117
OPTIMALITY OF THE TRANSLATION BOUNDS FOR LINEAR CONDUCTING COMPOSITES IN TWO AND THREE DIMENSIONS by Nathan Lee Albin A dissertation submitted to the faculty of The University of Utah in partial fulfillment of the requirements for the degree of Doctor of Philosophy Department of Mathematics The University of Utah May 2006

OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

OPTIMALITY OF THE TRANSLATION BOUNDS

FOR LINEAR CONDUCTING COMPOSITES IN

TWO AND THREE DIMENSIONS

by

Nathan Lee Albin

A dissertation submitted to the faculty ofThe University of Utah

in partial fulfillment of the requirements for the degree of

Doctor of Philosophy

Department of Mathematics

The University of Utah

May 2006

Page 2: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

Copyright c© Nathan Lee Albin 2006

All Rights Reserved

Page 3: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

THE UNIVERSITY OF UTAH GRADUATE SCHOOL

SUPERVISORY COMMITTEE APPROVAL

of a dissertation submitted by

Nathan Lee Albin

This dissertation has been read by each member of the following supervisory committeeand by majority vote has been found to be satisfactory.

Chair: Andrej Cherkaev

Aaron Fogelson

Graeme Milton

Klaus Schmitt

Andrejs Treibergs

Page 4: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

THE UNIVERSITY OF UTAH GRADUATE SCHOOL

FINAL READING APPROVAL

To the Graduate Council of the University of Utah:

I have read the dissertation of Nathan Lee Albin in its final formand have found that (1) its format, citations, and bibliographic style are consistent andacceptable; (2) its illustrative materials including figures, tables, and charts are in place;and (3) the final manuscript is satisfactory to the Supervisory Committee and is readyfor submission to The Graduate School.

Date Andrej CherkaevChair, Supervisory Committee

Approved for the Major Department

Aaron BertramChair/Dean

Approved for the Graduate Council

David S. ChapmanDean of The Graduate School

Page 5: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

ABSTRACT

We consider the long-standing problem of characterizing the set of all possible

effective tensors of two- and three-dimensional composites made of several isotropic

linearly conducting phases in prescribed volume fractions. For more than two

phases, a complete characterization of this set, the G-closure, is not known. The

results presented in this dissertation follow from and expand upon the work of a

number of people including Hashin and Shtrikman; Lurie and Cherkaev; Tartar;

Murat and Tartar; Milton; Milton and Kohn; and Gibiansky and Sigmund.

The “translation bound,” a generalization of the Hashin-Shtrikman bound,

consists of several inequalities that the effective tensors of any composite must

satisfy. The bound depends only on the G-closure parameters — the conductivities

of the phases and the relative volume fractions. It is independent of the layout of

the phases in the composite. The inequalities are known to be optimal for certain

parameters. It is also known that there exist parameters for which they are not

optimal. However, a complete characterization of the parameter set for which the

bound is optimal is still missing.

We use a systematic approach based on “field optimality conditions”. Using this

approach, we consider the general, anisotropic bound and prove that it is optimal

for a much larger range of parameters than was previously known. We do this

for three or more phases and in two or three dimensions by constructing laminate

composites that saturate one of the inequality bounds.

We illustrate a number of applications of this approach, including finite-rank

iterated laminates, special “block structures” resembling those of Gibiansky and

Sigmund, and infinite-rank laminates produced via a differential scheme. Finally,

we discuss a numerical algorithm that gives an approximation of the G-closure in

regions where the translation bounds may not be optimal.

Page 6: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

In loving memory of my mother,

Jana Dee Albin

Page 7: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

CONTENTS

ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

ACKNOWLEDGEMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi

CHAPTERS

1. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1.1 Why multimaterial composites? . . . . . . . . . . . . . . . . . . . . . . . . . 11.1.2 Why a systematic approach? . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.1.3 Why field optimality conditions? . . . . . . . . . . . . . . . . . . . . . . . . 2

1.2 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31.3 Multimaterial conducting mixtures: problem and notations . . . . . . . 7

1.3.1 The G-closure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2. SEQUENTIAL LAMINATES IN TWO DIMENSIONS . . . . . . . 10

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102.2 Known bounds on the G-closure . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.2.1 Outer bounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152.2.2 Inner bounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.2.2.1 Two-material optimal structures . . . . . . . . . . . . . . . . . . . . 172.2.2.1.1 The lamination formula. . . . . . . . . . . . . . . . . . . . . . . 18

2.2.2.2 Multimaterial optimal structures . . . . . . . . . . . . . . . . . . . . 182.2.2.2.1 Milton’s structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . 182.2.2.2.2 Lurie-Cherkaev multicoated spheres. . . . . . . . . . . . . . 192.2.2.2.3 Optimal structures for the Wiener bounds. . . . . . . . . 202.2.2.2.4 Milton-Kohn Matrix laminates. . . . . . . . . . . . . . . . . . 202.2.2.2.5 Gibiansky-Sigmund isotropic structures. . . . . . . . . . . 21

2.3 New optimal structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242.3.1 A convenient change of variable . . . . . . . . . . . . . . . . . . . . . . . . . 242.3.2 T-structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252.3.3 Coating preserves optimality . . . . . . . . . . . . . . . . . . . . . . . . . . . 262.3.4 Example: the optimality of two-material matrix laminates . . . . 272.3.5 Coated T-structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 292.3.6 T2-structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312.3.7 The set of optimal structures . . . . . . . . . . . . . . . . . . . . . . . . . . . 332.3.8 The problem with fixed volume fractions . . . . . . . . . . . . . . . . . . 34

Page 8: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

2.3.9 Applicability: volume fractions . . . . . . . . . . . . . . . . . . . . . . . . . 362.3.10 An inner bound of the G-closure . . . . . . . . . . . . . . . . . . . . . . . . 38

2.4 Fields in optimal structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 392.4.1 Local fields required by the translation bound . . . . . . . . . . . . . 392.4.2 Rank-one connection and the fields in the optimal structures . . 432.4.3 Constructing optimal laminates . . . . . . . . . . . . . . . . . . . . . . . . . 452.4.4 The general structure revisited . . . . . . . . . . . . . . . . . . . . . . . . . 45

2.4.4.1 The optimal T-structures . . . . . . . . . . . . . . . . . . . . . . . . . 462.4.4.2 The coating principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 492.4.4.3 The optimal T2-structures . . . . . . . . . . . . . . . . . . . . . . . . . 50

2.4.5 Four and more materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 512.5 A supplementary bound . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

2.5.1 The bound . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 522.5.2 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

2.6 Optimality versus attainability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

3. BLOCK STRUCTURES IN TWO AND THREE DIMENSIONS 61

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 613.2 Background: the structures of Gibiansky and Sigmund . . . . . . . . . . . 62

3.2.1 The structures introduced . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 623.2.2 Separation of variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 633.2.3 The structures revisited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

3.3 New optimal structures in two dimensions . . . . . . . . . . . . . . . . . . . . . 663.4 New optimal structures in three dimensions . . . . . . . . . . . . . . . . . . . 66

3.4.1 The lower translation bound . . . . . . . . . . . . . . . . . . . . . . . . . . . 673.4.2 Optimal structures for the lower bound . . . . . . . . . . . . . . . . . . . 69

3.4.2.1 Separation of variables . . . . . . . . . . . . . . . . . . . . . . . . . . . 743.4.2.2 Applicability: volume fractions . . . . . . . . . . . . . . . . . . . . . 75

3.4.3 The upper translation bound . . . . . . . . . . . . . . . . . . . . . . . . . . . 763.4.4 Optimal structures for the upper bound . . . . . . . . . . . . . . . . . . 77

4. A DIFFERENTIAL SCHEME: INFINITE-RANK LAMINATES 78

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 784.2 The differential scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

4.2.1 The equations for the volume fractions . . . . . . . . . . . . . . . . . . . 814.2.2 The equation for the effective tensor . . . . . . . . . . . . . . . . . . . . . 814.2.3 An optimal control problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

4.3 The modified differential scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . 834.3.1 Matrix laminates in replaced strips . . . . . . . . . . . . . . . . . . . . . . 844.3.2 The optimal controls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

4.4 The optimal structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 874.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

vii

Page 9: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

5. NUMERICALLY ESTIMATING THE G-CLOSURE . . . . . . . . 90

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 905.2 A finite scheme for laminate structures . . . . . . . . . . . . . . . . . . . . . . . 905.3 The optimization problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 925.4 Numerical results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

5.4.1 Numerical estimates of G-closures . . . . . . . . . . . . . . . . . . . . . . . 935.4.2 Numerical estimates of the isotropic bound . . . . . . . . . . . . . . . . 955.4.3 Fields in numerically optimized structures . . . . . . . . . . . . . . . . 95

6. CONCLUSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

viii

Page 10: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

LIST OF FIGURES

2.1 Bounds and optimal points: a summary of previous results. . . . . . . . . 12

2.2 Bounds and optimal points: a summary of new results. . . . . . . . . . . . . 12

2.3 Two-material structures optimal for the translation bound. . . . . . . . . . 17

2.4 Three-material structures optimal for the translation bound. . . . . . . . 19

2.5 Extremal structures of Gibiansky and Sigmund. . . . . . . . . . . . . . . . . . 22

2.6 The general three-material laminate which we optimize. . . . . . . . . . . . 24

2.7 A selection of new optimal structures. . . . . . . . . . . . . . . . . . . . . . . . . . 25

2.8 The set L(KT ) of coated T-structures. . . . . . . . . . . . . . . . . . . . . . . . . . 30

2.9 Optimal points for the lower translation bound for m2

m2+m3= 1

60. . . . . . 34

2.10 Optimal points for the upper translation bound for m2

m1+m2= 1

41. . . . . . 35

2.11 Domains of applicability of Theorem 2.1 and Theorem 2.6 in terms ofvolume fractions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

2.12 Inner and outer bounds on the G-closure for two different sets ofadmissible parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

2.13 Fields in optimal laminate structures. . . . . . . . . . . . . . . . . . . . . . . . . . 47

2.14 The generalized T-structure for four materials . . . . . . . . . . . . . . . . . . . 51

2.15 The structures from Theorem 2.9. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

3.1 The two-dimensional anisotropic optimal block structure. . . . . . . . . . . 67

3.2 The optimal three-dimensional block structure. . . . . . . . . . . . . . . . . . . 73

4.1 The differential scheme for isotropic composites in three dimensions. . 80

5.1 A numerical estimation of the G-closure for M1 = 0.4, M2 = 0.01. . . . 94

5.2 A comparison of numerical results to known inner bounds. . . . . . . . . . 94

5.3 The effective conductivity of numerically optimized isotropic resistorsfor m2 = 0.5(1 − m1). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

5.4 The effective conductivity of numerically optimized isotropic resistorsfor m2 = 0.1(1 − m1). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

5.5 The local fields in the numerically optimized structures used in Fig-ure 5.4 in the case when the bounds are known to be optimal . . . . . . . 97

Page 11: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

5.6 The local fields in the numerically optimized structures used in Fig-ure 5.4 in the case when the bounds are not known to be optimal . . . . 97

x

Page 12: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

ACKNOWLEDGEMENTS

This research was funded by the National Science Foundation through a VIGRE

grant to the University of Utah Department of Mathematics. I thank the NSF and

the Mathematics Department for this support.

It is also a great pleasure to thank the many people who have helped make this

dissertation possible.

I am forever indebted to my advisor, Prof. Andrej Cherkaev. His enthusiasm

for the subject captured me on our first meeting and his phenomenal intuition has

inspired me ever since. Without his frequent direction and advice, without the hours

and hours he spent sharing his knowledge of mathematics with me, without his

encouragement and gentle prodding, this dissertation would not have been possible.

I am likewise indebted to my “Italian mentor,” Prof. Vincenzo Nesi. His passion

for mathematics and his profound generosity are an inspiration to me. His advice,

his knowledge, and his uncanny ability to ask the right questions initiated a major

breakthrough in my research and have had an immeasurable effect in shaping this

dissertation ever since.

I would also like to thank my committee members for their support: Prof.

Aaron Fogelson, for the stimulating classes on fluids and numerical methods; Prof.

Graeme Milton, for sharing his knowledge of composite materials and numerical

algorithms; Prof. Klaus Schmitt, for the encouragement and advice he gave so

many times during my graduate career; and Prof. Andrejs Treibergs, for guiding

me through my first experience with graduate-level differential equations. I could

not have asked for a better or more supportive committee.

I owe special thanks to Prof. Arthur Sagle, for teaching me undergraduate

mathematics, for pointing me toward graduate school and for much advice and

encouragement throughout.

Page 13: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

I am also grateful for the help and encouragement of numerous friends and family

members. In particular my grandpa Gene for the use of a certain Chevy Lumina

during my years as a graduate student, and my grandpa Marvin for some very wise

advice when I needed it badly. I also thank my siblings, Heather, Seth, Joshua,

Hannah and Caleb, for frequent phone calls and many smiles. I am fortunate to

have such a loving and supportive family.

Special thanks go to my parents-in-law and good friends, Paul and Cynthia.

Their constant encouragement helped make this dissertation possible. To my

parents, Rick and Carol, I am eternally grateful. Their loving support and charity

have made my life better in countless ways.

Most importantly, I wish to thank my wife, Courtney, for her love, reassurance

and patience throughout this adventure.

xii

Page 14: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

CHAPTER 1

INTRODUCTION

In this dissertation we introduce a systematic approach for the construction

of optimal multimaterial conducting composites. The approach is based on the

fields optimality conditions associated with the “translation bounds.” Using this

approach we produce new laminate composites with properties beyond what was

previously known possible.

1.1 Motivation

1.1.1 Why multimaterial composites?

One of the key problems in multimaterial design is the characterization of

composite materials — fine-scale mixtures of two or more different materials. Such

mixtures are useful in a variety of physical applications because they tend to

exhibit properties that none of the constituent materials posses individually. While

the theory of homogenization has enjoyed great success in describing the effective

behavior of a given mixture, many very basic questions remain unanswered.

One such question is the characterization of composites formed from several

linear, isotropic, conducting materials with prescribed volume fractions. In both

the cases of two- and three-dimensional composites, a complete characterization

of composites in terms of their effective (homogenized) tensors exists only in the

case of two materials. Much remains unknown for composites of three and more

materials.

In a sense, the two-material problem always yields intuitively expected results.

To make a composite that conducts very well, one wraps the best conductor around

the worst. To make a composite that resists very well, one wraps the best resistor

around the worst. For some intermediate composite, one mixes these two extremes.

Page 15: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

2

Moreover, in the physical world it is unlikely that any composite can truly be

classified as a “two-material” composite. Due to impurities or gaps in the mixture,

one can typically identify at least three materials.

1.1.2 Why a systematic approach?

When a third material is added, the problem becomes much more complicated.

It is still not known how to construct the best conducting isotopic composite of

three conducting materials for all volume fractions of the materials. We show in

this dissertation that, when there are several materials in a composite, the materials

take on more complex roles than “best” and “worst.” A seemingly bad material

may not necessarily “hide” in the center of the composite. Instead it can take a

more active role: helping direct current into an intermediate material, for example.

The added complexity makes finding optimal structures much more difficult.

Thus, the need for a systematic approach becomes clear. Without it, we can

never be sure we are not “missing something.” The goal of this dissertation is to

expand the work toward a general systematic approach to classifying multimaterial

composites. The work presented here is an extension of previous results of Hashin

and Shtrikman (1962); Tartar (1979, 1985); Lurie and Cherkaev (1981, 1982);

Milton (1981); Milton and Kohn (1988); Gibiansky and Sigmund (2000). Moreover,

in some cases this approach poses new interesting and challenging mathematical

questions. We do not pursue this direction in this dissertation. However, our work

makes a more transparent mathematical connection with work in the field referred

to as “partial differential inclusions” (see for example Muller (1999); Dolzmann

(2003); Conti et al. (2005a,b)). While we focus on the problem of linear conductivity

here, we believe that the methods can be extended to a wide range of physical

phenomena.

1.1.3 Why field optimality conditions?

In this dissertation, we discuss several inequality bounds which must be satisfied

by the effective tensor of a multimaterial composite. Each of these bounds has

an associated set of conditions which must be satisfied by the local fields in any

Page 16: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

3

structure attaining the bound. These field optimality conditions are integral to the

derivation of the bounds. Such conditions can be seen, for example, in the works of

Voigt (1928); Reuss (1929) for the elasticity version of the arithmetic and harmonic

mean bounds. In this case, the optimality conditions require respectively that the

stresses and strains are constant throughout a composite that attains the bound.

Similar optimality conditions arise in the construction of the Hashin-Shtrikman

bounds (Hashin and Shtrikman (1962)) and in their anisotropic generalization, the

translation bounds (Lurie and Cherkaev (1982); Tartar (1979, 1985)). Furthermore,

the translation bounds are a specific case of a general class of bounds. Each of

these bounds has associated field optimality conditions which are satisfied by a

composite which is optimal for the bound. We invite the reader to refer to Milton

(2002) Section 25.3 for a discussion of these conditions and their importance in the

general case and to Section 25.5 for a discussion of the conditions for the translation

bound in particular. Field optimality conditions for various bounds have motivated

a number of optimal composites for the multimaterial and related problems. We

refer the reader to Milton (1981, 1986); Avellaneda et al. (1988); Avellaneda and

Milton (1989); Vigdergauz (1989); Nesi and Milton (1991); Avellaneda et al. (1996)

for some examples. These results provide an excellent argument for methods based

on field analysis.

1.2 Overview

In this dissertation, we consider the long-standing problem of bounding the

effective properties of composites made from several isotropic, linearly conduct-

ing materials given in prescribed proportions. The primary contributions of this

dissertation to the field of materials mathematics are as follows.

1. We prove that the so-called “translation bounds” are optimal for a much

larger range of parameters than previously known. We do this for two- and

three-dimensional composites of three or more materials.

2. We introduce a systematic approach to forming multimaterial conducting

composites that realize the translation bound. The approach is driven by

Page 17: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

4

the known “field optimality conditions” associated with the bounds. During

the construction we temporarily remove volume fraction constraints. We

exploit rank-one connection arguments to methodically generate new optimal

composites from old allowing the volume fractions to change as necessary

to keep optimality. Though the volume fractions of the final composite are

unconstrained in this process, the final composite is guaranteed to be optimal

for the bound (which depends on the volume fractions of the composite).

3. We introduce the “coating principle,” which allows us to generate a large

family of optimal composites from a given optimal composite by lamination.

This is a key ingredient in the construction of the optimal composites in this

dissertation.

4. Using (2), we methodically construct a large class of optimal hierarchical

laminates. The algorithm is straightforward and at each step produces an

optimal composite. The final volume fractions of the composite are left free

until the end of the procedure. These optimal laminates prove (1) in two

dimensions and for three or more materials.

5. Using (2), we analyze and then generalize the two-dimensional isotropic “block

structures” of Gibiansky and Sigmund. These structures are characterized by

the fact that the periodicity cell is partitioned into a number of rectangular

blocks and each of these blocks is filled with a composite (with several well-

separated infinitesimal length scales). After the “coating principle” is applied

the resulting anisotropic composites generate an equivalent set of effective

tensors to (4) in the three-material case.

6. Using (2), we construct the three-dimensional analogue of the “block struc-

tures” in (5). These structures along with the “coating principle” prove (1)

in the three-dimensional case for three materials.

7. We introduce a new controllable differential scheme for constructing com-

posites. The scheme differs from previous schemes in that it tracks fields in

Page 18: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

5

each material. We apply (2) to find optimal controls. The scheme produces

a family of composites parameterized by a real parameter µ ≥ 0. With

increasing µ, the family traces the transformation via infinitesimal inclusions

from an initial “seed” material to a final composite. Each of these composites

is optimal for the translation bound (for the appropriate volume fractions).

The scheme is applicable in two and three dimensions and for three or more

materials. The optimal composites it produces are equivalent to the compos-

ites produced in (4)–(6) when applied to the corresponding dimension and

number of materials.

8. In the case of three materials in two dimensions, we introduce a new sup-

plementary bound on the anisotropy of a periodic composite with smooth

interfaces which attains the translation bound. A subfamily of the structures

introduced in (4)–(7) satisfy this bound as equality.

9. We introduce a finite scheme that is used to numerically produce an inner

bound on the G-closure of three-material structures in two dimensions. The

results of the scheme are consistent with the analytical results in this disser-

tation and give an idea of the shape of the G-closure when the bounds are

not optimal.

10. We apply the scheme in (9) to the problem of finding the best resisting

isotropic, three-material composite in two dimensions. The results give an

indication of what the field optimality conditions may be for a bound that

improves the translation bound in this problem.

In the rest of Chapter 1, we introduce the problem and notations of effective

(homogenized) tensors and the G-closure set. This set characterizes all possible

effective tensors of composites made from the constituent materials in prescribed

proportions.

In Chapter 2, we discuss the G-closure problem for conducting materials in two

dimensions. For this problem, we consider the concepts of inner and outer bounds

Page 19: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

6

on the G-closure and the concept of optimal structures for an outer bound. The

“translation bound” is of particular interest in this paper and we give it special

treatment, reviewing what is known about it. In this chapter we introduce new

optimal structures for the translation bound and describe a general algorithm for

producing these structures. The set of effective tensors of the structures produced

by this algorithm is strictly larger than the corresponding set for all previously

known optimal structures.

In Chapter 3, we consider the three-dimensional version of the translation

bounds. We find new optimal structures with a “block” layout analogous to the

two-dimensional structures of Gibiansky and Sigmund (2000). Once again, the

new types of optimal structures produce a larger set of effective tensors than the

previously known optimal structures.

In Chapter 4, we use a modified differential scheme to find new optimal struc-

tures in a different setting. The differential scheme approach allows us to explore

a structure parameterized by a real number µ ≥ 0. The effective tensor of this

parameterized structure is given by a controllable differential equation which de-

scribes how the structure changes infinitesimally with µ. By choosing the controls

correctly, we use the scheme to produce optimal structures equivalent to those of

the previous two chapters.

Finally, in Chapter 5, we use a finite-dimensional analogue of the differential

scheme to numerically optimize a class of high-rank laminates. We consider the

case of three conducting materials in two dimensions. The generalization to more

materials and higher dimensions is obvious. This method gives a numerically

estimated inner bound on the G-closure. We compare the numerical results to

the outer bounds and known inner bounds. Furthermore, we consider the fields in

the numerically optimal isotropic structures and discuss their interesting properties

when the translation bounds are not realized.

Page 20: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

7

1.3 Multimaterial conducting mixtures:problem and notations

In this section, we recall the problems of homogenization and the G-closure,

fixing the notation. References for this section can be found, for example, in Dal

Maso (1993); Braides and Defranceschi (1998); Cherkaev (2000); Allaire (2002);

Milton (2002).

