28.08.2013 Views

Optimal Designs - Solid Mechanics

Optimal Designs - Solid Mechanics

Optimal Designs - Solid Mechanics

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Optimal</strong> <strong>Designs</strong><br />

- Structures and Materials - Problems and Tools -<br />

Pauli Pedersen<br />

Department of Mechanical Engineering, <strong>Solid</strong> <strong>Mechanics</strong><br />

Technical University of Denmark<br />

Nils Koppels Allè, Building 404, DK-2800 Kgs.Lyngby, Denmark<br />

email: pauli@mek.dtu.dk<br />

PRE-PRINT<br />

May 28, 2003


2<br />

<strong>Optimal</strong> <strong>Designs</strong><br />

- Structures and Materials - Problems and Tools -<br />

Copyright c○2003 by Pauli Pedersen,<br />

but feel free to copy for personal use.<br />

Colour-pages available from homepage<br />

ISBN 87-90416-06-6


Contents<br />

Contents 3<br />

1 Preface and Introduction 11<br />

1.0.1 Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

1.1 Words of the title . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13<br />

1.1.1 Specific models . . . . . . . . . . . . . . . . . . . . . . . . . . . 13<br />

1.1.2 Specific problems . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

1.2 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

1.3 The chapters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

2 Names and concepts 17<br />

2.1 Major classification of names and concepts . . . . . . . . . . . . . . . . 17<br />

2.2 Design variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.2.1 Alternative classifications . . . . . . . . . . . . . . . . . . . . . 18<br />

2.3 Design objective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

2.4 Design constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20<br />

2.5 Design space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

2.6 Optimization formulations . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

2.7 Optimization methods and procedures . . . . . . . . . . . . . . . . . . 25<br />

2.7.1 The sub-problems of optimal redesign . . . . . . . . . . . . . . 25<br />

2.7.2 The overall strategy . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

3 Design of trusses 31<br />

3.1 Theoretical and numerical background . . . . . . . . . . . . . . . . . . 32<br />

3.1.1 Common parameters for the examples . . . . . . . . . . . . . . 32<br />

3.1.2 Truss members with circular cross-section . . . . . . . . . . . . 33<br />

3.2 Size and topology for optimal plane trusses . . . . . . . . . . . . . . . 33<br />

3.2.1 A cantilever truss with 9 joints . . . . . . . . . . . . . . . . . . 33<br />

3.2.2 A cantilever truss with 17 joints . . . . . . . . . . . . . . . . . 38<br />

3


4 CONTENTS<br />

3.2.3 A bridge truss with 16 joints . . . . . . . . . . . . . . . . . . . 39<br />

3.2.4 A beam/bridge truss with 24 joints . . . . . . . . . . . . . . . . 41<br />

3.3 Size, supports and topology for<br />

optimal single load case space trusses . . . . . . . . . . . . . . . . . . . 44<br />

3.3.1 A tower . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

3.4 Size, shape and topology for optimal 2D-trusses . . . . . . . . . . . . . 44<br />

3.4.1 The cantilever truss with 7 to 12 joints . . . . . . . . . . . . . . 46<br />

3.4.2 The bridge truss with 16 joints . . . . . . . . . . . . . . . . . . 46<br />

3.5 Size and shape for optimal multi-purpose plane trusses . . . . . . . . . 46<br />

3.6 Size and shape for optimal multi-purpose space trusses . . . . . . . . . 50<br />

3.7 A large space truss with constraints on<br />

eigenfrequencies, displacements, stresses and buckling . . . . . . . . . 54<br />

3.8 Approach to topology optimization for<br />

multi-purpose space trusses . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

4 Continua of<br />

uniform energy density 55<br />

4.1 General comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55<br />

4.1.1 Design variables and design constraints . . . . . . . . . . . . . 55<br />

4.1.2 Presentation of results . . . . . . . . . . . . . . . . . . . . . . . 57<br />

4.2 Bars, beams and beam-bars in 2D-formulation . . . . . . . . . . . . . 57<br />

4.3 Bars: long, medium and short . . . . . . . . . . . . . . . . . . . . . . . 59<br />

4.4 Beam-bars: long, medium and short . . . . . . . . . . . . . . . . . . . 62<br />

4.5 Beams: long, medium and short . . . . . . . . . . . . . . . . . . . . . . 65<br />

4.6 ”Bridge”s with different load distributions and supports . . . . . . . . 65<br />

