate”two sets from each other. Separation of sets is the idea at the core of constrained op-timization. The tool that we use to attempt such separations is a continuous functioncalled a “hyperplane.” (You may not yet know it but for you this is an old idea if youhave studied at least intermediate-level economics.) Continuity is a crucial feature ofthe methods we use to separate sets. In fact, if you don’t understand continuity, thenyou cannot truly understand constrained optimization problems and their solutions.That is why this chapter matters and why all of the effort to understand its contentsis worth it.

Page 196: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.8. PROBLEMS 177

5.8 Problems

Definition 5.23. Let X be a subset of �n and let f : X �→ � be a function. Letα ∈ �. The level-α contour set of f is the set {x ∈ X | f(x) = α}.

Definition 5.24. Let X be a subset of �n and let f : X �→ � be a function. Letα ∈ �. The level-α upper contour set of f is the set {x ∈ X | f(x) ≥ α}.

Definition 5.25. Let X be a subset of �n and let f : X �→ � be a function. Letα ∈ �. The level-α lower contour set of f is the set {x ∈ X | f(x) ≤ α}.

Definition 5.26. Let X be a subset of �n and let f : X �→ � be a function. f isupper-semi-continuous on X if and only if all of the upper-contour sets of f are closedin En.

Definition 5.27. Let X be a subset of �n and let f : X �→ � be a function. f islower-semi-continuous on X if and only if all of the lower-contour sets of f are closedin En.

Problem 5.1. Give an example of a function that is upper-semi-continuous, but notcontinuous.

Problem 5.2. Give an example of a function that is lower-semi-continuous, but notcontinuous.

Problem 5.3. Let X be a subset of �n and let f : X �→ � be a function. Prove thatf is upper-semi-continuous on X if and only if −f is lower-semi-continuous on X.

Problem 5.4. Let X be a subset of �n and let f : X �→ � be a function. Provethat f is continuous on X if and only if f is both lower-semi-continuous on X andupper-semi-continuous on X.

Problem 5.5. Let X be a subset of �n and let f : X �→ � be a function. Prove thatf is continuous on X if and only if every upper-contour set and every lower-contourset of f is closed in En.

Problem 5.6. Figure 5.10 is Figure 5.3 to which has been added a horizontal linedrawn at the height that is the right-side limit of f(x) as x → x0 from above; i.e.limx→x+

0f(x). Claim: “The set {x | f(x) < limx→x+

0f(x)} = (−∞, x′) ∪ {x0}. This

set is not open with respect to E1, so f is not upper-semi-continuous at x = x0.” Isthis claim true or false? Explain.

Page 197: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

178 CHAPTER 5. CONTINUITY

Figure 5.10: The graph for Problem 5.6.

Problem 5.7. (Weierstrass’s Theorem for Upper-Semi-Continuous Functions) LetX be a nonempty and compact subset of �n and let f : X �→ � be an upper-semi-continuous function. Prove that f attains a maximum on X.

Problem 5.8. Let X be a nonempty and compact subset of �n and let f : X �→ �be a lower-semi-continuous function. Given an example that shows that f need notattain a maximum on X.

Problem 5.9. (Weierstrass’s Theorem for Lower-Semi-Continuous Functions) LetX be a nonempty and compact subset of �n and let f : X �→ � be a lower-semi-continuous function. Prove that f attains a minimum on X.

Problem 5.10. Let X be a nonempty and compact subset of �n and let f : X �→ �be an upper-semi-continuous function. Given an example that shows that f need notattain a minimum on X.

Problem 5.11. (Weierstrass’s Theorem for Continuous Functions) Let X be a non-empty and compact subset of �n and let f : X �→ � be a continuous function. Provethat f attains both a maximum and a minimum on X.

Page 198: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.8. PROBLEMS 179

Problem 5.12. The sets X and Y are

X = {hammer, saw, nail, drill, pencil, screw} and Y = {bird, bee, snail}.

(i) Confirm that the collection of sets

TX = {∅, {saw}, {nail}, {hammer}, {nail, drill}, {nail, saw}, {nail, hammer},{saw, hammer}, {saw, nail, drill}, {saw, nail, hammer},{nail, drill, hammer}, {nail, drill, saw, hammer}, X}

is a topology on X and that the collection of sets

TY = {∅, {bird}, {bee}, {snail}, {bird, bee}, {bird, snail}, {bee, snail}, Y }

is a topology on Y .(ii) Define the mapping f : X �→ Y by

f(saw) = bird

f(drill) = bee

f(nail) = snail

f(hammer) = bird.

What is the domain of f? What is the range of f? What is the image of f? Is f afunction? Is f surjective? Is f injective? Is f bijective? What is the graph of f? Inwhat space is the graph of f contained?(iii) Is f continuous relative to the topologies TX and TY ?

Problem 5.13. Consider the correspondence F : �+ �→ � that is

F (x) = ±√x; i.e. F (x) = {−

√x, +

√x } for x ≥ 0.

(i) What is the image of this correspondence? Is its image the same as its codomain?Is the correspondence surjective? Is it injective? Is it bijective? What is the graphof the correspondence? Is the graph a closed set and, if so, then closed with respectto which topology?(ii) Is the correspondence F convex-valued? Is the graph of F a convex set?(iii) Is the correspondence F closed-valued on �+?(iv) Is the correspondence F compact-valued on �+? Is the range of F compact? Isthe graph of F a compact set?

Page 199: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

180 CHAPTER 5. CONTINUITY

(v) Prove that F is upper-hemi-continuous on �+ with respect to E2.

(vi) Prove that F is lower-hemi-continuous on �+ with respect to E2.

(vii) Prove that F is continuous on �+ with respect to E2.

(viii) Can the Closed Graph Theorem be used to establish that F is upper-hemi-continuous on �+ with respect to E2? Why or why not?

Problem 5.14. Let X ⊂ �n and Y ⊂ �m. Let F : X �→ Y and G : X �→ Y becorrespondences. Define the correspondences Θ : X �→ Y and Ψ : X �→ Y by

Θ(x) = F (x) ∪G(x) and Ψ(x) = F (x) ∩G(x) ∀ x ∈ X.

(i) If F and G are convex-valued on X, then is Θ convex-valued on X?

(ii) If F and G are convex-valued on X, then is Ψ convex-valued on X?

(iii) If F and G are compact-valued on X, then is Θ compact-valued on X?

(iv) If F and G are compact-valued on X, then is Ψ compact-valued on X?

(v) If F and G are upper-hemi-continuous on X, then is Θ upper-hemi-continuouson X?

(vi) If F and G are lower-hemi-continuous on X, then is Θ lower-hemi-continuouson X?

(vii) If F and G are upper-hemi-continuous on X, then is Ψ upper-hemi-continuouson X?

(viii) If F and G are lower-hemi-continuous on X, then is Ψ lower-hemi-continuouson X?

(ix) If F and G are continuous on X, then is Θ continuous on X?

(x) If F and G are continuous on X, then is Ψ continuous on X?

(xi) If the graphs of F and G are closed with respect to En+m, then is the graph ofΘ closed and, if so, then is Θ upper-hemi-continuous on X? If your answer is that Θneed not be upper-hemi-continuous, then state any additional condition(s) that willensure that Θ is upper-hemi-continuous.

(xii) If the graphs of F and G are closed with respect to En+m, then is the graph ofΨ closed and, if so, then is Ψ upper-hemi-continuous on X? If your answer is that Ψneed not be upper-hemi-continuous, then state any additional condition(s) that willensure that Ψ is upper-hemi-continuous.

Page 200: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.9. ANSWERS 181

5.9 Answers

Answer to Problem 5.1. Three examples are

g(x) =

{x , for 0 ≤ x < 2

x+ 1 , for 2 ≤ x ≤ 4,

h(x) =

{x+ 2 , for 0 ≤ x ≤ 2

3− x , for 2 < x ≤ 4,

and i(x) =

⎧⎪⎨⎪⎩x , for 0 ≤ x < 2

5 , for x = 2

x , for 2 < x ≤ 4.

Answer to Problem 5.2. Three examples are

u(x) =

{x , for 0 ≤ x ≤ 2

x+ 1 , for 2 < x ≤ 4,

v(x) =

{x+ 2 , for 0 ≤ x < 2

3− x , for 2 ≤ x ≤ 4,

and w(x) =

⎧⎪⎨⎪⎩x , for 0 ≤ x < 2

1 , for x = 2

x , for 2 < x ≤ 4.

Answer to Problem 5.3.

Proof. The statement that, for every α ∈ �, the set {x ∈ X | f(x) ≥ α} is closed inXis equivalent to the statement that, for every −α ∈ �, the set {x ∈ X | −f(x) ≤ −α}is closed in X.

Answer to Problem 5.4.

Proof. Let’s start by taking as given that f is both upper and lower-semi-continuouson X. Take any x0 ∈ X. Then take any open ball in the codomain of f thatcontains f(x0). That is, take any δ > 0 and consider the interval O = (f(x0) −δ, f(x0) + δ) in the codomain of f . We need to show that the inverse image of O is aset that is open in the domain of f at x0.

f−1(O) = f−1(f(x0)− δ, f(x0) + δ)

= {x ∈ X | f(x) < f(x0) + δ} ∩ {x ∈ X | f(x0)− δ < f(x)}.

Page 201: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

182 CHAPTER 5. CONTINUITY

The set {x ∈ X | f(x) < f(x0) + δ} is open in the domain of f because f is upper-semi-continuous on X. The set {x ∈ X | f(x0)− δ < f(x)} is open in the domain off because f is lower-semi-continuous on X. The intersection of two open sets is anopen set, so f−1(O) is open at x0. x0 is an arbitrary choice from X, so f is continuouson X.

Now let’s take as given that f is continuous on X. Take any x0 ∈ X and anyδ > 0 and consider the open ball in the range of f that is the interval O = (f(x0)−δ, f(x0)+δ). The inverse image of O is open in the domain of f , since f is continuousat x0. That is, the set

f−1(O) = {x ∈ X | f(x0)− δ < f(x) < f(x0) + δ}= {x ∈ X | f(x) < f(x0) + δ} ∩ {x ∈ X | f(x) > f(x0)− δ}

is open in the domain of f , requiring that the sets {x ∈ X | f(x) < f(x0) + δ} and{x ∈ X | f(x) > f(x0) − δ} are open in the domain of f . This establishes that f isboth upper and lower-semi-continuous at x0. x0 is an arbitrary choice from X, so fis both upper and lower-semi-continuous on X.

Answer to Problem 5.5.

Proof. f is continuous on X if and only f is both upper and lower semi-continuouson X. But f is upper semi-continuous on X if and only if all of the upper contoursets of f are closed in En, and f is lower-semi-continuous on X if and only if all ofthe lower contour sets of f are closed in En.

Answer to Problem 5.6: The claim is false. The set

{x | f(x) < limx→x+

0

f(x)} = (−∞, x′).

This set is open with respect to E1.

Answer to Problem 5.7.

Proof. Let f = supx∈X f(x). It must be shown that there is a value x ∈ X for whichthe corresponding value of f is f .

Since f = supx∈X f(x), there exists in X a sequence {xn}∞n=1 such that

f(xn) → f. (5.17)

Since X is compact, this sequence must contain a convergent subsequence {xnk}∞k=1.

Let x = limk→∞ xnk. x ∈ X because X is a closed set. Since f is upper-semi-

continuous on X,limk→∞

f(xnk) ≤ f(x). (5.18)

Page 202: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.9. ANSWERS 183

Comparing (5.17) and (5.18) shows that f(x) = f .

Answer to Problem 5.8. Let X = [0, 4] and let the function be

v(x) =

{x , for 0 ≤ x ≤ 2

5− x , for 2 < x ≤ 4.

This function is lower-semi-continuous on [0, 4] and does not attain a maximum on[0, 4].

Answer to Problem 5.9.

Proof. Let f = infx∈X f(x). It must be shown that there is a value x ∈ X for whichthe corresponding value of f is f .

Since f = infx∈X f(x), there exists in X a sequence {xn}∞n=1 such that

f(xn) → f. (5.19)

Since X is compact, this sequence must contain a convergent subsequence {xnk}∞k=1.

Let x = limk→∞ xnk. x ∈ X because X is a closed set. Since f is lower-semi-

continuous on X,limk→∞

f(xnk) ≥ f(x). (5.20)

Comparing (5.19) and (5.20) shows that f(x) = f .

Answer to Problem 5.10. Let X = [0, 4] and let the function be

t(x) =

{5− x , for 0 ≤ x ≤ 2

x , for 2 < x ≤ 4.

This function is upper-semi-continuous on [0, 4] and does not attain a minimum on[0, 4].

Answer to Problem 5.11. A function is continuous if and only if it is both upperand lower-semi-continuous so the result follows from the answers to problems 5.7and 5.9.

Answer to Problem 5.12.(i) Any union of sets in TX is a set in TX and any finite intersection of sets in TX isa set in TX . Hence TX is a topology on X. Any union of sets in TY is a set in TY andany finite intersection of sets in TY is a set in TY . Hence TY is a topology on Y . Infact, TY is the discrete topology on Y since it is the collection of all possible subsetsformed from the elements of Y .

Page 203: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

184 CHAPTER 5. CONTINUITY

(ii) The domain of f is

{x ∈ X | f(x) = ∅} = {saw, drill, nail, hammer}.

The range/image of f is f(X) = {bird, bee, snail}.f is a function because f(x) is a singleton for every x in the domain of f . The

codomain and the image of f are the same set, so f is surjective. f is not injectivesince the element “bird” in the image of f is mapped to from more than one domainelement; f(saw) = f(hammer) = bird. Thus f is not bijective.

The graph of f is the set

Gf = {(x, y) | x ∈ X and y = f(x)}= {(saw, bird), (drill, bee), (nail, snail), (hammer, bird)}.

This set is contained in the direct product of the domain and codomain,

X × Y = {(hammer, bird), (hammer, bee), (hammer, snail),

(saw, bird), (saw, bee), (saw, snail),

(nail, bird), (nail, bee), (nail, snail),

(drill, bird), (drill, bee), (drill, snail),

(pencil, bird), (pencil, bee), (pencil, snail),

(screw, bird), (screw, bee), (screw, snail)}.

(iii) No. The sets {bee} and {bird, bee} are open with respect to TY . The inverseimages of these sets are

f−1(bee) = {drill} and

f−1({bird, bee}) = {saw, drill, hammer}.

But neither of the sets {drill} and {saw, drill, hammer} is open with respect to TX .Therefore f is not continuous relative to TX and TY .

Answer to Problem 5.13.

(i) The range/image is the real line, �. This is also the codomain of the corre-spondence. Hence the correspondence is surjective. The correspondence is injectivebecause any image set {−√

x,+√x } is mapped to from exactly one x in the domain

of the correspondence; i.e. for any x′, x′′ ∈ � with x′ = x′′, F (x′) ∩ F (x′′) = ∅. The

Page 204: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.9. ANSWERS 185

correspondence is both surjective and injective, so it is bijective. The graph of thecorrespondence is the set

GF = {(x, y) | x ∈ �+, y ∈ {−√x,+

√x }}.

GF is a subset of �+ × � and is closed with respect to the Euclidean topology on�+ ×�.(ii) The correspondence is not convex-valued. For any x = 0, the image of f consistsof the two distinct points −√

x and +√x in �; this is not a convex set. Similarly,

the graph is not a convex subset of �+ �.(iii) F (0) = {0} is a singleton and so is closed in E1. For any x = 0, F (x) is a setconsisting of two distinct points in �. Any such set is closed in E1. Hence F (x) isclosed in E1 for every element in the domain of F , meaning that the correspondenceis closed-valued with respect to E1.

(iv) F (0) = {0} is obviously both closed and bounded, and hence compact, withrespect to E1. For any x = 0, F (x) is closed with respect to E1 and the numberM(x) = +2

√x is a finite upper bound on the Euclidean distance between the points

in the set. Hence F (x) is bounded with respect to E1 and is therefore compact withrespect to E1. Since F (x) is compact with respect to E1 for every x in the domainof F , F is compact-valued with respect to E1. The range of F is �, which is notbounded, so the range is not compact with respect to E1. The graph of F , GF , isa subset of �+ × � that is not bounded, so it is not compact with respect to theEuclidean topology on �+ ×�.(v) Proof 1. F is upper-hemi-continuous on �+ if and only if the upper-inverseof any set O that is open in � is a set V that is open in �+, such that x ∈ Vimplies that F (x) ⊂ O. So take a nonempty subset O in the range � of F thatis open with respect to E1. Take any y ∈ O. Then there is a ball with a positiveradius ε(y) that is centered on y and is entirely contained in O. This ball is the setBdE(y, ε(y)) = {y′ ∈ � | y − ε(y) < y′ < y + ε(y)}.

y − ε(y) <√x < y + ε(y) ⇔ (y − ε(y))2 < x < (y + ε(y))2,

so the upper-inverse of F of BdE(y, ε(y)) is

F−1(BdE(y, ε(y)))+ = {x | x ∈ �+, F (x) ⊂ BdE(y, ε(y))}={x | x ∈ �+, (y − ε(y))2 < x < (y + ε(y))2

}.

Page 205: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

186 CHAPTER 5. CONTINUITY

This is an interval in �+ that is open with respect to the Euclidean topology on �+.The upper-inverse of the entire set O is the union of the upper-inverses of all of theballs comprising O:

F−1(O)+ =⋃y∈O

F−1(BdE(y, ε(y)))+.

Since it is the union of sets that are open with respect to the Euclidean topologyon �+, F

−1(O)+ is open with respect to this topology. Therefore F is upper-hemi-continuous on �+.

Proof 2. Since F is a compact-valued correspondence that maps from the Eu-clidean metric space (�+, dE) to the Euclidean metric space (�, dE), F is upper-hemi-continuous at x0 ∈ �+ if and only if, for every sequence {xn}∞n=1 with xn ∈ �+ for alln ≥ 1 and xn → x0, every sequence {yn}∞n=1 with yn ∈ F (xn) for all n ≥ 1 contains aconvergent subsequence {ynk

}∞k=1 with ynk→ y0 ∈ F (x0).

Take any x0 ∈ �+ and any sequence {xn}∞n=1 with xn ∈ �+ for all n ≥ 1 andxn → x0. For each n ≥ 1, F (xn) = {−√

xn,+√xn }. Select the sequence yn = +

√xn

for all n ≥ 1. As xn → x0, yn → y0 =√x0 ∈ F (x0). Similarly, select the sequence

yn = −√xn for all n ≥ 1. As xn → x0, yn → y0 = −√

x0 ∈ F (x0). Thus Fis upper-hemi-continuous at x0. But x0 is an arbitrary choice from �+, so F isupper-hemi-continuous on �+ with respect to the Euclidean metric topologies on �+

and �.(vi) Proof 1. F is lower-hemi-continuous at x0 ∈ �+ if and only if the lower-inverseof any set O that is open with respect to E1 with F (x0) ∩ O = ∅ is a set V that isopen with respect to the Euclidean topology on �+, such that x ∈ V implies thatF (x) ∩ O = ∅. So take a nonempty subset O in the range � of F that is open withrespect to E1, such that {−√

x0,+√x0 } ∩ O = ∅. Then either y−0 = −√

x0 ∈ O, ory+0 = +

√x0 ∈ O, or both. Suppose y+0 ∈ O. Then there is a ball with a positive

radius ε(y+0 ) that is centered on y+0 and is entirely contained in O. This ball is theset BdE(y

+0 , ε(y

+0 )) = {y′ ∈ � | y+0 − ε(y+0 ) < y′ < y+0 + ε(y+0 )}.

y+0 − ε(y+0 ) < +√x < y+0 + ε(y+0 ) ⇔ (y+0 − ε(y+0 ))

2 < x < (y+0 + ε(y+0 ))2,

so the lower-inverse of BdE(y+0 , ε(y

+0 )) is

F−1(BdE(y+0 , ε(y

+0 )))

={x | x ∈ �+, F (x) ∩BdE

(y+0 , ε

(y+0))

= ∅}

={x | x ∈

((y+0 − ε

(y+0))2

,(y+0 + ε

(y+0))2)

, {−√x,+

√x } ∩O = ∅

}.

Page 206: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.9. ANSWERS 187

The lower-inverse is not empty since it contains x0. Moreover the lower-inverse is aninterval in �+ that is an open set. The lower-inverse of the entire set O is the unionof the lower-inverses of all of the balls comprising O:

F−1(O)− =⋃y∈O

F−1 (BdE (y, ε(y)))−.

Since it is the union of sets that are open, it too is an open set. Therefore F islower-hemi-continuous on �+.

Proof 2. Since F is a compact-valued correspondence that maps from the Eu-clidean metric space (�+, dE) to the Euclidean metric space (�, dE), F is lower-hemi-continuous at x0 ∈ �+ if and only if, for every sequence {xn}∞n=1 with xn ∈ �+ forall n ≥ 1 and xn → x0 and for every y0 ∈ F (x0), there is a sequence {yn}∞n=1 withyn ∈ F (xn) for every n ≥ 1 and yn → y0.

Take any x0 ∈ �+ and any sequence {xn}∞n=1 with xn ∈ �+ for all n ≥ 1 andxn → x0. F (x0) = {−√

x0,+√x0 }. For each n ≥ 1, F (xn) = {−√

xn,+√xn }.

Select y+0 = +√x0. The sequence yn = +

√xn for all n ≥ 1 converges to y+0 as

n → ∞. Select y−0 = −√x0. The sequence yn = −√

xn converges to y−0 as n → ∞.Thus F is lower-hemi-continuous at x0. x0 is an arbitrary choice from �+ so F islower-hemi-continuous on �+.

(vii) Proof. F is continuous on �+ if and only if F is both upper and lower-hemi-continuous on �+. Therefore, from parts (v) and (vi), F is continuous on �+.

(viii) No. The Closed Graph Theorem requires the range of F to be a compact set.The range of F is �, which is not compact with respect to E1.

Answer to Problem 5.14.(i) F,G convex-valued on X ⇒ Θ is convex-valued on X.Proof. For any x ∈ X, the sets F (x) and G(x) are convex subsets of �m. The unionof convex sets need not be a convex set, so Θ(x) need not be a convex subset of �m.Thus Θ need not be a convex-valued correspondence.

(ii) F,G convex-valued on X ⇒ Ψ is convex-valued on X.Proof. For any x ∈ X, the sets F (x) and G(x) are convex subsets of �m. Theintersection of convex sets is a convex set, so Ψ(x) is a convex subset of �m. Thus Ψis a convex-valued correspondence.

(iii) F,G compact-valued on X ⇒ Θ is compact-valued on X.Proof. For any x ∈ X, the sets F (x) and G(x) are compact, with respect to Em,subsets of �m. The union of two compact sets is a compact set, so Θ(x) is a compactsubset of �m. Thus Θ is a compact-valued correspondence.

Page 207: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

188 CHAPTER 5. CONTINUITY

(iv) F,G compact-valued on X ⇒ Ψ is compact-valued on X.Proof. For any x ∈ X, the sets F (x) and G(x) are compact, with respect to Em,subsets of �m. The intersection of two compact sets is a compact set, so Ψ(x) is acompact subset of �m. Thus Ψ is a compact-valued correspondence.

(v) F,G upper-hemi-continuous on X ⇒ Θ is upper-hemi-continuous on X.Proof. Take any set O that is open with respect to Em in the codomain of Θ withΘ(x) ⊂ O for some x ∈ X. The upper-inverse of O is

Θ−1(O)+ = {x ∈ X | Θ(x) ⊂ O}= {x ∈ X | (F (x) ∪G(x)) ⊂ O}= {x ∈ X | F (x) ⊂ O} ∪ {x ∈ X | G(x) ⊂ O}= F−1(O)+ ∪G−1(O)+.

Since F is upper-hemi-continuous the upper-inverse image of F of O, F−1(O)+, is aset that is open in X. Similarly, G−1(O)+ is open in X. Therefore Θ−1(O)+ is openin X, establishing that Θ is upper-hemi-continuous on X.

(vi) F,G lower-hemi-continuous on X ⇒ Θ is lower-hemi-continuous on X.Proof. Take any set O that is open with respect to Em in the codomain of Θ withΘ(x) ∩O = ∅ for some x ∈ X. The lower-inverse of O is

Θ−1(O)− = {x ∈ X | Θ(x) ∩O = ∅}= {x ∈ X | (F (x) ∪G(x)) ∩O = ∅}= {x ∈ X | F (x) ∩O = ∅} ∪ {x ∈ X | G(x) ∩O = ∅}= F−1(O)− ∪G−1(O)−.

Since F is lower-hemi-continuous, the lower-inverse image of F of O, F−1(O)−, is aset that is open in X. Similarly, G−1(O)− is open in X. Therefore Θ−1(O)− is openin X, establishing that Θ is lower-hemi-continuous on X.

(vii) F,G upper-hemi-continuous on X ⇒ Ψ is upper-hemi-continuous on X.Proof. Take any set O that is open with respect to Em in the codomain of Ψ. Theupper-inverse of O is

Ψ−1(O)+ = {x ∈ X | Ψ(x) ⊂ O}= {x ∈ X | (F (x) ∩G(x)) ⊂ O}= {x ∈ X | F (x) ⊂ O} ∩ {x ∈ X | G(x) ⊂ O}= F−1(O)+ ∩G−1(O)+.

Page 208: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

5.9. ANSWERS 189

Since F is upper-hemi-continuous, the upper-inverse image of O, F−1(O)+, is a setthat is open in X. Similarly, G−1(O)+ is open in X. Therefore Ψ−1(O)+ is open inX, establishing that Ψ is upper-hemi-continuous on X.

(viii) F,G lower-hemi-continuous on X ⇒ Ψ is lower-hemi-continuous on X.

Proof. Take any set O that is open with respect to Em in the codomain of Ψ withΨ(x) ∩O = ∅ for some x ∈ X. The lower-inverse of O is

Ψ−1(O)− = {x ∈ X | Ψ(x) ∩O = ∅}= {x ∈ X | (F (x) ∩G(x)) ∩O = ∅}= {x ∈ X | F (x) ∩O = ∅} ∩ {x ∈ X | G(x) ∩O = ∅}= F−1(O)− ∩G−1(O)−.

Since F is lower-hemi-continuous, the lower-inverse image of F of O, F−1(O)−, is aset that is open in X. Similarly, G−1(O)− is open in X. Therefore Ψ−1(O)− is openin X, establishing that Ψ is lower-hemi-continuous on X.

(ix) F,G continuous on X ⇒ Θ is continuous on X.

Proof. Both F and G are continuous, and thus both upper-hemi-continuous andlower-hemi-continuous, on X. Therefore, from (v) and (vi), Θ is both upper-hemi-continuous and lower-hemi-continuous, and thus is continuous, on X.

(x) F,G continuous on X ⇒ Ψ is continuous on X.

Proof. Both F and G are continuous, and thus both upper-hemi-continuous andlower-hemi-continuous, on X. Therefore, from (vii) and (viii), Ψ is both upper-hemi-continuous and lower-hemi-continuous, and thus is continuous, on X.

(xi) The graph of Θ is the set

GΘ = {(x, y) | x ∈ X, y ∈ Θ(x)}= {(x, y) | x ∈ X, y ∈ F (x) ∪G(x)}= {(x, y) | x ∈ X, y ∈ F (x)} ∪ {(x, y) | x ∈ X, y ∈ G(x)}= GF ∪GG.

GΘ is the union of two closed sets and is therefore closed. By itself this does not ensurethat Θ is upper-hemi-continuous on X. If, however, the range of Θ is a compact set,then Θ is upper-hemi-continuous on X if and only if the graph of Θ is a closed subsetof X × Y .

Page 209: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

190 CHAPTER 5. CONTINUITY

(xii) The graph of Ψ is the set

GΨ = {(x, y) | x ∈ X, y ∈ Ψ(x)}= {(x, y) | x ∈ X, y ∈ F (x) ∩G(x)}= {(x, y) | x ∈ X, y ∈ F (x)} ∩ {(x, y) | x ∈ X, y ∈ G(x)}= GF ∩GG.

GΨ is the intersection of two closed sets and is therefore closed. By itself this doesnot ensure that Ψ is upper-hemi-continuous on X. If, however, the range of Ψ is acompact set, then Ψ is upper-hemi-continuous on X if and only if the graph of Ψ isa closed subset of X × Y .

Page 210: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

Chapter 6

Hyperplanes and Separating Sets

The whole idea of constrained optimization is to separate sets. If this seems odd toyou, then consider the following familiar (to an economist) examples.

The main idea in demand theory is that a consumer chooses a consumption bundlethat she most prefers from the bundles she can afford to purchase. The typicalpicture is displayed in Figure 6.1. The straight line that is the budget constraintseparates two sets, S1 and S2. In this context S1 = {x ∈ �2

+ | x � x∗} is the setof consumption bundles that the consumer prefers at least as much as the bundle x∗

and S2 = {x ∈ �2+ | p1x1 + p2x2 ≤ y} is the consumer’s budget set. The set

H = {x ∈ �2+ | p1x1 + p2x2 = y} that is the budget constraint is the hyperplane in

�2+ that separates the sets S1 and S2.

One of the ideas central to the theory of the firm is that any firm seeks to producea given quantity, y units, of its product at the smallest possible total productioncost. Again the typical picture is displayed in Figure 6.1. This time S1 = {x ∈ �2

+ |f(x) ≥ y} is the set of input bundles that provide at least y units of output (f isthe firm’s production function) and S2 = {x ∈ �2

+ | w1x1 + w2x2 ≤ w1x∗1 + w2x

∗2}

is the set of input bundles that cost no more than does the input bundle x∗ thatprovides y output units. w1 and w2 are the per-unit prices of the inputs. The setH = {x ∈ �2

+ | w1x1+w2x2 = w1x∗1+w2x

∗2} is the isocost line that is the hyperplane

in �2+ that separates the sets S1 and S2.

You are probably more familiar with these problems being posed to you as con-strained optimization problems; e.g. a consumer wishing to maximize her utilitysubject to a budget constraint, or a firm wishing to minimize its cost of producing aspecified quantity of its product. This should suggest to you the intimate relationshipexisting between constrained optimization problems and the separation of sets. It is

191

Page 211: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

192 CHAPTER 6. HYPERPLANES AND SEPARATING SETS

because of this relationship that in this chapter we study how and when a hyper-plane may be used to separate sets. Let’s start by seeing what is meant by the termhyperplane.

Figure 6.1: A hyperplane separating two sets.

6.1 Hyperplanes

Definition 6.1. Given p ∈ �n with p = 0 and ‖p‖E <∞ and given α ∈ �, the set

H(p, α) = {x ∈ �n | p·x = α} (6.1)

is the hyperplane in �n with normal p and level α.

This should be familiar to you. In �2 a hyperplane is a (one-dimensional) straightline with the equation p1x1 + p2x2 = α. In �3 a hyperplane is a (two-dimensional)plane with the equation p1x1 + p2x2 + p3x3 = α. In �n a hyperplane is typically aplane of dimension n− 1.

The normal p of a hyperplane is the gradient vector with respect to x of thefunction p·x; i.e.

∇x p·x =

(∂(p1x1 + · · ·+ pnxn)

∂x1, . . . ,

∂(p1x1 + · · ·+ pnxn)

∂xn

)= (p1, . . . , pn) = pT.

Page 212: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

6.2. SEPARATIONS OF SETS 193

If we take any two points x′ and x′ + Δx in the hyperplane then p·x′ = α andp·(x′+Δx) = α, so p·Δx = 0. Thus p is orthogonal (i.e. “normal”) to any vector Δxthat is a “perturbation” from any one point in the hyperplane to any other point inthe hyperplane.

Like any function mapping to the real line, the function p·x has contour sets,upper-contour sets and lower-contour sets. For a given α, the level-α contour set ofp·x is the hyperplane itself. The upper and lower-contour sets of the hyperplane arecalled half-spaces because a hyperplane splits a space into two pieces. As you will seein the next section, the proper language is that a hyperplane separates a space intotwo half-spaces.

Definition 6.2 (Half-spaces). Given p ∈ �n with p = 0 and ‖p‖E < ∞ and givenα ∈ �, the set

UH (p, α) = {x ∈ �n | p·x ≥ α}is the upper half-space of the hyperplane H(p, α) = {x ∈ �n | p·x = α} and the set

LH (p, α) = {x ∈ �n | p·x ≤ α}

is the lower half-space of H(p, α).

A half-space in �n is a closed and weakly convex subset. By the way, in thischapter we consider only subsets of the real vector spaces �n, so terms such as open,closed, bounded, and compact are all made with respect to the Euclidean metrictopology on �n.

6.2 Separations of Sets

What do we mean by “separating two sets by a hyperplane”? All it means is thatone set lies entirely within one of the hyperplane’s half-spaces and the other set liesentirely within the other half-space.

Definition 6.3 (Separation of Sets). Let S1 and S2 be nonempty subsets of �n. LetH(p, α) be a hyperplane in �n. Then H(p, α) separates S1 and S2 if

p·x ≥ α ≥ p·y ∀ x ∈ S1, ∀ y ∈ S2. (6.2)

Figure 6.2 displays some examples of separated sets. Notice that when two closedsets are separated they may still have one or more boundary points in common (see

Page 213: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

194 CHAPTER 6. HYPERPLANES AND SEPARATING SETS

Figure 6.2: Separated sets.

panel A of the figure). If, however, one or both of two separated sets is an open set,then the two sets must be disjoint (see panel C of the figure).

Also notice from panel D of the figure that separation allows the possibility thatthe separated sets may both be contained in the separating hyperplane; i.e. S1∪S2 ⊆H is allowed by Definition 6.3. Since some may regard this as not real separation,we may insist that it not be allowed. Applying this restriction gives us the idea ofproperly separating the sets S1 and S2.

Definition 6.4 (Proper Separation of Sets). Let S1 and S2 be nonempty subsetsof �n. Let H(p, α) be a hyperplane in �n. Then H(p, α) properly separates S1 andS2 if(a) H(p, α) separates S1 and S2, and(b) S1 ∪ S2 is not a subset of H(p, α).

S1 and S2 are properly separated in panels A, B and C of Figure 6.2 but are onlyseparated in panel D of the figure.

Then again you may wish to argue that two sets X and Y are not separatedunless the value of p·x for any x ∈ X is strictly larger than is the value of p·y forany y ∈ Y . For example, in panel A of Figure 6.2, the sets S1 and S2 have boundary

Page 214: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

6.2. SEPARATIONS OF SETS 195

points in common, so you might consider these sets to be not really separated. Thisis a reasonable view to take, and it is called strict separation.

Definition 6.5 (Strict Separation of Sets). Let S1 and S2 be nonempty subsets of �n.Let H(p, α) be a hyperplane in �n. Then H(p, α) strictly separates S1 and S2 if

p·x > α > p·y ∀ x ∈ S1, ∀ y ∈ S2. (6.3)

Figure 6.3 displays some examples of strict separations of sets. Notice that thesets S1 and S2 in panels B and D of Figures 6.2 and 6.3 are the same, but thehyperplanes are different. Thus the same pair of sets may be separated or strictlyseparated depending upon the position of the separating hyperplane. Notice also thatthe panels C of Figures 6.2 and 6.3 are the same. In the two panels C, the sets S1 andS2 are strictly separated because the sets are open. Therefore the sets are disjointand neither set has a point in common with the separating hyperplane: S1 ∩ S2 = ∅,S1 ∩H = ∅ and S2 ∩H = ∅.

Figure 6.3: Strictly separated sets.

Even so, you could say that in your opinion the sets in panel C are “not reallyseparated” because there are points in the two sets that are arbitrarily close to eachother. You might argue that for two sets to really be separated there should be

Page 215: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

196 CHAPTER 6. HYPERPLANES AND SEPARATING SETS

a strictly positive distance between their closest points. This is the idea of strongseparation.

Definition 6.6 (Strong Separation of Sets). Let S1 and S2 be nonempty subsets of �n.Let H(p, α) be a hyperplane in �n. Then H(p, α) strongly separates S1 and S2 if, forsome ε > 0,

p·x ≥ α + ε > α− ε ≥ p·y ∀ x ∈ S1, ∀ y ∈ S2. (6.4)

The sets in panels A, B, and D of Figure 6.3 are strongly separated. The sets inpanel C of the figure are only strictly separated.

In the following definition, the closure of a set S is denoted by S. S is the setS together with all of the boundary points of S. If S is a closed set, then S = S.Otherwise, S ⊂ S.

Definition 6.7 (Supporting Hyperplanes and Points of Support). Let S ⊆ �n be anonempty set. Let H(p, α) be a hyperplane in �n. If x′ ∈ S is such that p·x′ = αand either p·x ≥ α for every x ∈ S or p·x ≤ α for every x ∈ S, then x′ is a point ofsupport in H for S and H is a hyperplane supporting to S.

Put into words, this definition says that a point x′ in a hyperplane is a point ofsupport for a set S if, first, S is contained entirely within one of the hyperplane’shalf-spaces (which half-space, upper or lower, does not matter) and, second, x′ is aboundary point for S.

A point of support always belongs to the supporting hyperplane and may or maynot also belong to the set being supported. In panel A of Figure 6.2, all of the pointsin the interval indicated by the thick line are points of support in the hyperplane forboth of the closed sets S1 = S1 and S2 = S2. The points in the lower boundary of therectangle S1 that are not also members of the pentahedron S2 are points of supportin the hyperplane for S1 only. The hyperplane supports both S1 and S2. In panel Bthere are two points of support in H, one for each of the closed sets S1 = S1 andS2 = S2, and the hyperplane supports both sets. In panel C the sets S1 and S2 areboth open. There is a common point of support in the hyperplane for both sets. Thispoint of support is not a member of either set, but it does belong to the closures S1

and S2 (it is a boundary point for both sets) of the sets. In panel D of Figure 6.3, thehyperplane strictly separates the sets S1 and S2, but there are no points of supportbecause neither S1 nor S2 has a boundary point belonging to the hyperplane. Thisis an example of the fact that a hyperplane that strictly separates two sets may notsupport either set. Panel D is also an example of the fact that a hyperplane thatstrongly separates two sets contains no points of support.

Page 216: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

6.3. SEPARATING A POINT FROM A SET 197

6.3 Separating a Point from a Set

The place to begin our discussion is with the relatively simple problem of separating aset from a point that does not belong to the set. All of the more sophisticated resultson separating sets with hyperplanes rely upon the answer to this initial problem.

Theorem 6.1 (Separating a Singleton Set from a Closed Set). Let S ⊂ �n be anonempty set that is closed with respect to En. Let x0 ∈ �n with x0 /∈ S. Then

(i) there exists x′ ∈ S, such that 0 < dE(x0, x′) ≤ dE(x0, x) ∀ x ∈ S.

If, additionally, S is a convex set, then

(ii) x′ is unique and there exists a hyperplane with normal p ∈ �n, with p = 0 and‖p‖E <∞, such that

p·x > p·x0 ∀ x ∈ S.

For now let’s concern ourselves with only part (i) of the theorem. The point x0 isnot a member of the set S. Since S is closed, x0 is not a member of the boundary ofS and so must lie some positive distance away from S; thus 0 < dE(x0, x) for everyx ∈ S. Part (i) also says that the closed set S must contain at least one point x′

that is closest to x0. A picture makes clear why these statements are true, so look atFigure 6.4.

When we use the Euclidean metric to measure distances in �2, a set of pointsequally distant from x0 is a circle. The common distance of these points from x0 isthe radius of the circle. It is clear, then, that there will be a distance at which one ofthese circles just touches the boundary of S, possibly at more than one point, as isdemonstrated in Figure 6.4. x′ and x′′ are thus both points in S that are as close aspossible to the point x0; i.e. dE(x0, x

′) = dE(x0, x′′) ≤ dE(x0, x) for all x ∈ S. Since

neither x′ nor x′′ is the same as x0, we know that this minimal distance is strictlypositive; i.e. dE(x0, x

′) = dE(x0, x′′) > 0. Now that we know what part (i) says, let’s

prove it.

Proof of (i). Choose any value for a radius μ > 0, such that the closed ball

BdE(x0, μ) = {x ∈ �n | dE(x0, x) ≤ μ} ∩ S = ∅.

BdE(x0, μ) and S are both closed sets, so their intersection BdE(x0, μ)∩S (the shadedregion in Figure 6.4) is also a closed set. BdE(x0, μ)∩S is also a bounded set becauseit is completely contained in the ball BdE(x0, μ). Hence BdE(x0, μ)∩S is a nonemptyand compact set.

Page 217: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

198 CHAPTER 6. HYPERPLANES AND SEPARATING SETS

Figure 6.4: Part (i) of Theorem 6.1.

The Euclidean metric function dE(x0, x) is a continuous function of x. There-fore, by Weierstrass’s Theorem, the function attains a minimum at some point x′ ∈BdE(x0, μ) ∩ S. That is, there is a point x′ in BdE(x0, μ) ∩ S that is least distantfrom x0:

ε = dE(x0, x′) ≤ dE(x0, x) ∀ x ∈ BdE(x0, μ) ∩ S. (6.5)

ε > 0 because x′ ∈ S and x0 /∈ S imply that x′ = x0.

Now consider any point x ∈ S that is not in the intersection BdE(x0, ε)∩S. Thenx /∈ BdE(x0, ε), and so dE(x0, x) > dE(x0, x

′).

Now let’s think about part (ii) of the theorem. The picture to have in mind nowis Figure 6.5.

Now that S is a convex set, the closed ball centered at x0 that only just touchesS can touch in only one place, x = x′. It is easy to see that there will be lots ofhyperplanes that separate the singleton closed set {x0} from the closed set S. Inparticular there will be a separating hyperplane passing through the point x′ (see thefigure). For now, don’t worry about either the point x′′ or the line joining it to thepoint x′. The proof is not complicated. We will construct a particular hyperplaneand then show that it has the properties stated in part (ii). The normal of thehyperplane is the vector p = x′ − x0 shown in the figure. The level of the hyperplane

Page 218: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

6.3. SEPARATING A POINT FROM A SET 199

Figure 6.5: Part (ii) of Theorem 6.1.

is α = p·x′ = (x′−x0)·x′. The figure displays the hyperplane as the solid straight linelabeled as p·x = α. It should be obvious from the figure that every point in S lies inthe upper-half-space of the hyperplane, while x0 lies strictly inside the hyperplane’slower-half-space, so p·x = (x′ − x0)·x ≥ α > p·x0 = (x′ − x0)·x0. Now let’s provepart (ii).

Proof of Part (ii). Choose p = x′ − x0 and α = p·x′ = (x′ − x0)·x′. p = 0 since,from part (i), x′ = x0. Also, ‖p‖E = ‖x′ − x0‖E < ∞ since there is only a finitedistance between x′ and x0.

We will start by proving that p·x0 < α.

p·x0 = (x′ − x0)·x0 = (x′ − x0)·(x0 − x′ + x′)

= −(x′ − x0)·(x′ − x0) + (x′ − x0)·x′.

x′ = x0 so (x′ − x0)·(x′ − x0) > 0. Therefore

p·x0 < (x′ − x0)·x′ = p·x′ = α.

Now we will show that p·x ≥ α for every x ∈ S. Choose any x′′ ∈ S. We must showthat p·x′′ ≥ α = p·x′. Since S is a convex set and x′ ∈ S, for any value of θ ∈ (0, 1]

Page 219: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

200 CHAPTER 6. HYPERPLANES AND SEPARATING SETS

(notice that we choose not to allow θ = 0) the point x(θ) = θx′′ + (1− θ)x′ ∈ S; seeFigure 6.5. In part (i) we established that x′ is at least as close to x0 as is other point(x(θ), for instance) in S, so

dE(x0, x′) = ‖x′ − x0‖E

≤ dE(x0, x(θ)) = ‖θx′′ + (1− θ)x′ − x0‖E = ‖θ(x′′ − x0) + (1− θ)(x′ − x0)‖E.

Squaring both sides of this equality gives us

(x′ − x0)·(x′ − x0)

≤ θ2(x′′ − x0)·(x′′ − x0) + 2θ(1− θ)(x′′ − x0)·(x′ − x0) + (1− θ)2(x′ − x0)·(x′ − x0).

That is,

0 ≤ θ2(x′′ − x0)·(x′′ − x0) + 2θ(1− θ)(x′′ − x0)·(x′ − x0)− θ(2− θ)(x′ − x0)·(x′ − x0).

Because θ > 0 (remember, θ = 0 is not allowed), we may divide by θ and obtain

0 ≤ θ(x′′−x0)·(x′′−x0)+2(1−θ)(x′′−x0)·(x′−x0)− (2−θ)(x′−x0)·(x′−x0). (6.6)

The right-hand side of (6.6) is a continuous function of θ, so by allowing θ → 0, welearn that

0 ≤ (x′ − x0)·(x′′ − x0)− (x′ − x0)·(x′ − x0) = (x′ − x0)·(x′′ − x′).

Since p = x′ − x0, this is the statement that p·x′′ ≥ p·x′ = α.All that is left to do is to show that x′ is unique. Suppose it is not unique; i.e.

suppose that there is another point x = x′, such that dE(x0, x′) = dE(x0, x). Notice

that the closed ball BdE(x0, ε) is a strictly convex set. Since x′, x ∈ BdE(x0, ε), forany θ ∈ (0, 1) the point x = θx′ + (1 − θ)x lies in the interior of BdE(x0, ε), and sodE(x0, x) < ε. Since S is a convex set, x ∈ S also. But then neither x′ nor x is apoint in S that is closest to x0. Contradiction. Hence x

′ is unique.

We now need to generalize part (ii) by removing the condition that S must bea closed set, so consider the more general problem of separating a point x0 fromany (closed, open, or neither) nonempty and convex subset S in �n. Whenever weconsider a theorem for a more general problem, we should anticipate that the resultof the theorem will be less informative than for a special case of the problem, so whatdo you expect the more general theorem’s statement to be? Think about it. (Hint:Draw a picture for the case when S is an open set.)

The change is that the strict inequality in part (ii) becomes only a weak inequality.Here is the theorem.

Page 220: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

6.3. SEPARATING A POINT FROM A SET 201

Theorem 6.2 (Separating a Point from Any Nonempty Convex Set). Let S ⊂ �n benonempty and convex. Let x0 ∈ �n with x0 /∈ S. Then there exists a hyperplane withnormal p ∈ �n, with p = 0 and ‖p‖E <∞, such that

p·x ≥ p·x0 ∀ x ∈ S.

Why is there no longer a part (i) as in Theorem 6.1? Remember that, in our proofof part (i) of Theorem 6.1, we used Weierstrass’s Theorem to establish the existenceof the point x′ in S that is closest to x0. To do so we needed to know that theintersection BdE(x, μ) ∩ S was a closed and bounded set. In the more general case,we know that the intersection is bounded, but it need not be a closed set. Becauseof this, a point such as x′ may not exist (why not?).

Why does the strict inequality in part (ii) of Theorem 6.1 change to only a weakinequality in the more general case? Have a look at Figure 6.6. If the set S is open,then the point x0 /∈ S could be a boundary point for S. In such a case, all we cansay is that p·x ≥ α = p·x0.

Figure 6.6: Case (ii), with x0 a boundary point of the open set S.

The proof of the theorem has to consider two possible cases: (i) x0 is not aboundary point of S, and (ii) x0 is a boundary point of S (as is displayed in Figure 6.6).In the proof we will again use S to denote the closure of S.

Page 221: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

202 CHAPTER 6. HYPERPLANES AND SEPARATING SETS

Proof. Case (i): x0 /∈ S. S is a closed set, so, by part (ii) of Theorem 6.1, thereexists a hyperplane with normal p ∈ �n, where p = 0 and ‖p‖E < ∞, such thatp·x > p·x0 for every x ∈ S. Since S ⊆ S, it follows that p·x > p·x0 for every x ∈ S.

Case (ii): x0 ∈ S and x0 /∈ S. Thus x0 is a boundary point for S. Consider anyopen ball centered at x0. This ball must contain points in S and points not in S.From the points in the ball that are not in S, we can form a sequence {xm}∞m=1 thatconverges to x0. Each of these points xm is exterior to the closed and convex set S, so,for each point in turn, we can apply part (ii) of Theorem 6.1. This tells us that for eachpoint xm there is a hyperplane with normal pm ∈ �m, with pm = 0 and ‖pm‖E <∞,such that pm·x > pm·xm for every x ∈ S. So now we have two sequences, {xm}∞m=1

and its companion {pm}∞m=1. This gives us a sequence of hyperplanes, as displayedin Figure 6.6. We want to show that there is a subsequence of these hyperplanesthat converges to a hyperplane that supports S at x0. We cannot guarantee thatthe sequence {pm}∞m=1 converges, but, as you will see in a moment, it must possess aconvergent subsequence.

Without any loss of generality, we will restrict ourselves to vectors pm with“lengths” (actually norms: distances from the origin) that are not greater than unity:‖p‖E ≤ 1 (this amounts merely to rescaling units of measurement). Then the vectorspm all belong to the compact ball centered at the origin 0 ∈ �n with a radius of unity.Any sequence in a compact set possesses a convergent subsequence. Thus {pm}∞m=1

possesses a convergent subsequence that we will denote by {pmk}∞k=1. Denote the limitof this subsequence by p0; p

mk → p0. Accompanying this subsequence is the subse-quence {xmk}∞k=1. The parent sequence {xm}∞m=1 converges to x0, so any subsequencemust also converge to x0: x

mk → x0.The function p·x is jointly continuous with respect to p and x, so

limk→∞

(pmk ·xmk) =(limk→∞

pmk

)·(limk→∞

xmk

)= p0·x0. (6.7)

Also, for each k = 1, 2, . . ., pmk ·x > pmk ·xmk for every x ∈ S, so

limk→∞

(pmk ·x) ≥ limk→∞

(pmk ·xmk) ∀ x ∈ S. (6.8)

Comparing (6.7) and (6.8) shows that

limk→∞

(pmk ·x) =(limk→∞

pmk

)·x = p0·x ≥ p0·x0 ∀ x ∈ S. (6.9)

Since S ⊆ S, it follows from (6.9) that

p0·x ≥ p0·x0 ∀ x ∈ S.

Page 222: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

6.4. SEPARATING A SET FROM A SET 203

6.4 Separating a Set from a Set

Theorems 6.1 and 6.2 describe the separation of two sets, but one of the sets is merelya singleton. We need to describe the separations of more general types of sets. Foreconomists one of the most useful of the more general separation theorems is due toHermann Minkowski. His theorem tells us that any pair of nonempty, disjoint, andconvex sets can be separated by a hyperplane. The result is geometrically obvious (seeall of the panels in Figures 6.2 and 6.3) and underlies much of constrained optimizationtheory and the theory of general equilibrium that primarily was developed by KennethArrow, Gerard Debreu, and Lionel McKenzie.

Theorem 6.3 (Minkowski). Let S1 and S2 be nonempty, convex and disjoint subsetsof �n. Then there exists a hyperplane with a normal p ∈ �n, where p = 0 and‖p‖E <∞, and a level α ∈ �, such that

p·x ≥ α ≥ p·y ∀ x ∈ S1 and ∀ y ∈ S2.

The proof is just a simple application of Theorem 6.2. Here is the idea. The setsS1 and S2 have no common element, so ask yourself “Does the set S1+(−S2) containthe origin, 0?” Think about it. If you have forgotten what is meant by the negative−S of a set S, then have a look in Section 2.5.

If 0 ∈ S1 + (−S2), then there is an element x ∈ S1 and an element −x ∈ −S2.But then x ∈ S1 and x ∈ S2, in which case the sets S1 and S2 are not disjoint. Thus0 /∈ S1+(−S2). Now set x0 = 0 in Theorem 6.2 and we learn that there is a hyperplanewith a normal p, such that p·(x + (−y)) ≥ p·0 = 0 for all x + (−y) ∈ S1 + (−S2).This is the same as saying that p·x ≥ p·y for all x ∈ S1 and for all y ∈ S2. It thenfollows that infx∈S1 p·x ≥ supy∈S2

p·y, so to complete the result all we need to do ischoose any value for α that is in the interval [supy∈S2

p·y, infx∈S1 p·x]. That’s it!

6.5 How Do We Use What We Have Learned?

We can use what we have learned here in two distinct and valuable ways. The firstis to use separating hyperplanes to implement computer algorithms for numericalcomputation of solutions to constrained optimization problems; this is a topic notdiscussed in this book. The second is to use hyperplanes to separate sets calledconvex cones. This is a method used to solve constrained optimization problems inwhich we may use the calculus. If you have already had some exposure to constrainedoptimization problems, then they probably were problems that you solved using some

Page 223: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

204 CHAPTER 6. HYPERPLANES AND SEPARATING SETS

form of differentiation. When you solved such a problem, you were in fact usinga hyperplane to separate two convex cones. This particular procedure is the onemost often used when the solution sought to a constrained optimization problemis algebraic, rather than numeric. For example, if we attempt to solve a consumerdemand problem not for numerical values of quantities demanded, but instead fordemand functions, then typically we will use some form of differentiation to derivethese functions. Or perhaps you want to solve a firm’s profit maximization problemfor an output supply function and input demand functions, rather than numericalvalues of quantity supplied and input quantities demanded. This is the approachmost often taken by economists, so it is the approach that is emphasized in thechapters that follow. The next task, therefore, is look at the special sets that we callcones and convex cones. These are easy ideas, so don’t worry.

Page 224: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

6.6. PROBLEMS 205

6.6 Problems

Problem 6.1. Consider the choice problem of a consumer who has an endowmentω = (10, 5), faces prices p = (3, 2), and has the direct utility function U(x1, x2) =

x1/31 x

1/32 .

(i) Compute the consumer’s ordinary demands. Then compute the consumer’s vectorz of net trades. On a diagram, accurately draw the vectors p and z. Prove that p andz are orthogonal.

(ii) Now suppose that the consumer’s utility function is U(x1, x2, x3) = x1/31 x

1/32 x

1/33 .

The consumer’s endowment is now ω = (10, 5, 4) and prices are p = (3, 2, 2). What isthe consumer’s vector z of net trades? Prove that z is orthogonal to p. Draw p andz on a three-dimensional diagram.

(iii) For each of the two above cases in turn, what hyperplane separates which sets,and what are the supports?

Problem 6.2. Why is it that, when we define a hyperplane in �n, we insist thatthe hyperplane’s normal vector p is not the n-dimensional zero vector? Answer thisquestion by asking yourself what type of set would be a “hyperplane” with a zerovector for its normal.

Problem 6.3. Consider the closed and convex subset of �2 that is X = {(x1, x2) |2x21 + x22 ≤ 9}. The boundary of this set is an ellipse centered at (0, 0). Also considerthe point x0 = (8, 2).

(i) Show that x0 is exterior to the set X.

(ii) Using the Euclidean metric to measure distances, discover the point x′ ∈ X thatis closest to x0.

(iii) Discover the hyperplane that supports X at the point x′.

Problem 6.4. A separation theorem states that, if X is a closed, convex, andnonempty subset of �n and x0 /∈ X, then there exists a vector p ∈ �n, p = 0with ‖p‖E <∞ and a scalar α ∈ �, such that

p·x0 < α ≤ p·x ∀ x ∈ X. (6.10)

Is the result valid if X is not a closed set?

Problem 6.5. Minkowski’s separating hyperplane theorem states that, “If X and Yare nonempty, convex, and disjoint subsets of �n, then there exists p ∈ �n, p = 0,‖p‖E < ∞, and α ∈ �, such that p·x ≤ α for every x ∈ X and α ≤ p·y for every

Page 225: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

206 CHAPTER 6. HYPERPLANES AND SEPARATING SETS

y ∈ Y .” The sets X and Y are disjoint and at least weakly convex, so why is it thatMinkowski did not strengthen his result to conclude that, for the given conditions,either

(a) p·x ≤ α < p·y ∀ x ∈ X, ∀ y ∈ Y , or(b) p·x < α ≤ p·y ∀ x ∈ X, ∀ y ∈ Y ?

Problem 6.6. Let X1 and X2 be two nonempty and convex subsets of �n. Thehyperplane with normal p ∈ �n, p = 0, ‖p‖E <∞ and level α supports X1 at x

∗1 and

supports X2 at x∗2. Let X = X1 +X2. Does the hyperplane with normal p and level2α support X at x∗ = x∗1 + x∗2? Prove your answer.

6.7 Answers

Answer to Problem 6.1. The consumer’s utility function is U(x1, x2) = x1/31 x

1/32 ,

her endowment is ω = (10, 5), and she faces prices p = (3, 2).

(i) The consumer’s ordinary demand functions are

x∗1(p1, p2, ω1, ω2) =p1ω1 + p2ω2

2p1and x∗2(p1, p2, ω1, ω2) =

p1ω1 + p2ω2

2p2,

so the consumer’s net ordinary demand functions are

z1(p1, p2, ω1, ω2) = x∗1(p1, p2, ω1, ω2)− ω1 =p1ω1 + p2ω2

2p1− ω1,

and z2(p1, p2, ω1, ω2) = x∗2(p1, p2, ω1, ω2)− ω2 =p1ω1 + p1ω2

2p2− ω2.

Evaluated for (ω1, ω2) = (10, 5) and (p1, p2) = (3, 2), the ordinary and net ordinaryquantities demanded are

x∗1 =20

3and x∗2 = 10, and z1 =

20

3− 10 = − 10

3and z2 = 10− 5 = 5.

Figure 6.7 displays the standard rational choice picture in which the highest availableutility is achieved at the point (x∗1, x

∗2) = (20/3, 10) on the budget constraint (the

separating hyperplane) that supports both the budget set S1 and the set S2 of bundlesthat are weakly preferred to (x∗1, x

∗1). These are the two sets that are separated by

the hyperplane. Figure 6.8 displays the vector of net trades z = (z1, z2) = (−10/3, 5)and the price vector p = (3, 2) that is the normal of the separating hyperplane. TheEuclidean inner product of these vectors is zero, proving that p and z are orthogonal.

Page 226: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

6.7. ANSWERS 207

Figure 6.7: The budget constraint is a hyperplane that separates the budget setS1 from the set S2 of consumption bundles that are weakly preferred to the bundle(20/3, 10).

(ii) The consumer’s ordinary demand functions are

x∗1(p1, p2, p3, ω1, ω2, ω3) =p1ω1 + p2ω2 + p3ω3

3p1,

x∗2(p1, p2, p3, ω1, ω2, ω3) =p1ω1 + p2ω2 + p3ω3

3p2,

and x∗3(p1, p2, p3, ω1, ω2, ω3) =p1ω1 + p2ω2 + p3ω3

3p3,

so the consumer’s net ordinary demand functions are

z1(p1, p2, p3, ω1, ω2, ω3) = x∗1(p1, p2, p3, ω1, ω2, ω3)− ω1 =p1ω1 + p2ω2 + p3ω3

3p1− ω1,

z2(p1, p2, p3, ω1, ω2, ω3) = x∗2(p1, p2, p3, ω1, ω2, ω3)− ω2 =p1ω1 + p2ω2 + p3ω3

3p2− ω2,

z3(p1, p2, p3, ω1, ω2, ω3) = x∗3(p1, p2, p3, ω1, ω2, ω3)− ω3 =p1ω1 + p2ω2 + p3ω3

3p3− ω3.

Page 227: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

208 CHAPTER 6. HYPERPLANES AND SEPARATING SETS

Figure 6.8: The hyperplane’s normal p and the net trade vector z are orthogonal.

Evaluated for (ω1, ω2, ω3) = (10, 5, 4) and (p1, p2, p3) = (3, 2, 2), the ordinary and thenet ordinary quantities demanded are

x∗1 =16

3, x∗2 = 8, x∗3 = 8 and z1 =

16

3−10 = − 14

3, z2 = 8−5 = 3, z3 = 8−4 = 4.

Both panels of Figure 6.9 display the standard rational choice picture in whichthe highest achievable utility is achieved at the point (x∗1, x

∗2, x

∗3) = (16/3, 8, 8) on

the budget constraint (the separating hyperplane) that supports both the budget setand the set of weakly preferred bundles. These are the two sets separated by thehyperplane. Figure 6.10 displays the net trade vector z = (z1, z2, z3) = (−14/3, 3, 4)and the price vector p = (3, 2, 2) that is the normal of the separating hyperplane.The Euclidean inner product of these vectors is zero, so p and z are orthogonal.(iii) For part (i), the budget hyperplane H = {x ∈ �2

+ | 3x1 + 2x2 = 40} separatesthe budget set B = {x ∈ �2

+ | 3x1 + 2x2 ≤ 40} from the set WP = {x ∈ �2+ |

U(x1, x2) ≥ (200/3)1/3}. The point of support is (x∗1, x∗2) = (20/3, 10); see Figure 6.7.For part (ii), the budget hyperplaneH = {x ∈ �3

+ | 3x1+2x2+2x3 = 48} separatesthe budget set B = {x ∈ �3

+ | 3x1 + 2x2 + 2x3 ≤ 48} from the set WP = {x ∈ �3+ |

U(x1, x2, x3) ≥ (1024/3)1/3} of consumption bundles that are weakly preferred to thebundle (x∗1, x

∗2, x

∗3) = (16/3, 8, 8). The point of support is (x∗1, x

∗2, x

∗3) = (16/3, 8, 8);

Page 228: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

6.7. ANSWERS 209

Figure 6.9: The budget constraint is a hyperplane that separates the budget set fromthe set of consumption bundles that are weakly preferred to the rationally chosenbundle (16/3, 8, 8).

Page 229: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

210 CHAPTER 6. HYPERPLANES AND SEPARATING SETS

see both panels of Figure 6.9.

Answer to Problem 6.2. A hyperplane in �n with a normal p = 0 would be a setof the form

H = {(x1, . . . , xn) | 0× x1 + · · ·+ 0× xn = α},

where α is the level of the hyperplane. Necessarily, α = 0. So what is the subset of�n that is the “hyperplane”

H = {(x1, . . . , xn) | 0× x1 + · · ·+ 0× xn = 0}?

The answer is “All of �n,” which is not a set that we have in mind when we talk ofa hyperplane.

Figure 6.10: The net trade vector z and the normal p of the separating (budgetconstraint) hyperplane are orthogonal.

Answer to Problem 6.3.

(i) For x0 = (x1, x2) = (8, 2), the value of 2x21 + x22 = 132 > 9, so (8, 2) is exterior tothe set X.

Page 230: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

6.7. ANSWERS 211

(ii) Clearly x′ = (x′1, x′2) is a boundary point for the set X, so 2 (x′1)

2 + (x′2)2 = 9.

The Euclidean distance between the point x0 = (8, 2) and the point (x′1, x′2) is

dE((x′1, x

′2), (8, 2)) = +

√(x′1 − 8)2 + (x′2 − 2)2,

so the problem to be solved is the equality-constrained minimization problem

minx1,x2

+√

(x1 − 8)2 + (x2 − 2)2 subject to g(x1, x2) = 2x21 + x22 = 9.

Figure 6.11: Finding the point (x′1, x′2) in X that is closest to the point x0 = (8, 2).

One way to solve this problem is to use the constraint 2x21 + x22 = 9 to substitutefor, say, x2 as a function of x1 in the problem’s objective function and then to usedifferentiation to seek values for x1 that may be the global minimizing solution tothe problem. For an alternative approach, consider Figure 6.11. The figure showsthat the point we seek, (x′1, x

′2), is not only on the elliptical boundary of X but is

also a point at which the marginal rate of substitution (slope) along that boundarymust be the same as the marginal rate of substitution (slope) along the boundary

Page 231: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

212 CHAPTER 6. HYPERPLANES AND SEPARATING SETS

h(x1, x2) = (x1 − 8)2 + (x2 − 2)2 = r2 of the closed ball B with radius r centered at(8, 2) with (x′1, x

′2) contained in its circumference. These marginal rate of substitutions

are

MRSg(x1, x2) =dx2dx1

|g≡9 = − 2x1x2

and MRSh(x1, x2) =dx2dx1

|h≡r2 = − x1 − 8

x2 − 2.

Since these two marginal rates of substitution are equal at (x′1, x′2),

2x′1x′2

=x′1 − 8

x′2 − 2⇒ x′2 =

4x′1x′1 + 8

.

Since (x′1, x′2) is contained in the boundary of the ellipse,

9 = 2 (x′1)2+

(4x′1x′1 + 8

)2

⇒ 2 (x′1)4+ 32 (x′1)

3+ 135 (x′1)

2 − 144x′1 − 576 = 0.

This quartic has four roots. Two of them are complex numbers, so they can beignored. One of the other two roots is negative, so we can ignore that one too. Thelast root is x′1 ≈ 2·04187, implying that

x′2 ≈4× 2·041872·04187 + 8

= 0·81334.

The point in X that is closest to x0 = (8, 2) is thus (x′1, x′2) ≈ (2·04187, 0·81334).

(iii) The normal vector that we seek is the vector p = x′ − x0 ≈ (2·04187, 0·81334)−(8, 2) = (−5·95813,−1·18666). x′ is contained in the hyperplane (denoted by H inFigure 6.11), so the level of this hyperplane is

α = p·x′ ≈ (−5·95813,−1·18666)·(2·04187, 0·81334) = −13·13088.

It is clear from Figure 6.11 that there are many other hyperplanes that strictly sep-arate x0 from the set X.

Answer to Problem 6.4. The answer is “No”; i.e. the result cannot be extended toeach and every not-closed subset of �n. The result is true for some of these not-closedsubsets, but it is not true for all such subsets.

Let’s start by thinking of a case in which the result is true for a subset X that isnot closed but not open with respect to En. For example, let x0 be the point (1, 1)and let X = [2, 3] × [2, 3] but with the point (3, 3) removed from the set. Draw thepicture. X is not closed, does not contain x0, and the result is true in this case. Andthe result would be true even if we changed X to the open set (2, 3)× (2, 3).

Page 232: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

6.7. ANSWERS 213

The situation that is of most concern is when we have X as a not-closed set (notnecessarily an open set) that has x0 as a boundary point, so that the two sets X and{x0} are disjoint but “arbitrarily close” to each other. For example, x0 = (1, 1) andX = [1, 2]× [1, 2] but with the point (1, 1) removed from X. We will show that in thiscase there does not exist even one pair (p, α), such that the hyperplane with normalp ∈ �2 (with p = 0 and ‖p‖E <∞) and level α satisfies p·x0 < α and p·x ≥ α for allx ∈ X; i.e. the result is false in this particular case.

Consider any p and any α for which x0 and X are not in the same half-space ofthe hyperplane. Draw a picture (if it makes it easier for you, then use p = (1, 1)for your diagram). Your picture should indicate to you that the only way in which ahyperplane can separate x0 = (1, 1) from X is if the hyperplane contains x0. But thenif we do as required and choose a level α, such that p·x0 < α, the hyperplane passesthrough the interior of X, in which case there are some points in X close enoughto the boundary point (1, 1) for which p·x < α. This violates the requirement thatα ≤ p·x for every x ∈ X. In this example, then, the result is false.

Answer to Problem 6.5. The example provided in the answer to problem 6.4 is acounterexample to both assertions (a) and (b).

Answer to Problem 6.6. The answer is “No” in general. Here is a counterexample.Think of �2. Let X1 be the closed square with sides of length one starting at theorigin and ending at (1,1); i.e. X1 = {(x1, x2) | 0 ≤ x1 ≤ 1, 0 ≤ x2 ≤ 1}. Let X2

be the closed square with sides of length one starting at (1,1) and ending at (2,2);i.e. X2 = {(x1, x2) | 1 ≤ x1 ≤ 2, 1 ≤ x2 ≤ 2}. Draw these two squares. Their sumX1 +X2 is the square with sides of length two starting at (1,1) and ending at (3,3).Consider the hyperplane with normal p = (1, 1) and level α = 2; i.e. p1x1+ p2x2 = 2.This hyperplane weakly separates X1 from X2. Notice that X1 and X2 are not inthe same half-spaces defined by the hyperplane. The points of support are the samepoint: x∗1 = x∗2 = (1, 1). So x∗1+x

∗2 = (1, 1)+(1, 1) = (2, 2). But (2,2) is in the interior

of X1 + X2 and so cannot be a point of support for X1 + X2. Use your diagram tocheck that this is true.

What follows is the case where the result is true.

To Be Proved. Let X1 and X2 be two nonempty and convex subsets of �n. Ahyperplane with normal p ∈ �n, p = 0, ‖p‖E <∞ and level α supports X1 at x

∗1 and

supports X2 at x∗2. The hyperplane with normal p and level 2α supports X1 +X2 atx∗ = x∗1 + x∗2 if and only if X1 and X2 are in the same half-space of the hyperplane.

Proof. We will begin by showing that, if the sets X1 and X2 are in the samehalfspace of the hyperplane with level α, then x∗ is a point of support for X1 + X2

Page 233: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

214 CHAPTER 6. HYPERPLANES AND SEPARATING SETS

in the hyperplane with level 2α. Since the hyperplane supports X1 and X2 and sincethese sets are in the same half-space of the hyperplane, it must be, without loss ofgenerality, that

p·x1 ≤ α ∀ x1 ∈ X1 and p·x2 ≤ α ∀ x2 ∈ X2 (6.11)

andp·x∗1 = α = p·x∗2. (6.12)

Therefore, from (6.11) and (6.12), for all x1 ∈ X1 and all x2 ∈ X2,

p·(x1 + x2) ≤ 2α and p·(x∗1 + x∗2) = 2α.

That is, the hyperplane with normal p and level 2α supports the set X1 + X2 atx∗ = x∗1 + x∗2.

Now suppose that the hyperplane supports X1 at x∗1, supports X2 at x∗2, andsupports X1 +X2 at x∗ = x∗1 + x∗2. Without loss of generality, suppose that

p·x ≤ 2α ∀ x ∈ X1 +X2. (6.13)

Since x∗1 and x∗2 are members of the hyperplane,

p·x∗1 = α = p·x∗2. (6.14)

Suppose that p·x′1 > α for some x′1 ∈ X1. Then, from (6.14), x′ = x′1 + x∗2 ∈ X1 +X2

but p·x′ > 2α. This contradicts (6.13). Similarly, if p·x′′2 > α for some x′′2 ∈ X2, thenx′′ = x∗1 + x′′2 ∈ X1 + X2 but p·x′′ > 2α. This also contradicts (6.13). Therefore itmust be that p·x1 ≤ α for every x1 ∈ X1 and that p·x2 ≤ α for every x2 ∈ X2. Thatis, X1 and X2 must lie in the same half-space of the hyperplane.

If X1, . . . , Xn were all nonempty and convex subsets of �n, if a hyperplane withnormal p and level α supported these sets at x∗1, . . . , x

∗n, respectively, and if X =∑n

i=1Xi, would the hyperplane with normal p and level nα support X at∑n

i=1 x∗i ?

Prove your answer.

Answer. Again, and for the same reasons, the general answer is “No.” The coun-terexample given above is a proof. Given below is the case for which the answer istrue.

To Be Proved. Let X1, . . . , Xn be nonempty and convex subsets of �n. A hyper-plane with normal p ∈ �n, p = 0, ‖p‖E < ∞ and level α supports Xi at x∗i fori = 1, . . . , n. Let X =

∑ni=1Xi. Then the hyperplane with normal p and level nα

supports X at x∗ =∑n

i=1 x∗i if and only if X1, . . . , Xn are in the same half-space of

the hyperplane.

Page 234: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

6.7. ANSWERS 215

Proof. Since the hyperplane with level α supports X1, . . . , Xn and since these setsare in the same half-space of the hyperplane, it must be, without loss of generality,that

p·xi ≤ α ∀ xi ∈ Xi, ∀ i = 1, . . . , n (6.15)

andp·x∗i = α ∀ i = 1, . . . , n. (6.16)

Therefore, from (6.15) and (6.16), for all xi ∈ Xi and for all i = 1, . . . , n,

p·(x1 + · · ·+ xn) ≤ nα and p·(x∗1 + · · ·+ x∗n) = nα. (6.17)

That is, the hyperplane with normal p and level nα supports the set X at x∗ =x∗1 + · · ·+ x∗n.

Now suppose that the hyperplane with level α supports X1, . . . , Xn at x∗1, . . . , x∗n,

respectively, and that the hyperplane with level nα supports X = X1 + · · · +Xn atx∗ = x∗1 + · · ·+ x∗n. Without loss of generality, suppose that

p·x ≤ nα ∀ x ∈ X. (6.18)

Since x∗1, . . . , x∗n are members of the hyperplane with level α,

p·x∗1 = · · · = p·x∗n = α. (6.19)

Suppose that p·x′1 > α for some x′1 ∈ X1. Then, from (6.19), x′ = x′1+x∗2+· · ·+x∗n ∈ X

but p·x′ > nα. This contradicts (6.18). The same argument can be made in turnfor each of X2, . . . , Xn. Therefore, for each i = 1, . . . , n it must be that p·xi ≤ α forevery xi ∈ Xi; i.e. X1, . . . , Xn must all lie in the same half-space of the hyperplanewith level α.

Page 235: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

Chapter 7

Cones

The most famous theorems to do with the solution of differentiable constrained opti-mization problems are the Karush-Kuhn-Tucker Theorem and Fritz John’s Theorem.Both results describe whether or not certain gradient vectors are contained in setscalled convex cones, so we need to understand what are these sets. This chapter ex-plains cones and convex cones, and then provides explanations and proofs of Farkas’sLemma and Gordan’s Lemma. These lemmas are really important results that areused in many applications of linear algebra. For us, their primary significance is thatthey can be used to prove the theorems by Karush, Kuhn, and Tucker, and by FritzJohn. Descriptions of these two essential theorems, along with their proofs, are pro-vided in Chapter 8. Let’s get started by discovering the simple structure of the setwe call a cone.

7.1 Cones

Definition 7.1 (Cone). Let a1, . . . , am ∈ �n. The cone with vertex at the origin0 ∈ �n that is generated by the vectors a1, . . . , am is the set

K(a1, . . . , am) = {x ∈ �n | x = μai, μ ≥ 0, i = 1, . . . ,m}.

The definition says that to construct a cone we start with any one of the givenvectors ai and then collect all of the nonnegative multiples μai of ai. What does thisgive us? Think about it – draw a picture for n = 2 and work it out before readingfurther.

Ready? OK then. Since μ can be zero, we know that the origin 0 is a memberof the cone. And since μ can be any positive number at all, we collect every possible

216

Page 236: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

7.1. CONES 217

positive multiple of ai. Collect all these points together, and we have constructed theray from the origin that passes through the point ai. We do this in turn for eachof the given vectors a1, . . . , am, so the cone is the set that is the union of the raysfrom the origin that pass through the points a1, . . . , am. Here is a simple example.Suppose we take the vectors a1 = (−1, 2) and a2 = (2, 3). What is the cone generatedby these vectors? Have a look at Figure 7.1. Is it the set displayed in the left-handpanel or is it the set displayed in the right-hand panel? Think before you answer.

Figure 7.1: A cone and a convex cone.

The cone K = {x ∈ �2 | x = μ(−1, 2) or x = μ(2, 3), μ ≥ 0} is the set displayedin the left-hand panel. It is the union of the two rays shown and, particularly, it doesnot contain any of the points in the shaded area shown on the right-hand panel. Ihope it is clear that a cone is a set that is closed with respect to the Euclidean metrictopology.

Why do we say the cone has a “vertex at the origin”? In other words, why is the“pointy end” of the cone located at the origin? Well, we can define a cone with avertex at any vector v simply by taking a cone K with its vertex at the origin andthen adding the vector v to every element of K. Such cones are called affine cones.For example, take the cone displayed in the left-hand panel of Figure 7.2, then “moveit” by adding the vector v = (1, 1) to every element of the cone. You now have theset that is displayed in the right-hand panel of Figure 7.2. This is an affine cone withits vertex at v = (1, 1). The only cones that we will consider are those with verticesat the origin.

An Exercise. Is the singleton set {(0, 0)} a cone with its vertex at (0, 0) in �2?

Page 237: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

218 CHAPTER 7. CONES

Figure 7.2: A cone and an affine cone.

Answer. Yes.

Another Exercise. Is any other singleton set other than {(0, 0)} a cone with itsvertex at (0, 0) in �2?

Answer. No.

7.2 Convex Cones

The set displayed in the right-hand panel of Figure 7.1 is not the cone generated bythe two vectors displayed in that diagram. It is, instead, a convex cone with its vertexat the origin.

Definition 7.2 (Convex Cone). Let K be a cone in �n. The set

CK = {x ∈ �n | x = y + z for all y, z ∈ K}

is the convex cone generated by the cone K.

So what do we have here? We start with a cone, just like the one in the left-handpanel of Figure 7.1. Then we take any two points in the cone. Adding them alwaysgives us a point in the set called the convex cone. For example, the points (−1, 2)and (2, 3) belong to the cone. The sum of these points, (1, 5), does not belong tothe cone. Instead it lies in the shaded interior of the set displayed in the right-handpanel of Figure 7.1. Try it for yourself. Pick any two points from the cone in the

Page 238: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

7.2. CONVEX CONES 219

left-hand panel, add them together, and confirm that you have a member of the set inthe right-hand panel. The convex cone generated from a cone is therefore the originalcone in union with all of the points lying between the rays that constitute the originalcone. Thus a convex cone is always a (weakly) convex set. The set displayed in theright-hand panel of Figure 7.1 is the convex cone generated by the cone displayed inthe left-hand panel.

Equivalently, we can define the convex cone generated by a cone K to be theconvex hull of K; i.e. the convex cone generated by a cone K is the smallest convexset that contains K.

Notice that a cone need not be a convex cone, but a convex cone is always a cone.It is obvious in Figure 7.1 that the cone displayed in the left-hand panel is not aconvex set and so is not a convex cone. The convex cone in the right-hand panelis a cone. Why? Because for any point in the convex cone, the ray from the originpassing through that point is entirely in the convex cone. Thus the convex cone is acone. Particularly, it is a cone that is generated by the infinitely many vectors ai thatare rays between the displayed (boundary) vectors. Although Figure 7.1 displays thetypical case, it is nevertheless possible for a cone and the convex cone generated byit to be the same set. Can you see when this will be true? Think it over. Remember,a convex cone is always a weakly convex set.