Consider a d-dimensional, periodic, multimaterial structure for d ≥ 2. The unit

periodicity cell Ω = [0, 1]d is partitioned into N disjoint sets Ω1, . . . , ΩN such that

N⋃

i=1

Ωi = Ω.

The relative volume fractions of each part, mi = |Ωi| (where |Ωi| represents the

area of the set Ωi), satisfy

mi ≥ 0 ∀i = 1, . . . , N,N∑

i=1

mi = |Ω| = 1. (1.1)

The Ωi are assumed to be filled by materials with isotropic conductivity tensors

Ki = kiI for i = 1, . . . , N (1.2)

where I is the d-by-d identity matrix. We assume the conductivities are ordered so

that

0 < k1 < · · · < kN . (1.3)

The conductivity equations applied to the periodicity cell are written as

div(K(x)∇u(x)) = 0 in Ω,

Ω

∇u(x) dx = e, x = (x1, . . . , xd) ∈ Rd (1.4)

where K : Ω → K1, . . . , KN is the conductivity tensor defined by

K(x) =

K1 if x ∈ Ω1,...

KN if x ∈ ΩN ,

(1.5)

K1, . . . , KN are given by (1.2), and where e is the prescribed average field induced

by distant external sources.

Page 21: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

8

Assume that the periodicity cell with material layout defined by K(x) is subject

to the average field e. The energy stored in the material is defined as

W (K, e) = infu∈H1

#(Ω)+e·x

Ω

∇u(x) · K(x)∇u(x) dx

where H1#(Ω) is the space of locally H1 functions on R

d which are Ω-periodic and

have zero mean. The infimum is taken over functions that can be split into an affine

part, e · x, plus a periodic oscillating part:

u(x) = e · x + osc(x),

Ω

∇u(x) dx = e.

Notice that the affine part, e · x, is prescribed by the loading. The minimization is

taken over the variable, oscillating part, osc(x).

The structure defined by the partition Ωi is associated with its effective tensor

Keff, the conductivity tensor of homogeneous material that stores the same energy

as the mixture under the same homogeneous loading. That is,

e · Keff e = infu∈H1

#(Ω)+e·x

Ω

∇u(x) · K(x)∇u(x) dx ∀e ∈ Rd.

In order to completely determine Keff, it suffices to consider the response of the

same structure to d orthogonal loadings

e = rjej, j = 1, . . . , d (1.6)

where r1, . . . , rd ∈ R are the magnitudes of the loadings and e1, . . . , ed is the

canonical basis of Rd. If we assume that Keff is oriented so that its eigenvectors

are e1, . . . , ed, then the response refers to the sum of the energies of the separate

loadings:d∑

j=1

W (K, rjej) =

d∑

j=1

λjr2j , ∀r1, . . . , rd ∈ R (1.7)

where λ1, . . . , λd are the eigenvalues of Keff. This functional can be conveniently

rewritten in terms of two-by-two matrices. We write E = diag(r1, . . . , rd). Given

Page 22: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

9

any d potentials U = (u1, . . . , ud), we define the gradient matrix as the matrix

whose rows are the gradients of the uj:

DU = (DUij), DUij =∂ui

∂xj∀i, j = 1, . . . , d.

The sum of energies (1.7) can then be written as

W(K, E) = infU∈H1

#(Ω)d+Ex

Ω

〈DU(x) K(x), DU(x)〉 dx (1.8)

where 〈·, ·〉 is the inner product defined on two-by-two matrices by

〈A, B〉 = tr(ABT ).

The effective tensor Keff is the unique (symmetric) tensor satisfying the relation

〈E Keff, E〉 = infU∈H1

#(Ω)d+Ex

Ω

〈DU(x) K(x), DU(x)〉 dx ∀E ∈ Rd×d.

1.3.1 The G-closure

The G-closure was introduced by Lurie and Cherkaev (1981). We think of the

volume fractions, mi = |Ωi|, and the material properties ki as parameters of the

problem. If m = (m1, . . . , mN ) and k = (k1, . . . , kN) satisfy (1.1) and (1.3), then we

say m and k are admissible parameters for the G-closure problem. The closure of

the set of all possible Keff available for parameters m and k is called the G-closure.

Specifically, the G-closure G(m; k) is defined as

G = G(m; k) = Keff : K as in (1.5); |Ωi| = mi, i = 1, . . . , N.

Observe that the variable in this problem is the partition of Ω into the Ωi, since

K and thus Keff are defined through the Ωi. Each partition defines a material

structure through (1.5). The question is: what is the set of possible Keff which can

be obtained by some partition into Ωi subject to the volume constraints |Ωi| = mi?

Note that for a given K∗ ∈ G(m; k), there need not exist a partition of Ω into Ωi

and associated K as in (1.5) such that Keff = K∗. Indeed, one often shows that a

particular K∗ lies in the G-closure by finding a sequence of structures Kǫ such

that Kǫeff → K∗. We use the notation Keff versus K∗ to distinguish these concepts.

Page 23: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

CHAPTER 2

SEQUENTIAL LAMINATES IN TWO

DIMENSIONS

2.1 Introduction

In this chapter, we discuss new results for the long-standing problem of the

optimality of the so-called “translation bounds”1 on the effective properties of

two-dimensional composites of several isotropic conducting materials in prescribed

proportions. The bounds are known not to be optimal for all choices of parameters

— the constituent conductivities and prescribed proportions. However, it is still

not clear what the conditions of optimality are. In this chapter, we show that

the bounds are optimal for a much larger range of parameters than was previously

known.

The primary thrust of the chapter is two-fold. First, we introduce new optimal

structures which prove that the translation bounds are optimal for anisotropic

composites outside the region of parameters previously known. Second, we in-

troduce a general algorithm for constructing these structures and understanding

their optimality by looking at sufficient conditions on the local electrical fields.

In Section 2.2, we discuss known bounds on the G-closure, G, prior to the

results of this chapter. The set G can be associated with a compact subset of R2

by mapping tensors in G to their eigenvalue pairs. The bounds are divided into

two categories: inner bounds and outer bounds. An outer bound is a subset of the

plane B ⊇ G. The outer bound we discuss in this chapter is the intersection of the

Wiener bound (also known as the harmonic and arithmetic mean bound) and the

1The bounds are also frequently referred to as the Hashin-Shtrikman bounds, as they generalizethe pioneering work of Hashin and Shtrikman (1962).

Page 24: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

11

translation bound discussed in Section 2.2.1. An inner bound on the other hand, is

a subset of the plane L ⊆ G. Inner bounds are constructed by exhibiting a set of

structures whose effective tensors form the set L. In this chapter, we do not attempt

to improve the outer bound in is full generality (though we do prove a limit on the

optimality of the bound for a special class of structures in Section 2.5). Instead,

we produce a new inner bound which includes previously known inner bounds on

the G-closure.

In Section 2.3, we construct this improved inner bound, using a class of finite-

but high-rank laminate structures. In particular, we extend the known region of

optimality of the translation bound. For definiteness, we primarily discuss the

lower bound defined by (2.4) for three-material mixtures. The upper bound is

dealt with analogously. The generalization to four and more materials is discussed

in Section 2.4.5. Generalizations to other problems such as two-dimensional linear

elasticity and three-dimensional linear conductivity are possible. In particular, we

discuss the problem of three-dimensional conductivity in Chapter 3.

The result of the section is that the “translation bound” is now known to be

optimal in the anisotropic case for a wider range of volume fractions m1, m2, m3

(m1 + m2 + m3 = 1), where mi > 0 is the fixed relative amount of the ith material

in the three-material G-closure problem. We assume the materials have isotropic

conductivity tensors Ki = kiI for i = 1, 2, 3 where I is the identity matrix. We label

the materials so that 0 < k1 < k2 < k3. For the previous results to be applicable

(see Section 2.2.2.2 and Theorem 2.1 in particular), the condition m1 ≥ 2Θ(1−m2)

must be satisfied. The constant Θ is defined in (2.9) and depends only on the

material properties Ki. In this section, we introduce new anisotropic structures

which attain the translation bound (2.4) with an improved applicability condition:

2Θ(√

m2 − m2) < m1 < 1. (2.1)

(Note that√

m2 − m2 < 1 − m2 if m2 ∈ (0, 1).)

As a first illustration of the results described in Section 2.3, consider Fig-

ures 2.1–2.2, which represent the outer bounds and their optimality in the plane of

Page 25: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

12

Figure 2.1. The bounds (2.2)-(2.5) with previously known optimal points (left).A magnification of the upper-left corner (right).

Figure 2.2. The bounds (2.2)-(2.5) including new optimal points (left). Amagnification of the upper-left corner (right).

eigenvalues of an effective tensor K∗. The figures represent the case

k1 = 1, k2 = 2, k3 = 5

and

m1 = 0.4, m2 = 0.01, m3 = 0.59.

We have chosen an example where m2 is small to illustrate the extremeness of the

improvement in this case. While the amount of improvement depends on the values

of the parameters, our results always improve on previously known results if the

volume fractions satisfy (2.1).

Page 26: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

13

First consider Figure 2.1. The thin dashed lines represents the Wiener bounds

(2.2) and (2.3) while the thin curved lines represent the translation bounds (2.4)

and (2.5). The thick portion of the upper bound represents known optimal points

on the bound as proved in Milton and Kohn (1988) (see Theorem 2.1). The single

isotropic point marked on the lower bound is the point proved optimal by Gibiansky

and Sigmund (2000) (see Theorem 2.2). The thick portions of the Wiener bounds

represent the optimal structures described by Cherkaev and Gibiansky (1996) (see

Theorem 2.9). The right side of the figure shows a magnification of the top-left

corner in order to show that the Wiener bounds are tighter than the translation

bounds near this corner and that the Cherkaev-Gibiansky structures do not quite

reach the translation bounds. Figure 2.2 shows the status after the results of the

present chapter are added.

In Section 2.4, we illustrate an algorithm for constructing the structures of

the previous section by examining the local electrical fields in each constituent

material. This approach is quite different from the traditional methods, which

rely on physical intuition and/or numerical optimization to suggest structures for

a given external loading and fixed volume fractions. Instead, the approach offered

in this section leaves the external loading and volume fractions free and instead

uses standard techniques for solving differential inclusions to produce structures

consistent with the pointwise field requirements. In this sense, our approach is

more systematic than traditional methods. After a composite has been produced

in this fashion, a straightforward calculation finds the external loading and volume

fraction constraints for which the structure is optimal. The algorithm resembles

the methods used to solve a number of problems involving gradients supported on

certain sets of matrices. We refer the reader to Muller (1999); Dolzmann (2003);

Conti et al. (2005a,b) for some examples.

Section 2.5 addresses the optimality of the translation bounds from a different

direction. It is known that the bounds cannot be optimal for all values of the

volume fractions. However, the exact conditions for optimality are still unknown.

In this section, we consider the lower bound (2.4) for three materials. We derive a

Page 27: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

14

supplementary inequality bound that a special type of structure (periodic structures

with smooth interfaces) must satisfy if it attains the bound. The supplementary

bound limits the degree of anisotropy of such a structure. In particular, we show

that no periodic structure with smooth interfaces can attain the bound if

m1 < 2Θ(√

m2 − m2)

(compare to (2.1)). While this does not answer the optimality question in general,

we include this section for two reasons. First, the T 2-structures — which we describe

in Section 2.3 — satisfy the bound as equality. Second, the bound is a useful

application of the field analysis discussed in Section 2.4. Indeed, the bound is

obtained exactly by finding an inequality that the fields in the special structures

must satisfy pointwise. The bound can also be applied to finite-rank laminates

using properties of correctors as described in Briane (1994). However, no proof is

currently known for general structures.

2.2 Known bounds on the G-closure

The set G(m; k) is known to be a closed and bounded subset of the two-by-

two symmetric matrices, R2×2sym. It is also rotationally invariant, so it suffices to

consider the projection of the set into the two-dimensional plane of eigenvalues. As

mentioned in the introduction, there is a two-fold strategy to characterize G(m; k):

inner and outer bounds. By an outer bound, we mean a set B = B(m; k) such

that B(m; k) ⊇ G(m; k). On the other hand, by an inner bound, we mean a set

L = L(m; k) such that L(m; k) ⊆ G(m; k). Of course, if the sets L and B can

be constructed so that L = B, the we have characterized the entire G-closure. In

this section, we discuss known inner and outer bounds on the G-closure. As we

mentioned in the introduction, this chapter does not improve the outer bounds in

general, but instead produces a larger inner bound for admissible parameters m

and k satisfying certain conditions.

Page 28: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

15

2.2.1 Outer bounds

In this chapter, we consider the outer bound B defined in the eigenvalue plane by

a set of inequalities. In particular, any tensor K∗ ∈ G(m; k) satisfies the following

inequality bounds which depend only on the parameters m and k. The set B(m; k)

of two-by-two symmetric tensors that satisfy these inequalities forms an outer bound

on the G-closure.

1. The Wiener bounds are

λmin(K∗) ≥

(N∑

i=1

mi

ki

)−1

, (2.2)

λmax(K∗) ≤

N∑

i=1

miki. (2.3)

where λmin(K∗) and λmax(K

∗) are the minimum and maximum eigenvalues of

K∗ respectively.

2. The translation bounds are

tr K∗ − 2k1

det K∗ − k21

≤ 2

N∑

i=1

mi

ki + k1, (2.4)

tr K∗ − 2kN

det K∗ − k2N

≥ 2

N∑

i=1

mi

kN + ki. (2.5)

The translation bounds are not optimal for all values of the parameters m and k.

Intuitively, we see this from the fact that the formulas for the bounds still depend

on k1 (respectively kN) when m1 = 0 (respectively mN = 0). Indeed, for m1 or

mN near 0, there are better bounds (see Talbot et al. (1995); Nesi (1995)). Despite

several results in the area, it is still not known for which parameters the translation

bounds are optimal. In the rest of this chapter, we address this issue and extend

the known range of parameters for which they are.

Page 29: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

16

2.2.2 Inner bounds

In this section we summarize the known inner bounds on the G-closure prior

to this chapter. These bounds are found by proving that certain tensors K∗ lie

in G(m; k). One way to construct such a tensor is to exhibit a partition of Ω

into subsets Ωi with |Ωi| = mi so that with K defined as in (1.5), one has Keff =

K∗. However, such “exact structures” are often difficult to construct. A simpler

method is to construct a sequence of partitions Ωǫiǫ>0 so that |Ωǫ

i | = mi and the

corresponding effective tensors converge: Kǫeff → K∗ as ǫ → 0. We know from

homogenization theory that K∗ is then in G(m; k). This technique is simpler than

that of finding an exact structure because one can use simple formulas for computing

K∗ without explicitly computing any of the Kǫeff. Laminates are a common example

of such structures and are the structures we exploit in this chapter.

We focus on optimal structures — that is, structures (such as laminates) which

produce an effective tensor K∗ ∈ ∂G(m; k). We call the sequence of partitions an

optimal structure and the associated K∗ an optimal point on the G-closure. This

notion is based on the fact that the G-closure can be completely characterized

by its boundary. Indeed, G(m; k) is closed, bounded and simply connected (see

for example Cherkaev (2000)). The trick to finding optimal structures is to use

both inner and outer bounds. If we have L ⊆ G(m; k) ⊆ B and if there exists

K∗ ∈ L∩∂B then K∗ is optimal. Notice that for B defined by (2.2)-(2.4), K∗ ∈ ∂Bif and only if at least one of the inequalities (2.2)-(2.4) is satisfied as equality. To

differentiate these cases, we will say that K∗ is optimal for whichever inequality is

satisfied as equality.

Furthermore, we will sometimes need to distinguish the case when there exists an

effective tensor Keff = K∗ ∈ ∂B from the case when there exists a limit of effective

tensors Kǫeff → K∗ ∈ ∂B. In the former case, we will say the bound B is attainable

at the point K∗. In the latter case, we will say the bound is optimal at the point

K∗. Clearly, B is optimal at any point for which it is attainable. However, there

are examples of bounds which are optimal but not attainable. Section 2.6 discusses

such an example. This distinction is essentially only important in Section 2.5.

Page 30: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

17

2.2.2.1 Two-material optimal structures

The case of two-material optimal structures (N = 2) is completely known due

to the work of Hashin and Shtrikman (1962); Lurie and Cherkaev (1982); Tartar

(1979, 1985). The optimal isotropic structures were constructed by Hashin and

Shtrikman (1962). They used the coated spheres construction (see Figure 2.3a):

a circle filled with K2 surrounded by a concentric annulus filled with K1. When

this construction is placed into an infinite plane with the conductivity K∗, K1 ≤K∗ ≤ K2 and a constant field is applied at infinity, the volume fraction of the circle

in the structure can be chosen to keep the outside field homogeneous. Thus, the

effective conductivity of the structure is K∗. The periodicity cell Ω can then be filled

with infinitely many homeothetic coated circles (on infinitely many length scales).

Hashin and Shtrikman showed that the effective conductivity of such structure is

optimal for the isotropic version of (2.4) (which they discovered as well).

The optimal anisotropic two-material structures were found by Lurie and Cherkaev

(1982) together with the translation bound (2.4) for N = 2 (see Figure 2.3b). The

structures are iterated laminates.

Remark 2.1 The figures in this chapter should be thought of as schematic rep-

resentation of the actual structures. In reality, the laminate structures must have

well-separated scales for the results to apply. For example, in Figure 2.3b, one

should imagine that the vertical strips of K1 and K2 are interleaved at a scale ǫ2,

while the horizontal strips of this laminate with K1 are interleaved at a scale ǫ. The

effective tensor K∗ is obtained by sending ǫ → 0.

= K2

= K1

(a) (b)

Figure 2.3. Two-material structures optimal for the translation bound.

Page 31: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

18

2.2.2.1.1 The lamination formula. Recall the lamination formula of two

materials, KA and KB, with volume fraction of KA equal to m and normal of

lamination given by n. The effective tensor of the laminate is given by

K∗ = L(KA, KB, n, m) = mKA + (1 − m)KB −N (2.6)

where

N = m(1 − m)(KB − KA)n[nT (mKB + (1 − m)KA)n]−1

nT (KB − KA)

The optimal two-material structures use an iteration of this formula (see Tar-

tar (1985)). First K1 and K2 are laminated in some proportion, then this new

“auxiliary” material is laminated with K1 in the direction orthogonal to the first

lamination. The effective properties can be computed via

K∗ = L(K1, L(K1, K2, n1, c1), n2, c2) (2.7)

where n1 and n2 are orthogonal and c1, c2 ∈ [0, 1]. The fractions c1 and c2 are related

by (1 − c1)(1 − c2) = m2. These structures and their dual version (interchanging

K1 with K2 and m1 with m2) describe all optimal structures (they are optimal

for either (2.4) or (2.5)). Both classes of structures degenerate into laminates (for

example, when c1 = 0, c2 = 1 − m2).

Remark 2.2 While the optimal structures described above are sequences whose

effective conductivities converge to a point on the bound, there do exist exact

geometries whose effective conductivities attain each point of the two-dimensional

translation bound. We refer the reader to Vigdergauz (1989, 1999); Grabovsky and

Kohn (1995); Astala and Nesi (2003) for more details.

2.2.2.2 Multimaterial optimal structures

We now recall several optimal multimaterial structures. These structures are

illustrated in Figure 2.4.

2.2.2.2.1 Milton’s structure. The first type of isotropic, multimaterial

(N ≥ 3) structures which were proved optimal for the translation bound were

Page 32: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

19

described by Milton (1981) (see Figure 2.4a). His construction is as follows. The

amount m1 is split into two parts m′1 and m′′

1 so that the coated circles structures

from K1 and K2 (in the proportionsm′

1

m′

1+m2and m2

m′

1+m2, respectively) and from K1

and K3 (in the proportionsm′′

1

m′′

1+m3and m3

m′′

1+m3, respectively) have the same effective

conductivity. Obviously, any mixture of these structures has the conductivity of

each of them. This mixture is optimal for the bound (2.4) (K∗ is isotropic). All

mixtures of this form clearly satisfy

K1 ≤ K∗ ≤ K2

since the mixture of K1 and K2 must lie in this range. Such a construction is

possible as long as there is enough of material k1. Specifically, this construction

requires that

m1 ≥ 2Θ(1 − m2) (2.8)

where Θ is a constant defined below in (2.9). Similar structures are optimal for the

opposite bound (2.5) with K3 taking the role of the “coating,” and K1 and K2 the

inclusions.

2.2.2.2.2 Lurie-Cherkaev multicoated spheres. An inner bound for the

G-closure problem was found in Lurie and Cherkaev (1985) by posing an additional

assumption that the structure is of the type of multicoated circles (see Figure 2.4b)

and then solving the corresponding optimal control problem. Their construction

is geometrically different from Milton’s, but the effective conductivities of both

structures coincide in the range of parameters where (2.8) holds. The structures

are not optimal for the translation bound if (2.8) is violated.

= K3

= K2

= K1

(a) (b) (c)

Figure 2.4. Three-material structures optimal for the translation bound.

Page 33: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

20

2.2.2.2.3 Optimal structures for the Wiener bounds. Cherkaev and

Gibiansky (1996) introduced a class of three-material anisotropic structures with

the property that they are optimal for one of the Wiener bounds (2.2) or (2.3), but

not the other.2 In Section 2.6, we use this construction to illustrate the difference

between the attainability and optimality of an outer bound.

2.2.2.2.4 Milton-Kohn Matrix laminates. The matrix laminates intro-

duced by Milton and Kohn (1988) combine the idea of Milton (1981) with the

two-material anisotropic structures (2.7) (see Figure 2.4c). The amount m1 of K1 is

divided into two parts, which are used to form two different mixtures of materials,

one of K1 and K2 and the other of K1 and K3. The mixtures are chosen to be

optimal for the corresponding two-material G-closure problem and have effective

tensors given by

K∗ = L(K1, L(K1, K2, n1, c1), n2, c2)

K∗′ = L(K1, L(K1, K3, n3, c3), n4, c4)

where

nT1 n2 = 0, nT

3 n4 = 0, c1, c2, c3, c4 ∈ [0, 1].

From the results of two-material structures, we know both K∗ and K∗′ are optimal

structures. Furthermore, if the parameters can be chosen so that K∗ = K∗′, then

the linearity of the bounds (2.4) with respect to the volume fractions allows us

to mix the two constructions together in any way we wish to obtain an opti-

mal three-material structures. These constructions also require that (2.8) hold.

These structures are more general than those previously discussed: they include

anisotropic structures as well as isotropic. The results can be summarized as follows.

Theorem 2.1 (Milton-Kohn) Let m = (m1, m2, m3) and k = (k1, k2, k3) be

admissible parameters for the G-closure problem such that

m1 ≥ 2Θ(1 − m2),

2The analogous two-material structures are geometrically impossible — that is, if one of thebounds is satisfied as equality for some K∗ then so is the other.