4.7 A ”knee” domain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

4.8 Biaxially loaded hole . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

4.9 A foundation problem . . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

4.10 Uniformly loaded, non-rectangular cantilever . . . . . . . . . . . . . . 75<br />

4.11 The numerical procedure . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

4.12 Summing up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78<br />

4.12.1 Influence from the maximum thickness constraint . . . . . . . . 78<br />

4.12.2 Relation to free material design . . . . . . . . . . . . . . . . . . 78<br />

4.12.3 Relation to non-linear and non-isotropic material . . . . . . . . 79<br />

4.12.4 The specific examples . . . . . . . . . . . . . . . . . . . . . . . 79<br />

5 Design of beams and frames 81<br />

5.1 Explicit analytical optimal designs . . . . . . . . . . . . . . . . . . . . 82<br />

5.1.1 Statically determined elementary cases . . . . . . . . . . . . . . 83<br />

5.2 Implicit ”analytical” optimal designs . . . . . . . . . . . . . . . . . . . 85


CONTENTS 5<br />

5.2.1 Side constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

5.2.2 Statically indetermined cases . . . . . . . . . . . . . . . . . . . 85<br />

5.3 Beam design with eigenfrequency constraints . . . . . . . . . . . . . . 85<br />

5.3.1 Sensitivity analysis for eigenfrequencies . . . . . . . . . . . . . 85<br />

5.3.2 <strong>Optimal</strong> design with one eigenfrequency constraint . . . . . . . 87<br />

5.3.3 <strong>Optimal</strong> design with several eigenfrequency constraints . . . . . 87<br />

5.3.4 <strong>Optimal</strong>ity criterion with only a single constraint . . . . . . . . 87<br />

5.3.5 Sensitivity analysis for Timoshenko beam theory . . . . . . . . 89<br />

5.3.6 Optimization with different gradients . . . . . . . . . . . . . . . 90<br />

5.4 2D-Frame design with many load cases . . . . . . . . . . . . . . . . . . 91<br />

5.4.1 Design problem, analysis and sensitivity analysis . . . . . . . . 92<br />

5.4.2 Only stress constraints and influence from stiffness of columns 93<br />

5.4.3 Including displacement constraints . . . . . . . . . . . . . . . . 98<br />

5.5 3D-Frame design with optimal joint positions . . . . . . . . . . . . . . 98<br />

5.5.1 Design problem, analysis and sensitivity analysis . . . . . . . . 99<br />

5.5.2 <strong>Optimal</strong> design of a dome frame . . . . . . . . . . . . . . . . . 100<br />

5.5.3 <strong>Optimal</strong> design of a mobile crane frame . . . . . . . . . . . . . 103<br />

6 Design of laminates and plates 107<br />

6.1 Laminate analysis and rotational transformations . . . . . . . . . . . . 108<br />

6.1.1 Rotational transformations of stress and strain vectors . . . . . 108<br />

6.1.2 Rotational transformations of constitutive matrices . . . . . . . 109<br />

6.1.3 Alternative description by practical parameters . . . . . . . . . 110<br />

6.1.4 Laminate stiffnesses and lamination parameters . . . . . . . . . 113<br />

6.2 Elastic energy sensitivity analysis . . . . . . . . . . . . . . . . . . . . . 115<br />

6.3 <strong>Optimal</strong> orientation of material in<br />

simply supported plates in bending . . . . . . . . . . . . . . . . . . . . 116<br />