Figured it out? If not, then here is another clue. Think of a cone that consists ofjust one ray – say, the ray from the origin passing through (1, 1). What do you noticeabout this cone? Now try again to figure out the answer.

Let’s see how you did. I hope you noticed that the cone consisting of just oneray is a convex set. The convex cone generated by a cone is the smallest convex setthat contains the cone, so the cone and the convex cone generated by it are the samewhen the cone is its own convex hull. This happens if and only if the convex hull hasan empty interior. There are just three possibilities – what are they?

One is that the cone consists of a single ray from the origin. The second is thatthe cone consists of two rays, each pointing in the exactly opposite direction to theother (so the cone is a hyperplane through the origin). The third possibility is thatthe cone is the singleton set {0}.

Often the same convex cone can be generated by many cones. Think again of thecone displayed in the left-hand panel of Figure 7.1, and now add a third ray, one thatlies between the original two rays; e.g. the ray that is the vertical axis. The convexcone generated by this new cone is still the set displayed in the right-hand panel.Add as many more rays as you want to the cone in the left-hand panel, but makesure each new ray lies between the original two. Each of these new cones generates

Page 239: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

220 CHAPTER 7. CONES

the same convex cone, the one in the right-hand panel.

Sometimes adding a new ray to a cone can dramatically alter the convex coneit generates. Look again at the cone in the left-hand panel of Figure 7.1, then addto it the ray from the origin that passes through (0,−1). What is the convex conegenerated by this new cone? Think about it – what is the smallest convex set thatcontains the new cone?

The answer is �2. Adding just one new vector can greatly increase the size of theconvex cone.

Is it clear to you that, like a cone, a convex cone is a set that is closed with respectto the Euclidean topology?

What use are cones? You may be surprised to discover that most of the theory ofdifferentiable constrained optimization makes heavy use of cones. Most of the majortheorems to do with constrained optimization are statements describing something todo with cones. For example, the famous Karush-Kuhn-Tucker Theorem of constrainedoptimization theory is a description of conditions under which a particular vector liesinside a particular convex cone. So is its companion, the Fritz John Theorem. Thetwo really useful theorems to do with cones that we need to understand are Farkas’sLemma and Gordan’s Lemma. Both are examples of “theorems of the alternative.”There are lots of such theorems and they all take the same form. Each is a statementthat says here are two alternatives, A and B. One of them must be true, and whenA is true, B is false and conversely. The alternatives A and B are thus mutuallyexclusive. Here is a rather silly, but valid, theorem of the alternative: “Alternative Ais that the Earth is flat. Alternative B is that the Earth is not flat. The theorem isthat A is true if and only if B is false, and that one of A and B must be true.” It istime for us to have a close look at the two lemmas mentioned above. Let’s start withthe simpler of the two.

7.3 Farkas’s Lemma

Gyuala Farkas (also known as Julius Farkas) was a Hungarian physicist and mathe-matician who in 1902 published the result named after him. Here it is.

Lemma 7.1 (Farkas’s Lemma). Given vectors a1, . . . , am, b ∈ �n with b = 0, one andonly one of the following two statements is true.

(i) There exist λ1, . . . , λm ∈ �, all nonnegative and not all zero, such that

b = λ1a1 + · · ·+ λma

m. (7.1)

Page 240: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

7.3. FARKAS’S LEMMA 221

(ii) There exists x ∈ �n, such that

a1·x ≥ 0, . . . , am·x ≥ 0 and b·x < 0. (7.2)

When I first read this lemma, my reaction was something like, “What on earthis this talking about and why should I care?” Perhaps you are feeling the same way,so I suggest that we defer proving the result until we understand what it says. Goodnews – what it says is actually very simple. A picture will help. Look at Figure 7.3.

Figure 7.3: Farkas’s two alternatives.

The left-hand panel shows the vector b as a member of the convex cone generatedby a1 and a2. This is alternative (i), which says that b can be written as a nonnegativelinear combination of a1 and a2. Any vector constructed as a nonnegative linearcombination of two other vectors lies on a ray from the origin that is between therays containing the two other vectors (see Figure 2.4 – use the parallelogram rule ofvector addition). Since alternative (ii) is true when (i) is false, (ii) must describe thecase when b is not a member of the convex cone generated by a1 and a2. This isdisplayed in the right-hand panel. I understand that it still might not be obvious toyou that (7.2) describes this situation, so have a look at Figure 7.4, which is the sameas Figure 7.3 except that I have added dashed lines that are orthogonal to each of b,a1, and a2.

The “closed” (solid dots at the ends) dotted half-circles show the regions of pointsx where a1·x ≥ 0 and a2·x ≥ 0. The “open” (hollow dots at the ends) dotted half-circle shows the regions of points x where b·x < 0. In the left-hand panel, nowhere

Page 241: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

222 CHAPTER 7. CONES

Figure 7.4: Farkas’s second alternative.

do all of these three half-circles overlap each other, showing that, when (i) is true,there does not exist any x for which (7.2) is true. In the right-hand panel, thereis a (shaded) region where all of the half-circles overlap; there are many points xsatisfying (7.2), so (ii) is true. This is so because b is not a member of the convexcone generated by a1 and a2. Thus, (ii) is true if and only if (i) is false.

Before you proceed to the proof of Farkas’s Lemma, try the following little exercise.Suppose that alternative (ii) is altered by changing b·x < 0 to b·x ≤ 0. Is it nowpossible for both alternatives (i) and (ii) to be true? Draw yourself some diagramslike those in Figure 7.4 and discover why the answer is “Yes.”

Now it is time to prove the lemma. Fear not. There is nothing hard in the proof.In fact, the heart of the proof is displayed in the right-hand panel of Figure 7.4. Thispanel is reproduced in Figure 7.5. Notice that, when (i) is false, the point b is outsidethe convex cone generated by a1 and a2. The singleton set {b} is nonempty, convex,and closed. So is the convex cone. The convex cone and {b} are disjoint. We knowtherefore (see Theorem 6.1) that there must be a hyperplane that separates the twosets. One of these hyperplanes is the hyperplane with normal x that passes throughthe origin – see the dotted line labelled H in Figure 7.5. This is the hyperplaneH = {z ∈ �2 | x·z = 0} and it should be clear that x·b < 0 and x·z ≥ 0 for all points

Page 242: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

7.3. FARKAS’S LEMMA 223

Figure 7.5: Farkas’s second alternative is that a hyperplane through the origin sepa-rates the singleton set {b} from the convex cone generated by a1 and a2.

on the right-hand side (i.e. the upper-half-space) of the hyperplane; this includes allpoints in the cone. Set z equal to a1 and a2 in particular and these inequalities arealternative (ii) of Farkas’s Lemma.

Proof of Farkas’s Lemma. We will show that (ii) is true if and only if (i) is false.

We will begin by showing that, if (i) is false, then (ii) is true, so take as a factthat there do not exist scalars λ1, . . . , λm that are all nonnegative and not all zerothat allow us to write the vector b as the linear combination (7.1). Use CK to denotethe convex cone generated by a1, . . . , am. Since (7.1) is a statement that b is a vectorcontained in CK and since (7.1) is false, b must be a vector that lies outside CK .That is, the singleton set {b} is disjoint from CK . Both {b} and CK are nonempty,closed, and convex sets, so by Theorem 6.1 there exists a hyperplane with a normalx ∈ �n, such that x = 0, ‖x‖E < ∞, x·b < 0 and x·z ≥ 0 for every z ∈ CK . Sincea1, . . . , am ∈ CK , we have proved that (i) is false implies that (ii) is true.

Now we will show that (ii) true implies that (i) is false, so take as a fact that there

Page 243: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

224 CHAPTER 7. CONES

exists at least one x ∈ �n, such that (7.2) is true. If (7.1) is also true, then

x·b = λ1x·a1 + · · ·+ λmx·am. (7.3)

x·a1 ≥ 0, . . . , x·am ≥ 0 because (ii) is true. If (i) is also true, then λ1 ≥ 0, . . . , λm ≥0 and so the right-hand side of (7.3) must be nonnegative; i.e. x·b ≥ 0, whichcontradicts that (ii) is true. Hence (i) cannot be true when (ii) is true.

Farkas’s Lemma has a wide variety of valuable applications. It is, in fact, oneof the fundamental results of linear algebra. The only application described in thisbook is to the proof of the famous Karush-Kuhn-Tucker Theorem on necessary opti-mality conditions for differentiable constrained optimization problems. This is foundin Chapter 8, where you will find that the Karush-Kuhn-Tucker Theorem assumessomething called a “constraint qualification.” This is a bit of a nuisance when solv-ing some types of constrained optimization problems. Fortunately another statementabout necessary optimality conditions is available to us that does not require that anyconstraint qualification is satisfied. This is Fritz John’s Theorem. It too is presentedand proved in the constrained optimization chapter. The reason for mentioning ithere is that its proof relies on a different theorem of the alternative, called Gordan’sLemma, that also describes things to do with convex cones. Because the Fritz JohnTheorem is so useful, we should understand Gordan’s Lemma before we depart ourinspection of cones and their uses.

7.4 Gordan’s Lemma

We mentioned earlier that, like Farkas’s Lemma, Gordan’s Lemma is a theorem of thealternative. Also like Farkas’s Lemma, Gordan’s Lemma describes when a particularvector does and does not lie inside a particular convex cone.

Lemma 7.2 (Gordan’s Lemma). Given vectors a1, . . . , am ∈ �n, one and only oneof the following two statements is true.

(i) There exist λ1, . . . , λm ∈ �, all nonnegative and not all zero, such that

λ1a1 + · · ·+ λma

m = 0. (7.4)

(ii) There exists x ∈ �n, such that

a1·x > 0, . . . , am·x > 0. (7.5)

Page 244: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

7.4. GORDAN’S LEMMA 225

Figure 7.6: Gordan’s two alternatives.

Figure 7.6 helps to make sense of what seems at first glance to mean not much.The figure is drawn for the case of m = 3 in �2.

Suppose that (i) is true. Then we know, first, that there exist scalars λ1 ≥ 0,λ2 ≥ 0 and λ3 ≥ 0 that are not all zero. Therefore at least one of them is strictlypositive. We will suppose that at least λ1 > 0. Second, we know that

λ1a1 + λ2a

2 + λ3a3 = 0. (7.6)

Since λ1 = 0, we can rewrite (7.6) as

−a1 = λ2λ1a2 +

λ3λ1a3. (7.7)

Now let’s see – what does (7.7) say? We know that λ2/λ1 ≥ 0 and λ3/λ1 ≥ 0, so

λ2λ1a2 +

λ3λ1a3

is a nonnegative linear combination of a2 and a3 and so is a vector contained in theconvex cone generated by a2 and a3. (7.7) says that this vector is the negative of a1.

Page 245: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

226 CHAPTER 7. CONES

The left-hand panel of Figure 7.6 displays this case. If alternative (i) is false, then−a1 does not lie in the convex cone generated by a2 and a3. This is displayed in theright-hand panel of Figure 7.6. But I think I hear you muttering that you do not seewhy the dot products displayed in (7.5) tell you that we have the situation displayedin the right-hand panel. Things will look a bit clearer to you if I redraw Figure 7.6,so have a look at Figure 7.7. All I have done is add to Figure 7.6 the dotted linesthat are orthogonal to a1, a2, and a3.

Figure 7.7: A different perspective of Gordan’s two alternatives.

Start with the left-hand panel, where alternative (i) is true. Each semi-circleindicates the regions (open half-spaces, actually) in which there are points x thatform a strictly positive inner product with a vector ai (each semi-circle is “open” atits ends since at the ends the inner products are zero). You can see that nowheredo all three semi-circles overlap each other. In other words, there is no point x thatforms a strictly positive inner product with all of a1, a2, and a3. If alternative (i) istrue, then alternative (ii) must be false.

Now look at the right-hand panel, where alternative (i) is false since −a1 doesnot lie in the convex cone generated by a2 and a3. Observe that now the three semi-circles all overlap in the shaded region. The inner product of any point x in the

Page 246: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

7.4. GORDAN’S LEMMA 227

shaded region with each of a1, a2, and a3 is strictly positive. Hence alternative (ii) istrue.

I hope you now have an idea of what Gordan’s Lemma says. It probably stillseems to you that the lemma is not of much use, but I promise you it really is andthat it is worth the effort of proving. Let’s do it.

Proof of Gordan’s Lemma. We will begin by proving that, when (ii) is false, (i)must be true. The statement that (ii) is false is the statement that there does notexist even one x ∈ �n, such that a1·x > 0, . . . , am·x > 0. In other words,

for every x ∈ �n, ai·x ≤ 0 for at least one i ∈ {1, . . . ,m}. (7.8)

Define the set

V = {(a1·x, . . . , am·x) | x ∈ �n}. (7.9)

Notice that any v ∈ V has m, not n, elements a1·x up to am·x and so belongs to �m,not �n. The n-dimensional vectors a1, . . . , am are fixed and given. V is the collectionof all of the m-dimensional vectors formed by computing the inner products of thevectors a1, . . . , am with every n-dimensional vector x ∈ �n. Since (ii) is false, weknow from (7.8) that there is no strictly positive vector in V ; i.e. v � 0 ∀ v ∈ V .(Don’t make the mistake of thinking that every vector in V consists only of negativeor zero elements. A vector in V can have positive elements. But every vector in Vmust have at least one element that is either zero or negative.)

Now define the set

P = {y ∈ �m | y � 0}. (7.10)

You need to be clear that V and P are both subsets of �m (not of �n). Notice thatany vector y ∈ P has only strictly positive elements. It follows that V and P aredisjoint. Notice that 0m /∈ P and that 0m ∈ V (because (a1·x, . . . , am·x) = 0m whenwe choose x = 0n).

Now is the time for you to try drawing the sets P and V in �2. This is animportant exercise, so take your time, think about it, and then create your picture.P is just �2

++ – that’s easy. Now try drawing V . I’ll give you a warning – mostpeople get it wrong, and that is why I am urging that you try your very best now sothat I can later use your picture to help you out. Go on – give it a serious try beforereading further. (Hint: Remember that 0m ∈ V .)

It is obvious that P is nonempty, convex, and open with respect to Em. It issimilarly obvious that V is nonempty, and it is reasonably easy to see that V is closedin Em, but perhaps it is not so obvious that V is a convex subset of Rm. Why is

Page 247: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

228 CHAPTER 7. CONES

V a convex subset of �m? Take any two vectors v′, v′′ ∈ V . Then there are vectorsx′, x′′ ∈ �n, such that

v′ = (a1·x′, . . . , am·x′) and v′′ = (a1·x′′, . . . , am·x′′). (7.11)

For any θ ∈ [0, 1], write

v(θ) = θv′ + (1− θ)v′′

= θ(a1·x′, . . . , am·x′) + (1− θ)(a1·x′′, . . . , am·x′′)= (a1·(θx′ + (1− θ)x′′), . . . , am·(θx′ + (1− θ)x′′)). (7.12)

θx′+(1−θ)x′′ ∈ �n since �n is a convex space and x′, x′′ ∈ �n. Therefore v(θ) ∈ �m,establishing that V is a convex subset of �m.

Now look at the picture you drew of P and V in �2. You probably have drawnone of the diagrams displayed in Figure 7.8.

Figure 7.8: Where is the set V ?

If you drew the diagram in the left-hand panel that says V is everything in �2

except �2++, then you have made the mistake made by most first-timers (including

me). Remember that V is a convex space, so it cannot be the nonconvex set displayedin the left-hand panel. What about the middle panel, which suggests that V is �2

−,the negative orthant of �2? No. Why not? Because vectors in V are allowed to havesome positive elements. So we are looking for a convex space that contains the originand does not intersect with P = �2

++. There it is, in the right-hand panel. V is quitesmall. It is only a hyperplane passing through the origin and not intersecting with�2

++. Take a minute and check out that there is no other possibility for V . Thentry to draw both P and V in �3. You will discover that P = �3

++ and that V isa (two-dimensional) hyperplane passing through the origin of the three-dimensional

Page 248: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

7.4. GORDAN’S LEMMA 229

space �3 and not intersecting �3++. I hope it has become clear that in the general

case V is an (m − 1)-dimensional hyperplane in �m that passes through the origin0m and does not intersect �m

++. If this is clear to you, then the hard part is over andwe are ready to complete our argument.

P and V are both nonempty and convex subsets of �m. P and V are disjoint.Hence we can apply Minkowski’s Separation Theorem and conclude that there existsat least one (normal) vector λ ∈ �m with λ = 0m and ‖λ‖E <∞, and a level α, suchthat

λ·v ≤ α for every v ∈ V and λ·y ≥ α for every y ∈ P. (7.13)

Recollect that 0m ∈ V , so it is immediate from (7.13) that the separating hy-perplane’s level α ≥ 0. Look back at the right-hand panel of Figure 7.8. V is thehyperplane passing through the origin. What is the level of any such hyperplane? Itmust be zero. (Recall that a hyperplane is a set {x ∈ �m | p·x = α}. p is the “normal”of the hyperplane and α is its “level.”) This suggests that perhaps α = 0 in (7.13).This is indeed the case, so we are now about to show that α > 0 is impossible. Theargument is by a simple contradiction. Suppose that α > 0. Then we are assertingthat there is a strictly positive upper bound, α, to the values of inner products λ·vfor all of the elements v ∈ V , so there must be in V an element v′ for which λ·v′ > 0.For any μ > 0, v′ ∈ V implies that μv′ ∈ V (because v′ = (a1·x′, . . . , am·x′) for somex′ ∈ �n and x′ ∈ �n implies that μx′ ∈ �n, so μv′ = (a1·μx′, . . . , am·μx′) ∈ V ). Bychoosing μ to be as large as we want, we can cause λ·μv′ to become as large as welike, so there cannot be a strictly positive upper bound on the values of these innerproducts. Contradiction. Hence α > 0. Since α ≥ 0, we must conclude that α = 0.

So now we know that λ·v ≤ 0 for all v ∈ V . The next step is to show that λ·v < 0is impossible. Suppose that there is a v ∈ V for which λ·v < 0. Then there is anx ∈ �n such that v = (a1·x, . . . , am·x). x ∈ �n implies that −x ∈ �n and, therefore,that (a1·(−x), . . . , am·(−x) = −v ∈ V . But then λ·(−v) > 0, which contradicts thatλ·v ≤ 0 for every v ∈ V . Thus there is no v ∈ V for which λ·v < 0. Consequently wehave that λ·v = 0 for every v ∈ V . This establishes that the set V is a hyperplane in�m passing through the origin.

Let’s take a moment, collect our thoughts, and see what we still have to do. Wehave established that there exists a λ ∈ �m, such that λ·v = 0 for every v ∈ V . Byconstruction, an element in V is a vector v = (a1·x, . . . , am·x) for some x ∈ �n. Wehave yet to show that λ1 ≥ 0, . . . , λm ≥ 0, that at least one λi > 0, and that (7.4) istrue. The good news is that none of these is difficult to do.

First, not every λi = 0 because Minkowski’s Theorem tells us that λ = 0m.

Page 249: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

230 CHAPTER 7. CONES

Second, suppose λi < 0 for at least one i ∈ {1, . . . ,m}. We know that P containsvectors y that have strictly positive values μ for their ith element and have all otherelements equal to unity, where μ can be arbitrarily large. λ·y =

∑nj=1j �=i

λj + λiμ, so

λ·y < 0 for a large enough value of μ, contradicting that λ·y ≥ 0 for every y ∈ P .Therefore λi ≥ 0 for every i = 1, . . . ,m.

All that is left to show is that (7.4) is true. We are free to choose any x ∈ �n toconstruct an element of V , so let’s choose x′′ = λ1a

1 + · · ·+ λmam. Then

0 = λ·(a1·x′′, . . . , am·x′′)

= (λ1, . . . , λm)

⎛⎜⎝a

1·x′′...

am·x′′

⎞⎟⎠

= λ1a1·x′′ + · · ·+ λma

m·x′′

= (λ1a1 + · · ·+ λma

m)·x′′

= (λ1a1 + · · ·+ λma

m)·(λ1a1 + · · ·+ λmam). (7.14)

Hence

λ1a1 + · · ·+ λma

m = 0n. (7.15)

We are there; when (ii) is false, (i) is true.

That was quite a long argument, so you will be relieved to know that proving that(ii) is false when (i) is true is quick and simple. Suppose that (i) is true. Then thereis at least one λi > 0. Without loss of generality, suppose that λ1 > 0. Then we canrearrange (7.4) to

a1 = −λ2λ1a2 − · · · − λm

λ1am (7.16)

and so obtain that

a1·x = −λ2λ1a2·x− · · · − λm

λ1am·x. (7.17)

Suppose that (ii) is true also. Then there is at least one x ∈ �n for which a1·x > 0,a2·x > 0, . . . , am·x > 0. Then the right-hand side of (7.17) is not positive, contra-dicting that a1·x > 0. Hence when (i) is true, (ii) must be false.

I will make no effort here to convince you that all the hard work of proving thevaluable lemmas of Farkas and Gordan is worth it. All I can do at this moment isask you to trust me that it is so and to wait until we are working on constrainedoptimization problems to justify that trust.

Page 250: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

7.5. REMARKS 231

7.5 Remarks

We have reached the end of what we might call preparatory materials. The nextchapter introduces the type of constrained optimization problem mostly encounteredby economists. Its solution uses all of the contents of the chapters on hyperplanes andcones. It will turn out, as a result of the work already done to understand hyperplanesand cones, that understanding the methods used to solve constrained optimizationproblems is not a big step from where we are now. In fact, mostly all we need todo is apply what we have already learned to the particular context of constrainedoptimization problems. In the next chapter we will understand and prove two ofthe major theorems that commonly are employed to solve constrained optimizationproblems. These are the first-order necessity results discovered by Karush, Kuhn,and Tucker, and by Fritz John. We will use both theorems to solve a variety ofproblems and draw lots of pictures to be sure that we see the quite simple geometrythat explains the theorems and their applications. It is quite enjoyable, I think.

After that there remains a task that is very important to economists. Economistsare particularly interested in the properties of solutions to constrained optimiza-tion problems. For example, an economist is interested in whether or not quantitiesdemanded decrease as prices increase, and how quantities demanded change as con-sumers’ disposable incomes change. Similarly, economists want to know the ratesat which quantities supplied of products change as prices change. Answering suchinquiries requires that the properties of a constrained optimization problem’s opti-mal solution be discovered. Determining the properties of optimal solutions needsmore information that is provided by either the Karush-Kuhn-Tucker or Fritz JohnTheorems. We also need information provided by a second-order necessary condition.Statements of this condition are usually made using something called a Lagrangefunction, so, after the next chapter, we will examine Lagrange functions and thencarefully examine the second-order necessity condition. We are then ready to learnhow to establish the properties of optimal solutions, a task that occupies the lastthree chapters of this book.

So, let’s go. Let’s, at last, get to grips with constrained optimization problems.

Page 251: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

232 CHAPTER 7. CONES

7.6 Problems

Problem 7.1.

a1 =

(21

), a2 =

(01

), a3 =

(−12

), and b =

(10

). (7.18)

(i) Draw the convex cone with vertex at the origin that is generated by the vectorsa1, a2, and a3.

(ii) Add the vector b to your diagram.

(iii) Which of Farkas’s alternatives is true?

(iv) If the first of Farkas’s alternatives is true, then solve Aλ = b for λ for at leasttwo solutions and confirm for each that λ ≥ 0. If the second alternative is true, thendiscover at least two x ∈ �2 for which Ax ≥ 0 and b·x < 0.

(v) Now set b = ( 11 ) and repeat parts (i) to (iv) above.

Problem 7.2.

a1 =

(21

)and a2 =

(−6−3

).

(i) What is the convex cone generated by the vectors a1 and a2?

(ii) Does there exist a vector x ∈ �2 such that a1·x ≥ 0 and a2·x ≥ 0? If not, thenwhy not? If so, then what is the complete collection of such vectors x?

Problem 7.3. Our statement of Farkas’s Lemma is: Let a1, . . . , am ∈ �n and letb ∈ �n, b = 0. Then exactly one of the following two alternatives is true:

(i) There exist λ1, . . . , λm ∈ �, all nonnegative and not all zero, such that

b =m∑j=1

λjaj.

(ii) There exists x ∈ �n such that

b·x < 0 and aj·x ≥ 0 for all j = 1, . . . ,m.

Use Farkas’s Lemma to show that, if A is an (n×m)-matrix and b ∈ �n, b = 0 ,then either there is at least one nonzero and nonnegative solution λ to the linearequation system Aλ = b or there exists an x ∈ �n for which xTA ≥ 0 and b·x < 0.

Page 252: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

7.7. ANSWERS 233

Problem 7.4. a2, a3 ∈ �2 are nonzero and negatively collinear vectors. a1 ∈ �2 is anonzero vector that is not collinear with a2 or a3. Explain why it is, in such a case,that

(i) there do not exist λ2 ≥ 0 and λ3 ≥ 0, not both zero, such that a1 = λ2a2 + λ3a

3,and why

(ii) there exist λ1 ≥ 0, λ2 ≥ 0 and λ3 ≥ 0, not all zero, such that λ1a1+λ2a

2+λ3a3 =

0.

Problem 7.5. Let a1, a2, a3 ∈ �2 be nonzero vectors. Suppose that there exists anx ∈ �2 such that a1·x = 0, a2·x > 0 and a3·x > 0. Explain why, for this circumstance,

(i) there do not exist scalars λ1 = 1 and λ2 and λ3, both nonnegative and not bothzero, such that λ1a

1 = λ2a2 + λ3a

3, and

(ii) there do not exist scalars μ1 = 1 and μ2 and μ3, both nonnegative and not bothzero, such that μ1a

1 + μ2a2 + μ3a

3 = 0.

Draw two accurate and well-labeled diagrams to illustrate your answers to parts (i)and (ii).

7.7 Answers

Answer to Problem 7.1.

(i) and (ii) See the upper panel of Figure 7.9.

(iii) and (iv) The vector b = (1, 0)T is not contained in the convex cone that isgenerated by the vectors a1, a2 and a3. Therefore alternative 2 of Farkas’s Lemma istrue and alternative 1 is false. That is, there do not exist scalars λ1, λ2, and λ3, allnonnegative and not all zero, such that

λ1

(21

)+ λ2

(01

)+ λ3

(−12

)=

(10

). (7.19)

Equivalently, there exists at least one point (x1, x2) such that

a1·x = (2, 1)

(x1x2

)≥ 0, a2·x = (0, 1)

(x1x2

)≥ 0, a3·x = (−1, 2)

(x1x2

)≥ 0,

and b·x = (1, 0)

(x1x2

)< 0.

(7.20)

Page 253: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

234 CHAPTER 7. CONES

Figure 7.9: The graphs for problem 7.1.

The solutions to (7.19) are λ2 = 2 − 5λ1 and λ3 = 2λ1 − 1. λ2 ≥ 0 if λ1 ≤ 2/5,while λ3 ≥ 0 if λ1 ≥ 1/2. Thus there do not exist values for λ1, λ2, and λ3 that areall nonnegative, not all zero, and satisfying (7.19). x = (−1, 3) is one of many pointssatisfying (7.20).

(v) In the lower panel of Figure 7.9 the vector b = (1, 1)T lies in the convex conegenerated by the vectors a1, a2, and a3, so alternative 1 of Farkas’s Lemma is trueand alternative 2 is false. There must exist nonnegative scalars λ1, λ2, and λ3, notall zero, such that

λ1

(21

)+ λ2

(01

)+ λ3

(−12

)=

(11

). (7.21)

Page 254: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

7.7. ANSWERS 235

Equivalently, there does not exist any point (x1, x2) such that

a1·x = (2, 1)

(x1x2

)≥ 0, a2·x = (0, 1)

(x1x2

)≥ 0, a3·x = (−1, 2)

(x1x2

)≥ 0,

and b·x = (1, 1)

(x1x2

)< 0.

(7.22)

The solutions to (7.21) are λ2 = 3 − 5λ1 and λ3 = 2λ1 − 1. λ2 ≥ 0 if λ1 ≤ 3/5,while λ3 ≥ 0 if λ1 ≥ 1/2. Thus (7.21) is satisfied by values of λ1, λ2, and λ3 that areall nonnegative and not all zero for 1/2 ≤ λ1 ≤ 3/5. There is no point (x1, x2) thatsatisfies (7.22).

Answer to Problem 7.2.

(i) The convex cone is the hyperplane with normal p = (1,−2) and level α = 0.

(ii) x is any vector that is orthogonal to the hyperplane; e.g. x = p = (1,−2). Thecomplete set of such vectors is the set {(x1, x2) | (x1, x2) = μ(1,−2) ∀ μ ∈ �}.Answer to Problem 7.3. Expanded, the matrix equation

Aλ =

⎛⎜⎜⎜⎝a11 a12 · · · a1ma21 a22 · · · a2m...

.... . .

...an1 an2 · · · anm

⎞⎟⎟⎟⎠⎛⎜⎜⎜⎝λ1λ2...λm

⎞⎟⎟⎟⎠ = b =

⎛⎜⎜⎜⎝b1b2...bn

⎞⎟⎟⎟⎠

is

a11λ1 + a12λ2 + · · ·+ a1mλm = b1

a21λ1 + a22λ2 + · · ·+ a2mλm = b2...

...

an1λ1 + an2λ2 + · · ·+ anmλm = bn.

This is the same as the vector equation

λ1

⎛⎜⎜⎜⎝a11a21...an1

⎞⎟⎟⎟⎠+λ2

⎛⎜⎜⎜⎝a12a22...an2

⎞⎟⎟⎟⎠+· · ·+λm

⎛⎜⎜⎜⎝a1ma2m...

anm

⎞⎟⎟⎟⎠ = λ1a

1+λ2a2+· · ·+λmam =

⎛⎜⎜⎜⎝b1b2...bn

⎞⎟⎟⎟⎠ . (7.23)

Page 255: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

236 CHAPTER 7. CONES

Farkas’s Lemma states that there exist scalars λ1, . . . , λm, all nonnegative and not allzero, that solve (7.23) if and only if there does not exist any point x = (x1, · · · , xn)satisfying

b·x < 0 and aj·x = x·aj ≥ 0 ∀ j = 1, . . . ,m.

That is, if and only if there does not exist an x satisfying

b·x < 0 and (x1, x2, · · · , xn)

⎛⎜⎜⎜⎝a11 a12 · · · a1ma21 a22 · · · a2m...

.... . .

...an1 an2 · · · anm

⎞⎟⎟⎟⎠ ≥

⎛⎜⎜⎜⎝00...0

⎞⎟⎟⎟⎠ .

Answer to Problem 7.4.

(i) Since a2 and a3 are nonzero and negatively collinear vectors, the convex cone thatthey generate is a hyperplane through the origin. The vector a1 is not contained inthis cone since a1 is nonzero and is not collinear with a2 or a3. Therefore a1 cannotbe written as any linear combination λ2a

2 + λ3a3, no matter what are the weights λ2

and λ3.

(ii) Since a2 and a3 are nonzero and negatively collinear, we may write a2 + μa3 = 0for some μ > 0. Now write λ3/λ2 = μ with λ2 > 0 so that we have λ3 = μλ2 > 0.Then λ2a

2 + λ3a3 = 0. Now set λ1 = 0, and we have λ1a

1 + λ2a2 + λ3a

3 = 0 withλ1 ≥ 0, λ2 ≥ 0, λ3 ≥ 0 and not all of λ1, λ2, and λ3 zero.

Answer to Problem 7.5. In both panels of Figure 7.10, the set S of points xfor which a2·x > 0 and a3·x > 0 is the shaded region bordered by the hyperplanesa2·x = 0 and a3·x = 0. The set contains no points belonging to either hyperplane.

(i) Examine the upper panel of Figure 7.10. The set of points a1, such that a1 =λ2a

2+λ3a3 with λ2 ≥ 0, λ3 ≥ 0 and λ2, λ3 not both zero, is the convex cone K, minus

the origin, that is bordered by the rays containing a2 and a3. The question is, “Whyis there not even one point x in the set S and one vector a1 in the cone K that areorthogonal?” It should be clear from the figure that the vectors that are orthogonalto at least one of the vectors in the cone K are all contained in the subsets of �2

labeled “Region I” and “Region II.” These subsets contain vectors in the hyperplanesa2·x = 0 and a3·x = 0. But S does not, so there are no vectors common to the set Sand either Region I or Region II.

(ii) Examine the lower panel of Figure 7.10. The set of points a1, such that a1 +λ2a

2+λ3a3 = 0 with λ2 ≥ 0, λ3 ≥ 0 and λ2, λ3 not both zero, is labeled −K since this

set of points is exactly the negative of the set K discussed in part (i). The question

Page 256: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

7.7. ANSWERS 237

is, “Why is there not even one point x in the set S and one vector a1 in the cone−K that are orthogonal?” The vectors orthogonal to at least one of the vectors inthe cone −K are contained in the subsets of �2 labeled “Region I” and “Region II.”These subsets contain vectors in the hyperplanes a2·x = 0 and a3·x = 0. But S doesnot, so there are no vectors common to the set S and either Region I or Region II.

Figure 7.10: Diagrams for problem 7.5.

Page 257: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

Chapter 8

Constrained Optimization, Part I

8.1 The Constrained Optimization Problem

Constrained optimization is the economist’s primary means of modeling rationalchoice, the fundamental underpinning of modern economics. This chapter is thereforecrucial to your understanding of most economic theories. The good news, and it isvery good news, is that the core ideas of constrained optimization are rather obvious.Let’s get started by setting out what is meant by the term constrained optimizationproblem.

Definition 8.1 (Constrained Optimization Problem). Let C be a nonempty set. Letf : C �→ � be a function. Then the problem

maxx∈C

f(x) ≡ maxx

f(x) subject to x ∈ C (8.1)

is a constrained optimization problem.

The set of all of the choices from which the decisionmaker may select is denotedby C and is called the “choice set,” the “constraint set,” or the “feasible set.” Con-straint set is not a good name for C because C is not a set of constraints. The othertwo names are perfectly reasonable, so let’s follow the common convention of callingC the feasible set ; i.e. a choice x is feasible, meaning available for selection, if andonly if it satisfies all of the constraints imposed upon choices. A common economicexample is the set of consumption bundle choices x = (x1, . . . , xn) that are feasiblebecause they are affordable for a consumer who has $y to spend and faces pricesp = (p1, . . . , pn);

C(p1, . . . , pn, y) = {(x1, . . . , xn) | x1 ≥ 0, . . . , xn ≥ 0, p1x1 + · · ·+ pnxn ≤ y}.

238

Page 258: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.1. THE CONSTRAINED OPTIMIZATION PROBLEM 239

Another example is the set of production plans (x1, . . . , xn, y) that are feasible froma production technology that is described by a production function g. This usuallyis called a technology set,

T (g) = {(x1, . . . , xn, y) | x1 ≥ 0, . . . , xn ≥ 0, 0 ≤ y ≤ g(x1, . . . , xn)}.

The production function g : �n+ �→ �+ specifies the maximal output level y that the

firm can obtain from an input vector (x1, . . . , xn), where xi is the quantity used ofinput i. In this chapter it will always be presumed that C is a set of real numbers orvectors of real numbers, C ⊆ �n, but understand that this does not have to be so. Inprinciple C can contain any elements that a decisionmaker may choose between; e.gcars, books, songs or jobs.

The function f ranks feasible choices by desirability. Since (8.1) is a maximizationproblem, a feasible choice x′ ∈ C is strictly preferred to another feasible choice x′′ ∈ Cif and only if f(x′) > f(x′′). Similarly, f(x′) = f(x′′) if and only if x′ and x′′ areequally preferred, in which case we say that the decisionmaker is indifferent betweenx′ and x′′. If f(x′) ≥ f(x′′), then either x′ and x′′ are equally preferred or x′ is strictlypreferred to x′′. We state this more briefly by saying that x′ is weakly preferred to x′′.The function f thus expresses the decisionmaker’s objective, which is to select fromthe set C of available choices one that is as preferred as is possible, in the sense thatthe value of f is made as large as possible. For this reason f is typically calledproblem (8.1)’s objective function.

In case you are wondering if we have to consider minimization problems separatelyfrom maximization problems, the answer is “No.” The reason is easy to see. Considerthe minimization problem

minx∈C

h(x), (8.2)

and now write f(x) ≡ −h(x). Presto! You have the maximization problem (8.1)because any x that minimizes h(x) must maximize −h(x). Any minimization problemcan be written as a maximization problem and vice versa.

Let’s be clear that there are plenty of constrained optimization problems that donot have solutions. For example, can you achieve a maximum for f(x) = x on the setC = {x | x ≥ 0}? Of course not. Yet there is a unique solution to the similar problemof achieving a maximum for f(x) = x on the set C = {x | x ≤ 0}. Right?

Sometimes the feasible set C is finite; i.e. C contains only finitely many elements,such as {1, 2, 3}. Any constrained optimization problem with a finite feasible setalways has a maximizing (and also a minimizing) solution because there must alwaysbe at least one member of that set that provides a highest (lowest) possible value for

Page 259: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

240 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

the objective function. For example, if C = {x1, x2, . . . , xn}, then there will be just nvalues f(x1), f(x2), . . . , f(xn) for the objective function and one of these must be atleast as big as the others. Whichever one, or more, of the x-values that provides themaximum achievable value for f is an optimal solution to the problem.

We talk of feasible solutions and of optimal solutions to a constrained optimizationproblem. The term “feasible solution” means the same as “feasible choice”; i.e. ifx ∈ C, then x is a feasible solution or, equivalently, a feasible choice. The term“optimal solution” can be thought of in two ways, global and local.

A globally optimal solution is a feasible choice that provides the highest value forf that is achievable anywhere in the feasible set C.

Definition 8.2 (Globally Optimal Solution). x∗ ∈ C is a globally optimal solutionto problem (8.1) if and only if f(x∗) ≥ f(x) for every x ∈ C.

Put simply, x∗ is a most-preferred feasible choice among all of the feasible choices.Sometimes a constrained optimization problem has just one optimal solution, in whichcase we say that the problem’s optimal solution is unique. For example, the problemmax f(x) = x, where x ∈ C = {1, 2, 3}, has the unique optimal solution x∗ = 3. Otherproblems have nonunique, ormultiple, optimal solutions. The problem max f(x) = x2,where x ∈ C = {−2,−1, 0, 1, 2}, has two globally optimal solutions, x∗ = −2 andx∗∗ = 2. The problem max f(x) = cos x, where x ∈ C = �, has infinitely manyglobally optimal solutions: x∗ = nπ for n = 0,±2,±4, . . ..

A locally optimal solution is a feasible choice that provides the highest value forf in at least a locality (i.e. a neighborhood) in the feasible set. A locally optimalsolution could also be globally optimal (a globally optimal solution must also be alocally optimal solution), but need not be so. Consider the objective function

f(x) =

⎧⎪⎪⎪⎨⎪⎪⎪⎩x , x ≤ 1

2− x , 1 < x ≤ 2

x− 2 , 2 < x < 4

6− x , x ≥ 4

and take the feasible set to be C = �. Then this function has two local maxima,x = 1 and x = 4. One of these, x = 4, is also the global maximum.

Definition 8.3 (Locally Optimal Solution). x′ ∈ C is a locally optimal solutionto problem (8.1) in a nonempty neighborhood N ⊆ C if and only if x′ ∈ N andf(x′) ≥ f(x) for every x ∈ N .

Page 260: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.2. THE TYPICAL PROBLEM 241

The earlier example of f(x) = cosx is interesting because, when the feasible setis the whole of the real line �, this function has infinitely many local maxima, eachof which is also a global maximum.

8.2 The Typical Problem

There are lots of special cases we could consider, but economists mostly ponder onlya few of them, so it is only to these cases that I suggest we turn our attention. Thecase of most interest has two crucial properties. We will call this case the “typicalproblem.”

Definition 8.4 (The Typical Problem). The typical constrained optimization prob-lem is

maxx∈C

f(x), (8.3)

where f : C �→ � is continuous on C, C = ∅, C ⊂ �n, and C is compact with respectto En.

What does all this mean? Let’s start with C. First of all, we are restrictingourselves to feasible choices that are real numbers or vectors of real numbers, so therewill be no problems such as choosing between colors or pieces of furniture. Second,since we are considering only subsets of real numbers, a set is compact (with respectto the Euclidean topology on En) if and only if it is both closed and bounded. Ifyou have already worked your way through Chapter 3, then this is old stuff. If youhave not, then it will suffice for now to know that a set is bounded if it is completelycontained in a sphere with only a finite radius, meaning that no part of the setcontinues on forever in some direction. And a nonempty set is closed (again, withrespect to the Euclidean topology on En) if and only if it contains all of the pointson its boundaries.

What does it mean to say that the objective function f is continuous? If youhave worked through Chapter 5, then you already know this. If not, then for nowunderstand that a function f that maps from some subset X of real numbers x tothe real line is continuous on X if there is never an instance of an extremely smallshift from one x ∈ X to another causing a sudden jump in the value of f . Putin other words, any very small movement from one x to another must cause onlya very small change to the value of f . (This is loose language, and I urge you toread Chapter 5 if you have not done so. Continuity of functions and other types ofmappings is a really important idea.) Let’s also note here that, while any differentiable

Page 261: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

242 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

function is continuous, there are lots of continuous functions that are not everywheredifferentiable. An example is

f(x) =

{2 ; x < 2

x ; 2 ≤ x.

This function is continuous everywhere on �, but it is not differentiable at x = 2.One of the nice features of the typical problem described above is that Weier-

strass’s Theorem for continuous functions assures us that the problem always pos-sesses a globally optimal (not necessarily unique) solution.

8.3 First-Order Necessary Conditions

Now let’s get to the main matter. What we seek is a complete description of theconditions that have to be true at an optimal solution to a constrained optimizationproblem. That is, we seek to discover statements that necessarily are true as a conse-quence of the information that some point x∗ is an optimal solution to problem (8.3).The information provided by these necessarily true statements is what we will use tosolve the problem.

Two types of conditions are necessarily true at a local maximum. These are first-order necessary conditions and second-order necessary conditions. In this chapterwe will concern ourselves only with first-order necessary conditions. Second-ordernecessary conditions are described in Chapter 10.

We will consider only problems of the following form.

maxx1,...,xn

f(x1, . . . , xn)

subject to g1(x1, . . . , xn) ≤ b1...

... (8.4)

gm(x1, . . . , xn) ≤ bm,

where all of the functions f, g1, . . . , gm are continuously differentiable. One immediateimplication of this assumption is that the functions g1, . . . , gm are continuous. This,in turn, ensures that the feasible set

C = {x ∈ �n | g1(x) ≤ b1, . . . , gm(x) ≤ bm} = ∩mj=1{x ∈ �n | gj(x) ≤ bj}

is a closed set. Why? Prove it.

Page 262: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.3. FIRST-ORDER NECESSARY CONDITIONS 243

Figure 8.1: x∗ is a local maximum for f on C.

Have a look at Figure 8.1. The figure displays a compact feasible set C and apoint x∗ ∈ C that solves problem (8.4). At x∗ there are two binding constraints, g1and g2 (there might also be other slack constraints, but they can be ignored – onlybinding constraints matter), meaning that g1(x

∗) = b1 and that g2(x∗) = b2. The

figure displays three gradient vectors (which should really be displayed with respectto the origin, not with respect to x∗ – see Section 2.14). ∇f(x∗) is the value atx∗ of the objective function’s gradient vector. This vector is orthogonal to smallperturbation vectors Δx between x∗ and nearby points x∗ + Δx in the contour set{x ∈ �n | f(x) = f(x∗)}. The curve {x ∈ �n | g1(x) = b1} is the level-b1 contourset for the first binding constraint function g1, so the value at x = x∗ of its gradientvector, ∇g1(x∗), is orthogonal to small perturbations Δx between x∗ and nearbypoints x∗+Δx in the contour set {x ∈ �n | g1(x) = b1}. Similarly, the value ∇g2(x∗)of the gradient vector of the second constraint g2 that binds at x = x∗ is a vector thatis orthogonal to small perturbations Δx between x∗ and nearby points x∗ + Δx inthe contour set {x ∈ �n | g2(x) = b2}. The important feature of the figure is that thevector ∇f(x∗) lies in the convex cone that is generated by the vectors ∇g1(x∗) and∇g2(x∗). Remember, from alternative (i) of Farkas’s Lemma, that this means that

Page 263: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

244 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

there must exist nonnegative numbers, call them λ1 and λ2, that are not both zerosuch that ∇f(x∗) = λ1∇g1(x∗) + λ2∇g2(x∗). As you will soon see, this statement isthe famous Karush-Kuhn-Tucker necessary condition for x∗ to be a local maximumfor f on C.

Figure 8.2: x∗ is not a local maximum for f on C.

Why is the condition necessary? Well, suppose the condition is not true, meaningthat the objective function’s gradient vector ∇f(x∗) does not lie in the convex conegenerated by ∇g1(x∗) and ∇g2(x∗). This case is displayed in Figure 8.2. Because∇f(x∗) is orthogonal to small perturbation vectors Δx between x∗ and nearby pointsx∗+Δx in the contour set {x ∈ �n | f(x) = f(x∗)}, the consequence of ∇f(x∗) lyingoutside the convex cone is that the f(x) = f(x∗) contour set “cuts inside” the feasibleset C. Consequently there are feasible points such as x = x′ for which f(x′) > f(x∗),contradicting that x∗ is a local maximum for f on C. This is really all there is to theKarush-Kuhn-Tucker Theorem, except for a detail called a “constraint qualification”that for now we will put aside. Albert Tucker and Harold Kuhn published theirtheorem in 1951 (see Kuhn and Tucker 1951) and for some time were regarded as theoriginators of the result. Only later did it become widely known that William Karushhad already established the result in a Master of Science thesis written in 1939 atthe University of Chicago (see Karush 1939) and, also, that a similar result, the Fritz

Page 264: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.3. FIRST-ORDER NECESSARY CONDITIONS 245

John Theorem, had been published in 1948.

Theorem 8.1 (Karush-Kuhn-Tucker Necessity Theorem). Consider the problem

maxx∈�n

f(x) subject to gj(x) ≤ bj for j = 1, . . . ,m,

where the objective function f and the constraint functions g1, . . . , gm are all contin-uously differentiable with respect to x, and the feasible set

C = {x ∈ �n | gj(x) ≤ bj ∀ j = 1, . . . ,m}

is not empty. Let x∗ be a locally optimal solution to the problem. Let k denote thenumber of constraints that bind at x = x∗. Let I(x∗) be the set of indices of theconstraints that bind at x = x∗.

If k = 0, then I(x∗) = ∅ and ∇f(x∗) = 0.If 1 ≤ k ≤ m and if C satisfies a constraint qualification at x∗ then there exist

nonnegative scalars λ∗1, . . . , λ∗m, not all zero, such that λ∗j ≥ 0 for all j ∈ I(x∗), λ∗j = 0

for all j /∈ I(x∗),

∇f(x∗) = λ∗1∇g1(x∗) + · · ·+ λ∗m∇gm(x∗) and (8.5)

λ∗j(bj − gj(x∗)) = 0 ∀ j = 1, . . . ,m. (8.6)

(8.5) is called the Karush-Kuhn-Tucker necessary condition. The m conditions(8.6) are the complementary slackness conditions. The proof of this famous theoremis but a simple application of Farkas’s Lemma. Before we get to the proof, let medraw to your attention that the theorem assumes something called a “constraintqualification” that we have not yet examined. In most cases it is not good practiceto state a theorem that contains an as-yet unexplained idea, but I’m going to do itthis time and leave explaining what is meant by the term constraint qualification tothe next section. I think that, overall, things will be clearer for you if I do this.

Proof. If k = 0, then none of the constraints bind at x = x∗, and so I(x∗) = ∅; i.e.x = x∗ is a point in the interior of C. Obviously, then, necessarily ∇f(x∗) = 0.

Suppose instead that 1 ≤ k ≤ m. Thus I(x∗) = ∅. Without loss of generality,label the binding constraints by j = 1, . . . , k, so I(x∗) = {1, . . . , k}. Since x∗ is alocally optimal solution to the problem, it must be true that f(x∗) is at least as largeas is the value of f at any nearby feasible choice x∗ +Δx; i.e. for small enough Δxit must be true that

f(x∗) ≥ f(x∗ +Δx) ∀ Δx such that gj(x∗ +Δx) ≤ gj(x

∗) = bj ∀ j = 1, . . . , k. (8.7)

Page 265: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

246 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

Suppose that (8.5) is false; i.e. there do not exist scalars λ1, . . . , λk, all nonnegativeand not all zero, such that

∇f(x∗) = λ∗1∇g1(x∗) + · · ·+ λ∗k∇gk(x∗). (8.8)

The feasible set C satisfies a constraint qualification, so Farkas’s Lemma may beapplied (an explanation of this mysterious assertion is provided in Section 8.6). (8.8)is alternative (i) in Farkas’s Lemma, with aj = −∇gj(x∗) and b = −∇f(x∗). Sincealternative (i) is false, it follows from alternative (ii) of the lemma that there exists aΔx such that

−∇g1(x∗)·Δx ≥ 0, . . . , −∇gk(x∗)·Δx ≥ 0, −∇f(x∗)·Δx < 0.

That is, there exists a Δx such that

∇g1(x∗)·Δx ≤ 0, . . . , ∇gk(x∗)·Δx ≤ 0, ∇f(x∗)·Δx > 0. (8.9)

Since each of f, g1, . . . , gk is continuously differentiable, for small enough Δx,

f(x∗ +Δx) = f(x∗) +∇f(x∗)·Δx and

gj(x∗ +Δx) = gj(x

∗) +∇gj(x∗)·Δx ∀ j = 1, . . . , k.

Therefore, for small enough Δx, (8.9) is the statement that

f(x∗ +Δx) > f(x∗), g1(x∗ +Δx) ≤ g1(x∗) = b1, . . . , gk(x

∗ +Δx) ≤ gk(x∗) = bk.

(8.10)(8.10) is a statement that there is a feasible choice x∗ +Δx that is near to x∗ and

provides a higher objective function value than does x∗. But then x∗ is not a localmaximum; (8.10) contradicts (8.7). Thus (8.8) must be true if x∗ is a locally optimalsolution. Now choose λ∗k+1 = · · · = λ∗m = 0 for j = k + 1, . . . ,m, and we have shownthat (8.5) necessarily is true if x∗ is a locally optimal solution.

(8.6) follows from the facts that bj − gj(x∗) = 0 for j = 1, . . . , k and λ∗j = 0 for

j = k + 1, . . . ,m.

A few remarks need to be made before we go any further. First, notice that nomention has been made of a Lagrange function (this will be examined later), so if youare under the impression that you need a Lagrange function either to write down orto solve the Karush-Kuhn-Tucker conditions, then get rid of that impression – it isfalse.

Second, the theorem does not list all of the necessary optimization conditions.There is also a second-order necessary condition that will be described in Chapter 10.

Page 266: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.4. WHAT IS A CONSTRAINT QUALIFICATION? 247

Third, there may be more than one solution (x∗1, . . . , x∗n, λ

∗1, . . . , λ

∗m) to the first-

order optimality conditions (8.5) and (8.6). These conditions are only necessarilysatisfied at any locally optimal solution. It is possible that other feasible choices,some perhaps locally optimal and some perhaps not, may also satisfy the first-orderconditions.

Fourth, the theorem is subject to the requirement that the problem’s feasible setC “satisfies a constraint qualification at x∗.” It means that the Karush-Kuhn-TuckerTheorem cannot be applied to every constrained optimization problem of the type(8.4). We will examine the idea of a constraint qualification in the next section.

Last, the proof of the theorem uses the information that the problem satisfiesa constraint qualification when applying Farkas’s Lemma. Why this is so I cannotexplain until we know what is a constraint qualification. So, for now, tuck this remarkaway, and I will explain it when we reach Section 8.6.

8.4 What Is a Constraint Qualification?

In the present context the word “qualification” means “restriction,” so any constraintqualification is a restriction on the properties of the problem’s constraint functionsg1, . . . , gm. Constraint qualifications have nothing to do with the problem’s objectivefunction f . Every one of the many constraint qualifications has the same purpose,which is to ensure that any locally optimal solution to problem (8.4) does not occur ata point in the feasible set C where the Karush-Kuhn-Tucker condition does not exist.Such points are called irregularities. More precisely, a feasible point x ∈ C is said tobe irregular if and only if the Karush-Kuhn-Tucker condition does not exist at thatpoint. Equivalently, a feasible point is regular if and only if the Karush-Kuhn-Tuckercondition does exist at that point. The most commonly discussed example (there areothers) of an irregularity is a cusp. What is a cusp and why does it matter? Have alook at Figure 8.3.

Once again there is displayed a compact feasible set C, but, as you can see, thepoint x∗ now lies at a “needle point” vertex on the boundary of C that is formed bythe tangency at x = x∗ of two or more feasible set boundaries. Such a point is calleda cusp. Suppose that x∗ is a local maximum for f on C. Why does it matter thatx∗ is a cusp? Look again at the figure and think about the values at x = x∗ of thegradient vectors ∇g1(x) and ∇g2(x) of the constraints that bind at x = x∗. Sincethe constraints’ contour sets {x ∈ �2 | g1(x) = b1} and {x ∈ �2 | g2(x) = b2} aretangential at the cusp, what do you know about the gradient vectors ∇g1(x∗) and

Page 267: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

248 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

Figure 8.3: The feasible set C has a cusp at x = x∗.

∇g2(x∗)? Think about it – this is the crucial point! Remember that the gradientvectors are orthogonal to small perturbation vectors Δx from x∗ to nearby pointsx∗ +Δx in the contour sets.

The consequence of the tangency at x∗ of the curves g1(x) = b1 and g2(x) = b2 isthat ∇g1(x∗) and ∇g2(x∗) are negatively collinear ; i.e. the vectors point in exactlyopposite directions. Here is the next important question for you. What is the convexcone that is generated by the negatively collinear vectors∇g1(x∗) and∇g2(x∗)? Don’tproceed to the next paragraph until you have done your best to answer.

The convex cone is the hyperplane containing both vectors, shown as a dottedline in Figure 8.3 (remember that this hyperplane actually passes through the ori-gin). The objective function’s gradient vector ∇f(x) evaluated at x∗ does not liein the hyperplane. Consequently there do not exist any numbers λ1 and λ2 suchthat ∇f(x∗) = λ1∇g1(x∗) + λ2∇g2(x∗). The Karush-Kuhn-Tucker first-order neces-sary condition (8.5) may not be true at a locally optimal solution that is a cusp (orany other type of irregular point) in the feasible set of the constrained optimizationproblem.

This raises a troubling question: “For what particular types of constrained opti-

Page 268: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.5. FRITZ JOHN’S THEOREM 249

mization problems can we be sure that the Karush-Kuhn-Tucker necessary conditionwill exist at the locally optimal solutions to the problems?” The conditions thatidentify these types of problems are the constraint qualifications – that is their solepurpose.

A cusp is not the only type of irregularity that can occur. Here is an example (dueto Slater) of another. Consider the simple problem of maximizing f(x) = x subject tothe constraint g(x) = x2 ≤ 0 for only real values of x. Then the feasible set consistsof just one point, x = 0, which must therefore be the optimal solution. Evaluatedat x = 0, the gradients of f and g are f ′(0) = 1 and g′(0) = 0. Clearly there isno value for λ such that f ′(0) = λg′(0), so the Karush-Kuhn-Tucker condition fails.There is only one constraint in this problem, so there is no cusp in the feasible set.What’s going on here? Why does the Karush-Kuhn-Tucker condition fail? Beforewe answer this question, suppose we alter the constraint to g(x) = x2 ≤ ε for anyε > 0 that you wish. The feasible set becomes the interval (−√

ε,+√ε ), not just a

single point, and the optimal solution becomes x = +√ε. Evaluated at x = +

√ε, the

gradients of f and g are f ′(+√ε ) = 1 and g′(+

√ε ) = +2

√ε, so there is a positive

value, λ = 1/2√ε, satisfying f ′(+

√ε ) = λg′(+

√ε ). Adding some extra elements

to the feasible set seems to have made the irregularity go away. What do we makeof all this? Look back at the proof of the Karush-Kuhn-Tucker Necessity Theorem.Particularly, note that in the proof we used as our notion of local optimality that thevalue of the objective function f at x = x∗ had to be at least as large as the valueof f at any nearby feasible value x = x∗ + Δx. Thus we assumed the existence ofneighboring feasible values. The theorem does not apply to a case where there are nosuch neighbors, as in Slater’s example.

8.5 Fritz John’s Theorem

This raises a few practical problems. First, how would we discover if a problem’sfeasible set contains an irregular point? If the problem has only a few variables andonly a few constraints, then it is probably not too hard to find out. But looking fora possible irregularity in the boundaries of the feasible set of a problem with manyvariables and many constraints requires a search that is neither simple nor rapid.Second, an irregular point does not present a difficulty unless it is also a locallyoptimal solution to the problem. Now we are in a nasty situation. If the optimalsolution lies at an irregularity, then we cannot use the Karush-Kuhn-Tucker NecessityTheorem to solve for the value of the optimal solution, but without using the theorem

Page 269: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

250 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

locating the optimal solution will require an awful lot of hit-and-miss calculations.What are we to do? Perhaps we need a different tool than the Karush-Kuhn-TuckerTheorem? Indeed we do, and fortunately for us the tool exists. It is Fritz John’sTheorem (see John 1948), a result that applies even if the problem’s optimal solutionis at an irregular point. It is one of the most important results available on solvingdifferentiable constrained optimization problems. Let’s have a look at it.

Theorem 8.2 (Fritz John’s Necessity Theorem). Consider the inequality constrainedproblem

maxx∈�n

f(x) subject to gj(x) ≤ bj for j = 1, . . . ,m,

where the objective function f and the constraint functions g1, . . . , gm are all contin-uously differentiable with respect to x, and the feasible set

C = {x ∈ �n | gj(x) ≤ bj ∀ j = 1, . . . ,m}

is not empty. Let x∗ be a locally optimal solution to the problem. Let k denote thenumber of constraints that bind at x = x∗. Let I(x∗) be the set of indices of theconstraints that bind at x = x∗.

If k = 0, then I(x∗) = ∅ and ∇f(x∗) = 0.If 1 ≤ k ≤ m, then I(x∗) = ∅ and there exist nonnegative scalars λ∗0, λ

∗1, . . . , λ

∗m,

not all zero, such that λ∗0 ≥ 0, λ∗j ≥ 0 for all j ∈ I(x∗), λ∗j = 0 for all j /∈ I(x∗),

λ∗0∇f(x∗) = λ∗1∇g1(x∗) + · · ·+ λ∗m∇gm(x∗) and (8.11)

λ∗j(bj − gj(x∗)) = 0 ∀ j = 1, . . . ,m. (8.12)

This is a really good moment for you to reread the Karush-Kuhn-Tucker NecessityTheorem and to compare it word-by-word to Fritz John’s Theorem. Go ahead. I’llwait until you are done.

I’m sure that you have noticed that the m conditions (8.12) are the same comple-mentary slackness conditions (8.6) described in the Karush-Kuhn-Tucker Theorem.(8.11) is the Fritz John necessary condition and it is the same as the Karush-Kuhn-Tucker necessary condition (8.5) except that there is an extra multiplier λ∗0 in frontof the gradient of f at x∗. Indeed, if λ∗0 = 1, then Fritz John’s necessary condition isexactly the Karush-Kuhn-Tucker necessary condition. Notice that none of the com-plementary slackness conditions involves λ∗0. Also notice that Fritz John’s Theoremdoes not require that the problem’s feasible set C satisfies a constraint qualification.What this means is that Fritz John’s Theorem can be applied to problems in whicha local maximum x∗ is an irregular point, while the Karush-Kuhn-Tucker Theorem

Page 270: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.5. FRITZ JOHN’S THEOREM 251

cannot. It is hard to see what all this is about and why it matters without trying acouple of examples, so let’s do that now.

We will start with a simple problem with a feasible set that does not contain acusp or any other type of irregular point:

maxx1, x2

5x1 + x2 subject to g1(x1, x2) = −x1 ≤ 0,

g2(x1, x2) = −x2 ≤ 0 and g3(x1, x2) = x1 + 2x2 ≤ 2.

Figure 8.4: The optimal solution x∗ = (2, 0) is a regular point.

Figure 8.4 displays the problem. The optimal solution is (x∗1, x∗2) = (2, 0). Con-

straint g1 is slack at (2, 0) and constraints g2 and g3 are binding at (2, 0), so I(x∗) ={2, 3}. At (2, 0) the values of the gradients of f , g2 and g3 are

∇f(2, 0) = (5, 1), ∇g2(2, 0) = (0,−1) and ∇g3(2, 0) = (1, 2).

Page 271: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

252 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

Fritz John’s necessary condition (8.11) is thus

λ0 × 5 = λ2 × 0 + λ3 × 1

and λ0 × 1 = λ2 × (−1) + λ3 × 2,

resulting in the infinite collection of solutions λ2 = 9λ0 and λ3 = 5λ0 in which λ0 cantake any positive value (positive, not zero, because we do not want all of λ0, λ1, λ2,and λ3 to be zero). The Karush-Kuhn-Tucker necessary condition (8.5) is

5 = λ2 × 0 + λ3 × 1

and 1 = λ2 × (−1) + λ3 × 2,

resulting in the unique solution λ2 = 9 and λ3 = 5. This is the particular FritzJohn solution that results from choosing λ0 = 1. Indeed, so long as the problem’soptimal solution is a regular point, then we might as well choose λ0 = 1 and applythe Karush-Kuhn-Tucker Theorem. But now consider a similar problem in which theoptimal solution is an irregular point, a cusp, in the feasible set:

maxx1,x2

5x1 + x2 subject to g1(x1, x2) = −x1 ≤ 0,

g2(x1, x2) = −x2 ≤ 0 and g3(x1, x2) = x2 + (x1 − 2)3 ≤ 0.

Figure 8.5 displays the problem. The optimal solution is again (x∗1, x∗2) = (2, 0).

Confirm for yourself that this point is a cusp in the problem’s feasible set. Constraintg1 is slack and constraints g2 and g3 bind at (2, 0), so I(x∗) = {2, 3}. At (2, 0) thevalues of the gradients of f , g2 and g3 are

∇f(2, 0) = (5, 1), ∇g2(2, 0) = (0,−1), and ∇g3(2, 0) = (0, 1).

The Fritz John necessary condition (8.11) is

λ0 × 5 = λ2 × 0 + λ3 × 0

and λ0 × 1 = λ2 × (−1) + λ3 × 1.

There is an infinity of solutions to these equations, all of the form λ0 = 0 and λ2 = λ3.The two important things to notice here are that, first, there is at least one solutionto the Fritz John conditions and, second, for every solution λ0 = 0. The Karush-Kuhn-Tucker necessary condition (8.5) is

5 = λ2 × 0 + λ3 × 0

and 1 = λ2 × (−1) + λ3 × 1.

Page 272: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.5. FRITZ JOHN’S THEOREM 253

Figure 8.5: The feasible set has a cusp at the optimal solution x∗ = (2, 0).

These are equations with no solution. This demonstrates the extra generality ofFritz John’s necessary optimality condition. It can be used whether the problem’sfeasible set has an irregular point or not. If the problem’s optimal solution lies atan irregularity, then Fritz John’s necessary condition reports this by showing thatnecessarily λ0 = 0. But what does λ0 = 0 mean? Well, what is the whole idea of theKarush-Kuhn-Tucker necessary condition? It is that, at any regular optimal solutionto a problem, there must exist a non-zero and nonnegative vector λ = (λ1, . . . , λm),such that ∇f(x∗) =∑m

j=1 λj∇g(x∗). Only if this is so does the value of the objectivefunction’s gradient vector at x∗, ∇f(x∗), lie inside the convex cone defined by thevalues ∇gj(x∗) of the gradient vectors of the constraints j ∈ I(x∗) that bind at x∗.When there is a cusp at x∗, this convex cone typically does not contain ∇f(x∗), sothe solutions to λ0∇f(x∗) =

∑mj=1 λj∇g(x∗) all require that λ0 = 0. It is, in effect,

a statement that ∇f(x∗) cannot be written as a nonnegative and nonzero linearcombination of any of the ∇gj(x∗) for j ∈ I(x∗). Or, if you prefer, only by multiplying∇f(x∗) by zero to create the zero vector can we create a vector λ0∇f(x∗) that iscontained in the hyperplane through the origin that is the convex cone generated by

Page 273: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

254 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

the negatively collinear gradient vectors of the constraints that bind at a cusp.But why does λ∗0 = 0 tell us that there is an irregularity at x = x? Let’s think of

the case in which two constraints, say g1 and g2, are binding and form a cusp at x.If λ∗0 = 0, then Fritz John’s necessary condition is

λ∗0∇f(x∗) = 0 = λ∗1∇g1(x∗) + λ∗2∇g2(x∗),which tells us that

∇g1(x∗) = −λ∗2

λ∗1∇g2(x∗).

So λ∗0 = 0 only if the vectors ∇g1(x∗) and ∇g2(x∗) are collinear and point in opposite(because −λ∗2/λ∗1 ≤ 0) directions. Consequently the two vectors ∇g1(x∗) and ∇g2(x∗)generate a hyperplane when x = x is a cusp.

Now let’s think of Slater’s example. Why is it that necessarily λ∗0 = 0 in that case?In Slater’s example the feasible set contains only the point x∗ = 0 and, at that point,∇f(0) = 1 and ∇g(0) = 0. The only way to satisfy Fritz John’s necessary conditionthat there exist scalars λ∗0 and λ∗1, both nonnegative and not both zero, satisfyingλ∗0∇f(0) = λ∗1∇g(0) is with λ∗0 = 0. λ∗0 = 0 says that ∇f(0) is not a positive multipleof ∇g(0), which is, for Slater’s example, what the Karush-Kuhn-Tucker condition(impossibly) attempts to demand.

The Karush-Kuhn-Tucker necessary condition is the special case of the Fritz Johnnecessary condition in which λ0 = 1, so it would be entirely reasonable for you toconjecture that, since we used Farkas’s Lemma to prove the Karush-Kuhn-TuckerTheorem, we will again use Farkas’s Lemma to prove Fritz John’s Theorem. Thisis not so. To prove Fritz John’s Theorem we will instead use Gordan’s Lemma. Itis probably a good idea to take a few minutes to review Gordan’s Lemma beforeexamining the following proof.

Proof of Fritz John’s Necessity Theorem. If k = 0, then none of the constraintsbind at x = x∗; i.e. I(x∗) = ∅ and x = x∗ is interior to C. Obviously, then, necessarily∇f(x∗) = 0.

Suppose instead that 1 ≤ k ≤ m. Without loss of generality, label the bindingconstraints by j = 1, . . . , k. Then I(x∗) = {1, . . . , k}. Suppose that there do not existλ∗0, λ

∗1, . . . , λ

∗k, all nonnegative and not all zero, such that

λ∗0∇f(x∗) + λ∗1(−∇g1(x∗)) + · · ·+ λ∗k(−∇gk(x∗)) = 0. (8.13)

Then alternative (i) of Gordan’s Lemma is false and so, from alternative (ii) of thelemma, there exists Δx such that

∇f(x∗)·Δx > 0, −∇g1(x∗)·Δx > 0, . . . , −∇gk(x∗)·Δx > 0. (8.14)

Page 274: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.6. CONSTRAINT QUALIFICATIONS AND FARKAS’S LEMMA 255

That is,∇f(x∗)·Δx > 0, ∇g1(x∗)·Δx < 0, . . . , ∇gk(x∗)·Δx < 0. (8.15)

This is a statement that, for small enough Δx, there exists a feasible point x∗ +Δxnear to x∗ and in the interior of the feasible set for which f(x∗ +Δx) > f(x∗). Thiscontradicts that x∗ is a locally optimal solution. Therefore, because x∗ is a locallyoptimal solution, necessarily there exist λ∗0, λ

∗1, . . . , λ

∗k, all nonnegative and not all

zero, such thatλ∗0∇f(x∗) = λ∗1∇g1(x∗) + · · ·+ λ∗k∇gk(x∗). (8.16)

Now choose λ∗k+1 = · · · = λ∗m = 0, and both (8.11) and (8.12) are established.

Let’s notice a detail in the proof that is easily overlooked. Notice that the in-equalities ∇gj(x∗)·Δx < 0 in (8.15) are strict, not weak. So the proof relies upon theexistence of feasible points x∗ +Δx near to x∗ that are in the interior of the feasibleset; i.e. nearby points x∗ + Δx for which gj(x

∗ + Δx) = gj(x∗) + ∇gj(x∗)·Δx =

bj +∇gj(x∗)·Δx < bj (note: < bj, not ≤ bj) because ∇gj(x∗)·Δx < 0. For this to bepossible, the feasible set must locally have a nonempty interior, requiring that noneof the constraints that bind at x∗ is an equality constraint. That is why the titleof John’s paper is Extremum Problems with Inequalities as Subsidiary Conditions(emphasis added). The Fritz John necessary condition does not apply to constrainedoptimization problems with even one equality constraint that binds at a locally opti-mal solution.

The Karush-Kuhn-Tucker Theorem is more general than Fritz John’s Theoremin that the Karush-Kuhn-Tucker Theorem applies to problems with any mixture ofequality and inequality constraints, but is less general than Fritz John’s Theorem inthat it applies only to problems with feasible sets satisfying a constraint qualification.This is because the proof of the Karush-Kuhn-Tucker Theorem uses Farkas’s Lemma,which does not imply the Karush-Kuhn-Tucker first-order necessary condition at afeasible solution unless a constraint qualification is satisfied at that point. Why thisis so is explained in the next section.

8.6 Constraint Qualifications and Farkas’s Lemma

Now it is time to clear up a point that I deferred when explaining the proof of theKarush-Kuhn-Tucker Necessity Theorem. Do you recollect that, in the proof, I said,“The feasible set C satisfies a constraint qualification so Farkas’s Lemma may beapplied.”? Look again at the statement of Farkas’s Lemma so that you are sure ofwhat it says. In that statement do you see any caveat that the lemma can be used

Page 275: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

256 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

only if a constraint qualification is satisfied? No, you do not. So why insist uponthere being a constraint qualification satisfied before making use of the lemma?

The difficulty is not with Farkas’s Lemma. It is with the constrained optimizationproblem. The proof of the Karush-Kuhn-Tucker Theorem aims to show that, if x = x∗

is a maximum for f in a neighborhood of feasible points around x∗, then necessarilythe Karush-Kuhn-Tucker condition (8.5) is satisfied. But, as we have seen, this is notalways true. Particularly, if x = x∗ is some sort of irregular point in the feasible setC then the condition will not be satisfied. What does this have to do with Farkas’sLemma?

Figure 8.6: The feasible set has a cusp at the optimal solution. Farkas’s secondalternative is true.

Let’s suppose that the optimal solution x∗ is a cusp. Look at Figure 8.6. Be-cause the gradients of the binding constraints are negatively collinear, the set ofperturbations Δx from x∗, such that ∇g1(x∗)·Δx ≤ 0 and ∇g2(x∗)·Δx ≤ 0, is

Page 276: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.7. PARTICULAR CONSTRAINT QUALIFICATIONS 257

H = {Δx | ∇g1(x∗)·Δx = 0 and ∇g1(x∗)·Δx = 0}. This set is a hyperplane throughthe origin. The perturbations Δx in the hyperplane create a straight line locus ofpoints x∗ +Δx that is displayed as a dashed line in the figure. There are points suchas x∗+Δx′ in the locus that are in the feasible set, as well as points such as x∗+Δx′′

that are outside the feasible set. Notice that

∇g1(x∗)·Δx′′ = 0, ∇g2(x∗)·Δx′′ = 0 and ∇f(x∗)·Δx′′ > 0,

so we have found at least one value for Δx that satisfies alternative (ii) of Farkas’sLemma (with a1 = −∇g1(x∗), a2 = −∇g2(x∗) and b = −∇f(x∗)). Hence alterna-tive (i) is false, and it is alternative (i) upon which we rely to establish the Karush-Kuhn-Tucker necessary optimality condition. Thus, to establish the Karush-Kuhn-Tucker result, we must confine ourselves to constrained optimization problems inwhich alternative (ii) is false. These are exactly the problems wherein the optimalsolution is a regular point. One way to do this is to consider only problems in whichthe feasible set of the problem never has a locally maximal solution at an irregularpoint. The constraint qualifications tell us which are some of these problems.

8.7 Particular Constraint Qualifications

There are many constraint qualifications. This section presents just a representativefew of them. So long as you understand what a constraint qualification does, you cansafely skip this section. But there is some merit in reading at least the first few ofthe qualifications explained below since they are in common use by economists.

Slater’s Constraint Qualification. The feasible set C of the constrained optimiza-tion problem is compact with respect to En, is convex, and has a nonempty interior.

The feasible set is thus required to be as is displayed, for example, in Figures 8.1and 8.4, and not as is displayed in Figure 8.6. The nonempty interior requirementmeans that the feasible set can be neither empty nor a singleton (thus avoiding thedifficulty demonstrated in Slater’s example). The convexity of the set together withits nonempty interior means that there cannot be a cusp at any point in the set.Karlin’s Constraint Qualification, stated below, is equivalent to Slater’s, even thoughit might not seem so at a first glance.

Karlin’s Constraint Qualification. For any p ∈ �m+ with p = 0 and ‖p‖E < ∞

there exists an x′ ∈ C, such that p·(b− g(x′)) =∑m

j=1 pj(bj − gj(x′)) > 0.

Page 277: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

258 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

Figure 8.7: The Slater/Karlin Constraint Qualification.

It is not hard to see why Slater’s and Karlin’s statements are equivalent. The ideais illustrated in Figure 8.7 and is as follows. Take any point x on the boundary of thefeasible set C. Then gj(x) = bj for at least one j ∈ {1, . . . , n} and gk(x) ≤ bk for allk ∈ {1, . . . , n}, k = j. There will be a normal p for which the hyperplane p·x = p·xis tangential to C at x. Since p ∈ �m

+ and p = 0m, every one of p1, . . . , pm ≥ 0 withat least one strictly positive. Hence p·(b − g(x′)) ≥ 0. Karlin’s requirement is thatthere must be another point x′ that is in the interior of C, for then gj(x

′) < bj forevery j = 1, . . . ,m, causing p·(b− g(x′)) > 0; i.e. the interior of C is not empty, as isrequired by Slater’s Qualification. Why does Karlin’s Qualification ensure that C isa convex set? It says that all of the interior points of C lie in the lower half-spaces ofhyperplanes that are tangential to (supporting to) the feasible set C. In other words,the interior of C must be the intersection of all such possible lower half-spaces. Half-spaces are convex sets, so their intersection is a convex set, Slater’s other requirement,as you can see from Figure 8.7.

Since it seems that all we really need to do is insist that the gradients of theconstraints that bind at x = x are linearly independent, we can construct a constraintqualification by demanding exactly that. This is the Rank Constraint Qualification.

Rank Constraint Qualification. If the number of constraints that bind at x

Page 278: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.7. PARTICULAR CONSTRAINT QUALIFICATIONS 259

is k ≥ 1, then the rank of the Jacobian matrix consisting of the values at x of thegradients of the constraints that bind at x is k.

The usual shorthand notation for this Jacobian is

J(x) = (∇gj(x))j∈I(x) .Think of a problem with, say, five variables x1, . . . , x5 and four constraints labeled

by j = 1, 2, 3, 4. Suppose that the fourth constraint is slack at x = x. Then the indexset for the binding constraints is I(x) = {1, 2, 3} and the Jacobian matrix is

J(x) =

⎛⎝∇g1(x)∇g2(x)∇g3(x)

⎞⎠∣∣∣∣

x=x

=

⎛⎜⎜⎜⎜⎜⎜⎝

∂g1(x)

∂x1

∂g1(x)

∂x2

∂g1(x)

∂x3

∂g1(x)

∂x4

∂g1(x)

∂x5∂g2(x)

∂x1

∂g2(x)

∂x2

∂g2(x)

∂x3

∂g2(x)

∂x4

∂g2(x)

∂x5∂g3(x)

∂x1

∂g3(x)

∂x2

∂g3(x)

∂x3

∂g3(x)

∂x4

∂g3(x)

∂x5

⎞⎟⎟⎟⎟⎟⎟⎠∣∣∣∣∣x=x

,

where every derivative is evaluated at x = x. The rank of this matrix is at most 3.Suppose that x is a cusp. Then at least two of the gradient vectors ∇gj(x) forj = 1, 2, 3 are collinear, and so are linearly dependent, so the rank of J(x) is at most2, not 3. Conversely, suppose the rank of J(x) = 3. Then all of the gradient vectorsare linearly independent. Hence x cannot be a cusp. Look back at Figure 8.6, wherex = x∗ is a cusp, and observe that

I(x∗) = {1, 2}, J(x∗) =

⎛⎜⎜⎝∂g1(x)

∂x1

∂g1(x)

∂x2∂g2(x)

∂x1

∂g2(x)

∂x2

⎞⎟⎟⎠∣∣∣∣∣x=x∗

and Rank J(x∗) = 1 < 2.

I wish I could tell you that every constraint qualification is as easy to comprehendas are the above three, but, sad to say, some of them take quite a bit of puzzling over.Even more unfortunately, there is such a constraint qualification that is too importantto be ignored. This is the Kuhn-Tucker Constraint Qualification, usually abbreviatedto KTCQ. To make sense of it, we need to understand the terms constrained directionand constrained path. Then I will state the KTCQ and, with the help of a picture ortwo, we will figure out what it says and how it ensures that any feasible solution thatsatisfies it is a regular point.

Definition 8.5 (Constrained Direction). Let x′ ∈ C. Let I(x′) denote the set ofindices of the constraints g1(x) ≤ b1, . . . , gm(x) ≤ bm that bind at x = x′. A vectorΔx ∈ �n is a constrained direction from x′ if ∇gj(x′)·Δx ≤ 0 ∀ j ∈ I(x′).

Page 279: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

260 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

The set of constrained directions Δx from x′ is a closed convex cone that we willdenote by

CD(x′) = {Δx ∈ �n | ∇gj(x′)·Δx ≤ 0 ∀ j ∈ I(x′)}.