Page 34: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

21

where

Θ =k1(k3 − k2)

(k2 + k1)(k3 − k1)≤ 1

2. (2.9)

Then there exists a family of structures with the given volume fractions which are

optimal for (2.4). The effective tensors of this family cover a connected subset of the

translation bound curve which includes the isotropic point. The most anisotropic

structure of this family has an effective tensor with eigenvalues given by

λ1 = νk1 + (1 − ν)k2, λ2 =k1k2

(1 − ν)k1 + νk2

where ν is defined to be

ν = m1 −2Θ

1 − 2Θm3.

Remark 2.3 Consult the thick, solid line in Figure 2.1 for an illustration of the

family structures for the other bound, (2.5). The family in the theorem covers a

segment of the bound near the isotropic point.

Remark 2.4 As in the two-material case, there also exist exact geometries which

attain certain points of the translation bound. In this case, however, it is known

that there exist G-closure parameters for which the bounds are not even optimal.

Once again, we refer the reader to Astala and Nesi (2003) for more details on these

exact structures.

After the paper by Milton and Kohn, the development slowed down. The

combination of two facts: the known nonoptimality of the translation bound as

m1 → 0 and the “natural” limit, K∗ ≤ k2I, of the known optimal structures

provocatively suggested that there were no other structures optimal for the bound.

However, in a surprising development 12 years later, Gibiansky and Sigmund (2000)

discovered new isotropic structures that are optimal for the translation bound for

smaller values of m1 than given in Theorem 2.1.

2.2.2.2.5 Gibiansky-Sigmund isotropic structures. Gibiansky and Sig-

mund (2000) announced a new construction that significantly increased the set

of optimal points of the translation bounds, (2.4) and (2.5). The paper focused

Page 35: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

22

mainly on the problem of bulk moduli, but the results easily apply to the con-

ductivity problem as they describe in their Section 5.3. Their structures were

the surprising result of a numerical simulation. Using a “topology optimization”

algorithm developed earlier by Sigmund, the authors searched for optimal structures

by computer. We refer the reader to Figures 4–9 in their paper for examples of the

fascinating structures selected by the procedure. In particular, Figure 6 in their

paper illustrates a structure which violates the condition (2.8) of Theorem 2.1 but

which numerically appears to satisfy the translation bound. When the authors

attempted to replace the computer output with a similar, but simpler structure

for which the effective properties could be analytically computed, the simplified

structure was optimal for the translation bounds. In Figure 2.5, we illustrate a

special case of their structure.

Instead of iterated laminates or coated spheres, they used a construction (also

used earlier by Sigmund (2000)) which resembles the work of Marino and Spagnolo

(1969). In the latter paper, the authors consider a completely different issue.

However, they introduce (among other things), conductivities b(x) in d dimensions

which take the form

b(x) = b1(x1)b2(x2) · · · bd(xd) (2.10)

and study some special cases. Roughly speaking, Gibiansky and Sigmund consider

conductivities which have this form but on several different scales and only in an

approximate sense.

Reinterpreting their results slightly, we divide the cell of periodicity into four

rectangular subdomains. Two opposite squares are occupied by K2 and K3, and

= K3

= K2

= K1

Figure 2.5. Extremal structures of Gibiansky and Sigmund.

Page 36: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

23

the remaining rectangles are filled with laminates from K1 and K3. The effec-

tive conductivity of the laminate depends on the volume fraction of materials in

it. This conductivity (or, equivalently, the volume fractions in the laminate) is

chosen in such a way that the conductivity equation (1.4) permits a separation

of the variables if the average fields are homogeneous. Because of this feature,

the solution is analytic, and so are the effective properties. Using Maple, the

authors then found that the structures are optimal for the translation bound (2.4).

The result is amazing because the structure is a mathematical approximation of a

numerical approximation of the optimization problem. The authors also described

more complicated structures that were optimal for larger values of m1 and which

coincided with the previously known structures at the point K∗ = K2. Their results

are summarized by the following theorem, which lowers the minimum value of m1

for which the bound (2.4) is optimal.

Theorem 2.2 (Gibiansky-Sigmund) Let m = (m1, m2, m3) and k = (k1, k2, k3)

be admissible parameters for the G-closure problem such that

2Θ(√

m2 − m2) ≤ m≤2Θ(1 − m2),

where Θ is given in (2.9). Then there exists an isotropic structure with the given

volume fractions and optimal for the bound (2.4).

Remark 2.5 The results of Gibiansky and Sigmund raise an interesting question.

If the volume fractions satisfy the inequalities

2Θ(√

m2 − m2) ≤ m1 < 2Θ(1 − m2),

is the isotropic point on the translation bound attainable?

In the next section, we introduce a class of anisotropic laminate structures

which contains structures with the same properties as the those of Gibiansky and

Sigmund. In Section 2.4, we analyze the fields in optimal structures and obtain a

clear picture of their features.

Page 37: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

24

2.3 New optimal structures

In this section, we construct a family of optimal laminate structures. These

structures are all particular cases of the structure illustrated in Figure 2.6, an

orthogonal laminate of high rank with six design parameters and five well-separated

scales (see Remark 2.1). In this section, we choose the structural parameters so

that the structure satisfies the translation bound (2.4) as equality. We begin with

degenerate cases and work toward the structure in full generality. We delay until

Section 2.4 the discussion of why we should expect such a structure to be optimal.

The reader more interested in the reasons the structures are optimal than in the

structures themselves is invited to skip directly to Section 2.4.

2.3.1 A convenient change of variable

The material K1 and its volume fraction m1 play a special role in the bound

(2.4) and in the associated optimal structures. For this reason, it is convenient

to introduce (and fix) the relative fractions of the other two materials. Given

m1, m2, m3 > 0 with m1 + m2 + m3 = 1, define p ∈ (0, 1) by

p =m2

m2 + m3. (2.11)

Note that it follows that

1 − p =m3

m2 + m3

andm2

m3

=p

1 − p.

Using p-notation, the translation bound (2.4) for three material mixtures is rewrit-

ten as

= K3

= K2

= K1

Figure 2.6. The general three-material laminate which we optimize.

Page 38: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

25

1

2· tr K∗ − 2k1

det K∗ − k21

≤ m1

2k1+ (1 − m1)

(p

k2 + k1+

1 − p

k3 + k1

)

.

We think of p ∈ (0, 1) as a parameter of the problem. With p fixed, we write the

requirement that a structure is optimal for (2.4) as

m1 =

12· tr K∗−2k1

det K∗−k21

−(

pk2+k1

+ 1−pk3+k1

)

12k1

−(

pk2+k1

+ 1−pk3+k1

) . (2.12)

2.3.2 T-structures

Figure 2.7a, b and c illustrate several special cases of the general structure

illustrated in Figure 2.6. The simplest optimal three-material structure in this

section, shown in Figure 2.7a, is the T-structure. It is assembled as a sequence of

laminates which depends upon two parameters. First, K1 and K3 are laminated

with normal in the x1-direction. Then, the resulting structure is laminated with K2

with the normal in the x2-direction. The effective properties are found by iterating

the formula (2.6)

KT = L

(

K2, L

(

K1, K3, n1,m1

m1 + m3

)

, n2, m2

)

where n1 = (1, 0)T and n2 = (0, 1)T .

Theorem 2.3 Let t ∈ (0, 1) and 0 < k1 < k2 < k3. Then there exist volume

fractions m1, m2, m3 > 0 so that p = t with p given by (2.11) and such that the

= K3

= K2

= K1

(a) (b) (c)

Figure 2.7. A selection of structures optimal for (2.4). The T-structure (a), theT-structure with one layer of “coating”(b), and the T 2-structure (c)

Page 39: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

26

T-structure with these volume fractions is optimal for the translation bound (2.4).

The values of the volume fractions are

m1 =Θ(1 − p)

1 − pΘ, m2 =

p(1 − Θ)

1 − pΘ, m3 =

(1 − p)(1 − Θ)

1 − pΘ

where Θ is defined in (2.9). The eigenvalues λ1, λ2 of the optimal T-structure are

computed to be

λ1 =(1 − Θ)pk2 + (1 − p)k3β

(1 − Θ)p + (1 − p)

λ2 =(1 − Θ)pk2 + (1 − p)k2

(1 − Θ)p + (1 − p)β

where

β =k2 + k1

k3 + k1

. (2.13)

It may seem surprising that we have found that there is always an optimal T-

structure for any p. This happens because we consider structures with fixed relative

volume fractions of K2 and K3 but with arbitrary fraction of K1.

2.3.3 Coating preserves optimality

In order to describe the variety of the optimal structures, we make the following

observation.

Theorem 2.4 (The Coating Principle) If a structure with effective conductiv-

ity K∗ is optimal for the translation bound (2.4), then all structures obtained by

laminating it with material K1 are also optimal for (2.4), though with different

volume fractions. The laminating can be iterated several times with various normals

so that the original structure is “coated” by K1.

Proof : It is enough to apply the lamination formula (2.6) to K∗ (with volume

fractions m1, m2 and m3) and K1, specifying the normal of lamination, n, and the

volume fraction, c, of K1. This lamination produces a new material with effective

tensor

K∗′ = L(K∗, K1, n, c).

Page 40: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

27

Substituting K∗′ into the bound (2.4) along with the updated volume fractions

m′1 = 1 − c + cm1, m′

2 = cm2, m′3 = cm3

and using the fact that K∗ satisfies the bound as equality, one can verify that K∗′

also satisfies the bound as equality, with the new volume fractions.

The theorem states that laminating optimal structures for the translation bound

with K1 preserves the optimality, though it changes the volume fractions. This

observation allows us to restrict ourselves to the description of only extremal struc-

tures that attain the bound (2.4). By extremal structures, we mean structures that

contain the minimal amount of K1.

2.3.4 Example: the optimality oftwo-material matrix laminates

As a particular example, the coating principle can be used to prove the opti-

mality of the two-material second-rank laminates discussed in Section 2.2.2.1. This

case can be considered a special case of the three-material problem with

p =m2

m2 + m3= 1

(that is, m3 = 0). Begin with Ω filled with pure K2 so that m1 = 0 and m2 = 1.

These volume fractions clearly satisfy the requirement on the ratio p. Furthermore,

it is easy to check that (2.12) holds for this structure since K∗ = K2 so

12· tr K∗−2k1

det K∗−k21

−(

pk2+k1

+ 1−pk3+k1

)

12k1

−(

pk2+k1

+ 1−pk3+k1

) =

12· 2k2−2k1

k22−k2

1

−(

1k2+k1

+ 0k3+k1

)

12k1

−(

1k2+k1

+ 0k3+k1

)

=1

k2+k1− 1

k2+k1

12k1

− 1k2+k1

= 0 = m1.

Therefore, the block of pure K2 is optimal for (2.4). Applying the coating principle

once, we then find that any lamination of K2 and K1 is also optimal. Finally,

applying the principle a second time, we find that the structures illustrated in

Figure 2.3b are optimal.

The coating principle also plays an important role in the analysis of multima-

terial mixtures. Notice that the coating changes the volume fractions, mi, but

Page 41: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

28

it preserves the value of p. Since coating increases the value of m1, the principle

allows to look for the optimal structures with the lowest value of m1. Every optimal

structure generates a set of optimal coated structures. The set L(K∗) of optimal

structures obtained by coating is a domain in the plane of eigenvalues of the effective

tensor K∗. The two boundary components of this set correspond to the laminates

of the anisotropic, generating material K∗ and K1 with normal parallel to one of

the eigenvectors of K∗. To derive the equations for the boundaries, let K∗ be given

with volume fractions described by m1 and p. Let λ1 and λ2 be the eigenvalues of

K∗.

The boundary of L(K∗) is found from the lamination formula, (2.6), by lami-

nating K∗ with K1 with volume fractions c and 1 − c, respectively:

K∗′ = L(K∗, K1, n, c)

where n is parallel to an eigenvector of K∗. The laminate contains K1 in the fraction

m′1 = 1 − c + cm1, and the value of p is preserved, p′ = p. If n is chosen parallel

to the eigenvector associated with λ1, then the eigenvalues λ′1 and λ′

2 of K∗′ are

parameterized by c as

B1(K∗) =

((c

λ1+

1 − c

k1

)−1

, cλ2 + (1 − c)k1

)

: c ∈ [0, 1]

, (2.14)

while if n is parallel to the other eigenvector, the new eigenvalues are parameterized

by c as

B2(K∗) =

(

cλ1 + (1 − c)k1,

(c

λ2

+1 − c

k1

)−1)

: c ∈ [0, 1]

. (2.15)

We define Λ(K∗) ⊂ R2 to be the closed set bounded by B1(K

∗) ∪ B2(K∗) defined

in (2.14) and (2.15) where λ1 and λ2 are the eigenvalues of K∗. We define

L(K∗) = K∗′ ∈ R2×2sym : K∗′ has eigenvalues in Λ(K∗) (2.16)

where R2×2sym denotes the two-by-two symmetric matrices. The following is an im-

mediate corollary of Theorem 2.4.

Page 42: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

29

Corollary 2.1 Let t ∈ (0, 1). Let k = (k1, k2, k3) and m = (m1, m2, m3) be

admissible parameters of the G-closure problem such that m2 = t(m2+m3). Suppose

K∗ ∈ G(m; k). Then, for any effective tensor K∗′ ∈ L(K∗), there exist admissible

volume fractions m′ = (m′1, m

′2, m

′3) such that m′

2 = t(m′2 + m′

3) and such that

K∗′ ∈ G(m′; k).

2.3.5 Coated T-structures

From the optimal T-structure, we obtain a set of optimal structures by coating

with K1. This set, L(KT ), is shaded in the eigenvalue plane in Figure 2.8 for the

parameters

k1 = 1, k2 = 2, k3 = 5, p =1

60.

It is convenient to represent an anisotropic material by two symmetric points

(λ1, λ2) and (λ2, λ1) in the plane of eigenvalues to avoid ordering them. Particularly,

the optimal T-structure is represented by two points, both labeled KT . The domain

L(KT ) of optimal structures as defined in (2.16) is the union of two lens-shaped

regions in the plane. The boundaries of this set are the laminate curves. Recall

that rather than fixing volume fractions, we fix the value p, which in turn fixes the

ratio of m2 to m3. The figure also includes some dotted curves of constant volume

fraction. Those closer to K1 indicate larger values of m1 than those farther away.

Any point where one of these curves intersects the region L(KT ) is an optimal point

for the translation bound (2.4) with the volume fractions given through m1 and p.

Remark 2.6 Observe the change in topology of the intersection of constant vol-

ume fraction curves with L(KT ). For large values of m1, the intersection is a

connected portion of the curve. As m1 decreases, the intersection suddenly becomes

disconnected (specifically, when the constant volume fraction curve passes through

the point K2). Letting m1 continue to decrease, one sees that the two connected

components of the intersection shrink to points and then vanish (when the curve

passes through the points KT ). For m1 lower than this, the intersection is empty.

Page 43: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

30

λ1

λ2

b

b

b

b

b

b

K1

K3

KT

KT

KGS

K2

m1 = 0.8

m1 = 0.2469

m1 = 0.05488

Figure 2.8. The shaded region denotes the set L(KT ) of optimal structures formedby coating the optimal T-structure.

The two outer curves from KT to K1 represent the most anisotropic structures

of the class of coated T-structures. The pair of eigenvalues λ1, λ2 along these curves

are parameterized by

λ1 =

(

ν1

k1

+ (1 − ν)(1 − Θ)p + (1 − p)

(1 − Θ)pk2 + (1 − p)k3β

)−1

(2.17)

λ2 = νk1 + (1 − ν)(1 − Θ)pk2 + (1 − p)k2

(1 − Θ)p + (1 − p)β(2.18)

where the constants Θ and β are defined in (2.9) and (2.13). We can parameterize

the inner curves analogously. The parameter ν ∈ [0, 1] along the outer curve

controls the amount of material K1 added to the T-structure. The volume fractions

of the final structure depending on ν and p are given by

m1 = ν + (1 − ν)Θ(1 − p)

1 − pΘ, m2 = (1 − ν)

p(1 − Θ)

1 − pΘ, m3 = (1 − ν)

(1 − p)(1 − Θ)

1 − pΘ.

The more isotropic curves cross the line of isotropy at exactly λ1 = λ2 = k2 and do

so when

Page 44: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

31

ν =(1 − p)(k3β − k2)

(1 − Θ)p(k2 − k1) + (1 − p)(k3β − k1).

At this point, we find that m1 = 2Θ(1 − m2) (see Theorem 2.1!) The region of

intersection (shaded darker) of the two lenses was proved to be optimal by Milton

and Kohn (1988) using the laminated structures described in Section 2.2.2.2. The

remaining portion of the region of L(KT ) represents the first new optimal structures

of this chapter. We will improve this region later. Observe that the only optimal

isotropic structures found in L(KT ) were already known to be optimal. However,

this construction has introduced a relatively large set of new optimal anisotropic

structures. Furthermore, this construction proves the optimality of the bound (2.4)

in a region of anisotropic points for a smaller value of m1 than was previously known

possible: m1 ≥ Θ(1−m2) rather than m1 ≥ 2Θ(1−m2). The coated T-structures

are a generalization of the Milton-Kohn structures in the sense that the latter always

have effective tensors in L(KT ). For reference, we have also indicated by the dashed

line from K2 to KGS the optimal isotropic structures introduced by Gibiansky and

Sigmund. This line intersects the set L(KT ) only at K2.

2.3.6 T2-structures

We now enlarge the class of optimal structures with a set of structures which

connects the points KT and KGS in Figure 2.8. We laminate the T-structure with a

laminate from K1 and K3 in the orthogonal direction as illustrated in Figure 2.7c.

The effective tensors of such T 2-structures are found from the iterative procedure

KT2 = L(KT , K ′13, n1, ω2),

KT = L(K2, K13, n2, ω1),

K ′13 = L (K1, K3, n2, ν

′) ,

K13 = L (K1, K3, n1, ν) .

(2.19)

Here, ν and ω1 are the parameters of the original T-structure. ν ′ is the relative

fraction of K1 in the additional K1-K3 laminate, and ω2 is the relative amount of

the T-structure compared to the additional laminate in the final T 2-structure.

The properties depend on four structural parameters: ν, ν ′, ω1, ω2 that all vary

in [0, 1] and are subject to the constraint that fixes p. (Recall that m1 is treated as a

Page 45: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

32

variable.) Observe that the T 2-structures are a generalization of the T-structures.

(If ω2 = 0 in the equations above, then KT2 = KT .) Moreover, we show that

they describe a curve of anisotropic structures between the T-structures and the

isotropic structures of Gibiansky and Sigmund.

Theorem 2.5 Let t ∈ (0, 1) and let admissible parameters m = (m1, m2, m3) and

k = (k1, k2, k3) be given such that

p =m2

m2 + m3

= t

and

2Θ−p + 2p2Θ − pΘ +

p2Θ2 − 2p2Θ + p

(1 − 2pΘ)2 ≤ m1 ≤Θ(1 − p)

1 − pΘ. (2.20)

Then there exists a T 2-structure with the given volume fractions and optimal for

the bound (2.4). These structures vary between the anisotropic T-structure and the

isotropic point of Gibiansky and Sigmund (2000). The optimal volume fractions

(see (2.19)) in the structure satisfy

ν = Θ, ν ′ = ω1Θ

where Θ is defined in (2.9) and ω1, ω2 satisfy

ω1 + ω2 =1

Θ(m1 + 2Θm2), ω1ω2 = m2. (2.21)

The effective tensors have eigenvalues λ1, λ2, written in terms of ω1 and ω2 as

λ1 =ω1k2(k3 + k1) + (1 − ω1)k3(k2 + k1)

ω2(k3 + k1) + (1 − ω2)(k2 + k1), (2.22)

λ2 =ω2k2(k3 + k1) + (1 − ω2)k3(k2 + k1)

ω1(k3 + k1) + (1 − ω1)(k2 + k1). (2.23)

Notice that the relation between the effective properties of optimal mixtures is

symmetric to the interchanging of ω1 with ω2 in spite of the nonsymmetric iterative

procedure.

Page 46: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

33

When m1 equals its upper bound in (2.20), then one of ω1 or ω2 must be equal to

1. Thus, the T 2-structure degenerate into the T-structure for this volume fraction.

On the other hand, when m1 equals its lower bound, then we find

ω1 = ω2 =1

2−√

1

4− m1

and we obtain an isotropic structure whose volume fractions satisfy the equality

m1 = 2Θ(√

m2 − m2) (compare to Theorem 2.2!). The set of T 2-structures is a

generalization of both the T-structures and the Gibiansky-Sigmund structure with

minimal amount of m1 (see Figure 2.9).

Remark 2.7 We show in Section 2.4 how to find these structures by analysis

of the fields. In particular, we show how the parameters ω1, ω2, ν and ν ′ are

naturally determined and explain the remarkable properties of the local fields in

optimal structures that lead us to the construction.

2.3.7 The set of optimal structures

Applying the coating principle to the extremal T 2-structures, we obtain a variety

of optimal structures because each T 2-structure can be coated, increasing the

amount m1 but keeping relative fraction p. The union of all the L(KT2) (see

(2.16)), for each KT2 with the given value of p, forms a set of structures optimal

for the translation bound (2.4). This set is illustrated in the eigenvalue plane in

Figure 2.9 for parameters

k1 = 1, k2 = 2, k3 = 5, p =m2

m2 + m3

=1

60.

The set of optimal structures is bounded by the solid boundary, which is the union

of coated T-structures (the curves between K1 and KT ) and the T 2-structures

(the curve passing through KGS). The closed region bounded by the dashed

lines represent the previously known optimal structures of Milton and Kohn, and

Gibiansky and Sigmund.

Curves of constant volume fraction are indicated by the dotted lines. In partic-

ular the curve passing through KT represents the case m1 = Θ(1 − m2) while the

Page 47: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

34

λ1

λ2

b

b

b

b

b

b

K1

K2

K3

KGSKT

KT

m1 = 0.8

m1 = 0.4

m1 = 0.2469

m1 = 0.15

m1 = 0.05488

Figure 2.9. Optimal points for the lower translation bound for m2

m2+m3= 1

60.

The curve between KT and KGS corresponds to the optimal T 2-structures. Thecurve between KT and K1 corresponds to the coated T-structures illustrated inFigure 2.7b. The dashed line from K2 to KGS corresponds to the structures ofGibiansky and Sigmund. The dashed curve passing through K1 and K2 forms theboundary for the structures of Milton and Kohn.

curve passing through KGS represents the case m1 = 2Θ(√

m2 − m2). Recalling

Remark 2.6, we see that the topology of the intersection of the curves m1 = const

and the new region of optimal structures remains connected for all volume frac-

tions satisfying m1 > 2Θ(√

m2 − m2). As m1 decreases below this amount, the

intersection shrinks to a single point and then vanishes for all smaller m1.