6.4 Laminate point-wise design . . . . . . . . . . . . . . . . . . . . . . . . 120<br />

6.4.1 Thickness distribution only . . . . . . . . . . . . . . . . . . . . 122<br />

6.4.2 Orientations also . . . . . . . . . . . . . . . . . . . . . . . . . . 123<br />

6.5 Strength optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . 124<br />

6.5.1 Global design parameters . . . . . . . . . . . . . . . . . . . . . 125<br />

7 Shapes of minimum stress concentration 129<br />

7.1 Statement of the actual problems . . . . . . . . . . . . . . . . . . . . . 130<br />

7.2 The 2D-fillet in tension . . . . . . . . . . . . . . . . . . . . . . . . . . 131<br />

7.2.1 Circular connection fillet . . . . . . . . . . . . . . . . . . . . . . 132<br />

7.2.2 Super-circular connection fillet . . . . . . . . . . . . . . . . . . 135<br />

7.2.3 Extended length of the fillet . . . . . . . . . . . . . . . . . . . . 135<br />

7.2.4 Multi-parameter optimal shape design . . . . . . . . . . . . . . 135


6 CONTENTS<br />

7.3 The 3D-fillet in tension, bending and torsion . . . . . . . . . . . . . . 138<br />

7.4 The 2D-hole subjected to biaxial stress . . . . . . . . . . . . . . . . . . 144<br />

7.4.1 The classic solution for stress concentration around small elliptical<br />

holes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144<br />

7.4.2 Numerical solutions for finite domains . . . . . . . . . . . . . . 146<br />

7.5 The 2D-hole in an orthotropic material . . . . . . . . . . . . . . . . . . 148<br />

7.6 The 2D-hole in a non-linear elastic material . . . . . . . . . . . . . . . 148<br />

7.7 The 3D-cavity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151<br />

7.8 Shape design in materials for maximum bulk modulus . . . . . . . . . 151<br />

7.9 Stress release at a crack tip . . . . . . . . . . . . . . . . . . . . . . . . 151<br />

7.9.1 Influence from the external load and size of the hole . . . . . . 154<br />

7.9.2 Influence from material non-isotropy . . . . . . . . . . . . . . . 155<br />

7.9.3 Influence from material power law non-linearity . . . . . . . . . 155<br />

7.9.4 Influence from the allowable domain of the hole . . . . . . . . . 157<br />

8 Material design 159<br />

8.1 Free material design . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160<br />

8.2 Design of density distribution . . . . . . . . . . . . . . . . . . . . . . . 161<br />

8.2.1 ”Black and white” designs . . . . . . . . . . . . . . . . . . . . . 162<br />

8.2.2 An illustrative example . . . . . . . . . . . . . . . . . . . . . . 163<br />

8.3 Design of a two parameter cell model . . . . . . . . . . . . . . . . . . . 170<br />

8.4 Design of material with a single hole . . . . . . . . . . . . . . . . . . . 172<br />

8.4.1 Influence from the volume constraint . . . . . . . . . . . . . . . 173<br />

8.4.2 Influence from boundary conditions . . . . . . . . . . . . . . . 174<br />

8.4.3 Isotropic cases with influence from Poisson’s ratio . . . . . . . 178<br />

8.4.4 Influence from non-linear elasticity . . . . . . . . . . . . . . . . 180<br />

8.5 Design with multi-physics constraints . . . . . . . . . . . . . . . . . . . 184<br />

9 Bone mechanics and damage evolution 185<br />

9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185<br />

9.1.1 Points of view . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185<br />

9.1.2 Contents of the chapter . . . . . . . . . . . . . . . . . . . . . . 187<br />

9.2 Bone layouts from an optimization process . . . . . . . . . . . . . . . . 187<br />

9.3 Finite element analysis and load modelling . . . . . . . . . . . . . . . . 188<br />

9.4 Homogenization and inverse homogenization . . . . . . . . . . . . . . . 190<br />

9.5 Bone layout based on<br />

one parameter optimization . . . . . . . . . . . . . . . . . . . . . . . . 191<br />