Figure 8.8: The affine closed convex cone x′ + CD(x′) for the regular point x = x′.

The shaded region of Figure 8.8 displays the points x′ + Δx created by such acone. A convex cone of constrained directions does not have to create points x′ +Δxonly in the problem’s feasible set C, as is illustrated in Figure 8.8. Note that theset of points x′ + Δx “points inwards into C from x′ ” when x′ is a regular feasiblepoint. However, if x′ is an irregular point, a cusp perhaps, then the set of constraineddirections is still a closed cone but it is rather different in shape. Figure 8.9 displaysthe cone when x′ is a cusp. Notice, because this is important, that in this case thecone is a hyperplane containing points near to x′ that are in C and points near to x′

that are outside C.Do you remember Slater’s example, in which the feasible set is only a singleton?

What is the closed convex cone of constrained directions in such a case? Look againat Figure 8.8 and suppose this time that the two constraints are equalities instead ofweak inequalities; i.e. the constraints are g1(x1, x2) = b1 and g2(x1, x2) = b2. Thenthe feasible set is only the singleton set {x′}. What now is the set x′ + CD(x′)? I’ll

Page 280: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.7. PARTICULAR CONSTRAINT QUALIFICATIONS 261

Figure 8.9: The affine closed convex cone x′ + CD(x′) for the cusp x = x′.

give you a clue. Look again at Definition 8.5. Notice that the movements Δx do nothave to result in feasible points x′+Δx. So, where in Figure 8.8 is the set x′+CD(x′)?

The answer is that x′+CD(x′) is exactly where the figure shows it to be. Turningthe feasible set into a singleton does not alter the position of the cone because thepoints x′ +Δx do not have to be feasible.

Now it is time to turn to the second idea we need, the idea of a constrained path.

Definition 8.6 (Constrained Path). Let x′ ∈ C. Let Δx ∈ �n, Δx = 0. A con-strained path from x′ in the direction Δx is a function h : [0, 1) �→ �n that isfirst-order continuously differentiable and has the properties that

(i) h(0) = x′,(ii) h′(0) = αΔx where α > 0 is a scalar, and

(iii) h(t) ∈ C for all t ∈ [0, 1).

A constrained path from x′ in the direction Δx is a smooth (because it is differ-entiable) curve that starts at x = x′, is entirely contained in C, and is tangential at x′

to the vector Δx. For a given x′ and Δx, there can be many such paths. Figure 8.10displays two constrained paths, h1(t) and h2(t).

Page 281: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

262 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

Figure 8.10: h1 and h2 are constrained paths from x = x′. h3 is not.

Notice that, while x′ ∈ C, it is possible that some of the points x = x′ + αΔx arenot in C. You see an example of this in Figure 8.10, where x = x′+Δx1 /∈ C. However,the path h1(t) is entirely within C, converges to x = x′, and becomes tangential to theline from x′ to x′+Δx1 as it converges to x′. Thus the curve h1(t) is a constrained pathfrom x′ in the direction Δx1. Similarly, the path h2(t) is entirely within C, convergesto x = x′, and becomes tangential to the line from x′ to x′+Δx2 as it converges to x′.Thus the curve h2(t) is a constrained path from x′ in the direction Δx2. The curveh3(t) is not a constrained path because, as the curve approaches x′ in the directionof Δx3, the curve is forced outside of C in order to be tangential to the line from x′

to x′ + Δx3. Be clear that the value of a constrained path h at a given value of t isa vector in �n; i.e. h(t) = (h1(t), . . . , hn(t)) where each of h1(t), . . . , hn(t) is a realnumber. In Figure 8.10, for example, n = 2, h1(0) = x′ = (x′1, x

′2) and for 0 < t < 1,

Page 282: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.7. PARTICULAR CONSTRAINT QUALIFICATIONS 263

h1(t) = (h11(t), h12(t)) is a point (x1, x2) on the line h1(t). At last we are ready to

understand the Kuhn-Tucker Constraint Qualification. Here it is.

The Kuhn-Tucker Constraint Qualification (KTCQ). Let x′ ∈ C be a pointfor which at least one of the constraints g1(x) ≤ b1, . . . , gm(x) ≤ bm binds. LetI(x′) denote the set of indices of the constraints that bind at x = x′. Let Δx ∈ �n

be a constrained direction from x′. There exists a constrained path from x′ in thedirection Δx.

Figure 8.11: The Kuhn-Tucker Constraint Qualification is satisfied at x = x′.

Most people find this statement about as clear as mud when reading it for the firsttime. What it says, however, is not too hard to figure out by looking at two pictures.Start with Figure 8.11. As you can see there, x = x′ is a regular point. The set ofindices of the constraints that bind at x′ is I(x′) = {1, 2}. The closed convex coneCD(x′) containing all of the constrained directions from x′ added to x′ is the shadedarea x′+CD(x′). Take any point x′+Δx you like in x′+CD(x′), such as x′+Δx1 orx′ + Δx2, and there is always at least one constrained path from x′ in the directionof Δx. Thus, whenever x′ is a regular feasible point the KTCQ is satisfied.

Now have a look at Figure 8.12, where the point x′ is a cusp. The closed convexcone of constrained directions is the hyperplane CD(x′) = {Δx | ∇g1(x′)·Δx =

Page 283: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

264 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

0 and ∇g2(x′)·Δx = 0} through the origin. The set x′ + CD(x′) is the straight linepassing through x′ and containing points such as x′ +Δx and x′ +Δx.

Figure 8.12: The Kuhn-Tucker Constraint Qualification is not satisfied at x = x′.

If we select a constrained direction such as Δx for which x′ +Δx is in the “insideC” part of the line x′ + CD(x′), then we can always find at least one constrainedpath, such as h(t), from x′ in the direction Δx – see the figure. But if we selecta constrained direction such as Δx for which x′ + Δx is in the “outside C” part ofthe line x′ + CD(x′), then we cannot find even one constrained path from x′ in thedirection Δx – see the curve h(t) in the figure. Even though this smooth (remember,it is differentiable) curve starts inside C, it must eventually pass outside of C in orderto approach x′ from the direction Δx. Therefore the curve h(t) is not a constrainedpath. By insisting that there must be at least one constrained path from x′ for everyconstrained direction from x′, the KTCQ insists that x′ is not a cusp.

Page 284: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.8. REMARKS 265

Let’s again consider Slater’s example, in which the feasible set is a singleton. Howexactly is this prohibited by the KTCQ? Look again at Definition 8.6 and notice thatany constrained path must be entirely within the feasible set C = {x′}. Thereforeh(t) ≡ x′ (i.e. h is constant-valued) for all t ∈ [0, 1). Hence its derivative is thezero vector for all t, including at t = 0. Thus h′(0) = 0. Part (ii) of Definition 8.6insists that, for a constrained direction Δx = 0 from x′ and for α > 0, we can writeh′(0) = αΔx. This is impossible when h′(0) = 0, so the KTCQ demands also thatthe feasible set not be a singleton.

As I said earlier, the constraint qualifications explained in this section are buta few of many. I selected the ones described above because of their usefulness toeconomists, but it may happen that none of the qualifications described here meetsyour needs. Do not despair – there are many other qualifications and one of themmay be what you require.

8.8 Remarks

We have reached the end of our discussion of the first-order optimality conditions forconstrained optimization problems. A lot has been covered, and it is easy to losesight of the essential points when trying to deal with all of the details. So let’s take amoment to review what we have learned and to set the stage for what still awaits us.

The crucial idea is that, for most of the constrained optimization problems that areof interest to economists, a feasible point x∗ is locally optimal only if the value at x∗ ofthe gradient of the objective function is a vector that is contained in the convex conegenerated by the values at x∗ of the gradient vectors of the constraint functions thatbind at x∗. This is the meaning of the Karush-Kuhn-Tucker and Fritz John necessaryfirst-order optimality conditions. For each constraint, binding or not, another set ofconditions has to be true at a local optimum. These are the complementary slacknessconditions. Each necessary condition must individually be true at a local optimum;if any one of them is not true at x∗, then x∗ cannot be a local optimum.

We also learned that the Karush-Kuhn-Tucker Theorem and the Fritz John The-orem each applies to different classes of constrained optimization problems. Thegood news about the Karush-Kuhn-Tucker Theorem is that it applies to constrainedoptimization problems with any mixture of equality and/or inequality constraints.The bad news is that there may exist feasible, possibly optimal, points at which theKarush-Kuhn-Tucker necessary condition does not exist. We attempt to identify prob-lems that do not contain such irregular points by insisting that the problems satisfy

Page 285: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

266 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

at least one of the many constraint qualifications. The Karush-Kuhn-Tucker Theo-rem can be applied only to such problems. There is good and bad news about FritzJohn’s Theorem also. The bad news is that the theorem applies only to constrainedoptimization problems with only inequality constraints – no equality constraints areallowed. The good news is that, for such problems, the Fritz John necessary condi-tion is well defined at any feasible point, irregular or not. The two theorems thuscomplement each other. Neither is a special case of the other.

All that we have done in this chapter is subject to one obvious weakness. Thefirst-order local optimality conditions that we have discovered may be satisfied alsofor points that are not local maxima (local minima, for example), so we now need todiscover a method for distinguishing between local maxima and points that are notlocal maxima even though they too solve the necessary first-order optimality condi-tions. This is the central matter of Chapter 10. The ideas explained in Chapter 10make extensive use of Lagrange functions, so Chapter 9 explains these functions andsome of their valuable uses.

Page 286: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.9. PROBLEMS 267

8.9 Problems

Problem 8.1. Here is a constrained optimization problem:

maxx1, x2

f(x1, x2) = 4x1 + x2 subject to g1(x1, x2) = −2− 2x1 +(x1 − 20)3

500+ x2 ≤ 0,

g2(x1, x2) = 2x1 − x2 ≤ 0, g3(x1, x2) = −x1 ≤ 0, and g4(x1, x2) = −x2 ≤ 0.

(i) Try using the Karush-Kuhn-Tucker necessary condition to solve the problem.Hint: The globally optimal solution is x∗ = (x∗1, x

∗2) = (30, 60).

(ii) Try using the Fritz John necessary condition to solve the problem.(iii) What difference, if any, is there in the outcomes of the approaches used in

parts (i) and (ii)?(iv) Draw the feasible set. See for yourself that the point x∗ = (30, 60) is regular

and that the gradient ∇f(x∗) = (4, 1) of the objective function at x∗ = (30, 60) iscontained in the convex cone that is generated by the gradients ∇g1(x∗) = (−7/5, 1)and ∇g2(x∗) = (2,−1) of the constraints that bind at x∗ = (30, 60).

(v) Is the Slater/Karlin constraint qualification satisfied for the constrained op-timization problem’s feasible set? Is the rank constraint qualification satisfied at thefeasible solution (x1, x2) = (30, 60)?

Problem 8.2. The problem presented below is the same as problem 8.1 except thatthe constraint function g1 has been slightly changed:

maxx1, x2

f(x1, x2) = 4x1 + x2 subject to g1(x1, x2) = −2x1 +(x1 − 20)3

500+ x2 ≤ 0,

g2(x1, x2) = 2x1 − x2 ≤ 0, g3(x1, x2) = −x1 ≤ 0, and g4(x1, x2) = −x2 ≤ 0.

(i) Try using the Karush-Kuhn-Tucker necessary condition to solve the problem.Hint: The globally optimal solution is x∗ = (x∗1, x

∗2) = (20, 40).

(ii) Try using the Fritz John necessary condition to solve the problem.(iii) What difference, if any, is there in the outcomes of the approaches used in

parts (i) and (ii)?(iv) Draw the feasible set. See for yourself that the point x∗ = (20, 40) is irregular

and that the gradient ∇f(x∗) = (4, 1) of the objective function at x∗ = (20, 40) is notcontained in the convex cone that is generated by the gradients ∇g1(x∗) = (−2, 1)and ∇g2(x∗) = (2,−1) of the constraints that bind at x∗ = (20, 40).

(v) Is the Slater/Karlin constraint qualification satisfied for the constrained op-timization problem’s feasible set? Is the rank constraint qualification satisfied at thefeasible solution (x1, x2) = (20, 40)?

Page 287: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

268 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

Problem 8.3. Here is another constrained optimization problem:

maxx1, x2

f(x1, x2) = −4(x1 − 2)2 + x2 subject to g1(x1, x2) = (x1 − 2)3 + x2 ≤ 8,

g2(x1, x2) = (x1 − 2)3 − x2 ≤ −8, g3(x1, x2) = −x1 ≤ 0, and g4(x1, x2) = −x2 ≤ 0.

(i) Try using the Karush-Kuhn-Tucker necessary condition to solve the problem.Hint: The globally optimal solution is x∗ = (x∗1, x

∗2) = (2, 8). Draw the feasible set

and the contour set for the objective function that contains the point (2, 8).(ii) Why is it, in this problem, that the Karush-Kuhn-Tucker necessary condition

is well-defined at the global optimum even though the global optimum is a cusp? Addto your diagram of the feasible set the gradients, evaluated at (2, 8), of the objectivefunction and the two binding constraints. What do you see?

(iii) Try using the Fritz John necessary condition to solve the problem.

Problem 8.4. Consider the constrained optimization problem:

maxx∈C

f(x), where C = {x ∈ �n | gj(x) ≤ bj for j = 1, . . . ,m}.

f, g1, . . . , gm are all C1 functions defined on �n. x∗ denotes a locally optimal solutionto this problem.

(i) Is satisfaction of a constraint qualification at x∗ a condition that is necessaryfor this problem to have a locally optimal solution x∗? Explain.

(ii) Is satisfaction of a constraint qualification at x∗ a condition that is sufficientfor this problem to have a locally optimal solution x∗? Explain.

(iii) Is satisfaction of a constraint qualification at x∗ a condition that is necessaryfor the Karush-Kuhn-Tucker necessary condition to be well-defined at x∗? Explain.

(iv) Is satisfaction of a constraint qualification at x∗ a condition that is sufficientfor the Karush-Kuhn-Tucker necessary condition to be well-defined at x∗? Explain.

Problem 8.5. Consider the following constrained optimization problem:

maxx1, x2

f(x1, x2) = +√x1 + x2 subject to g1(x1, x2) = x21 + x22 ≤ 16,

g2(x1, x2) = x1 + x2 ≤ 5, g3(x1, x2) = −x1 ≤ 0, and g4(x1, x2) = −x2 ≤ 0.

Note: When evaluating f , use only the positive value of the square root of x1.(i) Draw a diagram that displays the feasible set

C = {(x1, x2) | x1 ≥ 0, x2 ≥ 0, x21 + x22 ≤ 16, x1 + x2 ≤ 5}.

Page 288: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.10. ANSWERS 269

(ii) What properties does C possess?(iii) Must this constrained optimization problem have an optimal solution?(iv) Add a typical contour set for the objective function to your diagram.(v) Use the Karush-Kuhn-Tucker necessary local maximization conditions to locatepossible local maxima for f on C.

Problem 8.6. Consider the following constrained optimization problem:

maxx1, x2

f(x1, x2) = x21 −33

4x1 + 21 + x2 subject to g1(x1, x2) = −√

x1 − x2 ≤ −6,

g2(x1, x2) = x1 + 4x2 ≤ 20, g3(x1, x2) = −x1 ≤ 0, and g4(x1, x2) = −x2 ≤ 0.

(i) Draw a diagram that displays the problem’s feasible set.(ii) The optimal solution to this problem is (x∗1, x

∗2) = (4, 4). Can you write down

the Karush-Kuhn-Tucker necessary optimality condition at this point? Explain.

Problem 8.7. Consider the following constrained optimization problem:

maxx1, x2

f(x1, x2) = x1 + x2 subject to g1(x1, x2) = (x1 − 4)2 + x2 ≤ 7,

and g2(x1, x2) = (x1 − 5)3 − 12x2 ≤ −60.

(i) Draw a diagram that displays the feasible set.(ii) Write down the Karush-Kuhn-Tucker necessary optimality condition togetherwith the complementary slackness conditions.(iii) Solve the problem.

8.10 Answers

Answer to Problem 8.1.(i) See Figure 8.13. At x∗ = (30, 60), constraints g3 and g4 are slack. Constraintsg1 and g2 bind at x∗ = (30, 60). The gradients of the objective function and theconstraint functions g1 and g2 are

∇f(x) = (4, 1), ∇g1(x) =(−2 +

3(x1 − 20)2

500, 1

), and ∇g2(x) = (2,−1).

Evaluated at x∗ = (30, 60), these gradients are

∇f(x∗) = (4, 1), ∇g1(x∗) =(− 7

5, 1

), and ∇g3(x∗) = (2,−1).

Page 289: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

270 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

The gradients ∇g1(x∗) and ∇g2(x∗) of the binding constraints are not collinear, so thefeasible point x∗ = (30, 60) is regular. Hence we may use the Karush-Kuhn-Tuckerfirst-order necessary condition and write that

(4, 1) = λ1

(− 7

5, 1

)+ λ2(2,−1) ⇒ λ∗1 = 10, λ∗2 = 9.

The values λ∗1 and λ∗2 are both nonnegative and at least one is strictly positive.

Figure 8.13: The feasible set, the optimal solution, and the gradient vectors forproblem 8.1.

(ii) Fritz John’s first-order necessary condition is that

λ0(4, 1) = λ1

(− 7

5, 1

)+ λ2(2,−1) ⇒ λ∗1 = 10λ0, λ

∗2 = 9λ0.

Clearly there is an infinity of positive values for λ0 that cause both λ∗1 and λ

∗2 also to

have positive values.

Page 290: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.10. ANSWERS 271

(iii) The Karush-Kuhn-Tucker approach insists that the objective function multiplierλ0 = 1. In the Fritz John approach, λ0 can be assigned any positive value, and theother multiplier values are then determined conditional upon the value assigned to λ0.

(iv) See Figure 8.13.

(v) The feasible set C has a nonempty interior but is not a convex set, so the S-later/Karlin constraint qualification is not satisfied by this problem’s feasible set.The Jacobian matrix formed by the values at (x1, x2) = (30, 60) of the gradient vec-tors of the binding constraints is

Jg(30, 60) =

(∇g1(x1, x2)∇g2(x1, x2)

) ∣∣∣∣(x1,x2)=(30,60)

=

(−7/5 12 −1

).

The row vectors of this matrix are linearly independent of each other. Equivalently,the rank of this matrix is 2, equal to the number of constraints that bind at (30, 60).So the rank constraint qualification is satisfied at (30, 60), establishing that (x1, x2) =(30, 60) is a regular feasible solution.

Answer to Problem 8.2.

(i) See Figure 8.14. At x∗ = (20, 40) constraints g3 and g4 are slack. Constraintsg1 and g2 bind at x∗ = (20, 40). The gradients of the objective function and theconstraint functions g1 and g2 are

∇f(x) = (4, 1), ∇g1(x) =(−2 +

3(x1 − 20)2

500, 1

), and ∇g2(x) = (2,−1).

Evaluated at x∗ = (20, 40), these gradients are

∇f(x∗) = (4, 1), ∇g1(x∗) = (−2, 1) , and ∇g3(x∗) = (2,−1).

The gradients∇g1(x∗) and∇g2(x∗) of the binding constraints are negatively collinear,so the feasible point x∗ = (20, 40) is irregular (a cusp, actually). Hence it may not bepossible to use the Karush-Kuhn-Tucker first-order necessary condition. If we try todo so, we get

(4, 1) = λ1 (−2, 1) + λ2(2,−1).

That is,

4 = −2λ1 + 2λ2 and 1 = λ1 − λ2.

These are inconsistent equations, so there no values for λ∗1 and λ∗2 satisfying theKarush-Kuhn-Tucker necessary condition.

Page 291: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

272 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

Figure 8.14: The feasible set, the optimal solution, and the gradient vectors forproblem 8.2.

(ii) Fritz John’s first-order necessary condition is that there are nonnegative scalarsλ0, λ1 and λ2, not all zero, such that

λ0(4, 1) = λ1 (−2, 1) + λ2(2,−1) ⇒ λ∗1 = 10λ0, λ∗2 = 9λ0.

That is,4λ0 = −2λ1 + 2λ2 and λ0 = λ1 − λ2.

The family of solutions is λ0 = 0 and λ1 = λ2 > 0 (because not all of λ0, λ1 and λ2are zero). That is, there is an infinity of solutions satisfying the Fritz John necessarycondition. In each of these solutions λ0 = 0.(iii) The Karush-Kuhn-Tucker necessary condition does not exist at x∗ = (20, 40).The condition assumes (a constraint qualification) that the gradient of the objectivefunction lies within the convex cone generated by the gradients of the constraints thatbind at x∗ = (20, 40). But this is not so. The Fritz John necessary condition pointsthis out by reporting that necessarily λ0 = 0.(iv) See Figure 8.14.

Page 292: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.10. ANSWERS 273

(v) The feasible set C has a nonempty interior but is not a convex set, so theSalter/Karlin constraint qualification is not satisfied by this problem’s feasible set.The Jacobian matrix formed by the values at (x1, x2) = (240, 0) of the gradient vectorsof the binding constraints is

Jg(30, 60) =

(∇g1(x1, x2)∇g2(x1, x2)

)∣∣∣∣(x1,x2)=(20,40)

=

(−2 12 −1

).

The row vectors of this matrix are linearly dependent; each is the negative of theother. Thus the rank of this matrix is 1 < 2, the number of constraints that bind at(20, 40); the rank constraint qualification is not satisfied at (x1, x2) = (20, 40).

Answer to Problem 8.3.(i) See Figure 8.15. At x∗ = (2, 8), constraints g3 and g4 are slack. Constraints g1and g2 bind at x∗ = (2, 8). The feasible set is displayed in Figure 8.15.

The gradients of the objective function and the constraint functions g1 and g2 are

∇f(x) = (−8(x1 − 2), 1), ∇g1(x) = (3(x1 − 2)2, 1), and ∇g2(x) = (3(x1 − 2)2,−1).

Evaluated at x∗ = (2, 8), these gradients are

∇f(x∗) = (0, 1), ∇g1(x∗) = (0, 1), and ∇g2(x∗) = (0,−1).

Notice that the gradients ∇g1(x∗) and ∇g2(x∗) of the binding constraints are nega-tively collinear. The Karush-Kuhn-Tucker first-order necessary condition is that thereare nonnegative values λ1 and λ2, not both zero, such that

(0, 1) = λ1(0, 1) + λ2(0,−1).

There are many solutions, all of the form (λ1, λ2) = (λ1, λ1−1). One such solution, forexample, is λ1 = 2 and λ2 = 1. These values for λ∗1 and λ∗2 are both nonnegative andat least one is strictly positive. Thus, even though the feasible (and globally optimal)point (2, 8) is a cusp, the Karush-Kuhn-Tucker necessary condition is satisfied. Forthis problem, then, the point (2, 8) is a regular point, even though it is a cusp. If thisseems contradictory, then remember that any constraint qualification is a conditionsufficient, but not necessary, for the Karush-Kuhn-Tucker necessary condition to besatisfied at a local optimum.(ii) The gradients (0, 1) and (0,−1) of the two constraints that bind at (2, 8) arenegatively collinear, so the convex cone they generate is the hyperplane that is the

Page 293: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

274 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

Figure 8.15: The feasible set, the optimal solution, and the gradient vectors forproblem 8.3.

set {(x1, x2) | x1 = 0}. In this particular problem, the value (0, 1) of the gradientvector of the objective function at (2, 8) happens to lie in this hyperplane. This iswhy the Karush-Kuhn-Tucker condition is true. The Karush-Kuhn-Tucker conditionmakes no sense only when the gradient of the objective function is not contained inthe convex cone that is generated by the gradient vectors of the binding constraints.

(iii) The Fritz John first-order necessary condition is that there exist nonnegativescalars λ0, λ1 and λ2, not all zero, such that

λ0(0, 1) = λ1(0, 1) + λ2(0,−1).

There is an infinity of nonnegative values for λ0, λ1, and λ2, not all zero, that satisfythe Fritz John condition. Each solution is of the form (λ0, λ1, λ2) = (λ0, λ1, λ1 −λ0), with λ0 ≥ 0 and λ1 ≥ λ0. An example is (λ0, λ1, λ2) = (1, 2, 1). Another is

Page 294: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.10. ANSWERS 275

(λ0, λ1, λ2) = (0, 1, 0). Notice that, even though (2, 8) is a cusp, it is not necessarythat λ0 = 0. Why? λ0 = 0 necessarily occurs when the gradient of the objectivefunction is not contained in the hyperplane generated by the convex cone of thegradients of the constraints that bind. In this problem, however, the gradient vector(0, 1) of the objective function at (2, 8) happens to be contained in the convex conegenerated by the gradient vectors of the constraints that bind. Consequently it is notnecessary that λ0 = 0.

Answer to Problem 8.4.(i) No. A constraint qualification makes no statement about the existence of anoptimal solution to the constrained optimization problem. The only thing that aconstraint qualification makes a statement about is the shape of the boundary of thefeasible set. Existence of an optimal solution to the constrained optimization problemis not an issue dealt with by any constraint qualification. The only difficulty is that,at an irregular feasible point, there is no guarantee that the Karush-Kuhn-Tuckerfirst-order necessary condition is defined.

Some constraint qualifications impose restrictions upon the entirety of the feasi-ble set. An example is the Slater/Karlin constraint qualification. Other constraintqualifications impose a restriction only upon a particular point in the feasible set. Anexample is the rank constraint qualification. So suppose that C contains an irregularpoint, but not at the maximizing solution to the constrained optimization problem.Then a constraint qualification such as Slater/Karlin will fail, while the rank qualifica-tion will be satisfied at the optimal solution to the constrained optimization problem,so the Karush-Kuhn-Tucker first-order necessary condition will be well-defined at thisoptimal solution and can be used to locate the solution.(ii) No. Consider the constrained optimization problem:

maxx1, x2

x1 + x2 subject to x1 ≥ 0 and x2 ≥ 0.

The feasible set is C = {(x1, x2) | x1 ≥ 0, x2 ≥ 0}. This set contains only regularpoints, and so satisfies each and every constraint qualification. Even so, there isno optimal solution to this constrained optimization problem because the objectivefunction is strictly increasing and the feasible set is unbounded above.(iii) No. Again consider problem 8.3. The constrained optimization problem thereinhas an optimal solution at a cusp. No constraint qualification applies at such apoint. Yet, at this optimal solution, the Karush-Kuhn-Tucker necessary first-ordercondition is well-defined. Again, the crucial point is that satisfaction at any locallyoptimal solution of any one of the constraint qualifications is a condition sufficient,

Page 295: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

276 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

but not necessary, for that locally optimal solution to be a regular point, a point atwhich the Karush-Kuhn-Tucker condition is well-defined.(iv) Yes. By definition, if a constraint qualification applies at a locally optimalsolution, then that point is regular and the Karush-Kuhn-Tucker first-order conditionis well-defined at that locally optimal solution.

Answers to Problem 8.5.(i) S1 = {(x2, x2) | x21 + x22 ≤ 16} is the disc centered at the origin with a radius of 4.S2 = {(x1, x2) | x1 + x2 ≤ 5} is the lower half-space in �2 of the hyperplane withnormal (1, 1) that contains the points (0, 5) and (5, 0). S3 = {(x1, x2) | x1 ≥ 0} is theupper half-space in �2 of the hyperplane with normal (1, 0) that includes the origin(0, 0). S4 = {(x1, x2) | x2 ≥ 0} is the upper half-space in �2 of the hyperplane withnormal (0, 1) that includes the origin (0, 0). The feasible set C = S1 ∩ S2 ∩ S3 ∩ S4 isthe shaded region displayed in the upper panel of Figure 8.16.(ii) Because g1(x1, x2) = x21 + x22 is continuous on �2, the set S1 is closed in E2.Similarly, S2, S3, and S4 are also closed in E2. Therefore C = S1 ∩ S2 ∩ S3 ∩ S4 isclosed in E2.

S1 is a strictly convex set. S2, S3, and S4 are weakly convex sets. ThereforeS1 ∩ S2 ∩ S3 ∩ S4 is a (weakly) convex set.

S2, S3, and S4 are not bounded sets, but S1 is bounded in E2. Therefore C =S1 ∩ S2 ∩ S3 ∩ S4 is bounded in E2.

Since C is both closed and bounded in E2, C is compact in E2.Finally, it is easy to see that C is not empty and has a nonempty interior.

(iii) Yes. The objective function is continuous over the feasible set, and the feasibleset is nonempty and is compact in E2. Weierstrass’s Theorem confirms that theremust exist at least one element in C that achieves a maximum for f on C.(iv) For level k = 0, the objective function’s contour set is the singleton containingonly the origin. For any level k > 0, the objective function’s level-k contour set is

S(k) = {(x1, x2) | +√x1 + x2 = k, x1 ≥ 0, x2 ≥ 0}.

This is the set described by the curve x2 = k −√x1 in �2

+. It is easy to see that theupper-contour set is always a weakly convex subset of �2

+. Therefore f is a quasi-concave function on �2

+. The lower panel of Figure 8.16 displays two of the objectivefunction’s contour sets.(v) The gradient vectors of the objective function f and of the constraint functionsg1 and g2 are

∇f(x1, x2) =(

1

2√x1, 1

), ∇g1(x1, x2) = (2x1, 2x2) , and ∇g2(x1, x2) = (1, 1).

Page 296: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.10. ANSWERS 277

Figure 8.16: Problem 8.5’s feasible set and two objective function contour sets.

Because the objective function is strictly increasing in both x1 and x2 the obviousplaces to first check for a local maximum are the two intersections of the bound-aries of the constraints g1 and g2. The solutions to x1 + x2 = 5 and x21 + x22 = 16are the points (x1, x2) =

((5−

√7 )/2, (5 +

√7 )/2

)≈ (1·177, 3·823) and (x1, x2) =(

(5 +√7 )/2, (5−

√7 )/2

)≈ (3·823, 1·177). The values at (1·177, 3·823) of the above

gradient vectors are

∇f(1·177, 3·823) = (0·461, 1), ∇g1(1·177, 3·823) = (2·354, 7·646),and ∇g2(1·177, 3·823) = (1, 1).

These gradient vectors are displayed in Figure 8.17. The gradient vector of the objec-

Page 297: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

278 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

tive function is contained in the interior of the convex cone generated by the gradientvectors of the two binding constraint functions, so there must exist two strictly posi-tive numbers λ1 and λ2, such that

(0·461, 1) = λ1(2·354, 7·646) + λ2(1, 1).

These numbers are λ1 = 0·102 and λ2 = 0·221. Thus the Karush-Kuhn-Tuckernecessary condition is satisfied at (x1, x2) = (1·177, 3·823), which might, therefore,be a local maximum for f on C.

Figure 8.17: The gradients of f , g1 and g2 at (x1, x2) = (1·177, 3·823) and at (x1, x2) =(3·823, 1·177).

Evaluated at (3·823, 1·177), the above gradient vectors are

∇f(3·823, 1·177) = (0·256, 1), ∇g1(3·823, 1·177) = (7·646, 2·354),and ∇g2(3·823, 1·177) = (1, 1).

Page 298: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.10. ANSWERS 279

These gradient vectors are also displayed in Figure 8.17. The gradient vector of theobjective function is not contained in the convex cone generated by the gradientvectors of the two constraint functions. That is, the Karush-Kuhn-Tucker necessitycondition is not satisfied at (x1, x2) = (3·823, 1·177). Necessarily, therefore, this pointis not a local maximum for f on C.

How do we know that (x1, x2) = (1·177, 3·823) is indeed a local maximum? Whyis it not a local minimum, for example? The objective function’s upper-contour setsare convex. The feasible set C is convex. Add to this that the objective functionf is strictly increasing and that its upper-contour set is locally strictly convex at(1·177, 3·823), and it follows immediately that (1·177, 3·823) is a local maximum andis the only local maximum for f on C. Thus (x1, x2) = (1·177, 3·823) is the globalmaximum for f on C.Answer to Problem 8.6.(i) See Figure 8.18. The feasible set is the singleton C = {(4, 4)}.

Figure 8.18: The singleton feasible set C and the gradient vectors for problem 8.6.

(ii) At (x1, x2) = (4, 4), constraints g1 and g2 bind, and constraints g3 and g4 are

Page 299: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

280 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

slack. The gradient vectors of the objective function and the binding constraints are

∇f(x1, x2) =(2x1 −

33

4, 1

), ∇g1(x1, x2) =

(− 1

2√x1, −1

), ∇g2(x1, x2) = (1, 4).

The values at (x1, x2) = (4, 4) of these gradient vectors are

∇f(4, 4) =(− 1

4, 1

), ∇g1(4, 4) =

(− 1

4, −1

), and ∇g2(4, 4) = (1, 4).

The gradient vectors of the binding constraints are negatively collinear, so the convexcone that they generate is the hyperplane with normal (−4, 1) that includes the origin.The objective function’s gradient vector (−1/4, 1) is not contained in this cone, so theKarush-Kuhn-Tucker necessary condition is not defined at the point (x1, x2) = (4, 4).If we nevertheless attempt to solve the Karush-Kuhn-Tucker necessary condition,then we are attempting to find nonnegative scalars λ1 and λ2, not both zero, suchthat

∇f(4, 4) =(− 1

4, 1

)= λ1∇g1(4, 4) + λ2∇g2(4, 4) = λ1

(− 1

4, −1

)+ λ2(1, 4).

That is, we are attempting to solve

− 1

4= − 1

4λ1 + λ2 and 1 = −λ1 + 4λ2.

These are inconsistent equations. This is a statement that the Karush-Kuhn-Tuckercondition does not exist at the optimal solution to the problem. The problem is anexample of the difficulty pointed out by Slater, which is why the proof of the Karush-Kuhn-Tucker Theorem assumes that there are points in the problem’s feasible setthat are neighboring to the locally maximal point. That assumption is not met in thepresent problem, so the Karush-Kuhn-Tucker necessary condition fails.

Answer to Problem 8.7.(i) The problem’s feasible set is displayed in Figure 8.19. The optimal solution is(x∗1, x

∗2) = (9/2, 27/4). The only constraint that binds at the optimal solution is g1.

(ii) The complete collection of first-order necessary maximization conditions consistsof the Karush-Kuhn-Tucker condition and the two complementary slackness condi-tions. The Karush-Kuhn-Tucker condition is that there exist nonnegative scalars λ1and λ2, not both zero, that at a local maximum (x1, x2) satisfy

∇f(x1, x2) = λ1∇g1(x1, x2) + λ2∇g2(x1, x2).

Page 300: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

8.10. ANSWERS 281

Figure 8.19: The feasible set, the optimal solution, and the gradient vectors forproblem 8.7.

That is,

(1, 1) = λ1(2(x1 − 4), 1) + λ2(3(x1 − 5)2,−12). (8.17)

The complementary slackness conditions are

λ1(7− (x1 − 4)2 − x2) = 0 and λ2(−60− (x1 − 5)3 + 12x2) = 0. (8.18)

Suppose that we cannot draw Figure 8.19 and so have to rely upon the necessaryoptimality conditions (8.17) and (8.18) to locate possible local maxima. How do weuse these equations? There are four possibilities. One possibility is that both of theconstraints are slack. But this makes no sense since the objective function is strictlyincreasing.

Another possibility is that both constraints bind. To locate such a point we must

Page 301: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

282 CHAPTER 8. CONSTRAINED OPTIMIZATION, PART I

simultaneously solve

(x1 − 4)2 + x2 = 7 and (x1 − 5)3 − 12x2 = −60.

There are two solutions; (x1, x2) ≈ (1·860, 2·420) and (x1, x2) ≈ (5·412, 5·006). Wemust check if the Karush-Kuhn-Tucker condition (8.17) is satisfied at either one, orboth, of these solutions.

At (x1, x2) = (1·860, 2·420), the Karush-Kuhn-Tucker condition (8.17) is

(1, 1) = λ1(−4·28, 1) + λ2(29·6,−12) ⇒ λ1 = −1·91 and λ2 = −0·243.

λ1 < 0 and λ2 < 0 inform us that, at (x1, x2) = (1·860, 2·420), the gradient vector ofthe objective function is not contained in the convex cone that is generated by thegradient vectors at (x1, x2) = (1·860, 2·420) of constraints 1 and 2. Thus (x1, x2) =(1·860, 2·420) cannot be a local maximum for the constrained optimization problem.

At (x1, x2) = (5·412, 5·006), the Karush-Kuhn-Tucker condition is

(1, 1) = λ1(2·824, 1) + λ2(0·509,−12) ⇒ λ1 = 0·364 and λ2 = −0·053.

λ2 < 0 informs us that, at (x1, x2) = (5·412, 5·006), the gradient vector of the objectivefunction is not contained in the convex cone that is generated by the gradient vectorsat (x1, x2) = (5·412, 5·006) of constraints 1 and 2. Thus (x1, x2) = (5·412, 5·006)cannot be a local maximum for the constrained optimization problem.

The next possibility is that only constraint g2 binds at the problem’s optimalsolution. If so, then g1 is slack and so, necessarily, λ1 = 0. But then, from (8.17),we observe that necessarily 1 = −12λ2, implying that λ2 < 0, which violates theKarush-Kuhn-Tucker necessary condition.

The only remaining possibility is that only constraint g1 binds at the problem’soptimal solution. If so, then g2 is slack and so, necessarily, λ2 = 0. Then, from (8.17)we observe that necessarily λ1 = 1. OK so far. Using λ1 = 1 in (8.17) shows thatnecessarily 1 = 2(x1 − 4), requiring that x1 = 9/2. Since constraint g1 is presumedto be binding, it then follows that

7 =

(9

2− 4

)2

+ x2 ⇒ x2 =27

4.

We have determined that there is exactly one point, (x1, x2) = (9/2, 27/4), thatsatisfies the necessary first-order optimality conditions. This point must, therefore,be the problem’s optimal solution.

Page 302: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

Chapter 9

Lagrange Functions

9.1 What Is a Lagrange Function?

At no point in our Chapter 8 discussion of first-order necessary optimality conditionsfor differentiable constrained optimization problems did we refer to or use a Lagrangefunction. Yet many discussions of solving constrained optimization problems intro-duce a Lagrange function at an early stage and use it intensively. Is a Lagrangefunction necessary for such a discussion? Obviously not, since we did not use it. Butthere must be something useful about such a function since, otherwise, few wouldmention it when describing how to solve constrained optimization problems. So letus now discover some of the valuable properties and uses of Lagrange functions. InChapter 10 we will see that Lagrange functions are very helpful when deriving andusing second-order optimality conditions. Chapters 12 and 13 explain the usefulnessof Lagrange functions in comparative statics analysis.

It may interest you to know that the spectacular mathematician whom we knowas Joseph-Louis Lagrange was born in Turin, Italy, in 1736 as Giuseppe Luigi La-grangia. Lagrange changed his name when he left Italy as a young man to live andwork in Prussia. Eventually he moved to France, where he narrowly avoided serioustrouble with the French revolutionaries. He died in 1810, shortly after receiving oneof France’s greatest honors.

Throughout the book I have used lower-case letters like f and g to denote functionsand upper-case letters like F and G to denote correspondences. I’m going to makean exception to this convention by using capital L to denote Lagrange functions. Ipromise I will not use L to denote anything other than a Lagrange function. Why dothis? Because Lagrange is a proper noun (it’s a person’s name) and, to my knowledge,

283

Page 303: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

284 CHAPTER 9. LAGRANGE FUNCTIONS

in the English language every proper noun starts with a capital letter.

Definition 9.1 (Lagrange Functions). The Karush-Kuhn-Tucker Lagrange functionfor the inequality constrained optimization problem

maxx1,...,xn

f(x1, . . . , xn) subject to gj(x1, . . . , xn) ≤ bj for j = 1, . . . ,m (9.1)

is the function L : �n ×�m+ �→ � that is

L(x1, . . . , xn, λ1, . . . , λm) = f(x1, . . . , xn) +m∑j=1

λj(bj − gj(x1, . . . , xn)), (9.2)

where λ1, . . . , λm ≥ 0.

The Fritz-John Lagrange function for the inequality constrained problem (9.1) isthe function L : �n �m+1

+ �→ � that is

L(x1, . . . , xn, λ0, λ1, . . . , λm) = λ0f(x1, . . . , xn) +m∑j=1

λj(bj − gj(x1, . . . , xn)), (9.3)

where λ0, λ1, . . . , λm ≥ 0.

The nonnegative scalar variables λ0, λ1, . . . , λm are called the Lagrange multiplier-s. Notice that, for the Lagrange functions, the values of x1, . . . , xn are unconstrained ;the “x-part” of the domain of the Lagrange functions is all of �n. In contrast thevalues of the multipliers λ0, λ1, . . . , λm are restricted to being nonnegative (note: mul-tipliers can take negative values in problems with equality constraints). Some of theproperties of the Lagrange function L are inherited from the properties of its com-ponent functions f, g1, . . . , gm. For example, if each of these component functions isdifferentiable with respect to x1, . . . , xn, then so too will be L.

To avoid a lot of duplication, from here onwards the discussion will assume thata constraint qualification applies so that we can set λ0 = 1 and talk only of theKarush-Kuhn-Tucker Lagrange function.

The immediate appeal of a Lagrange function is that it offers a convenient way towrite down the Karush-Kuhn-Tucker necessary condition. Unfortunately, sometimesfalse assertions are made about the uses and properties of Lagrange functions. Be-cause these functions are so useful, it is important to be clear about what is true andwhat is falsely claimed about them. Probably the best way to begin this discussion isto start with what Lagrange told us, in 1788, about the functions named after him.

Page 304: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

9.1. WHAT IS A LAGRANGE FUNCTION? 285

The particular problem in which Lagrange was interested is an equality constrainedoptimization problem; i.e. a problem of the form

maxx1,...,xn

f(x1, . . . , xn) subject to gj(x1, . . . , xn) = bj for j = 1, . . . , k, (9.4)

in which n > k and a constraint qualification holds. The Lagrange function for sucha problem is

L(x1, . . . , xn, λ1, . . . , λk)

= f(x1, . . . , xn) + λ1(b1 − g1(x1, . . . , xn)) + · · ·+ λk(bk − gk(x1, . . . , xn)).(9.5)

The first-order derivatives of this function are

∂L

∂x1=

∂f

∂x1− λ1

∂g1∂x1

− · · · − λk∂gk∂x1

...... (9.6)

∂L

∂xn=

∂f

∂xn− λ1

∂g1∂xn

− · · · − λk∂gk∂xn

and∂L

∂λ1= b1 − g1(x1, . . . , xn)

...... (9.7)

∂L

∂λk= bk − gk(x1, . . . , xn).

Notice that the value of each of the k partial derivatives in (9.7) is zero for any feasible(not necessarily optimal) point (x1, . . . , xn) because the constraints in problem (9.4)are equalities. Now suppose that (x∗1, . . . , x

∗n) is a local optimum for problem (9.4).

Then each of the n partial derivatives in (9.6) is zero, too, because the Karush-Kuhn-Tucker condition is that this necessarily is so. What have we discovered? We havediscovered that any locally optimal solution to problem (9.4) necessarily is a criticalpoint for the problem’s Lagrange function.

Theorem 9.1 (Lagrange). If (x∗, λ∗) = (x∗1, . . . , x∗n, λ

∗1, . . . , λ

∗k) is a locally optimal

solution to problem (9.4), then (x∗, λ∗) is a critical point of the Lagrange function forproblem (9.4).

The theorem does not assert that the Lagrange function is maximized or minimizedat (x∗, λ∗). Such claims are false, in fact. All the theorem asserts is that the n + kdirectional slopes of the Lagrange function must all be zero at (x∗, λ∗).

Page 305: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

286 CHAPTER 9. LAGRANGE FUNCTIONS

Can we apply Lagrange’s Theorem to the inequality constrained problem (9.1)?Yes, but with one caution. We must allow for the possibility that some of the prob-lem’s constraints are slack at a locally optimal solution. So suppose that (x∗, λ∗) isa locally optimal solution for problem (9.1) and that at this point k of the m con-straints bind, where k ∈ {1, . . . ,m}. Without loss of generality, suppose that theindices of the constraints that bind at x∗ are j = 1, . . . , k. Then the locally optimalsolution (x∗1, . . . , x

∗n) to the inequality constrained problem (9.1) with its accompany-

ing multipliers (λ∗1, . . . , λ∗k, 0, . . . , 0︸ ︷︷ ︸

m−k zeros

) is a locally optimal solution, with accompanying

multipliers (λ∗1, . . . , λ∗k), to the equality constrained problem (9.4). It is to this equal-

ity constrained problem that Lagrange’s Theorem is applied.

9.2 Revisiting theKarush-Kuhn-TuckerCondition

The Karush-Kuhn-Tucker condition is the statement that the Lagrange function’sgradient vector in the x-space is the n-dimensional zero vector; i.e.

∇xL(x, λ) =

(∂L

∂x1, . . . ,

∂L

∂xn

)=

(∂f

∂x1−

m∑j=1

λj∂gj(x)

∂x1, . . . ,

∂f

∂xn−

m∑j=1

λj∂gj∂xn

)

= 0n. (9.8)

This allows a restatement of the Karush-Kuhn-Tucker Necessity Theorem.

Theorem 9.2 (Restatement of the Karush-Kuhn-Tucker Necessity Theorem). Con-sider the problem

maxx∈�n

f(x) subject to gj(x) ≤ bj for j = 1, . . . ,m, (9.9)

where the objective function f and the constraint functions g1, . . . , gm are all contin-uously differentiable with respect to x, and the feasible set

C = {x ∈ �n | gj(x) ≤ bj ∀ j = 1, . . . ,m}

is not empty. The problem’s Lagrange function is

L(x, λ) = f(x) + λ1(b1 − g1(x)) + · · ·+ λm(bm − gm(x)).

Page 306: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

9.3. SADDLE-POINTS FOR LAGRANGE FUNCTIONS 287

Let x∗ be a locally optimal solution to the problem. If a constraint qualification issatisfied at x∗, then there exist nonnegative values λ∗1, . . . , λ

∗m, not all zero, such that

∂L(x;λ∗)∂xi

|x=x∗ = 0 for all i = 1, . . . , n (9.10)

and λ∗j(bj − gj(x∗)) = 0 for all j = 1, . . . ,m. (9.11)

Any critical point for the Lagrange function with respect to x is a point at whichthe slopes of L in the x1, . . . , xn directions are all zero. This is what (9.10) states.But how do we know that such a value of x is feasible? The complementary slacknessconditions eliminate infeasible values of x; i.e. x∗ is feasible only if x∗ satisfies (9.11).Why?

Suppose that, at some locally optimal solution x∗, at least one (let’s say the first)of the complementary slackness conditions is not satisfied. That is, λ∗1(b1 − g1(x

∗)) =0. Let’s suppose, then, that λ∗1(b1 − g1(x

∗)) < 0. Since λ∗1 ≥ 0, this implies thatg1(x

∗) > b1. But then x∗ is not a feasible solution. So now let’s suppose thatλ∗1(b1 − g1(x

∗)) > 0. Then λ∗1 > 0 and g1(x∗) < b1. But λ

∗1 > 0 only if constraint g1 is

binding at x∗, so we have a contradiction.

There is nothing to be gained by writing down a Lagrange function and thendifferentiating it to derive the Karush-Kuhn-Tucker and complementary slacknessconditions. We might as well directly write down these first-order necessary opti-mality conditions, just as we did throughout Chapter 8. Nevertheless, there are atleast four important reasons to consider the Lagrange function. First, if the Lagrangefunction possesses a particular type of saddle-point, then the “x-part” of the saddle-point is the global solution to the constrained optimization problem. Second, theLagrange function is very useful when considering the primal-dual version of the con-strained optimization problem. What this means is explained in Chapter 12. Third,as we will see in Chapter 10, the Lagrange function simplifies the task of consideringsecond-order optimality conditions. Fourth, the Lagrange function simplifies derivingsome types of comparative statics results about the properties of optimal solutions toconstrained optimization problems. This is explained in Chapters 12 and 13.

9.3 Saddle-Points for Lagrange Functions

Saddle-points were discussed in Section 2.19. If you have not read this discussion, orhave forgotten what you read, then you should review this material before proceedingfurther in this section.

Page 307: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

288 CHAPTER 9. LAGRANGE FUNCTIONS

Because the Lagrange function for problem (9.4) is a function of n+ k variables,it can possess many types of saddle-points (x, λ) = (x1, . . . , xn, λ1, . . . , λk). Oneparticular type is especially interesting. This is the saddle-point in which the Lagrangefunction is maximized with respect to the choice variables x1, . . . , xn at the point(x1, . . . , xn), when the values of the multipliers are fixed at λ1, . . . , λk, and, also,is minimized with respect to λ1, . . . , λk at the point (λ1, . . . , λk), when the valuesof the choice variables are fixed at x1, . . . , xn. We will call this an “x-max, λ-minsaddle-point.”

Definition 9.2 (x-max, λ-min Saddle-Point). (x, λ) = (x1, . . . , xn, λ1, . . . , λk) is alocal x-max, λ-min saddle-point for the Lagrange function (9.5) in a neighborhoodN(x, λ) ⊂ �n ×�k

+ if and only if, for all (x, λ) ∈ N(x, λ),

Lr(x1, . . . , xn; λ1, . . . , λk)≤L(x1, . . . , xn, λ1, . . . , λk)≤Lr(λ1, . . . , λk; x1, . . . , xn).(9.12)

The function Lr(x; λ) denotes the Lagrange function restricted by having the val-ues of the multipliers fixed at λ1, . . . , λk. The function L

r(λ; x) denotes the Lagrangefunction restricted by having the values of the choice variables fixed at x1, . . . , xn.The superscripts r denote restricted.

For an example of another type of saddle-point we might consider one in whichthe Lagrange function is minimized with respect to the choice variables x1, . . . , xnat the point (x1, . . . , xn) when the values of the multipliers are fixed at λ1, . . . , λkand, also, is maximized with respect to λ1, . . . , λk at the point (λ1, . . . , λk) when thevalues of the choice variables are fixed at x1, . . . , xn. We will call this an “x-min,λ-max saddle-point”.

Definition 9.3 (x-min, λ-max Saddle-Point). (x, λ) = (x1, . . . , xn, λ1, . . . , λk) is alocal x-min, λ-max saddle-point for the Lagrange function (9.5) in a neighborhoodN(x, λ) ⊂ �n ×�k

+ if and only if, for all (x, λ) ∈ N(x, λ),

Lr(x1, . . . , xn; λ1, . . . , λk)≥L(x1, . . . , xn, λ1, . . . , λk)≥Lr(λ1, . . . , λk; x1, . . . , xn).(9.13)

In both of the above definitions, if the neighborhood N(x, λ) = �n×�k+ (i.e. if (9.12)

or (9.13) is true for all x ∈ �n and for all λ ∈ �k+) then the saddle-point is global.

Let’s have a look at an example of each of these two types of saddle-points.Consider the constrained optimization problem that is to maximize f(x) = +4

√x

for x ∈ [0, 2]. Write the problem as

maxx

f(x) = +4√x subject to g1(x) = −x ≤ 0 and g2(x) = x ≤ 2. (9.14)

Page 308: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

9.3. SADDLE-POINTS FOR LAGRANGE FUNCTIONS 289

The globally optimal solution is x∗ = 2. Constraint g1 is slack at the optimal solution,so we can ignore it. Constraint g2 binds at the optimal solution with a multipliervalue λ∗2 = +

√2, so (9.14) has the same optimal solution as the equality constrained

problem

maxx

f(x) = +4√x subject to g2(x) = x = 2. (9.15)

The Lagrange function for this equality constrained problem is

L(x, λ2) = +4√x+ λ2(2− x). (9.16)

Figure 9.1 displays the graph of the Lagrange function (9.16). The point (x∗, λ∗2) =(2,

√2 ) is a global x-max, λ-min saddle-point for this Lagrange function.

Figure 9.1: The Lagrange function L(x, λ2) = 4√x + λ2(2 − x) has a global x-max,

λ-min saddle-point at (x, λ2) = (2,√2 ).

Now consider the similar problem that is to maximize f(x) = x2 for x ∈ [0, 2].Write the problem as

maxx

f(x) = x2 subject to g1(x) = −x ≤ 0 and g2(x) = x ≤ 2. (9.17)

Page 309: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

290 CHAPTER 9. LAGRANGE FUNCTIONS

The globally optimal solution is again x∗ = 2. Constraint g1 is slack at the optimalsolution. Constraint g2 binds at the optimal solution with a multiplier value λ∗2 = 4,so problem (9.17) has the same optimal solution as the equality constrained problem

maxx

f(x) = x2 subject to g2(x) = x = 2. (9.18)

The Lagrange function for this problem is

L(x, λ2) = x2 + λ2(2− x). (9.19)

Figure 9.2 displays the graph of the Lagrange function (9.19). The point (x∗, λ∗2) =(2, 4) is a global x-min, λ-max saddle-point for this Lagrange function.

Figure 9.2: The Lagrange function L(x, λ2) = x2 + λ2(2 − x) has a global x-min,λ-max saddle-point at (x, λ2) = (2, 4).

At this point you should be asking, “Is this stuff about saddle-points for La-grange functions important?” The answer is “Yes” because any point that satisfies aconstraint qualification and is a solution to the Karush-Kuhn-Tucker condition andthe complementary slackness conditions for a constrained optimization problem is asaddle-point of some sort for the Lagrange function.

Page 310: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

9.4. SADDLE-POINTS AND OPTIMALITY 291

Let’s turn to contradicting two erroneous statements that seem to have developedsome standing among economists. It is sometimes asserted that, to locate a localmaximum for a constrained optimization problem, we must locally “maximize theLagrange function.” This is never true. Any local maximum (x, λ) for the constrainedoptimization problem is a critical point for the Lagrange function, but this criticalpoint is always a saddle-point of some sort for the Lagrange function and so is nevera local maximum for the Lagrange function.

Another assertion is that a local maximum (x, λ) for a constrained optimizationproblem “maximizes the restricted Lagrange function” Lr(x; λ), in which the valuesof the multipliers of the Lagrange function are fixed at λ = (λ1, . . . , λk). Althoughthis assertion can be true for particular problems, it is not true in general. Againconsider problem (9.15). The optimal solution to the problem is (x, λ) = (2,

√2 ).

From (9.16), the restricted Lagrange function L(rx; λ) for this problem is

Lr(x;λ =

√2)= 4

√x+

√2 (2− x).

This function achieves a global maximum at x = 2, so, in this instance, the restrictedLagrange function is indeed maximized at x = 2. However, now consider againproblem (9.18). The optimal solution to the problem is (x, λ) = (2, 4), so, from(9.19), the restricted Lagrange function Lr(x; λ) is

Lr(x;λ = 4) = x2 + 4(2− x).

This function achieves a global minimum at x = 2, contradicting the assertion thatthe restricted Lagrange function is always locally maximized at a locally optimalsolution to the constrained optimization problem.

9.4 Saddle-Points and Optimality

We have established that a local optimum for a constrained optimization problemis a saddle-point of some sort for the problem’s Lagrange function. Must a saddle-point for a problem’s Lagrange function be a local optimum for that problem? Theanswer is “No.” However, there does exist one rather remarkable result of this type:if the saddle-point is a global x-max, λ-min saddle-point, then the “x-part” of thesaddle-point is a global optimum for the constrained optimization problem. Here isthe theorem.

Page 311: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

292 CHAPTER 9. LAGRANGE FUNCTIONS

Theorem 9.3 (Kuhn-Tucker Saddle-Point Theorem). If the Lagrange function for theconstrained optimization problem (9.1) possesses a global x-max, λ-min saddle-point(x∗, λ∗), then x∗ is a globally optimal solution to problem (9.1).

Proof. The proof proceeds in two steps. The first step is to show that the globalx-max, λ-min saddle-point property implies that x∗ is a feasible solution. The secondstep is then to show that x∗ is a globally optimal solution.

(x∗, λ∗) is a global x-max, λ-min saddle-point for L, so L(x∗, λ∗) ≤ Lr(λ; x∗) forall λ ∈ �m

+ . That is,

f(x∗) +m∑j=1

λ∗j(bj − gj(x∗)) ≤ f(x∗) +

m∑j=1

λj(bj − gj(x∗))

for all λj ≥ 0 and for all j = 1, . . . ,m. Hence

m∑j=1

λ∗j(bj − gj(x∗)) ≤

m∑j=1

λj(bj − gj(x∗)) ∀ λj ≥ 0, ∀ j = 1, . . . ,m. (9.20)

Let’s be clear that the left-hand side of (9.20) is a constant since x∗ and λ∗ are givenvectors. The right-hand side of (9.20) is a linear function only of λ1, . . . , λm and eachis nonnegative (see Definition 9.1).

(9.20) cannot be true if, for even one j, we have gj(x∗) > bj since then bj−gj(x∗) <

0, and by choosing λj to be as large as we like, we can make the right-hand side of(9.20) as negative as we wish. This contradicts that the right-hand side of (9.20)has a finite lower bound, which is the (constant) value of the left-hand side of (9.20).Therefore gj(x

∗) ≤ bj for all j = 1, . . . ,m. Thus x∗ is a feasible solution.

Now we need to show that x∗ is a globally optimal solution. Since λ∗j ≥ 0 andbj − gj(x

∗) ≥ 0 for all j = 1, . . . ,m, it follows that

m∑j=1

λ∗j(bj − gj(x∗)) ≥ 0. (9.21)

We are now going to show that the sum in (9.21) cannot be strictly positive and thusmust be exactly zero. The argument is by contradiction, so suppose that

m∑j=1

λ∗j(bj − gj(x∗)) = k > 0. (9.22)

Page 312: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

9.5. CONCAVE PROGRAMMING PROBLEMS 293

Now choose λj = λ∗j/2 for each j = 1, . . . ,m. λ∗j ≥ 0 shows that λj ≥ 0 also. Butthen

m∑j=1

λj(bj − gj(x∗)) =

k

2< k =

m∑j=1

λ∗j(bj − gj(x∗)).

This contradicts that λ∗1, . . . , λ∗m are values of λ1, . . . , λm that minimize Lr(λ; x∗).

Thereforem∑j=1

λ∗j(bj − gj(x∗)) = 0. (9.23)

Since (x∗, λ∗) is a global x-max, λ-min saddle-point for L, Lr(x;λ∗) ≤ L(x∗, λ∗) forall x ∈ �n; i.e.

f(x) +m∑j=1

λ∗j(bj − gj(x)) ≤ f(x∗) +m∑j=1

λ∗j(bj − gj(x∗)) = f(x∗) (9.24)

by (9.23). But bj−gj(x) ≥ 0 for any feasible x, so, f(x) ≤ f(x∗) everywhere on C.There is much that is attractive about this result. First, it does not require that

any of the functions f, g, . . . , gm be differentiable. Second, it is not required that thefeasible set of the constrained optimization problem satisfy a constraint qualification.Third, no particular curvatures are demanded of f or any of g1, . . . , gm. Against all ofthese attributes is the demanding requirement that the problem’s Lagrange functionhave a global x-max, λ-min saddle-point.

Notice that the proof of the theorem comes very close to establishing (see (9.23))a useful but less important result that merits emphasis by stating it separately.

Theorem 9.4 (x-max, λ-min Saddle-Point Implies Complementary Slackness). If theLagrange function for the constrained optimization problem (9.1) possesses a globalx-max, λ-min saddle-point (x∗, λ∗), then λ∗j(bj − gj(x

∗)) = 0 for each j = 1, . . . ,m.

Proof. By Theorem 9.3, x∗ is optimal and therefore feasible, so λ∗j ≥ 0 and bj −gj(x

∗) ≥ 0 for all j = 1, . . . ,m. Hence λ∗j(bj − gj(x∗)) ≥ 0 for all j = 1, . . . ,m. If, for

any j, the product λj(bj−gj(x∗)) > 0, then it follows from (9.23) that λk(bk−gk(x∗)) <0 for some k = j. Contradiction. Hence λ∗j(bj−gj(x∗)) = 0 for each j = 1, . . . ,m.

9.5 Concave Programming Problems

The usual method for solving a differentiable constrained optimization problem startsby solving the first-order local optimality conditions. Examining the second-order

Page 313: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

294 CHAPTER 9. LAGRANGE FUNCTIONS

conditions then eliminates solutions that are not local optima. Each local optimum isthen evaluated to determine the global optimum. Examining second-order conditionsis typically a lot of work, so it would be nice if the first-order conditions were bothnecessary and sufficient for a local optimum. This is not so in general, but, there isone important type of constrained optimization problem for which it is so. This isthe concave programming problem.

Definition 9.4 (Concave Programming Problem). A constrained optimization prob-lem

maxx∈�n

f(x) subject to gj(x) ≤ bj for j = 1, . . . ,m (9.25)

is a concave programming problem if f is concave with respect to x and each ofg1, . . . , gm is convex with respect to x.

Theorem 9.5 (Solution to a Concave Programming Problem). Consider a concaveprogramming problem (9.25) in which the functions f, g1, . . . , gm are all differentiablewith respect to x1, . . . , xn and the interior of the feasible set is not empty. Let (x∗, λ∗)be a solution to the problem’s first-order optimality conditions. Then (x∗, λ∗) is aglobal x-max, λ-min saddle-point for the problem’s Lagrange function and x∗ is aglobally optimal solution to the problem. If f is strictly concave with respect to x,then x∗ is the unique global solution to the problem.

This result should not surprise you. Let’s think about it. Suppose, for a moment,that the problem is unconstrained. Then the problem’s Lagrange function is justL(x1, . . . , xn) = f(x1, . . . , xn) and, at a local maximum x = x∗ for f , the Karush-Kuhn-Tucker condition, ∇f(x∗) = 0, necessarily is satisfied. This same conditionis also sufficient for x = x∗ to be a global maximum for f because f is a concavefunction of x. Now let’s put the constraints back. The problem’s Lagrange functionis

L(x1, . . . , xn, λ1, . . . , λm) = f(x1, . . . , xn) +m∑j=1

λj(bj − gj(x1, . . . , xn)).

f is concave with respect to x. Each of −g1, . . . ,−gm is concave with respect to x(the negative of a convex function is a concave function) and λ1 ≥ 0, . . . , λm ≥ 0.So L is a nonnegative linear combination of functions that are concave with respectto x. L is therefore concave with respect to x. A local optimum to any differentiableconstrained optimization problem necessarily is a critical point for the problem’sLagrange function. Hence the gradient of the Lagrange function with respect to xis a zero vector. But L is concave with respect to x, so the first-order zero gradient

Page 314: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

9.5. CONCAVE PROGRAMMING PROBLEMS 295

condition is sufficient for the restricted Lagrange function Lr(x;λ∗) to be globallymaximized at x = x∗ (i.e. the left-hand inequality in (9.12) is satisfied). Last, theother restricted Lagrange function,

Lr(λ; x∗) = f(x∗) + λ1(b1 − g1(x∗)) + · · ·+ λm(bm − g(x∗)),

is just an affine function of λ1, . . . , λm. Each λj ≥ 0 and each bj−gj(x∗) ≥ 0, since x∗

is a feasible solution. Thus the smallest possible value of Lr(λ; x∗) is achieved whenλ1(b1 − g1(x

∗)) + · · ·+ λm(bm − g(x∗)) = 0. The complementary slackness conditionsstate that this necessarily is so and thus the right-hand inequality in (9.12) is alsoestablished.

Proof of Theorem 9.5. We are given that the first-order optimality conditions aretrue at the point (x∗, λ∗). Our task is to show that this information implies that x∗

is the globally optimal solution to problem (9.25).

We will start by establishing that the problem’s feasible set satisfies Slater’s con-straint qualification. Each of the constraint functions gj is convex with respect tox and so is continuous and quasi-convex with respect to x. Therefore each of thelower-contour sets {x ∈ �n | gj(x) ≤ bj} is a closed and convex set. The problem’sfeasible set,

C = {x ∈ �n | g1(x) ≤ b1, . . . , gm(x) ≤ bm} = ∩mj=1{x ∈ �n | gj(x) ≤ bj},

is therefore an intersection of finitely many closed and convex sets and thus is aclosed and convex set. Its interior is not empty by hypothesis, so Slater’s constraintqualification is satisfied and the Karush-Kuhn-Tucker condition is well-defined at anyx ∈ C.

For each j = 1, . . . ,m the function −gj(x) is concave with respect to x and λj ≥ 0.The Lagrange function

L(x, λ) = f(x) +m∑j=1

λj(bj − gj(x))

is therefore a nonnegative linear combination of functions that are concave with re-spect to x and so is a function that is concave with respect to x.

(x∗, λ∗) is a solution to the Karush-Kuhn-Tucker condition, so

∂Lr(x;λ∗)∂xi

|x=x∗ = 0 ∀ i = 1, . . . , n.

Page 315: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

296 CHAPTER 9. LAGRANGE FUNCTIONS

Because Lr(x;λ∗) is concave with respect to x, it follows that Lr(x;λ∗) ≤ L(x∗, λ∗)for all x ∈ �n and for all x ∈ C in particular.

Also, since λ∗j(bj − gj(x∗)) = 0 and λj(bj − gj(x

∗)) ≥ 0 for all j = 1, . . . ,m,

L(x∗, λ∗) = f(x∗) = f(x∗) +m∑j=1

λ∗j(bj − gj(x∗))

≤ f(x∗) +m∑j=1

λj(bj − gj(x∗)) = Lr(λ; x∗) ∀ λ ∈ �m

+ .

We have now established that (x∗, λ∗) is an x-max, λ-min saddle-point for theproblem’s Lagrange function, so Theorem 9.3 informs us that x∗ is a globally optimalsolution to the problem.

Suppose that f is strictly concave with respect to x. The argument establishingthat x∗ is then unique is by contradiction. Suppose that (x′, λ′) and (x′′, λ′′), withx′ = x′′, are both solutions to the problem’s first-order optimality conditions. Thenboth x′ and x′′ are globally optimal solutions, and thus feasible solutions, to theproblem. C is a convex set with a nonempty interior, so, for any θ ∈ (0, 1), x(θ) =θx′ + (1 − θ)x′′ ∈ C and f(x(θ)) > f(x′) = f(x′′) because f is strictly concave withrespect to x. This contradicts the global optimality of x′ and x′′, so there is just onesolution to the problem’s first-order optimality conditions and thus a unique globaloptimal solution to the problem.

Theorem 9.5 is easily generalized. Let’s think using simple geometry. What is theidea at the heart of the above result? It is that the feasible set C is a convex set witha nonempty interior, and that the objective function f is concave with respect to x.Then any value for x ∈ C that locally maximizes f , must be a global maximum for fon C. Right? A less restrictive way to ensure that C is convex is to insist that eachof the constraint functions g1, . . . , gm is quasi-convex with respect to x (a functiondefined on �n is quasi-(strictly)-convex if and only if all of its lower-contour sets are(strictly) convex sets). Then, once again, C is the intersection of convex sets andso is a convex set. There is no need to insist upon the stronger condition that theconstraint functions be convex with respect to x (a quasi-convex function does nothave to be a convex function). The result we obtain by this reasoning is as follows.

Theorem 9.6. Consider a problem (9.1) in which the objective function f is differ-entiable and concave with respect to x1, . . . , xn, the constraint functions g1, . . . , gm areall differentiable and quasi-convex with respect to x1, . . . , xn, and the interior of thefeasible set C is not empty. Let C ⊂ �n

+. Let (x∗, λ∗) be a solution to the problem’s

Page 316: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

9.6. REMARKS 297

first-order optimality conditions. Then x∗ is a globally optimal solution to the prob-lem. If f is strictly concave with respect to x, then x∗ is the unique globally optimalsolution to the problem.

If you want to read a proof, then try the original in Arrow and A. Enthoven. 1961.Notice that, although the conditions of this theorem are less restrictive than those ofTheorem 9.5 in that the constraint functions now need to be only quasi-convex, theconditions are more restrictive in that only nonnegative vectors x are allowed; i.e.C ⊂ �n

+. Also notice that there is no mention of an x-max, λ-min saddle-point. Thisis for a very good reason: once we generalize from convex to quasi-convex constraintfunctions, the problem’s Lagrange function may not have an x-max, λ-min saddle-point.

9.6 Remarks

We have explored just a few uses of Lagrange functions. Least important is that La-grange functions let us use almost mindless differentiation to write down the Karush-Kuhn-Tucker necessary condition for a local constrained optimum. Lagrange func-tions have more valuable uses that will be explored in the following chapters. Thefirst of these, and arguably the most important, is explained in the next chapter,where we will discover that the information that a point (x∗, λ∗) is a locally opti-mal solution implies a valuable necessary second-order condition that so far we haveneither explored nor exploited. This second-order condition allows us to establish avariety of properties possessed by optimal solutions. These are the properties thatmost economists rely upon when applying economic theory. They allow us, for ex-ample, to predict the effects of a change to a tax rate or an interest rate. Chapters11, 12, and 13 explain how to obtain such results.

Page 317: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

298 CHAPTER 9. LAGRANGE FUNCTIONS

9.7 Problems

Problem 9.1. Consider a rational consumer with a Cobb-Douglas direct utility func-tion who chooses an affordable consumption bundle. The problem is

maxx1, x2

U(x1, x2;α) = xα1xα2 subject to g1(x1, x2) = −x1 ≤ 0,

g2(x1, x2) = −x2 ≤ 0 and g3(x1, x2) = p1x1 + p2x2 ≤ y.(9.26)

For α1 > 0, p1 > 0, p2 > 0, and y > 0, the optimal solution (x∗1(p1, p2, y), x∗2(p1, p2, y))

is interior to �2+; i.e. x

∗1(p1, p2, y) > 0 and x∗2(p1, p2, y) > 0. Moreover, this optimal

solution is unique. Consequently the optimal solution to the inequality constrainedproblem (9.26) is the same as the unique optimal solution to the equality constrainedproblem

maxx1, x2

U(x1, x2;α) = (x1x2)α subject to g3(x1, x2) = p1x1 + p2x2 = y. (9.27)

The first-order necessary optimality conditions for (9.27) are(α (x∗1)

α−1 (x∗2)α , α (x∗1)

α (x∗2)α−1) = λ∗3(p1, p2) and p1x

∗1 + p2x

∗2 = y, (9.28)

resulting in the Cobb-Douglas ordinary demand functions

x∗1(p1, p2, y) =y

2p1and x∗2(p1, p2, y) =

y

2p2. (9.29)

The associated value for the multiplier λ3 is

λ∗3(p1, p2, y, α) =(α

p1

)α(α

p2

)α(2α

y

)1−2α

. (9.30)

The Lagrange function for the equality constrained problem (9.27) is

L(x1, x2, λ3;α) = (x1x2)α + λ3(y − p1x1 − p2x2). (9.31)

Lagrange’s Theorem is that the point (x∗1(p1, p2, y), x∗2(p1, p2, y), λ

∗3(p1, p2, y, α)) nec-

essarily is a critical point for the Lagrange function (9.31). We now know that thiscritical point is some type of saddle-point for this Lagrange function. Let us explorejust what type of saddle-point it might be. Is it, for example, an x-max, λ-minsaddle-point?(i) If 0 < α < 1

2, then is (x∗1(p1, p2, y), x

∗2(p1, p2, y), λ

∗3(p1, p2, y)) an x-max, λ-min

saddle-point for the Lagrange function (9.31)?

Page 318: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

9.8. ANSWERS 299

(ii) If 1 < α, then is (x∗1(p1, p2, y), x∗2(p1, p2, y), λ

∗3(p1, p2, y)) an x-max, λ-min saddle-

point for the Lagrange function (9.31)?

(iii) If 12< α < 1, then is (x∗1(p1, p2, y), x

∗2(p1, p2, y), λ

∗3(p1, p2, y)) an x-max, λ-min

saddle-point for the Lagrange function (9.31)?

Problem 9.2. Theorem 9.3 takes as given only an x-max, λ-min type of saddle-point. Suppose instead that you know that a point (x∗, λ∗) for some constrainedoptimization problem is an x-min, λ-max saddle-point for the problem’s Lagrangefunction. Is there a theorem, analogous to Theorem 9.3, that applies to such a case?If so, then what is the theorem? If not, then why not?

Problem 9.3. We are given that a point (x∗, λ∗) = (x∗1, . . . , x∗n, λ

∗1, . . . , λ

∗m) is a

saddle-point for the Lagrange function for the constrained optimization problem (9.1).

(i) If the saddle-point is not an x-max, λ-min saddle-point, then can x∗ be the globallyoptimal solution to problem (9.1)? Explain.

(ii) Suppose that x∗ is the globally optimal solution to problem (9.1)? Must (x∗, λ∗)be an x-max, λ-min saddle-point? Explain.

9.8 Answers

Answer to Problem 9.1. We must consider in turn each one of the restrictedLagrange functions Lr(x1, x2;λ

∗3) and Lr(λ3; x

∗1, x

∗2). We need to determine if these

functions are minimized or maximized at the saddle-point.

(i) The answer is “Yes.” The Hessian matrix for the restricted Lagrange functionLr(x1, x2;λ

∗3) is

HLr(x1,x2;λ∗3)(x1, x2) =

⎛⎝−α(1− α)xα−2

1 xα2 α2 (x1x2)α−1

α2 (x1x2)α−1 −α(1− α)xα1x

α−22

⎞⎠ . (9.32)

This is a symmetric matrix, so to determine if the matrix’s quadratic form has adefiniteness property all we need do is check the signs of the first and second principalminors of the matrix. These minors are

M1 = −α(1− α)xα−21 xα2 and M2 = α2(1− 2α)(x1x2)

2(α−1).

If 0 < α < 12, then M1 < 0 and M2 > 0, establishing that Lr(x1, x2;λ

∗3) is a

strictly concave function of x1 and x2 (or, equivalently, that the Hessian matrix is

Page 319: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

300 CHAPTER 9. LAGRANGE FUNCTIONS

negative-definite for any x1 > 0 and for any x2 > 0). Any critical point for such afunction must be a global (note: global, not just local) maximum with respect to x1and x2. Thus we have that, if 0 < α < 1

2,

Lr(x1, x2;λ∗3) < Lr(x∗1, x

∗2;λ

∗3) = L(x∗1, x

∗2, λ

∗3) ∀ x1, x2 ∈ �, (x1, x2) = (x∗1, x

∗2).(9.33)

The restricted Lagrange function Lr(λ3; x∗1, x

∗2) = (x∗1x

∗2)

α+λ3(y− p1x∗1− p2x

∗2) ≡

(x∗1x∗2)

α is constant-valued with respect to λ3, so

Lr(λ3; x∗1, x

∗2) = Lr(λ∗3; x

∗1, x

∗2) = L(x∗1, x

∗2, λ

∗3) ∀ λ3 ≥ 0. (9.34)

(9.33) and (9.34) establish that the point (x∗1(p1, p2, y), x∗2(p1, p2, y), λ

∗3(p1, p2, y, α)) is

an x-max, λ-min saddle-point for the Lagrange function (9.31) when 0 < α < 12.

(ii) The answer is “No.” If 1 < α, then M1 > 0 and M2 < 0, establishing (see The-orem 2.11) that the Hessian matrix of the restricted Lagrange function Lr(x1, x2;λ

∗3)

is indefinite and that the restricted Lagrange function Lr(x1, x2;λ∗3) is neither maxi-

mized nor minimized with respect to both x1 and x2 at (x1, x2) = (x∗1, x∗2) (indeed, the

function does not possess a maximum in �2+ with respect to both x1 and x2). In fact, if

1 < α, then the point (x∗1, x∗2) is a saddle-point (not a maximum and not a minimum)

with respect to x1 and x2 for the restricted Lagrange function Lr(x1, x2;λ∗3). Thus the

point (x∗1(p1, p2, y), x∗2(p1, p2, y), λ

∗3(p1, p2, y, α)) is not an x-max, λ-min saddle-point

for the (unrestricted) Lagrange function when 1 < α.(iii) The answer is “No.” If 1

2< α < 1, then M1 < 0 and M2 < 0, establish-

ing (see Theorem 2.11) that the Hessian matrix of the restricted Lagrange functionLr(x1, x2;λ

∗3) is indefinite and that the restricted Lagrange function Lr(x1, x2;λ

∗3) is

neither maximized nor minimized with respect to x1 and x2 at (x1, x2) = (x∗1, x∗2).

Instead the point (x∗1, x∗2) is again a saddle-point (not a maximum and not a mini-

mum) with respect to x1 and x2 for the restricted Lagrange function Lr(x1, x2;λ∗3)

and so, again, the point (x∗1(p1, p2, y), x∗2(p1, p2, y), λ

∗3(p1, p2, y, α)) is not an x-max,

λ-min saddle-point for the (unrestricted) Lagrange function when 12< α < 1.