2.3.8 The problem with fixed volume fractions

Until now, we have considered a problem of fixed p = m2/(m2 + m3). The clas-

sical G-closure problem, however, asks that we fix the volume fractions. Obtaining

this information from our results is quite straightforward because the pair (m1, p)

uniquely determines all volume fractions.

As an example, we find optimal structures in G(m; k) for the parameters

k1 = 1, k2 = 2, k3 = 5

and

Page 48: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

35

m1 = 0.4, m2 = 0.01, m3 = 0.59

as in Figures 2.1–2.2. To accomplish this, we need only examine Figure 2.9 (notice

that m2/(m2 + m3) = 1/60) and the corresponding figure for the upper bound

with m2/(m1 + m2) = 1/41, Figure 2.10. The optimal points of the lower bound

marked by the thick portion of the curve in Figure 2.2 are the intersection of the

curve of constant m1 = 0.4 with the optimal region shown in Figure 2.9. The dot

in Figure 2.1 marks the point where this curve intersects the dashed Gibiansky-

Sigmund line. Similarly, the optimal points marked by the thick curve on the

upper bound in Figure 2.2 are where the line of constant volume fraction m3 = 0.59

intersects the set of optimal points in Figure 2.10. The thick portion of the upper

bound in Figure 2.1 marks the intersection with the Milton-Kohn region.

Look again at Figure 2.9. As long as m1 ≥ 2Θ(√

m2 − m2) (For the given

parameters, m1 ≥ 0.05488.), the intersection of the curve m1 = const and the

region of attainable points is a connected subset of the curve which includes the

isotropic point. Thus, the intersection is uniquely defined by the most anisotropic

λ1

λ2b

b

b

b b

b

K1

K2

K3

KGS

KTKT

m3 = 0.9

m3 = 0.59

m3 = 0.1749

m3 = 0.04616

Figure 2.10. Optimal points for the upper translation bound for m2

m1+m2= 1

41. The

curve between KT and KGS corresponds to the optimal T 2-structures. The curvebetween KT and K3 corresponds to coated T-structures. The dashed line from K2

to KGS corresponds to the structures of Gibiansky and Sigmund. The dashed curvepassing through K3 and K2 forms the boundary for the structures of Milton andKohn.

Page 49: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

36

point of this subset. For m1 ≥ Θ(1−m2) (For the given parameters, m1 ≥ 0.2469.)

this most anisotropic point is a coated T-structure. For m1 ≤ Θ(1 − m2), it is a

T 2-structure. We summarize this in the following theorem.

Theorem 2.6 Let the volume fractions m1, m2, m3 > 0, and material properties

0 < k1 < k2 < k3 be given admissible parameters. Define p as in (2.11) and Θ as

in (2.9). Then the following hold.

(i) If m1 > Θ(1 − m2), then (2.4) is optimal. There exists a set of opti-

mal points on the bound which includes the isotropic point and whose most

anisotropic member is that given by the coated T-structure with eigenvalues

(2.17) and (2.18) where ν is chosen to satisfy the volume fraction constraints:

ν =1 − pΘ

1 − Θ

(

m1 −Θ(1 − p)

1 − pΘ

)

.

(ii) If 2Θ(√

m2−m2) ≤ m1 ≤ Θ(1−m2), then (2.4) is optimal. There exists a

set of optimal points on the bound which includes the isotropic point and whose

most anisotropic member is that given by the optimal T 2-structure satisfying

the volume fraction constraints. In particular, the parameters ω1 and ω2 for

this most anisotropic structure can be found by solving simultaneously the

equations (2.21).

2.3.9 Applicability: volume fractions

In Figure 2.11, we illustrate the difference between the applicability of Theo-

rem 2.1 and Theorem 2.6 in the case of isotropic structures. In both figures we take

k1 = 1, k2 = 2, k3 = 5. By the definition of admissible volume fractions, we know

that in the m1m3-plane (note that m2 = 1 − m1 − m3), the volume fractions are

constrained to the region bounded by the axes m1 = 0 and m3 = 0 and by the line

m1 + m3 = 1.

Consider Theorem 2.1 for the lower bound (2.4) which we illustrate on the left

of Figure 2.11. It implies that there is an isotropic structure that attains the bound

if m1 ≥ 2Θ(1 − m2) where Θ is defined in (2.9). We have indicated the line m1 =

Page 50: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

37

m3

m1

m1 + m3 = 1

m3

m1

m1 + m3 = 1

Figure 2.11. Domain of applicability of Theorem 2.1 (left) and Theorem 2.6(right) in terms of volume fractions. (Here we consider the isotropic case fork1 = 1, k2 = 2, k3 = 5.) The volume fractions are physically constrained to theregion bounded by the axes m1 = 0 and m3 = 0 and the line m1 + m3 = 1. Thethick solid curves depict m1 = 2Θ(1−m2) (left) and m1 = 2Θ(

√m2 −m2) (right).

The thick dashed curves are similar but for the upper bound (2.5). The shadedregions indicate the region of parameters for which each theorem provides structureswhich attain both bounds.

2Θ(1−m2) by the thick solid curve passing from the origin (m2 = 1, m1 = m3 = 0)

to the line m1 + m3 = 1. (The point of intersection is m2 = 0, m1 = 2Θ, m3 =

1 − 2Θ.) For any admissible volume fractions which lie to the right of this curve,

Theorem 2.1 implies that there exists an isotropic structure that is optimal for the

lower bound (2.4). The thick dashed line gives similar information for the upper

bound (2.5). For any admissible volume fractions lying above this curve, there exists

an isotropic structure that satisfies the upper bound as equality. In particular, the

shaded region between the two curves indicates volume fractions for which both

bounds are optimal.

On the right of Figure 2.11, similar information is depicted for Theorem 2.6.

In particular, the thick solid curve represents m1 = 2Θ(√

m2 − m2). (Note the

endpoints m2 = 1, m1 = m3 = 0 and m2 = m1 = 0, m3 = 1.) The dashed

line represents the similar curve for the upper bound. The figure illustrates quite

clearly how powerful Theorem 2.6 is for small m2 (that is, for points near the line

m1 + m3 = 1). In this case, both bounds can be proved optimal for a range of

volume fractions including those where m1 or m3 are very close to zero.

Page 51: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

38

We should remark that the corresponding figure for Theorem 2.2 coincides with

the right side of Figure 2.11. This is because Theorem 2.2 and Theorem 2.6 are iden-

tical for isotropic structures. However, Theorem 2.2 does not apply to anisotropic

structures, while Theorem 2.6 certainly does. Similar figures to Figure 2.11 can

be produced for varying degrees of anisotropy. In particular, if the ratio of the

eigenvalues of K∗ is nearly one, Figure 2.11 changes only slightly. Theorem 2.6 can

be considered a generalization of Theorem 2.2 to anisotropic structures.

2.3.10 An inner bound of the G-closure

From Theorem 2.6 and the results of Cherkaev and Gibiansky (1996), we can

produce a naive inner bound of the G-closure by lamination. Consider Figure 2.12a.

Here we have plotted an inner and outer bound of the G-closure for the parameters

k1 = 1, k2 = 2, k3 = 5,

m1 = 0.104, m2 = 0.5, m3 = 0.396.

The union of the unshaded and shaded region is an outer bound for the G-closure

formed from the Wiener and translation bounds. In this case, m1 is very close

to 2Θ(√

m2 − m2) so that case (ii) of Theorem 2.6 applies. We mark the most

anisotropic effective tensor given by this theorem by T . We also mark the least

anisotropic effective tensor on the Wiener bound (Cherkaev and Gibiansky (1996))

by H . By laminating the structures with effective tensors T and H , we form a

family of structures which lie on the uppermost curve connecting the two points.

Since this curve necessarily lies in the G-closure for the given parameters, we find

that the unshaded region of the figure depicts an inner bound on the G-closure.

The shaded region shows what is still not known for these parameters. The bounds

do not determine whether these points belong to the G-closure or not. We remark

that there is a similar region of “unknown points” near the point marked U in the

Figure. However, for these parameters and at the scale of the figure, the region is

impossible to see.

Figure 2.12b is similar. We use the same conductivity parameters, but less

“extreme” values of the volume fractions:

Page 52: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

39

b

b

b

b

b

H

T

U

T ′

H ′

λ1

λ2

b

b

bb

H

T

T ′

H ′

λ1

λ2

(a) (b)

Figure 2.12. Inner and outer bounds on the G-closure for two different sets ofadmissible parameters. The bounds do not determine whether the points lying inthe shaded regions are a part of the G-closure or not.

m1 = 0.25, m2 = 0.5, m3 = 0.25.

There are very small (compared to the area of the G-closure) regions of unknown

points near the “corners” where the bounds intersect. That is, T and H are very

close to each other. In this case, both the inner and outer bounds are very close to

the G-closure itself.

If we allow m1 to decrease far enough, eventually this construction will not work.

When m1 = 2Θ(√

m2 − m2), the points T and T ′ of Figure 2.12a coincide. For

m1 < 2Θ(√

m2 −m2), we know of no structures which attain the lower translation

bound. In this case, a simple bound can be obtained by laminating the points H

and H ′ (that is, create a polycrystal of the anisotropic material H). In all cases,

a better inner bound could be obtained by more carefully mixing known optimal

points. This is a difficult problem, however, and we will not discuss it further.

2.4 Fields in optimal structures

2.4.1 Local fields required by the translation bound

For the reader’s convenience, we derive the optimality conditions for the transla-

tion bound below (see Lurie and Cherkaev (1982); Tartar (1979, 1985)). We sketch

the ideas of the derivation, focusing on the conditions on the fields inside each

material of an optimal structure.

Page 53: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

40

The lower bound uses the quasi-affineness of the determinant function

Ω

det DU dx = det E, ∀E ∈ R2×2, ∀U ∈ H1

#(Ω)2 + Ex

which we can verify for smooth functions by writing

det DU = div

(

u1

(∂u2

∂y,−∂u2

∂x

))

and applying the Divergence Theorem. The general result follows by approxima-

tion. The construction of the lower bound is as follows. We begin by adding and

subtracting the constant 2t det E for some t ∈ R.

W(K, E) = infU∈H1

#(Ω)2+Ex

Ω

(〈DU K, DU〉 + 2t det DU) dx − 2t detE.

Next, we relax the differential constraint on the field DU by replacing the set

H1#(Ω)2 + Ex with the set of F ∈ L2(Ω; R2×2) such that

ΩF dx = E. Thus, we

have

W(K, E) ≥ infF

Ω

(〈F K, F 〉 + 2t det F ) dx − 2t det E (2.24)

such that

Ω

F dx = E. (2.25)

This equation gives a family of bounds on W(K, E) parameterized by t. Here we

consider the case t = ±k1. We will choose t = k1 for the rest of this section.

The other case is analogous. It is not guaranteed that a minimizer, F , of the

right-hand side of (2.24) will be a gradient, but if it is (or if it can be approximated

in the appropriate sense by a sequence of gradients) then the translation bound

is optimal. In the rest of the section, we analyze the conditions of a minimizer,

F (with t = k1). Then we construct gradient fields DU ∈ H1#(Ω)2 + Ex which

approximate the minimizer F , proving that the bound is optimal in the cases we

discussed in Section 2.3.

To simplify the calculation, we use a rotation-invariant decomposition of the

quadratic forms on two-by-two matrices defined by Q1(F ) = |F |2 = 〈F, F 〉 and

Page 54: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

41

Q2(F ) = 2 detF (see Astala and Miettinen (1998)). Namely, we take the zero sets

H+ = (Q1 − Q2)−1(0) and H− = (Q1 + Q2)

−1(0). Then

R2×2 = H+ ⊕H−

where

H+ =

(a b

−b a

)

: a, b ∈ R

, H− =

(a bb −a

)

: a, b ∈ R

. (2.26)

In particular, we can write

F = F+ + F−, F+ ∈ H+, F− ∈ H−

where

F+ =1

2(F + cof F ), F− =

1

2(F − cof F ). (2.27)

Here cof is the linear operator on matrices which returns the cofactor matrix.

cof

(f11 f12

f21 f22

)

=

(f22 −f21

−f12 f11

)

.

It is easy to verify that

|F+|2 + |F−|2 = |F |2, |F+|2 − |F−|2 = 2 det F.

Differentiating the integrand in (2.24) with respect to F we find the conditions

for a minimizer:

F (x)K(x) + k1 cof F (x) = A, a.e. in Ω (2.28)

or, equivalently,

F+(x)(K(x) + K1) + F−(x)(K(x) − K1) = A, a.e. in Ω

where A is a constant matrix of Lagrange multipliers enforcing (2.25). In particular,

we find that in Ω1 (where K ≡ K1) we have

2k1F+ = A a.e. in Ω1 (2.29)

which implies that A− = 0 in Ω. For Ωi with i = 2, . . . , N , we then have

(ki + k1)F+ = A, F− = 0 a.e. in Ωi. (2.30)

Page 55: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

42

Using (2.29) and (2.30) we solve for A by noting that

E =

Ω

F dx = AN∑

i=1

mi

ki + k1

+

Ω1

F− dx

so that

A =

(N∑

i=1

mi

ki + k1

)−1

E+,

Ω1

F− dx = E−.

Thus, we have the following theorem.

Theorem 2.7 A vector field F ∈ L2(Ω; R2×2) is a minimizer of the right-hand-side

of (2.24) if and only if

(i) F+ = 1ki+k1

(N∑

j=1

mj

kj+k1

)−1

E+ a.e. in Ωi for i = 1, . . . , N .

(ii)∫

Ω1F− dx = E−.

(iii) F− = 0 a.e. in Ωi for i = 2, . . . , N .

In particular, we can explicitly write the lower bound as

W(K, E) ≥(

N∑

i=1

mi

ki + k1

)−1

|E+|2 − 2k1 det E, ∀E ∈ R2×2. (2.31)

The traditional form of the translation bound (2.4) is obtained by choosing E to

make the bound above as tight as possible, and thus eliminating the dependence of

the bound on the fields. However, we are more interested in Theorem 2.7 since it

tells us exactly what the fields in each material of an optimal structure are.

Specifically, notice that the value of F is constant with F− ≡ 0 in all but the

first material: Ω2, . . . , ΩN . Furthermore, F+ is fixed in the remaining material

Ω1. The only “freedom” we have in our choice of F is the values of F− in Ω1

which are arbitrary as long as we satisfy the constraint on the average, (ii). If the

translation bound is optimal, then the corresponding structures contain pointwise

fields which are (nearly) constant in Ω2, . . . , ΩN and constrained to belong to the

two-dimensional manifold

1

2k1

(N∑

j=1

mj

kj + k1

)−1

E+ + H−

in Ω1. Here H− is defined in (2.26).

Page 56: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

43

Remark 2.8 The problem of finding optimal structures has thus been reduced to a

differential inclusion problem: find U with a given average such that

DU ∈ K ⊂ R2×2 a.e. in Ω.

2.4.2 Rank-one connection and thefields in the optimal structures

As in Section 2.3, we deal with the class of orthogonal laminates — that is, the

class of laminates with mutually orthogonal normals. The effective properties of

this structures are found by iterating the equation (2.6) assuming that the normal

is equal either to n1 = (1, 0) or to n2 = (0, 1). It is enough to consider only diagonal

average fields (1.6) (see for example Milton (2002)) so that

E+ = αI, and A =

(N∑

i=1

mi

ki + k1

)−1

αI (2.32)

for some α ∈ R.

Remark 2.9 We choose laminates because the fields can be taken to be constant in

each layer, making them easy to work with. In actuality, there can be fluctuations,

especially near the boundaries. But with well-separated scales (see Briane (1994))

tending toward length zero, the fields may be assumed constant for the purpose of

computing the energy W(K, E). Technically speaking, what we will continue to call

“fields” are called “correctors” in the language of homogenization. The most elegant

mathematical formulation is given through the use of the so-called gradient Young

measures. The interested reader is referred to Muller (1999) for an introduction to

the subject in the context of materials science.

The fields of neighboring layers in a laminate structure are rank-one connected.

In other words, two neighboring layers with respective fields FA and FB must satisfy

det(FA − FB) = 0. (2.33)

On the other hand, the fields in each material in an optimal structure are given by

Theorem 2.7 with F = DU . We now analyze these conditions, showing how they

guide us in creating optimal structures.

Page 57: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

44

The magnitude of the average field does not affect the effective properties,

therefore it is convenient to rescale the fields in the theorem so that

α = k1

N∑

i=1

mi

ki + k1, E+ = αI, A+ = k1I (2.34)

in (2.32). Thus, in an optimal structure, the field DU satisfies

DU+ =k1

ki + k1I a.e. in Ωi for i = 1, . . . , N.

Since trF = tr F+, we can rephrase the conditions of Theorem 2.7 as follows.

(P1) DU = DUT and tr DU = 1 a.e. in Ω1.

(P2)∫

Ω1DU− dx = E−.

(P3) DU = k1

ki+k1I a.e. in Ωi for i = 2, . . . , N .

Remark 2.10 We refer the reader to Grabovsky (1996) for a detailed derivation

of similar conditions for composites of two generally anisotropic linear elastic ma-

terials. The isotropic case is covered in (3.26) and (3.27) of his paper which the

reader may wish to compare to (P1)–(P3) above with N = 2. The same paper

also makes the connection between optimal composites in linear conductivity and in

special cases of linear elasticity. We further remark that conditions (P1)–(P3) are

a special case of (25.31) in Milton (2002).

One immediately observes that the fields in all materials except K1 are not

in rank-one connection with each other and therefore are incompatible, because

(P1)–(P3) and (2.33) are contradictory: if FA = k1

k2+k1I and FB = k1

k3+k1I then

det(FA − FB) 6= 0. In particular, no optimal laminate can contain a layer of K2

laminated with K3.

In a sense, K1 must be used as a glue between layers to ensure compatibility.

The volume fraction of K1 therefore cannot be too small, which indicates that the

translation bound (2.4) cannot be optimal for laminates if m1 is smaller than a

critical value.

Page 58: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

45

2.4.3 Constructing optimal laminates

We now describe an algorithm for constructing laminates which are optimal for

(2.4). We always assume that the fields in the layers of the laminate satisfy (P1)–

(P3) and we find ways to join the materials by rank-one connection into laminate

structures. Since we require that (P1)–(P3) hold at all times in the process, the

final structures are necessarily optimal.

We leave the average field, E, and the volume fractions free until the end of

the process, when they are computed from the construction. In other words, by

following the procedure described in this section, we are guaranteed to produce

optimal structures, but we do not know the parameters of the optimization problem

(the average field and volume fractions) until the structure is complete. However,

at the end of the process, we have an algorithm to find all G-closure parameters.

2.4.4 The general structure revisited

Consider again the general structure illustrated in Figure 2.6. Notice that there

are six design parameters (the volume fractions in the laminate layers). Since the

structure is an orthogonal laminate and since we assume the average field E is

diagonal, we have that the local fields are also diagonal. Notice that material K2

appears in lamination only once, material K3 twice, and material K1 four times.

Thus, we have 6 design variables and 14 field variables. To make the structure

optimal, we must fill in the fields in the pure material components satisfying a

number of conditions. Let us count these conditions.

First, we have rank-one connections between the fields and the currents in each

layer of the laminate. This gives two condition per layer, since the tangential

component of the field and the normal component of the current must be continuous

across the interfaces. There are six laminate interfaces, so we have 12 continuity

conditions.

For optimality, we also need to satisfy (P1)–(P3). These give two conditions for

each layer of K2 and K3 and one condition for each layer of K1. Thus, (P1)–(P3)

impose 10 more conditions.

Page 59: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

46

By our count, we have 20 free variables and 22 conditions they must satisfy.

In fact, we observe that the general structure is a coated T 2-structure. The

T 2-structure has two free design parameters (ω1 and ω2). The coating introduces

two more parameters, so we actually have four degrees of freedom in spite of the

seemingly over-determined system. As we will show in the following sections, some

of the constraints are satisfied “for free” if the others are satisfied. In particular,

when we choose parameters to satisfy the rank-one connections and (P1)–(P3)

for the fields in the general structure, the currents are automatically rank-one

connected. This removes six constraints from the list above and we are left with 20

variables and 16 constraints, providing the four degrees of freedom we observe.

This is not as surprising as it may at first seem. Indeed, assume we have a

partition of Ω into Ωi with associated conductivity tensor K defined through (1.5).

Furthermore, assume that we find U ∈ H1#(Ω)2 + Ex that satisfies the optimality

condition (2.28) with F = DU . Then, by taking the divergence of both sides

of (2.28) and using the fact that div(cof DU) = 0, we find that U satisfies the

PDE div(K DUT ) = 0 automatically. The analogous statement for laminates is

that it is enough to check the jump conditions only of the piecewise constant field

approximating DU . We obtain “for free” the corresponding conditions for the

piecewise constant field approximating DU K.

2.4.4.1 The optimal T-structures

The construction of optimal laminates preserves the fields in the layers according

to (P1)–(P3). By restricting ourselves to orthogonal laminates and diagonal average

fields, we guarantee that the fields in each layer will be diagonal, which allows us

to simplify notation. We represent the diagonal two-by-two matrices as

M(α, β) =

(α 00 β

)

and associate M(α, β) with the point (α, β) in the plane. Observe that M(α1, β1)

and M(α2, β2) are rank-one connected if and only if (α2 −α1)(β2 − β1) = 0. In the

plane, this means that (α1, β1) and (α2, β2) lie on the same horizontal or vertical

line.

Page 60: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

47

We illustrate the construction of optimal structures in Figure 2.13a and b. We

begin with the T-structure. The discussion is accompanied by Figure 2.13a. In this

figure, the points E2 and E3 represent the fixed fields given by (P3). The line l

represents the line of constant trace given by (P1). We wish to construct a laminate

which has its internal fields lying on the set E2 ∪ E3 ∪ l.

First, we look for a rank-one connection between the materials K1 and K3. The

admissible fields for K1 lie on the line l while the admissible field for K3 is the point

E3. Let us laminate in the x1-direction, which means the field in K1 must lie on

the intersection of l and the horizontal line through the fixed field E3. The optimal

fields E1 and E3 are

E1 = M

(k3

k3 + k1

,k1

k3 + k1

)

, E3 = M

(k1

k3 + k1

,k1

k3 + k1

)

,

which ensures that

(E1 − E3) ·(

01

)

=k1

k3 + k1− k1

k3 + k1= 0.

As we mentioned in Section 2.4.4, the rank-one connection condition on the currents

is automatic. So far, we have only ensured the condition on the fields. However,

note that

(k1E1 − k3E3) ·(

10

)

=k1k3

k3 + k1

− k1k3

k3 + k1

= 0

so the currents are also rank-one connected.

x

yb

b

bbc

r

E3

E2

E1

l

E13

ET

x

yb

b

b

b

bc rsr r

E3

E2

E′

1

E′′

1 l

E′

13ETET2

ECT

(a) (b)

Figure 2.13. Fields in optimal laminate structures.