9.6 Multi-parameter evolution . . . . . . . . . . . . . . . . . . . . . . . . . 192<br />

9.7 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196


CONTENTS 7<br />

10 Identification (estimation) and inverse problems 197<br />

10.1 Estimation of orthotropic material parameters . . . . . . . . . . . . . 197<br />

10.1.1 Identification formulation . . . . . . . . . . . . . . . . . . . . . 199<br />

10.1.2 The experimental setup . . . . . . . . . . . . . . . . . . . . . . 200<br />

10.1.3 Plate eigenfrequency analysis and sensitivity analysis . . . . . . 202<br />

10.1.4 Example results . . . . . . . . . . . . . . . . . . . . . . . . . . 203<br />

10.1.5 Uncertainties and optimal experiments . . . . . . . . . . . . . . 204<br />

10.1.6 Comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206<br />

10.2 Laminate inverse problems . . . . . . . . . . . . . . . . . . . . . . . . . 207<br />

10.2.1 Laminate with given bending stiffness . . . . . . . . . . . . . . 207<br />

10.2.2 A redesign procedure . . . . . . . . . . . . . . . . . . . . . . . . 208<br />

10.3 Material identification and inverse homogenization . . . . . . . . . . . 209<br />

10.3.1 Homogenization . . . . . . . . . . . . . . . . . . . . . . . . . . 209<br />

10.3.2 Inverse homogenization . . . . . . . . . . . . . . . . . . . . . . 209<br />

11 Alternative Descriptions of Constitutive Parameter 211<br />

11.1 Stress/strain relations . . . . . . . . . . . . . . . . . . . . . . . . . . . 211<br />

11.2 Two-dimensional formulations . . . . . . . . . . . . . . . . . . . . . . . 211<br />

11.3 Three-dimensional formulations . . . . . . . . . . . . . . . . . . . . . . 214<br />

11.3.1 Spectral decomposition . . . . . . . . . . . . . . . . . . . . . . 214<br />

11.3.2 Constitutive parameters by energy densities . . . . . . . . . . . 214<br />

11.4 Bulk modulus and alternative isotropic parameters . . . . . . . . . . . 216<br />

11.4.1 Three-dimensional case . . . . . . . . . . . . . . . . . . . . . . 216<br />

11.4.2 Two-dimensional case . . . . . . . . . . . . . . . . . . . . . . . 216<br />

11.4.3 Relations between other alternative moduli . . . . . . . . . . . 217<br />

11.5 Summing up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218<br />

12 Effective stress/strain and energy densities 219<br />

12.1 Analysis by secant formulation . . . . . . . . . . . . . . . . . . . . . . 219<br />

12.1.1 Non-dimensional compliance matrix . . . . . . . . . . . . . . . 220<br />

12.1.2 The von Mises effective stress . . . . . . . . . . . . . . . . . . . 221<br />

12.1.3 The Hill strength measure . . . . . . . . . . . . . . . . . . . . . 222<br />

12.2 Strain energy density and<br />

stress energy density . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223<br />

12.3 Summing up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224<br />

13 Elastic potentials, relations and sensitivities 225<br />

13.1 Elastic potentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225<br />

13.2 Derivatives of elastic potentials . . . . . . . . . . . . . . . . . . . . . . 226<br />

13.2.1 Constant constitutive matrix, a specific case . . . . . . . . . . . 227


8 CONTENTS<br />

13.2.2 Constant volume, a specific case . . . . . . . . . . . . . . . . . 227<br />

13.3 Summing up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228<br />

14 Some necessary conditions for optimality 229<br />

14.1 Non-constrained problems . . . . . . . . . . . . . . . . . . . . . . . . . 229<br />

14.2 Problems with a single constraint . . . . . . . . . . . . . . . . . . . . . 230<br />

14.3 Size optimization for stiffness and strength . . . . . . . . . . . . . . . . 231<br />

14.3.1 Size design with optimal stiffness . . . . . . . . . . . . . . . . . 231<br />

14.3.2 Size design with optimal strength . . . . . . . . . . . . . . . . . 232<br />

14.4 Shape optimization for stiffness and strength . . . . . . . . . . . . . . 232<br />

14.4.1 Shape design with optimal stiffness . . . . . . . . . . . . . . . . 232<br />

14.4.2 Shape design with optimal strength . . . . . . . . . . . . . . . 234<br />

14.5 Conditions with a<br />

simple shape parametrization . . . . . . . . . . . . . . . . . . . . . . . 235<br />

14.5.1 Possible iterative procedure . . . . . . . . . . . . . . . . . . . . 239<br />

14.5.2 Extended design space . . . . . . . . . . . . . . . . . . . . . . . 239<br />

14.6 Summing up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240<br />

15 The ultimate optimal material 241<br />

15.1 The individual constitutive parameters . . . . . . . . . . . . . . . . . . 241<br />

15.2 Sensitivity analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241<br />

15.3 Final optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242<br />

15.4 Numerical aspects and comparison with isotropic material . . . . . . . 243<br />

15.5 Summing up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244<br />

16 Conditions for statically determinate trusses 245<br />

16.1 Theorem for the single load case . . . . . . . . . . . . . . . . . . . . . 246<br />

16.2 Proof with supports as further unknowns . . . . . . . . . . . . . . . . 246<br />

16.2.1 Force equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . 246<br />

16.2.2 The total cost with all forces . . . . . . . . . . . . . . . . . . . 247<br />

16.2.3 The change in total cost . . . . . . . . . . . . . . . . . . . . . . 248<br />

16.2.4 Monotonous behaviour and new basis solution . . . . . . . . . 248<br />

16.3 <strong>Optimal</strong>ity condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249<br />