Answer to Problem 9.2. We are given a Lagrange function

L(x1, . . . , xn, λ1, . . . , λk) = f(x1, . . . , xn) +k∑

j=1

λj(b− gj(x1, . . . , xn)) (9.35)

and that (x∗, λ∗) is a global x-min, λ-max saddle-point for this function. That is, forall x ∈ �n and for all λ ∈ �k

+,

Lr(x1, . . . , xn;λ∗1, . . . , λ

∗k)≥L(x∗1, . . . , x∗n, λ∗1, . . . , λ∗k)≥Lr(λ1, . . . , λk; x

∗1, . . . , x

∗n).(9.36)

Page 320: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

9.8. ANSWERS 301

The proof of Theorem 9.3 is in two parts, the first of which is a demonstrationthat the λ-min part of the x-max, λ-min saddle-point property implies that x∗ isa feasible solution. Let’s see if the λ-max part of the x-min, λ-max saddle-pointproperty implies the same thing. The right-side inequality in (9.36) is

f(x∗) +k∑

j=1

λ∗j(bj − gj(x∗)) ≥ f(x∗) +

k∑j=1

λj(bj − gj(x∗)) ∀ λj ≥ 0, j = 1, . . . , k,

sok∑

j=1

λ∗j(bj − gj(x∗)) ≥

k∑j=1

λj(bj − gj(x∗)) for all λ1 ≥ 0, . . . , λk ≥ 0. (9.37)

The left side of (9.37) is a fixed number. Just as in the proof of Theorem 9.3, wenow attempt to show that this fixed number is zero. Suppose it is not. Particularly,suppose that

∑kj=1 λ

∗j(bj − gj(x

∗)) > 0. Then, since λ∗1 ≥ 0, . . . , λ∗k ≥ 0, each of

λj = μλ∗j ≥ 0 for j = 1, . . . , k, where μ > 0. By making μ large enough, wemake the right side of (9.37) larger than the fixed value of the left side of (9.37).Hence

∑kj=1 λ

∗j(bj − gj(x

∗)) ≤ 0. Now suppose that∑k

j=1 λ∗j(bj − gj(x

∗)) < 0. Sinceλj ≥ 0 for all j = 1, . . . , k, there must be at least one value of j for which gj(x

∗) >bj. But then x∗ is not feasible. So, if x∗ is a feasible solution to the constrainedoptimization problem, then necessarily

∑kj=1 λ

∗j(bj − gj(x

∗)) ≥ 0. Necessarily, then,∑kj=1 λ

∗j(bj − gj(x

∗)) = 0.This information together with the left side inequality in (9.36) tells us that

f(x) +k∑

j=1

λ∗j(bj − gj(x)) ≥ f(x∗) +k∑

j=1

λ∗j(bj − gj(x∗)) = f(x∗). (9.38)

For feasible choices x,∑k

j=1 λ∗j(bj − gj(x)) ≥ 0, so we cannot deduce from (9.38) that

f(x∗) ≥ f(x). There is no analogue to Theorem 9.3 for an x-min, λ-max saddle-pointfor the Lagrange function (9.35).

Answer to Problem 9.3.(i) Yes. Examples are problem (9.17) and problem (9.26) with 1 < α.(ii) No. Theorem 9.3 is a sufficiency statement. It says that, if (x∗, λ∗) is a global x-max, λ-min type of saddle-point for the constrained optimization problem’s Lagrangefunction, then x∗ is a globally optimal solution to the problem. The theorem doesnot say the reverse. That is, the theorem does not say that, if x∗ is a globally optimalsolution to the constrained optimization problem, then the saddle-point at (x∗, λ∗)for the problem’s Lagrange function must be an x-max, λ-min type of saddle-point.Problem (9.17) is a counterexample to this false assertion.

Page 321: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

Chapter 10

Constrained Optimization, Part II

Chapter 8 described the first-order conditions that necessarily are true at a localmaximum to the constrained optimization problem

maxx1, ... , xn

f(x1, . . . , xn) subject to g1(x1, . . . , xn) ≤ b1, . . . , gm(x1, . . . , xn) ≤ bm, (10.1)

in which the objective function f and the constraint functions g1, . . . , gm are all con-tinuously differentiable with respect to each of x1, . . . , xn. In this chapter we willrequire these functions to all be twice-continuously differentiable. Reminder: A func-tion is twice-continuously differentiable with respect to x1, . . . , xn if and only if allof the function’s second-order derivatives are continuous functions of x1, . . . , xn. Them+ 1 first-order conditions consist of the Karush-Kuhn-Tucker condition

∇f(x∗) = λ∗1∇g1(x∗) + · · ·+ λ∗m∇gm(x∗), (10.2)

where the nonnegative scalars λ∗1, . . . , λ∗m are not all zero, and the complementary

slackness conditionsλ∗j(bj − gj(x

∗)) = 0 ∀ j = 1, . . . ,m. (10.3)

There is an additional condition that necessarily is true at a local optimum, asecond-order condition that will shortly be described. Following that, this chapterdiscusses sets of conditions that are sufficient for a feasible point to be a local opti-mum.

10.1 Necessary vs. Sufficient Conditions

Let’s take a minute to review the fundamental differences between necessary andsufficient optimality conditions. The basic difference is as follows. Necessary opti-

302

Page 322: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.1. NECESSARY VS. SUFFICIENT CONDITIONS 303

mality conditions are conditions that individually have to be true at an optimum,so the flow of the logic is, because x is a local optimum, we know that each one ofthe conditions A, B, and so on must be (is necessarily) true at x. Equivalently, ifany one of the conditions A, B, and so on is not true at x = x, then x is not alocal optimum. Necessary local optimality conditions are individually true when x isa local optimum. In contrast, a set of conditions all true at x = x is sufficient forx = x to be a local optimum if, collectively, these conditions ensure that x = x is alocal optimum. Equivalently, if x = x is not a local optimum then at least one of thesufficiency conditions must not be true at x. It is easy to get all of this confused, sotake some time to think it through and understand it before proceeding further.

We are going to start by making a very general statement about what necessarily istrue at some local optimum x = x. This will be our general “statement of necessity.”

Statement of Necessity. If f is locally maximized on C at x = x, then the valueof f at x is at least as large as the value of f at every point x+Δx in C that is closeto x. Put more formally, if x is a local maximizer for f on C, then necessarily thereis an ε > 0 for which

f(x+Δx) ≤ f(x) ∀ x+Δx ∈ BdE(x, ε) ∩ C. (10.4)

If (10.4) is false, then x is not a local maximizer.

f, g1, . . . , gm are continuously differentiable with respect to x1, . . . , xn, so, for smallenough ε (and thus for small enough perturbations Δx), we can use the first-orderversion of Taylor’s Theorem to write

f(x+Δx) ≈ f(x) +∇f(x)·Δx (10.5)

and, for each j = 1, . . . ,m, to write

gj(x+Δx) ≈ gj(x) +∇gj(x)·Δx. (10.6)

(10.5) and (10.6) let us rewrite (10.4) as: If x is a local maximizer for f on C, then

∇f(x)·Δx ≤ 0 for all Δx for which ∇gj(x)·Δx ≤ 0 ∀ j ∈ I(x), (10.7)

where I(x) is the set of the indices of the constraints that bind at x = x and thesize (i.e. norm) of any perturbation Δx is less than ε. Statement (10.7) was usedto derive the first-order necessary optimization conditions explained in Chapter 8.A single statement of necessity results in a single complete list of conditions thatindividually are necessarily true when x is a local optimum.

Page 323: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

304 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

Statements of Sufficiency. Notice the use of the plural; i.e. “statements,” not“statement.” Typically we can find more than one set of conditions that ensure somepoint is a local optimum. For example, I could make the statement, “if x∗ is a globaloptimum, then x∗ is a local optimum.” I could also say that, “if the value at x∗ ∈ Cof the function f is larger than the values of f at every other feasible point x in aneighborhood about x∗, then x∗ is a local optimum for f on C.” I’m sure that youcan dream up a few more such sufficiency statements for yourself. My points are,first, that there can be more than one statement that is sufficient for x∗ to be a localoptimum and, second, that each such statement will typically consist of a differentcollection of conditions that together are sufficient for x∗ to be a local optimum. Wellthen, which one of the sufficiency statements shall we choose? The statement thatusually is employed is the following:

Particular Statement of Sufficiency. If the value of f at x is strictly larger thanthe value of f at every point x+Δx in C with Δ = 0 that is close to x, then x locallymaximizes f on C. Put more formally, if for some ε > 0, f(x∗ + Δx) < f(x∗) forevery perturbation Δx = 0, such that x∗ + Δx ∈ BdE(x

∗, ε) ∩ C, then x∗ is locallymaximal for f on C.

From this statement we will later in this chapter derive a particular set of condi-tions that jointly ensure that x∗ is a local optimum. I emphasize “jointly” because theconditions that are sufficient for local maximality are collectively but not individuallysufficient. Think of the simple problem of maximizing f(x) = −x2 on �. To solvethis problem most of us would compute the first-derivative, f ′(x) = −2x, and thensolve f ′(x) = 0 to obtain x∗ = 0. Next we would check that the value at x∗ = 0 of thesecond-order derivative, f ′′(0) ≡ −2, is negative. Would you assert that f ′′(x) < 0is sufficient for x = x to be a local optimum for f(x) = −x2? I hope not, since yourassertion would be wrong (after all, f ′′(x) = −2 at, for example, x = 1 also, butx = 1 is not a local optimum). A correct statement is, “f ′(x∗) = 0 and f ′′(x∗) < 0are together sufficient for x∗ = 0 to be a local maximum for f(x) = −x2.” Thuswe do not usually talk of an individual condition as being enough to ensure that apoint x∗ is a local optimum. Instead we talk of a set of conditions that together aresufficient for local optimality.

10.2 Second-Order Necessary Conditions

Let’s briefly review some basic calculus. Suppose you have a twice-differentiablefunction f : � �→ � with a local maximum at x = x∗. You are asked to locate x∗.

Page 324: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.2. SECOND-ORDER NECESSARY CONDITIONS 305

There are no constraints. How would you do it? I imagine that you would obtainthe function’s first-order derivative f ′(x) and solve f ′(x) = 0. The solutions to thisequation are the critical points of the function. Let’s suppose there is only one criticalpoint, x = x. I imagine further that you would then obtain the function’s second-orderderivative f ′′(x) and evaluate it at x = x. If f ′′(x) < 0, then you would conclude thatx is x∗, the value of x that locally maximizes the value of f . If that is a reasonabledescription of what you would do, then I ask you now to consider carefully what isthe complete list of conditions that necessarily, individually are true when x = x is alocal maximum for f . What is this list?

We know that f ′(x) = 0 is a necessary condition. So is f ′′(x) ≤ 0. Why? Supposeinstead that f ′′(x) > 0. Then x is not a local maximum for f , is it? Hence, if xis a local maximizer for f , then necessarily f ′′(x) > 0. The reason becomes clearwhen we use a second-order Taylor’s series (this is why we require order-2 continuousdifferentiability for f) to approximate f(x +Δx) in our necessity statement; i.e. wewrite f(x+Δx) ≤ f(x) as

f(x) + f ′(x)Δx+1

2f ′′(x) (Δx)2 ≤ f(x).

Since necessarily f ′(x) = 0, the local maximality of x tells us that necessarily

f ′′(x) ≤ 0. (10.8)

What if f : �n �→ �? Again, using a second-order Taylor’s series lets us writef(x+Δx) ≤ f(x), for all x+Δx ∈ BdE(x, ε) and small ε > 0, as

f(x) +∇f(x)·Δx+ 1

2ΔxTHf (x)Δx ≤ f(x).

Necessarily ∇f(x) = 0, so it is also necessary that

ΔxTHf (x)Δx ≤ 0 ∀ Δx such that x+Δx ∈ BdE(x, ε). (10.9)

Hf is the Hessian matrix of f . ΔxTHf (x)Δx is the quadratic function of Δx1, . . . ,Δxnthat is called the “quadratic form of the matrix Hf .” The matrix is the orderedcollection of all of the second-order derivatives of f ; i.e.

Hf (x) =

⎛⎜⎜⎜⎜⎜⎝∂2f(x)

∂x21· · · ∂2f(x)

∂x1∂xn...

. . ....

∂2f(x)

∂xn∂x1· · · ∂2f(x)

∂x2n

⎞⎟⎟⎟⎟⎟⎠ . (10.10)

Page 325: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

306 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

A matrix with a quadratic form that never takes a positive value is called a negative-semi-definite matrix, so (10.9) says that, if x is a local maximizer for f , then neces-sarily the Hessian matrix of f evaluated at x = x is negative-semi-definite. (10.8) isexactly this statement for n = 1.

(10.9) has a simple meaning. The necessary condition ∇f(x) = 0 says that at xthe graph of f is flat in every direction. (10.9) says that, if x is a local maximum, thennecessarily in a small neighborhood around x the graph of f must not in any direction“curl upwards.” Look at Figure 10.1. There, x = (1, 1) is a local maximizer for f .The slope of the graph of f in any direction is zero at x and for any small movementin any direction away from the point x the graph of f does not curl upwards.

Figure 10.1: The first-order and second-order necessary conditions that are impliedby f being maximized at x = (1, 1).

Now look at Figure 10.2. Again the slope of the graph of f at x = (1, 1) is zeroin any direction, but this time in every small neighborhood around x the graph curlsupwards in at least one direction. Thus x cannot be a maximizer for f .

All of our discussion so far has been of unconstrained optimization problems.Let’s now examine constrained optimization problems. Our problem is again (10.1).

Page 326: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.2. SECOND-ORDER NECESSARY CONDITIONS 307

Figure 10.2: f satisfies the first-order necessary condition but not the second-ordernecessary condition at x = (1, 1), which is therefore not a local maximum for f .

Imagine that we have solved the first-order optimality conditions and that x is oneof the solutions. Again use I(x) to denote the set of the indices of the constraintsthat bind at x. If I(x) = ∅ and x is a local maximizer for f , then we are dealingwith an unconstrained solution, and the necessary local optimality conditions are just∇f(x) = 0 and (10.9), so consider the case in which at least one constraint binds atx; i.e. I(x) = ∅. Without loss of generality, suppose that the first k constraints bindat x; i.e. I(x) = {1, . . . , k} with 1 ≤ k ≤ m.

For a constrained optimization problem such as (10.1), our necessity statement(10.4) becomes, “If x is a local maximizer for f subject to g1(x) ≤ b1, . . . , gm(x) ≤ bm,then necessarily there is an ε > 0 for which

f(x+Δx) ≤ f(x) ∀ Δx such that x+Δx ∈ BdE(x, ε)

and gj(x+Δx) ≤ bj ∀ j = 1, . . . ,m.”(10.11)

The last m−k constraints are slack at x; i.e. gj(x) < bj for j = k+1, . . . ,m. We areconsidering only problems in which each constraint function is continuous with respectto x, so there are values for ε that are so close to zero that the neighborhood BdE(x, ε)is so small that every constraint that is slack at x is also slack for every feasible point

Page 327: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

308 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

x + Δx ∈ BdE(x, ε). Within this small neighborhood, then, we can ignore the slackconstraints. Hence (10.11) informs us that the following, more special, statement isalso true: “If x is a local maximizer for f subject to g1(x) ≤ b1, . . . , gm(x) ≤ bm, thennecessarily there is an ε > 0 for which

f(x+Δx) ≤ f(x) ∀ Δx such that x+Δx ∈ BdE(x, ε)

and gj(x+Δx) ≤ bj ∀ j = 1, . . . , k.”(10.12)

For small ε, a second-order Taylor’s series for f lets us write (10.12) as

∇f(x)·Δx+ 1

2ΔxTHf (x)Δx ≤ 0 ∀ Δx such that x+Δx ∈ BdE(x, ε)

and gj(x+Δx) ≤ bj ∀ j = 1, . . . , k.(10.13)

In this constrained case, the Karush-Kuhn-Tucker condition is not ∇f(x) = 0. It is

∇f(x) =m∑j=1

λj∇gj(x) =k∑

j=1

λj∇gj(x) (because λk+1 = · · · = λm = 0), (10.14)

where λ1, . . . , λk are the values of the multipliers for the binding constraints obtainedfrom the Karush-Kuhn-Tucker and complementary slackness conditions evaluatedat x. Using (10.14) in (10.13) gives us, “If x is a local maximizer for f subject tog1(x) ≤ b1, . . . , gm(x) ≤ bm, then necessarily there is an ε > 0 for which

λ1∇g1(x)·Δx+ · · ·+ λk∇gk(x)·Δx+1

2ΔxTHf (x)Δx ≤ 0 (10.15)

∀ Δx such that x+Δx ∈ BdE(x, ε) and gj(x+Δx) ≤ bj ∀ j = 1, . . . , k.” (10.16)

(10.15) necessarily is true for two types of perturbations Δx: those that satisfy(10.16) with < and those that satisfy (10.16) with =. From here onwards we willconsider only the perturbations that satisfy (10.16) with equality.

Second-order Taylor’s series approximations let us write the feasibility-preservingrequirements gj(x+Δx) = gj(x) = bj for j = 1, . . . , k as

gj(x+Δx) = gj(x) +∇gj(x)·Δx+1

2ΔxTHgj(x)Δx = gj(x) = bj for j = 1, . . . , k.

That is,

∇gj(x)·Δx = − 1

2ΔxTHgj(x)Δx for each j = 1, . . . , k. (10.17)

Page 328: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.2. SECOND-ORDER NECESSARY CONDITIONS 309

So (10.15) and (10.17) give us the still more special necessity statement, “If x isa local maximizer for f subject to g1(x) ≤ b1, . . . , gm(x) ≤ bm, then necessarily thereis an ε > 0 for which

ΔxT[Hf (x)− λ1Hg1(x)− · · · − λkHgk(x)

]Δx ≤ 0 ∀ Δx such that (10.18)

x+Δx ∈ BdE(x, ε) and gj(x+Δx) = bj for j = 1, . . . , k.” (10.19)

The matrix inside the square brackets in (10.18) is the Hessian matrix (with respectto x1, . . . , xn only) of the restricted Lagrange function

Lr(x1, . . . , xn; λ1, . . . , λk) = f(x1, . . . , xn) +k∑

j=1

λj(bj − gj(x1, . . . , xn)) (10.20)

for problem (10.1) with the multipliers fixed at the values that satisfy the Karush-Kuhn-Tucker condition at x; i.e. λj = λj for j = 1, . . . , k and λj = 0 for j =k + 1, . . . ,m.

Why is (10.18) a statement that the function Lr(x + Δx; λ) does not locallyincrease its value above its value Lr(x; λ) at x for perturbations satisfying (10.19)?To a second-order approximation (remember, the multiplier values are fixed),

Lr(x+Δx; λ) = Lr(x; λ) +∇xLr(x; λ)|x=x·Δx+

1

2ΔxTHLr(x; λ)|x=xΔx. (10.21)

The Karush-Kuhn-Tucker condition is necessarily satisfied at x; i.e.

∇f(x) =k∑

j=1

λj∇gj(x) ⇔ ∇xLr(x; λ)|x=x = 0n. (10.22)

Thus the statement that Lr(x+Δx; λ) ≤ Lr(x; λ) implies both of the statements

∇xLr(x; λ)|x=x = 0 and ΔxTHLr(x; λ)|x=xΔx ≤ 0. (10.23)

The second of these two statements is exactly (10.18), since

Hf (x)− λ1Hg1(x)− · · · − λkHgk(x) = HLr(x; λ). (10.24)

Page 329: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

310 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

Written out in full, this square order-n matrix is

HLr(x1, . . . , xn; λ1, . . . , λk) =

⎛⎜⎜⎜⎜⎜⎝∂2Lr(x; λ)

∂x21. . .

∂2Lr(x; λ)

∂xn∂x1...

. . ....

∂2Lr(x; λ)

∂x1∂xn. . .

∂2Lr(x; λ)

∂x2n

⎞⎟⎟⎟⎟⎟⎠

=

⎛⎜⎜⎜⎜⎜⎝∂2f(x)

∂x21−

k∑j=1

λj∂2gj(x)

∂x21. . .

∂2f(x)

∂xn∂x1−

k∑j=1

λj∂2gj(x)

∂xn∂x1...

. . ....

∂2f(x)

∂x1∂xn−

k∑j=1

λj∂2gj(x)

∂x1∂xn. . .

∂2f(x)

∂x2n−

k∑j=1

λj∂2gj(x)

∂x2n

⎞⎟⎟⎟⎟⎟⎠ . (10.25)

Let’s be clear what we have discovered so far. It is that the original necessitystatement (10.11) implies the more specialized statement that, “If x is a local maxi-mizer for f subject to g1(x) ≤ b1, . . . , gm(x) ≤ bm, then necessarily there is an ε > 0for which

ΔxTHLr(x; λ)|x=xΔx ≤ 0 ∀ Δx such that (10.26)

x+Δx ∈ BdE(x, ε) and gj(x+Δx) = bj for all j = 1, . . . , k, (10.27)

where Lr(x; λ) is given by (10.20).”Be careful not to interpret (10.26) and (10.27) as stating that the Hessian matrix

of the optimization problem’s Lagrange function necessarily is negative-semi-definiteat a local constrained optimum x. Such a statement is wrong for two reasons. First,the problem’s Lagrange function L(x1, . . . , xn, λ1, . . . , λm) is a function in which all ofx1, . . . , xn and λ1, . . . , λm are variable. The restricted Lagrange function mentionedin (10.26) is a function only of x1, . . . , xn, since the values of λ1, . . . , λk, λk+1, . . . , λmare fixed at λ1, . . . , λk, 0, . . . , 0. Second, (10.26) is necessarily true only for the s-mall enough perturbations Δx that satisfy (10.27). A statement that the Hessianof Lr(x; λ) is a negative-semi-definite matrix requires that (10.26) be true for everypossible Δx, whether it satisfies (10.27) or not.

Keep in mind that the necessary second-order condition (10.26) depends upon thenecessary first-order Karush-Kuhn-Tucker condition and also upon the complemen-tary slackness conditions. (10.26) depends upon the Karush-Kuhn-Tucker conditionbecause it was used to rewrite the necessity statement from (10.13) to (10.15). (10.26)

Page 330: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.3. GEOMETRY OF THE SECOND-ORDER NECESSARY CONDITION 311

depends upon the complementary slackness conditions because they, along with theKarush-Kuhn-Tucker condition, were used to compute the values λ1, . . . , λk.

What we have learned in this section provides the following theorem.

Theorem 10.1 (Necessary Second-Order Local Optimality Condition). Let x be alocally optimal solution to problem (10.1). Let I(x) = ∅ be the set of indices of theconstraints that strictly bind at x. Let λj for j ∈ I(x) be the solution at x to theKarush-Kuhn-Tucker condition

∇f(x) =∑j∈I(x)

λj∇gj(x)

and the complementary slackness conditions

λj(bj − gj(x)) = 0 for j ∈ I(x).

Let

Lr(x1, . . . , xn; λ1, . . . , λk) = f(x1, . . . , xn) +∑j∈I(x)

λj(bj − gj(x1, . . . , xn)).

Then necessarily there is an ε > 0, such that ΔxTHLr(x; λ)|x=xΔx ≤ 0 for all Δxsuch that x+Δx ∈ BdE(x, ε) and gj(x+Δx) = bj ∀ j ∈ I(x).

This second-order necessary local optimality condition has many valuable uses, aswe will later see.

10.3 The Geometry of the Second-Order

Necessary Condition

What is the geometric interpretation of the second-order necessary condition (10.26)and (10.27)? Until now we have said that a point x is a local maximum for a functionf only if the slopes of f at x in every direction are all zero and only if the graph of fdoes not “curl upwards” in any direction for a small neighborhood around x. But thisstatement was made for an unconstrained optimization problem. It does not applyto a constrained optimization problem.

Some examples will be instructive. Consider, first, a problem we have met before:

maxx1, x2

f(x1, x2) = x21x22 subject to g1(x1) = −x1 ≤ 0,

g2(x2) = −x2 ≤ 0 and g3(x1, x2) = x1 + x2 ≤ 2.(10.28)

Page 331: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

312 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

Figure 10.3: The second-order necessary condition for problem (10.28).

The panels in Figure 10.3 display the problem. x = (1, 1) is a locally optimal solution.Only the third constraint binds at x = (1, 1), so the Karush-Kuhn-Tucker conditionis

∇f(x) = (2x1x22, 2x

21x2) = (2, 2) = λ3∇g3(x) = λ3(1, 1) ⇒ λ3 = 2 > 0. (10.29)

The perturbations Δx = (Δx1,Δx2) that preserve feasibility and keep the bindingconstraint binding in a small neighborhood about (1, 1) must satisfy

b3 = 2 = g3((1, 1) + Δx) = g3(1, 1) +∇g3(1, 1)·Δx+1

2ΔxTHg3(x)|x=(1,1)Δx,

which, since g3(1, 1) = 2, ∇g3(1, 1) = (1, 1) and Hg3(1, 1) = ( 0 00 0 ), is the requirement

that each Δx satisfies

0 = Δx1 +Δx2 ⇒ Δx1 = −Δx2. (10.30)

The second-order necessary local maximality condition (10.18) is therefore

(Δx1,Δx2)

((2x22 4x1x24x1x2 2x21

) ∣∣∣∣x=(1,1)

− λ3

(0 00 0

) ∣∣∣∣λ3=2

)(Δx1Δx2

)

= (Δx1,Δx2)

(2 44 2

)(Δx1Δx2

)= 2(Δx21 + 4Δx1Δx2 +Δx22

)(10.31)

= − 4 (Δx1)2 ≤ 0 for Δx1 = −Δx2 (see (10.30)). (10.32)

Page 332: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.3. GEOMETRY OF THE SECOND-ORDER NECESSARY CONDITION 313

What does it mean? Around (1, 1), the feasibility-preserving perturbations that keepthe constraint x1 + x2 = 2 binding are perturbations from (1,1) only along the linex1 + x2 = 2; thus Δx1 = −Δx2. Think, then, of a small interval on that line around(1, 1), as is indicated in both panels of Figure 10.3 by the short thicker lines. For thepoints (x1, x2) in this interval, the equation of the objective function is

ψ(x1) ≡ f(x1, x2 = 2− x1) = x21(2− x1)2 = x41 − 4x31 + 4x21. (10.33)

The values at (x1, x2) = (1, 1) of the first-order and second-order derivatives of thisfunction are ψ′(x1) = 4x31−12x21+8x1 = 0 for x1 = 1 and ψ′′(x1) = 12x21−24x1+8 =−4 < 0 for x1 = 1. Thus, over this small interval, the first-order necessary and thesecond-order necessary local optimality conditions are true. Notice that the Lagrangefunction

L(x1, x2, λ1, λ2, λ3) = x21x22 + λ1x1 + λ2x2 + λ3(2− x1 − x2)

becomesLr(x1, x2; 0, 0, 2) = x21x

22 + 2(2− x1 − x2)

once the function is restricted by setting λ1 = λ1 = 0, λ2 = λ2 = 0, and λ3 = λ3 = 2.Thus the restricted Lagrange function

Lr(x1, x2; 0, 0, 2) ≡ ψ(x1)

on the set {(x1, x2) | x1 + x2 = 2}.Now consider a new problem that is similar to the first, differing only in that the

third constraint has been changed:

maxx1, x2

f(x1, x2) = x21x22 subject to g1(x1) = −x1 ≤ 0,

g2(x2) = −x2 ≤ 0, and g3(x1, x2) = x21 + x22 ≤ 2.(10.34)

The panels in Figure 10.4 display the problem. x = (1, 1) is again a locally optimalsolution. Only the third constraint binds at x, so the Karush-Kuhn-Tucker conditionis

∇f(x) = (2x1x22, 2x

21x2) = (2, 2) = λ3∇g3(x) = λ3(2x1, 2x2) = λ3(2, 2)

⇒ λ3 = 1 > 0. (10.35)

The feasibility-preserving perturbations Δx = (Δx1,Δx2) that keep the third con-straint binding in a small neighborhood about (1, 1) must again satisfy

b3 = 2 = g3((1, 1) + Δx) = g3(1, 1) +∇g3(1, 1)·Δx+1

2ΔxTHg3(x)|x=(1,1)Δx,

Page 333: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

314 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

Figure 10.4: The second-order necessary condition for problem (10.34).

which, since g3(1, 1) = 2, ∇g3(1, 1) = (2, 2), and Hg3(1, 1) ≡ ( 2 00 2 ), is now the require-

ment that each Δx satisfies

0 = 2Δx1 + 2Δx2 + (Δx1)2 + (Δx2)

2 . (10.36)

The second-order necessary local maximality condition (10.18) is that

Q = (Δx1,Δx2)

((2x22 4x1x24x1x2 2x21

) ∣∣∣∣x=(1,1)

− λ3

(2 00 2

) ∣∣∣∣λ3=1

)(Δx1Δx2

)

= (Δx1,Δx2)

(0 44 0

)(Δx1Δx2

)= 8Δx1Δx2 ≤ 0 (10.37)

for every small enough perturbation (Δx1,Δx2) that satisfies (10.36). Is this so?Rearranging (10.36) gives

Δx2 = −1 +√2− (Δx1 + 1)2. (10.38)

Using this for Δx2 in (10.37) gives

Q = 8Δx1

(−1 +

√2− (Δx1 + 1)2

).

Page 334: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.3. GEOMETRY OF THE SECOND-ORDER NECESSARY CONDITION 315

It is easy to see that Q = 0 for Δx1 = 0 and that Q < 0 for Δx1 ∈ [−1, 0)∪(0,√2−1],

so there do exist small enough, strictly positive values for ε for which the second-ordernecessary condition is satisfied. Another way to verify this is to recognize that theobjective function defined only on the set {(x1, x2) | x21 + x22 = 2} is the function

θ(x1) ≡ f

(x1, x2 =

√2− x21

)= x21(2− x21).

θ′(x1) = 4x1(1−x21) = 0 for x1 = 0 and x1 = 1 (x1 = −1 violates the first constraint).But θ′′(x1) = 4(1− 3x21) > 0 for x1 = 0, while θ′′(1) < 0. Thus θ(x1) achieves a localmaximum at x1 = 1 (thus x2 = 1) on the set {(x1, x2) ∈ �2

+ | x21 + x22 = 2}. Becauseat x = (1, 1) the Lagrange multiplier values are λ1 = 0, λ2 = 0, and λ3 = 1, therestricted Lagrange function is

Lr(x1, x2; λ1, λ2, λ3) = x21x22 + 2− x21 − x22 ≡ θ(x1)

on the set {(x1, x2) ∈ �2+ | x21 + x22 = 2}.

The necessary conditions (10.32) and (10.37) are not the same. Why not? Afterall, the two problems have the same objective function f(x1, x2) = x21x

22, the same

locally optimal solution x = (1, 1), only one constraint binds, and the gradients of thebinding constraint in each example point in exactly the same direction at x = (1, 1).The second-order necessary condition for the first problem tells us that the objectivefunction f(x1, x2) = x21x

22 restricted to points (x1, x2) = (1 + Δx1, 1 + Δx2) on the

path x1 + x2 = 2 necessarily is weakly concave in a small enough neighborhood onthis path about the point (1, 1). The second-order necessary condition for the secondproblem tells us that the objective function f(x1, x2) = x21x

22 restricted to points

(x1, x2) = (1 + Δx1, 1 + Δx2) on the path x21 + x22 = 2 necessarily is weakly concavein a small enough neighborhood on this different path about the point (1, 1). Theleft-hand panel of Figure 10.3 displays the objective function’s graph over the pathx1 + x2 = 2. The left-hand panel of Figure 10.4 displays the objective function’sgraph over the path x21 + x22 = 2. Clearly the two graphs are different, even thoughthey both achieve the same maximum value f(1, 1) = 1 at the same point (1, 1).

It is sometimes asserted that the necessary second-order condition says that theobjective function of the constrained optimization problem is locally (at least weakly)concave at a local optimum. This is false, as the two above examples demonstrate.For either problem, look at the graph of the objective function above the line fromthe origin to the point (1, 1). This is the line along which x1 = x2, so the objectivefunction has the equation f(x1) = x41 along this path. This is a strictly convex

Page 335: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

316 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

function, including at and near to the point (1, 1). What is true is that, “at a localoptimum x, the objective function necessarily is locally (at least weakly) concavealong any path from x that keeps binding all of the constraints that bind at x.”

Another false assertion is that the necessary second-order condition says that therestricted Lagrange function of the constrained optimization problem is locally (atleast weakly) concave at a local optimum. Both of the above examples demonstratethat this is not true. Once again, for either problem, look at the graph of the objectivefunction above the line x1 = x2 from the origin to the point (1, 1). In the first example,the restricted Lagrange function Lr(x1, x2; λ) with x1 = x2 is

Lr(x1, x1; 0, 0, 2) = x41 + 2(2− 2x1) = x41 − 4x1 + 4.

This is a strictly convex function for any x1 > 0, including at and near to x1 = 1.Similarly, in the second example, the restricted Lagrange function Lr(x1, x2; λ) withx1 = x2 is

Lr(x1, x1; 0, 0, 1) = x41 + 2− 2x21.

This too is a strictly convex function at and near to x1 = 1. Again, what is true isthat, “at a local optimum x, the problem’s restricted Lagrange function Lr(x1, x2; λ)necessarily is locally (at least weakly) concave along any path from x that keepsbinding all of the constraints that bind at x.”

Why are the above two statements in quotation marks so similar even though thefirst refers to the problem’s objective function and the second refers to the problem’srestricted Lagrange function? It is because the two statements are the same state-ment. Along any path from x that keeps binding all of the constraints that bind atx, the problem’s objective function f(x1, x2) and the problem’s restricted Lagrangefunction Lr(x1, x2; λ) are the same. Why?

A third example will further improve your understanding of the necessary second-order condition. The problem is:

maxx1, x2

f(x1, x2) = (x1 + 1)2 + (x2 + 1)2 subject to

g1(x1) = −x1 ≤ 0, g2(x2) = −x2 ≤ 0, and g3(x1, x2) = x21 + x22 ≤ 2.(10.39)

Figure 10.5 displays the problem. x = (1, 1) is a locally optimal solution. Clearlyλ1 = λ2 = 0. The Karush-Kuhn-Tucker condition is

∇f(1, 1) = (2(x1 + 1), 2(x2 + 1))|x=(1,1) = (4, 4) = λ3∇g3(1, 1)= λ3(2x1, 2x2)|x=(1,1) = λ3(2, 2).

Page 336: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.3. GEOMETRY OF THE SECOND-ORDER NECESSARY CONDITION 317

Figure 10.5: The second-order necessary condition for problem (10.39).

Thus λ3 = 2 > 0. The restricted Lagrange function is

Lr(x; λ) = Lr(x1, x2; 0, 0, 2) = (x1 + 1)2 + (x2 + 1)2 + 2(2− x21 − x22).

The Hessian matrix of this function is constant-valued on �2+, and is

HLr(x; λ) ≡(−2 00 −2

).

The matrix is negative-definite (thus negative-semi-definite), so the necessary second-order condition is satisfied. The objective function f(x1, x2) = (x1+1)2+(x2+1)2 isstrictly convex everywhere on �2

+, so this example makes the point that the necessarysecond-order condition does not imply that the constrained optimization problem’sobjective function must be even weakly concave. However, the problem’s objectivefunction constrained to the set {(Δx1,Δx2) | (1+Δx1)

2+(1+Δx2)2 = 2} is concave.

Let’s see why. From (10.38),

Γ(Δx1) ≡ f(1 + Δx1, 1 + Δx2) = (2 + Δx1)2 + (2 + Δx2)

2

= 6 + 2Δx1 + 2√

2− (1 + Δx1)2.

Page 337: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

318 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

The first-order derivative of this function is

Γ′(Δx1) = 2− 2(1 + Δx1)√2− (1 + Δx1)2

.

Γ′(Δx1) = 0 for Δx1 = 0. The second-order derivative of Γ is

Γ′′(Δx1) = − 4

(2− (1 + Δx1)2)3/2< 0

for −1−√2 < Δx1 <

√2−1. Thus, as we saw in the first two examples, the problem’s

restricted Lagrange function, with the multipliers fixed at the values λ that satisfy theproblem’s first-order optimality conditions at x, necessarily is at least weakly concavein a neighborhood about the locally optimal point x when the function is confined tothe set of feasible points that keep binding all of the constraints that bind at x.

All of the examples so far have in common that only one constraint binds at thelocal optimum. The next, and last, example changes that. Figure 10.6 displays thesolution to the problem:

maxx1, x2

f(x1, x2) = x1x2 subject to g1(x1) = −x1 ≤ 0, g2(x2) = −x2 ≤ 0,

g3(x1, x2) = x1 + 2x2 ≤ 3, and g4(x1, x2) = 2x1 + x2 ≤ 3.(10.40)

Once again x = (1, 1) is a locally optimal solution. At this point both theconstraints g3 and g4 bind. Evaluated at (1, 1), the gradients of f , g3 and g4 are∇f(1, 1) = (1, 1), ∇g3(1, 1) = (1, 2), and ∇g4(1, 1) = (2, 1). The Karush-Kuhn-Tucker condition is thus

(1, 1) = (λ3, λ4)

(1 22 1

)⇒ λ3 = λ4 =

1

3.

The problem’s restricted Lagrange function Lr(x; λ) for this local optimum is therefore

Lr(x1, x2; 0, 0, 1/3, 1/3) = x1x2 +1

3(3− x1 − 2x2) +

1

3(3− 2x1 − x2)

= x1x2 − x1 − x2 + 2.

Evaluated at x = (1, 1), the Hessian matrix of this restricted Lagrange function is

HLr(x; λ)|x=(1,1) =

(0 11 0

).

Page 338: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.3. GEOMETRY OF THE SECOND-ORDER NECESSARY CONDITION 319

Figure 10.6: The only perturbation that keeps both constraints binding is (0, 0).

The second-order necessary condition is that there is an ε > 0 such that

(Δx1,Δx2)

(0 11 0

)(Δx1Δx2

)= 2Δx1Δx2 ≤ 0 (10.41)

∀ (Δx1,Δx2) satisfying (1 + Δx1, 1 + Δx2) ∈ BdE((1, 1), ε),

1 + Δx1 + 2(1 + Δx2) = 3 ⇒ Δx1 = −2Δx2, (10.42)

and 2(1 + Δx1) + 1 + Δx2 = 3 ⇒ 2Δx1 = −Δx2. (10.43)

The only values for Δx1 and Δx2 that satisfy (10.42) and (10.43) are Δx1 = Δx2 = 0.Thus there are no feasibility-preserving perturbations Δx to consider in this case –the necessary second-order condition is true, but only trivially. How would we knowwhen we are dealing with such a case? The answer is not complex. In our presentexample we have two linearly independent affine constraints g3 and g4 that bind at(1, 1). Since we have only two choice variables, x1 and x2, there is just one pointat which these two constraints bind, so the only feasibility-preserving perturbationsare the “non-perturbations” Δx1 = Δx2 = 0. In each of our three earlier examples,there are nontrivial perturbations that keep the binding constraint binding. If, asin our last example, the constraints that bind at a local optimum x do so only at

Page 339: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

320 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

a single point, then the only allowed feasibility-preserving perturbations are Δx1 =· · · = Δxn = 0, and the second-order necessary condition is true, but only triviallyand uninformatively.

10.4 Second-Order Necessary Condition Testing

Suppose we have solved the first-order optimality conditions for a constrained opti-mization problem and that x is one of the solutions. How can we check if the value atx = x of the Hessian matrix of the function Lr(x; λ) is negative-semi-definite for allsmall enough perturbations Δx from x that keep binding all of the constraints thatbind at x? You have already seen some simple examples of how to do it. In generalyou need, first, to find for a small enough ε > 0 the set of perturbations that keepbinding all of the constraints that bind at x. This set is

Δ(x, ε) ≡ {Δx ∈ �n | gj(x+Δx) = bj ∀ j ∈ I(x) and x+Δx ∈ BdE(x, ε)}.

Then you must determine if the function Lr(x+Δx; λ) is at least weakly concave onthe set Δ(x, ε). With complicated problems, you may need to do this numerically.But there is a special case in which this computation is relatively easy. It is the casein which all of the constraints that bind at a local optimum are linear functions ofx1, . . . , xn; i.e. each binding constraint gj is of the form gj(x) = aj1x1 + · · ·+ ajnxn,where aj1, . . . , ajn are constants. The requirement that a perturbation Δx keeps sucha constraint binding is

bj = gj(x+Δx) = gj(x) +∇gj(x)·Δx = bj +∇gj(x)·Δx⇒ ∇gj(x)·Δx = (aj1, . . . , ajn)·(Δx1, . . . ,Δxn) = 0. (10.44)

Hence, whenever all of the constraints that bind at a local optimum x are linearfunctions of x, statement (10.26) and (10.27) becomes, “If x is a local maximizer forf subject to g1(x) ≤ b1, . . . , gm(x) ≤ bm, then necessarily there is an ε > 0 for which

ΔxTHLr(x; λ)|x=xΔx ≤ 0 (10.45)

∀ Δx such that x+Δx ∈ BdE(x, ε), and ∇gj(x)·Δx = 0 for all j = 1, . . . , k.”(10.46)

The condition that ∇gj(x)·Δx = 0 for all j = 1, . . . , k is the same as Jg(x)Δx = 0k,where Jg(x) is the value at x = x of the Jacobian matrix formed from the gradients of

Page 340: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.4. SECOND-ORDER NECESSARY CONDITION TESTING 321

the constraints that bind at x = x. Checking if (10.45) is true subject to the condi-tions (10.46) is relatively easy thanks to a theorem we will see shortly. The theoremrelies upon determinants formed from a matrix called a bordered Hessian that we willdenote by H. H is the Hessian with respect to x1, . . . , xn of the restricted Lagrangefunction Lr(x; λ), augmented, on the left and top sides, by a border consisting ofthe gradient vectors of the constraints that bind at x = x. The border is thus theJacobian matrix consisting of the gradient vectors of the constraints that bind at x.If no constraint binds at x = x, then there is no border, and the “bordered” Hessianis the unbordered Hessian of the restricted Lagrange function.

Definition 10.1 (Bordered Hessian Matrix). Let x ∈ C. Let I(x) denote the set ofindices of the constraints g1, . . . , gm that bind at x. Let HLr(x; λ) denote the value atx = x of the bordered Hessian of Lr(x; λ).

If I(x) = ∅, then HLr(x; λ) = Hf (x). If I(x) = ∅, then let I(x) = {1, . . . , k} for1 ≤ k ≤ m. Then

HLr(x; λ) =

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

0 · · · 0∂g1(x)

∂x1· · · ∂g1(x)

∂xn...

. . ....

......

0 · · · 0∂gk(x)

∂x1· · · ∂gk(x)

∂xn∂g1(x)

∂x1· · · ∂gk(x)

∂x1

∂2Lr(x; λ)

∂x21· · · ∂2Lr(x; λ)

∂x1∂xn...

......

. . ....

∂g1(x)

∂xn· · · ∂gk(x)

∂xn

∂2Lr(x; λ)

∂xn∂x1· · · ∂2Lr(x; λ)

∂x2n

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠

=

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

0 . . . 0∂g1(x)

∂x1. . .

∂g1(x)

∂xn...

. . ....

......

0 . . . 0∂gk(x)

∂x1. . .

∂gk(x)

∂xn∂g1(x)

∂x1. . .

∂gk(x)

∂x1

∂2f(x)

∂x21−

k∑j=1

λj∂2gj(x)

∂x21. . .

∂2f(x)

∂xn∂x1−

k∑j=1

λj∂2gj(x)

∂xn∂x1...

......

. . ....

∂g1(x)

∂xn. . .

∂gk(x)

∂xn

∂2f(x)

∂x1∂xn−

k∑j=1

λj∂2gj(x)

∂x1∂xn. . .

∂2f(x)

∂x2n−

k∑j=1

λj∂2gj(x)

∂x2n

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠

,

(10.47)

Page 341: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

322 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

where all of the derivatives are evaluated at x = x.

The bordered Hessian is a square order-(k + n) matrix that is made up of fourblocks. The upper-left k × k block is the order-k zero matrix, 0k. The upper-rightk × n block is the Jacobian matrix Jg(x), evaluated at x = x, that is formed by thegradient vectors of the constraints that bind at x = x. The lower-left n× k block isthe transpose of this Jacobian, Jg(x)

T . The lower-right n× n block is HLr(x; λ), thevalue at x = x of the Hessian matrix of Lr(x; λ). It is often notationally convenientto write the bordered Hessian in its “block form,” which is

HLr(x; λ) =

⎛⎝ 0k Jg(x)

Jg(x)T HLr(x; λ)

⎞⎠∣∣∣∣

x=x

. (10.48)

Sometimes the bordered Hessian is written with the borders at the right side andthe bottom of the matrix; i.e.

HLr(x; λ) =

⎛⎝HLr(x; λ Jg(x)

T

Jg(x) 0k

⎞⎠∣∣∣∣

x=x

.

This form of the matrix is just as useful as (10.48), so choose whichever you wish. Ifyou choose this second form, then you will need to rewrite for yourself the determi-nants that I will introduce to you in a moment (doing so is a simple exercise that willhelp to develop your understanding).

It will help if we work with an example. Let’s use the problem:

maxx∈�3

f(x1, x2, x3) = x1x2x3 subject to g1(x1) = −x1 ≤ 0, g2(x2) = −x2 ≤ 0,

g3(x3) = −x3 ≤ 0, and g4(x1, x2, x3) = 3x1 + x2 + x3 ≤ 9. (10.49)

Verify for yourself that (x1, x2, x3) = (1, 3, 3) is a locally optimal solution to thisproblem, that only the fourth constraint binds at x = (1, 3, 3), and that the associatedvalues of the multipliers are λ1 = λ2 = λ3 = 0, and λ4 = 3. The value at x = (1, 3, 3)of the bordered Hessian matrix for this problem is therefore

HLr(x; λ) =

⎛⎜⎜⎝0 3 1 13 0 x3 x21 x3 0 x11 x2 x1 0

⎞⎟⎟⎠ =

⎛⎜⎜⎝0 3 1 13 0 3 31 3 0 11 3 1 0

⎞⎟⎟⎠ . (10.50)

Page 342: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.4. SECOND-ORDER NECESSARY CONDITION TESTING 323

Strange as it may seem, the properties of the bordered Hessian matrix describecompletely the necessary second-order local optimality condition (10.45) and (10.46).The information that we require to test the condition is provided by the signs ofparticular determinants formed from the elements of the bordered Hessian. Thesedeterminants are called the bordered principal minors of the bordered Hessian.

Definition 10.2 (Bordered Principal Minor). For i = 1, . . . , n, let πi = {l1, . . . , li}be any permutation of i integers from the set {1, . . . , n}. Let

A(πi) =

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

∂g1∂xl1

∂g1∂xl2

· · · ∂g1∂xli

∂g2∂xl1

∂g2∂xl2

· · · ∂g2∂xli

......

...

∂gk∂xl1

∂gk∂xl2

· · · ∂gk∂xli

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠.

Let

B(πi) =

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

∂2Lr(x; λ)

∂x2l1

∂2Lr(x; λ)

∂xl2∂xl1· · · ∂2Lr(x; λ)

∂xli∂xl1

∂2Lr(x; λ)

∂xl1∂xl2

∂2Lr(x; λ)

∂x2l2· · · ∂2Lr(x; λ)

∂xli∂xl2...

.... . .

...

∂2Lr(x; λ)

∂xl1∂xli

∂2Lr(x; λ)

∂xl2∂xli· · · ∂2Lr(x; λ)

∂x2li

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠.

The order-(k + i) determinant

M(πi) = det

(0k A(πi)

A(πi)T B(πi)

)(10.51)

is an order-(k + i) bordered principal minor of HLr(x; λ).

Consider again the bordered Hessian matrix (10.50). For this matrix, k = 1and n = 3, so i ∈ {1, 2, 3}. Let’s start with i = 1. The 3P1 = 3 ordered ways ofselecting one value from the set {1, 2, 3} are the permutations π1 = {1}, π1 = {2},

Page 343: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

324 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

and π1 = {3}. Thus there are three bordered principal minors of order-2:

M({1}) = det

⎛⎜⎜⎜⎝

0∂g4∂x1

∂g4∂x1

∂2Lr(x; λ)

∂x21

⎞⎟⎟⎟⎠∣∣∣∣x1=1x2=3x3=3

= det

(0 33 0

)= −9,

M({2}) = det

⎛⎜⎜⎜⎝

0∂g4∂x2

∂g4∂x2

∂2Lr(x; λ)

∂x22

⎞⎟⎟⎟⎠∣∣∣∣x1=1x2=3x3=3

= det

(0 11 0

)= −1,

and M({3}) = det

⎛⎜⎜⎜⎝

0∂g4∂x3

∂g4∂x3

∂2Lr(x; λ)

∂x23

⎞⎟⎟⎟⎠∣∣∣∣x1=1x2=3x3=3

= det

(0 11 0

)= −1.

(10.52)

For i = 2, there are 3P2 = 6 ordered ways of selecting two values from the set {1, 2, 3}.These are π2 = {1, 2}, π2 = {2, 1}, π2 = {1, 3}, π2 = {3, 1}, π2 = {2, 3}, andπ2 = {3, 2}. The bordered principal minor created from the permutation π2 = {3, 2}is

M({3, 2}) = det

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎝

0∂g4∂x3

∂g4∂x2

∂g4∂x3

∂2Lr(x; λ)

∂x23

∂2Lr(x; λ)

∂x3∂x2

∂g4∂x2

∂2Lr(x; λ)

∂x2∂x3

∂2Lr(x; λ)

∂x22

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎠∣∣∣∣∣x1=1x2=3x3=3

= det

⎛⎝0 1 11 0 11 1 0

⎞⎠ = 2.

(10.53)

Page 344: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.4. SECOND-ORDER NECESSARY CONDITION TESTING 325

The other five bordered principal minors of order-3 are:

M({1, 2}) = det

⎛⎝0 3 13 0 31 3 0

⎞⎠ = 18, M({2, 1}) = det

⎛⎝0 1 31 0 33 3 0

⎞⎠ = 18,

M({1, 3}) = det

⎛⎝0 3 13 0 31 3 0

⎞⎠ = 18, M({3, 1}) = det

⎛⎝0 1 31 0 33 3 0

⎞⎠ = 18,

and M({2, 3}) = det

⎛⎝0 1 11 0 11 1 0

⎞⎠ = 2.

(10.54)

For i = 3, there are 3P3 = 6 ordered ways to select three numbers from the set{1, 2, 3}, so there are six bordered principal minors of order-4, but each has the samevalue. Why? When i = n, any permutation is a sequence of exchanges of the positionsof full rows of the bordered Hessian accompanied by the same sequence of exchangesof the positions of full columns of the matrix. Each time you exchange the positionof any two rows, you change the sign, but not the size, of the determinant of theHessian. The same is true whenever you exchange the position of any two columnsof the matrix. Thus any such sequences of pairs of exchanges (one row exchangeaccompanied by one column exchange) leaves the value of the determinant of thebordered Hessian unchanged, so we need compute only the value of any one of theHessian’s order-4 bordered principal minors, such as

det

⎛⎜⎜⎝0 3 1 13 0 3 31 3 0 11 3 1 0

⎞⎟⎟⎠ = −27. (10.55)

How do we use the bordered principal minors to see if the necessary second-order local optimality condition (10.45) and (10.46) is true when all of the bindingconstraints are linear? The answer is supplied in the following theorem (see Debreu1952).

Theorem 10.2 (Second-Order Necessary Condition Test). Let x be locally optimal forthe constrained optimization problem (10.1). If the bordered Hessian matrix HLr(x; λ)is symmetric, then there exists ε > 0 such that the quadratic form ΔxTHLr(x; λ)Δx ≤0 for all perturbations Δx satisfying x+Δx ∈ BdE(x, ε) and ∇gj(x)·Δx = 0 for everyj ∈ I(x), if and only if, for all i = k+1, . . . , n, every bordered principal minor (10.51)of order-(k + i) is either zero or has the same sign as (−1)i.

Page 345: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

326 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

The theorem says that we do not need to check all of the bordered principalminors. Instead we must check the signs of only the minors with orders from 2k + 1to k + n. Recall that, in our example, k = 1 and n = 3. Thus we must check thesigns of only the minors of order-3 and order-4. The six bordered principal minors oforder-3 are all positive (see (10.53) and (10.54)) and so meet the condition of all beingeither zero or having the same sign as (−1)2. Also, all six bordered principal minorsof order-4 are negative and so meet the condition of being either zero or having thesame sign as (−1)3 (see (10.55)). Thus the second-order necessary local maximizationcondition is satisfied for (x1, x2, x3) = (1, 3, 3) (as it must be, since this is the globallyoptimal solution to problem (10.49)).

Note that Theorem 10.2 applies only to symmetric Hessian matrices. Recall thatwe have assumed that all of the functions f, g1, . . . , gm are twice-continuously differ-entiable with respect to x. This ensures that both the Hessian matrix H and thebordered Hessian matrix H are symmetric.

When all of the constraints that bind at x = x are linear with respect to x, youmay see the bordered Hessian matrix written as

H(x; λ) =

⎛⎝ 0k Jg(x)

JTg (x) Hf (x)

⎞⎠∣∣∣∣

x=x

, (10.56)

with the lower-right block being the Hessian matrix of the objective function f . Whenall of the binding constraints are linear in x, the restricted Lagrange function’s Hessianmatrix is the same as the Hessian matrix of the objective function, making (10.56)and (10.48) the same matrix.

10.5 Sufficient Conditions

The usual particular statement of conditions that together are sufficient for x to bea local maximizer for f on C is that every point x+Δx that is close to x, but not x,gives an objective function value f(x + Δx) that is strictly less than f(x). That is,“If for some ε > 0

f(x+Δx) < f(x) ∀ Δx = 0 such that x+Δx ∈ BdE(x, ε) (10.57)

and gj(x+Δx) ≤ bj ∀ j = 1, . . . ,m, (10.58)

then x is a local maximizer for f on C.”

Page 346: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.5. SUFFICIENT CONDITIONS 327

Again let I(x) be the set of the indices of the constraints that bind at x = x. IfI(x) = ∅, then, as you already know, the conditions

∇f(x) = 0 and ΔxTHf (x)Δx < 0 for all Δx = 0 satisfying x+Δx ∈ BdE(x, ε)

for some ε > 0 are jointly sufficient for x to be either an unconstrained local optimumfor f or for x to be an interior locally optimal solution to the constrained optimizationproblem (10.1).

If I(x) = ∅, then we can replicate the reasoning used in Section 10.2 to discover asufficiency statement that uses a local strict concavity requirement for the problem’srestricted Lagrange function. Again suppose that I(x) = {1, . . . , k}, where 1 ≤k ≤ m. As before we use second-order Taylor’s series approximations for each off, g1, . . . , gk to write (10.57) and (10.58) as, “If, for some ε > 0,

∇f(x)·Δx+ 1

2ΔxTHf (x)Δx < 0 ∀ Δx = 0 such that (10.59)

x+Δx ∈ BdE(x, ε) and ∇gj(x)·Δx+1

2ΔxTHgj(x)Δx ≤ 0 ∀ j = 1, . . . , k, (10.60)

then x is a local maximizer for f on C.”Suppose that x is a regular point in the feasible set. Then x cannot be a local

optimum unless the first-order conditions are satisfied at x, so any set of conditionsthat together are sufficient for x to be a local optimum must include that there existnonnegative values λ1, . . . , λm ≥ 0, not all zero, satisfying

∇f(x) =m∑j=1

λj∇gj(x) and λj(bj − gj(x)) = 0 for all j = 1, . . . ,m. (10.61)

We already know that the complementary slackness conditions require λj = 0 forevery j /∈ I(x). Thus we obtain from (10.59), (10.60), and (10.61) the statement that,“If there exist values λ1, . . . , λk ≥ 0, not all zero, satisfying (10.61) and if, for someε > 0,

k∑j=1

λj∇gj(x)·Δx+1

2ΔxTHf (x)Δx < 0 ∀ Δx = 0 such that (10.62)

x+Δx ∈ BdE(x, ε) and ∇gj(x)·Δx+1

2ΔxTHgj(x)Δx ≤ 0 ∀ j = 1, . . . , k, (10.63)

then x is a local maximizer for f on C.”

Page 347: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

328 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

Once we specialize by considering only perturbations Δx that keep binding theconstraints that bind at x, the sufficiency statement becomes, “If there exist valuesλ1, . . . , λk ≥ 0 satisfying (10.61) and if, for some ε > 0,

k∑j=1

λj∇gj(x)·Δx+1

2ΔxTHf (x)Δx < 0 ∀ Δx = 0 such that (10.64)

x+Δx ∈ BdE(x, ε) and ∇gj(x)·Δx = −1

2ΔxTHgj(x)Δx ∀ j = 1, . . . , k, (10.65)

then x is a local maximizer for f on C.” This is the same as the statement, “If thereexist values λ1, . . . , λk ≥ 0 satisfying (10.61), and if, for some ε > 0,

ΔxT[Hf (x)− λ1∇g1(x)− · · · − λk∇gk(x)

]Δx < 0 ∀ Δx = 0 such that (10.66)

x+Δx ∈ BdE(x, ε) and gj(x+Δx) = bj ∀ j = 1, . . . , k, (10.67)

then x is a local maximizer for f on C.”The effort put into understanding the necessary second-order local optimality

condition should make it not difficult for you to understand what this sufficiencystatement asserts. First, the statement is a collection of conditions. Remove any onefrom the collection and you may not have a set of conditions sufficient for x to belocally optimal. Second, all of the necessary conditions are an essential part of anyset of collectively sufficient conditions. If it is not clear to you that this must beso, then think about it. Third, condition (10.67) is not quite the same as condition(10.27). Why not? The collectively sufficient conditions include that Δx = 0, whereasΔx = 0 is allowed in the necessary second-order condition. Fourth, condition (10.66)is a demand that, in small enough neighborhoods about x, the restricted Lagrangefunction Lr(x; λ) is strictly concave at every point x +Δx for which the constraintsthat bind at x also all bind at x + Δx. Note carefully that the demand is not thatLr(x; λ) is strictly concave at every neighboring point x+Δx. It is only that Lr(x; λ)is strictly concave at all neighboring point x+Δx with Δx satisfying (10.67).

You will find it helpful to reconsider the examples (10.28), (10.34), and (10.39).Verify for each example that the entire collection (10.61), (10.66), and (10.67) ofsufficiency conditions is satisfied, and therefore that each of the points x mentionedin those examples truly is a local optimum. The set of sufficiency conditions does notapply to example (10.40) because, for that example, the only allowed perturbationis Δx = (0, 0), which does not satisfy condition (10.67), which uses only nonzeroperturbations. This does not imply that the point x = (1, 1) is not a local optimum(it is), since it is only a set of sufficiency conditions that does not apply in this case.

Page 348: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.6. SECOND-ORDER SUFFICIENT CONDITION TESTING 329

10.6 Second-Order Sufficient Condition Testing

What is the procedure for applying the conditions (10.61), (10.66), and (10.67) inan effort to verify that a point x is a local optimum? It is almost the same as theprocedure for testing the necessary second-order local optimality condition. Onceagain the first step is to verify that there are small enough values of ε > 0 forwhich there are nonempty sets of nonzero perturbations that keep binding all of theconstraints that bind at x. That is, you must verify that, for small enough values ofε > 0, the set

Δ(x, ε) ≡ {Δx ∈ �n | Δx = 0, gj(x+Δx) = bj ∀ j ∈ I(x), x+Δx ∈ BdE(x, ε)} = ∅.

Then you must determine if the function L(x+Δx; λ) is strictly concave with respectto Δx on the set Δ(x, ε). With more complicated problems, it may be necessary todo this numerically.

As before, in the special case in which the only constraints that bind at x arelinear with respect to x, the testing of (10.66) and (10.67) can be done relativelyeasily by computing the values of bordered principal minors of the bordered Hessianof the function Lr(x; λ). In fact, only a few of the bordered principal minors need tobe evaluated. These are the successive bordered principal minors of the borderedHessian matrix.

Definition 10.3 (Successive Principal Bordered Minor). For i = 1, . . . , n, the ith

successive bordered principal minor of the bordered Hessian matrix (10.47) is thedeterminant consisting of the first k + i row and column elements of the matrix; i.e.

M i = det

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

0 · · · 0∂g1∂x1

· · · ∂g1∂xi

.... . .

......

...

0 · · · 0∂gk∂x1

· · · ∂gk∂xi

∂g1∂x1

· · · ∂gk∂x1

∂2Lr(x; λ)

∂x21· · · ∂2Lr(x; λ)

∂x1∂xi...

......

. . ....

∂g1∂xi

· · · ∂gk∂xi

∂2Lr(x; λ)

∂xi∂x1· · · ∂2Lr(x; λ)

∂x2i

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠

. (10.68)

Page 349: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

330 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

In example (10.49), the n = 3 successive bordered principal minors are, from(10.50),

M1 = det

(0 33 0

)= −9, M2 = det

⎛⎝0 3 13 0 31 3 0

⎞⎠ = 18,

and M3 = det

⎛⎜⎜⎝0 3 1 13 0 3 31 3 0 11 3 1 0

⎞⎟⎟⎠ = −27. (10.69)

Theorem 10.3 (Second-Order Sufficient Optimality Condition Test). Let x be afeasible solution for problem (10.1). Let I(x) = ∅ be the set of the indices of theconstraints that bind at x. If

(i) for each j ∈ I(x), λj solves the first-order necessary conditions at x with eachλj ≥ 0 and not all zero, and if

(ii) the bordered Hessian matrix HLr(x; λ) is symmetric, then there exists ε > 0,such that ΔxTHLr(x; λ)Δx < 0 for all perturbations Δx = 0 satisfying x + Δx ∈BdE(x, ε) and ∇gj(x)·Δx = 0 for every j ∈ I(x), if and only if, for each i = k +1, . . . , n, the successive bordered principal minor (10.68) of order-(k + i) is not zeroand has the same sign as (−1)i.

That is, the last n − k successive bordered principal minors are all nonzero andalternate in sign, with the sign of the last minor being the sign of (−1)n. In ourexample, k = 1 and n = 3, so the only successive bordered principal minors that need

to be examined are for i = 2, 3. M2 = 18 is nonzero and has the same sign as (−1)2.

M3 = −27 is nonzero and has the same sign as (−1)3. Therefore the second-ordersufficiency condition for a local maximum is satisfied.

Recollect that this second-order condition is, by itself, not sufficient for a pointto be locally maximal. It is the set consisting of this second-order condition alongwith the Karush-Kuhn-Tucker and the complementary slackness conditions that col-lectively is sufficient for a point to be locally maximal.

This concludes our examination of the essentials of constrained optimization prob-lems in which the objective and constraint functions are differentiable. It is worthmentioning that the essential ideas developed here can be extended without too muchadditional effort to problems with objective and constraint functions that are not alldifferentiable. The basic tool is Minkowski’s Separating Hyperplane Theorem. Thegeometries of the solutions of these problems are similar to those we have exploredfor our differentiable problems.

Page 350: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.7. REMARKS 331

10.7 Remarks

Well, now what? What do we do with the knowledge we have gained so far intosolving constrained optimization problems? What we have focused upon so far arethe questions of when does a problem have a solution and, if so, then how do welocate it? Next on the to-be-answered list is the question, “Given that a constrainedoptimization problem has an optimal solution, what are the properties of the optimalsolution?” This is the theme of Chapters 11, 12, and 13. Chapter 11 examines in areasonably general way the properties of optimal solution mappings and maximum-value functions. This chapter relies upon continuity a lot, so you will need to befamiliar with the contents of Chapter 5 to understand all that is presented. Chap-ters 12 and 13 use calculus to explain how to derive differential descriptions of theproperties of optimal solution functions and maximum-value functions. These chap-ters rely heavily upon the second-order necessary maximization condition explainedin the current chapter.

Page 351: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

332 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

10.8 Problems

Problem 10.1. A firm wishes to minimize its cost of producing a specified quantityy ≥ 0 of its product. The firm’s production function is y = x

1/21 x

1/22 , where x1 and x2

denote the quantities used by the firm of inputs 1 and 2. The per-unit input pricesare w1 > 0 and w2 > 0. Thus the firm’s problem is:

minx1, x2

w1x1 + w2x2 subject to x1 ≥ 0, x2 ≥ 0 and y ≤ x1/21 x

1/22 .

Rewritten as a maximization problem, this is

maxx1, x2

−w1x1 − w2x2 subject to g1(x1) = −x1 ≤ 0, g2(x2) = −x2 ≤ 0,

and g3(x1, x2) = −x1/21 x1/22 ≤ −y.

The cost-minimizing input bundle, the optimal solution to this problem, is the con-ditional input demand pair consisting of

x∗1(w1, w2, y) =

(w2

w1

)1/2

y and x∗2(w1, w2, y) =

(w1

w2

)1/2

y. (10.70)

The first and second constraints are slack at the optimal solution. The third con-straint strictly binds. The multipliers are λ∗1(w1, w2, y) = 0, λ∗2(w1, w2, y) = 0 andλ∗3(w1, w2, y) = 2(w1w2)

1/2 > 0.(i) Let Δx1 and Δx2 denote small perturbations of the variables x1 and x2 from thevalues x∗1 and x∗2. In order to keep the third constraint binding, what relationshipmust exist between Δx1 and Δx2?(ii) What is the problem’s objective function when its domain is confined to the smallperturbations discovered in part (i)? Rewrite this mapping as a function of only Δx1and denote it by f .(iii) Show that f is an at least weakly concave function of Δx1, for small valuesof Δx1.(iv) Write down the problem’s Lagrange function. Then write down the restrictedLagrange function Lr(x1, x2;λ

∗1, λ

∗2, λ

∗3).

(v) Write down the Hessian matrix HLr(x1, x2;λ∗1, λ

∗2, λ

∗3) of the restricted Lagrange

function.(vi) Let Δx = (Δx1,Δx2)

T . Write down the value at (x∗1, x∗2) of the quadratic form

ΔxTHLr(x∗1, x∗2;λ

∗1, λ

∗2, λ

∗3)Δx of the Hessian matrix. Use your answer to part (i) to

write this quadratic as a function of only Δx1. Then show that this function is never

Page 352: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.8. PROBLEMS 333

positive for small values of Δx1 that satisfy the relationship discovered in part (i),thereby confirming the necessary second-order local maximality condition.(vii) Is the binding constraint function linear with respect to x1 and x2? Is it validto evaluate the bordered principal permutation minors of the bordered Hessian ma-trix for this problem in order to deduce whether or not the second-order necessarycondition is true at (x∗1, x

∗2)?

Problem 10.2. Consider the problem:

maxx1, x2, x3

f(x1, x2, x3) = x1x2x3 subject to g1(x1) = −x1 ≤ 0, g2(x2) = −x2 ≤ 0,

g3(x3) = −x3 ≤ 0, g4(x1, x2, x3) = 2x1 + x2 + x3 ≤ 18, and

g5(x1, x2, x3) = x1 + 2x2 + x3 ≤ 18.

The optimal solution is (x∗1, x∗2, x

∗3) = (4, 4, 6). The associated multiplier values are

λ∗1 = λ∗2 = λ∗3 = 0, λ∗4 = 8, and λ∗5 = 8.(i) Let Δx1, Δx2, and Δx3 denote small perturbations of the variables x1, x2, andx3 from the values x∗1 = 4, x∗2 = 4, and x∗3 = 6. In order to keep the fourth and fifthconstraints binding, what relationships must exist between Δx1, Δx2, and Δx3?(ii) What is the problem’s objective function when its domain is confined to the smallperturbations discovered in part (i)? Rewrite this mapping as a function of only Δx1and denote it by f .(iii) Show that f is an at least weakly concave function of Δx1, for small valuesof Δx1.(iv) Write down the problem’s Lagrange function. Then write down the restrictedLagrange function Lr(x1, x2, x3;λ

∗1, λ

∗2, λ

∗3, λ

∗4, λ

∗5).

(v) Write down the Hessian matrix HLr(x1, x2, x3;λ∗1, λ

∗2, λ

∗3, λ

∗4, λ

∗5) of the restricted

Lagrange function.(vi) Let Δx = (Δx1,Δx2,Δx3)

T . Write down the value at (x∗1, x∗2, x

∗3) of the quadratic

form ΔxTHLr(x1, x2, x3;λ∗1, λ

∗2, λ

∗3, λ

∗4, λ

∗5)Δx of the Hessian matrix. Use your answer

to part (i) to write this quadratic as a function of only Δx1. Then show that thisfunction is never positive for small values of Δx1, thereby confirming the necessarysecond-order local maximality condition.(vii) Explain why, for the present problem, it is valid to use Theorem 10.2 to evaluatethe necessary second-order local maximality condition.(viii) What is the value at (x∗1, x

∗2, x

∗3) = (4, 4, 6) of the problem’s bordered Hessian

matrix?(ix) Apply Theorem 10.2 to the bordered Hessian matrix to confirm the necessarysecond-order local maximality condition.

Page 353: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

334 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

Problem 10.3. Consider again problem 10.2. Use the Karush-Kuhn-Tucker condi-tion, the five complementary slackness conditions, and the second-order sufficiencycondition to verify that (x1, x2, x3) = (4, 4, 6) is a local maximum for the problem.

Problem 10.4. Consider again problem 8.6. Another quick look at Figure 8.18 willbe helpful. The optimal solution is the point (x∗1, x

∗2) = (4, 4). Confirm the second-

order necessary local maximality condition at this point. Can the second-order localmaximality sufficiency condition be verified?

Problem 10.5. Consider again problem 8.7. Take another peek at Figure 8.19. Weearlier established that the optimal solution is the point (x∗1, x

∗2) = (4·5, 6·75). At

this point, the constraint g1(x1, x2) = (x1 − 4)2 + x2 ≤ 7 binds and the constraintg2(x1, x2) = (x1 − 5)3 − 12x2 ≤ −60 is slack. Another feasible, not optimal, solutionis the point (x1, x2) ≈ (5·412, 5·006), at which both of the constraints bind.(i) Confirm that the necessary second-order condition is true at the optimal solution(x∗1, x

∗2) = (4·5, 6·75). I know – it must be true, but check it anyway, just for the

practice.(ii) Is the necessary second-order condition true at the feasible solution (x1, x2) ≈(5·412, 5·006)? What do you learn from this?

10.9 Answers

Answer to Problem 10.1.(i) At the optimal solution only the third constraint binds, so the second-order nec-essary condition considers only small perturbations Δx1 and Δx2 from (x∗1, x

∗2) to

nearby input bundles (x∗1 + Δx1, x∗2 + Δx2) for which the third constraint continues

to bind; that is, Δx1 and Δx2 satisfy

g3(x∗1 +Δx1, x

∗2 +Δx2) = g3(x

∗1, x

∗2) = y;

i.e. (x∗1 +Δx1)1/2 (x∗2 +Δx2)

1/2 = (x∗1)1/2 (x∗2)

1/2 = y.

Squaring followed by a little rearranging gives

Δx2 = − x∗2 Δx1x∗1 +Δx1

. (10.71)

(ii) Evaluated for input bundles (x∗1+Δx1, x∗2+Δx2), the problem’s objective function

isf(x∗1 +Δx1, x

∗2 +Δx2;w1, w2) = −w1(x

∗1 +Δx1)− w2(x

∗2 +Δx2).

Page 354: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.9. ANSWERS 335

Substitution from (10.71) then shows the value of the objective function at inputbundles where Δx1 and Δx2 satisfy (10.71) to be

f(Δx1; x∗1, x

∗2, w1, w2) = −w1x

∗1 − w2x

∗2 − w1Δx1 + w2x

∗2 ×

Δx1x∗1 +Δx1

. (10.72)

(iii) The second-order derivative of f with respect to Δx1 is

d2f(Δx1; x∗1, x

∗2, w1, w2)

dΔx21= − 2w2x

∗1x

∗2

(x∗1 +Δx1)3 = − 2w2y

2((w2

w1

)1/2

y +Δx1

)3 , (10.73)

after substitution from (10.70). If Δx1 is small, then this second-derivative is notpositive, confirming the second-order local maximality condition that we already knewhad to be true as a consequence of the local optimality of the input bundle (10.70).

(iv) The problem’s Lagrange function is

L(x1, x2, λ1, λ2, λ3;w1, w2, y) = −w1x1 − w2x2 + λ1x1 + λ2x2 + λ3

(−y + x

1/21 x

1/22

).

The Lagrange function restricted by λ1 = λ∗1 = 0, λ2 = λ∗2 = 0, and λ3 = λ∗3 =2(w1w2)

1/2y is

Lr(x1, x2;λ∗1, λ

∗2, λ

∗3, w1, w2, y) = −w1x1 − w2x2 + 2(w1w2)

1/2y(−y + x

1/21 x

1/22

).

(v) The Hessian matrix of the restricted Lagrange function is

HLr(x1, x2;λ∗1, λ

∗2, λ

∗3) =

(w1w2)1/2y

2

⎛⎝−x−3/2

1 x1/22 x

−1/21 x

−1/22

x−1/21 x

−1/22 −x1/21 x

−3/22

⎞⎠ .

(vi) For perturbations Δx1 and Δx2, the quadratic form of the Hessian matrix of therestricted Lagrange function is

Q(Δx1,Δx2) =(w1w2)

1/2y

2(Δx1,Δx2)

⎛⎝−x−3/2

1 x1/22 x

−1/21 x

−1/22

x−1/21 x

−1/22 −x1/21 x

−3/22

⎞⎠(Δx1

Δx2

)

=(w1w2)

1/2y

2

(−x−3/2

1 x1/22 (Δx1)

2 + 2x−1/21 x

−1/22 Δx1Δx2 − x

1/21 x

−3/22 (Δx2)

2).

Page 355: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

336 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

Evaluating the Hessian matrix at (x∗1, x∗2) and considering only the perturbations that

satisfy (10.71) gives us the function of Δx1 that is

Q(Δx1) =(w1w2)

1/2y

2

(− (x∗1)

−3/2(x∗2)1/2(Δx1)

2+

2(x∗1)−1/2(x∗2)

−1/2Δx1

(− x∗2 Δx1x∗1 +Δx1

)− (x∗1)

1/2(x∗2)−3/2

(− x∗2 Δx1x∗1 +Δx1

)2)

= − (w1w2)1/2y

2

(1 +

2x∗1x∗1 +Δx1

+(x∗1)

2

(x∗1 +Δx1)2

)(x∗1)

−3/2(x∗2)1/2(Δx1)

2.

This function is nonpositive for small values of Δx1, confirming the second-order localmaximality condition that we already knew necessarily to be true as a consequenceof the local optimality of the input bundle (10.70).(vii) Theorem 10.2 is valid for small perturbations Δx1 and Δx2 from x∗1 and x

∗2 that

satisfy the linear relationship between Δx1 and Δx2 that is ∇g3(x∗1, x∗2)·(Δx1,Δx2) =0. This relationship is

−1

2

((x∗2x∗1

)1/2

,

(x∗1x∗2

)−1/2)·(Δx1,Δx2) = 0 ⇒ Δx2 =

x∗2x∗1

Δx1. (10.74)

The correct relationship between Δx1 and Δx1 is the nonlinear relationship (10.71).Theorem 10.2 should not be used for the present problem.

Answer to Problem 10.2.(i) At the optimal solution only the fourth and fifth constraints bind, so the second-order necessary condition considers only small perturbations Δx1, Δx2 and Δx3 from(x∗1, x

∗2, x

∗3) to nearby input bundles (x∗1 + Δx1, x

∗2 + Δx2, x

∗3 + Δx3) for which the

fourth and fifth constraints continue to bind; i.e. Δx1, Δx2 and Δx3 satisfy

2(x∗1 +Δx1) + (x∗2 +Δx2) + (x∗3 +Δx3) = 2x∗1 + x∗2 + x∗3 = 18,

and (x∗1 +Δx1) + 2(x∗2 +Δx2) + (x∗3 +Δx3) = x∗1 + 2x∗2 + x∗3 = 18,

⇒ 2Δx1 +Δx2 +Δx3 = 0 and Δx1 + 2Δx2 +Δx3 = 0.

From these equations we obtain that the small perturbations to be considered satisfy

Δx2 = Δx1 and Δx3 = −3Δx1. (10.75)

(ii) Evaluated for points (x∗1 + Δx1, x∗2 + Δx2, x

∗3 + Δx3), the problem’s objective

function is

f(x∗1 +Δx1, x∗2 +Δx2, x

∗3 +Δx3) = (x∗1 +Δx1)(x

∗2 +Δx2)(x

∗3 +Δx3).

Page 356: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.9. ANSWERS 337

Substitution from (10.75) then shows the value of the objective function at pointswhere Δx1, Δx2 and Δx3 satisfy (10.75) to be

f(Δx1; x∗1, x

∗2, x

∗3) = (x∗1+Δx1)(x

∗2+Δx1)(x

∗3−3Δx1) = (4+Δx1)

2(6−3Δx1). (10.76)

(iii) The second-order derivative of f with respect to Δx1 is

d2f(Δx1; x∗1, x

∗2, x

∗3)

dΔx21= −18(2 + Δx1).

If Δx1 is small, then this second-derivative is not positive, thus confirming the second-order local maximality condition that must be true as a consequence of the localoptimality of the point (x∗1, x

∗2, x

∗3) = (4, 4, 6).

(iv) The problem’s Lagrange function is

L(x1, x2, λ1, λ2, λ3, λ4, λ5) = x1x2x3 + λ1x1 + λ2x2 + λ3x3 + λ4(18− 2x1 − x2 − x3)

+ λ5(18− x1 − 2x2 − x3).

The Lagrange function restricted by λ1 = λ∗1 = 0, λ2 = λ∗2 = 0, λ3 = λ∗3 = 0,λ4 = λ∗4 = 8, and λ5 = λ∗5 = 8 is

Lr(x1, x2;λ∗1, λ

∗2, λ

∗3, λ

∗4, λ

∗5)=x1x2x3 + 8(18− 2x1 − x2 − x3) + 8(18− x1 − 2x2 − x3)

=x1x2x3 − 24x1 − 24x2 − 16x3 + 288.

(v) The Hessian matrix of the restricted Lagrange function is

HLr(x1, x2, x3;λ∗1, λ

∗2, λ

∗3, λ

∗4, λ

∗5) =

⎛⎝ 0 x3 x2x3 0 x1x2 x1 0

⎞⎠ .

Evaluated at the optimal solution (x∗1, x∗2, x

∗3) = (4, 4, 6), this is the matrix

HLr(4, 4, 6; 0, 0, 0, 8, 8) =

⎛⎝0 6 46 0 44 4 0

⎞⎠ .

(vi) For perturbations Δx1, Δx2, and Δx3, the quadratic form of the Hessian matrixof the restricted Lagrange function is

Q(Δx1,Δx2,Δx3) = (Δx1,Δx2,Δx3)

⎛⎝0 6 46 0 44 4 0

⎞⎠⎛⎝Δx1Δx2Δx3

⎞⎠

= 12Δx1Δx2 + 8Δx1Δx3 + 8Δx2Δx3.

Page 357: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

338 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

Considering only the perturbations that satisfy (10.75) gives us the function of Δx1that is

Q(Δx1) = 12(Δx1)2 − 24(Δx1)

2 − 24(Δx1)2 = −36(Δx1)

2.