Page 61: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

48

The average field upon laminating K1 with K3 is

E13 = νE1 + (1 − ν)E3 = M

(νk3 + (1 − ν)k1

k3 + k1,

k1

k3 + k1

)

while the average current is

J13 = νk1E1 + (1 − ν)k3E3 = M

(k1k3

k3 + k1,νk2

1 + (1 − ν)k1k3

k3 + k1

)

where ν is the relative fraction of K1 to K3. In Figure 2.13a, we indicate the point

E1. The dashed line connecting this point to E3 represents the path of the point

E13 as ν varies between 0 and 1. All of these fields are available to us in an optimal

laminate of K1 and K3 by appropriate choice of ν.

We choose ν so that we can laminate this new material with K2 in the x2-

direction (Refer to Figure 2.7a.), again satisfying (P1)–(P3). Thus, we need to

adjust ν so that E13 and E2 lie on the same vertical line in the plane, where

E2 = M(

k1

k2+k1, k1

k2+k1

)

. Solving for ν, we find

ν =k1(k3 − k2)

(k2 + k1)(k3 − k1)= Θ.

Note that Θ is defined in (2.9) and is a parameter of the optimal structures in the

previous section.

Once again, we can verify that the currents in the laminate layers are compatible.

When ν = Θ,

J13 = M

(k1k3

k3 + k1,

k1k2

k2 + k1

)

which is rank-one connected to the current k2E2. Similar calculations for all

other structures discussed in this chapter show that the condition on currents is

automatically satisfied each time we satisfy the condition on the fields, so for the

rest of this section, we will keep track only of the fields and not the currents in the

construction.

Setting ν = Θ brings the point E13 to the vertical line passing through E2

and we are free to laminate with as much K2 as we please. The average field in

the T-structure can lie anywhere on the dashed line connecting E13 and E2. The

only condition we have is the ratio of m1 to m3, leaving m2 free. In this way, the

Page 62: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

49

average field, E, depends on m2. We then find the volume fractions of the optimal

T-structure depending on the amount K2 parameterized by ω1:

m1 = Θ(1 − ω1), m2 = ω1, m3 = (1 − Θ)(1 − ω1). (2.35)

To find the effective properties, we need to find the average field,

ET = m1E1 + m2E2 + m3E3

=k1

k2 + k1

M

(

1,ω1(k3 + k1) + (1 − ω1)(k2 + k1)

k3 + k1

)(2.36)

and the average current,

JT = m1k1E1 + m2k2E2 + m3k3E3

=k1

k2 + k1

M

(ω1k2(k3 + k1) + (1 − ω1)k3(k2 + k1)

k3 + k1

, k2

)

.

Then we have that K∗ = E−1T JT . Substituting and simplifying, we get the eigen-

values of K∗:

λ1 =m2k2(k3 + k1) + (1 − m2)k3(k2 + k1)

k3 + k1

(2.37)

λ2 =k2(k3 + k1)

m2(k3 + k1) + (1 − m2)(k2 + k1). (2.38)

These expressions coincide with (2.22) and (2.23) for ω1 = m2, ω2 = 1. We can

now prove Theorem 2.3 by simple algebra using m2 = p(m2 + m3), (2.35), (2.37)

and (2.38).

2.4.4.2 The coating principle

It is now easy to see why the coating principle (Theorem 2.4) is true. Consider

Figure 2.13b. Starting from the optimal T-structure, ET , we look for compatible

fields for lamination. Notice that since the field in K1 can lie anywhere on the

line l, there are always two compatible fields: the intersection of the horizontal or

vertical line through ET and the line l. For this illustration, we have chosen the

Page 63: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

50

point E ′1 ∈ l which lies on the same horizontal line as ET . Specifically, from (2.36)

we find

E ′1 =

1

(k3 + k1)(k2 + k1)M(ω1k2(k3 + k1) + (1 − ω1)k3(k2 + k1),

ω1(k3 + k1) + (1 − ω1)(k2 + k1)).

By varying the amount of K1 which is added in this layer, we can obtain a new

optimal structure ECT which lies anywhere on the dashed line connecting ET and

E ′1. (Refer to Figure 2.7b.) Of course, this operation can be iterated. For example,

we could then laminate in the x2-direction with K1 by choosing the point on l which

intersects the vertical line through KCT and so on. In this way, we obtain a whole

family of optimal structures from a single optimal structure. Again, the coating

will change the average field, E, and volume fractions. However, the construction

must give an optimal structure for some values of these parameters which we can

compute.

2.4.4.3 The optimal T2-structures

Finally, we illustrate how to obtain the optimal T 2-structures. Begin with the

optimal T-structure, indicated by ET in Figure 2.13b. Notice that we can laminate

E3 with the point E ′′1 = M

(k1

k3+k1, k3

k3+k1

)

. By adjusting the volume fraction ν ′ of

K1 to K3 in this laminate, we can move the resulting average field

E ′13 = ν ′E ′

1 + (1 − ν ′)E3 = M

(k1

k3 + k1,ν ′k3 + (1 − ν ′)k1

k3 + k1

)

to any point on the dashed line connecting E3 and E ′′1 . In particular, if we choose

the volume fraction so that E ′13 is rank-one connected to ET — that is if we choose

ν ′ so thatν ′k3 + (1 − ν ′)k1

k3 + k1=

ω1(k3 + k1) + (1 − ω1)(k2 + k1)

k3 + k1

(see (2.36)) — then we can laminate the E ′13 and ET structures to obtain a T 2-

structure (see Figure 2.7c). It is easy to check that the correct value of ν ′ is ν ′ = ω1Θ

Page 64: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

51

with Θ defined in (2.9). We are then free to laminate ET with E ′13 in any fraction

we wish to obtain an optimal T 2-structure:

ET2 = ω2ET + (1 − ω2)E′13.

By computing the volume fractions, average field, and average current as we did

with the T-structure, one can verify all the statements of Theorem 2.5. We can

now apply the coating principle as many times as we wish to obtain a family of

optimal structures as described in Section 2.3.

2.4.5 Four and more materials

The generalization to N ≥ 4 is straightforward. In fact, Theorem 2.7 was proved

in the general case. The method for constructing optimal laminates is exactly the

same: we are given a set K of N−1 points and a two-dimensional plane in R2×2 and

we construct laminate structures whose internal fields are lie in K. As an example,

Figure 2.14a shows the rank-one construction of a four-material generalization of the

T-structure and Figure 2.14b shows the associated laminate structure. Laminates

of K1 and K3 (and K1 and K4) are brought into rank-one connection with material

K2 respecting (P1)–(P3), allowing for second-rank lamination. The improvements

of the applicability conditions for the method discussed in this chapter over previous

results become more pronounced with larger N .

x

y b

b

b

b

b

bc

bc

r

E4

E3

E2

l

= K4

= K3

= K2

= K1

(a) (b)

Figure 2.14. The rank-one construction of a generalized T-structure (marked bythe square) (a) and the generalized T-structure (b)

Page 65: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

52

2.5 A supplementary bound

In Section 2.4, we showed that a structure is optimal for the bound (2.4)

provided certain conditions (P1)–(P3) hold in an approximate sense. Furthermore,

we constructed optimal structures exactly by satisfying these conditions. The

present section addresses a new issue. Do there exist structures optimal for (2.4)

which are not equivalent to those we have found? (Here we will consider two

structures equivalent if they have the same volume fractions m = (m1, . . . , mN)

and generate the same K∗ ∈ G(m; k).) An answer is not yet known for general

K∗ ∈ G(m; k). In this section, we prove a supplementary bound that any effective

tensor, Keff, must satisfy if it attains the translation bound, provided the interfaces

between materials is smooth. The supplementary bound limits the anisotropy of

such a structure and has an interesting relationship to the T 2-structures discussed

in Section 2.3. For this section, we assume the average field E has been rescaled as

in (2.34).

2.5.1 The bound

Theorem 2.8 Let k = (k1, k2, k3) and m = (m1, m2, m3) be admissible parameters.

Define

γ =

3∑

i=1

mi

ki + k1, δ =

m1k3

k1(k3 + k1)2+

m2

(k2 + k1)2+

m3

(k3 + k1)2. (2.39)

Assume that there exists three open sets Ω1, Ω2, Ω3 with smooth boundaries such

that |Ωi| = mi and3⋃

i=1

Ωi = Ω = [0, 1]2 and assume that there exists a field U ∈

H1#(Ω)2 + Ex which satisfies conditions (i)–(iii) in Theorem 2.7 for F = DU and

for average field E ∈ R2×2 such that

E+ = k1

(3∑

i=1

mi

k1 + k1

)

I = k1γI.

Let λ1 and λ2 be the eigenvalues of the effective tensor Keff. Then one has

(λ1 − λ2

λ1λ2 − k21

)2

≤ 4(γ2 − δ). (2.40)

Page 66: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

53

Remark 2.11 To better understand the nature of (2.40), suppose we have labeled

the eigenvalues so that 0 < λ1 ≤ λ2 and that the right-hand side of (2.40) is

nonnegative. Since we assume that Keff is optimal, the eigenvalues must also satisfy

(see (2.4))λ1 + λ2 − 2k1

λ1λ2 − k21

= 2γ.

Using this together with the inequality (2.40), one finds an upper bound on λ2 in

terms of λ1:

(

γ −√

γ2 − δ)

λ2 ≤(

γ +√

γ2 − δ)

λ1 − 2k1

γ2 − δ.

Before proving Theorem 2.8, we need a bound on the pointwise fields in the

structure

Proposition 2.1 Under the hypotheses of Theorem 2.8, det DU satisfies the in-

equality

det DU ≥ k1k3

(k3 + k1)2in Ω1. (2.41)

Proof : Observe that (i)–(iii) in Theorem 2.7 implies that

det DU =1

4− 1

2|DU−|2 in Ω1 , (2.42)

det DU =k2

1

(ki + k1)2in Ωi for i = 2, 3.

Write the vector function U in terms of its components U = (u, v). Then

|DU−|2 = 12(ux − vy)

2 + 12(uy + vx)

2. Let w = ux − vy and z = uy + vx. Then

since u and v are harmonic in Ω1, so are w and z. We now show that − det DU is

subharmonic in Ω1. Indeed,

∆(− det DU) =1

2∆|DU−|2

=1

2∆(w2 + z2)

= w∆w + |∇w|2 + z∆z + |∇z|2

= |∇w|2 + |∇z|2 ≥ 0.

Therefore, det DU is smooth in Ω1 and satisfies the strong minimum principle there.

Page 67: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

54

On the other hand, the assumption of smooth boundaries implies that the vector

potential satisfies the transmission conditions. If t is the tangent to the boundary

at a point and n is the normal, we have DU · t and DU K · n are continuous

across an interface. As a consequence of the invariance under rotations of the

determinant function we get that the function K det DU must be continuous across

the interfaces.

Using the fact that DU is constant on Ω2 and Ω3 along with the continuity of

K det DU and the minimum principle, we find that

det DU ≥ min

k1k2

(k2 + k1)2,

k1k3

(k3 + k1)2

=k1k3

(k3 + k1)2in Ω1.

Proof of Theorem 2.8: Begin with (2.28), which implies

DU(x)K(x) + k1 cof DU(x) = A = k1I, in Ω.

Integrating this equation on Ω, we have (by the definition of Keff)

EKeff + k1 cof E = k1I

which we can solve for

E = k1cof Keff − k1I

det Keff − k21

.

Assume that the material has been oriented in such a way that Keff = diag(λ1, λ2)

for λ1, λ2 > 0. Then we find

E+ =1

2(E + cof E) =

k1

2

(λ1 + λ2 − 2k1

λ1λ2 − k21

)

I,

E− =1

2(E − cof E) =

k1

2

(λ2 − λ1

λ1λ2 − k21

)(1 00 −1

)

.

Note that E+ = k1γI and the above equation for E+ are consistent since we assume

that the translation bound is attained. That is,

E+ = k1γI =k1

2

(λ1 + λ2 − 2k1

λ1λ2 − k21

)

I ⇐⇒ λ1 + λ2 − 2k1

λ1λ2 − k21

= 23∑

i=1

mi

ki + k1

.

Page 68: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

55

To obtain the desired inequality, consider

2 detE = |E+|2 − |E−|2 = 2k21γ

2 − k21

2

(λ2 − λ1

λ1λ2 − k21

)2

. (2.43)

Using the quasiaffineness of the determinant, we also have

2 detE =

Ω

2 det DU(x) dx =

Ω1

2 det DU(x) dx +

Ω\Ω1

2 det DU(x) dx.

We know det DU in Ω\Ω1 and we use (2.41) in Ω1 to obtain

2 detE ≥ 2m1k1k3

(k3 + k1)2+

2m2k21

(2k2 + k1)2+

m3k21

(k3 + k1)2= 2k2

1δ.

Combining this with (2.43), we obtain (2.40).

2.5.2 Discussion

Theorem 2.8 requires that any “exact structure” with smooth interfaces satisfy

(2.40). One can check that the right-hand side of the inequality vanishes when

m1 = 2Θ(√

m2 − m2) and is negative for

0 ≤ m1 < 2Θ(√

m2 − m2).

Thus no structure with smooth interfaces can attain the bound for this range of

volume fractions. The validity of the supplementary bound for arbitrary structures

is not addressed here. However, observe that this critical relationship of m1 =

2Θ(√

m2 −m2) is exactly the relationship that holds between the volume fractions

of the most extreme version of the isotropic structures introduced in Gibiansky and

Sigmund (2000). (This structure is represented by the point KGS in Figure 2.9.)

It is also interesting to note that if we formally set Keff = diag(λ1, λ2) with λ1, λ2

given in (2.22),(2.23) and set the volume fractions accordingly, then (2.40) holds as

equality. Thus the T 2-structures — which have the smallest values of m1 among

all known optimal structures — while not exact, satisfy the supplementary bound

(2.40) as equality. For all other structures in this chapter which are optimal for

(2.4), the inequality is strict.

Page 69: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

56

2.6 Optimality versus attainability

Here we show the difference between the concepts of optimality and attainability

of an outer bound on the G-closure, as defined in Section 2.2.2. In particular, we

exhibit points on the Wiener bounds (2.2) and (2.3) which are optimal but not

attainable. First we recall the results published by Cherkaev and Gibiansky (1996).

These results can be summarized by the following theorem.

Theorem 2.9 (Cherkaev-Gibiansky) Let

0 < k1 < k2 < k3 and m1, m2, m3 > 0

such that m1 + m2 + m3 = 1 be admissible parameters to the G-closure problem.

Define

k =

3∑

i=1

miki, k =

(3∑

i=1

mi

ki

)−1

.

Then there exist numbers α and β (given explicitly below) such that

α < k and k < β

and such that

(i) The points on the closed line segment joining the points (k, k) and (k, α) are

optimal for the Wiener bound (2.2).

(ii) The points on the closed line segment joining the points (k, k) and (β, k) are

optimal for the Wiener bound (2.3).

The values of α and β are given by the following equations.

α =

α1 if m1 ≥ k1(k3−k2)k3(k2−k1)

m3,

α2 if m1 < k1(k3−k2)k3(k2−k1)

m3,

β =

β1 if m1 ≥ k3−k2

k2−k1m3,

β2 if m1 < k3−k2

k2−k1m3,

Page 70: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

57

where

α1 = µ1k1 +(1 − µ1)

2k2(k1k2 + k2k3 − k1k3)

(k3 − k2)(k2 − k1)m2 + (1 − µ1)k22

, µ1 = m1 −k1(k3 − k2)

k3(k2 − k1)m3,

α2 = µ2k3 +(1 − µ2)

2k2(k1k2 + k2k3 − k1k3)

(k3 − k2)(k2 − k1)m2 + (1 − µ2)k22

, µ2 = m3 −k3(k2 − k1)

k1(k3 − k2)m1,

β1 =

(ν1

k1+

(1 − ν1)2(k1 + k3 − k2)

(k3 − k2)(k2 − k1)m2 + (1 − ν1)k1k3

)−1

, ν1 = m1 −k3 − k2

k2 − k1m3,

β2 =

(ν2

k3

+(1 − ν2)

2(k1 + k3 − k2)

(k3 − k2)(k2 − k1)m2 + (1 − ν2)k1k3

)−1

, ν2 = m3 −k2 − k1

k3 − k2

m1.

Proof : The proof uses one of the iterated laminates illustrated in Figure 2.15.

The important parameter is the relative volume fraction, c, of K1 to K3 in the inner

laminate layer. To find α, one fixes c so that the laminate has the same conductivity

in the x1 direction as K2. This ensures that when this composite is laminated with

K2 in the x2 direction, the current does not jump across the interface. It is easy to

check that

c = cα =k1(k3 − k2)

k2(k3 − k1).

If

m1 ≥cα

1 − cα

m3,

then the structure in Figure 2.15a is used, otherwise the structure in Figure 2.15b

is used.

To find β, one instead chooses c so that the conductivity of the laminate in the

x2 direction is the same as K2. This ensures that when this composite is laminated

with K2 in the x2 direction, the field does not jump across the interface. In this

case, one can check that

c = cβ =k3 − k2

k3 − k1.

If

m1 ≥cβ

1 − cβ

m3,

then the structure in Figure 2.15a is used, otherwise the structure in Figure 2.15b

is used.

Page 71: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

58

= K3

= K2

= K1

(a) (b)

Figure 2.15. The structures from Theorem 2.9.

Remark 2.12 As a particular example, for the parameters of Figure 2.12b, one

has

ν1 =3

8and µ1 =

3

4,

so

m1 > ν1(1 − m2) and m1 < µ1(1 − m2).

Thus, we can calculate α and β from the theorem as

α =79

34≈ 2.3235 < k =

5

2= 2.5,

β =170

89≈ 1.9101 > k =

20

11≈ 1.8182.

On the other hand, we will now give a proof that the only attainable point on

the segments described in (i) and (ii) of the theorem is the point (k, k).

Theorem 2.10 Let

0 < k1 < k2 < k3 and m1, m2, m3 > 0

such that m1 + m2 + m3 = 1 be admissible parameters to the G-closure problem.

Define k and k as in the previous theorem. Then a point Keff is attainable for (2.2)

or (2.3) if and only if Keff has eigenvalues k and k.

Proof : The fact that (k, k) is attainable is easy; we simply use a first-rank

Page 72: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

59

lamination in the x1 direction. Choose the partition

Ω1 = (0, m1) × (0, 1),

Ω2 = (m1, m1 + m2) × (0, 1),

Ω3 = (m1 + m2, 1) × (0, 1).

Then Keff = diag(k, k).

Now assume that there exists a partition of Ω into Ωi such that for K defined

as in (1.5), the effective tensor Keff = diag(λ1, λ2) is such that λ2 = k. That is, we

have

infu∈H1

#(Ω)

Ω

(∇u(x) + e2) · K(x)(∇u(x) + e2) dx = k

where we denote by e1, e2 the canonical basis in R2. Then by uniqueness of weak

solutions to the PDE (1.4), we have u ≡ 0. It follows that for the K defined through

the hypothetical partition, we have

div(K(x)e2) = 0 or∂K(x1, x2)

∂x2

= 0

in the sense of distribution. It is well known that this implies the existence of a

distribution K = K(x1) such that K(x1, x2) = K(x1) as distributions.

Let us now compute the other eigenvalue λ1 by considering the orthogonal

applied field e1. We solve the equation

div(K(x)(∇v(x) + e1)) = 0 in Ω.

Define v ∈ H1#(Ω) such that

v(x) = v(x1), ∇v(x) =k

K(x1)− e1 .

Then

div(K(x)(∇v(x) + e1)) = div

(

K(x1) ·k

K(x1)

)

= 0.

Page 73: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

60

By uniqueness of the solution, we have that

infu∈H1

#(Ω)

Ω

(∇u(x) + e1) · K(x)(∇u(x) + e1) dx

=

Ω

(∇v(x) + e1) · K(x)(∇v(x) + e1) dx

=

Ω

k2

K(x1)dx = k2

3∑

i=1

mi

ki= k,

which proves that λ1 = k. A similar argument shows that if λ1 = k then λ2 = k.

Page 74: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

CHAPTER 3

BLOCK STRUCTURES IN TWO AND

THREE DIMENSIONS

3.1 Introduction

In this chapter, we construct optimal anisotropic two- and three-dimensional

structures that generalize the isotropic structures introduced by Gibiansky and

Sigmund (2000). The structures have a special “block” form in which the periodicity

cell, Ω = [0, 1]d with d = 2, 3 the dimension of the composite, is split into 2d

rectangular blocks with sides parallel to the sides of the cell. Each block contains

a composite of the constituent materials. By choosing special composites for each

block, we are able to easily compute the effective tensor of the overall composite

and ensure that the composite is optimal.

In Section 3.2, we recall the Gibiansky-Sigmund structures, which we use as

inspiration for the new structures introduced in this chapter. We have introduced

these structures already in the previous chapter, Section 2.2.2.2 and illustrated

a special case in Figure 2.5. In this section we discuss the structures in more

detail and explain the special characteristics that allow the conductivity PDE to

be solved by separation of variables. These concepts motivate the rest of the

chapter. In Section 3.3, we describe a simple modification of the Gibiansky-Sigmund

block structures that produces optimal anisotropic structures equivalent to the

T 2-structures introduced previously in Section 2.3.

In Section 3.4, we use the optimal two-dimensional block structures to motivate

a class of three-dimensional block structures optimal for the three-dimensional

version of the translation bounds. We refer the reader to Cherkaev (2000); Milton

(2002) for an overview of the history of the bounds optimal structures. The

Page 75: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

62

structures introduced in this section improve the region of parameters for which

the translation bounds are known optimal.

3.2 Background: the structures ofGibiansky and Sigmund

3.2.1 The structures introduced

In Figure 2.5, we have illustrated the structures introduced by Gibiansky and

Sigmund (2000). The reader should imagine that the two laminate layers are mixed

at an arbitrarily fine scale. This structure is optimal for the lower two-dimensional

translation bound (2.4) if the relative volume fractions in each rectangular block

are chosen correctly. In particular, one should choose the relative volume fractions

of K1 and K3 in the laminate layers to be Θ and 1 − Θ, respectively, where Θ is

the constant defined in (2.9) which depends only on the materials’ properties.

Remark 3.1 In the paper by Gibiansky and Sigmund (2000), the lower-left block

need not contain pure K2 but can also contain inclusions of “coated spheres” of

K1 and K3 so long as the relative amounts of K1 and K3 are chosen so that the

effective conductivity of the coated sphere is exactly K2. The results of this section

apply to these more general structures as well. In fact, by coating the structure in

Figure 2.5 with layers of K1 so that the final structure is isotropic, one can always

generate a structure equivalent to those structures with coated spheres inside the

block of K2. For simplicity we restrict the discussion to the case where no coated

spheres are present.