16.4 Alternative proofs and solution procedures . . . . . . . . . . . . . . . . 249<br />

16.5 A truss member model . . . . . . . . . . . . . . . . . . . . . . . . . . . 249<br />

16.5.1 Tensile bars, i.e. P>0 . . . . . . . . . . . . . . . . . . . . . . 250<br />

16.5.2 Slender columns, i.e. PL


CONTENTS 9<br />

16.5.6 Cost functions for the supports . . . . . . . . . . . . . . . . . . 252<br />

16.6 Summing up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252<br />

17 Orientation of orthotropic material 253<br />

17.1 Inherent practical problems . . . . . . . . . . . . . . . . . . . . . . . . 253<br />

17.2 From global to local optimality criterion . . . . . . . . . . . . . . . . . 254<br />

17.3 Multiplicity of extremum . . . . . . . . . . . . . . . . . . . . . . . . . 255<br />

17.4 The match problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257<br />

17.5 Numerical iterations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259<br />

17.6 Summing up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260<br />

18 Sensitivity analysis for dynamic problems 261<br />

18.1 Including non-conservative problems . . . . . . . . . . . . . . . . . . . 261<br />

18.2 Classification of dynamic systems and their behaviour . . . . . . . . . 262<br />

18.2.1 System classification . . . . . . . . . . . . . . . . . . . . . . . . 262<br />

18.2.2 Classification of the behaviour . . . . . . . . . . . . . . . . . . 263<br />

18.3 Sensitivity analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264<br />

18.3.1 General variational analysis . . . . . . . . . . . . . . . . . . . . 264<br />

18.3.2 Results with finite element formulation . . . . . . . . . . . . . 266<br />

18.3.3 Results with Galerkin modelling (global expansion) . . . . . . . 267<br />

18.3.4 Example of a cantilever beam . . . . . . . . . . . . . . . . . . . 268<br />

18.3.5 Multiple eigenvalues . . . . . . . . . . . . . . . . . . . . . . . . 270<br />

18.3.6 Double eigenvalue with only a single eigenvector . . . . . . . . 271<br />

18.4 Summing up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271<br />

19 Simplex and a modified simplex procedure 273<br />

19.1 Standard linear programming form . . . . . . . . . . . . . . . . . . . . 274<br />

19.2 Effect of changing the basis set . . . . . . . . . . . . . . . . . . . . . . 274<br />

19.3 Non-linear objective . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275<br />

19.4 Positive and negative variables . . . . . . . . . . . . . . . . . . . . . . 276<br />

19.5 Summing up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 278<br />

20 Linear programming reformulations with SLP 279<br />

20.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279<br />

20.2 Actual optimization problem after sign reformulations . . . . . . . . . 280<br />

20.3 From non-linear to sequential linear problem . . . . . . . . . . . . . . 280<br />

20.4 From inequalities to equalities . . . . . . . . . . . . . . . . . . . . . . . 281<br />

20.5 Obtaining non-negative variables . . . . . . . . . . . . . . . . . . . . . 281<br />

20.6 Move-limits and non-negative variables . . . . . . . . . . . . . . . . . . 282<br />

20.7 Problem constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282<br />

20.8 Total LP coefficient matrix for the constraints . . . . . . . . . . . . . . 284


10 CONTENTS<br />

20.9 The vector of cost coefficients . . . . . . . . . . . . . . . . . . . . . . . 284<br />

20.10 Sequential linear programming . . . . . . . . . . . . . . . . . . . . . . 285<br />

20.11 Summing up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285<br />

Bibliography 287<br />

Index 297


Chapter 1<br />

Preface and Introduction<br />

1.0.1 Preface<br />

This book grew from a Ph.D.course given at Politecnico di Milano in July 1999. For<br />