This function is nonpositive, confirming the second-order local maximality conditionthat must be true as a consequence of the local optimality of the point (x1, x2, x3) =(4, 4, 6).(vii) The relationships between the perturbations Δx1, Δx2, and Δx3 are all linear,so Theorem 10.2 may be used for the current problem.(viii) The bordered Hessian is

HLr(x1, x2, x3;λ∗1, λ

∗2, λ

∗3, λ

∗4, λ

∗5) =

⎛⎜⎜⎜⎜⎝0 0 2 1 10 0 1 2 12 1 0 x3 x21 2 x3 0 x11 1 x2 x1 0

⎞⎟⎟⎟⎟⎠ .

The value of the matrix at (x∗1, x∗2, x

∗3) = (4, 4, 6) is

HLr(4, 4, 6; 0, 0, 0, 8, 8) =

⎛⎜⎜⎜⎜⎝0 0 2 1 10 0 1 2 12 1 0 6 41 2 6 0 41 1 4 4 0

⎞⎟⎟⎟⎟⎠ . (10.77)

(ix) The number of choice variables is n = 3, so the set of integers to be permutedwhen designing the bordered principal permutation minors is {1, 2, 3}. The numberof binding constraints is k = 2. Theorem 10.2 informs us that we must check thevalues of each of the minors of orders i = k+1, . . . , n, so, for the present problem, weneed to check the values of only the order-3 minors. And, since each of the order-3permutation minors has the same value, we need to compute the value of only one ofthem. So let’s compute the determinant of (10.77). This is

M = det

⎛⎜⎜⎜⎜⎝0 0 2 1 10 0 1 2 12 1 0 6 41 2 6 0 41 1 4 4 0

⎞⎟⎟⎟⎟⎠ = −36.

Since this determinant has the same sign as (−1)i = (−1)3, the second-order necessarycondition for a local maximum is true.

Page 358: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

10.9. ANSWERS 339

Answer to Problem 10.3. The problem’s feasible set satisfies the Slater/Karlinconstraint qualification, so we are assured that the Karush-Kuhn-Tucker condition iswell-defined at any feasible solution. The Karush-Kuhn-Tucker condition evaluatedat (x1, x2, x3) = (4, 4, 6) is

∇f = (x2x3, x1x3, x1x2) = (24, 24, 16)

= λ1(−1, 0, 0) + λ2(0,−1, 0) + λ3(0, 0,−1) + λ4(2, 1, 1) + λ5(1, 2, 1),

where λ1 ≥ 0, λ2 ≥ 0, λ3 ≥ 0, λ4 ≥ 0, λ5 ≥ 0, and not all zero. The complementaryslackness conditions evaluated at (x1, x2, x3) = (4, 4, 6) are

λ1x1 = 0 = λ1 × 4 ⇒ λ∗1 = 0,

λ2x2 = 0 = λ2 × 4 ⇒ λ∗2 = 0,

λ3x3 = 0 = λ3 × 6 ⇒ λ∗3 = 0,

λ4(18− 2x1 − x2 − x3) = 0 = λ4 × 0 ⇒ λ∗4 ≥ 0,

and λ5(18− x1 − 2x2 − x3) = 0 = λ5 × 0 ⇒ λ∗5 ≥ 0.

The Karush-Kuhn-Tucker condition thus requires

24 = 2λ∗4 + λ∗5, 24 = λ∗4 + 2λ∗5 and 16 = λ∗4 + λ∗5 ⇒ λ∗4 = λ∗5 = 8.

The fourth and fifth constraints are strictly binding at (x1, x2, x3) = (4, 4, 6). Sincethese are constraints that are linear with respect to x1, x2, and x3, we may applyTheorem 10.3. The number of choice variables is n = 3. The number of bindingconstraints is k = 2. Therefore Theorem 10.3 tells us to evaluate only the order-3 successive bordered principal minor of the bordered Hessian matrix evaluated at(x1, x2, x3) = (4, 4, 6). From (10.77), this is

M = det

⎛⎜⎜⎜⎜⎝0 0 2 1 10 0 1 2 12 1 0 6 41 2 6 0 41 1 4 4 0

⎞⎟⎟⎟⎟⎠ = −36.

This minor is not zero and has the same sign as (−1)i = (−1)3, so the second-ordersufficiency condition is satisfied at (x1, x2, x3) = (4, 4, 6). Collectively, the Karush-Kuhn-Tucker condition, the complementary slackness conditions, and the second-order sufficiency condition establish that (x1, x2, x3) = (4, 4, 6) is locally maximal forthe constrained optimization problem.

Page 359: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

340 CHAPTER 10. CONSTRAINED OPTIMIZATION, PART II

Answer to Problem 10.4. The feasible set for problem 8.6 is a singleton, and twoconstraints are strictly binding at the optimal and only feasible solution (x1, x2) =(4, 4). Consequently the only perturbations Δx1 and Δx2 that continue to keep thebinding constraints binding are Δx1 = Δx2 = 0. Thus the second-order necessary lo-cal maximality condition is trivially true. The second-order local maximality sufficien-cy condition considers only nonzero perturbations that keep the binding constraintsbinding. There are no such perturbations, so the second-order local maximality suf-ficiency condition cannot be applied.

Answer to Problem 10.5. Consider first the point that is the optimal solution(x∗1, x

∗2) =

(412, 63

4

). The second-order condition considers small perturbations, Δx1

and Δx2, such that the binding constraint g1 continues to bind. That is, Δx1 andΔx2 are required to satisfy

g1(x∗1+Δx1, x

∗2+Δx2) = (x∗1+Δx1−4)2+x∗2+Δx2 = g1(x

∗1, x

∗2) = (x∗1−4)2+x∗2 = 7.

Rearranging gives

Δx2 = −(2x∗1 − 8 + Δx1)Δx1 = −(1 + Δx1)Δx1, (10.78)

since x∗1 = 412. Evaluated only for points (x∗1 +Δx1, x

∗2 +Δx2) nearby to (x∗1, x

∗2), the

problem’s objective function is

f(x∗1 +Δx1, x∗2 +Δx2) = x∗1 +Δx1 + x∗2 +Δx2 = 11

1

4+ Δx1 +Δx2.

When evaluated only for perturbations satisfying (10.78), the problem’s objectivefunction is

f(Δx1) = 111

4+ Δx1 − (1 + Δx1)Δx1 = 11

1

4− (Δx1)

2,

a function that obviously is concave with respect to Δx1. The second-order necessarylocal maximality condition is thus true, as it must be, because (x∗1, x

∗2) =

(412, 63

4

)is

locally maximal.Now let’s consider the feasible, not locally maximal, point (x1, x2) ≈ (5·412, 5·006).

At this point both constraints g1 and g2 bind. It should be clear from Figure 8.19that the only perturbations Δx1 and Δx2 that keep both constraints binding are the“nonperturbations” Δx1 = Δx2 = 0. For these perturbations is the necessary second-order local maximality condition true? Yes – only trivially, but it is true. What dowe learn from this? We get a reminder that, while necessary optimality conditionsmust be true at a local optimum, these same necessary optimality conditions may betrue at a point that is not a local optimum. Be careful.

Page 360: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

Chapter 11

Optimal Solutions and MaximumValues

11.1 Questions to Be Answered

Now that we have understood how to obtain solutions to differentiable constrainedoptimization problems, together with their maximum values, it is time to turn to thequestion of what properties might be possessed by optimal solution mappings andmaximum-value functions. This question can reasonably be divided into two parts.

The first part is the more technical question of whether or not the mappings thatare the optimal solution and the maximum value to some problem are functions,are continuous in some way, or are differentiable, and so on. And when can we besure that optimal solutions and maximum values exist? This chapter provides someanswers to these types of questions.

The second part is to examine how the values of the optimal solution and themaximum value for a constrained optimization problem alter with the values of theproblem’s parameters. For example, does the quantity demanded by a firm of someinput change when the price for the firm’s product changes? Does the quantityof a product supplied by a firm change if the rate of tax on its profit increases?And at what rates do these changes occur? These types of questions are answeredby a method called comparative statics analysis. Chapters 12 and 13 present threecomparative statics analysis methodologies.

341

Page 361: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

342 CHAPTER 11. OPTIMAL SOLUTIONS AND MAXIMUM VALUES

11.2 Solution Correspondences, Maximum Values

What do we mean when we talk of “the optimal solution to a constrained optimizationproblem”? Consider the problem maxx −ax2 + bx+ c, where a > 0, b, and c areparameters. Take a moment and think carefully about what you mean by the optimalsolution to this problem.

You probably have decided that the optimal solution is x = −b/2a. This is astatement that the optimal solution is a mapping. It is a mapping from the set ofparameter vectors (a, b, c) that are admissible to the problem. The optimal solutionis always a real number, so it maps to the set �. Consequently it is more informativeto write the optimal solution using a notation such as

x∗(a, b, c) = − b

2a(11.1)

that clearly asserts the dependence of the value of the problem’s optimal solutionupon the values of the problem’s parameters. Thus we write the optimal solutionmapping for this problem as x∗ : �++ × � × � �→ �, with x∗ defined by (11.1). Wewill find that much the same is true for the optimal solution mappings for constrainedoptimization problems; i.e. an optimal solution to a constrained optimization problemis a mapping from the parameters of the problem (the parameters of the objectivefunction, the parameters of the constraint functions, and the levels of the constraints)to a real number space.

You may recall a notational convention that I used in Section 2.7 to distinguishbetween quantities that are variables for a mapping and quantities that are parametersfor that mapping. The convention is that all of the quantities listed before (to the leftof) a semicolon are variables, while all of the quantities listed after (to the right of)the semicolon are parameters. So, if I write H(t, z), then, since there is no semicolon,I mean a mapping in which both t and z are variables. If I write instead H(t; z),then, since t is to the left of the semicolon and z is to its right, I mean that t is avariable, while z is a parameter. Writing H(z; t) means instead that z is a variable,while t is a parameter. For example, if H(t, z) = t2(z+2), then H(t; z = 3) = 5t2 andH(z; t = 2) = 4(z + 2). We will often call H(t; z) and H(z; t) restricted versions ofH(t, z), since H(t; z′) is H(t, z) subject to the restriction that z is fixed at the valuez′, and H(z; t′) is H(t, z) subject to the restriction that t is fixed at the value t′.

Let’s take a moment to refresh our memories of the notation used in earlier chap-ters. The constrained optimization problem that we spent a lot of time thinking

Page 362: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

11.2. SOLUTION CORRESPONDENCES, MAXIMUM VALUES 343

about in Chapters 8 and 10 is:

maxx1, ... , xn

f(x1, . . . , xn) subject to gj(x1, . . . , xn) ≤ bj for j = 1, . . . ,m. (11.2)

I would like to modify the notation used in (11.2) to make explicit the way that thisproblem and its solutions depend upon parameters, so let me rewrite (11.2) as

maxx∈�n

f(x;α) subject to gj(x; βj) ≤ bj for j = 1, . . . ,m. (11.3)

In (11.3), α is a vector list of the parameters of the objective function and βjis a vector list of the parameters of the jth constraint function. It is convenient toconcatenate (i.e. join together as a chain) these vectors, so let’s define

β ≡

⎛⎜⎝β1...βm

⎞⎟⎠ , b ≡

⎛⎜⎝ b1...bm

⎞⎟⎠ , and γ ≡

⎛⎝αβb

⎞⎠ .

The concatenation operator is usually denoted by a colon, so we can also write β =(β1 : · · · : βm) and γ ≡ (α : β : b). The set of all allowed values for the parametervector γ is called the problem’s parameter space. We will use Γ to denote the problem’sparameter space, so every allowed parameter vector γ ∈ Γ.

An example is the profit-maximization problem

maxy, x1, x2

py − w1x1 − w2x2 subject to y ≥ 0, x1 ≥ 0, x2 ≥ 0, and y−xβ1

1 xβ2

2 ≤ 0 (11.4)

that confronts a firm facing a given product price p > 0, given input prices w1 > 0and w2 > 0, and using a technology parameterized by β1 > 0, β2 > 0, and β1+β2 < 1.In terms of the notation used in (11.3), the functions in (11.4) are

f(y, x1, x2; p, w1, w2) = py − w1x1 − w2x2, g1(y) = −y,g2(x1) = −x1, g3(x2) = −x2, and g4(y, x1, x2; β1, β2) = y − xβ1

1 xβ2

2 .

Notice that some of the parameters of (11.4) are numerically specified (b1 = b2 = b3 =b4 = 0), while others (i.e. p, w1, w2, β1, and β2) are left as symbols without specifiedvalues. This indicates that the optimal solution correspondence is to be stated asa mapping of the parameters p, w1, w2, β1, and β2. With this in mind, then, theparameter vectors of interest for (11.4) are

α =

⎛⎝ pw1

w2

⎞⎠ and β =

(β1β2

), so γ =

⎛⎜⎜⎜⎜⎝pw1

w2

β1β2

⎞⎟⎟⎟⎟⎠ ,

Page 363: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

344 CHAPTER 11. OPTIMAL SOLUTIONS AND MAXIMUM VALUES

and the parameter space Γ = �3++ × {(β1, β2) | β1 > 0, β2 > 0, β1 + β2 < 1}.

The optimal solution to our example (11.4) is the vector of functions (y∗(γ), x∗1(γ),x∗2(γ), λ

∗1(γ), λ

∗2(γ), λ

∗3(γ), λ

∗4(γ)) that state the values of the input and output levels

that maximize the firm’s profit and the values of the constraints’ multipliers. Thesefunctions of p, w1, w2, β1, and β2 are

y∗(p, w1, w2, β1, β2) =

(pβ1+β2ββ1

1 ββ2

2

wβ1

1 wβ2

2

)1/(1−β1−β2)

, (11.5a)

x∗1(p, w1, w2, β1, β2) =

(pβ1−β2

1 ββ2

2

w1−β2

1 wβ2

2

)1/(1−β1−β2)

, (11.5b)

x∗2(p, w1, w2, β1, β2) =

(pββ1

1 β1−β1

2

wβ1

1 w1−β1

2

)1/(1−β1−β2)

, (11.5c)

λ∗1(p, w1, w2, β1, β2) = 0, (11.5d)

λ∗2(p, w1, w2, β1, β2) = 0, (11.5e)

λ∗3(p, w1, w2, β1, β2) = 0, (11.5f)

and λ∗4(p, w1, w2, β1, β2) = p. (11.5g)

Each of y∗, x∗1, x∗2, λ

∗1, λ

∗2, λ

∗3, and λ

∗4 is a mapping from the parameter space of the

firm’s problem to the nonnegative part of the real line.Let us return to our more general statement (11.3) of our constrained optimization

problem. A will be used to denote the set of values that can be taken by the parametervector α. For each j = 1, . . . ,m, the set of all possible values for βj will be denotedby Bj. b will denote the set of all possible values for the vector b. The problem’sparameter space is thus

Γ ≡ A× B1 × · · · ×Bm × b.

The feasible set C(β, b) ⊆ X for a particular problem (11.3) is determined only bythe problem’s constraints (not by the problem’s objective function), and so dependsupon the parameter vectors β and b but is independent of the parameter vector α thatdetermines only the objective function. The mapping from the problem’s parameterspace to the set of all possible feasible sets is called the problem’s feasible solutioncorrespondence.

Definition 11.1 (Feasible Set Correspondence). The feasible solution correspon-dence for problem (11.3) is the mapping C : ×m

j=1Bj × b �→ X that is

C(β, b) = {x ∈ X | g1(x; β1) ≤ b1, . . . , gm(x; βm) ≤ bm} for (β, b) ∈ ×mj=1Bj × b.

Page 364: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

11.2. SOLUTION CORRESPONDENCES, MAXIMUM VALUES 345

An example is the set of consumption bundles that are feasible (affordable) fora consumer with a budget of y who faces prices p1, . . . , pn for commodities 1, . . . , n.For this problem, the feasible set correspondence is the budget set correspondence

C(p1, . . . , pn, y) = {(x1, . . . , xn) | x1 ≥ 0, . . . , xn ≥ 0, p1x1 + · · ·+ pnyn ≤ y}

for (p1, . . . , pn, y) ∈ �n++ ×�+.

The feasible set correspondence for problem (11.4) is the firm’s technology set;i.e. it is the set of all technically feasible production plans,

C(β1, β2) ={(x1, x2, y) | x1 ≥ 0, x2 ≥ 0, 0 ≤ y ≤ xβ1

1 xβ2

2

}for all (β1, β2) ∈ {(β1, β2) | β1 > 0, β2 > 0, 0 < β1 + β2 < 1}.

The set of solutions that are optimal for problem (11.3) for a particular parametervector typically depends upon all of the parameter vectors α, β, and b. Particularvalues for β and b determine a particular feasible set. A particular value for α thendetermines the particular objective function for the problem, and so determines whichof the feasible solutions is/are optimal. For a particular γ ∈ Γ, the set of optimalsolutions to the particular constrained optimization problem parameterized by γ iscalled the optimal solution set. The mapping from the problem’s parameter space tothe set of optimal solutions is called the problem’s optimal solution correspondence.

Definition 11.2 (Optimal Solution Correspondence). The optimal solution corre-spondence for problem (11.3) is the mapping X∗ : Γ �→ X that is

X∗(α, β, b) = {x∗ ∈ C(β, b) | f(x∗;α) ≥ f(x;α) ∀ x ∈ C(β, b)} for (α, β, b) ∈ Γ.

A familiar economics example is the consumer budget allocation problem in whicha consumer with a Cobb-Douglas direct utility function chooses a most-preferredconsumption bundle from those he can afford. For two commodities the problem is

maxx1, x2

U(x1, x2) = xα11 x

α22 subject to x1 ≥ 0, x2 ≥ 0, and p1x1 + p2x2 ≤ y,

where α1 > 0, α2 > 0, p1 > 0, p2 > 0, and y ≥ 0. The optimal solution correspondenceis the mapping

X∗(α1, α2, p1, p2, y) =(x∗1(α1, α2, p1, p2, y), x

∗2(α1, α2, p1, p2, y)

)=

(α1y

(α1 + α2)p1,

α2y

(α1 + α2)p2

)

Page 365: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

346 CHAPTER 11. OPTIMAL SOLUTIONS AND MAXIMUM VALUES

from the parameter space Γ = {(α1, α2, p1, p2, y) | α1 > 0, α2 > 0, p1 > 0, p2 >0, y ≥ 0} = �4

++ �+ to the commodity space �2+.

For problem (11.4) the optimal solution correspondence is, from (11.5a), (11.5b),and (11.5c),

X∗(p, w1, w2, β1, β2)

=((y∗(p, w1, w2, β1, β2), x

∗1(p, w1, w2, β1, β2), x

∗2(p, w1, w2, β1, β2)

)=

⎛⎝(pβ1+β2ββ1

1 ββ2

2

wβ1

1 wβ2

2

)1/(1−β1−β2)

,

(pβ1−β2

1 ββ2

2

w1−β2

1 wβ2

2

)1/(1−β1−β2)

,

(pββ1

1 β1−β1

2

wβ1

1 w1−β1

2

)1/(1−β1−β2)⎞⎠ .

This is a mapping from the problem’s parameter space Γ = {(p, w1, w2, β1, β2) | p >0, w1 > 0, w2 > 0, β1 > 0, β2 > 0, 0 < β1 + β2 < 1} to the production planspace �3

+.What is the value provided by an optimal solution? It is the maximum value

that is achievable given the constraints on choices; i.e. it is the largest value for theproblem’s objective function that is possible from all of the choices in the feasible set.So it is natural that we call this maximum value exactly that: the maximum valuepossible for the specified values (α, β, b) of the problem’s parameters. The mappingso constructed is the problem’s maximum-value function.

Definition 11.3 (Maximum-Value Function). The maximum-value function for prob-lem (11.3) is the mapping v : Γ �→ � that is

v(α, β, b) = f(x∗(α, β, b);α) for x∗(α, β, b) ∈ X∗(α, β, b), ∀ (α, β, b) ∈ Γ.

The maximum-value mapping is a function; i.e. associated with each possiblevector of parameters there is only a single maximum value for the objective function.Why is this? Figure it out for yourself by asking what if it is not unique? For example,suppose for some particular vector of parameters there were two maximum values,say, v = 33 and v = 37. This is a contradiction. Why?

For the above example of the consumer’s budget allocation problem, the maximum-value function reveals how the consumer’s maximum achievable utility level dependsupon the problem’s parameters. Economists call this function the consumer’s indirectutility function. For our example it is the function v : �2

++ ×�+ �→ � that is

v(p1, p2, y) = U(x∗1(p1, p2, y), x∗2(p1, p2, y)) =

(α1y

(α1 + α2)p1

)α1(

α2y

(α1 + α2)p2

)α2

.

(11.6)

Page 366: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

11.3. EXISTENCE OF AN OPTIMAL SOLUTION 347

For problem (11.4) the maximum-value function reveals how the firm’s maximumachievable profit level depends upon the problem’s parameters. Economists call thisfunction the firm’s indirect profit function. For our example it is the functionv : �3

++ × {(β1, β2) | β1 > 0, β2 > 0, 0 < β1 + β2 < 1} �→ � that is

v(p, w1, w2, β1, β2)

= py∗(p, w1, w2, β1, β2)− w1x∗1(p, w1, w2, β1, β2)− w2x

∗2(p, w1, w2, β1, β2))

= (1− β1 − β2)

(pββ1

1 ββ2

2

wβ1

1 wβ2

2

)1/(1−β1−β2) (11.7)

once we substitute from (11.5a), (11.5b), and (11.5c), and simplify the resultingexpression.

11.3 Existence of an Optimal Solution

Not every constrained optimization problem has an optimal solution. An example isthe problem maxx f(x) = x subject to the constraint x < 1. Then there are otheroptimization problems that have optimal solutions for some values of their parametervectors γ and have no optimal solutions for other values of the parameter vectors. Foran example, think of the problem maxx f(x) = ax2 + 2x+ 3, where the parametera can be any real number. If a < 0, then the problem has the optimal solutionx∗(a) = 1/a. But if a ≥ 0, then the problem has no optimal solution. If, fora particular parameter vector γ, an optimization problem has no optimal solution,then the value of the optimal solution correspondence is the empty set; X∗(γ) = ∅.Other problems have many, even an infinity, of optimal solutions; e.g. maxx f(x) ≡ 2for 0 ≤ x ≤ 1 is a constrained optimization problem with an infinity of optimalsolutions. For which types of constrained optimization problems can we be sure thatthere is always at least one optimal solution? Are there problems in which there isalways a unique optimal solution?

The good news is that almost all of the constrained optimization problems en-countered by economists have an optimal solution. If the feasible set of the problemis nonempty, closed, and bounded with respect to the Euclidean metric topology onEn, and if the problem’s objective function is upper-semi-continuous over the feasibleset, then there exists at least one global maximizing solution to the problem. If theobjective function is instead lower-semi-continuous over the feasible set, then thereexists at least one global minimizing solution to the problem. In many cases we will

Page 367: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

348 CHAPTER 11. OPTIMAL SOLUTIONS AND MAXIMUM VALUES

not be interested in both types of solutions. For example, most firms want to avoidprofit-minimizing production plans.

Theorem 11.1 (Weierstrass’s Theorem for Upper-Semi-Continuous Functions). LetC be a nonempty subset of �n that is compact with respect to En. Let f : C �→ �be a function that is upper-semi-continuous on C. Then there exists x ∈ C such thatf(x) ≥ f(x) for every x ∈ C.

An example of the content of this theorem is displayed in Figure 11.1. The upper-semi-continuous function f displayed there must attain a global maximum on thecompact set that is the interval [a, b]. It is possible that an upper-semi-continuousfunction also attains local maxima, local minima, and a global minimum, but it isonly the existence of the global maximum that is assured.

Figure 11.1: On the compact interval [a, b], the upper-semi-continuous function fattains a global maximum at x = x∗, attains a local minimum, and does not attain aglobal minimum.

Proof of Theorem 11.1. Let f = supx∈C f(x). It must be shown that there is avalue x ∈ C for which the corresponding value of f is f .

Since f = supx∈C f(x), there exists in C a sequence {xn}∞n=1 such that

f(xn) → f. (11.8)

Page 368: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

11.3. EXISTENCE OF AN OPTIMAL SOLUTION 349

Since C is compact, this sequence must contain a convergent subsequence {xnk}∞k=1.

Let x = limk→∞ xnk. x ∈ C because C is a closed set.

Since f is upper-semi-continuous on C,

limk→∞

f(xnk) ≤ f(x). (11.9)

The sequences {f(xn)}∞n=1 and {f(xnk)}∞k=1 must converge to the same value, so it

follows from (11.8) and (11.9) that f ≤ f(x). But, by definition, f ≥ f(x) for everyx ∈ C, so it must be that f = f(x).

Theorem 11.2 (Weierstrass’s Theorem for Lower-Semi-Continuous Functions). LetC be a nonempty subset of �n that is compact with respect to En. Let f : C �→ �be a function that is lower-semi-continuous on C. Then there exists x ∈ C such thatf(x) ≤ f(x) for every x ∈ C.

The proof is very similar to that of Theorem 11.1, so take some time to under-stand the proof of Theorem 11.1 and then prove Theorem 11.2 for yourself – this isproblem 11.2.

Figure 11.2: On the compact interval [a, b], the lower-semi-continuous function fattains a global minimum at x = x∗, attains a local maximum, and does not attain aglobal maximum.

An example of the content of Theorem 11.2 is displayed in Figure 11.2. Thelower-semi-continuous function f displayed there must attain a global minimum on

Page 369: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

350 CHAPTER 11. OPTIMAL SOLUTIONS AND MAXIMUM VALUES

the compact set that is the interval [a, b]. It is possible that a lower-semi-continuousfunction also attains local maxima, local minima, and a global maximum, but it isonly the existence of the global minimum that is assured.

The most commonly employed existence result is the theorem that one obtainswhen the constrained optimization problem’s objective function is continuous, and sois both lower-semi-continuous and upper-semi-continuous on the problem’s feasibleset C. Combining Theorems 11.1 and 11.2 gives the following result.

Theorem 11.3 (Weierstrass’s Theorem for Continuous Functions). Let C be a non-empty subset of �n that is compact with respect to En. Let f : C �→ � be a functionthat is continuous on C. Then there exists x ∈ C such that f(x) ≥ f(x) for everyx ∈ C and, also, there exists x ∈ C such that f(x) ≤ f(x) for every x ∈ C.

The content of Weierstrass’s Theorem is displayed in Figure 11.3. Notice thatWeierstrass’s Theorem implies neither a unique global maximizing solution nor aunique global minimizing solution.

Figure 11.3: On the compact set [a, b], the continuous function f(x) attains a globalmaximum at x = x∗, and attains a global minimum at both x = x∗ and x = x∗∗.

The above theorems are sufficiency statements. It is possible that a lower-semi-continuous function can attain a global maximum and that an upper-semi-continuousfunction can attain a global minimum. It is also possible that a lower-semi-continuousfunction defined on a compact set does not attain even a local maximum and that an

Page 370: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

11.4. THEOREM OF THE MAXIMUM 351

upper-semi-continuous function defined on a compact set does not attain even a localminimum, as the following two examples demonstrate. Let C = [0, 4]. The function

v(x) =

{x, 0 ≤ x ≤ 2

5− x, 2 < x ≤ 4

is lower-semi-continuous on [0, 4] and does not attain even a local maximum on [0, 4].Again let C = [0, 4]. The function

t(x) =

{5− x, 0 ≤ x ≤ 2

x, 2 < x ≤ 4

is upper-semi-continuous on [0, 4] and does not attain even a local minimum on [0, 4].Now let us turn to discovering some of the properties that can be possessed by an

optimal solution correspondence.

11.4 Theorem of the Maximum

Let’s start with an example. Consider the budget allocation problem in which aconsumer wants to maximize his utility from consumption of quantities x1 and x2 ofcommodities 1 and 2. Commodity 2 is the numeraire commodity, so p2 = 1. Theconsumer regards the two commodities as perfect substitutes, so his direct utilityfunction is U(x1, x2) = x1 + x2. The consumer’s problem is

maxx1, x2

U(x1, x2) = x1 + x2 subject to x1 ≥ 0, x2 ≥ 0 and p1x1 + x2 ≤ y. (11.10)

The optimal solution correspondence is

X∗(p1, y) =

⎧⎪⎪⎪⎨⎪⎪⎪⎩

(y

p1, 0

), p1 < 1

{(x1, x2) | x1 ≥ 0, x2 ≥ 0, x1 + x2 = y} , p1 = 1

(0, y) , p1 > 1.

(11.11)

This optimal solution correspondence, which maps to �2+, is upper-hemi-continuous

with respect to p1 and is displayed in Figure 11.4. It says that when commodity 1is the less expensive (i.e. p1 < p2 = 1) the consumer purchases only commodity 1.When commodity 2 is the less expensive (i.e. p1 > p2 = 1) the consumer purchases

Page 371: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

352 CHAPTER 11. OPTIMAL SOLUTIONS AND MAXIMUM VALUES

Figure 11.4: The ordinary demand correspondence X∗(p1, y) is upper-hemi-continuous with respect to p1.

only commodity 2. When the commodities are equally expensive (i.e. p1 = p2 = 1)the consumer cares only about the total number of commodity units he acquires.

For p1 = 1, the consumer does not care which commodity he purchases, so theimage of the optimal solution correspondence for p1 = 1 is the set X∗(p1 = 1, y) ={(x1, x2) | x1 ≥ 0, x2 ≥ 0, x1+x2 = y}. The correspondence is thus not single-valuedeverywhere and so is not a function.

If p1 > 1, then the consumer purchases only commodity 2, so X∗(p1, y) = (0, y) forall p1 > 1. Therefore the limit of the images of the optimal solution correspondenceas p1 approaches unity from above is

limp1→1+

X∗(p1, y) = limp1→1+

(0, y)

= (0, y) ∈ {(x1, x2) | x1 ≥ 0, x2 ≥ 0, x1 + x2 = y} = X∗(1, y).

The limit limp1→1+ X∗(p1, y) is a member of the set X∗(p1 = 1, y).

If p1 < 1, then the consumer purchases only commodity 1, so X∗(p1, y) = (y/p1, 0)for all 0 < p1 < 1. Therefore the limit of the images of the optimal solution corre-

Page 372: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

11.4. THEOREM OF THE MAXIMUM 353

spondence as p1 approaches unity from below is

limp1→1−

X∗(p1, y) = limp1→1−

(y

p1, 0

)= (y, 0) ∈ {(x1, x2) | x1 ≥ 0, x2 ≥ 0, x1 + x2 = y} = X∗(1, y).

The limit limp1→1− X∗(p1, y) is also a member of the set X∗(p1 = 1, y).

Since limp1→1− X∗(p1, y) ∈ X∗(p1 = 1, y) and limp1→1+ X

∗(p1, y) ∈ X∗(p1 = 1, y),X∗ is upper-hemi-continuous at p1 = 1. X∗ is a continuous function for all p1 > 0other than p1 = 1 and so is upper-hemi-continuous for all p1 > 0.

It turns out that there is a large class of constrained optimization problems ofinterest to economists for which the optimal solution correspondence is upper-hemi-continuous. This very useful information is provided by the Theorem of the Maximum,a result proved in 1959 by Claude Berge. At first sight the statement of this importantresult appears daunting and confusing. But don’t worry – we will examine it part bypart and will discover that it is not too complicated after all.

Theorem 11.4 (Berge’s Theorem of the Maximum). Consider problem (11.3). Theproblem’s feasible set correspondence is C : ×m

j=1Bj × b → X, where

C(β, b) = {x ∈ X | gj(x; βj) ≤ bj ∀ j = 1, . . . ,m}.

The problem’s optimal solution correspondence is X∗ : A×mj=1 Bj × b → X, where

X∗(α, β, b) = {x∗ ∈ C(β, b) | f(x∗;α) ≥ f(x;α) ∀ x ∈ C(β, b)}.

The problem’s maximum-value function is v : A×mj=1 Bj × b → �, where

v(α, β, b) = f(x∗;α) ∀ x∗ ∈ X∗(α, β, b).

If f is jointly continuous with respect to (x, α) and if C is nonempty-valued,compact-valued, and continuous with respect to (β, b), then X∗ is nonempty-valuedand is upper-hemi-continuous with respect to (α, β, b), and v is continuous with re-spect to (α, β, b).

An example will help to make sense of the statement of the theorem, so let usthink of a small generalization to problem (11.10). Consider the problem

maxx1, x2

f(x1, x2;α) = αx1 + x2 subject to x1 ≥ 0, x2 ≥ 0, and p1x1 + x2 ≤ y. (11.12)

Page 373: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

354 CHAPTER 11. OPTIMAL SOLUTIONS AND MAXIMUM VALUES

We again take commodity 2 to be the numeraire, meaning that p2 = 1. Observe thatthe problem’s objective function is jointly continuous in (x1, x2, α) (joint continuityis more demanding than is continuity in each of x1, x2, and α individually – seeSection 5.6). The problem’s feasible solution correspondence is

C(p1, y) = {(x1, x2) | x1 ≥ 0, x2 ≥ 0 and p1x1 + x2 ≤ y}.

For given values for p1 > 0 and y ≥ 0, the problem’s feasible set is a nonempty andcompact subset of �2

+. Thus C is a nonempty-valued and compact-valued correspon-dence. Notice that the feasible set’s boundaries move smoothly (i.e. continuously)as the values of either p1 or y are changed. This is what we mean by the statementthat C(p1, y) is a correspondence that is continuous with respect to (p1, y).

The optimal solution (ordinary demand) correspondence is

X∗(α, p1, y) =

⎧⎪⎪⎪⎨⎪⎪⎪⎩

(y

p1, 0

), p1 < α

{(x1, x2) | x1 ≥ 0, x2 ≥ 0, αx1 + x2 = y} , p1 = α

(0, y) , p1 > α.

Figure 11.4 displays the optimal solution correspondence for α = 1. The importantfeature to notice is that the correspondence is upper-hemi-continuous with respect top1 > 0 and α > 0, and is fully continuous (thus upper-hemi-continuous) with respectto y ≥ 0.

The problem’s maximum-value (indirect utility) function is

v(α, p1, p2, y) = max

p1, 1

}y.

This is a continuous function of the problem’s parameters α > 0, p1 > 0 and y ≥ 0.

Proof of the Theorem of the Maximum. The proof is a somewhat lengthysequence of steps, but only two involve serious thought. To make it easier to follow,I have placed titles at the start of each new part of the proof. For notational brevitywe will once again use γ to denote the parameter vector (α, β, b) and use Γ to denotethe parameter space A×m

j=1 Bj × b.(i) Show X∗ is not empty-valued. What do we know about the feasible set? We

are given that, for any γ ∈ Γ (actually, for any given β and b), the feasible set C(β, b)is not empty and is a compact set. What do we know about the objective function?For any γ ∈ Γ (actually, for any given α), we know that the objective function

Page 374: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

11.4. THEOREM OF THE MAXIMUM 355

f(x;α) is continuous with respect to x. It follows from Weierstrass’s Theorem 11.3that there is at least one x∗ ∈ C(β, b) that achieves a maximum for f(x;α) overC(β, b); i.e. X∗(γ) = ∅. Since this is true for every γ ∈ Γ, X∗ is a nonempty-valuedcorrespondence.

(ii) Show X∗(γ) is a bounded set. The set of solutions that are optimal given γ isa subset of the set of solutions that are feasible given γ; i.e. X∗(α, β, b) ⊆ C(β, b). Cis a compact-valued correspondence, so for any given γ ∈ Γ the feasible set C(β, b) isa bounded set. Hence its subset X∗(γ) is also a bounded set.

(iii) Show X∗(γ) is a closed set. We need to show that the limit of any convergentsequence of elements in X∗(γ) is an element ofX∗(γ), so we start by taking a sequence{x∗n}∞n=1 of elements with x∗n ∈ X∗(γ) for every n ≥ 1 that converges to some point x0;i.e. x∗n → x0. The first thing we need to do is to show that x0 is a feasible solution;i.e. x0 ∈ C(β, b). Then we need to show that x0 is optimal; i.e. x0 ∈ X∗(γ).

Let’s show that x0 ∈ C(β, b). Each of the optimal solutions x∗n must be a feasiblesolution: x∗n ∈ C(β, b) for every n ≥ 1. But then x0 ∈ C(β, b) because x∗n → x0 andC(β, b) is a closed set.

Now we need to show that x0 ∈ X∗(γ). For the given value of α, the objectivefunction f(x;α) is continuous with respect to x, so

limn→∞

f(x∗n;α) = f( limn→∞

x∗n;α) = f(x0;α). (11.13)

All of the optimal solutions x∗n provide the same maximum value; denote it by v.That is, f(x∗n;α) = v for all n ≥ 1. Therefore (11.13) is the statement that

v = f(x0;α). (11.14)

That is, x0 provides the maximum value v so x0 is an optimal solution: x0 ∈ X∗(γ).Therefore X∗(γ) is a closed set.

(iv) Show X∗(γ) is a compact set. X∗(γ) is both closed and bounded, so the setis compact with respect to En.

(v) Show X∗ is upper-hemi-continuous on Γ. What is it that we need to provehere? Have a look at Figure 11.5. We need to consider a sequence {γn}∞n=1 of pa-rameter vectors in Γ that converges to a vector that we will denote by γ0; γn → γ0.For each value γn, there is a set of solutions X∗(γn) that are optimal when γ = γn.From each of the sets X∗(γ1), X∗(γ2), . . ., we take arbitrarily one element and so forma sequence x∗n, where x

∗n ∈ X∗(γn) for each n ≥ 1. We need to show that this se-

quence possesses a convergent subsequence (let’s call the limit x∗0), and then we needto show that x∗0 is a solution that is optimal when γ = γ0; i.e. we need to show thatx∗0 ∈ X∗(γ0).

Page 375: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

356 CHAPTER 11. OPTIMAL SOLUTIONS AND MAXIMUM VALUES

Figure 11.5: X∗(γ) is upper-hemi-continuous at γ = γ0.

We start with the sequence {γn}∞n=1 with γn → γ0 and the associated sequenceof optimal solution sets {X∗(γn)}∞n=1, from each of which a single element has beenselected arbitrarily to form the sequence {x∗n}∞n=1. The sequence {x∗n}∞n=1 need not beconvergent. However, it must contain at least one convergent subsequence. Why so?Well, again note that any optimal solution is a feasible solution, so x∗n ∈ C(βn, bn)(the set of feasible solutions when γ = γn) for each n ≥ 1. We are given thatC is a correspondence that is continuous, and thus is upper-hemi-continuous, withrespect to (β, b). Therefore the sequence {x∗n}∞n=1 of feasible solutions must possess asubsequence {x∗nk

}∞k=1 that converges to a value that we denote by x∗0 and, moreover,x∗0 ∈ C(β0, b0); i.e. x∗0 is a feasible solution when γ = γ0.

Now we need to show that x∗0 is an optimal solution when γ = γ0; i.e. we need toshow that

f(x∗0;α0) ≥ f(z0;α0) ∀ z0 ∈ C(β0, b0). (11.15)

Choose any z0 ∈ C(β0, b0). Since C is a continuous correspondence with respect to(β, b), it is lower-hemi-continuous and so there must exist a sequence {zn}∞n=1 with

Page 376: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

11.4. THEOREM OF THE MAXIMUM 357

zn ∈ C(βn, bn) for all n ≥ 1 that converges to z0. Now let’s create the subsequence{znk

}∞k=1 by selecting the elements from the {zn}∞n=1 sequence that correspond to the{x∗nk

}∞k=1 subsequence. Any subsequence of a convergent subsequence is convergent,so znk

→ z0. Now let’s remember that

f(x∗nk;αnk

) ≥ f(znk;αnk

) ∀ k ≥ 1, (11.16)

because x∗nkis optimal when γ = γnk

, while znkis merely feasible when γ = γnk

. f isjointly continuous in (x, α), so it follows from (11.16) that

f( limk→∞

x∗nk; limk→∞

αnk) = f(x∗0;α0) ≥ f( lim

k→∞znk

; limk→∞

αnk) = f(z0;α0). (11.17)

Thus x∗0 is optimal when γ = γ0; i.e. x∗0 ∈ X∗(γ0), establishing thatX∗ is upper-hemi-

continuous at γ0. But γ0 is an arbitrary choice from Γ, soX∗ is upper-hemi-continuouson Γ.

We have now completed the part of the theorem that deals with the properties ofthe optimal solution correspondence X∗. The rest of the proof establishes the claimedproperties of the maximum-value function.

(vi) Show v exists and is a function. By definition, v(γ) = f(x∗(γ);α). SinceX∗(γ) = ∅, the image set v(γ) = ∅. Suppose that, for some γ ∈ Γ, the set v(γ) isnot a singleton. Then v(γ) must contain at least two different values, say, v′ and v′′.Without loss of generality, suppose that v′ < v′′. But then v′ is not a maximum value.Contradiction. Thus v(γ) must be a singleton for each γ ∈ Γ; i.e. v is a function.

(vii) Show v is continuous on Γ. We must show that v(γn) → v(γ0) for anysequence {γn}∞n=1 that converges to γ0. So we start by taking any sequence {γn}∞n=1

that converges to some limit γ0. Then we look at the accompanying sequence ofmaximum values {v(γn)}∞n=1. This sequence must be bounded, both above and below,so the sequence has a least upper bound (the limit superior of the sequence; seeDefinition 4.6) and, therefore, contains a subsequence {v(γnk

)}∞k=1 that converges tothis least upper bound: v(γnk

) → lim sup v(γn). The accompanying subsequence{γnk

}∞k=1 must converge to γ0 because it is drawn from the sequence {γn}∞n=1 andγn → γ0. For each k ≥ 1, there is an optimal solution x∗(γnk

) ∈ X∗(γnk) providing

the maximum value v(γnk). Construct a sequence {x∗nk

}∞k=1 of these optimal solutionsby arbitrarily selecting one element from each of the sets X∗(γnk

). Since the optimalsolution correspondence X∗ is upper-hemi-continuous with respect to γ, the sequence{x∗nk

}∞k=1 must contain a subsequence that converges to some x∗0 ∈ X∗(γ0). Denotethis subsequence by {x∗nkj

}∞j=1; x∗nkj

→ x∗0. From the joint continuity of f with respect

Page 377: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

358 CHAPTER 11. OPTIMAL SOLUTIONS AND MAXIMUM VALUES

to (x, α),

limj→∞

v(γnkj) = lim

j→∞f(x∗nkj

;αnkj) = f( lim

j→∞x∗nkj

; limj→∞

αnkj) = f(x∗0;α0) = v(γ0).

(11.18)Thus

v(γ0) = lim sup v(γn). (11.19)

The bounded sequence {v(γn)}∞n=1 must also have a greatest lower bound (thelimit inferior of the sequence; see Definition 4.6) and so must contain a subsequence{v(γnt)}∞t=1 that converges to this greatest lower bound: v(γnt) → lim inf v(γn). Re-peating the above argument for this subsequence shows that

v(γ0) = lim inf v(γn). (11.20)

(11.19) and (11.20) together show that the sequence {v(γn)}∞n=1 does converge andthat its limit is v(γ0), so we have established that v is continuous at γ = γ0. But γ0is an arbitrary choice from Γ, so v is continuous everywhere on Γ.

11.5 Comparative Statics Analysis

Economists are usually interested in the directions of change caused to economic vari-ables when some other quantity, a policy variable perhaps, changes. Will an increasein a tax rate cause an increase in tax revenue, for example? These questions concernhow the optimal solutions to, and maximum-value functions of, constrained optimiza-tion problems change as the parameters of the problems are changed. In this chapterwe have not been able to address such questions because insufficient information isprovided about the ways in which the objective and constraint functions are alteredby changes to parameters. The following two chapters develop comparative static-s analysis, a name given to any method that determines how optimal solutions ormaximum values are altered by changes to parameters. These two chapters explaindifferential comparative statics analysis methods only. Another method, called mono-tone comparative statics analysis, which does not rely upon differentiability, is notpresented in this book because of the amount of preparation needed to comprehendit.

Differentiable comparative statics analyses can be divided into two types. A com-parative statics analysis that determines only the directions in which optimal solutionsor maximum values change when a parameter changes is called “qualitative.” Suchan analysis might predict that the quantity demanded of coffee decreases when the

Page 378: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

11.5. COMPARATIVE STATICS ANALYSIS 359

price of coffee rises, but it would not say how quickly. A comparative statics anal-ysis that determines the values of rates of change of optimal solutions or maximumvalues with respect to changes to a parameter is called “quantitative.” Clearly aquantitative analysis is more informative than is a qualitative analysis, but under-taking a quantitative analysis requires a lot more information than does a qualitativeanalysis. Whether this additional information is available to you will in large partdetermine which type of analysis you are able to undertake. Chapter 12 explains howto undertake quantitative analyses. Chapter 13 explains how, under less demandingconditions, to conduct qualitative analyses.

Page 379: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

360 CHAPTER 11. OPTIMAL SOLUTIONS AND MAXIMUM VALUES

11.6 Problems

Problem 11.1. Provide an example of an upper-semi-continuous function that is notcontinuous and that attains both a global maximum and a global minimum. Thenprovide an example of a lower-semi-continuous function that is not continuous andattains both a global maximum and a global minimum. Why is it that these examplesdo not contradict Theorems 11.1 and 11.2?

Problem 11.2. Have another look at the proof provided for Weierstrass’s Theo-rem for upper-semi-continuous functions (Theorem 11.1). Then prove Weierstrass’sTheorem for lower-semi-continuous functions (Theorem 11.2).

Problem 11.3. Consider a consumer with nonconvex preferences that are represent-ed by the direct utility function U(x1, x2) = x21 + x22. The consumer faces given perunit prices p1 > 0 and p2 = 1 for commodities 1 and 2, and has a budget of $y > 0to spend.(i) What is the consumer’s ordinary demand correspondence?(ii) Prove that the ordinary demand correspondence is upper-hemi-continuous withrespect to p1.(iii) Derive the consumer’s maximum-value (i.e. indirect utility) function v(p1, y).Show that this function is continuous with respect to p1. Is this maximum-valuefunction also differentiable with respect to p1?

Problem 11.4. Consider the constrained optimization problem

maxx∈�n

f(x;α) subject to gj(x; βj) ≤ bj for j = 1, . . . ,m,

where α ∈ A, β1 ∈ B1, . . . , βm ∈ Bm, and bj ∈ � for j = 1, . . . ,m. The objectivefunction f : �n×A→ � is jointly continuous with respect to (x, α). Each constraintfunction gj : �n × Bj → � is jointly continuous with respect to (x, βj), for j =1, . . . ,m. Let β = (β1, . . . , βm)

T, let b = (b1, . . . , bm)T, and let γ = (α : β : b)T.

Let B = B1 × · · · × Bm and let Γ = A × B × �m. The feasible-set correspondenceC : B ×�m �→ �n is

C(β, b) = {x ∈ �n | g1(x; β1) ≤ b1, . . . , gm(x; βm) ≤ bm}.

For all (β, b) ∈ B × �m, suppose that C(β, b) is a bounded subset of �n with anonempty interior. The optimal solution correspondence X∗ : Γ �→ �n is

X∗(γ) = {x∗ ∈ �n | f(x∗;α) ≥ f(x;α) ∀ x ∈ C(β, b)}.

Page 380: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

11.6. PROBLEMS 361

(i) Must X∗(γ) = ∅ for every γ ∈ Γ? Prove your answer.

(ii) Suppose, additionally, that every one of the constraint functions gj is quasi-convexwith respect to x ∈ �n and that the objective function f is quasi-strictly-concave withrespect to x ∈ �n. For any given γ ∈ Γ, what is the cardinality of the set X∗(γ); i.e.how many elements are there in the set X∗(γ)? Prove your answer.

(iii) Suppose that every one of the constraint functions gj is quasi-convex with respectto x ∈ �n and that the objective function f is quasi-strictly-concave with respect tox ∈ �n. Is the optimal solution correspondence X∗ a continuous function on Γ?Prove your answer.

Problem 11.5. Consider the equality constrained optimization problem

maxx∈�n

f(x;α) subject to h(x; β, b) ≡ b− g(x; β) = 0.

Write γ = (α, β, b) ∈ A × B × � = Γ. Suppose that f is jointly concave in (x, α) ∈�n ×A and that h(x; β, b) is jointly concave in (x, β, b) ∈ X ×B×�. Prove that theproblem’s maximum-value function v(γ) is concave on Γ.

Problem 11.6. Most readers will have had some experience with basic game theory.The usual games discussed are Nash games, so-called because the notion of equilibri-um used in these games was developed by John Nash. Nash proved that at least oneNash equilibrium exists for any game consisting of a finite number of players, eachwith von Neumann-Morgenstern preferences (i.e. preferences represented by expectedutility functions), and in which each player possesses a finite number of pure actions.The argument that Nash gave has two major parts to it. The first part is a demon-stration that each player’s “best-response correspondence” is upper-hemi-continuouswith respect to the (typically mixed) actions chosen by all of the other players. Thesecond part, which we will not discuss here, is an application of Kakutani’s Fixed-Point Theorem. Nash games are now a standard part of modeling strategic economicenvironments, so it is useful to understand why it is that each Nash player’s best-response correspondence is upper-hemi-continuous with respect to the actions chosenby the other players. You might think that this is going to be difficult, but it is not.The argument consists of a sequence of quite simple steps, and the problem is brokenup into parts, one part for each step.

We will start with a description of the game. There are n ≥ 2 players. Let’s thinkabout Player 1’s problem. Player 1 can choose from m1 pure actions a11, . . . , a

1m1

.These might be actions such as “sulk,” “tantrum,” “throw something,” and so on.Let’s use A1 to denote the set of pure actions: A1 = {a11, . . . , a1m1

}. Player 1 is allowed

Page 381: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

362 CHAPTER 11. OPTIMAL SOLUTIONS AND MAXIMUM VALUES

Figure 11.6: The unit simplices of dimensions one, two, and three, and some examplesof mixed actions.

to make randomized choices over the pure actions available to her so, for example,she might choose “sulk” with probability 3/4, “tantrum” with probability zero, and“throw something” with probability 1/4. Such a randomization over Player 1’s pureactions is a mixed action for Player 1. It is a probability distribution over all ofthe m1 pure actions available for Player 1 to choose from. So, for Player 1, a mixedaction is a discrete probability distribution

π1 = (π11, . . . , π

1m1

), where π11 ≥ 0, . . . , π1

m1≥ 0 and π1

1 + · · ·+ π1m1

= 1,

where π1i is the probability that she will select her ith pure action a1i . The collection

of all mixed actions available to Player 1 is her mixed action set

Σ1 ={(π11, . . . , π

1m1

)| π1

1 ≥ 0, . . . , π1m1

≥ 0 and π11 + · · ·+ π1

m1= 1}.

This set is the m1-dimensional unit simplex. The one, two, and three-dimensionalunit simplices are displayed in Figure 11.6. A pure action a1i is the probabilisticallydegenerate mixed action in which π1

i = 1 and πij = 0 for all j = 1, . . . ,mi; j = i.

Page 382: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

11.6. PROBLEMS 363

Player 1’s payoff depends upon the pure actions that end up being chosen by allof the players in the game. Let’s use a to denote a complete list of these pure actions,so a = (a1, . . . , an) is an ordered list of the pure actions chosen by Players 1 to n. Wecan also write this list as a = (a1, a−1), where a−1 = (a2, . . . , an) is the list of the pureactions chosen by all of the players other than Player 1; i.e. the superscript “−1”means “excluding Player 1”. Player 1’s payoff function is g1(a).

In a Nash game, each Player i selects a strategy using the belief (this belief is calledthe Nash conjecture) that whatever action is chosen by Player i does not cause anyof the other players to change the actions already selected by them. So, for example,Player 1 believes that she can alter her choice of pure action from “tantrum” to “throwsomething” without causing any of Players 2 to n to alter their chosen actions. Inother words, Player 1 treats the list of actions taken by the other players as parametric;i.e. fixed and given.

We will use π−1 = (π2, . . . , πn) to denote the ordered list of the mixed actionschosen by all of the other Players 2 to n. For a given list π−1, and a given mixedaction choice π1 by Player 1, what is Player 1’s expected payoff? Consider a given listof strategies π1, π2, . . . , πn. Then the probability that Player 1 chooses pure actiona1j1 is π1

j1, the probability that Player 2 chooses pure action a2j2 is π2

j2, and so on.

Thus the probability that the particular list of pure actions chosen by the players isa1j1 , a

2j2, . . . , anjn is

Pr(a1j1 , a

2j2, . . . , anjn

)= π1

j1× π2

j2× · · · × πn

jn .

This is the probability that Player 1’s payoff is g1(a1j1 , a

2j2, . . . , anjn

), so the payoff

expected by Player 1 when the list of strategies for the players is π1, π2, . . . , πn is

f 1(π1; π−1) =

m1∑j1=1

m2∑j2=1

· · ·mn∑jn=1

g1(a1i , a2j2, . . . , anjn)× π1

j1× π2

j2× · · · × πn

jn

=

m1∑j1=1

{m2∑j2=1

· · ·mn∑jn=1

g1(a1i , a2j2, . . . , anjn)× π2

j2× · · · × πn

jn

}︸ ︷︷ ︸

independent of π11 ,...,π

1m1

×π1j1. (11.21)

Player 1 treats the other players’ strategies as fixed and given, and so believes thequantities in the { } braces to be given numbers. Thus Player 1 believes that herexpected payoff function (11.21) is linear with respect to each of π1, . . . , π1

m1.

Player 1 chooses her action π1 so as to maximize her expected payoff given π−1;i.e. Player 1’s problem is:

maxπ11 , ... , π

1m1

∈Σ1f 1(π1; π−1).

Page 383: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

364 CHAPTER 11. OPTIMAL SOLUTIONS AND MAXIMUM VALUES

The optimal solution to this problem is π∗1(π−1), the set of mixed actions (probabilitydistributions) that each maximize Player 1’s expected payoff given the (parametric,she believes) list π−1 of the other players’ actions. This optimal solution is Player 1’sbest-response correspondence.(i) Prove that π∗1(π−1) = ∅, for any list π−1 of the other players’ (mixed) actions.(ii) Prove that π∗1 is an upper-hemi-continuous correspondence of π−1.

Problem 11.7. The 2× 2 table below is the payoff matrix for a Nash game that isplayed by two firms. One firm, the incumbent, is initially the only firm present ina market. The other firm, the entrant, decides between entering the market (action“Enter”) and not entering (action “Stay Out”). The incumbent chooses between theactions of “Fight”ing entry and “Yield”ing to entry. The first element in each payoffpair is the entrant’s payoff. The second is the incumbent’s payoff.(i) Compute the best-response correspondences for the two firms.

Incumbent

Fight Yield

Enter (-1, -1) (1, 1)

Entrant

Stay Out (0, 4) (0, 4)

(ii) Are the best-response correspondences upper-hemi-continuous? Are they convex-valued? What are the domains and ranges of the correspondences?(iii) What is the game’s best-response correspondence? What is its domain andrange? What is its graph? Explain why its graph is closed. With respect to whichspace is the graph closed?

11.7 Answers

Answer to Problem 11.1. The function

f(x) =

{3− (x− 1)2 , if 0 ≤ x ≤ 2

(x− 3)2 , if 2 < x ≤ 4

Page 384: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

11.7. ANSWERS 365

is displayed in Figure 11.7. f is discontinuous at x = 2, is upper-semi-continuous onthe interval [0, 4], attains a global maximum at x = 1, and attains a global minimumat x = 3.

Figure 11.7: Graph of f(x) for problem 11.1.

The function −f(x) is discontinuous at x = 2, is lower-semi-continuous on theinterval [0, 4], attains a global minimum at x = 1, and attains a global maximum atx = 3.

Theorem 11.1 states that an upper-semi-continuous function with a domain thatis a compact set must attain a global maximum. The theorem does not say that sucha function cannot also attain a global minimum. Similarly, Theorem 11.2 states thatan lower-semi-continuous function with a domain that is a compact set must attain aglobal minimum. The theorem does not say that such a function cannot also attaina global maximum.

Answer to Problem 11.2.

Proof of Theorem 11.2. Let f = infx∈C f(x). It must be shown that there is avalue x ∈ C for which the corresponding value of f is f .

Since f = infx∈C f(x), there exists in C a sequence {xn}∞n=1 such that

f(xn) → f. (11.22)

Since C is compact, this sequence must contain a convergent subsequence {xnk}∞k=1.

Let x = limk→∞ xnk. x ∈ C because C is a closed set.

Page 385: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

366 CHAPTER 11. OPTIMAL SOLUTIONS AND MAXIMUM VALUES

Since f is lower-semi-continuous on C,

limk→∞

f(xnk) ≥ f(x). (11.23)

The sequences {f(xn)}∞n=1 and {f(xnk)}∞k=1 must converge to the same value, so it

follows from (11.22) and (11.23) that f ≥ f(x). But, by definition, f ≤ f(x) forevery x ∈ C, so it must be that f = f(x).

Answer to Problem 11.3.(i) The consumer’s ordinary demand correspondence is

X∗(p1, y) =

⎧⎪⎪⎪⎨⎪⎪⎪⎩

(y

p1, 0

), if 0 < p1 < 1

{(y, 0), (0, y)} , if p1 = 1

(0, y) , if 1 < p1.

(11.24)

(ii) For 0 < p1 < 1 and 1 < p1, the demand correspondence is single-valued and con-tinuous, and thus is upper-hemi-continuous for p1 = 1. The demand correspondenceis not continuous at p1 = 1.

Take any sequence {pn}∞n=1 with pn ≥ 1 for all n ≥ 1 and pn → 1. Thecorresponding sequence of ordinary demands is {(0, y)}∞n=1, which is convergent to(0, y) ∈ X∗(p1=1, y). Now take any sequence {pn}∞n=1 with 0 < pn < 1 for all n ≥ 1and pn → 1. Then the corresponding sequence of ordinary demands is {(y/p1, 0)}∞n=1,which converges to (y, 0) ∈ X∗(p1 = 1, y). Thus X∗(p1, y) is upper-hemi-continuousat p1 = 1. This, together with the above paragraph, establishes that X∗(p1, y) isupper-hemi-continuous with respect to p1 for p1 > 0.(iii) The consumer’s indirect utility is

v(p1, y) =

⎧⎪⎪⎪⎨⎪⎪⎪⎩

(y

p1

)2

, if 0 < p1 < 1

y2 , if p1 = 1

y2 , if 1 < p1

=

⎧⎪⎨⎪⎩(y

p1

)2

, if 0 < p1 < 1

y2 , if 1 ≤ p1.

As p1 → 1−, (y/p1)2 → y2, so v(p1, y) is continuous for all p1 > 0.v is not differentiable at p1 = 1. Why? For p1 > 1, the value of ∂v/∂p1 is zero.

For 0 < p1 < 1, the value of ∂v/∂p1 = −2y2/p31, so, as p1 → 1−, ∂v/∂p1 → −2y2 = 0.Thus v is not differentiable with respect to p1 at p1 = 1. This example makes thepoint, that while the Theorem of the Maximum gives conditions that are sufficient

Page 386: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

11.7. ANSWERS 367

for a maximum-value function to be continuous, these same conditions are generallynot sufficient for the maximum-value function to be differentiable.

Answer to Problem 11.4.(i) Proof. Yes. Take any γ ∈ Γ. The feasible set C(β, b) = ∅. Also,

C(β, b) = {x ∈ �n | g1(x; β1) ≤ b1, . . . , gm(x; βm) ≤ bm}

=m⋂j=1

{x ∈ �n | gj(x; βj) ≤ bj}.

Since each constraint function gj is continuous jointly with respect to (x, βj), it iscontinuous with respect to x alone for fixed βj. Therefore gj is lower-semi-continuouson �n for any given βj, and so the sets {x ∈ �n | gj(x; βj) ≤ bj} are all closed in En.Hence C(β, b) is the intersection of a finite number of sets that are closed in En, andso it too is a closed subset of En. Since it is given above that any feasible set isnot empty and is bounded in En, it follows that C(β, b) is a compact and nonemptysubset of En.

The objective function f is jointly continuous with respect to (x, α), and so iscontinuous with respect to x alone for any fixed value of α. Hence, for any givenγ′ ∈ Γ, the constrained optimization problem is to maximize a function f(x;α′) thatis continuous with respect to x on a nonempty and compact subset C(β′, b′) of �n.Weierstrass’s Theorem asserts that there is at least one maximizing (optimal) solutionto this problem; i.e. X∗(γ′) = ∅. Since γ′ is an arbitrary choice from Γ, X∗(γ) = ∅for every γ ∈ Γ.(ii) X∗(γ) is a singleton; i.e. #X∗(γ) ≡ 1 on Γ.Proof. Take any γ ∈ Γ. Since gj(x; βj) is quasi-convex with respect to x, the lower-contour set {x ∈ � | gj(x; βj) ≤ bj} is a convex subset of �n. Hence C(β, b) is theintersection of convex sets, and so is a convex subset of �n.

Since f(x;α) is quasi-strictly-concave with respect to x ∈ �n, the upper-contourset {x ∈ �n | f(x;α) ≥ k} is a strictly convex subset of �n if it has a nonemptyinterior, or is weakly convex if it is a singleton or is empty.

It has been established that X∗(γ) = ∅, so #X∗(γ) ≥ 1. Suppose #X∗(γ) > 1.Then there exist at least two optimal solutions x′, x′′ ∈ X∗(γ) with x′ = x′′. Denotethe maximum value provided by these solutions by

v = f(x′;α) = f(x′′;α).

Since x′ and x′′ are optimal solutions, they are also feasible solutions, meaning thatx′, x′′ ∈ C(β, b), a convex set. So, for any μ ∈ [0, 1], x(μ) = μx′ + (1 − μ)x′′ ∈ C(γ);

Page 387: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

368 CHAPTER 11. OPTIMAL SOLUTIONS AND MAXIMUM VALUES

i.e. x(μ) is a feasible solution. But, since f is quasi-strictly-concave with respect tox, the objective function value provided by x(μ) is

f(μx′ + (1− μ)x′′;α) > f(x′;α) = f(x′′;α) = v.

This contradicts the optimality of x′ and x′′, so #X∗(γ) > 1. Hence #X∗(γ) = 1.(iii) Proof. Yes. Since #X∗(γ) ≡ 1 on Γ, X∗ is a function. Since X∗ is upper-hemi-continuous on Γ, it is a continuous function.

Answer to Problem 11.5. Choose γ′, γ′′ ∈ Γ with γ′ = γ′′. Use x∗(γ′) to denote asolution that is optimal for γ′ = (α′, β′, b′) and use x∗(γ′′) to denote a solution thatis optimal for γ′′ = (α′′, β′′, b′′). Then

v(γ′) = f(x∗(γ′);α′) and h(x∗(γ′); β′, b′) ≥ 0, (11.25)

and v(γ′′) = f(x∗(γ′′);α′′) and h(x∗(γ′′); β′′, b′′) ≥ 0. (11.26)

Since h is jointly concave in (x, β, b), for any μ ∈ [0, 1],

h(μx∗(γ′) + (1− μ)x∗(γ′′), μβ′ + (1− μ)β′′, μb′ + (1− μ)b′′)

≥ μh(x∗(γ′); β′, b′) + (1− μ)h(x∗(γ′′); β′′, b′′) ≥ 0(11.27)

by (11.25) and (11.26). Hence μx∗(γ′) + (1 − μ)x∗(γ′′) is a feasible point when γ =μγ′ + (1− μ)γ′′. Then

v(μγ′ + (1− μ)γ′′) = f(x∗(μγ′ + (1− μ)γ′′);μα′ + (1− μ)α′′)

≥ f(μx∗(γ′) + (1− μ)x∗(γ′′);μα′ + (1− μ)α′′) (11.28)

≥ μf(x∗(γ′);α′) + (1− μ)f(x∗(γ′′);α′′) (11.29)

= μv(γ′) + (1− μ)v(γ′′),

where (11.28) follows from the fact that x = μx∗(γ′) + (1 − μ)x∗(γ′′) need not beoptimal for γ = μγ′ + (1 − μ)γ′′, and where (11.29) follows from the joint concavityof f with respect to (x, α).

Answer to Problem 11.6. Player 1’s problem

maxπ1∈Σ1

f 1(π1; π−1).

is to select a mixed action π1 from her set Σ1 of available mixed actions so as tomaximize her expected payoff, given the list π−1 of the other players’ chosen mixedactions.

Page 388: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

11.7. ANSWERS 369

(i) The problem’s feasible set is the unit simplex

Σ1 ={(π11, . . . π

1m1

)| π1

1 ≥ 0, . . . , π1m1

≥ 0, π11 + · · ·+ π1

m1= 1}.

This set is not empty, is convex, and is compact with respect to Em1 . The problem’sobjective function f 1(π1; π−1) is linear, and therefore continuous, with respect to eachof π1

1, . . . , π1m1

. Weierstrass’s Theorem asserts that there is a mixed action π∗1 ∈ Σ1

that maximizes Player 1’s expected payoff given the list π−1 of the mixed actionschosen by Players 2 to n; i.e. π∗1 (π−1) = ∅.(ii) Does Player 1’s choice problem satisfy the conditions of Berge’s Theorem of theMaximum? The problem’s objective function (11.21) is jointly continuous with re-spect to each of π1

1, . . . , π1m1

, each of π21, . . . , π

2m2

, and so on, up to each of πn1 , . . . , π

nmn

.The value of the problem’s feasible set correspondence is constant at Σ1 because,by assumption, in a Nash game the pure actions available to any player cannot bechanged by any other player(s). In this trivial sense, then, the problem’s feasible setcorrespondence is continuous with respect to the parameter (from Player 1’s perspec-tive) vector π−1. As well, the set Σ1 is not empty, and is compact with respect to Em1 .Thus, Berge’s Theorem may be applied to Player 1’s problem, establishing that theproblem’s optimal solution correspondence π∗1 is upper-hemi-continuous on the setΣ2 × · · · × Σn that contains all of the possible lists of mixed actions for Players 2to n.

Answer to Problem 11.7. Let py denote the probability that the incumbent willchoose action “Yield.” Then 1− py is the probability that the incumbent will chooseaction “Fight.” Let pe denote the probability that the potential entrant will chooseaction “Enter.” Then 1− pe is the probability that the potential entrant will chooseaction “Stay Out.” In other words, the incumbent’s mixed strategy is σI = (py, 1−py)and the potential entrant’s mixed strategy is σE = (pe, 1− pe).

(i) Given the incumbent’s strategy, the expected value to the potential entrant, E, ofchoosing action “Enter” is

EVEEnter = 1× py + (−1)× (1− py) = 2py − 1.

The expected value to the potential entrant of choosing action “Stay Out” in responseto the incumbent’s strategy is

EVEStay Out = 0.

So the potential entrant’s best-response correspondence is

Page 389: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

370 CHAPTER 11. OPTIMAL SOLUTIONS AND MAXIMUM VALUES

σ∗E = (p∗e, 1− p∗e) =

⎧⎪⎨⎪⎩(0, 1) , if 0 ≤ py < 1/2

(α, 1− α) for any α ∈ [0, 1] , if py = 1/2

(1, 0) , if 1/2 < py ≤ 1.

(11.30)

Similarly, the incumbent’s best-response correspondence is

σ∗I = (p∗y, 1− p∗y) =

{(β, 1− β) for any β ∈ [0, 1] , if pe = 0

(1, 0) , if 0 < pe ≤ 1.(11.31)

Figure 11.8 displays the graphs of the best-response correspondences.

Figure 11.8: The entrant’s and incumbent’s best-response correspondences.

(ii) σ∗E maps from the strategy space of the incumbent into the strategy space ofthe potential entrant. That is, σ∗E : S �→ S, where S = {(p, 1 − p) | 0 ≤ p ≤ 1} isthe unit simplex in �2. Similarly, σ∗I maps from the strategy space of the potentialentrant into the strategy space of the incumbent. That is, σ∗E : S �→ S.

When 0 ≤ py < 1/2, σ∗E((py, 1 − py)) is the singleton set containing only thevertex (0, 1) of the simplex S. This is a compact and weakly convex subset of �2.When 1/2 < py ≤ 1, σ∗E((py, 1− py)) is the singleton set containing only the vertex(1, 0) of the simplex S. This too is a compact and weakly convex subset of �2. Whenpy = 1/2, σ∗E((py, 1 − py)) is the interval that is the edge of the simplex S between

Page 390: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

11.7. ANSWERS 371

the vertex (0, 1) and the vertex (1, 0). This is a compact and weakly convex subsetof �2. Since the image of σ∗E is always a compact and convex subset of �2, σ∗E is acompact-valued and convex-valued correspondence. Similarly, σ∗I is a convex-valued,compact-valued correspondence also.

Both σ∗E and σ∗I are upper-hemi-continuous correspondences. There are twoways to demonstrate this. First, we can use Berge’s Theorem of the Maximum.Second, we can use the Closed Graph Theorem. To apply Berge’s Theorem, noticethat the incumbent’s problem is to maximize an expected value function that iscontinuous in the probabilities pe and py. Next, the incumbent’s choice set is theunit simplex in �2, which is a compact set. Last, this unit simplex choice set isindependent of the potential entrant’s strategic choice, so, in this trivial sense, theincumbent’s choice set is a compact-valued and continuous correspondence of thepotential entrant’s strategy choice. Berge’s Theorem now applies and we concludethat the incumbent’s best-response correspondence is upper-hemi-continuous on thepotential entrant’s strategy space.

To apply the Closed Graph Theorem, notice from (11.30) that the union of all ofthe images of σ∗E is just the one-dimensional unit simplex, which is compact in �2.The graph of σ∗E is the set

graph(σ∗E) = {(p∗e, 1− p∗e, py) | p∗e = 0 for 0 ≤ py < 1/2, p∗e ∈ [0, 1] for py = 1/2,

and p∗e = 1 for 1/2 < py ≤ 1}.

This set is closed in �3 so the Closed Graph Theorem informs us that the potentialentrant’s best-response correspondence is upper-hemi-continuous on the unit sim-plex that is the incumbent’s strategy space. Similarly, the incumbent’s best-responsecorrespondence is upper-hemi-continuous on the unit simplex that is the entrant’sstrategy space.(iii) The game’s best-response correspondence σ∗ lists the individual players’ best-response correspondences; i.e. σ∗ = (σ∗E, σ∗I). σ∗ maps from the game’s strategyspace, which is the direct product of the two players’ strategy spaces, to the samespace; i.e. σ∗ : S2 �→ S2. Written out in detail, σ∗ is

Page 391: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

372 CHAPTER 11. OPTIMAL SOLUTIONS AND MAXIMUM VALUES

σ∗((pe, 1− pe), (py, 1− py)) =⎧⎪⎪⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎪⎪⎩

(0, 1), (β, 1− β) ∀ β ∈ [0, 1] , if 0 ≤ py < 1/2 and pe = 0

(0, 1), (1, 0) , if 0 ≤ py < 1/2 and 0 < pe ≤ 1

(α, 1− α) ∀ α ∈ [0, 1], (β, 1− β) ∀ β ∈ [0, 1] , if py = 1/2 and pe = 0

(α, 1− α) ∀ α ∈ [0, 1], (1, 0) , if py = 1/2 and 0 < pe ≤ 1

(1, 0), (β, 1− β) ∀ β ∈ [0, 1] , if 1/2 < py ≤ 1 and pe = 0

(1, 0), (1, 0) , if 1/2 < py ≤ 1 and 0 < pe ≤ 1.

Figure 11.9: The entry game’s best-response correspondence.

The graph of this correspondence is closed in �4. A lower-dimensional drawing ofthe game’s best-response correspondence can be constructed in (pe, py)-space, whichis [0, 1]2; see Figure 11.9. It is constructed by superimposing on each other the abovediagrams for (11.30) and (11.31) (see Figure 11.8 – one of the diagrams has to berotated by 90◦ to have both diagrams on the same axes).

Page 392: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

Chapter 12

Comparative Statics Analysis,Part 1

12.1 What Is Comparative Statics Analysis?

Comparative statics analysis is the comparison of the values taken by the optimalsolution correspondence to a static optimization problem for different parameter val-ues. Understand that comparative statics analysis has nothing to do with the studyof dynamic problems. When we talk of changes in a dynamic problem or its solution,we are speaking of changes that occur through time. Time does not exist in a staticproblem.

maxx∈�n

f(x;α) subject to gj(x; βj) ≤ bj for all j = 1, . . . ,m. (12.1)

If you have not already done so, then you should read Section 11.2 to understand thenotation being used in (12.1). The feasible set correspondence, the optimal solutioncorrespondence, and the maximum-value function for the problem are

C(β, b) ≡ {x ∈ �n | gj(x; βj) ≤ bj ∀ j = 1, . . . ,m},X∗(γ) = X∗(α, β, b) ≡ {x∗ ∈ C(β, b) | f(x∗;α) ≥ f(x;α) ∀ x ∈ C(β, b)},

and v(γ) ≡ f(x∗;α) where x∗ ∈ X∗(α, β, b).

In the example that is problem (11.4), the parameters not assigned numericalvalues are the elements of the parameter vector γ = (p, w1, w2, β1, β2). These arethe parameters that are of interest for comparative statics analyses of the problem.The typical comparative statics experiment is where one of a problem’s parameters is

373

Page 393: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

374 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

varied. The experimenter then observes the changes caused to the problem’s optimalsolution(s). For example, if in problem (11.4) the value of the firm’s product’s pricep is altered, then the experimenter observes the consequences of a slight change to p.These will be changes to the quantity of product supplied by the firm, changes to thequantities used of the firm’s inputs, changes to the values of the multipliers of theconstraints that bind at the profit-maximizing solution to the firm’s problem, and achange to the firm’s maximum profit level. Similarly, the experimenter could slightlyalter the price w1 of input 1 and observe the consequences. Using such experiments toanalyze the properties of the firm’s profit-maximizing production plan is a comparativestatics analysis of the firm’s problem.

More generally, two different values γ′ and γ′′ of the parameter vector γ definetwo parametrically different versions of problem (12.1), and typically cause two d-ifferent values X∗(γ′) and X∗(γ′′) of the optimal solution correspondence and twodifferent values v(γ′) and v(γ′′) of the maximum-value function. Comparing thesetwo optimal solution values and the two maximized values of the objective functionhelps us to determine the properties of both the optimal solution correspondenceand the maximum-value function. Sometimes this is easily done. In problem (11.4),for example, we were able to discover explicit algebraic equations for the problem’soptimal solution; see (11.5). These equations directly tell us everything there is toknow about the properties of the optimal solution to problem (11.4). For example,we observe from (11.5a) that the quantity y∗ that the firm supplies of its productfalls as the price w2 of input 2 increases. This is a qualitative comparative staticsresult. Again from (11.5a), by differentiation we can determine the value of the rateat which the firm’s quantity supplied decreases as w2 increases. This is a quantita-tive comparative statics result. So comparative statics analyses can be divided intotwo types, qualitative and quantitative. In a qualitative analysis, we are interestedonly in the directions of change of the parts of the optimal solution as values of theproblem’s parameters are changed. In a quantitative analysis we want to know aswell the numerical magnitudes of the changes or rates of change. In this chapter wewill consider how to obtain quantitative comparative statics results. In the followingchapter we will confine ourselves to obtaining qualitative comparative statics results.

12.2 Quantitative Comparative Statics Analysis

Let us begin with a simple example. Consider the problem

maxx

f(x;α1, α2, α3) = −α1x2 + α2x+ α3 with α1 > 0. (12.2)

Page 394: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.2. QUANTITATIVE COMPARATIVE STATICS ANALYSIS 375

The necessary first-order optimality condition is

df(x;α1, α2, α3)

dx|x=x∗(α1,α2,α3) = −2α1x

∗(α1, α2, α3) + α2 = 0

⇒ x∗(α1, α2, α3) =α2

2α1

.(12.3)

Because we have an explicit equation for the problem’s optimal solution, it is easyto discover all of the properties of the optimal solution. For instance, the optimalsolution decreases as α1 increases, increases as α2 increases, and is independent of α3.

More commonly, however, there will not exist an algebraically explicit optimalsolution. What can we learn in such a case? Consider the problem

maxx≥0

f(x; γ) = x(1− ln (γ + x)). (12.4)

We will restrict the parameter γ to values from zero to unity: γ ∈ [0, 1]. The necessaryfirst-order optimality condition is

0 =df(x; γ)

dx= 1− ln (γ + x)− x

γ + xat x = x∗(γ). (12.5)

It is not possible to rewrite this equation so that the optimal solution mapping x∗(γ)is stated as an explicit function of γ. Yet, for any given value for γ ∈ [0, 1], there is anoptimizing value x∗(γ) for x; see Figure 12.1. In this case, then, the optimal solutionto the problem is defined only implicitly by the first-order maximization condition.For such cases the Implicit Function Theorem (stated below) is often used to obtaindifferential quantitative comparative statics results. The procedure is straightforward.The key is to realize that, for any given value of γ, at the corresponding optimalvalue x∗(γ) for x, the first-order maximization condition (12.5) necessarily has thevalue zero. Therefore, when evaluated only at optimal values of x, the first-ordermaximization condition (12.5) is identically zero:

0 ≡ 1− ln (γ + x∗(γ))− x∗(γ)γ + x∗(γ)

∀ γ ∈ [0, 1]. (12.6)

The right-hand side of (12.6) is a function only of γ (because x∗(γ) is a function ofγ), so, in recognition of this, let us write (12.6) as

I(γ) ≡ 1− ln (γ + x∗(γ))− x∗(γ)γ + x∗(γ)

≡ 0 ∀ γ ∈ [0, 1]. (12.7)

Page 395: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

376 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

Figure 12.1: The graph of the implicit optimal solution x∗(γ) to problem (12.4).