Let us now assume that the mixture in each rectangular block has been homog-

enized so that the conductivity tensor, K, in Ω = [0, 1]2 is piecewise constant with

the form

K(x) =

k2I if x ∈ (0, ω)× (0, ω)

k3I if x ∈ (ω, 1)× (ω, 1)

diag(µ1, µ2) if x ∈ (0, ω)× (ω, 1)

diag(µ2, µ1) if x ∈ (ω, 1)× (0, ω).

(3.1)

Here ω ∈ (0, 1) is the parameter that defines the sizes of the rectangular blocks,

and µ1 and µ2 are the conductivities of the laminates of K1 and K3. Specifically,

Page 76: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

63

µ1 =

k1+

1 − Θ

k3

)−1

, µ2 = Θk1 + (1 − Θ)k3. (3.2)

We will soon see that we are able to solve the conductivity equation with this

conductivity tensor by separation of variables. First, we must take a short detour

and examine the implications of a conductivity tensor of special form related to

that of Marino and Spagnolo (1969) discussed in the previous chapter.

3.2.2 Separation of variables

Suppose that the conductivity, K, in d dimensions is diagonal on Ω = [0, 1]d

with

K(x) = diag(λ1(x), λ2(x), . . . , λd(x)) (3.3)

such that λ1, . . . , λd have the form

λi(x) = λ0i (xi)λ

′i(x1, . . . , xi−1, xi+1, . . . , xd), for i = 1, . . . , d. (3.4)

A particular case of such a λ resembles (2.10). However, we need the more general

case for the three-dimensional structures to follow.

Now we show that this hypothesis allows us to separate variables to solve the

PDE

div(DU(x) K(x)) = 0 in Ω,

U ∈ H1#(Ω)d + Ex

(3.5)

with E ∈ Rd×d and diagonal: E = diag(r1, . . . , rd). Begin with the ansatz U =

(u1(x1), . . . , ud(xd)) so that

DU = diag

(du1

dx1

, . . . ,dud

dxd

)

.

Then the ith row of the PDE has the form

0 =∂

∂xi

(

λi(x)dui

dxi(xi)

)

= λ′i(x1, . . . , xi−1, xi+1, . . . , xd)

d

dxi

(

λ0i (xi)

dui

dxi(xi)

)

.

Thus, the solution we seek is the U ∈ H1#(Ω)d +Ex that solves the system of ODEs

d

dxi

(

λ0i (xi)

dui

dxi(xi)

)

= 0 in (0, 1), for i = 1, . . . , d.

In particular, if K is piecewise constant, then U is piecewise affine. Now let us

return to the two-dimensional structures in question.

Page 77: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

64

3.2.3 The structures revisited

To understand the relationship of the separation of variables approach to the

structures of Gibiansky and Sigmund, let us consider (3.5) with d = 2 and K given

by (3.1). Assume there is a piecewise affine solution U such that DU is diagonal

and

DU(x) =

A1 if x ∈ (0, ω) × (0, ω)

A3 if x ∈ (ω, 1) × (ω, 1)

A2 if x ∈ (0, ω) × (ω, 1)

A′2 if x ∈ (ω, 1) × (0, ω).

Since we can always rescale the average field, let us assign the field A1 as in the

previous chapter:

A1 =k1

k2 + k1

I.

Then the transmission conditions immediately imply that

A2 = diag

(k1

k2 + k1,

k1k2

µ2(k2 + k1)

)

,

A′2 = diag

(k1k2

µ2(k2 + k1),

k1

k2 + k1

)

.

Now we see the potential for trouble. Since it shares boundaries with both A2

and A′2, A3 has four transmission conditions to satisfy. Looking at the interfaces

separately, we obtain the following two forms for A3.

A3 = diag

(µ1k1

k3(k2 + k1),

k1k2

µ2(k2 + k1)

)

,

A3 = diag

(k1k2

µ2(k2 + k1),

µ1k1

k3(k2 + k1)

)

.

For these to be consistent, we must have µ1µ2 = k2k3. But by multiplying µ1 and

µ2 given in (3.2), we find that this condition is satisfied.

From this, we find the decomposition

K(x) = diag(λ11(x1)λ12(x2), λ21(x1)λ22(x2)).

Page 78: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

65

Indeed, if µ1µ2 = k2k3, then (3.1) admits the decomposition

λ11(x1) =µ1√k3

χ(0,ω)(x1) +√

k3 χ(ω,1)(x1),

λ12(x2) =µ2√k3

χ(0,ω)(x2) +√

k3 χ(ω,1)(x2),

λ21(x1) =µ2√k3

χ(0,ω)(x1) +√

k3 χ(ω,1)(x1),

λ22(x2) =µ1√k3

χ(0,ω)(x2) +√

k3 χ(ω,1)(x2).

Now let us analyze the fields inside the structure. By our earlier assumption,

A1 =k1

k3 + k1

I.

Now consider A3. With µ1 and µ2 given by (3.2), we find

A3 =k1

k3 + k1I.

Turning our attention to the laminate layer A2, we find that

A2 = diag

(k1

k2 + k1

,k1

k3 + k1

)

.

In this laminate, material K1 has volume fraction Θ while K3 has volume fraction

1−Θ. Thus, the average field A2 is achieved through lamination as the average of

the two approximately constant fields of these materials, A21 and A23 respectively.

By writing

A2 = ΘA21 + (1 − Θ)A23

and satisfying the transmission conditions in this laminate, we find that the fields

inside K1 and K3 of the laminate layer are, respectively,

A21 = diag

(k3

k3 + k1

,k1

k3 + k1

)

,

A23 = diag

(k1

k3 + k1,

k1

k3 + k1

)

.

We see immediately that these fields satisfy (P1)–(P3) of the previous chapter,

which explains the optimality of these structures.

Here, then, is the connection between the construction of Gibiansky and Sig-

mund and the special form of a conductivity tensor that allows for the separation

Page 79: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

66

of variables. The following two requirements on the laminate layers of K1 and K3

are equivalent!

(i) The relative volume fractions of K1 and K3 are chosen so that a separation

of variables is possible in the conductivity equation when each rectangular

block is considered homogenized.

(ii) The relative volume fractions of K1 and K3 are chosen so that the local fields

approximately satisfy the optimality conditions (P1)–(P3) of the previous

chapter.

3.3 New optimal structuresin two dimensions

It is not difficult to see that the same arguments can be used to prove the

optimality of a more general form of (3.1). Indeed, with the same laminate of K1

and K3 as before, consider the following modified version of K.

K(x) =

k2I if x ∈ (0, ω1) × (0, ω2)

k3I if x ∈ (ω1, 1) × (ω2, 1)

diag(µ1, µ2) if x ∈ (0, ω1) × (ω2, 1)

diag(µ2, µ1) if x ∈ (ω1, 1) × (0, ω2).

Now there are two parameters ω1, ω2 ∈ (0, 1). K still has the correct form to admit

the separation of variables solution. Furthermore, the calculation of the piecewise

constant approximate fields is unchanged. Thus, such a structure is optimal for any

ω1 and ω2. In fact, it can be shown that these structures are exactly equivalent to

the T 2-structures of the previous chapter. Figure 3.1 illustrates such a structure.

3.4 New optimal structuresin three dimensions

In the rest of this chapter, we will use the idea behind this generalization of the

Gibiansky-Sigmund structure as inspiration for optimal three-dimensional block

structures. These structures are optimal for the lower or upper translation bounds

in three dimensions. These bounds are similar to the two dimensional bounds in

their derivation. In both cases, we make use of a “translator”, φ : R3×3 → R, (the

Page 80: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

67

ω2

ω1

= K3

= K2

= K1

Figure 3.1. The two-dimensional anisotropic optimal block structure.

analogue of det DU in the two-dimensional case) with a convexity property of the

form

φ

(∫

Ω

F (x) dx

)

≤∫

Ω

φ(F (x)) dx

for all F in a certain class of vector fields (curl-free for the lower bound, divergence-

free for the upper bound).

3.4.1 The lower translation bound

The lower translation bound in three dimensions is derived similarly to the

bound (2.4). In this case, the dimension d = 3, so the periodicity cell Ω is the

unit three-dimensional cube and the field matrix E lies in R3×3. In place of the

“translator” det DU from the two-dimensional case, we use a sum of two-by-two

minors:

(trE)2 − tr E2 = 2(e11e22 − e12e21) + 2(e11e33 − e13e31) + 2(e22e33 − e23e32).

This function has the property that

Ω

((trDU)2 − trDU2

)dx = (trE)2 − tr E2, ∀E ∈ R

3×3, ∀U ∈ H1#(Ω)3 + Ex.

This can be seen by integration by parts or by the fact that (trE)2 − trE2 is

polyconvex and therefore quasiconvex. In this case, it is convenient to decompose

R3×3 into the direct sum

R3×3 = M1 ⊕M2 ⊕M3

Page 81: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

68

where

M1 = αI : α ∈ R , M2 =E ∈ R

3×3 : E = ET , tr E = 0

,

M3 =E ∈ R

3×3 : E = −ET

The projections of a matrix E onto these orthogonal subspaces are

E1 :=1

3tr E I, E2 :=

1

2

(E + ET

)− E1 and E3 :=

1

2

(E − ET

)

respectively. It is easy to verify the following identities.

|E|2 = |E1|2 + |E2|2 + |E3|2,

(trE)2 = 3|E1|2,

trE2 =⟨E, ET

⟩= 〈E1 + E2 + E3, E1 + E2 − E3〉

= |E1|2 + |E2|2 − |E3|2.

Remark 3.2 This decomposition was used by Nesi and Rogora (2004) to show that

certain functions (“translators” in our language) are the most efficient for bounds

in homogenization. These translators are the ones used in this chapter to derive

the translation bounds.

As in the two-dimensional case, we fix the matrix E and add and subtract our

translator:

〈E Keff, E〉 = infU∈H1

#(Ω)3+Ex

Ω

(〈DU K, DU〉 + t

((trDU)2 − tr DU2

))dx

−t((trE)2 − tr E2

).

We then relax the problem by replacing U with F ∈ L2(Ω; R3×3) such that∫

ΩF dx =

E so that

〈E Keff, E〉 ≥ infF :

R

ΩF dx=E

Ω

(〈F K, F 〉 + t

((trF )2 − trF 2

))dx

−t((trE)2 − tr E2

).

(3.6)

Thus, using the identities calculated above, a minimizer F must satisfy

K(x) F (x) + 2tF1(x) − tF2(x) + tF3(x) = A a.e. in Ω (3.7)

Page 82: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

69

for some constant A ∈ R3×3. Alternatively, we can use the specific form of K given

in (1.5) and the identities calculated above to rewrite the bound as

〈E Keff, E〉 ≥ infF :

R

ΩF dx=E

N∑

i=1

Ωi

[(ki + 2t)|F1|2 + (ki − t)|F2|2 + (ki + t)|F3|2

]dx

−t((trE)2 − tr E2

).

This bound is not trivially −∞ only if t ∈ [−k1/2, k1]. The bound we consider is

for the choice t = k1. For this choice, we obtain a theorem similar to Theorem 2.7

Theorem 3.1 F is a minimizer of the right-hand-side of (3.6) with t = k1 if and

only if

(i) F1 = 1ki+2k1

(N∑

j=1

mj

kj+2k1

)−1

E1 a.e. in Ωi for i = 1, . . . , N .

(ii) F3 = 1ki+k1

(N∑

j=1

mj

kj+k1

)−1

E3 a.e. in Ωi for i = 1, . . . , N .

(iii)∫

Ω1F2 dx = E2.

(iv) F2 = 0 a.e. in Ωi for i = 2, . . . , N .

Thus, we can write the lower bound as

〈E Keff, E〉 ≥(

n∑

i=1

mi

ki + 2k1

)−1

|E1|2 +

(n∑

i=1

mi

ki + k1

)−1

|E3|2

−k1

((trE)2 − tr E2

), ∀E ∈ R

3×3.

Once again, however, we are more interested in the fields than in the optimal value

of the bound. As in the two-dimensional case, the only freedom in the fields in a

structure which attains this bound is in the value of DU2 in the first material.

3.4.2 Optimal structures for the lower bound

Now we are in a position to produce three-dimensional three-material structures.

Assuming that E is diagonal and properly scaled, the conditions of Theorem 3.1

become the following.

(P1) DU = DUT and tr DU = 1 a.e. in Ω1.

Page 83: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

70

(P2)∫

Ω1DU2 dx = E2 = 1

2(E + ET ) − 1

3tr E I.

(P3) DU = k1

ki+2k1I a.e. in Ωi for i = 2, 3.

As in the two-dimensional case, we will see that the transmission conditions on

DU K are free if we satisfy those on DU . Indeed, assume that U ∈ H1#(Ω)3 + Ex

and that DU satisfies (3.7) with F = DU , so

DU(x) K(x) + 2tDU1(x) − tDU2(x) + tDU3(x) = A a.e. in Ω. (3.8)

Note that

2tDU1 − tDU2 + tDU3 = t trDU I − tDUT .

Taking the divergence of each terms on the right-hand side respectively, we find

(div(t tr DU I))i = t3∑

j=1

∂xj

(trDU I)ij

= t∂

∂xi

(tr DU I)ii

= t

3∑

j=1

∂2uj

∂xi∂xj

and

(div(tDUT )

)

i= t

3∑

j=1

∂xj

(DUT )ij

= t3∑

j=1

∂xj

(DU)ji

= t

3∑

j=1

∂2uj

∂xi∂xj.

Thus, by taking divergence of both sides of (3.8), we find that U automatically

solves the PDE

div(DU(x) K(x)) = 0.

The corresponding statement for laminates is that the jump conditions on currents

are “free” if we satisfy the jump conditions on the fields.

Page 84: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

71

In this chapter, we construct structures in which the unit cube is divided

into eight blocks. Each of the eight blocks will contain either a pure material

or a laminate. If we consider each block to be homogenized, then we will find

conductivity tensors K with the form (See Figure 3.2 for an illustration of the final

structure.)

K(x) =

k2I if x ∈ (0, ω1) × (0, ω2) × (0, ω3),

k3I if x ∈ (ω1, 1) × (ω2, 1) × (ω3, 1),

diag(µ1, µ1, µ2) if x ∈ (0, ω1) × (0, ω2) × (ω3, 1),

diag(µ1, µ2, µ1) if x ∈ (0, ω1) × (ω2, 1) × (0, ω3),

diag(µ2, µ1, µ1) if x ∈ (ω1, 1) × (0, ω2) × (0, ω3),

diag(η1, η2, η2) if x ∈ (0, ω1) × (ω2, 1) × (ω3, 1),

diag(η2, η1, η2) if x ∈ (ω1, 1) × (0, ω2) × (ω3, 1),

diag(η2, η2, η1) if x ∈ (ω1, 1) × (ω2, 1) × (0, ω3).

(3.9)

We perform this construction in the “easy” direction, considering the fields

and ensuring that they satisfy (P1)–(P3) above. There are three blocks which

share a border with pure material K2. For definiteness, let us consider the block

(0, ω1)×(0, ω2)×(ω3, 1). We construct a laminate of K1 and K3 in this block whose

local fields satisfy (P1)–(P3) and whose average field is in rank one connection with

the field in K2 given by (P3). This is a matrix laminate.

First we need a field for K1 which is compatible with the field in K3. We

laminate first in the direction e1 = (1, 0, 0)T , so the required field, E1, in K1 is

E1 = diag

(k3

k3 + 2k1,

k1

k3 + 2k1,

k1

k3 + 2k1

)

.

Laminating this with the field E3 = k1

k3+k1I in K3 and with relative volume fraction

ν1, we obtain a composite with average field

E13 = diag

(ν1k3 + (1 − ν1)k1

k3 + 2k1

,k1

k3 + 2k1

,k1

k3 + 2k1

)

.

We know we need this field to be rank-one connected to the field in K2, so we

choose ν1 so thatν1k3 + (1 − ν1)k1

k3 + 2k1=

k1

k2 + 2k1.

Page 85: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

72

Thus, we find

ν1 =k1(k3 − k2)

(k2 + 2k1)(k3 − k1)≤ 1

3(3.10)

and

E13 = diag

(k1

k2 + 2k1,

k1

k3 + 2k1,

k1

k3 + 2k1

)

.

Now we will laminate this composite with K1 again in the e2 = (0, 1, 0)T

direction. The compatible field for K1 is

E ′1 = diag

(k1

k2 + 2k1

,k1k2 + k1k3 + k2k3

(k3 + 2k1)(k2 + 2k1),

k1

k3 + 2k1

)

.

Laminating this in volume fraction ν2 with E13, we obtain a new composite with

average field

E113 = diag

(k1

k2 + 2k1,ν2(k1k2 + k1k3 + k2k3) + (1 − ν2)k1(k2 + 2k1)

(k3 + 2k1)(k2 + 2k1),

k1

k3 + 2k1

)

.

We bring this into rank-one connection with the field in K2 by choosing ν2 so that

ν2(k1k2 + k1k3 + k2k3) + (1 − ν2)k1(k2 + 2k1)

(k3 + 2k1)(k2 + 2k1)=

k1

k2 + 2k1.

Thus,

ν2 =k1(k3 − k2)

k1k3 + k2k3 − 2k21

=ν1

1 − ν1≤ 1

2(3.11)

and

E113 = diag

(k1

k2 + 2k1,

k1

k2 + 2k1,

k1

k3 + 2k1

)

.

Now we have all the components of the structure and we can construct it by

placing the matrix laminate with average field E113 along with its two orthogonal

rotations into the three blocks neighboring K2. Furthermore, since E13 is rank one

connected to both E113 and to E3, we can fill the remaining three blocks with this

laminate and its two orthogonal rotations. In particular, the laminate with average

field E13 can be placed in the block (0, ω1)× (ω2, 1)× (ω3, 1). The final structure is

illustrated in Figure 3.2. For illustrative convenience, the matrix laminate has

been replaced by an equivalent “coated cylinders” construction. These should

be visualized as a three-dimensional extrusion of the two-dimensional structures

Page 86: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

73

1 − ω2

ω3

ω2

ω1

ω1

ω3

= K3

= K2

= K1

Figure 3.2. The optimal three-dimensional structure (left) and the details of the“hidden” block (right).

illustrated in Figure 2.3a. The block on the right of the figure shows the details of

the “hidden” block behind the block of pure K2.

Now we can compute the volume fractions of the structure. Note that by (3.11),

we get that the volume fraction in the matrix laminate is

ν2 + (1 − ν2)ν1 =ν1

1 − ν1+

ν1(1 − 2ν1)

1 − ν1= 2ν1.

With a little simplification, the volume fractions can be shown to be

m1 = ν1(ω1 + ω2 + ω3 − 3ω1ω2ω3),

m2 = ω1ω2ω3,

m3 = 1 − m2 − m3.

Computing the effective tensor, K∗, is also straightforward, since we know the

approximate piecewise constant field in each piece of the structure. Keeping track

of the transmission conditions further helps simplify the calculation. For example,

consider the eigenvalue λ1 in the e1 direction. We see that

DU(x) · e1 =

k1

k2+2k1if x1 ∈ (0, ω1),

k1

k3+2k1if x1 ∈ (ω1, 1)

and

Page 87: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

74

DU(x) K(x) · e1 =

k1k2

k2+2k1if (x2, x3) ∈ (0, ω2) × (0, ω3),

k1k3

k3+2k1if (x2, x3) ∈ (ω2, 1) × (ω3, 1),

k1(k1k2+k1k3+k2k3)(k2+2k1)(k3+2k1)

if (x2, x3) ∈ (0, ω2) × (ω3, 1),k1(k1k2+k1k3+k2k3)(k2+2k1)(k3+2k1)

if (x2, x3) ∈ (ω2, 1) × (0, ω3).

We define the average field, E, and average current, J , to be

E =

Ω

DU(x) dx and J =

Ω

DU(x) K(x) dx.

From the above calculation, we find that E and J satisfy

E · e1 =ω1k1(k3 + 2k1) + (1 − ω1)k1(k2 + 2k1)

(k2 + 2k1)(k3 + 2k1),

J · e1 =k1k3(k2 + 2k1) − k2

1(k3 − k2)(ω2 + ω3)

(k2 + 2k1)(k3 + 2k1).

Thus,

λ1 =J · e1

E · e1=

k3(k2 + 2k1) − k1(k3 − k2)(ω2 + ω3)

ω1(k3 + 2k1) + (1 − ω1)(k2 + 2k1). (3.12)

Similarly, we can find the other two eigenvalues in the e2 and e3 directions respec-

tively:

λ2 =k3(k2 + 2k1) − k1(k3 − k2)(ω1 + ω3)

ω2(k3 + 2k1) + (1 − ω2)(k2 + 2k1),

λ3 =k3(k2 + 2k1) − k1(k3 − k2)(ω1 + ω2)

ω3(k3 + 2k1) + (1 − ω3)(k2 + 2k1).

In particular, one can verify that when ω1 = ω2 = ω3, then the structure is

isotropic (λ1 = λ2 = λ3 = λ) and is optimal for the three-dimensional version of

the Hashin-Shtrikman bound:

1

λ + 2k1=

3∑

i=1

mi

ki + 2k1.

3.4.2.1 Separation of variables

Now let us return to the tensor K as in (3.9) with the composites inside each

block of Ω considered homogenized already. From the construction above, it is

straightforward to find

µ1 =k1k2 + k1k3 + k2k3

k3 + 2k1, µ2 =

k2(k3 + 2k1)

k2 + 2k1,

η1 =k3(k2 + 2k1)

k3 + 2k1

, η2 =k1k2 + k1k3 + k2k3

k2 + 2k1

.

Page 88: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

75

We now show that this K admits a decomposition as in (3.3) and (3.4) so that we

can solve the PDE by separation of variables. Indeed, if we define

λ01(x1) = k2χ(0,ω1) + µ2χ(ω1,1)

and

λ′1(x2, x3) =

1 if (x2, x3) ∈ (0, ω2) × (0, ω3),µ1

k2if (x2, x3) ∈ (ω2, 1) × (0, ω3),

µ1

k2if (x2, x3) ∈ (0, ω2) × (ω3, 1),

η1

k2if (x2, x3) ∈ (ω2, 1) × (ω3, 1),

then λ1(x) = λ01(x1)λ

′1(x2, x3) exactly because the following two identities hold.

µ1µ2

k2= η2 and

η1µ2

k2= k3.

The other two eigenvalues are analogous. Once again, we see that the conditions for

a separation of variables are identical to those of optimality. Note that λ1 does not

have a decomposition as in (2.10). The more general decomposition is necessary.