more than 30 years I have been involved with optimal designs and identifications, but<br />

this was the first chance to put together such a course. From July 1999 until finishing<br />

the book many new aspects have been examined and are included.<br />

My personal line through these problems started out with beam design in 1967,<br />

followed by a Ph.D. study concentrated on truss design. From 1970 the importance<br />

of combinations with the finite element method became clear with focus on design for<br />

minimum stress concentration, first 2D-problems and later 3D-problems. In fact, the<br />

interest in the finite element method itself took over.<br />

Later, i.e., from 1980, vibrational problems were my primary area of research,<br />

with focus on non-conservative dynamic stability. Related sensitivity analysis was<br />

(and still is) a fascinating subject. Since 1985 the design of composite materials<br />

like laminates was given a lot of attention. The step to material design is a natural<br />

extension with inverse homogenization and bone mechanics now in focus.<br />

It is a challenge trying to put all this together, so that the main results can be<br />

communicated to graduate students and engineers in practice. This is a first layout<br />

that will be freely available on the Internet, thus making colour presentations and<br />

even animations possible (later). Critics and suggestions for improvements are most<br />

welcome, e.g. by email to pauli@mek.dtu.dk Email<br />

pauli@mek.dtu.dk<br />

11


12 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ....<br />

Thanks to students and colleagues without whom the job was not done. For<br />

critical comments on early manuscripts I thank Alexander Seyranian from Moscow<br />

State University, Russia; John E. Taylor from University of Michigan, Ann Arbor,<br />

USA; Ciro A. Soto from Ford research laboratory in Dearborn, USA; Alija Picuga from<br />

Mostar, Bosnia; and my young colleagues at Aalborg University Niels L. Pedersen and<br />

Jan Stegmann.<br />

Lyngby, January 2003<br />

Pauli Pedersen


Preface and Introduction 13<br />

1.1 Words of the title<br />

<strong>Optimal</strong><br />

designs<br />

This book is about optimal designs. Optimization and optimization methods are<br />

necessary tools but are not put in focus. There are a number of good books that<br />

cover methods to a large extent. From the optimal design point of view we mention:<br />

(Banichuk 1983), (Save, Prager and Sacchi 1985), (Arora 1989), (Brandt 1989),<br />

(Rozvany 1989), (Borkowski, Jendo and Reitman 1990), (Haftka, Gurdal and Kamat<br />

1990), (Kirch 1993), (Bendsøe 1995), and (Bendsøe and Sigmund 2003). In the<br />

present book we focus on the resulting designs and the possible general knowledge<br />

connected to these. Identification<br />

The subject of identification, or in more weak form estimation, is also of major<br />

importance, but to make the title short it is not included in the title. Identification is Inverse<br />

a subset of inverse problems, that from the author’s point of view are of even greater problems<br />

practical importance than the optimal design problems. Traditionally, engineers analyse<br />

a given structure (or material) to determine a response. In inverse problem we<br />

seek a structure that gives a wanted response and in identification this response may<br />

be obtained experimentally. Stated in other words, we seek a model that matches<br />

measured values. Theory and tools<br />

Theory and tools (procedures) are included as the last 10 chapters and may serve<br />

as a kind of appendices. The hope is that the reader, motivated by the many examples<br />

in the first chapters, is interested in the more general aspects of optimal design.<br />

1.1.1 Specific models<br />

In the second line of the title, structures and materials are mentioned in parallel.<br />

Most research on optimal design is related to structures or we may be more specific<br />

and say structural models such as bars, rods, trusses, beams, frames, grids, plates,<br />

Structural<br />

models<br />

shells, axisymmetric models or in general 2D- and 3D-continuum models. In recent Material<br />

years a major focus has been put on design of materials and again we should say on<br />

models<br />

material models. Points or<br />

domains<br />

There is not a clear distinction between structure and material. Is a laminate<br />

a structure or a material? In reality even a traditional material is a structure if we<br />

look close enough. The distinction could be if we model what is related to a point (a<br />

material point) or we model what is related to a domain (a structural domain). Inverse<br />

Methods of homogenization transfer modelling from domain to point. Modelling<br />

a material point as a finite domain (a cell model) we have a structural/continuum<br />

problem, and the material design is then related to structural/continuum design.<br />