The identity mapping I(γ) is constant-valued (at zero) so its rate of change withrespect to γ must be zero. Accordingly, using the Chain Rule to differentiate (12.7)with respect to γ gives

dI(γ)

dγ= −

γ + (2γ + x∗(γ))dx∗(γ)dγ

(γ + x∗(γ))2= 0. (12.8)

The numerator of (12.8) necessarily is zero, which tells us that the rate of changewith respect to γ of the value of the implicit optimal solution x∗(γ) is

dx∗(γ)dγ

= − γ

2γ + x∗(γ). (12.9)

So what have we discovered? We have an expression for the rate at which the optimalsolution to problem (12.4) changes as the problem’s parameter γ changes. But sincewe do not have an algebraic equation for the optimal solution x∗(γ), we do not havean algebraic equation for this rate of change. Have we, then, learned anything newat all? Well, yes – actually we have learned a great deal. First, notice from (12.7),that if γ = 0, then we can easily determine that the value of the optimal solution is

Page 396: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.3. IMPLICIT FUNCTION THEOREMS 377

Table 12.1: x∗(γ) and its order 2 Taylor’s series approximation.

γ 0 0·001 0·01 0·1 0·2Δγ 0 0·001 0·01 0·1 0·2

x∗(γ) 1 0·9999995 0·99995 0·9956 0·984

1− (Δγ)2

21 0·9999900 0·99990 0·9900 0·960

unity: x∗(0) = 1. So then, from (12.9), we know that the rate of change for x∗(γ)with respect to γ is zero at γ = 0 (check that this is so in Figure 12.1). And if wedifferentiate (12.9), then we obtain

d2x∗(γ)dγ2

=

γdx∗(γ)dγ

− x∗(γ)

(2γ + x∗(γ))2= − (γ + x∗(γ))2

(2γ + x∗(γ))3(12.10)

after substitution from (12.9). Since x∗(0) = 1 we have that, at γ = 0, the value ofthe second derivative of x∗(γ) at γ = 0 is −1. Now we can use an order-2 Taylor’sseries to assert that, for values of γ = 0+Δγ that are close to γ = 0, the value of theoptimal solution x∗(γ) is close to

x∗(0 + Δγ) ≈ x∗(0) +dx∗(γ)dγ

|γ=0Δγ +1

2

d2x∗(γ)dγ2

|γ=0(Δγ)2 = 1− (Δγ)2

2.

Table 12.1 lists numerically computed values of x∗(γ) and of the approximation 1 −(Δγ)2/2 for various values of Δγ. The table shows that the approximation for x∗(γ)is quite accurate for small values of Δγ. All of this is achieved without having aformula for the optimal solution x∗(γ) – that’s quite impressive.

12.3 Implicit Function Theorems

The argument we used in going from (12.7) to (12.8) and then to (12.9) is an explana-tion of the use of the Implicit Function Theorem for a single equation. The theoremis a result often used by economists when undertaking comparative statics analyses.

Page 397: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

378 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

Theorem 12.1 (Implicit Function Theorem for a Single Equation). Let I(x, γ1, . . . , γk)be a C1 function defined in a neighborhood about a point (x′, γ′1, . . . , γ

′k). Let c denote

the value of I at the point (x′, γ′1, . . . , γ′k); i.e. I(x

′, γ′1, . . . , γ′k) = c. If

∂I(x, γ1, . . . , γk)

∂x= 0 at (x′, γ′1, . . . , γ

′k),

then there exists an ε > 0 and a C1 function x = g(γ1, . . . , γk) defined over the ballBdE ((γ′1, . . . , γ

′k) , ε), such that

(i) I(g(γ1, . . . , γk), γ1, . . . , γk) ≡ c on BdE ((γ′1, . . . , γ′k) , ε),

(ii) x′ = g(γ′1, . . . , γ′k), and

(iii) for j = 1, . . . , k, the rate of change of g with respect to γj at (γ′1, . . . , γ′k) is

∂g(γ1, . . . , γk)

∂γj

∣∣γ1=γ′

1

...γk=γ′

k

= − ∂I(x, γ1, . . . , γk)/∂γj∂I(x, γ1, . . . , γk)/∂x

∣∣x=x′γ1=γ′

1

...γk=γ′

k

.

Let’s take careful note of some features of this result. First, the function I iscontinuously differentiable, not just differentiable, to order 1 with respect to each ofx, γ1, . . . , γk. In our example above, the function is I(x, γ) = 1− ln (γ + x)−γ/(γ+x)(see (12.5)). The first-order derivatives of I(x, γ) are

∂I(x, γ)

∂γ= − γ

(γ + x)2and

∂I(x, γ)

∂x= − 2γ + x

(γ + x)2.

Both of these derivative functions are continuous in a neighborhood about (x, γ) =(1, 0), so I is continuously differentiable to order 1 over the neighborhood. Moreoverthe value at (x, γ) = (1, 0) of ∂I/∂x = −1 = 0. The theorem then assures us that,in some (possibly small) neighborhood centered on (x, γ) = (1, 0), there exists acontinuously differentiable to order 1 implicit function x = g(γ), such that I(γ) ≡I(g(γ), γ) ≡ 0 over the neighborhood and the rate of change of g(γ) at γ = 0 is

dg(γ)

dγ|γ=0 = − ∂I(x, γ)/∂γ

∂I(x, γ)/∂x

∣∣x=1γ=0

= − 0

−1= 0,

just as we determined from (12.9).In most of the problems that interest economists, there are several choice variables

x1, . . . , xn to be solved for in terms of several parameters γ1, . . . , γk. The ImplicitFunction Theorem extends to these more general cases. I’ll state this version of thetheorem and then explain what it says. The theorem looks a bit ugly but don’t bedeterred. It is just talking about when we can solve linear equation systems, a topicwe considered in Section 2.3.

Page 398: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.3. IMPLICIT FUNCTION THEOREMS 379

Theorem 12.2 (Implicit Function Theorem for an Equation System).Let I1(x1, . . . , xn, γ1, . . . , γk), . . . , In(x1, . . . , xn, γ1, . . . , γk) be C1 functions defined ina neighborhood about a point (x′1, . . . , x

′n, γ

′1, . . . , γ

′k). Let ci denote the value of Ii at

the point (x′1, . . . , x′n, γ

′1, . . . , γ

′k); i.e. Ii(x

′1, . . . , x

′n, γ

′1, . . . , γ

′k) = ci for i = 1, . . . , n.

Let det JI(x′, γ′) denote the value at (x′1, . . . , x

′n, γ

′1, . . . , γ

′k) of the determinant of the

Jacobian matrix

JI(x, γ) =

⎛⎜⎜⎜⎜⎜⎝∂I1∂x1

· · · ∂I1∂xn

.... . .

...

∂In∂x1

· · · ∂In∂xn

⎞⎟⎟⎟⎟⎟⎠ .

If det JI(x′, γ′) = 0, then, there exists an ε > 0, and C1 functions

x1 = g1(γ1, . . . , γk), . . . , xn = gn(γ1, . . . , γk)

defined over the ball BdE ((γ′1, . . . , γ′k) , ε) , such that

(i) Ii(g1(γ1, . . . , γk), . . . , gn(γ1, . . . , γk), γ1, . . . , γk) ≡ ci on BdE ((γ′1, . . . , γ′k) , ε) for

each i = 1, . . . , n,(ii) x′i = gi(γ

′1, . . . , γ

′k) for i = 1, . . . , n, and

(iii) for each j = 1, . . . , k, the values of the rates of change of xi with respect to γj at(γ′1, . . . , γ

′k) are the solution to

JI(x′, γ′)

⎛⎜⎜⎜⎜⎝∂x1∂γj...

∂xn∂γj

⎞⎟⎟⎟⎟⎠ = −

⎛⎜⎜⎜⎜⎝∂I1∂γj...∂In∂γj

⎞⎟⎟⎟⎟⎠∣∣∣∣∣x1=x′

1

...xn=x′

nγ1=γ′

1

...γk=γ′

k

.

Let’s look at a simple example in which we must solve for two choice variablevalues. Consider the problem

maxx1, x2

f(x1, x2; γ) = −(x1 − γ)(x2 − 2γ) = −x1x2 + 2γx1 + γx2 − 2γ2.

The optimal solution is (x∗1(γ), x∗2(γ)) = (γ, 2γ). The comparative-statics results are

dx∗1(γ)/dγ = 1 and dx∗2(γ)/dγ = 2. How do we obtain these results using the abovetheorem? The first-order conditions, necessarily true at the optimal solution, are

∂f(x1, x2; γ)

∂x1= −x2 + 2γ = 0 and

∂f(x1, x2; γ)

∂x2= −x1 + γ = 0

Page 399: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

380 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

at (x1, x2) = (x∗1(γ), x∗2(γ)). Let’s change our notation to match that used in the above

theorem. Then we have that, at the point (x1, x2, γ) = (x∗1(γ), x∗2(γ), γ), necessarily

I1(x1, x2, γ) ≡ −x2 + 2γ = c1 = 0 and I2(x1, x2, γ) ≡ −x1 + γ = c2 = 0.

The Jacobian matrix of this equation system is

JI(x1, x2, γ) =

⎛⎜⎜⎝∂I1∂x1

∂I1∂x2

∂I2∂x1

∂I2∂x2

⎞⎟⎟⎠ =

(0 −1−1 0

).

The vector ⎛⎜⎜⎝∂I1∂γ

∂I2∂γ

⎞⎟⎟⎠ =

(21

).

The value of the determinant of the Jacobian matrix is −1 = 0, so the Implicit Func-tion Theorem says that the comparative statics quantities dx∗1(γ)/dγ and dx∗2(γ)/dγare the solution to

JI(x1, x2, γ)

⎛⎜⎜⎝dx∗1(γ)dγ

dx∗2(γ)dγ

⎞⎟⎟⎠ =

(0 −1−1 0

)⎛⎜⎜⎝dx∗1(γ)dγ

dx∗2(γ)dγ

⎞⎟⎟⎠ = −

(21

). (12.11)

The solution is as we discovered earlier: dx∗1(γ)/dγ = 1 and dx∗2(γ)/dγ = 2. Why?What’s going on here? It’s simple. We know that, if we substitute into the first-order conditions the optimizing values x∗1(γ) and x

∗2(γ) for x1 and x2, then the values

of the first-order condition have to be zeros; i.e. the first-order conditions becomezero-valued identities,

I1(x∗1(γ), x

∗2(γ), γ) = −x∗2(γ) + 2γ ≡ 0 and I2(x

∗1(γ), x

∗2(γ), γ) = −x∗1 + γ ≡ 0.

Both are constant (at zero) so their rates of change with respect to γ must be zero;i.e. the values of the derivatives of these identities with respect to γ must both bezero. Using the Chain Rule, then, we get

∂I1∂x1

× dx∗1(γ)dγ

+∂I1∂x2

× dx∗2(γ)dγ

+∂I1∂γ

= 0× dx∗1(γ)dγ

+ (−1)× dx∗2(γ)dγ

+ 2 = 0

and∂I2∂x1

× dx∗1(γ)dγ

+∂I2∂x2

× dx∗2(γ)dγ

+∂I2∂γ

= (−1)× dx∗1(γ)dγ

+ 0× dx∗2(γ)dγ

+ 1 = 0.

Page 400: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.4. STARTING POINTS 381

These are exactly the same as their matrix version, (12.11). The nonzero value of thedeterminant of the Jacobian matrix is an insistence that the n first-order conditions,one with respect to each of x1, . . . , xn, yield n linearly independent equations in the ncomparative statics quantities ∂x∗1(γ)/∂γ, . . . , ∂x

∗n(γ)/∂γ for which we wish to solve.

The discussion in this section is intended only as an explanation of how the Im-plicit Function Theorem may be used to compute the values of comparative staticsquantities. Nothing here is a proof.

12.4 Starting Points

Generally, the goal of a differential comparative statics analysis is to obtain values,for a particular given value of the parameter vector γ, of the derivatives ∂x∗i /∂γk and∂λ∗j/∂γk that measure quantitatively the rate of change of the optimal value x∗i ofthe ith choice variable and the rate of change of the value λ∗j of the jth constraint’smultiplier as the value of the kth parameter γk changes. The good news is that theprocedure used to obtain these results is not difficult. It is, in fact, just the procedureused above in our last example. The bad news is that often the algebra required islengthy and tedious.

There are two starting points for obtaining quantitative differential comparativestatics results for a constrained optimization problem. These are the problem’s first-order necessary optimality conditions and the problem’s maximum-value function.These two starting points provide different types of comparative statics results, as weshall see. In Section 12.6 we will see what types of results we can obtain from theproblem’s first-order conditions. In Section 12.7 we will see the types of results thatwe can obtain from the problem’s maximum-value function.

12.5 Assumptions

To undertake differentiable comparative statics analyses of problem (12.1), we imposeconditions sufficient for the problem’s optimal solution correspondence to be a vectorof differentiable functions.

Assumption 1 (Unique Solution). X∗(γ) is a singleton for every γ ∈ Γ. Inrecognition of this the optimal solution function will be written as x∗ : Γ �→ �n,where

f(x∗(γ);α) > f(x;α) ∀ x ∈ C(β, b), x = x∗(γ).

Page 401: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

382 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

Assumption 2 (Constraint Qualification). For every γ ∈ Γ, the optimal solutionx∗(γ) is a regular point in C(β, b).Assumption 3 (Twice-Continuous Differentiability). f, g1, . . . , gm are all twice-continuously differentiable with respect to each choice variable xi and with respectto each parameter γk.

These assumptions are collectively sufficient for the optimal solution (x∗(γ), λ∗(γ))to be the unique, continuously differentiable solution to the first-order optimalityconditions for problem (12.1). That is, there exist functions x∗i : Γ → � for i =1, . . . , n and λ∗j : Γ �→ �+ for j = 1, . . . ,m, such that, for any given γ ∈ Γ, the values(x∗1(γ), . . . , x

∗n(γ), λ

∗1(γ), . . . , λ

∗m(γ)) uniquely solve

∂f(x;α)

∂xi

∣∣x=x∗(γ) =

m∑j=1

λ∗j(γ)∂gj(x; βj)

∂xi

∣∣x=x∗(γ) ∀ i = 1, . . . , n, (12.12)

λ∗j(γ)(bj − gj(x∗(γ); βj)) = 0 ∀ j = 1, . . . ,m, (12.13)

with λ∗j(γ) ≥ 0 ∀ j = 1, . . . ,m, (12.14)

and gj(x∗) ≤ bj ∀ j = 1, . . . ,m. (12.15)

The maximum-value function v : Γ → � for the constrained optimization problem is

v(γ) ≡ f(x∗(γ);α) +m∑j=1

λ∗j(γ)(bj − gj(x∗(γ); βj)). (12.16)

Since each x∗i (γ) and each λ∗j(γ) is a function that is continuously differentiable withrespect to each parameter γk, the maximum-value function v is also a continuouslydifferentiable function of each γk.

12.6 Using the First-Order Conditions

We start by writing down the necessary first-order optimality conditions for the givenproblem. Then we note that these equations hold as identities when evaluated at theoptimal solution (x∗1(γ), . . . , x

∗n(γ), λ

∗1(γ), . . . , λ

∗m(γ)). Next we differentiate each of

these identities with respect to whichever one of the parameters γk that is of interestto us. This gives us n + m linear equations in the n + m unknowns that are thecomparative statics quantities ∂x∗1(γ)/∂γk, . . . , ∂λ

∗m(γ)/∂γk. So long as these n +m

equations are linearly independent, we can compute the values of these derivativesat the optimal solution to the problem. That is all there is to it. Let’s have a moredetailed look at the procedure.

Page 402: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.6. USING THE FIRST-ORDER CONDITIONS 383

Our constrained optimization problem is (12.1). For a given value of the parametervector γ, the optimal solution is (x∗(γ), λ∗(γ)) = (x∗1(γ), . . . , x

∗n(γ), λ

∗1(γ), . . . , λ

∗m(γ)).

Slack constraints are of no interest, so, without loss of generality, suppose that thebinding (both strictly binding and just-binding) constraints are those indexed byj = 1, . . . , �, where 1 ≤ � ≤ m. Consequently λ∗1(γ) ≥ 0, . . . , λ∗�(γ) ≥ 0 (> 0for each strictly binding constraint and = 0 for each just-binding constraint) andλ∗�+1(γ) = · · · = λ∗m(γ) = 0. The necessary first-order optimality conditions are the nKarush-Kuhn-Tucker conditions

∂f(x;α)

∂x1− λ1

∂g1(x; β1)

∂x1− · · · − λ�

∂g�(x; β�)

∂x1= 0

......

∂f(x;α)

∂xn− λ1

∂g1(x; β1)

∂xn− · · · − λ�

∂g�(x; β�)

∂xn= 0

(12.17)

and the � complementary slackness conditions

λ1(b1 − g1(x; β1)) = 0, . . . , λ�(b� − g�(x; β�)) = 0. (12.18)

These n + � equations are functions of x1, . . . , xn, λ1, . . . , λ� and of the parameterslisted in γ. The optimal solution (x∗1(γ), . . . , λ

∗�(γ)) is a list of functions of only the

parameters in γ, so when we substitute the optimal solution into (12.17) and (12.18),we obtain the following n+ � equations that are functions of only the parameters :

∂f(x;α)

∂x1|x=x∗(γ) − λ∗1(γ)

∂g1(x; β1)

∂x1|x=x∗(γ) − · · · − λ∗�(γ)

∂g�(x; β�)

∂x1|x=x∗(γ) = 0,

......

(12.19)

∂f(x;α)

∂xn|x=x∗(γ) − λ∗1(γ)

∂g1(x; β1)

∂xn|x=x∗(γ) − · · · − λ∗�(γ)

∂g�(x; β�)

∂xn|x=x∗(γ) = 0,

and λ∗1(γ)(b1 − g1(x∗(γ); β1)) = 0, . . . , λ∗�(γ)(b� − g�(x

∗(γ); β�)) = 0.

Page 403: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

384 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

Necessarily, each of these n+ � equations is always zero, so let us write them as

I1(γ) ≡∂f(x;α)

∂x1|x=x∗(γ) − λ∗1(γ)

∂g1(x; β1)

∂x1|x=x∗(γ) − · · · − λ∗�(γ)

∂g�(x; β�)

∂x1|x=x∗(γ)≡0,

...... (12.20)

In(γ) ≡∂f(x;α)

∂xn|x=x∗(γ) − λ∗1(γ)

∂g1(x; β1)

∂xn|x=x∗(γ) − · · · − λ∗�(γ)

∂g�(x; β�)

∂xn|x=x∗(γ)≡0,

In+1(γ) ≡ λ∗1(γ)(b1 − g1(x∗(γ); β1)) ≡ 0, . . . , In+�(γ) ≡ λ∗�(γ)(b� − g�(x

∗(γ); β�))≡0.

The n+ � comparative statics quantities ∂x∗1(γ)/∂γk, . . . , ∂λ∗�(γ)/∂γk are obtained by

differentiating this system of zero-valued identities with respect to γk, and then solvethe resulting n+ � linear equations in these quantities.

Consider the identity I1(γ) ≡ 0. Its value is constant, so its rate of change withrespect to any parameter γk is always zero: ∂I1(γ)/∂γk = 0 for all possible valuesof γ. From (12.20), differentiating I1(γ) ≡ 0 with respect to γk gives

0 =∂I1(γ)

∂γk

=∂2f(x;α)

∂x21|x=x∗(γ)

∂x∗1(γ)∂γk

+ · · ·+ ∂2f(x;α)

∂x1∂xn|x=x∗(γ)

∂x∗n(γ)∂γk

+∂2f(x;α)

∂x1∂γk|x=x∗(γ)

− λ∗1(γ)(∂2g1(x; β1)

∂x21|x=x∗(γ)

∂x∗1(γ)∂γk

+ · · ·+ ∂2g1(x; β1)

∂x1∂xn|x=x∗(γ)

∂x∗n(γ)∂γk

)...

... (12.21)

− λ∗�(γ)(∂2g�(x; β�)

∂x21|x=x∗(γ)

∂x∗1(γ)∂γk

+ · · ·+ ∂2g�(x; β�)

∂x1∂xn|x=x∗(γ)

∂x∗n(γ)∂γk

)

− λ∗1(γ)∂2g1(x; β1)

∂x1∂γk|x=x∗(γ) − · · · − λ∗�(γ)

∂2g�(x; β�)

∂x1∂γk|x=x∗(γ)

− ∂λ∗1(γ)∂γk

∂g1(x; β1)

∂x1|x=x∗(γ) − · · · − ∂λ∗�(γ)

∂γk

∂g�(x; β�)

∂x1|x=x∗(γ).

Let’s tidy up (12.21). Collect the terms with ∂x∗1(γ)/∂γk, then collect the terms with

Page 404: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.6. USING THE FIRST-ORDER CONDITIONS 385

∂x∗2(γ)/∂γk, and so on down to collecting the terms with ∂λ∗�(γ)/∂γk. This gives

0 =

(∂2f(x;α)

∂x21− λ∗1(γ)

∂2g1(x; β1)

∂x21− · · · − λ∗�(γ)

∂2g�(x; β�)

∂x21

)|x=x∗(γ)

∂x∗1(γ)∂γk

......

+

(∂2f(x;α)

∂xn∂x1− λ∗1(γ)

∂2g1(x; β1)

∂xn∂x1− · · · − λ∗�(γ)

∂2g�(x; β�)

∂xn∂x1

)|x=x∗(γ)

∂x∗n(γ)∂γk

+∂2f(x;α)

∂x1∂γk|x=x∗(γ) − λ∗1(γ)

∂2g1(x; β1)

∂x1∂γk|x=x∗(γ) − · · · − λ∗�(γ)

∂2g�(x; β�)

∂x1∂γk|x=x∗(γ)

− ∂g1(x; β1)

∂x1|x=x∗(γ)

∂λ∗1(γ)∂γk

− · · · − ∂g�(x; β�)

∂x1|x=x∗(γ)

∂λ∗�(γ)∂γk

. (12.22)

This still looks like an ugly expression, but in fact it is very simple. Remember thatall of the derivatives of f and g are being evaluated at the point x = x∗(γ), so theyare just numbers, as are the values λ∗j(γ). So (12.22) is actually a linear equation

0 = a11∂x∗1(γ)∂γk

+ · · ·+ a1n∂x∗n(γ)∂γk

+ c1 + a1,n+1∂λ∗1(γ)∂γk

+ · · ·+ a1,n+�∂λ∗�(γ)∂γk

(12.23)

with numerical coefficients

a11 =

(∂2f(x;α)

∂x21− λ∗1(γ)

∂2g1(x; β1)

∂x21− · · · − λ∗�(γ)

∂2g�(x; β�)

∂x21

)|x=x∗(γ)

......

a1n =

(∂2f(x;α)

∂xn∂x1− λ∗1(γ)

∂2g1(x; β1)

∂xn∂x1− · · · − λ∗�(γ)

∂2g�(x; β�)

∂xn∂x1

)|x=x∗(γ)

c1 =

(∂2f(x;α)

∂x1∂γk− λ∗1(γ)

∂2g1(x; β1)

∂x1∂γk− · · · − λ∗�(γ)

∂2g�(x; β�)

∂x1∂γk

)|x=x∗(γ)

a1,n+1 = − ∂g1(x; β1)

∂x1|x=x∗(γ), . . . , a1,n+� = − ∂g�(x; β�)

∂x1|x=x∗(γ)

and unknowns ∂x∗1(γ)/∂γk, . . . , ∂λ∗�(γ)/∂γk.

We repeat this process for the identity I2(γ) ≡ 0, and obtain another linearequation

0 = a21∂x∗1(γ)∂γk

+ · · ·+ a2n∂x∗n(γ)∂γk

+ c2 + a2,n+1∂λ∗1(γ)∂γk

+ · · ·+ a2,n+�∂λ∗�(γ)∂γk

(12.24)

Page 405: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

386 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

in which the numerical coefficients are

a21 =

(∂2f(x;α)

∂x1∂x2− λ∗1(γ)

∂2g1(x; β1)

∂x1∂x2− · · · − λ∗�(γ)

∂2g�(x; β�)

∂x1∂x2

)|x=x∗(γ)

......

a2n =

(∂2f(x;α)

∂xn∂x2− λ∗1(γ)

∂2g1(x; β1)

∂xn∂x2− · · · − λ∗�(γ)

∂2g�(x; β�)

∂xn∂x2

)|x=x∗(γ)

c2 =

(∂2f(x;α)

∂x2∂γk− λ∗1(γ)

∂2g1(x; β1)

∂x2∂γk− · · · − λ∗�(γ)

∂2g�(x; β�)

∂x2∂γk

)|x=x∗(γ)

a2,n+1 = − ∂g1(x; β1)

∂x2|x=x∗(γ), . . . , a2,n+� = − ∂g�(x; β�)

∂x2|x=x∗(γ)

and the unknowns are again ∂x∗1(γ)/∂γk, . . . , ∂λ∗�(γ)/∂γk. We continue this process

for each of the identities I3(γ) ≡ 0 to In(γ) ≡ 0. Each differentiation of an identitygives us another linear equation similar to (12.23) and (12.24).

Now we differentiate In+1(γ) ≡ 0 with respect to γk, knowing that the value ofthis derivative must be zero, and so obtain that

0 =∂λ∗1(γ)∂γk

(b1 − g1(x∗(γ); β1)) + λ∗1(γ)

∂b1∂γk

− λ∗1(γ)(∂g1(x; β1)

∂x1|x=x∗(γ)

∂x∗1(γ)∂γk

+ · · ·+ ∂g1(x; β1)

∂xn|x=x∗(γ)

∂x∗n(γ)∂γk

)

− λ∗1(γ)∂g1(x; β1)

γk|x=x∗(γ).

(12.25)

The first term on the right-hand side of (12.25) is zero because the first constraintbinds at x = x∗(γ). Thus (12.25) reduces to

0 = λ∗1(γ)(∂g1(x; β1)

∂x1|x=x∗(γ)

∂x∗1(γ)∂γk

+ · · ·+ ∂g1(x; β1)

∂xn|x=x∗(γ)

∂x∗n(γ)∂γk

)

+ λ∗1(γ)∂g1(x; β1)

γk|x=x∗(γ) − λ∗1(γ)

∂b1∂γk

.

(12.26)

∂b1/∂γk = 0 unless γk is b1, in which case ∂b1/∂γk = 1. ∂g1(x; β1)/∂γk is zero unlessγk is one of the parameters listed in the vector β1. (12.26) is a linear equation

an+1,1∂x∗1(γ)∂γk

+ · · ·+ an+1,n∂x∗n(γ)∂γk

= cn+1 (12.27)

Page 406: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.6. USING THE FIRST-ORDER CONDITIONS 387

in which the numerical coefficients are

an+1,1 = λ∗1(γ)∂g1(x; β1)

∂x1|x=x∗(γ), . . . , an+1,n = λ∗1(γ)

∂g1(x; β1)

∂xn|x=x∗(γ)

and cn+1 = −λ∗1(γ)∂g1(x; β1)

∂γk|x=x∗(γ) + λ∗1(γ)

∂b1∂γk

.

Now we repeat this process for each in turn of the identities In+2(γ) ≡ 0 to In+�(γ) ≡0. For example, differentiating the identity In+�(γ) ≡ 0 yields the linear equation

an+�,1∂x∗1(γ)∂γk

+ · · ·+ an+�,n∂x∗n(γ)∂γk

= cn+� (12.28)

in which the numerical coefficients are

an+�,1 = λ∗�(γ)∂g�(x; β�)

∂x1|x=x∗(γ), . . . , an+�,n = λ∗�(γ)

∂g�(x; β�)

∂xn|x=x∗(γ)

and cn+� = −λ∗�(γ)∂g�(x; β�)

∂γk|x=x∗(γ) + λ∗�(γ)

∂b�∂γk

.

Together, these differentiations create the linear simultaneous equation system⎛⎜⎜⎜⎜⎜⎜⎜⎝

a11 · · · a1n a1,n+1 · · · a1,n+�...

. . ....

......

an1 · · · ann an,n+1 · · · an,n+�

an+1,1 · · · an+1,n an+1,n+1 · · · an+1,n+�...

......

. . ....

an+�,1 · · · an+�,n an+�,n+1 · · · an+�,n+�

⎞⎟⎟⎟⎟⎟⎟⎟⎠

⎛⎜⎜⎜⎜⎜⎜⎜⎝

∂x∗1(γ)/∂γk...

∂x∗n(γ)/∂γk∂λ∗1(γ)/∂γk

...∂λ∗�(γ)/∂γk

⎞⎟⎟⎟⎟⎟⎟⎟⎠

=

⎛⎜⎜⎜⎜⎜⎜⎜⎝

c1...cncn+1...

cn+�

⎞⎟⎟⎟⎟⎟⎟⎟⎠.

(12.29)Every one of the coefficients aij is a number. Every one of the coefficients ci is a num-ber (many of them will be zeros). So long as the rows of the matrix (aij) are all linearlyindependent of each other, you now can use whatever is your favorite method for solv-ing linear simultaneous equation systems to discover the value at the optimal solutionto problem (12.1) of the comparative statics quantities ∂x∗1(γ)/∂γk, . . . , ∂λ

∗�(γ)/∂γk.

That is all there is to this method of conducting a quantitative comparative staticsanalysis of the optimal solution to problem (12.1). The process may be tedious butit is not difficult.

We will proceed to an example in a moment, but, before doing so, there is a reallyimportant point for you to realize. What exactly do we need in order to compute thenumerical values of the coefficients in the

(aij)-matrix on the left-hand side of (12.29)?

Page 407: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

388 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

We are given the functions f, g1, . . . , g� and the values of all of the parameters in thevector γ. We are also given the values x∗1(γ), . . . , λ

∗�(γ) of the optimal solution for the

given parameter values. This is all we need to compute the numerical value of everyone of the aij coefficients. So what don’t we need? Particularly, we do not need toknow the formulae (if they even exist, since they may be only implicitly defined) forthe optimal solution functions x∗1(γ), . . . , λ

∗�(γ). This is a huge advantage. It lets us

compute the values of all of the rates of change ∂x∗i (γ)/∂γk and ∂λ∗j(γ)/∂γk withouthaving equations for x∗1(γ), . . . , λ

∗�(γ).

For an example we will use the problem faced by a price-taking and production-cost-minimizing firm that uses two variable inputs and must produce at least a spec-ified quantity y ≥ 0 of its product. The firm’s production function is g(x1, x2) =

x1/21 x

1/22 , so the firm’s problem is

minx1, x2

w1x1 + w2x2 subject to x1 ≥ 0, x2 ≥ 0, and x1/21 x

1/22 ≥ y, (12.30)

where w1 and w2 are the given per unit prices of the firm’s inputs. The parametervector for the problem is

γ =

⎛⎝w1

w2

y

⎞⎠ .

The problem has three constraints. At the optimal solution

(x∗1(w1, w2, y), x∗2(w1, w2, y), λ

∗1(w1, w2, y), λ

∗2(w1, w2, y), λ

∗3(w1, w2, y)),

the first and second constraints are slack. Thus λ∗1(w1, w2, y) = λ∗2(w1, w2, y) = 0,and we consider only the third constraint during our comparative statics analysis ofthe problem’s optimal solution. The comparative statics quantities that we wish toevaluate are

∂x∗1(w1, w2, y)

∂γk,∂x∗2(w1, w2, y)

∂γk, and

∂λ∗3(w1, w2, y)

∂γk,

where γk is any one of the parameters w1, w2, or y. Let’s say that we want to see howthe optimal solution changes as the price of input 2 changes. Then we select γk = w2

and proceed to evaluate the comparative statics quantities

∂x∗1(w1, w2, y)

∂w2

,∂x∗2(w1, w2, y)

∂w2

, and∂λ∗3(w1, w2, y)

∂w2

.

We start by rewriting problem (12.30) so that it is in the format of problem (12.1).That is, we rewrite problem (12.30) so that it is a maximization problem and has

Page 408: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.6. USING THE FIRST-ORDER CONDITIONS 389

constraints of the form g(x) ≤ b. To convert the problem from a minimizationproblem into a maximization problem, simply multiply the objective function bynegative one. To convert a ≥ constraint into a ≤ constraint, multiply it by negativeone. This gives us the problem

maxx1, x2

−w1x1 − w2x2 subject to − x1 ≤ 0, −x2 ≤ 0, and y − x1/21 x

1/22 ≤ 0. (12.31)

We write down the problem’s necessary first-order optimality conditions. Theseare

−w1 = − λ32x−1/21 x

1/22 , −w2 = − λ3

2x1/21 x

−1/22 , and 0 = λ3

(−y + x

1/21 x

1/22

).

Now we note that, for all parameter values w1 > 0, w2 > 0, and y > 0, theseequations are true when evaluated at the optimal solution x∗1(w1, w2, y), x

∗2(w1, w2, y),

and λ∗3(w1, w2, y). This allows us to write the identities

I1(w1, w2, y) ≡ −w1 +λ∗3(w1, w2, y)

2x∗1(w1, w2, y)

−1/2 x∗2(w1, w2, y)1/2 ≡ 0,

(12.32)

I2(w1, w2, y) ≡ −w2 +λ∗3(w1, w2, y)

2x∗1(w1, w2, y)

1/2 x∗2(w1, w2, y)−1/2 ≡ 0,

(12.33)

and I3(w1, w2, y) ≡ λ∗3(w1, w2, y)(−y + x∗1(w1, w2, y)

1/2x∗2(w1, w2, y)1/2)≡ 0.

(12.34)

Each identity is a function of only the parameters w1, w2, and y. Since each identity’svalue is constant, at zero, for any values w1 > 0, w2 > 0, and y > 0, we know that

∂I1(w1, w2, y)

∂w2

= 0,∂I2(w1, w2, y)

∂w2

= 0, and∂I3(w1, w2, y)

∂w2

= 0.

Page 409: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

390 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

So, from (12.32), (12.33), and (12.34), we obtain

∂I1(w1, w2, y)

∂w2

= 0 =1

2

∂λ∗3(·)∂w2

x∗1(·)−1/2x∗2(·)1/2 −λ∗3(·)4

x∗1(·)−3/2x∗2(·)1/2∂x∗1(·)∂w2

+λ∗3(·)4

x∗1(·)−1/2x∗2(·)−1/2 ∂x∗2(·)

∂w2

, (12.35)

∂I2(w1, w2, y)

∂w2

= 0 = −1 +1

2

∂λ∗3(·)∂w2

x∗1(·)1/2x∗2(·)−1/2 +λ∗3(·)4

x∗1(·)−1/2x∗2(·)−1/2 ∂x∗1(·)

∂w2

− λ∗3(·)4

x∗1(·)1/2x∗2(·)−3/2 ∂x∗2(·)

∂w2

, and (12.36)

∂I3(w1, w2, y)

∂w2

= 0 =∂λ∗3(·)∂w2

(−y + x∗1(·)1/2x∗2(·)1/2

)+λ∗3(·)2

x∗1(·)−1/2x∗2(·)1/2∂x∗1(·)∂w2

+λ∗3(·)2

x∗1(·)1/2x∗2(·)−1/2 ∂x∗2(·)

∂w2

. (12.37)

It will be easier to solve these equations for ∂x∗1(·)/∂w2, ∂x∗2(·)/∂w2, and ∂λ

∗3(·)/∂w2

if we first tidy up the equations. Multiply (12.35) by 4x∗1(·)3/2x∗2(·)1/2 and rearrangethe result to

−λ∗3(·)x∗2(·)∂x∗1(·)∂w2

+ λ∗3(·)x∗1(·)∂x∗2(·)∂w2

+ 2x∗1(·)x∗2(·)∂λ∗3(·)∂w2

= 0. (12.38)

Then multiply (12.36) by 4x∗1(·)1/2x∗2(·)3/2 and rearrange the result to

λ∗3(·)x∗2(·)∂x∗1(·)∂w2

− λ∗3(·)x∗1(·)∂x∗2(·)∂w2

+ 2x∗1(·)x∗2(·)∂λ∗3(·)∂w2

= 4x∗1(·)1/2x∗2(·)3/2. (12.39)

Last, note that the first term on the right-hand side of (12.37) is zero (because theconstraint binds at the optimal solution). Multiply the other two terms on the right-hand side of (12.37) by 2x∗1(·)1/2x∗2(·)1/2, divide by λ∗3(·) (you can do this becauseλ∗3(·) > 0), and rearrange the result to

x∗2(·)∂x∗1(·)∂w2

+ x∗1(·)∂x∗2(·)∂w2

= 0. (12.40)

(12.38), (12.39), and (12.40) are three linear equations in the three unknowns thatare the comparative statics quantities we wish to evaluate. So all that remains is to

Page 410: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.7. USING THE MAXIMUM-VALUE FUNCTION 391

solve these equations. You can verify for yourself that the solution is

∂x∗1(w1, w2, y)

∂w2

=y

λ∗3(w1, w2, y), (12.41)

∂x∗2(w1, w2, y)

∂w2

= − x∗2(w1, w2, y)3/2

x∗1(w1, w2, y)1/2λ∗3(w1, w2, y), (12.42)

and∂λ∗3(w1, w2, y)

∂w2

=

(x∗2(w1, w2, y)

x∗1(w1, w2, y)

)1/2

. (12.43)

To evaluate these comparative statics quantities, all we need to know are the valuesof x∗1(w1, w2, y), x

∗2(w1, w2, y), and λ

∗3(w1, w2, y) for the given values of w1, w2, and y.

Particularly, we do not need to have equations for any of x∗1(w1, w2, y), x∗2(w1, w2, y),

and λ∗3(w1, w2, y) to obtain the numerical values of all three comparative statics quan-tities.

What of economic interest do we learn from (12.41), (12.42), and (12.43)? It isthat a firm that wishes to keep its production costs of producing y units of its productas small as possible responds to a higher per unit price w2 for input 2 by increasing itsquantity demanded of input 1 at the rate that is the value of (12.41) and by reducingits quantity demanded of input 2 at the rate that is the value of (12.42). The higherinput price w2 also increases λ∗3(·), the firm’s marginal production cost, at the ratethat is the value of (12.43).

As an exercise, why don’t you repeat the analysis of the firm’s cost-minimizationproblem for the comparative statics experiment in which only the required productionlevel y is altered? Remember to provide an economic interpretation of these results.

12.7 Using the Maximum-Value Function

In the previous section we used the first-order optimality conditions together with theImplicit Function Theorem to obtain quantitative results about the rates of change ofthe optimal solution components x∗1(γ), . . . , λ

∗m(γ). In this section we use a different

information source to create a different type of comparative statics result. The newinformation source is the maximum-value function of the constrained optimizationproblem. The technique that we use to obtain these results employs a rather wonderfultool called the First-Order Envelope Theorem.

Our discussion will be made a little easier if we slightly alter the notation that wehave used so far for the constraints. From here onwards I would like to write

hj(x; βj, bj) ≡ bj − gj(x; βj) ≥ 0 for j = 1, . . . ,m,

Page 411: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

392 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

and, consequently, write our constrained optimization problem as

maxx∈�n

f(x;α) subject to hj(x; βj, bj) ≥ 0 for j = 1, . . . ,m. (12.44)

This notational change does not alter our problem at all and makes it easier to writedown some of the expressions that you will see shortly.

Let’s get started by recalling what is meant by the maximum-value function ofour constrained optimization problem.

Definition 12.1 (Maximum-Value Function). The maximum-value function for theconstrained optimization problem (12.44) is

v(γ) ≡ maxx∈�n

f(x;α) subject to hj(x; βj, bj) ≥ 0 for j = 1, . . . ,m (12.45)

≡ f(x∗(γ);α) (12.46)

≡ f(x∗(γ);α) +m∑j=1

λ∗j(γ)hj(x∗(γ); βj, bj). (12.47)

Notice that each of (12.45), (12.46), and (12.47) is identical to the other two ex-pressions. (12.45) and (12.46) are identical by the definition of an optimal solution.(12.46) is identical to (12.47) because, at an optimal solution to the problem, eachcomplementary slackness term λ∗j(γ)hj(x

∗(γ); βj, bj) necessarily is zero. (12.47) re-veals to us that the problem’s maximum-value function is exactly the same as theproblem’s Lagrange function evaluated at the problem’s optimal solution.

Two examples of maximum-value functions that we have already encountered arethe indirect utility function (11.6) and the indirect profit function (11.7).

The properties possessed by a maximum-value function depend upon the prop-erties of the constrained optimization problem to which it belongs. This chapterconsiders problems for which the maximum-value function is differentiable on theparameter space Γ.

The First-Order Envelope Theorem is often called “the Envelope Theorem,” butthere is more than one envelope theorem, so for clarity let’s call the theorem by itsfull name.

Theorem 12.3 (First-Order Envelope Theorem). Consider a problem (12.44) forwhich the optimal solution correspondence (x∗, λ∗) : Γ �→ �n × �m

+ is a functionthat is differentiable on Γ and for which the maximum-value function (12.45) is also

Page 412: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.7. USING THE MAXIMUM-VALUE FUNCTION 393

differentiable on Γ. Let I(x∗(γ)) be the set of indices of the constraints that bind atx = x∗(γ). Then

∂v(γ)

∂γk=∂f(x;α)

∂γk|x=x∗(γ) +

∑j∈I(x∗(γ))

λ∗j(γ)∂hj(x; βj, bj)

∂γk|x=x∗(γ). (12.48)

Let’s be sure to read this result correctly. It says that the rate of change of themaximum value with respect to a parameter γk can be computed in the followingway. We are given a constrained optimization problem that is defined in part by agiven parameter vector γ. First, we write down the problem’s Lagrange function.Second, in the Lagrange function, we fix the value of each xi at x

∗i (γ) and also fix

the value of each λj at λ∗j(γ). Each x∗i and each λ∗j is just a number, so this secondstep creates a restricted Lagrange function of only the parameter vector γ. Especiallyimportant is that this function depends upon the parameters γk only directly, meaningthat the dependency of the restricted Lagrange function upon γk is only through γk’seffect directly upon whichever of the functions f, g1, . . . , gm directly depend upon γk.Particularly, the restricted Lagrange function does not depend upon γk through thex∗i ’s and λ

∗j ’s (these are called the indirect effects on v through x∗ and λ∗) because we

have fixed their values. Last, we differentiate the restricted Lagrange function withrespect to γk.

It is easy to misunderstand what I have said in the preceding paragraph, so let’sagain use the firm’s profit-maximization problem (11.4) as an example. Supposeβ1 = β2 = 1

3, p = 3, and w1 = w2 = 1. From (11.5) the profit-maximizing solution

is (y∗, x∗1, x∗2, λ

∗1, λ

∗2, λ

∗3, λ

∗4) = (1, 1, 1, 0, 0, 0, 3). The first step is to write down the

problem’s Lagrange function, which is

L(y, x1, x2, λ1, λ2, λ3, λ4; p, w1, w2, β1, β2)

= py − w1x1 − w2x2 − λ1y − λ2x1 − λ3x2 + λ4(xβ1

1 xβ2

2 − y).

The second step is to substitute the numerical values of the solution that is optimal forthe parameter vector (p, w1, w2, β1, β2) =

(3, 1, 1, 1

3, 13

). Note: We do not substitute

the numerical values of the parameters. This gives us the restricted Lagrange functionof p, w1, w2, β1, and β2 that is

Lr(p, w1, w2, β1, β2)= p× 1− w1 × 1− w2 × 1− 0×1− 0×1− 0×1 + 3(1β1×1β2− 1

)= p− w1 − w2. (12.49)

The important point is that this function’s value at the particular parameter vectorvalue (p, w1, w1) = (3, 1, 1) is the firm’s maximum achievable profit for these prices.

Page 413: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

394 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

This maximum profit value is 3− 1− 1 = 1. Really? Let’s check by using (11.7). For(p, w1, w2, β1, β2) =

(3, 1, 1, 1

3, 13

), the value of the indirect profit function is

v(p, w1, w2, β1, β2) = (1− β1 − β2)

(pββ1

1 ββ2

2

wβ1

1 wβ2

2

)1/(1−β1−β2)

=1

3

(3× 1

3

1/3 × 13

1/3

11/3 × 11/3

)3

= 1,

(12.50)so (12.49) is indeed reporting the maximum profit that the firm can achieve for theparameter values (p, w1, w2, β1, β2) =

(3, 1, 1, 1

3, 13

). When applied to our example,

the First-Order Envelope Theorem claims that the rates of change of the firm’s max-imum profit level with respect to p, w1, and w2 are the values at (p, w1, w2, β1, β2) =(3, 1, 1, 1

3, 13

)of the partial derivatives of (12.49) with respect to each of p, w1, and w2.

These are

∂Lr(p, w1, w2, β1, β2)

∂p|(p,w1,w2,β1,β2)

=(3,1,1, 13 ,13)

= 1,∂Lr(p, w1, w2, β1, β2)

∂w1

|(p,w1,w2,β1,β2)

=(3,1,1, 13 ,13)

= −1,

and∂Lr(p, w1, w2, β1, β2)

∂w2

|(p,w1,w2,β1,β2)

=(3,1,1, 13 ,13)

= −1. (12.51)

We can check this by computing the partial derivatives of (12.50) with respect to p,w1, and w2, and evaluating them at (p, w1, w2, β1, β2) =

(3, 1, 1, 1

3, 13

). These values

are

∂v(p1, w1, w2, β1, β2)

∂p|(p,w1,w2,β1,β2)

=(3,1,1, 13 ,13)

=

(pβ1+β2ββ1

1 ββ2

2

wβ1

1 wβ2

2

)1/(1−β1−β2)

|(p,w1,w2,β1,β2)

=(3,1,1, 13 ,13)

= 1 =∂Lr(p, w1, w2, β1, β2)

∂p|(p,w1,w2,β1,β2)

=(3,1,1, 13 ,13),

∂v(p1, w1, w2, β1, β2)

∂w1

|(p,w1,w2,β1,β2)

=(3,1,1, 13 ,13)

= −β1(pββ1

1 ββ2

2

w1−β2

1 wβ2

2

)1/(1−β1−β2)

|(p,w1,w2,β1,β2)

=(3,1,1, 13 ,13)

= −1 =∂Lr(p, w1, w2, β1, β2)

∂w1

|(p,w1,w2,β1,β2)

=(3,1,1, 13 ,13),

and∂v(p1, w1, w2, β1, β2)

∂w2

|(p,w1,w2,β1,β2)

=(3,1,1, 13 ,13)

= −β2(pββ1

1 ββ2

2

wβ1

1 w1−β1

2

)1/(1−β1−β2)

|(p,w1,w2,β1,β2)

=(3,1,1, 13 ,13)

= −1 =∂Lr(p, w1, w2, β1, β2)

∂w2

|(p,w1,w2,β1,β2)

=(3,1,1, 13 ,13).

Page 414: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.7. USING THE MAXIMUM-VALUE FUNCTION 395

So, for our example at least, the First-Order Envelope Theorem seems to work. Noticethat it is much easier to differentiate (12.49) than it is to differentiate (12.50). Oneof the great advantages of the theorem is that it typically saves us from grindingthrough a lot of tedious algebra.

This algebraic simplification is due to our not having to compute the (indirect)effects caused by a change to a parameter changing the problem’s optimal solutionx∗1(γ), . . . , λ

∗m(γ) and thereby changing the value of the problem’s maximum value.

Why is this so? Why don’t we have to consider the changes caused to the maximumvalue v(γ) by changes to the optimal solution (x∗(γ), λ∗(γ)) when computing the rateof change of the maximum value? Because, as you will see in the proof of the theorem,all of the indirect effects cancel each other out – they sum to zero. This results ina huge simplification. Before we work through the proof, it will be helpful to see anexample of exactly how the theorem allows us to ignore the indirect effects, so, forone more time, let’s go back to our example of the profit-maximizing firm.

The Lagrange function for (11.4) is

L(y, x1, x2, λ1, λ2, λ3, λ4; p, w1, w2, β1, β2)

= py − w1x1 − w2x2 + λ1y + λ2x1 + λ3x2 + λ4(xβ1

1 xβ2

2 − y).

For brevity, we will use only y∗, x∗1, x∗2, λ

∗1, λ

∗2, λ

∗3, and λ∗4 to denote the optimal

values, rather than go to the bother of substituting in complicated equations suchas those on the right-hand sides of (11.5). Be clear that, for a given parametervector (p, w1, w2, β1, β2), the quantities y∗, . . . , λ∗4 are fixed values. The maximum-value function is

π∗(p, w1, w2, β1, β2)

≡ py∗ − w1x∗1 − w2x

∗2 + λ∗1y

∗ + λ∗2x∗1 + λ∗3x

∗2 + λ∗4

((x∗1)

β1(x∗2)β2 − y∗

). (12.52)

According to the First-Order Envelope Theorem, to compute the rate of change ofthe firm’s maximum profit with respect to, say, the parameter p, all we need todo is to compute the derivative of the right-hand side of (12.52) with respect to p(remembering that y∗, . . . , λ∗4 are numbers, and thus are constants). This easily givesus that

∂π∗(p, ·)∂p

= y∗(p, w1, w2, β1, β2). (12.53)

If, instead, you insist upon doing things the hard way by applying the chain rule of

Page 415: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

396 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

differentiation to (12.52), then you obtain

∂π∗(p, ·)∂p

= y∗(p, ·) + p∂y∗(p, ·)∂p

− w1∂x∗1(p, ·)∂p

− w2∂x∗2(p, ·)∂p

+∂λ∗1(p, ·)∂p

y∗(p, ·) + λ∗1(p, ·)∂y∗(p, ·)∂p

+∂λ∗2(p, ·)∂p

x∗1(p, ·)

+ λ∗2(p, ·)∂x∗1(p, ·)∂p

+∂λ∗3(p, ·)∂p

x∗2(p, ·) + λ∗3(p, ·)∂x∗2(p, ·)∂p

(12.54)

+∂λ∗4(p, ·)∂p

(x∗1(p, ·)β1x∗2(p, ·)β2 − y∗(p, ·)

)+ λ∗4(p, ·)

(β1x

∗1(p, ·)β1−1x∗2(p, ·)β2

∂x∗1(p, ·)∂p

+ β2x∗1(p, ·)β1x∗2(p, ·)β2−1 ∂x

∗2(p, ·)∂p

− ∂y∗(p, ·)∂p

).

What a mess! But (12.53) and (12.54) are the same thing, so all of the indirect effectterms that contain quantities such as ∂x∗/∂p, ∂y∗/∂p and ∂λ∗/∂p) in (12.54) mustsum to zero. How can this be? Let’s group the terms in (12.54) together as follows:

∂π∗(p, ·)∂p

= y∗(p, ·) + (p+ λ∗1(p, ·)− λ∗4(p, ·))∂y∗(p, ·)∂p

−(w1 − λ∗2(p, ·)− λ∗4(p, ·)β1x∗1(p, ·)β1−1x∗2(p, ·)β2

) ∂x∗1(p, ·)∂p

−(w2 − λ∗3(p, ·)− λ∗4(p, ·)β2x∗1(p, ·)β1x∗2(p, ·)β2−1

) ∂x∗2(p, ·)∂p

+∂λ∗1(p, ·)∂p

y∗(p, ·) + ∂λ∗2(p, ·)∂p

x∗1(p, ·) +∂λ∗3(p, ·)∂p

x∗2(p, ·) (12.55)

+∂λ∗4(p, ·)∂p

(x∗1(p, ·)β1x∗2(p, ·)β2 − y∗(p, ·)

).

The maximum value for the problem for a given parameter vector γ is the objectivefunction’s value at the problem’s optimal solution (x∗(γ), λ∗(γ)), where necessarilythe Karush-Kuhn-Tucker and the complementary slackness conditions are satisfied;

Page 416: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.7. USING THE MAXIMUM-VALUE FUNCTION 397

i.e. ⎛⎝ p−w1

−w2

⎞⎠ = λ∗1(p, ·)

⎛⎝−1

00

⎞⎠+ λ∗2(p, ·)

⎛⎝ 0−10

⎞⎠+ λ∗3(p, ·)

⎛⎝ 0

0−1

⎞⎠

+ λ∗4(p, ·)

⎛⎝ 1−β1x∗1(p, ·)β1−1x∗2(p, ·)β2

−β2x∗1(p, ·)β1x∗2(p, ·)β2−1

⎞⎠ , (12.56)

0 = λ∗1(p, ·)y∗(p, ·), (12.57)

0 = λ∗2(p, ·)x∗1(p, ·), (12.58)

0 = λ∗3(p, ·)x∗2(p, ·), (12.59)

and 0 = λ∗4(p, ·)(y∗(p, ·)− (x∗1(p, ·))β1(x∗2(p, ·))β2

). (12.60)

(12.56) says that the second, third, and fourth terms on the right-hand side of (12.55)are necessarily zero. (12.57) says that necessarily either λ∗1(p, ·) = 0, in which case∂λ∗1(γ)/∂p = 0, or else y∗(p, ·) = 0. Either way, the fifth term on the right-hand sideof (12.55) is necessarily zero. Similarly, (12.58), (12.59), and (12.60), respectively,inform us that necessarily the sixth, seventh, and last terms on the right-hand side of(12.55) are each zero. Thus all the indirect effect terms can be ignored, and we havethe conclusion provided earlier by the First-Order Envelope Theorem that

∂π∗(p, ·)∂p

= y∗(p, ·).

The proof of the First-Order Envelope Theorem is merely a restatement of the argu-ment we have just completed, albeit in a more general setting. Here it is.

Proof of the First-Order Envelope Theorem. Given γ, the primal problem’s op-timizing solution is x∗(γ) = (x∗1(γ), . . . , x

∗n(γ)) and the associated vector of Lagrange

multipliers is λ∗(γ) = (λ∗1(γ), . . . , λ∗m(γ)). From (12.46),

∂v(γ)

∂γk=∂f(x;α)

∂γk|x=x∗(γ) +

n∑i=1

∂f(x;α)

∂xi|x=x∗(γ) ×

∂x∗i (γ)∂γk

. (12.61)

Since x = x∗(γ) is an optimal solution, the Karush-Kuhn-Tucker condition must besatisfied at x = x∗(γ); i.e.

∂f(x;α)

∂xi|x=x∗(γ) = −

m∑j=1

λ∗j(γ)∂hj(x; βj, bj)

∂xi|x=x∗(γ) ∀ i = 1, . . . , n. (12.62)

Page 417: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

398 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

Substitution of (12.62) into (12.61) gives

∂v(γ)

∂γk=∂f(x;α)

∂γk|x=x∗(γ) −

n∑i=1

(m∑j=1

λ∗j(γ)∂hj(x; βj, bj)

∂xi|x=x∗(γ)

)∂x∗i (γ)∂γk

=∂f(x;α)

∂γk|x=x∗(γ) −

m∑j=1

λ∗j(γ)

(n∑

i=1

∂hj(x; βj, bj)

∂xi|x=x∗(γ)

∂x∗i (γ)∂γk

). (12.63)

Also necessarily satisfied at x∗(γ) are the complementary slackness conditions

λ∗j(γ)hj(x∗(γ); βj, bj) ≡ 0 for each j = 1, . . . ,m.

Consequently, for each j = 1, . . . ,m,

0 =∂λ∗j(γ)

∂γkhj(x

∗(γ); βj, bj) + λ∗j(γ)∂hj(x; βj, bj)

∂γk|x=x∗(γ)

+ λ∗j(γ)n∑

i=1

∂hj(x; βj, bj)

∂xi|x=x∗(γ)

∂x∗i (γ)∂γk

.

(12.64)

If the jth constraint is slack, then λ∗j(γ) = 0 and ∂λ∗(γ)/∂γk = 0. If the jth constraintbinds, then hj(x

∗(γ); βj, bj) = 0. Therefore, the first term of the right-hand side of(12.64) is zero for every j = 1, . . . ,m. Consequently, for each j = 1, . . . ,m,

λ∗j(γ)∂hj(x; βj, bj)

∂γk|x=x∗(γ) = −λ∗j(γ)

n∑i=1

∂hj(x; βj, bj)

∂xi|x=x∗(γ)

x∗i (γ)∂γk

. (12.65)

Substitution from (12.65) into (12.63) gives

∂v(γ)

∂γk=∂f(x;α)

∂γk|x=x∗(γ) +

m∑j=1

λ∗j(γ)∂hj(x; βj, bj)

∂γk|x=x∗(γ). (12.66)

I(x∗(γ)) is the set of indices of the constraints that bind at x = x∗(γ), so λ∗j(γ) = 0for each j /∈ I(x∗(γ)) and (12.66) is

∂v(γ)

∂γk=∂f(x;α)

∂γk|x=x∗(γ)+

∑j∈I(x∗(γ))

λ∗j(γ)∂hj(x; βj, bj)

∂γk|x=x∗(γ).

Now let’s get a clear understanding of the geometry of the First-Order EnvelopeTheorem. Figure 12.2 displays three curves labeled vr(γ; x) for three fixed values x′,

Page 418: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.7. USING THE MAXIMUM-VALUE FUNCTION 399

Figure 12.2: The restricted functions vr(γ; x′), vr(γ; x′′), and vr(γ; x′′′).

x′′, and x′′′ of x. The superscript “r” denotes “restricted.” For example, vr(γ; x′)is the function v(γ, x) restricted by fixing x at the value x′. There is one restrictedfunction vr(γ; x) for each of the three allowed values of x.

Take any value of γ < γ in Figure 12.2. Observe that the values of the threerestricted functions are then vr(γ; x′) > vr(γ; x′′) > vr(γ; x′′′). If your goal is toattain the highest of these values, then which value for x would you choose? Well,that would be x = x′. We have that the optimal solution to the problem

maxx

vr(x; γ) subject to x ∈ {x′, x′′, x′′′}

is x∗(γ) = x′ for γ < γ.

Now take any value of γ satisfying γ < γ < γ in Figure 12.2. The largest of thevalues of the three restricted functions is now vr(γ; x′′), so the optimal solution to theproblem

maxx

vr(x; γ) subject to x ∈ {x′, x′′, x′′′}

is x∗(γ) = x′′ for γ < γ < γ.

Page 419: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

400 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

Last, for any value of γ > γ, the largest of the values of the three restrictedfunctions is vr(γ; x′′′) and the optimal solution to the problem

maxx

vr(x; γ) subject to x ∈ {x′, x′′, x′′′}

is x∗(γ) = x′′′ for γ > γ.Put the above three paragraphs together and we have discovered that the optimal

solution for the problem maxx vr(x; γ) subject to x ∈ {x′, x′′, x′′′} is

x∗(γ) =

⎧⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎩

x′ , γ < γ

{x′, x′′} , γ = γ

x′′ , γ < γ < γ

{x′′, x′′′} , γ = γ

x′′′ , γ < γ.

The maximum-value function for the problem is

v(γ) =

⎧⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎩

vr(γ; x′) , for γ < γ

vr(γ; x′) = vr(γ; x′′) , for γ = γ

vr(γ; x′′) , for γ < γ < γ

vr(γ; x′′) = vr(γ; x′′′) , for γ = γ

vr(γ; x′′′) , for γ < γ.

In Figure 12.3 the graph of this maximum-value function is the thick black linethat runs along the tops of the graphs of the three restricted value functions vr(γ; x′),vr(γ; x′′), and vr(γ; x′′′). This “upper-most” graph encloses from above, or “envelops”from above, the graphs of all of the restricted value functions vr(γ; x), which is whythe graph of v(γ) is called the “upper envelope” of the graphs of the three vr(γ; x)functions.

In many problems the allowed values for x lie in some interval I, so let’s see howour above analysis extends to such a case. Look again at Figures 12.2 and 12.3, butnow think of the three restricted value functions that are displayed as being just threeof infinitely many such functions, one for each of the infinitely many values of x ∈ Ithat are now allowed. Each one of these infinitely many restricted functions has itsown graph, each similar to the three graphs that are displayed. Our problem remainsthe same. That is, for a given value of γ, we seek the value for x ∈ I that makes thevalue function vr(x; γ) as large as is possible. This is the problem

maxx

vr(x; γ) subject to x ∈ I. (12.67)

Page 420: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.7. USING THE MAXIMUM-VALUE FUNCTION 401

Figure 12.3: The upper envelope v(γ) generated by the restricted functions vr(γ; x′),vr(γ; x′′), and vr(γ; x′′′).

In this problem γ’s value is fixed. The solution to this problem is the value of x thatis optimal for γ; i.e. x∗(γ). The maximum value for problem (12.67) for the givenvalue of γ is therefore

v(γ) = vr(x∗(γ); γ). (12.68)

If we now vary γ over the entire parameter space Γ (this is the comparative staticsexperiment), but for each possible value of γ we still treat γ as a parameter in prob-lem (12.67), then we obtain both the entire optimal solution function x∗(γ) definedon Γ and the entire maximum-value function for problem (12.67), which is

v(γ) ≡ vr(x∗(γ); γ) ∀ γ ∈ Γ. (12.69)

The graph of this function is displayed in Figure 12.4 as the darker, uppermost ofthe four curves. This graph of the unrestricted maximum value v(γ) is the upperenvelope of all of the graphs of the family of the vr(γ; x) functions.

Now look again at Figure 12.4. Consider the particular value γ′ for γ. Here is thecrucial idea. As you can see from point A in the figure, for γ = γ′ the value v(γ′)

Page 421: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

402 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

Figure 12.4: The upper envelope v(γ) generated by the family of functions vr(γ; x).

of the (unrestricted) maximum-value function is the same as the value vr(γ′; x′) ofthe restricted function vr(γ; x′) at γ = γ′. Thus x∗(γ′) = x′. This tells us that therestriction x = x′ is just-binding. Put another way, when γ = γ′, the unrestrictedoptimizing choice for x is x = x′, so the restriction that x must be the number x′ iseffectively not a restriction when γ = γ′.

Not only that, but, at γ = γ′, the slope with respect to γ of the unrestrictedmaximum-value function v(γ) is the same as the slope with respect to γ of the restrictedvalue function vr(γ; x′) (see point A). That is,

∂v(γ)

∂γ|γ=γ′ =

∂vr(γ; x′)∂γ

|γ=γ′x′=x∗(γ′)

. (12.70)

Statement (12.70) is the First-Order Envelope Theorem. It says something remark-ably simple and useful: if you want to know the rate at which a maximum value v(γ)changes with respect to a parameter γ at a particular value γ′ of γ, then all you needto do is the simpler task of computing the rate at which the restricted value vr(γ; x)changes with respect to γ at γ = γ′, when x is fixed at the value x∗(γ′). There are no“indirect effects” (i.e. terms that measure how x∗(γ) changes as γ changes) because

Page 422: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.7. USING THE MAXIMUM-VALUE FUNCTION 403

x is fixed at the value x∗(γ′).Similarly, at the particular value γ′′ for γ, the value of x that makes the function

v(γ′′; x) largest is x = x′′, which is, therefore, the optimizing value of x when γ = γ′′:x′′ = x∗(γ′′). Consequently when γ = γ′′ restricting x to the value x′′ is a just-binding restriction. This causes (see point B in the figure) the value v(γ′′) of the(unrestricted) maximum-value function to be the same as the value vr(γ′′; x′′) of therestricted function vr(γ; x′′) at γ = γ′′. It also causes at γ = γ′′, the rate of changewith respect to γ of the unrestricted maximum value v(γ) to be the same as the rateof change with respect to γ of the restricted value vr(γ; x′′) (see point B); i.e.

∂v(γ)

∂γ|γ=γ′′ =

∂vr(γ; x′′)∂γ

|γ=γ′′x′′=x∗(γ′′)

.

Many applications of the First-Order Envelope Theorem are familiar to economist-s. Perhaps the best-known examples are Roy’s Identity, Shephard’s Lemma, andHotelling’s Lemma. It is very helpful to visualize how the First-Order Envelope The-orem provides these results, so let’s examine how Hotelling’s Lemma is derived fromthe theorem.

Theorem 12.4 (Hotelling’s Lemma). Consider a profit-maximizing, price-taking firmwith a direct profit function

π(y, x1, . . . , xn; p, w1, . . . , wn) = py − w1x1 − · · · − wnxn,

where p > 0 and wi > 0 for i = 1, . . . , n, and with a technology described by a twice-continuously differentiable and strictly concave production function y = g(x1, . . . , xn).Denote the firm’s profit-maximizing plan by (y∗(p, w), x∗1(p, w), . . . , x

∗n(p, w)), where

w = (w1, . . . , wn). Denote the firm’s indirect profit function by

π∗(p, w) ≡ py∗(p, w)− w1x∗1(p, w)− · · · − wnx

∗n(p, w).

Then

∂π∗(p, w)∂p

= y∗(p, w) (12.71)

and∂π∗(p, w)∂wi

= −x∗i (p, w) for i = 1, . . . , n. (12.72)

To understand this result and how it is derived from the First-Order EnvelopeTheorem, start by thinking of the following restricted version of the firm’s problem

Page 423: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

404 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

(you might recognize this restricted problem as a particular “short-run” problem thatis often described in microeconomics textbooks):

maxy, x2, ... , xn

πr(y, x2, . . . , xn; x1, p, w) = py − w1x1 − w2x2 − · · · − wnxn

subject to y ≥ 0, x2 ≥ 0, . . . , xn ≥ 0, and g(x1, x2, . . . , xn)− y ≥ 0, (12.73)

where the quantity used of input 1, x1, is fixed at x1 > 0. The solution is therestricted (short-run) production plan (yr(p, w; x1), x1, x

r2(p, w; x1), . . . , x

rn(p, w; x1)).

Denote the firm’s restricted indirect profit function by

πr(p, w; x1) ≡ pyr(p, w; x1)− w1x1 − w2xr2(p, w; x1)− · · · − wnx

rn(p, w; x1). (12.74)

There is one such function for each possible value x1, so we may speak of the family(i.e. the complete collection) of all of these restricted, or short-run, maximum-profitfunctions. Now realize that the firm’s unrestricted indirect profit function

π∗(p, w) ≡ maxx1≥0

πr(p, w; x1). (12.75)

This is a statement that the unrestricted indirect profit function π∗(p, w) is theupper envelope of the family of restricted indirect profit functions πr(p, w; x1). Theidea is illustrated in Figure 12.5 for just the (w1, π)-plane. In this plane the graphof each restricted profit function πr(p, w; x1) is a straight line with a slope −x1 withrespect to w1 (see (12.74)). Three such restricted profit functions’ graphs are dis-played, one for x1 fixed at the value x1

′, another for x1 fixed at the value x1′′, and the

last for x1 fixed at the value x1′′′. Also displayed is the graph of the unrestricted prof-

it function π∗(p, w1, w2). Consider the point (w′1, π

∗(p, w′1, ·)) = (w′

1, πr(p, w′

1, ·; x1′))(labeled A in the figure). At this point the restriction x1 = x1

′ only just-binds since,when w1 = w′

1, the firm chooses x1 = x1′ even if it is not restricted to do so; i.e.

x∗1(p, w′1, ·) = x1

′. So, if w1 = w′1, then the firm’s unrestricted maximum profit level

and the firm’s maximum profit level when it is restricted to x1′ units of input 1 are

the same. As well, the rates of change of these profit levels with respect to w1 arethe same; this is statement (12.72). As an exercise to test your understanding of thisreasoning, you should draw a picture similar to Figure 12.5 and examine why (12.71)is true. Start by considering the restricted version of the firm’s problem that is

maxx1, ... , xn

πr(x2, . . . , xn; y, p, w) = py − w1x1 − w2x2 − · · · − wnxn

subject to x1 ≥ 0, . . . , xn ≥ 0, and y ≤ g(x1, x2, . . . , xn), (12.76)

Page 424: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.8. QUALITATIVE VS. QUANTITATIVE ANALYSES 405

C

Figure 12.5: Hotelling’s result that ∂π∗(p, w)/∂w1 = −x∗1(p, w).

where the firm’s quantity supplied, y, is fixed at some value y > 0.

Here we have used the First-Order Envelope Theorem to obtain some select quanti-tative comparative statics results. This technique complements the Implicit Functionmethod that we used in the previous section. Which method should you use? Thatdepends upon the type of quantitative comparative statics result you wish to obtain.The First-Order Envelope Theorem provides results of the type ∂v(γ)/∂γk. The Im-plicit Function Theorem provides results of the types ∂x∗i (γ)/∂γk and ∂λ∗j(γ)/∂γk.

12.8 Qualitative vs. Quantitative Analyses

This chapter has explained two techniques commonly used for quantitative compar-ative statics analyses. The Implicit Function Theorem approach is particularly de-manding in that to use it you must have a lot of information about the objective and

Page 425: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

406 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

constraint functions of the constrained optimization problem that is being analyzed.But what if you don’t? Well, you could assume the missing information and providequantitative comparative statics results that are conditional upon your assumptionsbeing true. This is quite often done. If you do not like this idea, then you will haveto settle for qualitative comparative statics results. Qualitative analyses are less in-formative but they require much less information. Which approach is appropriate foryou will depend upon what questions you want answered and upon what you knowabout the properties of the objective and constraint functions of the problem youwish to analyze.

Chapter 13 explains an elegant method of deriving qualitative comparative-staticsresults from the information provided by the second-order necessary optimizationcondition.

Page 426: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.9. PROBLEMS 407

12.9 Problems

This problem set is presented in two parts. The first part presents problems that usethe first-order optimality conditions and the Implicit Function Theorem. The secondpart presents problems that use the maximum-value function and the First-OrderEnvelope Theorem.

Part 1: Problems that use the first-order optimality conditions and theImplicit Function Theorem.

Problem 12.1. At the start of this chapter, we considered the problem

maxx

f(x;α1, α2, α3) = α1x2 + α2x+ α3,

where the parameter α1 < 0. We solved the problem by computing the first-orderderivative of f with respect to x and then noting that, at the optimal solution x∗,this derivative necessarily has the value zero. That is, we solved

df(x;α1, α2, α3)

dx= 2α1x+ α2 = 0 (12.77)

and obtained that the optimal solution is the algebraically explicit function

x∗(α1, α2) = − α2

2α1

. (12.78)

Having this function makes it easy to deduce the comparative statics quantities

∂x∗(α1, α2)

∂α1

=α2

2α21

,∂x∗(α1, α2)

∂α2

= − 1

2α1

, and∂x∗(α1, α2)

∂α3

= 0. (12.79)

Suppose that you are not able to derive (12.78). Use the Implicit Function Theoremto derive the comparative statics quantities listed in (12.79).

Problem 12.2. Consider the consumer demand problem

maxx1, x2

U(x1, x2) = x1x2 subject to x1 ≥ 0, x2 ≥ 0, and p1x1 + p2x2 ≤ y.

You probably already know that the optimal solution is

x∗1(p1, p2, y) =y

2p1, x∗2(p1, p2, y) =

y

2p2, and λ∗(p1, p2, y) =

2p1p2y

,

Page 427: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

408 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

but suppose that this algebraically explicit solution is not known to you. Use theImplicit Function to obtain the comparative statics quantities

∂x∗1∂p1

,∂x∗2∂p1

, and∂λ∗

∂p1.

Problem 12.3. It is not difficult to discover the algebraically explicit optimal solution(x∗1(p1, p2, y), x

∗2(p1, p2, y), λ

∗(p1, p2, y)) to the consumer demand problem

maxx1, x2

U(x1, x2) =√x1 +

√x2

subject to x1 ≥ 0, x2 ≥ 0, and g(x1, x2) = p1x1 + p2x2 ≤ y,(12.80)

but suppose that you are unable to compute the equations of this optimal solution.You want to know the rates of change of the values of x∗1, x

∗2, and the budget con-

straint’s multiplier λ∗ as p1 changes; i.e. you wish to evaluate ∂x∗1/∂p1, ∂x∗2/∂p1, and

∂λ∗/∂p1. Use the Implicit Function Theorem to derive these quantities.

Part 2: Problems that use the maximum-value function and the First-Order Envelope Theorem.

Consider a family of functions ψ : X × Γ �→ �, where X ⊆ Rn is the space ofchoice variables and Γ ⊆ �s is the parameter space.

Definition: If for every γ ∈ Γ there exists an x∗(γ) ∈ X such that

Θ∗(γ) ≡ ψ(x∗(γ), γ) ≥ ψ(x, γ) ∀ x ∈ X,

then the function Θ∗ : Γ → � is the upper envelope of the family of functions ψ.

Definition: If for every γ ∈ Γ there exists an x∗(γ) ∈ X such that

Θ∗(γ) ≡ ψ(x∗(γ), γ) ≤ ψ(x, γ) ∀ x ∈ X,

then the function Θ∗ : Γ → � is the lower envelope of the family of functions ψ.

Problem 12.4. Consider the family of functions f : X × Γ → �, where X = �,Γ = [0, 8], and f(x, γ) = −x2 + γx+ 2.(i) What is the optimal solution x∗(γ) to the problem maxx f

r(x; γ) = −x2 + γx+ 2,in which γ ∈ [0, 8] is a parameter? The maximum-value function for this problem isv(γ) ≡ f r(x∗(γ); γ) = −x∗(γ)2 + γx∗(γ) + 2. Derive the function v(γ).(ii) For each of the particular values x = 1, 2, 3, write down the function f r(γ; x) inwhich x is parametric and γ is variable. Then, on the same diagram, plot the graphsof these three functions for γ ∈ [0, 8].(iii) Now add to your diagram the graph of the maximum-value function v(γ). Con-firm that v(γ) ≥ f r(γ; x) for γ ∈ [0, 8] with = for x = x∗(1), x = x∗(2), and x = x∗(3).

Page 428: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.9. PROBLEMS 409

Problem 12.5. Consider the family of functions f : X × Γ → �, where X = �,Γ = �++, and f(x, γ) = γ2x2 − γx+ 2.(i) What is the optimal solution x∗(γ) to the problem minx f

r(x; γ) = γ2x2 − γx+ 2in which γ > 0 is parametric? The minimum-value function is v(γ) ≡ f r(x∗(γ); γ) =γ2x∗(γ)2 − γx∗(γ) + 2. Derive the function v(γ).(ii) For each of the particular values x = 1

6, 14, 12, write down the function f r(γ; x) in

which x is parametric and γ is variable. Then, on the same diagram, plot the graphsof these three functions for γ ∈ (0, 5].(iii) Now add to your diagram the graph of the minimum-value function v(γ). Con-firm that v(γ) ≤ f r(γ; x) with = for x = x∗

(16

), x = x∗

(14

), and x = x∗

(12

).

The First-Order Upper-Envelope Theorem. Let X ⊆ �n be open with respectto En. Let Γ ⊆ �m be open with respect to Em. Let ψ : X × Γ �→ � be a functionthat is continuously differentiable with respect to (x, γ) ∈ X ×Γ. If, for every γ ∈ Γ,there exists x∗(γ) ∈ X such that

ψ(x∗(γ), γ) ≥ ψ(x, γ) ∀ x ∈ X,

then define the maximum-value mapping Θ∗ : Γ �→ � by

Θ∗(γ) = ψ(x∗(γ), γ) ∀ γ ∈ Γ.

Then Θ∗ is a continuously differentiable function on Γ and

∂Θ∗(γ)∂γk

=∂ψ(x, γ)

∂γk|x=x∗(γ).

The First-Order Lower-Envelope Theorem. Let X ⊆ �n be open with respectto En. Let Γ ⊆ �m be open with respect to Em. Let ψ : X × Γ �→ � be a functionthat is continuously differentiable with respect to (x, γ) ∈ X ×Γ. If, for every γ ∈ Γ,there exists an x∗(γ) ∈ X such that

ψ(x∗(γ), γ) ≤ ψ(x, γ) ∀ x ∈ X,

then define the minimum-value mapping Θ∗ : Γ �→ � by

Θ∗(γ) = ψ(x∗(γ), γ) ∀ γ ∈ Γ.

Then Θ∗ is a continuously differentiable function on Γ and

∂Θ∗(γ)∂γk

=∂ψ(x, γ)

∂γk|x=x∗(γ).

Page 429: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

410 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

Problem 12.6.

(i) Draw a simple diagram that illustrates the content of the First-Order Upper-Envelope Theorem and state in simple words the meaning of the result.

(ii) Prove the First-Order Upper-Envelope Theorem.

Problem 12.7.

(i) Draw a simple diagram that illustrates the content of the First-Order Lower-Envelope Theorem and state in simple words the meaning of the result.

(ii) Prove the First-Order Lower-Envelope Theorem.

Problem 12.8. Consider again the family of functions described in problem 12.4.

(i) Use the diagram that you constructed when answering parts (ii) and (iii) of prob-lem 12.4 to help you determine what information is provided by applying the First-Order Upper-Envelope Theorem to the envelope that is the maximum-value functionderived in part (iii) of problem 12.4.

(ii) Suppose that γ = 52. Use the First-Order Upper-Envelope Theorem to discover

x∗(52

), the optimal value for x when γ = 5

2. Once you have done so, return to the

diagram constructed for your answers to parts (ii) and (iii) of problem 12.4, andconfirm visually that you have correctly determined the value x∗

(52

).

(iii) Repeat part (ii), but with γ = 32.

Problem 12.9. Consider again the family of functions described in problem 12.5.

(i) Use the diagram that you constructed when answering parts (ii) and (iii) of prob-lem 12.5 to help you determine what information is provided by applying the First-Order Lower-Envelope Theorem to the envelope that is the minimum-value functionderived in part (iii) of problem 12.5.

(ii) Suppose that γ = 25. Use the First-Order Lower-Envelope Theorem to discover

x∗(25

), the optimal value for x when γ = 2

5. Once you have done so, return to the

diagram constructed for your answers to parts (ii) and (iii) of problem 12.5, andconfirm visually that you have correctly determined the value x∗

(25

).

(iii) Repeat part (ii), but with γ = 23.

Problem 12.10. The versions of the First-Order Envelope Theorems stated aboveare particular to constrained optimization problems in which the maximum-valuefunctions are differentiable, both with respect to the choice variables xi and theparameters γk. This question explores envelopes and their properties more generally.

(i) Prove that the maximum-value and minimum-value mappings Θ∗ and Θ∗ statedabove are functions.

Page 430: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.9. PROBLEMS 411

(ii) Provide a counterexample to the assertion that, if ψ(x, γ) is differentiable withrespect to (x, γ) ∈ X × Γ, then Θ∗ and Θ∗ must be differentiable with respect toγ ∈ Γ.

(iii) Prove that, if ψ(x, γ) is convex with respect to γ ∈ Γ, then Θ∗(γ) is convex withrespect to γ ∈ Γ.

(iv) Provide a counterexample to the assertion that, if ψ(x, γ) is convex with respectto γ ∈ Γ for given x ∈ X, then Θ∗(γ) is convex with respect to γ ∈ Γ.