3.4.2.2 Applicability: volume fractions

As we did with the two-dimensional structures, let us analyze the applicability

of this construction in terms of volume fraction. The optimal structures of Milton

(1981); Milton and Kohn (1988) applied to three materials in three dimensions

require that the volume fractions satisfy the inequality

3ν1(1 − m2) ≤ m1 ≤ 1

where ν1 is defined in (3.10). On the other hand, if ω1 = ω2 = ω3 in the above

structure, then we see that

m1 = 3ν1( 3√

m2 − m2) (3.13)

which lies outside this region. By using a generalization of the “coating principle”

of the previous chapter, we find a set of optimal anisotropic structures strictly larger

than was previously known. In particular, such structures are possible as long as

3ν1( 3√

m2 − m2) ≤ m1 ≤ 1.

Furthermore, similar three-dimensional constructions are also possible for any num-

ber of materials N ≥ 3. In all cases, one can obtain structures outside the realm

Page 89: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

76

of previously known optimal structures. In fact, as N increases, the improvement

over previous results becomes more and more pronounced.

3.4.3 The upper translation bound

Though the derivation of the upper bound is somewhat different, the calcula-

tions are very similar. In fact, we will see that one can produce optimal structures

for this bound by reversing the roles of K1 and K3 in the previous structures. First

we recall the derivation of the bound, which uses the dual variational principle and

the “translator” φ(B) = tr BBT +trB2− (tr B)2. This functional is A-quasiconvex

on 3 × 3 divergence-free matrices in the sense that

φ

(∫

Ω

B(x) dx

)

≤∫

Ω

φ(B(x)) dx (3.14)

for any divergence-free vector field B : Ω → R3×3. We can use this to obtain bounds

in the dual variational principle as follows. Let Ωi be a partition of Ω and let

K be the conductivity defined through (1.5). Then Keff is characterized by the

following principle.

⟨J (Keff)

−1, J⟩

= infB∈B

Ω

⟨B K−1, B

⟩dx ∀J ∈ R

3×3

where B is the set of divergence-free vector fields with mean J . Now we can add

and subtract φ(B) in the integral and use (3.14) to obtain

⟨J (Keff)

−1, J⟩≥ inf

B∈B

Ω

(⟨B K−1, B

⟩− tφ(B)

)dx + tφ(J) ∀J ∈ R

3×3, ∀t ≥ 0.

Once again, we obtain the bound by relaxing the differential constraint. This time,

we drop the constraint that B is divergence-free, replacing it instead with a vector

field F with average J . The analog of (3.7) for the upper bound is

K−1(x)F (x) + tF1(x) − 2tF2(x) = A a.e. in Ω.

Alternatively, we can rewrite the energy using the same identities as for the lower

bound, which gives

⟨J (Keff)

−1, J⟩≥ inf

F :R

ΩF=J

N∑

i=1

Ωi

[(1

ki+ t

)

|F1|2 +

(1

ki− 2t

)

|F2|2

+1

ki|F3|2

]

dx + tφ(J).

(3.15)

Page 90: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

77

This bound is nontrivial for 0 ≤ t ≤ 12kN

and we will analyze the case t = 12kN

. It

is not difficult to verify the following theorem.

Theorem 3.2 F is a minimizer of the right-hand-side of (3.15) with t = k1 if and

only if

(i) F1 = kikN

ki+2kN

(N∑

j=1

mjkjkN

kj+2kN

)−1

J1 a.e. in Ωi for i = 1, . . . , N .

(ii) F3 = ki

(N∑

j=1

mjkj

)−1

J3 a.e. in Ωi for i = 1, . . . , N .

(iii)∫

ΩNF2 dx = J2.

(iv) F2 = 0 a.e. in Ωi for i = 1, . . . , N − 1.

3.4.4 Optimal structures for the upper bound

Using the constitutive relation DU K = B and rescaling the average field E

as usual, we obtain the following conditions for a the fields DU = B K−1 in a

three-material optimal structure.

(P1′) DU = DUT and tr DU = 1 a.e. in Ω3.

(P2′)∫

Ω3DU2 dx = E2 = 1

2(E + ET ) − 1

3trE I.

(P3′) DU = k3

ki+2k3I a.e. in Ωi for i = 1, 2.

Note that if U ∈ H1#(Ω) + Ex satisfies (P1′)–(P3′) then a similar identity to

(3.8) holds. Specifically,

DU(x) K(x) + 2k3DU1(x) − k3DU2(x) + k3DU3(x) = k3I a.e. in Ω.

Which implies that U is a solution of the PDE and so we once again do not need

to worry about the transmission conditions on the currents DU K. We can repeat

the construction of the structure in Figure 3.2 interchanging the roles of K1 and K3

to find a family of structures optimal for the upper translation bound. By coating

with K3, we obtain a class of structures that generalizes and improves previous

results.

Page 91: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

CHAPTER 4

A DIFFERENTIAL SCHEME: INFINITE-

RANK LAMINATES

4.1 Introduction

In this chapter, we introduce a modified differential scheme and use it to produce

optimal infinite-rank laminates. The traditional differential scheme uses the strat-

egy of inserting infinitesimal inclusions into an existing composite and calculating

the increment of its effective properties. By repeatedly performing this inclusion,

one finds that the effective tensor of the structure changes according to an ordinary

differential equation. If the inclusions can be of several different types at each step,

the differential equation becomes controllable. We refer the reader to such papers

as Bruggemann (1935); Norris (1985); Lurie and Cherkaev (1985); Avellaneda et al.

(1988); Hashin (1988) for a more in-depth look at the general idea of differential

schemes.

In Section 4.2 we introduce a differential scheme for the three-material G-closure

problem using a special type of inclusions. The inclusions are composites of the

constituent materials and are always placed into infinitesimally thin strips parallel

to one of the sides of the periodicity cell. We derive the equations of the differential

scheme, which is controlled by the effective tensors and orientations of these strips.

Furthermore, we derive the equations for the volume fractions of the composite

which change as the scheme progresses. Finally, we introduce an optimal control

problem which is equivalent to finding effective tensors on the boundary of the

G-closure.

In Section 4.3, we modify the differential scheme to allow us to easily solve

the optimal control problem of the previous section in the case where the three-

dimensional translation bound is optimal. The strategy is to ensure that the

Page 92: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

79

optimality conditions on the fields can be satisfied at each infinitesimal step of

the differential scheme. We leave the parameters of the optimal control problem

free until the end of the scheme. Once the structure is produced, it is guaranteed

optimal since the optimality conditions on the fields can be satisfied. By tracking

the effective tensor and volume fractions throughout the scheme, we can easily find

the parameters of the optimal control problem that was solved.

In Section 4.4, we compute the effective tensor of the optimal structures pro-

duced in the previous section. We show that in the isotropic case, these structures

are exactly equivalent to the three-dimensional optimal block structures constructed

in the previous chapter. This is also true of the anisotropic case. The basic strategy

is very similar to the isotropic case, but the calculations are more complex, so we

do not prove it here.

In Section 4.5, we discuss the applicability of the differential scheme method to

other problems. Specifically, similar methods can be applied to the problem with

two dimensions (d = 2) and/or more materials (N ≥ 4). Furthermore, we point

the reader to a discussion of the differential scheme applied to three-dimensional

isotropic polycrystals.

4.2 The differential scheme

The differential scheme described in this section is a method of generating

effective tensors for an inner bound in the G-closure problem. We imagine a

sequence of d-dimensional structures in Ω = [0, 1]d parameterized by µ ≥ 0 with

effective tensors K∗(µ) and with volume fractions mi(µ) for i = 1, . . . , N . We

describe the scheme in d dimensions here and then specialize to d = 2, 3 in the

following sections.

The idea is to describe the infinitesimal transition between the tensors K∗(µ)

and K∗(µ + dµ) with dµ ≪ 1. The basic strategy is as follows. For µ ≥ 0, begin

with the homogenized material with effective tensor K∗(µ) filling Ω. For each

j = 1, . . . , d, we replace the infinitesimal strip

[0, 1] × · · · × [0, 1]︸ ︷︷ ︸

j−1 times

×[0, ρj(µ)dµ] × [0, 1] × · · · × [0, 1]︸ ︷︷ ︸

d−j times

Page 93: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

80

by some new material K∗j (µ). The tensors K∗

j (µ) and parameters ρj(µ) ≥ 0 are

controls of the construction. The ρj are subject to the constraint ρ1(µ) + · · · +

ρd(µ) = 1. Thus, after performing the d replacements, we have replaced a total

volume of dµ of the original material with new material. K∗(µ + dµ) is found by

homogenizing the new conductivity tensor in Ω. Figure 4.1 illustrates a special

example of this procedure for d = 3. Note that the order of replacement does not

matter since the pairwise intersections of the infinitesimal strips have volume of the

order o(dµ).

To simplify the calculations, we choose the K∗j to be diagonal for each µ and we

select the initial tensor K∗(0) to be diagonal. This ensures that K∗(µ) is diagonal

for each µ ≥ 0. We now specialize the scheme to the problem of multimaterial

composites.

Let 0 < k1 < · · · < kN be the constituent material properties and fix µ ≥ 0.

Let ρ1(µ), . . . , ρd(µ) ≥ 0 be such that ρ1(µ) + · · · + ρd(µ) = 1. Let νij(µ) ≥ 0 for

i = 1, . . . , N , j = 1, . . . , d be such that ν1j (µ) + · · · + νN

j (µ) = 1 for j = 1, . . . , d.

As described earlier, the ρi control the thickness of the layers replaced in each of

the d directions. The νij act as controls for the volume fractions in the replacement

layers where νij represents the relative amount of the material with conductivity Ki

present in the jth layer. Specifically, we choose

K∗j ∈ G

((ν1

j (µ), . . . , νNj (µ)) ; (k1, . . . , kN)

)∀j = 1, . . . , d.

1

3dµ

Kadd

Kcore

Figure 4.1. The differential scheme for isotropic composites in three dimensions.

Page 94: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

81

4.2.1 The equations for the volume fractions

Suppose that for some fixed µ ≥ 0, we have successfully constructed a composite

with effective tensor K∗(µ) and with volume fractions m1(µ), . . . , mN (µ). Let us

analyze the change in volume fraction after an infinitesimal volume dµ has been

replaced by the differential scheme described above. Since the volume of the original

material, K∗(µ), in the new composite is 1 − dµ, we find that

mi(µ + dµ) = (1 − dµ)mi(µ) + dµ

d∑

j=1

ρj(µ)νij(µ)

= mi(µ) + dµ

(d∑

j=1

ρj(µ)νij(µ) − mi(µ)

)

.

Taking dµ → 0, we find the differential equation satisfied by the volume fractions:

dmi

dµ=∑d

j=1 ρjνij − mi,

mi(0) = m0i

∀i = 1, . . . , N

where m0i is the initial volume fraction of the ith material in the “seed” composite

with effective tensor K∗(0). Thus, we find the volume fractions parameterized by

µ are

mi(µ) = m0i e

−µ +

∫ µ

0

d∑

j=1

ρj(ξ)νij(ξ)e

ξ−µ dξ ∀i = 1, . . . , N. (4.1)

4.2.2 The equation for the effective tensor

Now we consider the differential equation of the effective tensor K∗. By our

assumptions, we have ensured that K∗ and the K∗i are diagonal for every µ. To

simplify notation, we write

K∗(µ) = diag(λ1(µ), . . . , λd(µ))

and

K∗j (µ) = diag(ηj

1(µ), . . . , ηjd(µ)) ∀j = 1, . . . , d.

First consider the case when ρ1(µ) = 1 and ρ2(µ) = · · · = ρd(µ) = 0 so that only

one infinitesimal strip is replaced. In this case, the lamination formula shows that

λ1(µ + dµ) =

(1 − dµ

λ1(µ)+

η11(µ)

)−1

= λ1(µ) + dµ λ1(µ)

(

1 − λ1(µ)

η11(µ)

)

+ o(dµ)

Page 95: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

82

and

λl(µ + dµ) = (1 − dµ)λl(µ) + dµ η1l (µ) = λl(µ) + dµ(η1

l (µ) − λl(µ))

for l = 2, . . . , d. From this, it is straightforward to find the differential equation

satisfied by λ1 for general ρj :

dλ1

dµ= ρ1λ1

(

1 − λ1

η11

)

+∑d

j=1 ρj(ηj1 − λ1),

λ1(0) = λ01

(4.2)

where λ01 is the first eigenvalue of K∗(0). The equations for the other eigenvalues

are similar.

4.2.3 An optimal control problem

One way to characterize the optimal structures in the two- and three-dimensional

problems discussed in the previous chapters (see for example Cherkaev (2000)) is

to solve the following optimal control problem (or the related maximum problem).

min λ1(µf)

subject to the constraints of the differential equations above and

λj(µf) = βj ∀j = 2, . . . , d,

mi(µf) = Mi ∀i = 1, . . . , N

where the βj and Mi are problem parameters. The controls for the optimization

problem are the measurable functions ρj, νij : [0, µf ] → [0, 1] that are subject to the

constraints

N∑

j=1

ρj = 1 a.e. µ,

N∑

i=1

νij = 1 a.e. µ ∀i = j, . . . , d.

Rather than attack the problem directly at this point, we instead make the

following observation. If the effective tensor K∗(µf) is optimal for one of the bounds

Page 96: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

83

we have discussed in the previous chapters (with the volume fractions given by the

Mi), then K∗(µf) is a solution to the optimal control problems with the parameters

chosen appropriately. The next section shows how simplify the control problem by

ensuring that the structure saturates the translation bound. Such a structure is

necessarily optimal. After we have completed the construction, we can easily find

the βj and Mi of the optimal control problem the structure solves.

4.3 The modified differential scheme

In this section, we describe a convenient method for finding optimal multimate-

rial structures in three dimensions, using a modification of the differential scheme

of the last section. The scheme is modified to incorporate the optimality conditions

of the fields at each step. To keep the calculations simple, we will construct only

isotropic composites by requiring

ρ1(µ) = ρ2(µ) = ρ3(µ) =1

3, (4.3)

νi1(µ) = νi

2(µ) = νi3(µ) = νi(µ) for i = 1, 2, 3, (4.4)

K∗1(µ) = diag(ηn(µ), ηt(µ), ηt(µ)), (4.5)

K∗2(µ) = diag(ηt(µ), ηn(µ), ηt(µ)), (4.6)

K∗3(µ) = diag(ηt(µ), ηt(µ), ηn(µ)). (4.7)

The scheme generalizes easily to anisotropic structures, however, and produces a

set of structures equivalent to those of the previous chapter.

Consider a three-material translation-optimal composite. After rescaling the

average field, the fields in the phases are described by (P1)–(P3) of the previous

chapter, which we repeat here for the reader’s convenience.

(P1) DU = DUT and tr DU = 1 a.e. in Ω1.

(P2)∫

Ω1DU2 dx = E2 = 1

2(E + ET ) − 1

3tr E I.

(P3) DU = k1

ki+2k1I a.e. in Ωi for i = 2, 3.

The construction of optimal isotropic structures proceeds as follows. We begin with

an initial core of the second material K2, which we assume is subject to an average

Page 97: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

84

isotropic field k1

k2+2k1I. Then at each step in the differential scheme process we add

three orthogonal layers of a transversely isotropic composite, K13, of materials K1

and K3 formed by placing a cylindrical inclusion of K3 into a matrix of K1.

At the same time, we alter to average field so that the field in the core does

not change. We further choose the composite K13 to ensure that (P1)–(P3) can be

satisfied. The infinitesimal layers are always oriented so that the axis of cylindrical

inclusions in K13 coincides with normal to the layer.

Remark 4.1 While it is impossible to describe the infinite-rank laminate in finite

length scales, the following coated spheres structure is useful for visualizing its main

features. A spherical core of K2 is surrounded by a spherical layer of K1 stuffed

with radially-oriented conical inclusions of K3. The cones become thicker with the

increase of the radius.

4.3.1 Matrix laminates in replaced strips

The three-dimensional transversely isotropic extremal structure can be assem-

bled either as coated cylinders, or as second-rank matrix laminates, or as Vigdergauz-

type structures (see for example Cherkaev (2000); Milton (2002)). In all cases, the

volume fractions can be chosen in such a way that it is possible to satisfy (P1)–(P3).

We show the calculation using the matrix laminate structure here. We do this only

for the composite that will fill the strip [0, dµ/3] × [0, 1] × [0, 1]. The composites

for the other strips are simply rigid rotations of this composite (see (4.3)-(4.7)).

We begin by laminating K1 and K3 in the x2-direction. In order to satisfy

(P1)–(P3) and the rank-one connectedness condition, the fields in K1 and K3 must

be

E1 = diag

(k1

k3 + 2k1,

k3

k3 + 2k1,

k1

k3 + 2k1

)

and E3 =k1

k3 + 2k1I

respectively. Laminating these two fields with relative proportions ν11 and 1− ν11,

we obtain a laminate composite with average field

E13 = diag

(k1

k3 + 2k1

,ν11k3 + (1 − ν11)k1

k3 + 2k1

,k1

k3 + 2k1

)

.

Page 98: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

85

Now we laminate this composite with K1 in the x3 direction. The field in

this new layer of K1, which satisfies the rank-one connectedness condition and the

optimality conditions on the fields is

E ′1 = diag

(k1

k3 + 2k1,ν11k3 + (1 − ν11)k1

k3 + 2k1,ν11k1 + (1 − ν11)k3

k3 + 2k1

)

.

Laminating this with the composite with average field E13 in relative proportions

ν12 and 1 − ν12, we obtain a matrix laminate with average field

E113 = diag

(k1

k3 + 2k1,ν11k3 + (1 − ν11)k1

k3 + 2k1,(1 − ν12(1 − ν11))k1 + ν12(1 − ν11)k3

k3 + 2k1

)

.

We will keep the average field in the core proportional to the identity, so for E113

to be rank-one connected to the core, we need first of all that the fields in the x2

and x3 direction are equal:

ν11k3 + (1 − ν11)k1

k3 + 2k1=

(1 − ν12(1 − ν11))k1 + ν12(1 − ν11)k3

k3 + 2k1

which requires

ν12 =ν11

1 − ν11.

The field in the infinitesimal strip is controlled by the single parameter ν11 and is

given by

E113 = diag

(k1

k3 + 2k1,ν11k3 + (1 − ν11)k1

k3 + 2k1,ν11k3 + (1 − ν11)k1

k3 + 2k1

)

.

The volume fraction ν1 of K1 in this composite is given by

ν1 = ν12 + (1 − ν12)ν11 =ν11

1 − ν11+

ν11(1 − 2ν11)

1 − ν11= 2ν11. (4.8)

Furthermore, since we know piecewise approximations to the fields, we can find

the effective tensor for this composite to be

K∗1 = (ηn, ηt, ηt), (4.9)

ηn = 2ν11k1 + (1 − 2ν11)k3, (4.10)

ηt =k1(ν11k1 + (1 − ν11)k3)

ν11k3 + (1 − ν11)k1

. (4.11)

Page 99: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

86

4.3.2 The optimal controls

Now we are in a position to produce optimal controls for the problem described

in Section 4.2.3. As mentioned previously, we initiate the process with a pure block

of material K2

K∗(0) = K2, m2(0) = 1, m1(0) = m3(0) = 0.

Recall that to simplify calculations, we have chosen to keep K∗ isotropic for all µ,

so we may write K∗(µ) = λ(µ)I for a scalar function λ. In order to check (P1)–(P3)

for each µ, we also track the average field, E(µ), in the homogenized composite,

K∗(µ). In fact, it will be enough to assume E(µ) = γ(µ)I for all µ. To satisfy

(P1)–(P3) at µ = 0, we set

γ(0) =k1

k2 + 2k1

.

Suppose for some µ ≥ 0, we have constructed a composite in which we can

satisfy the field conditions (P1)–(P3) and that the average field for which these

conditions are satisfied is E(µ) = γ(µ)I = γI. To find the composites for the

replacement strips for the next infinitesimal step, we need to choose ν11 so that the

field E(µ) and E113 are rank-one connected. Solving, we find

ν11 =γ(k3 + 2k1) − k1

k3 − k1. (4.12)

Of course, we must make sure that we satisfy the physical constraints 0 ≤ ν11 ≤ 1,

which requiresk1

k3 + 2k1≤ γ ≤ k3

k3 + 2k1. (4.13)

We will show later that this constraint is always satisfied. For now, note that γ(0)

satisfies it and so we expect to be able to proceed for at least some small range of

µ.

Using (4.12) together with (4.9)-(4.11), we find

ηn = (1 − 2γ)(k3 + 2k1), (4.14)

ηt = k1

(k3 + k1

γ(k3 + 2k1)− 1

)

. (4.15)

Page 100: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

87

Provided we know γ(µ), then, we can find the optimal controls. By analyzing the

transmission conditions on the fields, we find that γ changes at each infinitesimal

step as

γ(µ + dµ) =

(

1 − dµ

3

)

γ(µ) +dµ

3

k1

k3 + 2k1

= γ(µ) + dµ

(1

3

(k1

k3 + 2k1

− γ(µ)

))

.

Thus, with the infinitesimal strips chosen in this way, γ satisfies the differential

equation

dγdµ

= 13

(k1

k3+2k1− γ)

,

γ(0) = k1

k2+2k1.

Solving, we find

γ(µ) = e−µ/3 k1

k2 + 2k1

+ (1 − e−µ/3)k1

k3 + 2k1

. (4.16)

This combined with (4.14) and (4.15) provides the optimal control for the problem

until the final structure is obtained at µ = µf provided (4.13) is satisfied for every

µ ∈ [0, µf ]. But these inequalities hold for any µ ≥ 0 since for such a µ, γ(µ) is a

convex sum of two quantities, each of which satisfies the inequalities.

4.4 The optimal structures

From (4.2), we know that K∗(µ) = λ(µ)I satisfies the differential equation

dλdµ

= 13λ(

1 − ληn

)

+ 23(ηt − λ),

λ(0) = k2

(4.17)

where ηn, ηt are functions of µ defined through (4.14) and (4.15) respectively which

depend on γ(µ) defined in (4.16). Suppose λ(µ) is a solution to this equation. Then

for any µf ≥ 0, the corresponding structure with effective tensor K∗(µf) = λ(µf)I

is optimal for the three dimensional translation bound.

Let us now make more precise the earlier statement that these structures are

equivalent to those discussed in the previous chapter. Consider the isotropic case

of (3.12) where ω1 = ω2 = ω3 = 3√

m2 and K∗ = λI. Then

λ =k3(k2 + 2k1) − 2 3

√m2 k1(k3 − k2)

3√

m2(k3 + 2k1) + (1 − 3√

m2)(k2 + 2k1). (4.18)

Page 101: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

88

By (4.1), we find

m2(µ) = e−µ.

Substituting this into (4.18), we find that the function

λ(µ) =k3(k2 + 2k1) − 2e−µ/3 k1(k3 − k2)

e−µ/3(k3 + 2k1) + (1 − e−µ/3)(k2 + 2k1)

solves (4.17).

The final quantity to check is the volume fraction m1. From (4.1), we find

m1(µ) =3k1(k3 − k2)

(k2 + 2k1)(k3 − k1)

(e−µ/3 − e−µ

)

=3k1(k3 − k2)

(k2 + 2k1)(k3 − k1)

(3√

m2(µ) − m2(µ))

.

Comparing this to (3.13), we see that the modified differential scheme produces

isotropic structures with exactly the same effective tensors and volume fractions as

the three-dimensional block structures described in the previous chapter.

4.5 Discussion

This construction is easily extended to larger numbers of materials (N > 3).

Choosing the initial core to be K2 was convenient, but we may start with any core

we wish as long as (P1)–(P3) are satisfied there. Additionally, instead of adding

a matrix laminate at any given step, we can also add a layer of pure K1 because

the optimal field in this material can always be brought into rank-one connection

with whatever field is in the core. As we mentioned before, there is no need to

require that the final structure be isotropic. Instead, we can choose the ρj ≥ 0 with

ρ1 + ρ2 + ρ3 = 1 and materials in each infinitesimal strip however we like as long

as we can satisfy the field requirements. The general scheme is as follows.

At any step and in any direction, we may add one of up to N different types of

layers: N − 1 types of composites of some Ki in a matrix of K1 or a layer of pure

K1. Some bookkeeping is required, but the idea is straightforward. At any step we

can add either pure K1 or any matrix laminate for which we can choose the volume

fractions to satisfy (P1)–(P3). The general scheme produces structures equivalent

to all currently known structures optimal for the translation bound. Furthermore,

Page 102: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

89

this scheme is not limited to dimension d = 3. Indeed for d = 2, the scheme also

produces structures equivalent to all currently known structures optimal for the

two-dimensional translation bound.

Finally, we remark that the differential scheme can be used for other types of

problems. In particular, we point the reader to Albin and Cherkaev (2006) where a

similar scheme was applied to the example of generating optimal three-dimensional

polycrystals. The result was consistent with the earlier analysis of the fields in these

structures due to Nesi and Milton (1991) and shows that the modified differential

scheme may be useful in a larger class of optimization problems.

Page 103: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

CHAPTER 5

NUMERICALLY ESTIMATING THE

G-CLOSURE

5.1 Introduction

In this chapter, we introduce a numerical method for estimating the G-closure

from inside. The method is a finite-dimensional version of the differential scheme

discussed in the last chapter.

In Section 5.2, we introduce the finite scheme, which produces hierarchical

laminates by alternately laminating the current composite to the side and to the

top with simple laminates and re-homogenizing. The evolution of the structures are

controlled at each step by six parameters. In Section 5.3, we rephrase the optimal

control problem from the previous chapter in terms of the finite scheme.

In Section 5.4, we illustrate some numerical results obtained by combining

the finite scheme with a nonlinear optimizer to seek a numerical solution to the

optimization problem. We apply the scheme to the G-closure problem of three ma-

terials in two dimensions and to the problem of finding the most resistive isotropic

structure for a range of volume fractions. In the latter case, we also plot the local

fields in the laminates and discuss some interesting properties they exhibit.

5.2 A finite scheme for laminate structures

In this section, we discuss an algorithm for obtaining a numerical inner bound

on the G-closure using a finite version of the differential scheme introduced in the

last chapter. We work with the two-dimensional problem and with three materials

whose isotropic conductivities are given by k1, k2, k3 > 0. The generalizations to

higher dimension and more materials are obvious.

Page 104: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

91

The basic step for the scheme consists of two parts and depends on six param-

eters: µ1, µ2, ν11 , ν

12 , ν

21 , ν

22 ∈ [0, 1] with the additional constraint

ν1k + ν2

k ≤ 1 k = 1, 2.

The νik define two auxiliary laminates in a similar way to the differential scheme.

Specifically, νik gives the volume fraction of the ith material in the kth auxiliary

laminate for i, k = 1, 2.

For k = 1, the materials are laminated with vertical normal n = (0, 1)T forming

a material with eigenvalues

η11 = ν11k1 + ν2

1k2 + (1 − ν11 − ν2

1)k3, η12 =

(ν1

1

k1+

ν21

k2+

1 − ν11 − ν2

1

k3

)−1

directed in the horizontal and vertical directions respectively. A similar construction

for k = 2 is done with all normals rotated 90 degrees. This gives a laminate

materials with eigenvalues

η21 =

(ν1

2

k1+

ν22

k2+

1 − ν12 − ν2

2

k3

)−1

, η22 = ν12k1 + ν2

2k2 + (1 − ν12 − ν2

2)k3

directed in the horizontal and vertical directions respectively.

The two parts of the basic step, which resembles a step of the differential scheme

without the assumption of the smallness of the change, are as follows.

(I) Laminate the current composite with the first laminate (eigenvalues η11 and

η12) in the x1 direction with relative fractions 1 − µ1 and mu1 respectively.

Homogenize the new composite.

(II) Laminate the current composite with the second laminate (eigenvalues η21

and η22) in the x2 direction with relative fractions 1−µ2 and µ2 respectively.

Homogenize the new composite.

This basic step can be repeated any number of times to produce a composite.

We will classify the different schemes obtained as N-step schemes. The one-step

scheme begins with a seed material and produces a new composite by performing (I)

and then (II); the two-step scheme begins with the seed and performs (I), then (II),

Page 105: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

92

then (I) and then (II); and so on. From a theoretical standpoint, the seed material

is irrelevant (unless it cannot be obtained by the N -step scheme to begin with)

since the choice of µ1 = 1 or µ2 = 1 are allowed in (I) and (II) which completely

replaces the existing material.

5.3 The optimization problem

Now, motivated by the differential scheme of the previous chapter, we fix num-

bers β > 0, M1, M2 ≥ 0 such that M1 + M2 ≤ 1 and consider the following

minimization problem (or the corresponding maximization problem)

min λ1

subject to the constraints

λ2 = β, m1 = M1, m2 = M2.

The controls of the problem are the values of the µk and νik at each step and λ1, λ2

are the eigenvalues of the effective tensor of the produced composite with volume

fractions given by m1, m2. λ1, λ2, m1, m2 are computed from the µk and νik through

the lamination formula applied by (I) and (II). For the constraints to be satisfied,

we must have that β lies inside the arithmetic-harmonic mean bounds:

(3∑

i=1

Mi

ki

)−1

≤ β ≤3∑

i=1

Miki

where M3 = 1 − M1 − M2.

5.4 Numerical results

In this section, we present numerical results for the optimization problem. The

finite scheme was implemented in the C programming language and the optimiza-

tion was performed by the nonlinear optimizer DONLP21. The initial guess was

chosen to be the T 2-structure (not necessarily optimal) with the correct volume

fractions. Experimentally, we found that the nonlinear optimizer worked best if we

1http://www.netlib.org/opt/donlp2/

Page 106: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

93

slowly increased the number of degrees of freedom. For this reason, we began by

optimizing the one-step scheme then used the results to seed the two-step scheme

and used this to seed the three-step scheme. In passing from the (N − 1)-step

scheme to the N -step scheme, we retain the previously optimized structure as a

small core (about 1% of the new material). The rest of the composite is made of

two additional laminate layers, which maintain the volume fraction constraints.

5.4.1 Numerical estimates of G-closures

In Figures 5.1–5.2, we show the results after the three-step scheme has been

numerically optimized in this way. In all cases the conductivities for the problem

are

k1 = 1, k2 = 2, k3 = 5,

and β ranged through 50 values between the arithmetic and harmonic mean. Both

the minimization and maximization problems were performed.

Figure 5.1 displays the results corresponding to Figures 2.2–2.1. Recall that the

volume fractions for this problem are

M1 = 0.4, M2 = 0.01.

The numerically optimized conductivity tensors are indicated by dots. The arith-

metic/harmonic mean bounds and the translation bounds are indicated by dashed

lines. Note that the translation bounds are difficult to see since they are covered

by the dots.

Figure 5.2 displays the results corresponding to Figure 2.12. The volume frac-

tions are

M1 = 0.104, M2 = 0.5 (5.1)

and

M1 = 0.25, M2 = 0.5

respectively. In addition to the numerically optimized points indicated by solid

dots, we also show the inner and outer bounds on the G-closure discussed in

Section 2.3.10. Observe that the numerical scheme does better than the naive

analytic inner bound.

Page 107: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

94

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

bbb

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

bbb

b

Figure 5.1. A numerical estimation of the G-closure for M1 = 0.4, M2 = 0.01.

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

bbbbb

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

bbbb

b

λ1

λ2

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

bbbbb

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

bbbb

λ1

λ2

(a) (b)

Figure 5.2. A comparison of numerical results to known inner bounds.

Page 108: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

95

5.4.2 Numerical estimates of the isotropic bound

In Figures 5.3–5.4, we consider a similar problem. In this case, m1 is allowed to

vary and the constraints in the earlier optimization problem are replaced by

λ2 = λ1, m2 = p(1 − m1)

for some p ∈ [0, 1]. This problem characterizes the most resisting isotropic com-

posite for the various volume fractions given through m1. The figures illustrate the

results for p = 0.5 and p = 0.1 respectively. For reference, we have also included

the limits of construction (minimum m1) for the results of Milton (1981); Gibiansky

and Sigmund (2000), marked by the points M and GS respectively. For values of

m1 > 0 and below the GS point, it is not known whether the translation bound is

optimal. Observe that the numerical scheme tracks the translation bound when it

is known to be optimal and then leaves the bound (but remains very close!) beyond

this point.

5.4.3 Fields in numerically optimized structures

Finally, we remark that the fields in the numerically optimized structures can be

found automatically by satisfying the jump conditions in each layer. In Figures 5.5–

5.5 we illustrate the results corresponding to the second case above (m2 = 0.1(1 −m1)). We have split the data into two parts. Figure 5.5 displays the fields when the

translation bound is known to be optimal. Figure 5.6 displays the other cases. The

data in both figures come from numerically optimizing structures for 50 equally-

spaced values of m1 in the respective range. The fields are computed by assuming

an isotropic average field scaled so that the average field in the third material is

the same for each composite.

Figure 5.5 is much as we expect from the field optimality conditions. The fields

tend to cluster about two points and a line. Once we move from this region of m1,

however, the fields that were free to move along the line suddenly appear to cluster

about two points while the fields in the second material gain a degree of freedom.

This gives some indication of where to look for new bounds on the G-closure. We

expect the field optimality conditions to have the following properties.

Page 109: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

96

u

u

M

GS

bbbbbbbbb

bb

bb

bb

bb

bb

bb

bb

bb

bb

bb

bb

bb

bb

bb

bb

b

bb

bb

b

b

b

b

b

b

m1

λ

Figure 5.3. The effective conductivity of numerically optimized isotropic resistorsfor m2 = 0.5(1 − m1). The points M and GS represent the limits of applicabilityof the results of Milton, and Gibiansky and Sigmund, respectively.

u

u

M

GS

bb

bb

bb

bb

bb

bb

bb

bb

bb

bb

bb

bb

bb

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

b

m1

λ

Figure 5.4. The effective conductivity of numerically optimized isotropic resistorsfor m2 = 0.1(1 − m1). The points M and GS represent the limits of applicabilityof the results of Milton, and Gibiansky and Sigmund, respectively.

Page 110: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

97

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

local diagonal fieldsconstant trace line

isotropic line

Figure 5.5. The local fields in the numerically optimized structures used inFigure 5.4 in the case when the bounds are known to be optimal

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

local diagonal fieldsconstant trace line

isotropic line

Figure 5.6. The local fields in the numerically optimized structures used inFigure 5.4 in the case when the bounds are not known to be optimal

Page 111: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

98

1. The field in K3 is constant and isotropic.

2. The field in K1 is always rank-one connected to the constant isotropic field

of K3

3. The field in K2 exhibits a degree of “freedom,” which allows it to take values

along some manifold. (This manifold should degenerate to the line required

by the two-material translation bounds when m1 = 0.)

Page 112: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

CHAPTER 6

CONCLUSION

We have seen that the use of the sufficient conditions on pointwise fields allows

us to methodically generate structures optimal for the translation bound from a very

general class of structures. We illustrated this in Chapter 2 for two-dimensional

iterated laminates, in Chapter 3 for a class of two- and three-dimensional block

structures and in Chapter 4 for a class of infinite rank-laminates generated by a

controllable differential scheme.

Each of these classes of structures is extremely rich and nonunique in the sense

that several different structures produce the same effective tensor. This makes it

difficult to guess the best members of the class. On the other hand, by requiring

that the sufficient conditions on the fields are satisfied, we easily generate such

members from each class. In fact, since we require the field optimality conditions

to hold at each step of the process, the approach cannot create a composite which

is not optimal. Furthermore, up to a “coating argument,” this method produces

equivalent optimal structures from each general class of structures. The method

not only reproduces previous results, but also shows that the G-closure for multi-

material conducting composites in two or three dimensions is strictly larger than

was previously known.

In Chapter 5, we introduced a simple numerical scheme for optimizing a class of

structures. This scheme proves very effective in several examples of estimating an

inner bound on the G-closure. Furthermore, when applied to the problem of best

resisting isotropic composites, it is consistent with known results and gives some

idea of field optimality conditions that would arise in a bound improving on the

translation bound.

Page 113: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

100

In Section 2.5, we presented some evidence that the classification of optimality of

the translation bounds for three-phase composites in the plane is nearly known. We

introduced a supplementary bound on the anisotropy of a structure which attains

the translation bound and satisfies certain conditions. We conjecture that this

bound can also be shown to hold for general three-phase composites. If this is true,

then the question of attainability of the translation bound for this problem will be

completely closed for isotropic composites.

The achievement of this dissertation has been to show that a very careful

analysis of the field optimality conditions in several general classes of laminate

composites provides a pronounced improvement of the known region of optimality

of the translation bounds. We have shown the effectiveness of such an approach for

the translation bounds in linear conductivity with isotropic phases. We believe that

such methods are also possible for a much wider range of physical phenomena. In

particular, as we mentioned earlier (Remark 2.10), the work of Grabovsky (1996)

allows these results to immediately apply to certain cases of linear elasticity. Our

hope is that the work presented in this dissertation will give rise to new systematic

methods for classifying G-closure problems.

Page 114: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

REFERENCES

N. Albin and A. Cherkaev (2006), Optimality conditions on fieldsin microstructures and controllable differential schemes, (preprint)http://www.math.utah.edu/∼albin/.

G. Allaire (2002), Shape optimization by the homogenization method, volume146 of Applied Mathematical Sciences, Springer-Verlag, New York.

K. Astala and M. Miettinen (1998), On quasiconformal mappings and 2-dimensional G-closure problems, Arch. Rational Mech. Anal., 143(3), pp. 207–240.

K. Astala and V. Nesi (2003), Composites and quasiconformal mappings: newoptimal bounds in two dimensions, Calc. Var. Partial Differential Equations,18(4), pp. 335–355.

M. Avellaneda, A. V. Cherkaev, L. V. Gibiansky, G. W. Milton, andM. Rudelson (1996), A complete characterization of the possible bulk andshear moduli of planar polycrystals, Journal of the Mechanics and Physics ofSolids, 44(7), pp. 1179–1218.

M. Avellaneda, A. V. Cherkaev, K. A. Lurie, and G. W. Milton(1988), On the effective conductivity of polycrystals and a three-dimensionalphase interchange inequality, Journal of Applied Physics, 63, pp. 4989–5003.

M. Avellaneda and G. W. Milton (1989), Optimal bounds on the effectivebulk modulus of polycrystals, SIAM J. Appl. Math., 49(3), pp. 824–837.

A. Braides and A. Defranceschi (1998), Homogenization of multiple integrals,volume 12 of Oxford Lecture Series in Mathematics and its Applications, TheClarendon Press Oxford University Press, New York.

M. Briane (1994), Correctors for the homogenization of a laminate, Adv. Math.Sci. Appl., 4(2), pp. 357–379.

D. A. G. Bruggemann (1935), Berechnung verschiedener physikalis-cher Konstanten von heterogenen Substanzen. I. Dielectrizitatkonstanten undLeitfahigkeiten der Mischkorper aus isotropen Substanzen, Annalen der Physik(1900), 22, pp. 636–679.

A. Cherkaev (2000), Variational methods for structural optimization, volume140 of Applied Mathematical Sciences, Springer-Verlag, New York.

Page 115: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

102

A. Cherkaev and L. Gibiansky (1996), Extremal structures of multiphase heatconducting composites, Int. J. Solids Structures, 33(18), pp. 2609–2623.

S. Conti, D. Faraco, and F. Maggi (2005a), A new approach to counterex-amples to L1 estimates: Korn’s inequality, geometric rigidity, and regularity forgradients of separately convex functions, Arch. Ration. Mech. Anal., 175(2), pp.287–300.

S. Conti, D. Faraco, F. Maggi, and S. Muller (2005b), Rank-one convexfunctions on 2 × 2 symmetric matrices and laminates on rank-three lines, Calc.Var. Partial Differential Equations, 24(4), pp. 479–493.

G. Dal Maso (1993), An introduction to Γ-convergence, Progress in NonlinearDifferential Equations and their Applications, 8, Birkhauser Boston Inc., Boston,MA.

G. Dolzmann (2003), Variational methods for crystalline microstructure—analysis and computation, volume 1803 of Lecture Notes in Mathematics,Springer-Verlag, Berlin.

L. V. Gibiansky and O. Sigmund (2000), Multiphase composites with extremalbulk modulus, J. Mech. Phys. Solids, 48(3), pp. 461–498.

Y. Grabovsky (1996), Bounds and extremal microstructures for two-componentcomposites: a unified treatment based on the translation method, Proc. Roy. Soc.London Ser. A, 452(1947), pp. 919–944.

Y. Grabovsky and R. V. Kohn (1995), Microstructures minimizing the energyof a two phase elastic composite in two space dimensions. II: the Vigdergauzmicrostructure, Journal of the Mechanics and Physics of Solids, 43(6), pp. 949–972.

Z. Hashin (1988), The differential scheme and its application to cracked materials,J. Mech. Phys. Solids, 36(6), pp. 719–734.

Z. Hashin and S. Shtrikman (1962), A variational approach to the theory ofthe effective magnetic permeability of multiphase materials, J. Appl. Phys., 35.

K. A. Lurie and A. V. Cherkaev (1981), G-closure of a set of anisotropicconducting media in the case of two dimensions, Doklady Akademii Nauk SSSR,259(2), pp. 328–331, In Russian.

K. A. Lurie and A. V. Cherkaev (1982), Exact estimates of conductivityof mixtures composed of two isotropical media taken in prescribed proportion,Technical Report Ref. 783, Physical-Technical Institute, Acad. Sci. USSR, InRussian. Published in English as Lurie and Cherkaev (1984).

K. A. Lurie and A. V. Cherkaev (1984), Exact estimates of conductivityof composites formed by two isotropically conducting media taken in prescribedproportion, Proc. Roy. Soc. Edinburgh Sect. A, 99(1-2), pp. 71–87.

Page 116: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

103

K. A. Lurie and A. V. Cherkaev (1985), Optimization of properties of multi-component isotropic composites, J. Optim. Theory Appl., 46(4), pp. 571–580.

A. Marino and S. Spagnolo (1969), Un tipo di approssimazione dell’operatoren∑

i,j=1

Di(aij(x)Dj) con operatorin∑

j=1

Dj(β(x)Dj), Ann. Scuola Norm. Sup. Pisa

(3), 23, pp. 657–673.

G. W. Milton (1981), Concerning bounds on the transport and mechanicalproperties of multicomponent composite materials, Appl. Phys., A 26, pp.125–130.

G. W. Milton (1986), Modelling the properties of composites by laminates,In Homogenization and effective moduli of materials and media (Minneapolis,Minn., 1984/1985), volume 1 of IMA Vol. Math. Appl., pages 150–174, Springer,New York.

G. W. Milton (2002), The theory of composites, volume 6 of Cambridge Mono-graphs on Applied and Computational Mathematics, Cambridge UniversityPress, Cambridge.

G. W. Milton and R. V. Kohn (1988), Variational bounds on the effectivemoduli of anisotropic composites, J. Mech. Phys. Solids, 36(6), pp. 597–629.

S. Muller (1999), Variational models for microstructure and phase transitions, InCalculus of variations and geometric evolution problems (Cetraro, 1996), volume1713 of Lecture Notes in Math., pages 85–210, Springer, Berlin.

V. Nesi (1995), Bounds on the effective conductivity of two-dimensional compositesmade of n ≥ 3 isotropic phases in prescribed volume fraction: the weightedtranslation method, Proc. Roy. Soc. Edinburgh Sect. A, 125(6), pp. 1219–1239.

V. Nesi and G. W. Milton (1991), Polycrystalline configurations that maximizeelectrical resistivity, J. Mech. Phys. Solids, 39(4), pp. 525–542.

V. Nesi and E. Rogora (2004), A complete characterization of jointly rank-r convex quadratic forms and applications to composite materials, (preprint)http://www.mat.uniroma1.it/∼nesi/ricerca.html.

A. N. Norris (1985), A differential scheme for the effective moduli of composites,Mechanics of Materials: An International Journal, 4, pp. 1–16.

A. Reuss (1929), Berechnung der Fliessgrenze von Mischkristallen auf GrundPlastizitatsbedingung fur Einkristalle, Zeitschrift fur Angewandte Mathematikund Mechanik, 9(1), pp. 49–58.

O. Sigmund (2000), A new class of extremal composites, J. Mech. Phys. Solids,48(2), pp. 397–428.

Page 117: OPTIMALITY OF THE TRANSLATION BOUNDS FOR ...albin/pubs/albin_2006_otb.pdfNathan Lee Albin This dissertation has been read by each member of the following supervisory committee and

104

D. R. S. Talbot, J. R. Willis, and V. Nesi (1995), On improving the Hashin-Shtrikman bounds for the effective properties of three-phase composite media,IMA J. Appl. Math., 54(1), pp. 97–107.

L. Tartar (1979), Estimation de coefficients homogeneises, In Computingmethods in applied sciences and engineering (Proc. Third Internat. Sympos.,Versailles, 1977), I, volume 704 of Lecture Notes in Math., pages 364–373,Springer, Berlin.

L. Tartar (1985), Estimations fines des coefficients homogeneises, In EnnioDe Giorgi colloquium (Paris, 1983), volume 125 of Res. Notes in Math., pages168–187, Pitman, Boston, MA.

S. B. Vigdergauz (1989), Regular structures with extremal elastic properties, J.Mech. Phys. Solids, 24(3), pp. 57–63.

S. B. Vigdergauz (1999), Energy-minimizing inclusions in a planar elasticstructure with macroisotropy, Struct. Opt., 17(2–3), pp. 104–112.

W. Voigt (1928), Lehrbuch der Kristallphysik, Teubner, Stuttgart; Leipzig.