(v) Prove that, if ψ(x, γ) is concave with respect to γ ∈ Γ, then Θ∗(γ) is concavewith respect to γ ∈ Γ.

(vi) Provide a counterexample to the assertion that, if ψ(x, γ) is concave with respectto γ ∈ Γ for given x ∈ X, then Θ∗(γ) is concave with respect to γ ∈ Γ.

Problem 12.11. A perfectly competitive firm’s “long-run” profit-maximization prob-lem is

maxy, x1, x2

py − w1x1 − w2x2 subject to y ≤ ψ(x1, x2).

ψ is the firm’s production function. y is the output level. x1 and x2 are the quantitiesused of inputs 1 and 2. p is the given per unit price of the firm’s product. w1 and w2

are the given per unit prices of inputs 1 and 2. Assume that there is a unique profit-maximizing plan (yl(p, w1, w2), x

l1(p, w1, w2), x

l2(p, w1, w2)) for any given (p, w1, w2) �

(0, 0, 0). Denote the multiplier for the production constraint by λl(p, w1, w2). Theproblem’s maximum-value function is the firm’s long-run indirect profit function

πl(p, w1, w2) ≡ pyl(p, w1, w2)− w1xl1(p, w1, w2)− w2x

l2(p, w1, w2)

+ λl(p, w1, w2)[ψ(xl1(p, w1, w2), x

l2(p, w1, w2))− yl(p, w1, w2)

].

(i) Use the First-Order Upper-Envelope Theorem to derive Harold Hotelling’s resultthat

∂πl(p, w1, w2)

∂p= yl(p, w1, w2).

(ii) Draw a carefully labeled diagram that explains Hotelling’s result. Give a simple,accurate verbal explanation of the result.

(iii) Prove that the firm’s long-run indirect profit function is a convex function of p.

Problem 12.12. A perfectly competitive firm’s “long-run” cost-minimization prob-lem is

minx1, x2

w1x1 + w2x2 subject to y ≤ ψ(x1, x2).

Page 431: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

412 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

ψ is the firm’s production function. y > 0 is a given output level. x1 and x2 arethe quantities used of inputs 1 and 2. w1 and w2 are the given per unit prices ofinputs 1 and 2. Assume that, for any given (w1, w2) � (0, 0), there is a unique cost-minimizing input bundle (xl1(w1, w2, y), x

l2(w1, w2, y)). Denote the multiplier for the

production constraint by μl(w1, w2, y). The problem’s minimum-value function is thefirm’s long-run cost function

cl(w1, w2, y) ≡ w1xl1(w1, w2, y) + w2x

l2(w1, w2, y)

+ μl(w1, w2, y)[ψ(xl1(w1, w2, y), x

l2(w1, w2, y))− y

].

(i) Use the First-Order Lower-Envelope Theorem to derive Shephard’s result that

∂cl(w1, w2, y)

∂wi

= xl(w1, w2, y) for i = 1, 2.

(ii) Draw a carefully labeled diagram that explains Shephard’s result. Give a simple,accurate verbal explanation of the result.(iii) Prove that the firm’s long-run cost function is a concave function of w1.

Problem 12.13. A consumer’s budget allocation problem is

maxx1, x2

U(x1, x2) subject to p1x1 + p2x2 ≤ y,

where U is the consumer’s direct utility function, y > 0 is the consumer’s bud-get, x1 and x2 are the consumption levels of commodities 1 and 2, and p1 and p2are the given per unit prices of commodities 1 and 2. Assume that, for any giv-en (p1, p2, y) � (0, 0, 0), there is a unique utility-maximizing consumption bundle(x∗1(p1, p2, y), x

∗2(p1, p2, y)). The problem’s maximum-value function is the consumer’s

indirect utility function

v(p1, p2, y) ≡ U(x∗1(p1, p2, y), x∗2(p1, p2, y))

+ λ∗(p1, p2, y)[y − p1x∗1(p1, p2, y)− p2x

∗2(p1, p2, y)].

(i) Use the First-Order Upper-Envelope Theorem to derive that

∂v(p1, p2, y)

∂pi= −λ∗(p1, p2, y)x∗i (p1, p2, y) for i = 1, 2, (12.81)

and∂v(p1, p2, y)

∂y= λ∗(p1, p2, y), (12.82)

and hence that∂v(p1, p2, y)/∂pi∂v(p1, p2, y)/∂y

= −x∗i (p1, p2, y) for i = 1, 2. (12.83)

Page 432: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.10. ANSWERS 413

(12.83) is known as Roy’s Identity, named after the French economist Rene Roy.(ii) Draw a carefully labeled diagram that explains (12.81) and (12.82). Give a simple,accurate verbal explanation of these results.

Problem 12.14. U(x1, . . . , xn) is a consumer’s direct utility function. The consumerfaces given positive prices p1, . . . , pn for commodities 1 to n. The consumer’s bud-get is y > 0. Assume that the consumer’s preferences are strictly monotonic andstrictly convex. Use e(p1, . . . , pn, u) to denote the consumer’s expenditure function,use hi(p1, . . . , pn, u) to denote the consumer’s Hicksian demand for commodity i, anduse x∗i (p1, . . . , pn, y) to denote the consumer’s ordinary demand for commodity i ;i = 1, . . . , n. For given p1, . . . , pn, and y, Slutsky’s equation is

∂x∗i∂pj

=∂hi∂pj

− x∗j∂x∗i∂y

for i, j = 1, . . . , n.

Derive Slutsky’s equation.

12.10 Answers

Answer to Problem 12.1. We start by noting that the first-order maximizationcondition (12.77) holds as a zero-valued identity if we evaluate (12.77) only at theproblem’s optimal solution x∗(α1, α2, α3). Thus we write

I(α1, α2, α3) ≡ 2α1x∗(α1, α2, α3) + α2 ≡ 0. (12.84)

Differentiation of this identity with respect to α1 yields

0 = 2x∗(α1, α2, α3) + 2α1∂x∗(α1, α2, α3)

∂α1

⇒ ∂x∗(α1, α2, α3)

∂α1

= − x∗(α1, α2, α3)

α1

.

(12.85)Similarly, differentiation of (12.84) with respect to α2 yields

0 = 2α1∂x∗(α1, α2, α3)

∂α2

+ 1 ⇒ ∂x∗(α1, α2, α3)

∂α2

= − 1

2α1

. (12.86)

Last, differentiation of (12.84) with respect to α3 yields

0 = 2α1∂x∗(α1, α2, α3)

∂α3

, so∂x∗(α1, α2, α3)

∂α3

= 0 (12.87)

because α1 = 0. You should use (12.79) to verify (12.85), (12.86), and (12.87).

Page 433: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

414 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

Answer to Problem 12.2. Consumption bundles with x1 = 0 or x2 = 0 can neverbe optimal, so the first-order optimality conditions are(∂U(x1, x2)

∂x1,∂U(x1, x2)

∂x2

) ∣∣∣∣x1=x∗

1(p1,p2,y)x2=x∗

2(p1,p2,y)

= (x∗2(p1, p2, y), x∗1(p1, p2, y)) = λ∗(p1, p2, y) (p1, p2)

and λ∗(p1, p2, y) (y − p1x∗1(p1, p2, y)− p2x

∗2(p1, p2, y)) = 0,

yielding the identities

I1(p1, p2, y) ≡ x∗2(p1, p2, y)− λ∗(p1, p2, y)p1 ≡ 0, (12.88)

I2(p1, p2, y) ≡ x∗1(p1, p2, y)− λ∗(p1, p2, y)p2 ≡ 0, (12.89)

and I3(p1, p2, y) ≡ λ∗(p1, p2, y) (y − p1x∗1(p1, p2, y)− p2x

∗2(p1, p2, y)) ≡ 0. (12.90)

Differentiation with respect to p1 gives

0 =∂x∗2(·)∂p1

− ∂λ∗(·)∂p1

p1 − λ∗(·), (12.91)

0 =∂x∗1(·)∂p1

− ∂λ∗(·)∂p1

p2, and (12.92)

0 =∂λ∗(·)∂p1

(y − p1x∗1(·)− p2x

∗2(·))− λ∗(·)

(x∗1(·) + p1

∂x∗1(·)∂p1

+ p2∂x∗2(·)∂p1

). (12.93)

The first term on the right-hand side of (12.93) is zero, so we have

∂x∗2(·)∂p1

− p1∂λ∗(·)∂p1

= λ∗(·), (12.94)

∂x∗1(·)∂p1

− p2∂λ∗(·)∂p1

= 0, (12.95)

and λ∗(·)p1∂x∗1(·)∂p1

+ λ∗(·)p2∂x∗2(·)∂p1

= − λ∗(·)x∗1(·). (12.96)

The solution to these linear equations is

∂x∗1(p1, p2, y)∂p1

= − p2λ∗ + x∗12p1

,∂x∗2(p1, p2, y)

∂p1=p2λ

∗ − x∗12p2

,

and∂λ∗(p1, p2, y)

∂p1= − p2λ

∗ + x∗12p1p2

.

(12.97)

Page 434: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.10. ANSWERS 415

Using the first-order condition (12.89) simplifies (12.97) to

∂x∗1(p1, p2, y)∂p1

= − p2λ∗

p1< 0,

∂x∗2(p1, p2, y)∂p1

= 0, and∂λ∗(p1, p2, y)

∂p1= − λ∗

p1< 0.

An increase to p1 results in a lower quantity demanded of commodity 1, no changeto the quantity demanded of commodity 2, and a lower value for the consumer’smarginal utility of expenditure. You should use the algebraically explicit optimalsolution stated in the problem to verify these results.

Answer to Problem 12.3. The problem’s first-order optimality conditions are thatthere exists a λ∗(·) ≥ 0 such that(

∂U(x1, x2)

∂x1,∂U(x1, x2)

∂x2

) ∣∣∣∣x1=x∗1(p1,p2,y)

x2=x∗2(p1,p2,y)

=

(1

2√x∗1(p1, p2, y)

,1

2√x∗2(p1, p2, y)

)

= λ∗(p1, p2, y) (p1, p2)

and λ∗(p1, p2, y) (y − p1x∗1(p1, p2, y)− p1x

∗2(p1, p2, y)) = 0,

yielding the identities

I1(p1, p2, y) ≡1

2√x∗1(p1, p2, y)

− λ∗(p1, p2, y)p1 ≡ 0, (12.98)

I2(p1, p2, y) ≡1

2√x∗2(p1, p2, y)

− λ∗(p1, p2, y)p2 ≡ 0, (12.99)

and I3(p1, p2, y) ≡ λ∗(p1, p2, y) (y − p1x∗1(p1, p2, y)− p2x

∗2(p1, p2, y)) ≡ 0. (12.100)

These are constant-valued identities, so

∂I1(p1, p2, y)

∂γ= 0,

∂I2(p1, p2, y)

∂γ= 0, and

∂I3(p1, p2, y)

∂γ= 0,

where γ is any one of the parameters p1, p2 or y. We are asked to set γ = p1.Differentiating each of (12.98), (12.99), and (12.100) with respect to p1 gives

0 = − 1

4x∗1(·)−3/2 ∂x

∗1(·)∂p1

− ∂λ∗(·)∂p1

p1 − λ∗(·), (12.101)

0 = − 1

4x∗2(·)−3/2 ∂x

∗2(·)∂p1

− ∂λ∗(·)∂p1

p2, and (12.102)

0 =∂λ∗(·)∂p1

(y − p1x∗1(·)− p2x

∗2(·))− λ∗(·)

(x∗1(·) + p1

∂x∗1(·)∂p1

+ p2∂x∗2(·)∂p1

).

(12.103)

Page 435: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

416 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

The first term on the right-hand side of (12.103) is zero because the budget constraintbinds. A little tidying up of (12.101), (12.102), and (12.103) gives

1

4x∗1(·)−3/2 ∂x

∗1(·)∂p1

+ p1∂λ∗(·)∂p1

= −λ∗(·), (12.104)

1

4x∗2(·)−3/2 ∂x

∗2(·)∂p1

+ p2∂λ∗(·)∂p1

= 0, (12.105)

and λ∗(·)p1∂x∗1(·)∂p1

+ λ∗(·)p2∂x∗2(·)∂p1

= −λ∗(·)x∗1(·). (12.106)

λ∗(·) > 0 because the budget constraint is strictly binding. This lets us divide (12.106)by λ∗(·) and we end up with

1

4x∗1(·)−3/2 ∂x

∗1(·)∂p1

+ p1∂λ∗(·)∂p1

= −λ∗(·), (12.107)

1

4x∗2(·)−3/2 ∂x

∗2(·)∂p1

+ p2∂λ∗(·)∂p1

= 0, (12.108)

and p1∂x∗1(·)∂p1

+ p2∂x∗2(·)∂p1

= −x∗1(·). (12.109)

The solution is

∂x∗1(p1, p2, y)∂p1

= −

p1x∗1

4(x∗2)3/2+ p22λ

p224(x∗1)3/2

+p21

4(x∗2)3/2

, (12.110)

∂x∗2(p1, p2, y)∂p1

= −p2

(1

4(x∗1)1/2− p1λ

∗)

p224(x∗1)3/2

+p21

4(x∗2)3/2

, (12.111)

and∂λ∗(p1, p2, y)

∂p1=

1

4(x∗2)3/2

(1

4(x∗1)1/2− p1λ

∗)

p224(x∗1)3/2

+p21

4(x∗2)3/2

. (12.112)

These expressions can be simplified. Squaring and cubing (12.98) gives

1

4x∗1= λ∗2p21 ⇒ x∗1 =

1

4p21λ∗2 and

1

8x∗13/2

= λ∗3p31 ⇒ 1

4x∗13/2

= 2λ∗3p31. (12.113)

Page 436: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.10. ANSWERS 417

Substituting from (12.113) into (12.110), (12.111), and (12.112) gives

∂x∗1(p1, p2, y)∂p1

= − (2p1 + p2)x∗1

p1(p1 + p2)< 0,

∂x∗2(p1, p2, y)∂p1

=p1x

∗1

p2(p1 + p2)> 0,

and∂λ∗(p1, p2, y)

∂p1= − p2λ

2p1(p1 + p2)< 0.

Without knowing the explicit equations for x∗1, x∗2, and λ

∗, we have discovered thatan increase to p1 results in a lower quantity demanded of commodity 1, an increaseto the quantity demanded for commodity 2, and a lower value for the consumer’smarginal utility of expenditure.

Answer to Problem 12.4.(i) The optimal solution to the problem maxx f

r(x; γ) = −x2 + γx+ 2, in which γ isparametric, necessarily satisfies the first-order maximization condition

0 =df r(x; γ)

dx|x=x∗(γ) = −2x∗(γ) + γ ⇒ x∗(γ) =

γ

2.

The second-order condition confirms that x∗(γ) is a local maximum for f r(x; γ). Sincethis is the function’s only critical point, the point is a global maximum for f r(x; γ).Consequently the maximum-value function for the problem is

v(γ) ≡ f r(x∗(γ); γ) = −x∗(γ)2 + γx∗(γ) + 2 =γ2

4+ 2.

(ii) and (iii). See Figure 12.6.x = 1 is an optimal solution if and only if x∗(γ) = 1; i.e. if and only if γ/2 = 1.

So x = 1 = x∗(γ) if and only if γ = 2. The value of the function f r(x; γ) when x = 1and γ = 2 is f r(1; 2) = −12 + 2× 1 + 2 = 3. The value of the envelope function v(γ)when γ = 2 is v(2) = 22/4 + 2 = 3. Thus the restricted function f r(γ; x = 1) is apart (just a point) of the envelope at γ = 2.

x = 2 is an optimal solution if and only if x∗(γ) = 2; i.e. if and only if γ/2 = 2.So x = 2 = x∗(γ) if and only if γ = 4. The value of the function f r(x; γ) when x = 2and γ = 4 is f r(2; 4) = −22 + 4× 2 + 2 = 6. The value of the envelope function v(γ)when γ = 2 is v(2) = 42/4 + 2 = 6. Thus the restricted function f r(γ; x = 2) is apart of the envelope at γ = 4.

x = 3 is an optimal solution if and only if x∗(γ) = 3; i.e. if and only if γ/2 = 3.So x = 3 = x∗(γ) if and only if γ = 6. The value of the function f r(x; γ) when x = 3and γ = 6 is f r(3; 6) = −32+6× 3+2 = 11. The value of the envelope function v(γ)

Page 437: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

418 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

Figure 12.6: The upper envelope for problem 12.4.

when γ = 6 is v(2) = 62/4 + 2 = 11. Thus the restricted function f r(γ; x = 3) is apart of the envelope at γ = 6.

Answer to Problem 12.5.(i) The optimal solution to the problem minx f

r(x; γ) = γ2x2 − γx+ 2, in which γ >0 is parametric, necessarily satisfies the first-order minimization condition

0 =df r(x; γ)

dx|x=x∗(γ) = 2γ2x∗(γ)− γ ⇒ x∗(γ) =

1

2γ.

The second-order condition confirms that x∗(γ) is a local minimum for f r(x; γ). Sincethis is the function’s only critical point, the point is a global minimum for f r(x; γ)and so the problem’s minimum-value function is

v(γ) ≡ f r(x∗(γ); γ) = γ2x∗(γ)2 − γx∗(γ) + 2 ≡ 7

4.

(ii) and (iii). See Figure 12.7.x = 1/6 is an optimal solution if and only if x∗(γ) = 1/6; i.e. if and only if

1/2γ = 1/6. So x = 1/6 = x∗(γ) if and only if γ = 3. The value of the function

Page 438: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.10. ANSWERS 419

Figure 12.7: The lower envelope for problem 12.5.

f r(x; γ) when x = 1/6 and γ = 3 is f r(1/6; 3) = 32×(1/6)2−3×1/6+2 = 7/4 = v(3).Thus the restricted function f r(γ; x = 1/6) is a part of the envelope at γ = 3.

x = 1/4 is an optimal solution if and only if x∗(γ) = 1/4; i.e. if and only if1/2γ = 1/4. So x = 1/4 = x∗(γ) if and only if γ = 2. The value of f r(x; γ)when x = 1/4 and γ = 2 is f r(1/4; 2) = 7/4 = v(2). Thus the restricted functionf r(γ; x = 1/4) is a part of the envelope at γ = 2.

x = 1/2 is an optimal solution if and only if x∗(γ) = 1/2; i.e. if and only if1/2γ = 1/2. So x = 1/2 = x∗(γ) if and only if γ = 1. The value of f r(x; γ)when x = 1/2 and γ = 1 is f r(1/2; 1) = 7/4 = v(1). Thus the restricted functionf r(γ; x = 1/2) is a part of the envelope at γ = 1.

Answer to Problem 12.6.(i) The diagram should be drawn with the horizontal axis measuring the valuesof a parameter γk and the vertical axis measuring the value of a maximum-valuefunction Θ∗(γk). A few values, say, x′, x′′, and x′′′, should be taken from the choiceset X. For each fixed value of x, a graph of ψ(x, γk) should be drawn. In this examplethere will be three such graphs, one for each of x′, x′′, and x′′′. The upper-envelope

Page 439: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

420 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

of these three graphs consists of the segments of these graphs that are highest; i.e.

Θ∗(γk) = max {ψ(x′, γk), ψ(x′′, γk), ψ(x′′′, γk)}.

Now notice that, wherever the envelope Θ∗ coincides with a particular ψ(x, γk), theslope with respect to γk of the envelope Θ

∗(γk) is the same as the slope with respect toγk of the ψ(x, γk) curve with the value of x fixed. This is the content of the theorem.(ii) Proof. For any γ ∈ Γ, there exists an x∗(γ) ∈ X for which

ψ(x∗(γ), γ) ≥ ψ(x, γ) ∀ x ∈ X.

Since ψ is continuously differentiable on X × Γ, necessarily

∂ψ(x, γ)

∂xi|x=x∗(γ) = 0 ∀ i = 1, . . . , n. (12.114)

Since (12.114) is always evaluated at the maximizing solution x∗(γ), (12.114) holdsas an identity. Since ψ is continuously differentiable on X × Γ, the Implicit FunctionTheorem asserts that x∗(γ) is a continuously differentiable function on Γ. Therefore

∂Θ∗(γ)∂γk

=∂ψ(x, γ)

∂γk|x=x∗(γ) +

n∑i=1

∂ψ(x, γ)

∂xi|x=x∗(γ) ×

∂x∗i (γ)∂γk

=∂ψ(x, γ)

∂γk|x=x∗(γ) by (12.114).

Answer to Problem 12.7.(i) The diagram should be drawn with the horizontal axis measuring the valuesof a parameter γk and the vertical axis measuring the value of a minimum-valuefunction Θ∗(γk). A few values, say x′, x′′, and x′′′, should be taken from the choiceset X. For each fixed value of x, a graph of ψ(x, γk) should be drawn. In this examplethere will be three such graphs, one for each of x′, x′′, and x′′′. The lower-envelope ofthese three graphs consists of the segments of these graphs that are lowest; i.e.

Θ∗(γk) = min {ψ(x′, γk), ψ(x′′, γk), ψ(x′′′, γk)}.

Now notice that, wherever the envelope Θ∗ coincides with a particular ψ(x, γk), theslope with respect to γk of the envelope Θ∗(γk) is the same as the slope with respect toγk of the ψ(x, γk) curve with the value of x fixed. This is the content of the theorem.(ii) Proof. For any γ ∈ Γ, there exists an x∗(γ) ∈ X for which

ψ(x∗(γ), γ) ≤ ψ(x, γ) ∀ x ∈ X.

Page 440: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.10. ANSWERS 421

Since ψ is continuously differentiable on X × Γ, necessarily

∂ψ(x, γ)

∂xi|x=x∗(γ) = 0 ∀ i = 1, . . . , n. (12.115)

Since (12.115) is always evaluated at the minimizing solution x∗(γ), (12.115) holdsas an identity. Since ψ is continuously differentiable on X × Γ, the Implicit FunctionTheorem asserts that x∗(γ) is a continuously differentiable function on Γ. Therefore

∂Θ∗(γ)∂γk

=∂ψ(x, γ)

∂γk|x=x∗(γ) +

n∑i=1

∂ψ(x, γ)

∂xi|x=x∗(γ) ×

∂x∗i(γ)∂γk

=∂ψ(x, γ)

∂γk|x=x∗(γ) by (12.115).

Answer to Problem 12.8. For the problem maxx fr(x; γ) = −x2 + γx+ 2, in which

γ is parametric, the maximum value is

v(γ) ≡ f r(x∗(γ); γ) = −x∗(γ)2 + γx∗(γ) + 2.

(i) The First-Order Upper-Envelope Theorem states that the two-step process thatderives the value of the rate of change of the maximum value v(γ) with respect to γat a particular value γ′ of γ is, first, to fix the value of x at the number x′ = x∗(γ′)and, second, to differentiate the resulting function with respect to γ. The first stepgives us the function of only γ (since x is fixed at the number x′) that is

f r(γ; x′) = − (x′)2 + γx′ + 2.

The value at γ = γ′ of the derivative of this function with respect to γ is

df r(γ; x′)dγ

|γ=γ′ = x′.

The First-Order Upper-Envelope Theorem informs us that this is the value at γ = γ′

of the rate of change with respect to γ of the maximum value v(γ); i.e.

dv(γ)

dγ|γ=γ′ =

df r(γ; x′)dγ

|γ=γ′ = x′ = x∗(γ′).

Since the maximum-value function is v(γ) = γ2/4 + 2, we have that

x∗(γ′) =d

(γ2

4+ 2

)|γ=γ′ =

γ′

2.

Page 441: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

422 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

This is the optimality result obtained in part (i) of problem 12.4. The point is thatthe same result can be obtained directly from a knowledge of only the maximum-valuefunction and the family of (restricted) functions that generate the envelope that isthe maximum-value function.

(ii) If γ = 52, then the corresponding optimal value for x is x∗

(52

)= 5/2

2= 5

4. The

restricted function in which the value of x is fixed at 54is f r

(γ; x = 5

4

)= −

(54

)2+ 5

4γ+

2 = 716+ 5

4γ. At γ = 5

2, the value of this restricted function is f r

(52; 54

)= 7

16+ 5

4× 5

2= 57

16,

which is the maximum value for γ = 52. So the envelope and the restricted function

are the one and the same for γ = 52. Therefore the slopes of these two functions are

the same for γ = 52, and this common slope is x∗

(52

)= 5

4. Add the graph of 7

16+ 5

4γ to

your diagram for parts (ii) and (iii) of problem 12.4, and you will observe a tangencywith the graph of the maximum-value function at γ = 5

2. The slope at the tangency

is 54.

(iii) If γ = 32, then the corresponding optimal value for x is x∗

(32

)= 3/2

2= 3

4. The

restricted function in which the value of x is fixed at 34is f r

(γ; x = 3

4

)= −

(34

)2+ 3

4γ+

2 = 2316

+ 34γ. At γ = 3

2, the value of this restricted function is f r

(32; 34

)= 41

16, which

is the maximum value for γ = 32. So the envelope and the restricted function are the

same for γ = 32. Therefore the slopes of these two functions are the same for γ = 3

2,

and this common slope is x∗(32

)= 3

4. Add the graph of 23

16+ 3

4γ to your diagram for

parts (ii) and (iii) of problem 12.4, and you will observe a tangency with the graphof the maximum-value function at γ = 3

2. The slope at the tangency is 3

4.

Answer to Problem 12.9. For the problem minx f(x; γ) = γ2x2 − γx+ 2, in whichγ is parametric, the minimum-value function is v(γ) ≡ 7

4.

(i) To apply the First-Order Lower-Envelope Theorem, we first select some value γ′

for γ. Next we fix the value of x at whatever is the number x′ = x∗(γ′). This givesus the function

f r(γ; x′) = γ2 (x′)2 − γx′ + 2.

We differentiate this function of only γ and evaluate the derivative at γ = γ′, obtaining

df r(γ; x′)dγ

|γ=γ′ = 2γ′ (x′)2 − x′.

The First-Order Lower-Envelope Theorem informs us that this is the value at γ = γ′

of the rate of change with respect to γ of the minimum value v(γ); i.e.

dv(γ)

dγ|γ=γ′ =

df r(γ; x′)dγ

|γ=γ′ = 2γ′ (x′)2 − x′ = 2γ′x∗(γ′)2 − x∗(γ′).

Page 442: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.10. ANSWERS 423

Since the minimum-value function is v(γ) ≡ 74we have that

2γ′x∗(γ′)2 − x∗(γ′) =d

(7

4

)|γ=γ′ = 0 ⇒ x∗(γ′) =

1

2γ′.

This is the optimality result obtained in part (i) of problem 12.5.(ii) If γ = 2

5, then the corresponding optimal value for x is x∗

(25

)= 5

4. The restricted

function in which the value of x is fixed at 54is f r

(γ; x = 5

4

)= 25

16γ2 − 5

4γ + 2. At

γ = 25, the value of this restricted function is f r

(25; 54

)= 7

4, which is the minimum

value for γ = 25. So the envelope and the restricted function are the same for γ = 2

5.

Therefore the slopes of these two functions are the same for γ = 25, and this common

slope is 2× 25× x∗

(25

)2 − x∗(25

)= 4

5× 25

16− 5

4= 0. Add the graph of 25

16γ2 − 5

4γ + 2 to

your diagram for parts (ii) and (iii) of problem 12.5, and you will observe a tangencywith the graph of the minimum-value function at γ = 2

5. The slope at the tangency

is zero.(iii) If γ = 2

3, then the corresponding optimal value for x is x∗

(23

)= 3

4. The restricted

function in which the value of x is fixed at 34is f r(γ; x = 3

4) = 9

16γ2 − 3

4γ + 2. At

γ = 23, the value of this restricted function is f r

(23; 34

)= 7

4, which is the minimum

value for γ = 23. So the envelope and the restricted function are the same for γ = 2

3.

Therefore the slopes of these two functions are the same for γ = 23, and this common

slope is 2× 23× x∗

(23

)2 − x∗(23

)= 4

3× 9

16− 3

4= 0. Add the graph of 9

16γ2 − 3

4γ +2 to

your diagram for parts (ii) and (iii) of problem 12.5, and you will observe a tangencywith the graph of the minimum-value function at γ = 2

3. The slope at the tangency

is zero.

Answer to Problem 12.10.(i) Proof. Suppose that Θ∗ is not a function on Γ. Then there exists at least onevalue γ′ ∈ Γ such that there are at least two distinct values θ, θ ∈ Θ(γ′). Eitherθ > θ or vice versa. Suppose θ > θ. But then θ is not a maximum value for γ′.Contradiction. Hence Θ∗ is a function on Γ.

A similar argument establishes that Θ∗ is also a function on Γ.(ii) Consider the family of functions ψ : [−1, 1]× [−1, 1] → [−1, 1], where

ψ(x, γ) = γx for x ∈ [−1, 1] and γ ∈ [−1, 1].

Every function in this family is continuously differentiable to any order. The family’supper and lower envelopes are

Θ∗(γ) = max−1≤x≤1

γx =

{−γ ;−1 ≤ γ < 0

+γ ; 0 ≤ γ ≤ 1and Θ∗(γ) = min

−1≤x≤1γx =

{+γ ;−1 ≤ γ < 0

−γ ; 0 ≤ γ ≤ 1.

Page 443: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

424 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

Neither Θ∗ nor Θ∗ is differentiable at γ = 0.

(iii) Proof. Take γ′, γ′′ ∈ Γ with γ′ = γ′′. Then, for any μ ∈ [0, 1],

Θ∗(μγ′ + (1− μ)γ′′) = ψ(x∗(μγ′ + (1− μ)γ′′), μγ′ + (1− μ)γ′′)

≤ μψ(x∗(μγ′ + (1− μ)γ′′), γ′) + (1− μ)ψ(x∗(μγ′ + (1− μ)γ′′), γ′′),

because ψ(x, γ) is convex with respect to γ. x∗(μγ′ + (1 − μ)γ′′) is not necessarilya maximizing value for x when γ = γ′, and x∗(μγ′ + (1 − μ)γ′′) is not necessarily amaximizing value for x when γ = γ′′; i.e.

ψ(x∗(μγ′+(1−μ)γ′′), γ′)≤ψ(x∗(γ′), γ′) and ψ(x∗(μγ′+(1−μ)γ′′), γ′′)≤ψ(x∗(γ′′), γ′′).

Therefore

Θ∗(μγ′+(1−μ)γ′) ≤ μψ(x∗(γ′), γ′)+(1−μ)ψ(x∗(γ′′), γ′′)=μΘ∗(γ′)+(1−μ)Θ∗(γ′′).

(iv) The example given as an answer to part (ii) is a counterexample.

(v) Proof. Take γ′, γ′′ ∈ Γ with γ′ = γ′′. Then, for any μ ∈ [0, 1],

Θ∗(μγ′ + (1− μ)γ′′) = ψ(x∗(μγ′ + (1− μ)γ′′), μγ′ + (1− μ)γ′′)

≥ μψ(x∗(μγ′ + (1− μ)γ′′), γ′) + (1− μ)ψ(x∗(μγ′ + (1− μ)γ′′), γ′′),

because ψ(x, γ) is concave with respect to γ. x∗(μγ′ + (1 − μ)γ′′) is not necessarilya minimizing value for x when γ = γ′, and x∗(μγ′ + (1 − μ)γ′′) is not necessarily aminimizing value for x when γ = γ′′; i.e.

ψ(x∗(μγ′+(1−μ)γ′′), γ′)≥ψ(x∗(γ′), γ′) and ψ(x∗(μγ′+(1−μ)γ′′), γ′′)≥ψ(x∗(γ′′), γ′′).

Therefore

Θ∗(μγ′+(1−μ)γ′′)≥μψ(x∗(γ′), γ′)+(1−μ)ψ(x∗(γ′′), γ′′)=μΘ∗(γ′)+(1−μ)Θ∗(γ′′).

(vi) The example given as an answer to part (ii) is a counterexample.

Answer to Problem 12.11.

(i) Applying the First-Order Upper-Envelope Theorem to the firm’s long-run indirectprofit function proves the result directly.

(ii) See Figure 12.8. The diagram’s horizontal axis measures the per unit price p ofthe firm’s product. The vertical axis measures the firm’s profit.

Page 444: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.10. ANSWERS 425

Figure 12.8: Hotelling’s result that ∂π∗(p, ·)/∂p = y∗(p, ·).

Consider the firm’s profit problem in which the firm is restricted to producing agiven quantity y ≥ 0 of its product. This “short-run” problem is

maxx1, x2

py − w1x1 − w2x2 subject to − ψ(x1, x2) ≤ −y.

This is the problem of locating an input bundle that minimizes the cost of producingy units of the firm’s product, but we will continue to consider this problem as one ofmaximizing profit. The solution will be denoted by (xs1(p, w1, w2; y ), x

s2(p, w1, w2; y )).

The multiplier for the production constraint will be denoted by λs(p, w1, w2; y ). Themaximum-value function for this restricted problem is the “short-run” indirect profitfunction

πs(p, w1, w2; y ) ≡ py − w1xs1(p, w1, w2; y )− w2x

s2(p, w1, w2; y )

+ λs(p, w1, w2; y )[−y + ψ(xs1(p, w1, w2; y ), xs2(p, w1, w2; y ))].

Notice that this function is affine with respect to p, because y is a given constant.Select, say, three values p′, p′′, and p′′′ for p, with p′ < p′′ < p′′′. Use y ′ to denote

the value for y that is long-run profit-maximizing when p = p′; i.e. y ′ = yl(p′, w1, w2).

Page 445: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

426 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

Similarly, use y ′′ and y ′′′ to denote the values for y that are long-run profit-maximizingfor p = p′′ and p = p′′′, respectively; y ′ < y ′′ = yl(p′′, w1, w2) < y ′′′ = yl(p′′′, w1, w2).

Plot the short-run profit function πs(p, w1, w2; y′) on the diagram. The graph is

a straight line with a positive slope of y ′ profit dollars per unit increase in p. Thegraph’s intercept is the value of the short-run indirect profit when p = 0. This is

πs(0, w1, w2; y′) = −w1x

s1(0, w1, w2; y

′)− w2xs2(0, w1, w2; y

′).

The graph of πs(p, w1, w2; y′′) is a straight line with a slope of y ′′ and a vertical

intercept of

πs(0, w1, w2; y′′) = −w1x

s1(0, w1, w2; y

′′)− w2xs2(0, w1, w2; y

′′)

< −w1xs1(0, w1, w2; y

′)− w2xs2(0, w1, w2; y

′),

since y ′′ > y ′. Thus the graph of πs(p, w1, w2; y′′) commences at a lower vertical

intercept than does the graph of πs(p, w1, w2; y′), and increases at a greater constant

positive slope. Similarly, the graph of πs(p, w1, w2; y′′′) commences at a still lower

vertical intercept, and increases at the even greater constant positive slope y′′′.The graph of the long-run indirect profit function is the upper envelope of affine

graphs such as the three described above. Therefore the long-run indirect profit func-tion’s slope with respect to p is the same as the slope of the short-run indirect profitfunction that is part of the envelope at that point. Thus, for the value p′ of p, theslope ∂πl(p, w1, w2)/∂p evaluated at p′ is the same as the slope ∂πs(p, w1, w2; y

′)/∂pevaluated at p′, because y ′ is the long-run profit-maximizing choice for y whenp = p′, meaning that the short-run profit function πs(p, w1, w2; y

′) is the envelopeπl(p, w1, w2) when p = p′. Therefore, at p = p′, the slope with respect to p of theenvelope πl(p, w1, w2) is the slope with respect to p of the short-run profit functionπs(p, w1, w2; y

′). That is, ∂πl(p, w1, w2)/∂p = y ′ = ∂πs(p, w1, w2; y′)/∂p at p = p′.

The same statements may be made for p = p′′ and for p = p′′′.(iii) Proof. The firm’s Lagrange function

L(y, x1, x2, λ; p, w1, w2) = py − w1x1 − w2x2 + λ(−y + ψ(x1, x2))

is jointly convex in (p, y). Applying the result stated in part (iii) of problem 12.10gives that the firm’s long-run indirect profit function is a convex function of p.

Answer to Problem 12.12.

(i) Applying the First-Order Lower-Envelope Theorem to the firm’s long-run costfunction immediately yields the result.

Page 446: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.10. ANSWERS 427

(ii) See Figure 12.9. Consider the short-run circumstance in which the firm has afixed input. Suppose this is input 1. Denote the level at which it is fixed by x1.Input 2 remains variable. The firm’s short-run total cost function is then

cs(w1, w2, y; x1) = w1x1 + w2x2 + λs(w1, w2, y; x1)(y − ψ(x2; x1)).

Figure 12.9: Shephard’s result that ∂cl(w1, w2, y)/∂w1 = xl1(w1, w2, y).

Figure 12.9 displays three short-run cost functions, one for x1 fixed at x1′, another

for x1 fixed at x1′′, and the other for x1 fixed at x1

′′′, where x1′ < x1′′ < x1

′′′. Eachshort-run cost function is affine with respect to w1, and has a slope of x1 when plottedwith w1 measured on the horizontal axis and cost measured on the vertical axis.

The graph of the short-run cost function cs(w1, w2, y; x1′) is a straight line with a

positive slope of x1′ cost dollars per unit increase in w1. The graph’s vertical intercept

is the value of the short-run variable input 2 cost when w1 = 0. This variable inputcost is w2x2

′, where y = ψ(x1′, x′2); i.e. x2

′ is the quantity of the variable input 2 that,combined with the fixed x1

′ units of input 1, provides y output units. This variableinput 2 requirement, and its cost, both decrease as the fixed quantity x1 of input 1increases.

Page 447: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

428 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

The graph of cs(w1, w2, y; x1′′) is a straight line with a greater slope of x1

′′ and alesser vertical intercept of w2x

′′2, where y = ψ(x1

′′, x′′2) and x′′2 < x′2 because x1

′′ > x1′.

Similarly, the graph of cs(w1, w2, y; x1′′′) is a straight line with a still greater slope of

x1′′′ and a still lesser vertical intercept of w2x

′′′2 , where y = ψ(x1

′′′, x′′′2 ). The graphof the long-run cost function is the lower envelope of affine graphs such as the threedescribed above. Therefore the long-run cost function’s slope with respect to w1 is thesame as the slope of the short-run cost function that is part of the envelope at thatpoint. Thus, for the value w′

1 of w1, the slope ∂cl(w1, w2, y)/∂w1 evaluated at w′

1 is thesame as the slope ∂cs(w1, w2, y; x1

′)/∂w1 evaluated at w′1, because x1

′ is the long-runcost-minimizing choice for x1 when w1 = w′

1, meaning that the short-run cost functioncs(w1, w2, y; x1

′) is the envelope cl(w1, w2, y) when w1 = w′1. Therefore, at w1 = w′

1,the slope with respect to w1 of the envelope cl(w1, w2, y) is the slope with respect tow1 of the short-run cost function cs(w1, w2, y; x1

′). That is, ∂cl(w1, w2, y)/∂w1 = x1′ =

∂cs(w1, w2, y; x1′)/∂w1 at w1 = w′

1. The same statements may be made for w1 = w′′1

and for w1 = w′′′1 .

(iii) Proof. The firm’s Lagrange function

L(x1, x2, λ;w1, w2, y) = w1x1 + w2x2 + λ(y − ψ(x1, x2))

is concave with respect to w1. Applying the result stated in part (v) of problem 12.10gives that the firm’s long-run cost function is a concave function of w1.

Answer to Problem 12.13.(i) Applying the First-Order Upper-Envelope Theorem to the indirect utility functionby differentiating with respect to pi gives (12.81) immediately. (12.82) is obtainedimmediately by applying the First-Order Upper-Envelope Theorem to the indirectutility function and differentiating with respect to y.(ii) We will begin with an explanation of (12.81), for i = 1. Consider the consumer’sbudget allocation problem modified by the restriction that the quantity consumed ofcommodity 1 is fixed at x1 > 0. Use (x1, x

c2(p1, p2, y; x1), λ

c(p1, p2, y; x1)) to denotethe optimal solution to this problem. The maximum-value function for the problemis the quantity-constrained indirect utility function

vc(p1, p2, y; x1) ≡ U(x1, xc2(p1, p2, y; x1))+λ

c(p1, p2, y; x1)(y−p1x1−p2xc2(p1, p2, y; x1)).

Note that vc is affine with respect to p1, with a slope of −λc(p1, p2, y; x1) x1. Theusual argument now applies. The upper envelope of all of the quantity-constrainedindirect utility functions vc is the indirect utility function v. Any point on the graph ofthis upper-envelope must be a point on the graph of one of the quantity-constrained

Page 448: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

12.10. ANSWERS 429

indirect utility functions, and so the slopes, at this point, of both curves are thesame. The slope of the indirect utility function is ∂v/∂p1 for some value p′1 of p1.The slope of the quantity-constrained indirect utility function that coincides with theindirect utility function at p1 = p′1 is −λ(p′1, p2, y; x′1) x′1, where x′1 is the same as thevalue x∗1(p

′1, p2, y) that is the consumer’s rational (with x1 free) choice of the quantity

demanded of commodity 1 when p1 = p′1. This explains (12.81).The explanation for (12.82) is analogous to the explanation of (12.81) once the

constraint on the consumer is changed from a fixed level of x1 to a fixed incomelevel y. Try it.

Answer to Problem 12.14. Let v(p1, . . . , pn, y) denote the consumer’s indirect util-ity function. For given values of p1, . . . , pn, and y, the consumer’s ordinary demand forcommodity i is x∗i (p1, . . . , pn, y). The consumer’s Hicksian demand for commodity i ishi(p1, . . . , pn, u), where u = v(p1, . . . , pn, y), and x

∗i (p1, . . . , pn, y) = hi(p1, . . . , pn, u).

Since the consumer’s (minimum-value) expenditure function is

e(p1, . . . , pn, u) ≡ p1h1(p1, . . . , pn, u) + · · ·+ pnhn(p1, . . . , pn)

+ λ∗(p1, . . . , pn, u)(u− U(h1(p1, . . . , pn, u), . . . , hn(p1, . . . , pn, u))),

it follows from the First-Order Lower-Envelope Theorem that

∂e

∂pi= hi(p1, . . . , pn, u); i = 1, . . . , n. (12.116)

u ≡ v(p1, . . . , pn, e(p1, . . . , pn, u)), so

0 =∂v

∂pi+∂v

∂y× ∂e

∂pi

which, from (12.116) and Roy’s Identity, asserts that, for y = e(p1, . . . , pn, u),

0 = −x∗i (p, y)×∂v

∂y+∂v

∂y× hi(p, u). (12.117)

The consumer’s marginal utility of expenditure ∂v/∂y is positive (thus, not zero)because the consumer’s preferences are strictly monotonic, so (12.117) simplifies to

hi(p, u) = x∗i (p, y) when y = e(p, u), (12.118)

where p = (p1, . . . , pn). That is,

hi(p, u) ≡ x∗i (p, e(p, u)) for i = 1, . . . , n.

Page 449: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

430 CHAPTER 12. COMPARATIVE STATICS ANALYSIS, PART I

Since this is an identity, the rates of change of the left and right sides are the same;

∂hi(p, u)

∂pj=∂x∗i (p, y)∂pj

|y=e(p,u) +∂x∗i (p, y)

∂y|y=e(p,u) ×

∂e(p, u)

∂pj. (12.119)

Using (12.116) and (12.118) in (12.119) gives Slutsky’s equation:

∂hi(p, u)

∂pj=∂x∗i (p, y)∂pj

|y=e(p,u)+x∗j(p, y)|y=e(p,u)×

∂x∗i (p, y)∂y

|y=e(p,u).

Page 450: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

Chapter 13

Differentiable Comparative Statics,Part 2

13.1 Qualitative Comparative Statics Analysis

The method explained in this chapter was devised by Eugene Silberberg in a paperentitled A Revision of Comparative Statics Methodology in Economics, or, How toDo Comparative Statics on the Back of an Envelope (see Silberberg 1974). Once youhave read and understood this chapter, you will realize that the title of Silberberg’spaper is an amusing double entendre.

The goal is to discover only the signs of the differential comparative statics quan-tities ∂x∗i /∂γk and ∂λ∗j(γ)/∂γk.

Silberberg’s method relies upon a comparison of two types of constrained op-timization problems: primal and primal-dual problems. Primal problems are theproblems we have talked about so far in this book. Primal-dual problems are newand are built in a rather clever way from families of primal problems. What thismeans is explained in the next section. Once we have understood what primal-dualproblems are, we will examine their optimal solutions. It turns out that the first-orderoptimality conditions to a primal-dual problem give us much of what we have alreadydiscussed. Particularly, we are given both the first-order optimality conditions forprimal problems and the results of applying the First-Order Envelope Theorem toprimal problems. Not only that, but this is all done in a very easy manner. Ofcourse we already have these results, even if we will now be able to obtain them ina different way. So what do we get from Silberberg’s approach that is new? Well,we use something old to get something new. The old thing is the necessary second-

431

Page 451: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

432 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

order optimality condition, but for the primal-dual problem. Rather amazingly, thiscondition gives us every qualitative comparative statics result there is for the primalproblems used to construct the primal-dual problem, and it does it in a relativelyeasy way.

So here is the plan. We start in the next section with an explanation of a primal-dual problem. In the following section, we will examine the first-order optimalityconditions for a primal-dual problem. The section after that gets to the heart of thematter by examining the necessary second-order optimality condition for the primal-dual problem.

There is more. Many times in economics you see statements such as, “When theprice of a firm’s product changes the firm adjusts the quantity that it supplies of itsproduct by more in the long-run than in a short-run.” Translated into more standardEnglish this statement is, “When the price of a firm’s product changes the firmadjusts the quantity that it supplies of its product by more when the firm is subjectto fewer constraints upon its choices of production plans.” Introductory textbooksoften paraphrase this with statements like, “the slope of the firm’s supply curve isflatter in the long-run than in a short-run” (remember, economists do this weird thingof plotting supply curves with the product price p measured on the vertical axis).Statements like these are about comparisons of comparative statics results. If theslopes of the firm’s long-run and short-run supply functions are ∂y�/∂p and ∂ys/∂p,then the assertion is that ∂ys/∂p < ∂y�/∂p. Such comparisons of comparative staticsresults nowadays are often called Le Chatelier phenomena, after the person who firstbrought them to the attention of scientists. In the last part of this chapter, we will seehow to obtain them by using a different type of envelope theorem, the Second-OrderEnvelope Theorem. Don’t worry. It’s quite a simple and obvious result.

13.2 Primal and Primal-Dual Problems

Consider again the primal constrained optimization problem

maxx1, ... , xn

f(x1, . . . , xn;α) subject to hj(x1, . . . , xn; βj, bj) ≥ 0 for j = 1, . . . ,m. (13.1)

For the purposes of this chapter, the crucial feature of a primal problem is that ithas choice variables x1, . . . , xn, and fixed quantities, the parameters γ1, . . . , γr. Let usdenote the number of parameters for this problem by r, so γ = (γ1, . . . , γr) and theproblem’s parameter space Γ ⊂ �r. Suppose that all of the functions f, h1, . . . , hm

Page 452: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.2. PRIMAL AND PRIMAL-DUAL PROBLEMS 433

are twice-continuously differentiable and that the solution correspondence is a func-tion; i.e. for any γ ∈ Γ, the optimal solution (x∗(γ), λ∗(γ)) to the primal problemis unique. The maximum-value function for the primal problem is the problem’sobjective function evaluated at x∗(γ). This is the function v : Γ �→ �, defined by

v(γ) ≡ f(x∗(γ);α). (13.2)

From the primal problem, Silberberg constructs another constrained optimizationproblem that he calls primal-dual. It is

maxx∈�n, γ∈Γ

θ(γ, x) ≡ f(x, α)− v(γ) (13.3)

subject to hj(x, βj, bj) ≥ 0 for j = 1, . . . ,m.

The semicolons in f and in each of h1, . . . , hm for the primal problem (13.1) have,in the primal-dual problem, been replaced with commas, indicating that in theprimal-dual problem, all of the r quantities in the vectors α, β and b are variables,not fixed parameters. The primal-dual problem thus has n + r choice variables;x1, . . . , xn, γ1, . . . , γr.

It appears at a casual glance that the primal and primal-dual problems havethe same constraints, but this is not so. Yes, the equations of the constraints are thesame for both problems, but even so, the constraints for the primal problem are “moreconstraining” than are the constraints for the primal-dual problem. Why? Supposethat the primal problem has just one constraint, say, h(x1, x2; β, b) = b−x1−βx2 ≥ 0(notice the semi-colon in h), where β > 0 and b ≥ 0. The set of feasible choices(x1, x2) for this primal problem is Cp(β, b) = {(x1, x2) | x1 + βx2 ≤ b} for particularfixed and given values for β and b, say, β = 2 and b = 3. The primal problem’sconstraint is thus a restriction on choices of only x1 and x2 for given particularvalues of β and b. Cp(2, 3) = {(x1, x2) | x1 + 2x2 ≤ 3} is a particular example of aprimal problem’s feasible set. The primal-dual problem’s version of this constraint ish(x1, x2, β, b) = b − x1 − βx2 ≥ 0 (notice that the semi-colon in h is replaced by acomma), meaning that the constraint is jointly on choices of all of x1, x2, β, and b.The primal-dual’s feasible set is thus

Cpd = {(x1, x2, β, b) | x1 + βx2 ≤ b, β > 0, b ≥ 0}.The complete set of values of x1 and x2 that, for one or more values of β and b, arefeasible for the primal-dual problem, is therefore⋃

β>0b≥0

Cp(β, b).

Page 453: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

434 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

This typically is a much larger set than a feasible set, such as Cp(2, 3), for the primalproblem with particular and fixed values for β and b.

Another interesting feature of the primal-dual problem is that it has no parame-ters. This may seem odd since we are so used to dealing with problems in which somequantities have fixed values (i.e. are parameters). But a moment’s thought shouldmake this idea appear less strange. Consider, for example, the primal problem

minx

(x− b)2. (13.4)

This problem has one choice variable, x, and one parameter, b. For any given valueof b, the unique minimizing solution is x∗(b) = b. There are many primal problems,each defined by a different given value of b, and with one optimal solution, x∗(b) = b.Now consider the primal-dual problem

minx, b

(x− b)2,

in which both x and b are choice variables. There are no parameters for this problemand, consequently, there is just one primal-dual problem. This one problem possessesan infinity of optimal solutions (x∗, b∗), in each of which x∗ = b∗. Notice also thatthis one primal-dual problem corresponds to infinitely many primal problems (13.4),one for each possible value of the (primal) parameter b. This is true in general. Afixed value for the parameter vector γ in problem (13.1) defines a particular primalproblem. There is one such primal problem for each possible value of γ. In theprimal-dual problem (13.3) the vector γ is allowed to vary (it’s not parametric) overall values admissible for it. Hence there is just one primal-dual problem.

This is a good moment for you to pause before reading further. It is very easy tomisunderstand what a primal-dual problem really is and how it is created from anentire family of primal problems. So before you plunge into reading about solvinga primal-dual problem, I recommend that you review this section to make sure youunderstand both the similarities and the differences between any family of primalproblems and the primal-dual problem created from that family.

13.3 Primal-Dual First-Order Conditions

For any particular γ ∈ Γ (thus, for any particular primal problem), v(γ) ≥ f(x; γ)for every x ∈ C(β, b), the primal problem’s feasible set. Therefore the primal-dualobjective function θ(γ, x) never takes a positive value. What is the greatest value

Page 454: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.3. PRIMAL-DUAL FIRST-ORDER CONDITIONS 435

that the primal-dual objective function can take? Think about it before proceedingto the next paragraph.

The answer is zero. Why? Well, what does an optimal solution to the primal-dualproblem look like? Take any value allowed in the primal-dual problem for the choicevector γ; i.e. take any γ ∈ Γ. Then solve the primal problem using γ as a fixedparameter vector. Now we have the point in �n+m+r that is (x∗(γ), λ∗(γ), γ). Thispoint is an optimal solution to the primal-dual problem because, at this point, thevalue of the primal-dual problem’s objective function is

θ(γ, x∗(γ)) = f(x∗(γ);α)− v(γ) = f(x∗(γ);α)− f(x∗(γ);α) = 0.

So, how many optimal solutions does the primal-dual problem have? Think about itbefore reading further.

There is one optimal solution (x∗(γ), λ∗(γ), γ) for every γ ∈ Γ. Since we haveassumed that the functions in the primal-dual problem are differentiable with respectto γ there are infinitely many values in Γ and, for each one, an optimal solution tothe primal-dual problem.

The necessary first-order optimization conditions must hold at any optimal solu-tion to the primal-dual problem. From (13.3), the Karush-Kuhn-Tucker componentof the first-order optimality conditions is that, at a primal-dual optimal solution(x∗(γ), λ∗(γ), γ), necessarily the scalars λ∗1(γ), . . . , λ

∗m(γ) are all nonnegative, not all

zero, and such that the gradient vector of the objective function θ is

∇γ,xθ(γ, x)|x=x∗(γ) = −m∑j=1

λ∗j(γ)∇γ,xhj(x, βj, bj)∣∣x=x∗(γ). (13.5)

∇γ,x denotes the gradient vector in which the gradients are measured in each one ofthe directions of the primal-dual problem’s variables γ1, . . . , γr and x1, . . . , xn, and inthat order, so the gradient vector of θ is

∇γ,xθ(γ, x) =

(∂θ(γ, x)

∂γ1, . . . ,

∂θ(γ, x)

∂γr,∂θ(γ, x)

∂x1, . . . ,

∂θ(γ, x)

∂xn

)

=

(∂f(x, α)

∂γ1− ∂v(γ)

∂γ1, . . . ,

∂f(x, α)

∂γr− ∂v(γ)

∂γr,∂f(x, α)

∂x1, . . . ,

∂f(x, α)

∂xn

)= (∇γθ(γ, x),∇xθ(γ, x)),

where the gradient of θ(γ, x) = f(x, α)− v(γ) in the directions of γ1, . . . , γr only is

∇γθ(γ, x) =

(∂f(x, α)

∂γ1− ∂v(γ)

∂γ1, . . . ,

∂f(x, α)

∂γr− ∂v(γ)

∂γr

)

Page 455: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

436 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

and the gradient of θ(γ, x) = f(x, α)− v(γ) in the directions of x1, . . . , xn only is

∇xθ(γ, x) =

(∂f(x, α)

∂x1, . . . ,

∂f(x, α)

∂xn

).

The gradient vectors of h1, . . . , hm similarly are r + n dimensional row vectors.Expanded, the Karush-Kuhn-Tucker condition (13.5) for the primal-dual problem is(13.6) and (13.7). Then we add the other necessary first-order conditions that arethe m complementary slackness conditions (13.8), so, in all, the first-order optimalityconditions for the primal-dual problem consist of

∂θ(x, γ)

∂xi|x=x∗(γ) =

∂f(x, α)

∂xi|x=x∗(γ) = −λ∗1(γ)

∂h1(x, β1, b1)

∂xi|x=x∗(γ) − · · ·

− λ∗m(γ)∂hm(x, βm, bm)

∂xi|x=x∗(γ) ∀ i = 1, . . . , n, (13.6)

∂θ(x, γ)

∂γk|x=x∗(γ) =

∂f(x, α)

∂γk|x=x∗(γ) −

∂v(γ)

∂γk= −λ∗1(γ)

∂h1(x, β1, b1)

∂γk|x=x∗(γ) − · · ·

− λ∗m(γ)∂hm(x, βm, bm)

∂γk|x=x∗(γ) ∀ k = 1, . . . , r, (13.7)

0 = λ∗j(γ)hj(x∗(γ), βj, bj) ∀ j = 1, . . . ,m, (13.8)

and hj(x∗(γ), βj, bj) ≥ 0 ∀ j = 1, . . . ,m. (13.9)

The Karush-Kuhn-Tucker and complementary slackness conditions for x and λ forthe primal-dual problem, (13.6) and (13.8), are exactly the same as the Karush-Kuhn-Tucker and complementary slackness conditions for x and λ for the primal problem fora given value of γ. If you think about it for a moment, you will realize that this mustbe so. After all, for any given parameter vector γ ∈ Γ, the primal problem’s optimalsolution (x∗(γ), λ∗(γ)) must also be the optimal solution to the primal-dual problemfor the same value of γ, and thus must be a solution to the primal-dual problem’snecessary first-order conditions for x and λ for that given value of γ. Perhaps thepoint will be clearer if you realize that, for a fixed value, say, γ′, of γ, the maximumvalue v(γ′) is a fixed number and the primal-dual problem is, for γ fixed at γ′,

maxx

f(x;α′)− v(γ′) subject to hj(x; β′j, b

′j) ≥ 0 for j = 1, . . . ,m.

This is exactly the primal problem parameterized by γ = γ′, other than that theconstant v(γ′) is subtracted from the objective function, so the first-order necessaryconditions with respect to x for this primal problem have to be the same as the

Page 456: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.4. THE PRIMAL-DUAL SECOND-ORDER CONDITION 437

first-order necessary conditions with respect to x for the primal-dual problem withγ = γ′.

Did you notice that (13.7) is a restatement of the First-Order Envelope Theorem?Why is the theorem a necessary implication of any optimal solution to the primal-dualproblem? Pick some value of γ, say, γ′. What (13.7) says is that, if we hold the valueof x fixed at x∗(γ′) then, when γ = γ′, an optimal solution (x∗(γ′), λ∗(γ′), γ′) to theprimal-dual problem has been reached, so v(γ′) = f(x∗(γ′); γ′); the primal problem’sobjective function’s maximized value when γ = γ′ is the same as the value at γ = γ′

of the envelope v(γ). Consequently the slopes with respect to γk of the envelope v(γ)and of the restricted function f r(γ; x= x∗(γ′)) must be the same at γk = γ′k. This,the statement of the First-Order Envelope Theorem, is the part of the primal-dualproblem’s Karush-Kuhn-Tucker condition for the γk choice variable. There is no suchcondition in the primal problem because therein the vector γ is not variable.

13.4 The Primal-Dual Second-Order Condition

Silberberg’s method concentrates upon the primal-dual problem’s necessary second-order maximization condition because it is the source of the qualitative comparativestatics results for the primal problems. The condition is that the Hessian matrixof the Lagrange function for the primal-dual problem must, when evaluated at aprimal-dual optimal solution (x∗(γ′), λ∗(γ′), γ′), be negative-semi-definite for all smallperturbations (Δx,Δγ) from (x∗(γ′), γ′) that keep binding all of the constraints thatbind at (x∗(γ′), γ′).

Let’s be clear about the meaning of a “small” perturbation (Δx,Δγ) from a point(x∗(γ′), γ′). The Δx part of the perturbation (Δx,Δγ) means that we consider points(x∗1(γ

′)+Δx1, . . . , x∗n(γ

′)+Δxn) that are “close” to the point (x∗1(γ

′), . . . , x∗n(γ′)). Let’s

say that for the objective function the list of parameters is (α1, . . . , αa) and that forthe jth constraint function the list of parameters is βj1, . . . , βjnj

, for j = 1, . . . ,m.Then the Δγ part of the perturbation (Δx,Δγ) means that we consider points

(α′1 +Δα1, . . . , α

′a +Δαa, β

′11 +Δβ11, . . . , β

′1n1

+Δβ1n1 , β′21 +Δβ21, . . . , β

′2n2

+Δβn2 ,

. . . , β′m1 +Δβm1, . . . , β

′mnm

+Δβmnm , b′1 +Δb1, . . . , b

′m +Δbm)

that are “close” to γ′ = (α′1, . . . , α

′a, β

′11, . . . , β

′1n1, . . . , β′

m1, . . . , β′mnm

, b′1, . . . , b′m).

The Lagrange function for the primal-dual problem is

Lpd(x, γ, λ) = f(x, α)− v(γ) + λ1h1(x, β1, b1) + · · ·+ λmhm(x, βm, bm). (13.10)

Page 457: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

438 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

Let Δ(βj, bj) denote a perturbation of the vector (βj, bj) of variables other thanx1, . . . , xn in the jth constraint. Then the primal-dual problem’s second-order neces-sary maximization condition is that, there is an ε > 0 for which

(Δx,Δγ)HLpd(x∗(γ′), λ∗(γ′), γ′)(ΔxΔγ

)≤ 0 ∀ (Δx,Δγ) such that (13.11)

(x∗(γ′) + Δx, γ′ +Δγ) ∈ BdE ((x∗(γ′), γ′) , ε) , and (13.12)(∇xhj(x, βj, bj),∇(βj ,bj)hj(x, βj, bj)

)∣∣(x,γ)=(x∗(γ′),γ′)

·(

ΔxΔ(βj, bj)

)= 0 ∀ j ∈ I(x∗(γ′), γ′),

(13.13)

where I(x∗(γ′), γ′) is the set of indices of the constraints that bind at (x, γ) =(x∗(γ′), γ′), and where the Lagrange function’s Hessian matrix with respect to theprimal-dual problem’s choice variables x and γ is the symmetric order-(n+ r) matrix

HLpd(x, λ, γ) =

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

∂2Lpd

∂x21· · · ∂2Lpd

∂x1∂xn

∂2Lpd

∂x1∂γ1· · · ∂2Lpd

∂x1∂γr...

. . ....

......

∂2Lpd

∂xn∂x1· · · ∂2Lpd

∂x2n

∂2Lpd

∂xn∂γ1· · · ∂2Lpd

∂xn∂γr

∂2Lpd

∂γ1∂x1· · · ∂2Lpd

∂γ1∂xn

∂2Lpd

∂γ21· · · ∂2Lpd

∂γ1∂γr...

......

. . ....

∂2Lpd

∂γr∂x1· · · ∂2Lpd

∂γr∂xn

∂2Lpd

∂γr∂γ1· · · ∂2Lpd

∂γ2r

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠

. (13.14)

We can write this sometimes large matrix more briefly in its block form, which is

HLpd(x, λ, γ) =

⎛⎝∇2

xxLpd ∇2

xγLpd

∇2γxL

pd ∇2γγL

pd

⎞⎠ . (13.15)

∇2xxL

pd denotes the n × n upper-left block of terms ∂2Lpd/∂xi∂xj for i, j = 1, . . . , nin the Hessian. ∇2

xγLpd denotes the n × r upper-right block of terms ∂2Lpd/∂xi∂γk

for i = 1, . . . , n and k = 1, . . . , r. ∇2γxL

pd denotes the r × n lower-left block of terms∂2Lpd/∂γk∂xi for k = 1, . . . , r and i = 1, . . . , n. ∇2

γγLpd denotes the r× r lower-right

block of terms ∂2Lpd/∂γj∂γk for j, k = 1, . . . , r.

Page 458: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.4. THE PRIMAL-DUAL SECOND-ORDER CONDITION 439

The necessary second-order maximization condition, then, is that the matrix(13.15), evaluated at (x, λ, γ) = (x∗(γ′), λ∗(γ′), γ′), must be negative-semi-definitefor all small perturbations (Δx,Δγ) from (x∗(γ′), γ′) that keep binding all of theconstraints that bind at (x∗(γ′), γ′). That is, necessarily,

(Δx,Δγ)

⎛⎝∇2

xxLpd ∇2

xγLpd

∇2γxL

pd ∇2γγL

pd

⎞⎠∣∣∣∣(x,λ,γ)=

(x∗(γ′),λ∗(γ′),γ′)

(ΔxΔγ

)≤ 0 (13.16)

for all small perturbations (Δx,Δγ) from (x∗(γ′), γ′) such that

(∇xhj(x, βj, bj),∇(βj ,bj)hj(x, βj, bj)

)∣∣(x,γ)=(x∗(γ′),γ′)

(Δx

Δ(βj, bj)

)= 0 ∀ j ∈ I(x∗(γ′), γ′).

(13.17)Silberberg then remarks that, in a comparative statics experiment, the only pertur-bations made are to the primal problem’s parameters, listed in the vector γ. Thatis, for any comparative statics experiment, we set the “x-part” of the perturbationvector to zero. Then, from (13.16) and (13.17) with Δx = 0n, it is necessarily truefor comparative statics experiments that

ΔγT(∇2

γγLpd)∣∣

(x,λ,γ)=(x∗(γ′),λ∗(γ′),γ′)

Δγ ≤ 0 (13.18)

for all small perturbations Δγ from γ′ such that

∇(βj ,bj)hj(x, βj, bj)∣∣(x,γ)=(x∗(γ′),γ′)

·Δ(βj, bj) = 0 ∀ j ∈ I(x∗(γ′), γ′). (13.19)

The statement that is (13.18) and (13.19) combined is necessarily true as a conse-quence of the point (x∗(γ′), λ∗(γ′), γ′) being an optimal solution to the primal-dualproblem (13.3). It is this statement that provides us with the qualitative comparativestatics properties of the optimal solution to the primal problem (13.1). Let’s see howto extract these results.

Page 459: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

440 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

13.5 Qualitative Comparative Statics Results

Have a close look at the matrix ∇2γγL

pd in (13.18). It is

∇2γγL

pd =

⎛⎜⎜⎜⎜⎜⎝∂2Lpd

∂γ21· · · ∂2Lpd

∂γ1∂γr...

. . ....

∂2Lpd

∂γr∂γ1· · · ∂2Lpd

∂γ2r

⎞⎟⎟⎟⎟⎟⎠∣∣∣∣∣(x,λ,γ)=(x∗(γ′),λ∗(γ′),γ′)

. (13.20)

What does a typical element of this matrix look like? From (13.10),

∂Lpd

∂γk=∂f(x, α)

∂γk− ∂v(γ)

∂γk+ λ1

∂h1(x, β1, b1)

∂γk+ · · ·+ λm

∂hm(x, βm, bm)

∂γm, (13.21)

so

∂2Lpd

∂γj∂γk=∂2f(x, α)

∂γj∂γk− ∂2v(γ)

∂γj∂γk+λ1

∂2h1(x, β1, b1)

∂γj∂γk+· · ·+λm

∂2hm(x, βm, bm)

∂γj∂γk. (13.22)

The jkth-element of the matrix (13.20) is the value of (13.22) at the primal-dualoptimal solution (x∗(γ′), λ∗(γ′), γ′). This is

∂2Lpd

∂γj∂γk

∣∣x=x∗(γ′)λ=λ∗(γ′)γ=γ′

=∂2f(x, α)

∂γj∂γk

∣∣x=x∗(γ′)α=α′

− ∂2v(γ)

∂γj∂γk

∣∣γ=γ′ + λ∗1(γ

′)∂2h1(x, β1, b1)

∂γj∂γk

∣∣x=x∗(γ′)(β1,b1)=(β′

1,b′1)

+ · · ·+ λ∗m(γ′)∂2hm(x, βm, bm)

∂γj∂γk

∣∣x=x∗(γ′)(βm,bm)=(β′

m,b′m)

. (13.23)

Now recall that, when evaluated at only primal-dual optimal solutions, the first-ordernecessary conditions (13.6), (13.7), and (13.8) are identically zero, so, from (13.7),

∂v(γ)

∂γk≡ ∂f(x, α)

∂γk|x=x∗(γ)

+λ∗1(γ)∂h1(x, β1, b1)

∂γk|x=x∗(γ)

+· · ·+λ∗m(γ)∂hm(x, βm, bm)

∂γk|x=x∗(γ).

Differentiating this identity with respect to γj gives

∂2v(γ)

∂γj∂γk=

n∑i=1

∂2f(x, α)

∂xi∂γk

∣∣x=x∗(γ) ×

∂x∗i (γ)∂γj

+∂2f(x, α)

∂γj∂γk

∣∣x=x∗(γ) +

m∑t=1

∂λ∗t (γ)∂γj

× ∂ht(x, βt, bt)

∂γk

∣∣x=x∗(γ) + (13.24)

m∑t=1

λ∗t (γ)

[∂2ht(x, βt, bt)

∂γj∂γk

∣∣x=x∗(γ) +

n∑i=1

∂2ht(x, βt, bt)

∂xi∂γk

∣∣x=x∗(γ′)(βt,bt)=(β′

t,b′t)

× ∂x∗i (γ)∂γj

∣∣γ=γ′

].

Page 460: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.5. QUALITATIVE COMPARATIVE STATICS RESULTS 441

Evaluating (13.24) at γ = γ′ and then using (13.24) to substitute for ∂2v(γ)/∂γj∂γkin (13.23) shows that value of the jkth-element of the matrix (13.20) is

∂2Lpd

∂γj∂γk

∣∣x=x∗(γ′)λ=λ∗(γ′)γ=γ′

= −n∑

i=1

[∂2f(x, α)

∂xi∂γk+

m∑t=1

λ∗t (γ′)∂2ht(x, βt, bt)

∂xi∂γk

] ∣∣∣∣x=x∗(γ′)γ=γ′

× ∂x∗i (γ)∂γj

−m∑t=1

∂λ∗t (γ)∂γj

× ∂ht(x, βt, bt)

∂γk

∣∣∣∣x=x∗(γ′)γ=γ′

. (13.25)

(13.25) states that every element in (13.20) is a linear combination of the values at x =x∗(γ′) and γ = γ′ of the comparative statics quantities ∂x∗1(γ)/∂γj, . . . , ∂x

∗n(γ)/∂γj

and ∂λ∗1(γ)/∂γj, . . . , ∂λ∗m(γ)/∂γj. Information about the signs of the elements in

the matrix (13.20) therefore gives us information about the signs of the comparativestatics quantities. What information is available to us?

First, a matrix is negative-semi-definite only if its main diagonal elements are allnonpositive;

∂2Lpd

∂γ2k≤ 0 for all k = 1, . . . , r. (13.26)

Second, the matrix is symmetric (because all of the functions f, g1, . . . , gm aretwice-continuously differentiable);

∂2Lpd

∂γj∂γk=

∂2Lpd

∂γk∂γjfor all j, k = 1, . . . , r; j = k. (13.27)

If the constraints that bind at the particular optimal solution (x∗(γ), λ∗(γ), γ) tothe primal-dual problem are affine with respect to γ1, . . . , γr (not necessarily affinewith respect to x1, . . . , xn), then we have a third way of extracting comparative stat-ics results. If the binding constraints are all affine with respect to γ1, . . . , γr, thenthe combination of (13.18) and (13.19) is true if and only if the bordered principalminors of the bordered Hessian constructed from the Hessian (13.20) and the gradi-ents, with respect to γ1, . . . , γr, of the constraints that bind at the optimal solution(x∗(γ), λ∗(γ), γ) have the properties described in Theorem 10.2.

Let’s examine this bordered Hessian. Without loss of generality, suppose thatthe constraints that bind at (x∗(γ′), λ∗(γ′), γ′) are indexed by j = 1, . . . , �. Thenthe bordered Hessian, evaluated at (x∗(γ′), λ∗(γ′), γ′), is the symmetric order-(�+ r)

Page 461: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

442 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

matrix

HLpd(x∗(γ′), λ∗(γ′), γ′) =

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

0 · · · 0∂h1∂γ1

· · · ∂h1∂γr

.... . .

......

...

0 · · · 0∂h�∂γ1

· · · ∂h�∂γr

∂h1∂γ1

· · · ∂h�∂γ1

∂2Lpd

∂γ21· · · ∂2Lpd

∂γ1∂γr...

......

. . ....

∂h1∂γr

· · · ∂h�∂γr

∂2Lpd

∂γr∂γ1· · · ∂2Lpd

∂γ2r

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠

∣∣∣∣∣(x,λ,γ)=(x∗(γ),λ∗(γ),γ)

.

(13.28)Theorem 10.2 states that, if all of the constraints j = 1, . . . , � are affine with respectto γ1, . . . , γr, then (13.18) and (13.19) together are equivalent to the requirementthat the bordered principal minors of the bordered Hessian (13.28) that are of ordersi = 2� + 1, . . . , r are either zero or have the same sign as (−1)i. This informationprovides still more comparative statics results.

13.6 An Example

Let’s apply Silberberg’s method of obtaining qualitative comparative-statics resultsto our primal problem example (11.4) of a price-taking, profit-maximizing firm. Themaximum-value function for the problem is the firm’s indirect profit function

π∗(p, w1, w2, β1, β2)

= py∗(p, w1, w2, β1, β2)− w1x∗1(p, w1, w2, β1, β2)− w2x

∗2(p, w1, w2, β1, β2),

so the primal-dual problem’s objective function is

θ(y, x1, x2, p, w1, w2, β1, β2) = py − w1x1 − w2x2 − π∗(p, w1, w2, β1, β2)

and the primal-dual problem itself is

maxp, w1, w2, β1, β2, y, x1, x2

θ(p, w1, w2, β1, β2, y, x1, x2) = py − w1x1 − w2x2 − π∗(p, w1, w2, β1, β2)

subject to y ≥ 0, x1 ≥ 0, x2 ≥ 0, and xβ1

1 xβ2

2 ≥ y. (13.29)

Page 462: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.6. AN EXAMPLE 443

The primal-dual problem’s Lagrange function is

Lpd(y, x1, x2, p, w1, w2, β1, β2, λ1, λ2, λ3, λ4)

= py − w1x1 − w2x2 − π∗(p, w1, w2, β1, β2) + λ1y + λ2x1 + λ3x2 + λ4(xβ1

1 xβ2

2 − y).

All of y, x1, x2, p, w1, w2, β1, β2, λ1, λ2, λ3, and λ4 are variables. The first-orderderivatives of Lpd with respect to the primal-dual variables p, w1, w2, β1, and β2 thatare parameters in the primal problem are

∂Lpd

∂p= y − ∂π∗(p, ·)

∂p,

∂Lpd

∂wi

= − xi −∂π∗(·, wi, ·)

∂wi

for i = 1, 2, (13.30)

and∂Lpd

∂βi= − ∂π∗(·, βi)

∂βi+ λ4

∂(xβ1

1 xβ2

2

)∂βi

for i = 1, 2. (13.31)

Each of these necessarily is zero when evaluated at any optimal solution to the primal-dual problem. (13.30) therefore gives us Hotelling’s Lemma:

y∗(p, w1, w2, β1, β2) =∂π∗(p, ·)∂p

and x∗i (p, w1, w2, β1, β2) = − ∂π∗(·, w1, ·)∂wi

for i = 1, 2.

(13.31) gives us the new result

∂π∗(·, βi)∂βi

= λ∗4(γ)∂(xβ1

1 xβ2

2

)∂βi

∣∣x=x∗(γ) = p

∂(xβ1

1 xβ2

2

)∂βi

∣∣x=x∗(γ) for i = 1, 2. (13.32)

An increase to βi increases the productivity of input i. (13.32) states that the rate atwhich the firm’s maximum profit increases as βi increases is the rate at which extrarevenue is provided due to higher production caused by the enhanced productivity ofinput i. (13.32) is thus the firm’s marginal valuation of the technology parameter βi.

In what follows I’m going to hold β2 fixed so we have to work with only 4 × 4matrices. Using (13.25) shows that the Hessian matrix of Lpd with respect to the

Page 463: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

444 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

primal-dual problem choice variables p, w1, w2, and β1 is

(∇2

γγLpd)=

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

−∂y∗

∂p

∂x∗1∂p

∂x∗2∂p

−p∑2i=1

(gxi β1×

∂x∗i∂p

)− gβ1

− ∂y∗

∂w1

∂x∗1∂w1

∂x∗2∂w1

−p∑2i=1

(gxi β1×

∂x∗i∂w1

)

− ∂y∗

∂w2

∂x∗1∂w2

∂x∗2∂w2

−p∑2i=1

(gxi β1×

∂x∗i∂w2

)

−∂y∗

∂β1

∂x∗1∂β1

∂x∗2∂β1

−p∑2i=1

(gxi β1×

∂x∗i∂β1

)

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠

, (13.33)

where gβ1 = ∂(xβ1

1 xβ2

2

)/∂β1 = xβ1

1 xβ2

2 ln x1, gx1 β1 = ∂2(xβ1

1 xβ2

2

)/∂x1∂β1 =

xβ1−11 xβ2

2 (β1 ln x1 + 1), and gx2 β1 = ∂2(xβ1

1 xβ2

2

)/∂x2∂β1 = xβ1

1 xβ2−12 (β2 ln x2 + 1).

(13.33) is a negative-semi-definite matrix only if each of its main diagonal elementsis nonpositive, so

∂y∗(p, w1, w2, β1, β2)

∂p≥ 0,

∂x∗i (p, w1, w2, β1, β2)

∂wi

≤ 0 for i = 1, 2, (13.34)

and p

⎡⎣∂2

(xβ1

1 xβ2

2

)∂x1∂β1

× ∂x∗1(·)∂β1

+∂2(xβ1

1 xβ2

2

)∂x2∂β1

× ∂x∗2(·)∂β1

⎤⎦ ∣∣∣∣∣x1=x∗

1x2=x∗

2

≥ 0. (13.35)

(13.34) states the well-known results that the quantity supplied by the firm is anondecreasing function of the price of the firm’s product and that the firm’s quantitydemanded of an input is a nonincreasing function of the price of that input. (13.35) isnot a well-known result. It says that the marginal value of the technology parameterβ1 is nondecreasing with respect to β1; i.e. the value to the firm of its technologyparameter β1 is a convex function of β1.

Page 464: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.6. AN EXAMPLE 445

The Hessian matrix is symmetric, so we learn also that

∂y∗(p, w1, w2, β1, β2)

∂wi

= − ∂x∗i (p, w1, w2, β1, β2)

∂pfor i = 1, 2, (13.36)

∂x∗1(p, w1, w2, β1, β2)

∂w2

=∂x∗2(p, w1, w2, β1, β2)

∂w1

, and (13.37)

∂y∗(p, w1, w2, β1, β2)

∂β1= p

2∑i=1

⎛⎝∂2

(xβ1

1 xβ2

2

)∂xi∂β1

∣∣x1=x∗

1x2=x∗

2

× ∂x∗i∂p

⎞⎠+

∂(xβ1

1 xβ2

2

)∂β1

∣∣x1=x∗

1x2=x∗

2

.

(13.38)

(13.36) states that the rate at which the firm’s quantity supplied decreases as theprice of input i rises necessarily is the same as the rate at which the firm’s quantitydemanded of input i rises as the price of the firm’s product rises. Did you knowthat? (13.37) states that the rate at which the firm’s quantity demanded of input iincreases as the price of input j rises is necessarily the same as the rate at which thefirm’s quantity demanded of input j rises as the price of input i rises. Understanding(13.38) requires a bit more work. The quantity of product supplied by the firm is

y∗(p, w1, w2, β1; β2) = x∗1(p, w1, w2, β1; β2)β1x∗2(p, w1, w2, β1; β2)

β2 ,

so

∂y∗(·)∂β1

=∂(xβ1

1 xβ2

2

)∂x1

∣∣x1=x∗

1(·)x2=x∗

2(·)× ∂x∗1(·)

∂β1+∂(xβ1

1 xβ2

2

)∂x2

∣∣x1=x∗

1(·)x2=x∗

2(·)× ∂x∗2(·)

∂β1

+∂(xβ1

1 xβ2

2

)∂β1

∣∣x1=x∗

1(·)x2=x∗

2(·).

(13.39)

The sum of the first two terms on the right side of (13.39) is the indirect effectof changing β1, through changes caused to the profit-maximizing quantities used ofinputs 1 and 2, on the firm’s quantity supplied. The last term in the right side of(13.39) is the direct effect on quantity supplied of changing β1. Setting the right sidesof (13.38) and (13.39) equal to each other gives us that

∂(xβ1

1 xβ2

2

)∂x1

∣∣x1=x∗

1(·)x2=x∗

2(·)× ∂x∗1(·)

∂β1+∂(xβ1

1 xβ2

2

)∂x2

∣∣x1=x∗

1(·)x2=x∗

2(·)× ∂x∗2(·)

∂β1

= p∂2(xβ1

1 xβ2

2

)∂x1∂β1

∣∣x1=x∗

1(·)x2=x∗

2(·)× ∂x∗1(·)

∂p+ p

∂2(xβ1

1 xβ2

2

)∂x2∂β1

∣∣x1=x∗

1(·)x2=x∗

2(·)× ∂x∗2(·)

∂p.

(13.40)

Page 465: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

446 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

The right side of (13.40) is the sum of the indirect effects of changing the productprice p, through changes caused to the profit-maximizing quantities used of inputs 1and 2, on the firm’s marginal revenue product of the technology variable β1. This,necessarily, is the same as the indirect effect of changing β1 on the firm’s quantitysupplied. You may question if this is a useful result. Whether or not this is sodepends upon the questions you want answered. The first point you should take fromthis example is that it demonstrates how Silberberg’s method allows for the extractionof comparative statics results that are not obvious. The second point is that (13.40)is a simple linear equation in the unknowns ∂x∗1/∂β1, ∂x

∗2/∂β1, ∂x

∗1/∂p and ∂x∗2/∂p.

Why? Because all of the derivatives of xβ1

1 xβ2

2 in (13.40) are evaluated at (x∗1, x∗2); i.e.

they are just numbers. (13.40) is a linear restriction that gives us information aboutthe signs possible for the four comparative statics quantities.

Can we use Theorem 10.2 to extract further comparative-statics results? Lookonce again at the primal-dual problem (13.29). What constraints bind at an optimalsolution to this problem? For any interesting solution we will have y > 0, x1 > 0, andx2 > 0, so the only binding constraint is the technology constraint: xβ1

1 xβ2

2 − y = 0.This depends on none of the parameters p, w1, and w2. And, although the constraintdoes depend upon the parameters β1 and β2, these dependencies are not affine. Sowe cannot apply Theorem 10.2 to this particular problem.

13.7 Second-Order Envelope Theorem

The First-Order Envelope Theorem tells us about maximum-value functions’ first-order derivatives. The Second-Order Envelope Theorem informs us about a com-parison of the second-order derivatives of two related maximum-value functions. Todiscuss this theorem with clarity, we must once again adjust the notation we use.

The primal constrained optimization problem that we are considering is

maxx∈�n

f(x;α) subject to hj(x; βj, bj) ≥ 0 for j = 1, . . . ,m. (13.41)

Denote the parameter vector for this m-constraint problem by mγ = (α,mβ,mb).Now choose a particular value for mγ. This selects the particular version of (13.41)that is defined by that particular parameter vector mγ. The optimal solution forthis particular problem will be denoted by (mx∗(mγ),mλ∗(mγ)). Use I(mx∗(mγ)) todenote the set of indices of the constraints that bind at x = mx∗(mγ). Without loss ofgenerality, suppose that the indices in I(mx∗(mγ)) are j = 1, . . . , �, where 1 ≤ � ≤ m.Then, given mγ, the optimal solution to (13.41) is the same as the optimal solution

Page 466: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.7. THE SECOND-ORDER ENVELOPE THEOREM 447

to the equality-constrained primal problem

maxx∈�n

f(x;α) subject to h1(x; β1, b1) = 0, . . . , h�(x; β�, b�) = 0. (13.42)

(13.42) is the �-constraint problem with the parameter vector �γ = (α, �β, �b). Denotethe set of all possible values of �γ by �Γ. Denote the optimal solution to the problem,for given �γ, by (�x

∗(�γ), �λ

∗(�γ)), where �x

∗(�γ) = (�x

∗1(

�γ), . . . , �x∗n(

�γ)) and �λ∗(�γ) =

(�λ∗1(

�γ), . . . , �λ∗�(

�γ)). The problem’s maximum-value function is

�v(�γ) ≡ f(�x∗(�γ);α) +

�∑j=1

�λ∗j(

�γ)hj(�x

∗(�γ); βj, bj). (13.43)

Now we add to (13.42) a new equality constraint, h�+1(x; β�+1, b�+1) = 0. Thisgives us the new, more constrained, primal problem

maxx∈�n

f(x;α) subject to h1(x; β1, b1) = 0, . . . , h�(x; β�, b�) = 0

and h�+1(x; β�+1, b�+1) = 0.(13.44)

We will call (13.44) the (�+ 1)-constraint problem with the parameter vector �+1γ =(�γ : β�+1, b�+1) ≡

(α : �β, β�+1 :

�b, b�+1

). �+1λ

∗�+1(

�+1γ) is the value of the multiplierassociated with the new constraint at the optimal solution (�+1x

∗(�+1γ), �+1λ

∗(�+1γ))

to problem (13.44). Be clear that the parameter vector for the (�+1)-constraint prob-lem is the same as the parameter vector for the �-constraint problem but concatenatedwith (augmented by) the parameters β�+1 and b�+1 from the new constraint.

The ideas presented in what follows are not difficult, but because the notation issomewhat intricate, it will be helpful to use an example as we proceed. The exampleis again problem (11.4), but simplified by setting β1 = β2 =

13. We have already noted

that, at any interesting optimal solution to this problem, the technology constraintbinds and all of the nonnegativity constraints are slack, so � = 1 and the optimalsolution to the inequality constrained problem (11.4) is the same as the optimalsolution to the 1-constraint equality constrained problem

maxy, x1, x2

py − w1x1 − w2x2 subject to h1(y, x1, x2) = x1/31 x

1/32 − y = 0.

In terms of the notation introduced above, the parameter vector for this problem is

Page 467: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

448 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

1γ = (p, w1, w2), the optimal solution function is (see (11.5))

⎛⎜⎜⎜⎜⎜⎜⎝

1y∗(p, w1, w2)

1x∗1(p, w1, w2)

1x∗2(p, w1, w2)

1λ∗(p, w1, w2)

⎞⎟⎟⎟⎟⎟⎟⎠ =

⎛⎜⎜⎜⎜⎜⎜⎝

p2/9w1w2

p3/27w21w2

p3/27w1w22

p

⎞⎟⎟⎟⎟⎟⎟⎠ ,

and the maximum-value function is (see (11.7) with β1 = β2 =13) the indirect profit

function

1v(1γ) = 1v(p, w1, w2) =p3

27w1w2

.

Before we can proceed further, we have to understand the crucial idea of a just-binding constraint. Think of a simple inequality constrained optimization problem

maxx∈�

f(x) subject to g(x) ≤ b.

The optimal solution is (x, λ) = (x∗( b ), λ∗( b )). We say that the inequality constraintg(x) ≤ b is nonbinding (or slack, ineffective, or not active) at x = x∗( b ) if g(x∗( b )) <b, and that the constraint is binding (or tight, effective, or active) at x = x∗( b ) ifg(x∗( b )) = b. Recall that λ∗(b) is the rate of change of the problem’s maximum valuewith respect to the level b of the constraint: λ∗(b) = ∂v(b)/∂b. When the constraint isslack, b can be changed over a small interval ( b−ε, b+ε) with the constraint remainingslack, so λ∗(b) ≡ 0 on this interval, and it is as if the constraint is not present in theproblem. When a constraint is binding for x = x∗( b ), there are two possibilities. Oneis that, for a small enough ε > 0, λ∗(b) > 0 for all values b ∈ ( b− ε, b+ ε), in whichcase the constraint is strictly binding for b = b; tightening the constraint by reducingb to slightly below b lowers the problem’s maximum value and relaxing the constraintby increasing the value of b to slightly above b raises the problem’s maximum value.The other possibility is that g(x∗( b )) = b with λ∗( b ) = 0, and g(x∗(b)) = b for b < bwith λ∗(b) > 0 for b < b. This second possibility is the just-binding for b = b case.What does it mean? When b = b, the rate of change of the problem’s maximum valueis zero, stating that raising the constraint’s level (i.e. relaxing the constraint) aboveb = b will make the constraint nonbinding. But any decrease in the constraint’s level(i.e. tightening the constraint) causes a lower maximum value for the problem. Thus,for both b = b and b > b, it is as if the constraint is not present in the problem, but,

Page 468: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.7. THE SECOND-ORDER ENVELOPE THEOREM 449

for any b < b, the constraint is strictly binding and the problem’s maximum value isreduced.

Here is an example. Consider the problem maxx f(x) = 4− x2 subject to theconstraint that x ≤ b (draw the problem). If b < 0, then the constraint strictly binds,the optimal solution is x∗(b) = b, the Lagrange multiplier is λ∗(b) = −2b > 0, and themaximum value is v(b) = 4− b2 < 4. If b > 0, then the constraint is nonbinding, theoptimal solution is x∗(b) = 0, the Lagrange multiplier is λ∗(b) = 0, and the maximumvalue is v(b) = 4. If b = 0, then the optimal solution is x∗(b = 0) = 0, the constraintbinds since x∗(b = 0) = b = 0, the Lagrange multiplier is λ∗(b = 0) = −2b = 0, themaximum value is v(b = 0) = 4, and any small tightening of the constraint (i.e. anydecrease in b to below zero) results in the constraint becoming strictly binding. Thusthe constraint x ≤ b is just-binding for b = 0.

What if the problem is equality constrained? Consider the problem

maxx∈�

f(x) subject to g(x) = b.

The optimal solution is (x, λ) =(x∗( b ), λ∗( b )

). The constraint is never slack because

it is an equality constraint. The constraint is therefore always binding, either just-binding or strictly binding. The constraint is just-binding at b = b if λ∗( b ) = 0 andλ∗(b) = 0 for b = b. That is, if b = b, then even though the constraint binds, becauseg(x∗( b )) = b, it is as if the constraint is not present, because the maximum valueis the same as for the unconstrained problem maxx∈� f(x). But if b = b, then theconstraint strictly binds, and a change to the level b of the constraint changes boththe optimal solution and the maximum value compared to when the constraint isabsent. Consider the problem maxx f(x) = 4− x2 subject to the equality constraintthat x = b. The only feasible, and thus optimal, solution is x∗(b) = b. If b = 0, thenthe constraint is strictly binding. If b = 0, then the constraint is just-binding becausethe optimal solution x∗(b = 0) = 0 is the same as the solution to the unconstrainedproblem maxx∈� f(x) = 4− x2, which is x∗∗ = 0. Thus, when b = 0, it is as if theconstraint is not there, even though the constraint x = b = 0 holds with equality; i.e.the constraint just-binds when b = 0.

Definition 13.1 (Just-Binding Equality Constraint). If the (� + 1)th constraint inproblem (13.44)

h�+1

(�+1x

∗(�γ0 : β0�+1, b

0�+1

); β0

�+1, b0�+1

)= 0 and �+1λ

∗�+1

(�γ0 : β0

�+1, b0�+1

)= 0

(13.45)

Page 469: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

450 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

and, for every ε > 0, the neighborhood BdE

(�γ0, ε

)contains an �γ such that

h�+1

(�+1x

∗ (�γ : β0�+1, b

0�+1

); β0

�+1, b0�+1

)= 0 and �+1λ

∗�+1

(�γ : β0

�+1, b0�+1

)= 0,

(13.46)then the (�+ 1)th constraint is just-binding for �+1γ0 =

(�γ0 : β0

�+1, b0�+1

).

The values β0�+1 and b0�+1 that parameterize the extra constraint will not alter in

what follows, so we can simplify our notation a little by stating that the (� + 1)th-constraint just binds for �γ = �γ0 and strictly binds for an �γ = �γ0 that is near to�γ0.

(13.45) says that, when �γ = �γ0, the extra constraint holds as an equality buthas no effect on the problem’s maximum value. (13.46) says that there is at least onevalue of �γ arbitrarily close to �γ0 for which the extra constraint strictly binds andchanges the problem’s maximum value.

Be clear that, for �γ = �γ0, it is typically the case that the optimal solution �x∗(�γ)

to the �-constraint problem differs from the optimal solution �+1x∗(�γ : β0

�+1, b0�+1

)to

the (� + 1)-constraint problem. When �γ = �γ0 the two optimal solutions are thesame: �x∗

(�γ0)= �+1x∗

(�γ0 : β0

�+1, b0�+1

).

The maximum-value function for the (�+ 1)-constraint problem is

�+1v(�γ; β0

�+1, b0�+1

)≡ f

(�+1x

∗ (�γ : β0�+1, b

0�+1

);α)

+�∑

j=1

�+1λ∗j

(�γ : β0

�+1, b0�+1

)hj

(�+1x

∗(�γ : β0

�+1, b0�+1); βj, bj

)+ �+1λ

∗�+1

(�γ : β0

�+1, b0�+1

)h�+1

(�+1x

∗(�γ : β0

�+1, b0�+1); β

0�+1, b

0�+1

).

(13.47)

Compare (13.43) and (13.47) before proceeding further.For a given value of �γ0, the maximum value achieved in the (� + 1)-constraint

problem is never larger than the maximum value achieved in the �-constraint problem(right?), but (13.45) and (13.46) say more. Particularly, (13.45) says that the max-imum values of the �-constraint and (� + 1)-constraint problems are the same when�γ = �+1γ0; i.e.

�v(�γ0)= �+1v

(�γ0 : β0

�+1, b0�+1

), (13.48)

and (13.46) says that there is at least one value �γ arbitrarily close to �γ0 for whichthe maximum value of the �-constraint problem is strictly greater than the maximumvalue of the (�+ 1)-constraint problem; i.e. there is an ε > 0 for which

�v(�γ)> �+1v

(�γ : β0

�+1, b0�+1

)for at least one �γ ∈ BdE

(�γ0, ε

); �γ = �γ0. (13.49)

Page 470: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.7. THE SECOND-ORDER ENVELOPE THEOREM 451

Why is this so? If an additional constraint is imposed upon a problem and thisextra constraint is just-binding at the problem’s optimal solution for some value �γ0

of the parameter vector �γ, then for �γ = �γ0 it is as if the additional constraintis not present. Consequently the two problems have the same feasible set and sohave the same optimal solution and maximum value. But when, for some nearbyparameter vector value �γ = �γ0, the additional constraint strictly binds, the moreconstrained problem’s feasible set is a strict subset of the less constrained problem’sfeasible set and does not contain the less constrained problem’s optimal solution.Consequently the maximum value achieved in the more constrained problem is lessthan the maximum value achieved in the less constrained problem. The significantimplication of this statement is that the function �v

(�γ)− �+1v

(�γ : β0

�+1, b0�+1

)must

be strictly convex with respect to �γ for values of �γ that are close to �γ0. Put moreformally, if the extra constraint is just-binding for �γ = �γ0, then, necessarily, (theFirst-Order Envelope Theorem)

�v(�γ0)= �+1v

(�γ0; β0

�+1, b0�+1

)and ∇�γ

�v(�γ)|�γ=�γ0 = ∇�γ

�+1v(�γ; β0

�+1, b0�+1

)|�γ=�γ0

and also necessarily (the Second-Order Envelope Theorem), there exists ε > 0 suchthat the function �v

(�γ)− �+1v

(�γ : β0

�+1, b0�+1

)is strictly convex with respect to �γ

on the neighborhood BdE

(�γ0, ε

).

Figure 13.1 displays the two statements in both of its panels. The panels displaygraphs of �v (·, γk) and �+1v

(·, γk : β0

�+1, b0�+1

), where γk is one of the elements of the

vector �γ. For γk = γ0k , the extra constraint is just-binding, so the maximum-valuefunctions have the same value, �v (·, γ0k) = �+1v

(·, γ0k : β0

�+1, b0�+1

), and the same slope,

∂ �v (·, γk) /∂γk = ∂ �+1v(·, γk : β0

�+1, b0�+1

)/∂γk, at γk = γ0k ; see points A and B in Fig-

ure 13.1. For γk = γ0k , the extra constraint is strictly binding, so over the interval (γ0k−ε, γ0k + ε) the upper-envelope function �v (·, γk) must be strictly less concave (hence,strictly more convex) with respect to γk than is the function �+1v

(·, γk : β0

�+1, b0�+1

).

So over this interval the function �v (·, γk) − �+1v(·, γk : β0

�+1, b0�+1

)must be strictly

convex with respect to γk.

Let’s return to our example. Think of given values 1γ0 = (p0, w01, w

02) for the pa-

rameters of the firm’s 1-constraint equality constrained problem. The optimal solutionvalue for x1 for these parameter values is 1x

∗1(p

0, w01, w

02) = p30/27(w

01)

2w02. Now im-

pose the additional constraint that x1 =1x

∗1 (

1γ0) = 1x∗1(p

0, w01, w

02) = p30/27(w

01)

2w02.

Page 471: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

452 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

Figure 13.1: The Second-Order Envelope Theorem.

Page 472: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.7. THE SECOND-ORDER ENVELOPE THEOREM 453

This gives the 2-constraint equality constrained problem

maxy, x1, x2

py − w1x1 − w2x2 subject to h1(y, x1, x2) = x1/31 x

1/32 − y = 0

and h2(x1;

1x∗1

(1γ0))

= 1x∗1

(1γ0)− x1 =

p30

27 (w01)

2w0

2

− x1 = 0.

This problem’s parameter vector is 2γ = (1γ : b2) =(1γ : 1x

∗1 (

1γ0)). The optimal

solution is(2y

∗(2γ) , 1x

∗1 (

1γ0) , 2x∗2 (

2γ)). The additional constraint just-binds for 1γ =

1γ0 = (p0, w01, w

02) and strictly binds for 1γ = 1γ0. The optimal solution is

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

2y∗ (1γ; 1x∗1 (1γ0))

2x∗1

(1γ; 1x

∗1 (

1γ0))

2x∗2

(1γ; 1x

∗1 (

1γ0))

2λ∗1

(1γ; 1x

∗1 (

1γ0))

2λ∗2

(1γ; 1x

∗1 (

1γ0))

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠=

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

2y∗ (p, w1, w2;

1x∗1 (

1γ0))

2x∗1

(p, w1, w2;

1x∗1 (

1γ0))

2x∗2

(p, w1, w2;

1x∗1 (

1γ0))

2λ∗1

(p, w1, w2;

1x∗1 (

1γ0))

2λ∗2

(p, w1, w2;

1x∗1 (

1γ0))

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠=

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

(p 1x

∗1 (

1γ0)

3w2

)1/2

1x∗1 (

1γ0)(p3 1x

∗1 (

1γ0)

27w32

)1/2

p(p3

27w21x∗1 (1γ0)

)1/2

− w1

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠

and the maximum-value function is (see (13.47))

2v(1γ : 1x

∗1

(1γ0))

= 2v(p, w1, w2 :

1x∗1

(1γ0))

= p 2y∗ (1γ : 1x

∗1

(1γ0))

− w11x

∗1

(1γ0)− w2

2x∗2

(1γ : 1x

∗1

(1γ0))

+ 2λ∗1

(1γ : 1x

∗1

(1γ0)) (

1x∗1

(1γ0)1/3 × 2x

∗2

(1γ : 1x

∗1

(1γ0))1/3 − 2y

∗ (1γ : 1x∗1

(1γ0)))

+ 2λ∗2

(1γ : 1x

∗1

(1γ0)) (

1x∗1

(1γ0)− 2x

∗1

(1γ : 1x

∗1

(1γ0)))

=2

3

(p3 1x

∗1 (

1γ0)

3w2

)1/2

− w11x

∗1

(1γ).

Notice that, for 1γ = 1γ0, the value 2λ∗2(p, w1, w2;

1x∗1 (

1γ0))of the multiplier of the

additional constraint is

2λ∗2

(p0, w0

1, w02;

1x∗1

(1γ0))

=

(p30

27w02

× 27 (w01)

2w0

2

p30

)1/2

− w01 = 0

and that 2λ∗2(p, w1, w2;

1x∗1 (

1γ0))= 0 for 1γ = 1γ0.

Page 473: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

454 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

Theorem 13.1 (The Second-Order Envelope Theorem). Consider a primal optimiza-tion problem with � equality constraints. If � = 0, then the problem is

maxx∈�n

f(x;α).

If � ≥ 1, then the problem is

maxx∈�n

f(x;α) subject to hj(x; βj, bj) = 0 for j = 1, . . . , �.

Denote the parameter vector for the �-constraint problem by �γ ∈ �Γ. Consider onlyproblems for which the optimal solution is a real-valued function

(�x

∗ (�γ) , �λ∗ (�γ))that is continuously differentiable on �Γ. The maximum-value function for the �-constraint problem is

�v(�γ)≡ f

(�x

∗ (�γ) ;α)+ �∑j=1

�λ∗j

(�γ)hj

(�x

∗ (�γ) ; βj, bj).The (� + 1)-constraint problem is the �-constraint problem with one additional con-straint h�+1 (x; β�+1, b�+1) = 0 that just-binds when �γ = �γ0, and for which the op-timal solution is a real-valued function

(�+1x

∗ (�γ : β0�+1, b

0�+1

), �+1λ

∗ (�γ : β0�+1, b

0�+1

))that is continuously differentiable on �Γ. The maximum-value function for the (�+1)-constraint problem is

�+1v(�γ : β0

�+1, b0�+1

)≡ f

(�+1x

∗ (�γ : β0�+1, b

0�+1

);α)

+�+1∑j=1

�+1λ∗j

(�γ : β0

�+1, b0�+1

)hj

(�+1x

∗ (�γ : β0�+1, b

0�+1

); βj, bj

).

Then there is an ε > 0 for which the function �v(�γ)− �+1v

(�γ : β0

�+1, b0�+1

)is strictly

convex with respect to �γ over the neighborhood BdE

(�γ0, ε

).

Proof. For small enough perturbations Δ�γ, a second-order Taylor’s series expansionof �v

(�γ)at �γ = �γ0 gives

�v(�γ0 +Δ�γ

)= �v

(�γ0)+∇�γ

�v(�γ)| �γ=�γ0

·Δ�γ +1

2Δ�γ

T∇2�γ�γ

�v(�γ)| �γ=�γ0

Δ�γ.

(13.50)

Page 474: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.7. THE SECOND-ORDER ENVELOPE THEOREM 455

Similarly

�+1v(�γ0 +Δ�γ : β0

�+1, b0�+1

)= �+1v

(�γ0 : β0

�+1, b0�+1

)+∇�γ

�+1v(�γ : β0

�+1, b0�+1

)| �γ=�γ0

·Δ�γ

+1

2Δ�γ

T∇2�γ�γ

�+1v(�γ : β0

�+1, b0�+1

)| �γ=�γ0

Δ�γ.

(13.51)

Since the (�+ 1)th-constraint is just-binding for �γ = �γ0,

�v(�γ0)= �+1v

(�γ0 : β0

�+1, b0�+1

)(13.52)

and �v(�γ0 +Δ�γ

)> �+1v

(�γ0 +Δ�γ : β0

�+1, b0�+1

)(13.53)

for all small enough Δ�γ = 0. By the First-Order Envelope Theorem,

∇�γ�v(�γ)| �γ=�γ0

= ∇�γ�+1v

(�γ : β0

�+1, b0�+1

)| �γ=�γ0

. (13.54)

Using (13.52), (13.53), and (13.54) in (13.50) and (13.51) gives

Δ�γT(∇2

�γ�γ�v(�γ)−∇2

�γ�γ�+1v

(�γ : β0

�+1, b0�+1

))| �γ=�γ0

Δ�γ > 0 (13.55)

for small enough Δ�γ. That is, there is an ε > 0 for which �v(�γ)−�+1v

(�γ : β0

�+1, b0�+1

)is strictly convex with respect to �γ for all �γ ∈ BdE

(�γ0, ε

).

For our example,

1v(p, w1, w2)−2v(p, w1, w2;

1x∗1

(1γ0))

=p3

27w1w2

− 2

3

(p3 1x

∗1 (

1γ0)

3w2

)1/2

+w11x

∗1

(1γ0).

(13.56)Choose any one of the parameters p, w1, or w2. I will pick p. You can try whatfollows for w1 and w2. The second partial derivative of (13.56) with respect to p is

∂2 (1v(p, ·)− 2v(p, ·))∂p2

=2p

9w1w2

− 1

2

(1x

∗1 (

1γ0)

3pw2

)1/2

. (13.57)

At (p, w1, w2) = (p0, w01, w

02) the value of (13.57) is

∂2 (1v(p, ·)− 2v(p, ·))∂p2

|1γ=(p0,w01 ,w

02)

=p0

6w01w

02

> 0,

Page 475: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

456 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

establishing that (13.56) is strictly convex with respect to p at and near to 1γ = 1γ0 =(p0, w0

1, w02). An evaluation at 1γ0 of the Hessian matrix of (13.56) shows that (13.56)

is jointly strictly convex with respect to p, w1, and w2 for values of (p, w1, w2) nearto (p0, w0

1, w02).

It is a simple matter to use the Second-Order Envelope Theorem to compare thecomparative statics results of a constrained optimization problem to the results of amore constrained version of that problem. We will see how to do this in the nextsection.

13.8 Le Chatelier’s Principle

Henri-Louis Le Chatelier (1850 -1936) was a highly regarded French industrial chemistwho made an observation that applies to a huge variety of circumstances. The mostfrustrating aspect of his principle is that, although it is a simple idea, it is not easy tostate it accurately in just a few words. Le Chatelier’s first published statement in 1884of his principle was a lengthy and confusing paragraph in the research journal of theFrench Academy of Sciences. Four years later he gave a much simpler restatement.A translation into English of his restatement is:

Every change of one of the factors of an equilibrium occasions a rear-rangement of the system in such a direction that the factor in questionexperiences a change in a sense opposite to the original change.

Perhaps you do not find even his restatement to be immediately comprehensible?

Le Chatelier’s main concern was with changes to the equilibria of chemical systems,but others have since tried to construct more useful statements of the principle andto make it applicable to the social sciences, including economics. Such a restatementis found in Milgrom and Roberts 1996. Put into words, it states that we are toconsider two optimization problems that are the same except that one is subject toan extra equality constraint. We start with an optimal solution (an equilibrium, in LeChatelier’s terminology) to the more restricted problem at which the extra constraintis just-binding, so this is also an optimal solution to the less restricted problem.Now we perturb a little the vector of the parameters that are common to the twoproblems. Le Chatelier’s Principle states that the changes caused to the optimalsolution of the less restricted problem are at least as large as the changes causedto the optimal solution of the more restricted problem. An economic example oftenused to illustrate the idea is that the changes that a profit-maximizing competitive

Page 476: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.8. LE CHATELIER’S PRINCIPLE 457

firm makes to its quantity supplied and input quantities demanded in response to achange to, say, the price of its product are at least as large in the (less-restricted)“long-run” as in any (more restricted) “short-run.” We will examine this applicationin a moment. The tool that we will use is the Second-Order Envelope Theorem.

The theorem states that, if the (�+1)th-constraint just-binds when �γ = �γ0, thenthe Hessian matrix of the function �v

(�γ)− �+1v

(�γ : β0

�+1, b0�+1

)with respect to the

parameter vector �γ of the �-constrained problem necessarily is positive-definite (notmerely positive-semi-definite) at �γ0 for all small enough perturbations Δ�γ from �γ0.That is,

Δ�γT(∇2

�γ�γ

(�v(�γ)− �+1v

(�γ : β0

�+1, b0�+1

)))|�γ=�γ0

Δ�γ > 0 (13.58)

for all small enough perturbations Δ�γ from 1γ0. What do we learn from this?

The typical element in the Hessian matrix in (13.58) is

∂2 �v(�γ)

∂γi∂γj− ∂2 �+1v

(�γ : β0

�+1, b0�+1

)∂γi∂γj

for i, j = 1, . . . , r. (13.59)

A matrix is positive-definite only if all of its main diagonal elements are strictlypositive, so, from (13.59),

∂2 �v(�γ)

∂γ2i>∂2 �+1v

(�γ : β0

�+1, b0�+1

)∂γ2i

for i = 1, . . . , r. (13.60)

To see how to use this information, let’s return to our example of a profit-maximizing firm. The original 1-constraint equality constrained problem is

maxy, x1, x2

py − w1x1 − w2x2 subject to h1(y, x1, x2) = x1/31 x

1/32 − y = 0.

The 2-constraint equality constrained problem is

maxy, x1, x2

py − w1x1 − w2x2 subject to h1(y, x1, x2) = x1/31 x

1/32 − y = 0

and h2(x1;

1x∗1

(p0, w0

1, w02

))= 1x

∗1

(p0, w0

1, w02

)− x1 = 0,

where 1x∗1 (p

0, w01, w

02) is the optimal value for x1 in the 1-constraint problem when

(p, w1, w2) = (p0, w01, w

02). We have already discovered from Hotelling’s Lemma (see

Page 477: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

458 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

Theorem 12.4) that

∂2 (1v(p, w1, w2))

∂p2=∂(1y

∗(p, w1, w2)

)∂p

and

∂2(2v(p, w1, w2 :

1x∗1 (p

0, w01, w

02)))

∂p2=∂(2y

∗ (p, w1, w2 :

1x∗1 (p

0, w01, w

02)))

∂p,

that

∂2 (1v (p, w1, w2))

∂w21

= − ∂(1x

∗1 (p, w1, w2)

)∂w1

and

∂2(2v(p, w1, w2 :

1x∗1 (p

0, w01, w

02)))

∂w21

= − ∂(2x

∗1

(p, w1, w2 :

1x∗1 (p

0, w01, w

02)))

∂w1

= 0,

and that

∂2 (1v (p, w1, w2))

∂w22

= − ∂(1x

∗2 (p, w1, w2)

)∂w2

and

∂2(2v(p, w1, w2 :

1x∗1 (p

0, w01, w

02)))

∂w22

= − ∂(2x

∗2

(p, w1, w2 :

1x∗1 (p

0, w01, w

02)))

∂w2

.

Therefore, from (13.60),

∂(1y

∗(p, w1, w2)

)∂p

>∂(2y

∗ (p, w1, w2 :

1x∗1 (p

0, w01, w

02)))

∂p, (13.61)

∂(1x

∗1 (p, w1, w2)

)∂w1

<∂(2x

∗1

(p, w1, w2 :

1x∗1 (p

0, w01, w

02)))

∂w1

= 0, (13.62)

and∂(1x

∗2 (p, w1, w2)

)∂w2

<∂(2x

∗2

(p, w1, w2 :

1x∗1 (p

0, w01, w

02)))

∂w2

. (13.63)

These are familiar comparisons of comparative statics results. (13.61) informs us thatrate-of-change of the quantity of product supplied by a competitive firm is strictlylarger in the “long-run” than it is in the “short-run” in which one of the firm’s inputlevels is fixed. (13.62) and (13.63) inform us that the rate-of-change of the quantitydemanded of a variable input by a competitive firm is strictly more negative in thelong-run than it is in the short-run in which one of the firm’s input levels is fixed.In Le Chatelier’s parlance, the long-run problem’s equilibrium adjusts by more soas to better overcome, or exploit, the consequences of parameter value changes. Inour parlance, the principle is that the rate at which any component of an optimal

Page 478: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.9. MORE COMPARISONS OF COMPARATIVE STATICS 459

solution changes with respect to a parameter common to a less constrained and amore constrained problem is larger in size for the less constrained problem than formore constrained problem. The principle is a necessary consequence of the additionalconstraint reducing the ways in which a problem’s choice variables can be adjustedto offset the adverse (maximum-value reducing) consequences of the additional con-straint.

13.9 More Comparisons of Comparative Statics

Le Chatelier-type results are not the only type of comparative statics results providedby the Second-Order Envelope Theorem.

The Hessian matrix in (13.58) is symmetric, so

∂2 �v(�γ)

∂γj∂γk− ∂2 �+1v

(�γ : β0

�+1, b0�+1

)∂γj∂γk

=∂2 �v

(�γ)

∂γk∂γj− ∂2 �+1v

(�γ : β0

�+1, b0�+1

)∂γk∂γj

for j, k = 1, . . . , r; j = k. In the context of our example, this gives us results such as

∂(1y

∗(p, w1, w2)

)∂w1

− ∂(2y

∗ (p, w1, w2 :

1x∗1 (p

0, w01, w

02)))

∂w1

= −∂(1x

∗1(p, w1, w2)

)∂p

,

as well as

∂(1y

∗(p, w1, w2)

)∂w2

− ∂(2y

∗ (p, w1, w2 :

1x∗1 (p

0, w01, w

02)))

∂w2

= − ∂(1x

∗2(p, w1, w2)

)∂p

+∂(2x

∗2

(p, w1, w2 :

1x∗1 (p

0, w01, w

02)))

∂pand

∂(1x

∗1(p, w1, w2)

)∂w2

= − ∂(1x

∗2(p, w1, w2)

)∂w1

+∂(2x

∗2

(p, w1, w2 :

1x∗1 (p

0, w01, w

02)))

∂w1

.

Whether or not these particular results are useful is debatable. All I wish to pointout to you is that they are necessarily true and are easily obtained.

A symmetric matrix is positive-definite if and only if all of its successive principalminors are strictly positive. This provides still more comparison of comparative stat-

ics results from the Hessian matrix(∇2

1γ1γ1v (1γ)−∇2

1γ1γ2v(1γ : β0

�+1, b0�+1

))|1γ=1γ0 .

These results have even more obscure economic interpretations than those above, soI will not present them, but you might like to derive some of them just to see whatthey look like.

Page 479: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

460 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

13.10 Problems

Problem 13.1. A firm uses quantities x1, x2, and x3 of three inputs to produce asingle output. The positive per-unit prices of the inputs are w1, w2, and w3. Thefirm’s production function is g(x1, x2, x3). The firm’s cost-minimization problem is

maxx1, x2, x3

−w1x1 − w2x2 − w3x3 subject to

h1(x1) = x1 ≥ 0, h2(x2) = x2 ≥ 0, h3(x3) = x3 ≥ 0,

and h4(x1, x2, x3; y) = g(x1, x2, x3)− y ≥ 0

(13.64)

for given values of w1, w2, w3, and y. Assume that g is twice-continuously differ-entiable and that g’s properties cause the nonnegativity constraints on x1, x2, andx3 to be slack at an optimal solution when y > 0. That is, the cost-minimizingbundle (x∗1(w1, w2, w3, y), x

∗2(w1, w2, w3, y), x

∗3(w1, w2, w3, y)) of the firm’s conditional

input demands contains only positive values when y > 0. Assume also that g’s prop-erties cause the Karush-Kuhn-Tucker condition to be well-defined at any optimalsolution to the primal problem (13.64).(i) There is a different primal problem (13.64) for each different value of the parametervector (w1, w2, w3, y). What is the primal-dual problem that is generated by thisfamily of primal problems?(ii) Are any of the constraints in the primal-dual problem different from the con-straints in the primal problems? Explain.(iii) Write down the first-order necessary maximization conditions for any one of theprimal problems.(iv) Write down the first-order necessary maximization conditions for the primal-dual problem. Compare these conditions to the necessary first-order maximizationconditions for any of the primal problems.(v) Derive the necessary second-order local maximization condition for any one ofthe primal problems (13.64).(vi) Derive the primal-dual problem’s necessary second-order maximization condi-tion.(vii) Use your answer to part (vi) to derive some comparative statics properties ofthe firm’s conditional input demand functions and its marginal cost of productionfunction.

Problem 13.2. Slutsky’s equation is the major quantitative comparative statics re-sult of neoclassical consumer demand theory. It is a quantitative result because itallows numerical computation of the rates-of-change of ordinary demand functions,

Page 480: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.11. ANSWERS 461

Hicksian demand functions, and income effects. However, one wonders what quali-tative comparative statics results might be implied by this theory of demand. Theanswer to this inquiry is obtained by applying Silberberg’s methodology to the con-sumer’s expenditure minimization problem. Set up the appropriate primal-dual prob-lem and use its necessary optimization conditions to discover the implied qualitativecomparative statics results.

Problem 13.3. The standard consumer expenditure minimization problem is

minx1, ... , xn

p1x1 + · · ·+ pnxn subject to x1 ≥ 0, . . . , xn ≥ 0 and U(x1, . . . , xn) ≥ u.

The parameters of this problem are p1, . . . , pn, and u. Let p = (p1, . . . , pn) andlet (p0, u0) denote an initial value for the parameter vector (p, u). h(p0, u0) is theimplied initial value of the expenditure-minimizing (Hicksian compensated demand)consumption vector, so, in particular, h01 = h1(p

0, y0) is the quantity of commodity 1that is initially demanded when the consumer faces the initial price vector p0 andachieves the utility level u0.

Use the Second-Order Envelope Theorem to compare the comparative statics re-sults for the consumer for small perturbations from (p0, y0) when the consumer is,and is not, subject to the extra constraint that x1 = h01.

13.11 Answers

Answer to Problem 13.1.

(i) The maximum-value function for the family of primal problems (13.64) is thenegative of the firm’s (long-run) cost of production function

−c�(w1, w2, w3, y) = −w1x∗1(w1, w2, w3, y)− w2x

∗2(w1, w2, w3, y)− w3x

∗3(w1, w2, w3, y),

so the primal-dual problem that is generated by the family of primal problems is

maxx1, x2, x3, w1, w2, w3, y

c�(w1, w2, w3, y)− w1x1 − w2x2 − w3x3

subject to h1(x1) = x1 ≥ 0, h2(x2) = x2 ≥ 0, h3(x3) = x3 ≥ 0,

and h4(x1, x2, x3, y) = g(x1, x2, x3)− y ≥ 0.

(13.65)

(ii) The first, second, and third constraints of the primal-dual problem are the sameas the first, second, and third constraints of any of the primal problems. The fourth

Page 481: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

462 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

constraint in the primal-dual problem differs from the fourth constraint in any of theprimal problems. Why? In the fourth constraint of any of the primal problems, thevalue of y is fixed, so this constraint confines choices of x1, x2, and x3 to the setSp(y) = {(x1, x2, x3) | g(x1, x2, x3) ≥ y}. y is a choice variable in the primal-dualproblem’s fourth constraint, so the constraint confines choices of x1, x2, x3, and y tothe set Spd = {(x1, x2, x3, y) | g(x1, x2, x3) − y ≥ 0}. For any given value of y > 0,the set Sp(y) is only a subset of Spd.(iii) The Karush-Kuhn-Tucker condition for the primal problem (13.64) is

(−w1,−w2,−w3) = −λp1(1, 0, 0)− λp2(0, 1, 0)− λp3(0, 0, 1)− λp4

(∂g(·)∂x1

,∂g(·)∂x2

,∂g(·)∂x3

)(13.66)

evaluated at (x1, x2, x3) = (x∗1(w1, w2, w3, y), x∗2(w1, w2, w3, y), x

∗3(w1, w2, w3, y)). We

are told that, at this maximizing solution, the first, second, and third constraints areslack, so λp1 = λp2 = λp3 = 0, making the Karush-Kuhn-Tucker condition

(w1, w2, w3) = λ∗p4 (w1, w2, w3, y)

(∂g(·)∂x1

,∂g(·)∂x2

,∂g(·)∂x3

) ∣∣∣∣x1=x∗1(w1,w2,w3,y)

x2=x∗2(w1,w2,w3,y)

x3=x∗3(w1,w2,w3,y)

(13.67)

with λ∗p4 (w1, w2, w3, y) > 0. The only useful complementary slackness condition is

g(x∗1(w1, w2, w3, y), x∗2(w1, w2, w3, y), x

∗3(w1, w2, w3, y)) = y. (13.68)

(iv) The Karush-Kuhn-Tucker condition for the primal-dual problem (13.65) is(−w1,−w2,−w3,

∂c�(·)∂w1

− x∗1(·),∂c�(·)∂w2

− x∗2(·),∂c�(·)∂w3

− x∗3(·),∂c�(·)∂y

)= − λpd1 (1, 0, 0, 0, 0, 0, 0)− λpd2 (0, 1, 0, 0, 0, 0, 0)− λpd3 (0, 0, 1, 0, 0, 0, 0)

− λpd4

(∂g(·)∂x1

,∂g(·)∂x2

,∂g(·)∂x3

, 0, 0, 0,−1

) ∣∣∣∣x1=x∗1(w1,w2,w3,y)

x2=x∗2(w1,w2,w3,y)

x3=x∗3(w1,w2,w3,y)

. (13.69)

The first three components of (13.69) are the primal problem’s necessary maximiza-tion conditions (13.67). Why? Because any primal-dual problem optimal solu-tion (x∗1(w1, w2, w3, y), x

∗2(w1, w2, w3, y), x

∗3(w1, w2, w3, y), w1, w2, w3, y) must, for the

same but fixed values of w1, w2, w3, and y, imply the same optimal solution valuesx∗1(w1, w2, w3, y), x

∗2(w1, w2, w3, y), and x

∗3(w1, w2, w3, y) for the primal problem that

is parameterized by these particular values for w1, w2, w3, and y. Since the non-negativity constraints are slack at an optimal solution, λ∗pd1 = λ∗pd2 = λ∗pd3 = 0 and

Page 482: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.11. ANSWERS 463

λ∗pd4 (w1, w2, w3, y) > 0, so the primal-dual problem’s Karush-Kuhn-Tucker conditionis (

−w1,−w2,−w3,∂c�(·)∂w1

− x∗1(·),∂c�(·)∂w2

− x∗2(·),∂c�(·)∂w3

− x∗3(·),∂c�(·)∂y

)

= − λ∗pd4 (·)(∂g(·)∂x1

,∂g(·)∂x2

,∂g(·)∂x3

, 0, 0, 0,−1

) ∣∣∣∣x1=x∗1(w1,w2,w3,y)

x2=x∗2(w1,w2,w3,y)

x3=x∗3(w1,w2,w3,y)

. (13.70)

The primal-dual problem’s Karush-Kuhn-Tucker condition contains a component foreach of the primal-dual choice variables w1, w2, w3, and y. These do not appearin the primal problems’ Karush-Kuhn-Tucker conditions since these quantities havefixed values in any primal problem. These components

∂c�(w1, w2, w3, y)

∂y= λ∗pd4 (w1, w2, w3, y) and

∂c�(w1, w2, w3, y)

∂wi

= x∗i (w1, w2, w3, y)

(13.71)for i = 1, 2, 3 reveal the First-Order Envelope Theorem result that is Shephard’sLemma, and also that the multiplier for the strictly binding production functionconstraint is the firm’s marginal cost of production function.(v) The primal problem’s restricted Lagrange function is

Lpr(x1, x2, x3;λ1=0, λ2=0, λ3=0, λ4=λ∗p4 (w1, w2, w3, y))

= − w1x1 − w2x2 − w3x3 + λ∗p4 (w1, w2, w3, y)(g(x1, x2, x3)− y).(13.72)

The necessary second-order local maximization condition for a primal problem is that,there is an ε > 0 such that

(Δx1,Δx2,Δx3)

⎛⎜⎜⎜⎜⎜⎜⎜⎝

∂2Lpr

∂x21

∂2Lpr

∂x1∂x2

∂2Lpr

∂x1∂x3∂2Lpr

∂x2∂x1

∂2Lpr

∂x22

∂2Lpr

∂x2∂x3∂2Lpr

∂x3∂x1

∂2Lpr

∂x3∂x2

∂2Lpr

∂x23

⎞⎟⎟⎟⎟⎟⎟⎟⎠∣∣∣∣∣x1=x∗

1(·)x2=x∗

2(·)x3=x∗

3(·)

⎛⎝Δx1Δx2Δx3

⎞⎠ = (13.73)

λ∗p4 (·)(Δx1,Δx2,Δx3)

⎛⎜⎜⎜⎜⎜⎜⎜⎝

∂2g

∂x21

∂2g

∂x1∂x2

∂2g

∂x1∂x3∂2g

∂x2∂x1

∂2g

∂x22

∂2g

∂x2∂x3∂2g

∂x3∂x1

∂2g

∂x3∂x2

∂2g

∂x23

⎞⎟⎟⎟⎟⎟⎟⎟⎠∣∣∣∣∣x1=x∗

1(·)x2=x∗

2(·)x3=x∗

3(·)

⎛⎝Δx1Δx2Δx3

⎞⎠ ≤ 0

Page 483: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

464 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

for all (Δx1,Δx2,Δx3) satisfying both g(x∗1(·)+Δx1, x

∗2(·)+Δx2, x

∗3(·)+Δx3) = y and

(x∗1(·)+Δx1, x∗2(·)+Δx2, x

∗3(·)+Δx3) ∈ BdE((x

∗1(·), x∗2(·), x∗3(·)), ε). This is a statement

that, over a set of values of (x1, x2, x3) that are “close” to the optimal solution toa primal problem (13.64) and satisfy the binding constraint g(x1, x2, x3) = y, theproduction function g is necessarily at least weakly concave.(vi) The primal-dual problem’s restricted Lagrange function is

Lpdr(x1, x2, x3, w1, w2, w3, y;λ1=0, λ2=0, λ3=0, λ4=λ∗pd4 (w1, w2, w3, y)) = (13.74)

c�(w1, w2, w3, y)− w1x1 − w2x2 − w3x3 + λ∗pd4 (w1, w2, w3, y)(g(x1, x2, x3)− y).

Let (Δx,Δγ) = (Δx1,Δx2,Δx3,Δw1,Δw2,Δw3,Δy). Then the primal-dual prob-lem’s necessary second-order local maximization condition is that, there is an ε > 0such that

(Δx,Δγ)

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

∂2Lpdr

∂x21

∂2Lpdr

∂x1∂x2

∂2Lpdr

∂x1∂x3

∂2Lpdr

∂x1∂w1

∂2Lpdr

∂x1∂w2

∂2Lpdr

∂x1∂w3

∂2Lpdr

∂x1∂y

∂2Lpdr

∂x2∂x1

∂2Lpdr

∂x22

∂2Lpdr

∂x2∂x3

∂2Lpdr

∂x2∂w1

∂2Lpdr

∂x2∂w2

∂2Lpdr

∂x2∂w3

∂2Lpdr

∂x2∂y

∂2Lpdr

∂x3∂x1

∂2Lpdr

∂x3∂x2

∂2Lpdr

∂x23

∂2Lpdr

∂x3∂w1

∂2Lpdr

∂x3∂w2

∂2Lpdr

∂x3∂w3

∂2Lpdr

∂x3∂y

∂2Lpdr

∂w1∂x1

∂2Lpdr

∂w1∂x2

∂2Lpdr

∂w1∂x3

∂2Lpdr

∂w21

∂2Lpdr

∂w1∂w2

∂2Lpdr

∂w1∂w3

∂2Lpdr

∂w1∂y

∂2Lpdr

∂w2∂x1

∂2Lpdr

∂w2∂x2

∂2Lpdr

∂w2∂x3

∂2Lpdr

∂w2∂w1

∂2Lpdr

∂w22

∂2Lpdr

∂w2∂w3

∂2Lpdr

∂w2∂y

∂2Lpdr

∂w3∂x1

∂2Lpdr

∂w3∂x2

∂2Lpdr

∂w3∂x3

∂2Lpdr

∂w3∂w1

∂2Lpdr

∂w3∂w2

∂2Lpdr

∂w23

∂2Lpdr

∂w3∂y

∂2Lpdr

∂y∂x1

∂2Lpdr

∂y∂x2

∂2Lpdr

∂y∂x3

∂2Lpdr

∂y∂w1

∂2Lpdr

∂y∂w2

∂2Lpdr

∂y∂w3

∂2Lpdr

∂y2

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠

(ΔxΔγ

)

= (Δx,Δγ)

( ∇2x,xL

pdr ∇2x,w,yL

pdr

∇2w,y,xL

pdr ∇2w,yL

pdr

)(ΔxΔγ

)≤ 0 (13.75)

for all (Δx1,Δx2,Δx3,Δw1,Δw2,Δw3,Δy) satisfying both

g(x∗1(·) + Δx1, x∗2(·) + Δx2, x

∗3(·) + Δx3) = y +Δy and

(x∗1(·) + Δx1, x∗2(·) + Δx2, x

∗3(·) + Δx3,Δw1,Δw2,Δw3,Δy)

∈ BdE((x∗1(·), x∗2(·), x∗3(·), w1, w2, w3, y), ε),

Page 484: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.11. ANSWERS 465

where

∇2x,xL

pdr = λ∗pd4 (·)

⎛⎜⎜⎜⎜⎜⎜⎜⎝

∂2g

∂x21

∂2g

∂x1∂x2

∂2g

∂x1∂x3∂2g

∂x2∂x1

∂2g

∂x22

∂2g

∂x2∂x3∂2g

∂x3∂x1

∂2g

∂x3∂x2

∂2g

∂x23

⎞⎟⎟⎟⎟⎟⎟⎟⎠,

∇2x,w,yL

pdr =

⎛⎝−1 0 0 0

0 −1 0 00 0 −1 0

⎞⎠ , ∇2

w,y,xLpdr =

⎛⎜⎜⎝−1 0 00 −1 00 0 −10 0 0

⎞⎟⎟⎠ ,

and

∇2w,yL

pdr =

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

∂2c�

∂w21

∂2c�

∂w1∂w2

∂2c�

∂w1∂w3

∂2c�

∂w1∂y

∂2c�

∂w2∂w1

∂2c�

∂w22

∂2c�

∂w2∂w3

∂2c�

∂w2∂y

∂2c�

∂w3∂w1

∂2c�

∂w3∂w2

∂2c�

∂w23

∂2c�

∂w3∂y

∂2c�

∂y∂w1

∂2c�

∂y∂w2

∂2c�

∂y∂w3

∂λ∗pd4

∂y

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠.

(vii) If Δγ = (Δw1,Δw2,Δw3,Δy) = (0, 0, 0, 0), then the primal-dual problem’snecessary second-order maximization condition is the same as for a primal prob-lem. If we wish to conduct a comparative statics experiment, then we set Δx =(Δx1,Δx2,Δx3) = (0, 0, 0) and obtain from the primal-dual problem’s necessarysecond-order maximization condition that, necessarily, there is an ε > 0 such that

(Δw1,Δw2,Δw3,Δy)∇2w,yL

pdr

⎛⎜⎜⎝Δw1

Δw2

Δw3

Δy

⎞⎟⎟⎠ ≤ 0

Page 485: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

466 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

for all (Δw1,Δw2,Δw3,Δy) ∈ BdE((0, 0, 0, 0), ε). From (13.71), this condition is

(Δw1,Δw2,Δw3,Δy)

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

∂x∗1∂w1

∂x∗1∂w2

∂x∗1∂w3

∂x∗1∂y

∂x∗2∂w1

∂x∗2∂w2

∂x∗2∂w3

∂x∗2∂y

∂x∗3∂w1

∂x∗3∂w2

∂x∗3∂w3

∂x∗3∂y

∂λ∗pd4

∂w1

∂λ∗pd4

∂w2

∂λ∗pd4

∂w3

∂λ∗pd4

∂y

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠

⎛⎜⎜⎝Δw1

Δw2

Δw3

Δy

⎞⎟⎟⎠ ≤ 0,

revealing that, necessarily,

∂x∗i∂wi

≤ 0 and∂x∗i∂y

=∂λ∗pd4

∂wi

for i = 1, 2, 3, and∂x∗i∂wj

=∂x∗j∂wi

for i, j = 1, 2, 3; i = j.

Answer to Problem 13.2. Use p to denote the price vector (p1, . . . , pn) and usex to denote the consumption bundle (x1, . . . , xn). Given p and a utility level u, theconsumer’s expenditure minimization problem is to choose x to minimize p·x subjectto U(x) ≥ u or, equivalently, to

maxx

−p·x subject to U(x) ≥ u, (13.76)

where U is strictly increasing. This primal problem’s Lagrange function is

Lp(x, λp) = −p·x+ λp(U(x)− u). (13.77)

For given (p, u) the optimal solution is the Hicksian demand vector h(p, u). Theconsumer’s expenditure function is the minimum value function

e(p, u) = p·h(p, u) + λ∗(p, u)(U(h(p, u))− u). (13.78)

Because the expenditure function is a minimum value function,

e(p, u)− p·x ≤ 0 for all x satisfying U(x) ≥ u.

Therefore the primal-dual problem is

maxx, p, y

e(p, u)− p·x subject to U(x) ≥ u (13.79)

Page 486: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.11. ANSWERS 467

and its Lagrange function is

Lpd(x, p, u, λpd) = e(p, u)− p·x+ λpd(U(x)− u). (13.80)

The first-order conditions for the primal-dual problem are

∂Lpd

∂xi= −pi + λpd

∂U

∂xi= 0; i = 1, . . . , n, (13.81)

∂Lpd

∂pi=

∂e

∂pi− xi = 0; i = 1, . . . , n, (13.82)

∂Lpd

∂u=∂e

∂u− λpd = 0, and (13.83)

U(x) = u. (13.84)

(13.81) and (13.84) are the first-order conditions for the primal problem (13.76).(13.82) and (13.83) are applications of the First-Order Envelope Theorem to (13.78);their ratio establishes Roy’s Identity. (13.83) shows that the value λ∗pd(p, u) > 0 ofthe multiplier of the utility constraint is the reciprocal of the consumer’s marginalutility of expenditure. These results are not new. What is new arises from the factthat the Hessian of Lpd is necessarily negative-semi-definite for feasibility-preservingperturbations of (x, p, u). For perturbations just of (p, u), then, the submatrix ofthe Hessian of Lpd consisting only of the second-order partial derivatives of Lpd withrespect to (p, u),

Hpdp,u =

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

∂2Lpd

∂p21

∂2Lpd

∂p2∂p1· · · ∂2Lpd

∂pn∂p1

∂2Lpd

∂u∂p1∂2Lpd

∂p1∂p2

∂2Lpd

∂p22· · · ∂2Lpd

∂pn∂p2

∂2Lpd

∂u∂p2...

.... . .

......

∂2Lpd

∂p1∂pn

∂2Lpd

∂p2∂pn· · · ∂2Lpd

∂p2n

∂2Lpd

∂u∂pn∂2Lpd

∂p1∂u

∂2Lpd

∂p2∂u· · · ∂2Lpd

∂pn∂u

∂2Lpd

∂u2

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠

is necessarily negative-semi-definite with respect to small feasibility-preserving per-turbations of (p, u) when it is evaluated at an optimal solution to the primal-dual

Page 487: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

468 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

problem. ∂e(p, u)/∂pi = hi(p, u), so, from (13.81), (13.82), and (13.83), the Hessian’svalue is

Hpdp,u =

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

∂h1∂p1

∂h1∂p2

· · · ∂h1∂pn

∂h1∂u

∂h2∂p1

∂h2∂p2

· · · ∂h2∂pn

∂h2∂u

......

. . ....

...

∂hn∂p1

∂hn∂p2

· · · ∂hn∂pn

∂hn∂u

∂λ∗pd

∂p1

∂λ∗pd

∂p2· · · ∂λ∗pd

∂pn

∂λ∗pd

∂u

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠

. (13.85)

Because this matrix is necessarily negative-semi-definite, the main diagonal elementsare all nonpositive, informing us that

∂hi(p, u)

∂pi≤ 0 for all i = 1, . . . , n. (13.86)

That is, any compensated demand curve slopes downwards or, equivalently, all pureown-price substitution effects are nonpositive.

If the consumer’s expenditure function is twice-continuously differentiable, thenthe matrix Hpd

p,u is symmetric:

∂hi(p, u)

∂pj=∂hj(p, u)

∂pifor all i, j = 1, . . . , n; i = j. (13.87)

This is a statement that the rates of cross-price pure-substitution effects betweencommodities i and j are the same. Symmetry also shows that the rate of changewith respect to pi of the rate of change of the consumer’s minimum expenditure withrespect to u is the same as the rate of change with respect to u of the consumer’sHicksian demand for commodity i:

∂pi

(∂e(p, u)

∂u

)=∂hi∂u

for i = 1, . . . , n.

The Slutsky matrix is the matrix of second-order partial derivatives with respect

Page 488: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.11. ANSWERS 469

to prices of the consumer’s expenditure function:

S(p, u) =

(∂2e(p, u)

∂pi∂pj

)=

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

∂h1∂p1

∂h1∂p2

· · · ∂h1∂pn

∂h2∂p1

∂h2∂p2

· · · ∂h2∂pn

......

. . ....

∂hn∂p1

∂hn∂p2

· · · ∂hn∂pn

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠. (13.88)

It is the matrix of the rates of change with respect to prices of Hicksian (compensated)quantities demanded. These are the slopes ∂hi/∂pi of Hicksian demand curves andthe rates of pure-substitution effects ∂hi/∂pj for i = j. The famous comparativestatics result that the Slutsky matrix is negative-semi-definite for small feasibility-preserving perturbations of p is easily obtained from (13.85), since Hpd

p,u is negative-semi-definite.

Answer to Problem 13.3. The original problem is

minx∈�n

p·x subject to U(x) ≥ u.

The optimal solution is the compensated demand vector (h1(p, u), . . . , hn(p, u)). Theminimum-value function for the problem is the consumer’s expenditure function

e(p, u) ≡ p1h1(p, u) + · · ·+ pnhn(p, u) + λ∗(p, u)(U(h(p, u))− u). (13.89)

Choose a price vector p0 � 0 and a utility level u0. Then the expenditure-minimizing consumption bundle is (h1(p

0, u0), . . . , hn(p0, u0)). Let h01 denote the nu-

merical value h1(p0, u0). The new problem is the original with the additional con-

straint x1 = h01; i.e.

minx∈�n

p·x subject to U(x) ≥ x and x1 = h01.

The additional constraint is just-binding when (p1, . . . , pn) = (p01, . . . , p0n), and u = u0.

The solution is the compensated demand vector hr(p, u) = (hr1(p, u), . . . , hrn(p, u)),

where the superscript r denotes the presence of the extra restriction. The minimum-value function for the new problem is the expenditure function

er(p, u;h01) ≡ p1hr1(p, u) + p2h

r2(p, u) + · · ·+ pnh

rn(p, u)

+ λ∗r(p, u)(U(hr(p, y))− u) + μr(p, u)(hr1(p, u)− h01).(13.90)

Page 489: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

470 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

The crucial facts are that

e(p0, u0) = er(p0, u0;h01) and e(p, u) < er(p, u;h01) for (p, u) = (p0, u0). (13.91)

That is, e(p, u) is the lower envelope formed from the family of functions er(p, u;h01)by minimizing er(p, u;h01) with respect to h01.

For small enough ε and any (p, u) ∈ Bde((p0, u0), ε), an order-2 Taylor’s series can

be used to write

e(p, u) = e(p0, u0) +∇e(p0, u0)·(p− p0, u− u0)

+1

2(p− p0, u− u0)Hp,u(p

0, u0)

(p− p0

u− u0

),

(13.92)

where Hp,u(p0, u0) is the value of the Hessian matrix for e at (p, u) = (p0, u0). Simi-

larly,

er(p, u) = er(p0, u0) +∇er(p0, u0)·(p− p0, u− u0)

+1

2(p− p0, u− u0)Hr

p,u(p0, u0)

(p− p0

u− u0

),

(13.93)

where Hrp,u(p

0, u0) is the value of the Hessian matrix for er at (p, u) = (p0, u0).

Combining (13.91), (13.89), and (13.93) shows that, necessarily,

∇e(p0, u0)·(p− p0, u− u0) +1

2(p− p0, u− u0)Hp,u(p

0, u0)

(p− p0

u− u0

)

< ∇er(p0, u0)·(p− p0, u− u0) +1

2(p− p0, u− u0)Hr

p,u(p0, u0)

(p− p0

u− u0

) (13.94)

for any (p, u) ∈ BdE((p0, u0), ε).

The kth element of the gradient vector of e is the partial derivative of e withrespect to pk. From (13.89) and the First-Order Envelope Theorem, this is

∂e(p, u)

∂pk= hk(p, u). (13.95)

The last element of the gradient vector of e is the partial derivative of e with respectto u. From (13.89) and the First-Order Envelope Theorem, this is

∂e(p, u)

∂u= λ∗(p, u). (13.96)

Page 490: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.11. ANSWERS 471

The kth element of the gradient vector of er is the partial derivative of er with respectto pk. From (13.89) and the First-Order Envelope Theorem, this is

∂er(p, u)

∂pk= hrk(p, u). (13.97)

The last element of the gradient vector of er is the partial derivative of er with respectto u. From (13.89) and the First-Order Envelope Theorem, this is

∂er(p, u)

∂u= λ∗r(p, u). (13.98)

When (p, u) = (p0, u0), the constraint x1 = h01 is only just binding, so μ∗r(p0, u0) = 0and, particularly, λ∗(p0, u0) = λ∗r(p0, u0) and h(p0, u0) = hr(p0, u0). Thus, from(13.93), (13.96), (13.97), and (13.98),

∇e(p, u) = ∇er(p, u) at (p, u) = (p0, u0) (13.99)

and so, from (13.94) and (13.99), necessarily

(p− p0, u− u0)Hp,u(p0, u0)

(p− p0

u− u0

)< (p− p0, u− u0)Hr

p,u(p0, u0)

(p− p0

u− u0

).

Rearranged, this is

(p− p0, u− u0)(Hp,u(p

0, u0)−Hrp,u(p

0, u0))(p− p0

u− u0

)< 0 (13.100)

for (p, u) ∈ BdE((p0, u0), ε); i.e. the matrix Hp,u(p

0, u0) − Hrp,u(p

0, u0) necessarily isnegative-definite (not merely semi-definite).

The Hessian matrix of e is

Hp,u =

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

∂2e

∂p21· · · ∂2e

∂p1∂pn

∂2e

∂p1∂u...

. . ....

...

∂2e

∂pn∂p1· · · ∂2e

∂p2n

∂2e

∂pn∂u

∂2e

∂u∂p1· · · ∂2e

∂u∂pn

∂2e

∂u2

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠

=

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

∂h1∂p1

· · · ∂h1∂pn

∂h1∂u

.... . .

......

∂hn∂p1

· · · ∂hn∂pn

∂hn∂u

∂λ∗

∂p1· · · ∂λ∗

∂pn

∂λ∗

∂u

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠

(13.101)

Page 491: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

472 CHAPTER 13. COMPARATIVE STATICS ANALYSIS, PART II

from (13.95) and (13.96). Similarly, the Hessian matrix of er is

Hrp,u =

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

∂2er

∂p21· · · ∂2er

∂p1∂pn

∂2er

∂p1∂u...

. . ....

...

∂2er

∂pn∂p1· · · ∂2er

∂p2n

∂2er

∂pn∂u

∂2er

∂u∂p1· · · ∂2er

∂u∂pn

∂2er

∂u2

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠

=

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

∂hr1∂p1

· · · ∂hr1∂pn

∂hr1∂u

.... . .

......

∂hrn∂p1

· · · ∂hrn∂pn

∂hrn∂u

∂λ∗r

∂p1· · · ∂λ∗r

∂pn

∂λ∗r

∂u

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠

(13.102)

from (13.97) and (13.98). From (13.101) and (13.102),

Hp,u(p0, u0)−Hr

p,u(p0, u0) =

⎛⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎝

∂h1∂p1

− ∂hr1∂p1

· · · ∂h1∂pn

− ∂hr1∂pn

∂h1∂u

− ∂hr1∂u

.... . .

......

∂hn∂p1

− ∂hrn∂p1

· · · ∂hn∂pn

− ∂hrn∂pn

∂hn∂u

− ∂hrn∂u

∂λ∗

∂p1− ∂λ∗r

∂p1· · · ∂λ∗

∂pn− ∂λ∗r

∂pn

∂λ∗

∂u− ∂λ∗r

∂u

⎞⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎠,

(13.103)where every term in the matrix is evaluated at (p0, u0).

The main diagonal elements of this matrix are necessarily strictly negative, so,necessarily,

∂hi∂pi

|(p,u)=(p0,u0) <∂hri∂pi

|(p,u)=(p0,u0) for i = 1, . . . , n (13.104)

and∂λ∗

∂u|(p,u)=(p0,u0) <

∂λ∗r

∂u|(p,u)=(p0,u0). (13.105)

(13.104) states that the rate at which the consumer changes her Hicksian quantitydemanded of commodity i as pi rises is more negative when the consumer is notsubject to the additional restriction h1 ≡ h01. Simply put, when the consumer isless restricted in her choices of consumption bundles, she is better able to reducethe adverse (higher expenditure) consequence of a larger pi, and does so by makingstrictly greater (Hicksian) substitutions compared to when she is subject to the extrarestriction of consuming a fixed quantity of commodity 1. This is what Le Chatelierwould have predicted.

Page 492: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

13.11. ANSWERS 473

What does (13.105) say? The quantities ∂λ∗/∂u and ∂λ∗r/∂u are the reciprocalsof the rates of change of the consumer’s marginal utilities of expenditure with respectto the utility level achieved. These are both positive quantities, so (13.105) statesthat the consumer’s marginal utility of expenditure increases at a slower rate, or,more likely, decreases at a faster rate when the consumer is subject to the additionalrestriction. This should be intuitively clear. If the consumer is more restricted inhow she can use an extra dollar of disposable income, then that extra dollar will notprovide her as much extra utility. This too is what Le Chatelier would have predicted.

If the two expenditure functions e and er are twice-continuously differentiable,then the matrices Hp,u and Hr

p,u are symmetric and so is the matrix Hp,u − Hrp,u.

Hence∂hi∂pj

− ∂hri∂pj

=∂hj∂pi

−∂hrj∂pi

for all i, j = 1, . . . , n; i = j. (13.106)

This states that, necessarily, the difference in the rates of the less and more restrictedcross-price pure-substitution effects for commodity i with respect to pj is the same asthe difference in the rates of the less and more restricted cross-price pure-substitutioneffects for commodity j with respect to pi. Similarly, for what it is worth, necessarily,

∂hi∂u

− ∂hri∂u

=∂λ∗

∂pi− ∂λ∗r

∂pifor i = 1, . . . , n. (13.107)

Page 493: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

Bibliography

Arrow, K., and A. Enthoven. 1961. “Quasi-Concave Programming.” Econometrica 29(4): 779–800.

Chiang, A., and K. Wainwright. 2005. Fundamental Methods of Mathematical Eco-nomics. McGraw-Hill.

Debreu, G. 1952. “Definite and Semidefinite Quadratic Forms.” Econometrica 20 (2):295–300.

Franklin, J. 2002. Methods of Mathematical Economics: Linear and Nonlinear Pro-gramming, Fixed-Point Theorems. Society for Industrial and Applied Mathemat-ics.

Hoy, M., J. Livernois, C. McKenna, R. Rees, and T. Stengos. 2011. Mathematics forEconomics. MIT Press.

John, F. 1948. “Extremum Problems with Inequalities as Subsidiary Conditions.” InStudies and Essays, Courant Anniversary Volume, 187–204. Wiley InterscienceNew York.

Karush, W. 1939. “Minima of Functions of Several Variables with Inequalities as SideConstraints.” M.Sc. Thesis, Department of Mathematics, University of Chicago.

Kuhn, H., and A. Tucker. 1951. “Nonlinear Programming.” In Proceedings of 2ndBerkeley Symposium on Mathematical Statistics and Probability, edited by J.Neyman, 481–492. University of California Press, Berkeley.

Milgrom, P., and J. Roberts. 1996. “The Le Chatelier Principle.” American EconomicReview 86 (1): 173–179.

Silberberg, E. 1974. “A Revision of Comparative Statics Methodology in Economics,or, How to Do Comparative Statics on the Back of an Envelope.” Journal ofEconomic Theory 7:159–172.

Simon, C., and L. Blume. 1994. Mathematics for Economists. W. W. Norton and Co.

474

Page 494: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

Index

actionmixed, 362pure, 361

addition of convex sets, 30addition of sets, 22affine cone, 217

ball, 111Banach space, 144Banach’s Fixed-Point Theorem, 143basis

topological, 107vector, 11

Berge’s Theorem of the Maximum, 353best-response correspondence, 364bijective mapping, 149block form, 438block form of a matrix, 322Bolzano-Weierstrass Theorem, 130bordered Hessian matrix, 321bordered principal minor, 323

Cauchy sequence, 132closed graph, 172closed graph in a metric space, 173Closed Graph Theorem, 173closure of a set, 196codomain of a mapping, 146cofactor, 18comparative statics analysis, 373, 374comparative statics experiment, 373

complementary slackness condition, 245

concatenation, 343

concave programming problem, 294

cone, 216

affine, 217

convex, 218

constrained direction, 259

constrained optimization problem, 238

concave programming, 294

primal, 432

primal-dual, 433

quasi-concave programming, 296

typical, 241

constrained path, 261

constraint qualification, 247

Karlin, 257

Kuhn-Tucker, 263

rank, 258

Slater, 257

constraint qualification and Farkas’s Lem-ma, 255

continuous correspondence, 165

continuous function, 156

continuous mapping, 151

contour set, 46, 177

contraction mapping, 139

convergence, 125, 132

convex combination, 26

convex cone, 14, 218

convex hull, 219

475

Page 495: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

476 INDEX

convex set, 26correspondence, 146

best-response, 364closed graph, 172closed-valued, 172compact-valued, 169continuous, 165domain, 147feasible solution, 344graph of, 172lower-hemi-continuous, 163optimal solution, 345range, 148upper-hemi-continuous, 162, 169

critical point, 70cusp, 247

derivativepartial, 33total, 38

determinant, 17differentiability

continuous, 44continuous to order r, 45to order r, 44

direct effect, 393direct product of convex sets, 30, 87direct product of sets, 20

envelopelower, 408upper, 400, 408

Envelope TheoremFirst-Order, 392, 402, 437First-Order Lower, 409First-Order Upper, 409Second-Order, 446, 451, 452, 454

epsilon neighborhood, 111

existence of an optimal solution, 347

Farkas’s alternatives, 221

Farkas’s Lemma, 220

Farkas’s Lemma and constraint qualifica-tion, 255

feasible choice, 238

feasible set, 238

feasible solution correspondence, 344

First-Order Envelope Theorem, 392

First-Order Lower Envelope Theorem, 409

first-order necessary conditions, 242

primal-dual problem, 434

First-Order Upper Envelope Theorem, 409

fixed-point, 144

Fritz John’s necessary condition, 250

Fritz John’s Necessity Theorem, 250

full rank matrix, 20

function, 146

affine, 85

continuous, 152, 156

contour set, 177

domain, 147

indirect profit, 347

indirect utility, 346

jointly continuous, 175

Lagrange, 284

Lagrange, restricted, 288

linear, 85

lower contour set, 177

lower-semi-continuous, 155, 177

maximum-value, 346, 392, 409

restricted, 399

minimum-value, 409

objective, 239

quasi-convex, 296

range, 147

Page 496: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

INDEX 477

restricted, 342upper contour set, 177upper-semi-continuous, 155, 177

globally optimal solution, 240Gordan’s alternatives, 225Gordan’s Lemma, 224gradient vector, 50, 55graph of a correspondence, 172graph of a mapping, 149

half-space, 193lower, 193upper, 193

Heine-Borel Theorem, 116Hessian matrix, 70Hotelling’s Lemma, 403, 411, 443hyperplane, 192

half-space, 193level, 192normal, 192supporting, 196

image of a correspondence, 148image of a function, 147Implicit Function Theorem

equation system, 379single equation, 377

implicit solution, 375indirect effect, 393indirect profit function, 347indirect utility function, 346infimum, 125, 132injective mapping, 149inner product, 51, 55intersection of convex sets, 29irregular feasible point, 247

Jacobian matrix, 259

joint continuity, 175jointly continuous function, 175just-binding constraint, 448, 449

Karlin’s constraint qualification, 257Karush-Kuhn-Tucker necessary condition,

245Karush-Kuhn-Tucker Necessity Theorem,

245, 286Kuhn-Tucker constraint qualification, 263Kuhn-Tucker Saddle-Point Theorem, 291

Lagrange function, 284restricted, 288, 291, 309

Lagrange function saddle-point, 288Lagrange multiplier, 284Lagrange’s Critical Point Theorem, 285Le Chatelier’s Principle, 456leading principal

minorand definiteness, 68

minor, 68submatrix, 68

lim inf, 130lim sup, 130limit inferior, 130limit of a sequence, 132limit superior, 130linear

combination, 10dependence, 12independence, 12

locally optimal solution, 240lower contour set, 177lower envelope, 408lower hemi-continuity, 161lower-inverse of a mapping, 161lower-semi-continuous function, 155, 177

Page 497: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

478 INDEX

mapping, 145

bijective, 149

codomain, 146

continuous, 151

contraction, 139

correspondence, 146

domain of a correspondence, 147

domain of a function, 147

function, 146

graph of, 149

image of a correspondence, 148

image of a function, 147

injective/one-to-one, 149

inverse image, 151

lower inverse image, 161

lower-hemi-continuous, 161

surjective/onto, 148

upper inverse image, 160

upper-hemi-continuous, 161

marginal rate of substitution, 47, 55

matrix, 14

block form, 322

bordered Hessian, 321

full rank, 20

Hessian, 70

Jacobian, 259

rank, 16

Slutsky, 468

square, 17

maximum-value function, 346, 392, 409

restricted, 399

metric, 79

Euclidean, 80

space, 82

space, Euclidean, 82

sup, 81

triangular inequality, 79

metrizable space, 152

minimum-value function, 409

Minkowski’s Separating Hyperplane The-orem, 203

minor, 18

bordered principal, 323

successive principal bordered, 329

mixed action, 362

mixed action space, 362

Nash conjecture, 363

Nash game, 361

natural numbers, 123

necessary and sufficient conditions, 302

necessary optimality condition

second-order, 304, 306, 308, 309, 311

negative of a set, 23

norm, 82

objective function, 239

one-to-one mapping, 149

onto mapping, 148

optimal solution, 240

optimal solution correspondence, 345

parameter space, 343, 344

partial derivative, 33

point of support, 196

power set, 106, 147

primal problem, 432

primal-dual problem, 433

pure action, 361

quadratic approximations of functions, 71

quadratic form, 61

definiteness, 64, 68

quasi-concave programming problem, 296

quasi-convex function, 296

Page 498: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

INDEX 479

range of a correspondence, 148

range of a function, 147

rank constraint qualification, 258

rank of a matrix, 16

ray, 217

real vector, 7

reference set, 104

regular feasible point, 247

Roy’s Identity, 413

saddle-point, 76, 78

x-max λ-min, 288

x-min λ-max, 288

Lagrange function, 288

Second-Order Envelope Theorem, 454

second-order necessary condition, 304, 311

geometry of, 304

primal-dual problem, 437

testing, 320, 325

second-order sufficiency condition

testing, 329, 330

separation of sets, 193

proper, 194

strict, 195

strong, 196

sequence, 123

bounded, 124

Cauchy, 132

convergent, 125, 132

finite, 124

infinite, 124

limit, 125, 132

monotonic, 128

subsequence, 129

sets

addition of, 22

addition of convex, 30, 86

bounded, 114

closed, 109

closure, 196

compact, 115, 119

complementary, 109

contour, 46

convex, 26, 87

direct product of, 20

direct product of convex, 30, 87

feasible, 238

intersection of bounded, 119

intersection of compact, 119

intersection of convex, 29, 86

negative of, 23

open, 109

separation, 193

proper, 194

strict, 195

strong, 196

strictly convex, 28

union of bounded, 119

union of compact, 119

Shephard’s Lemma, 412, 463

Silberberg’s method, 431

Slater’s constraint qualification, 257

Slater’s example, 249, 254, 265

Slutsky matrix, 468

Slutsky’s equation, 413

solution

feasible, 238

optimal

global, 240

local, 240

space, 6

complete, 137

complete metric, 137

normed, 83

Page 499: Beyond the Nation?: Immigrants’ Local Lives in Transnational Cultures

480 INDEX

parameter, 344

proper subspace, 86

subspace, 85

topological, 109

trivial vector, 93

vector, 7

spanning, 11

Spanning Theorem, 12

statement of necessity, 303

statements of sufficiency, 304

strictly convex set, 28

subsequence, 129

monotonic, 129

subspace, 85

proper, 86

successive principal bordered minor, 329

sufficiency statement for local optimality,326–328

sufficient optimality conditions, 326

supremum, 125

surjective mapping, 148

Taylor quadratic, 74

Taylor remainder, 74

Taylor’s Theorem, 74

testing the second-order condition, 320

testing the second-order necessary condi-tion, 325

testing the second-order sufficiency condi-tion, 329

theorem of the alternative, 220

Theorem of the Maximum, 353

topology, 105

basis, 107

coarse, 106

discrete, 107

Euclidean metric, 113

finer, 118generation from a basis, 107indiscrete, 105metric space, 111refined, 106, 118topological space, 109

total derivative, 38

unit simplex, 362upper contour set, 177upper envelope, 400, 408upper inverse of a mapping, 160upper-hemi-continuity, 161, 169upper-hemi-continuity in metric spaces, 170upper-semi-continuous function, 155, 177

vectorgradient, 50, 55real, 7

vector space, 7trivial, 93

Weierstrass’s Theoremcontinuous function, 178lower-semi-continuous function, 178upper-semi-continuous function, 178