Thus we are directly lead to what is named inverse homogenization.<br />

homogenization


14 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ....<br />

1.1.2 Specific problems<br />

The further words of the title, - problems and tools -, again stress the fact that formulation<br />

and solution of specific problems are put in focus. We want to learn by solving,<br />

hopefully well-defined problems, and from this extrapolate our possibilities. This<br />

General<br />

means that the problems are not just examples to illustrate a theory or a procedure.<br />

knowledge However, the specific problems often teach us about or indicate that general<br />

knowledge can be obtained. Then the focus is naturally turned in that direction,<br />

but the more theoretical aspects and proofs are placed in chapters that more serve as<br />

appendices, say the theory behind statically determined optimal trusses, the theory<br />

behind uniformly stressed continuum domains, and the theory related to shapes with<br />

uniform energy density.<br />

1.2 References<br />

Not proper<br />

references The book is intended to be self-containing, although this naturally depends on the<br />

readers background. A list of references is given, but not with proper references to<br />

the large international literature. The book is mainly based on ”in house” research,<br />

often carried out in cooperation with students and colleagues, also outside Denmark.<br />

A number of extensive review papers are available, for early reviews see (Wasiutynski<br />

and Brandt 1963), (Sheu and Prager 1968), (Niordson and Pedersen 1973-74),<br />

(Olhoff 1980), (Olhoff and Taylor 1983), and for a more recent review see (Eschenauer<br />

and Olhoff 2001). Also the books mentioned in the beginning of this introduction have<br />

valuable lists of references. A book with all proper references would be difficult to<br />

write, and hopefully, the reader will agree and appreciate the present style.<br />

Glossary<br />

Trusses<br />

Uniform<br />

energy<br />

density<br />

1.3 The chapters<br />

Like most areas of research, optimal design has its own names and concepts. Chapter<br />

2 may serve as a glossary without being ordered. Then already in chapter 3 we go to<br />

a specific problem, statically determined trusses.<br />

Besides being the most simple structural model it is also the most effective structure,<br />

at least when it is up to us to define the word effective. A number of rather<br />

different aspects are covered including the account for local stability. The proof for<br />

the statically determined, one load case is given in chapter 16.<br />

In chapter 4 continua of uniform energy density are designed, restricted to 2Dproblems.<br />

The theoretical knowledge of uniform elastic energy density is applied in a<br />

simple recursive procedure that iteratively gives an optimal continuum, optimal with<br />

respect to stiffness as well as to strength. In chapters 12, 13, 14, and 15 the theory


Preface and Introduction 15<br />

behind these examples shows the extension to non-isotropic elasticity, to power law<br />

non-linear elasticity, and to free material design. Energy<br />

Energy principles of mechanics play a major role in the derivations, but even principles<br />

without knowing the details of these derivations, the resulting conclusions can be<br />

used, say in a standard finite element program which then directly extends from<br />

analysis to synthesis. Beams<br />

From continuum we go back to the one-dimensional models of static beam problems<br />

in chapter 5, where the knowledge from energy principles is used to obtain some<br />

analytical solutions for optimal design. More extended problems, like frames (made<br />

of beams) must be solved numerically. On the other hand, the extension to dynamic<br />

problem for the frames is not a big step, assuming that results from eigenvalue analysis<br />

and eigenvalue sensitivity analysis are known. In agreement with the general Vibrations<br />

idea of the book, this theory is described later in chapter 18, so chapter 5 can directly<br />

discuss the solution to some specific problems.<br />

and frames


16 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...


Chapter 2<br />

Names and concepts<br />

2.1 Major classification of names and concepts<br />

Like most areas of research, optimal design has its own names and concepts. This<br />

chapter may serve as a glossary without being ordered. From this chapter we go<br />

directly to specific problems. The major classes of concepts are: Glossary<br />

• Design VARIABLES<br />

• Design OBJECTIVE<br />

• Design CONSTRAINTS and Design SPACE<br />

<strong>Optimal</strong> design problems may have one solution, several solutions or no solution.<br />

Even when there is only one solution, the optimal design problem may be presented<br />

in many different formulations. In addition a given formulation may be solved with<br />

many different methods and procedures. Thus we add as major classes of concepts:<br />

• Optimization FORMULATIONS<br />

• Optimization METHODS and PROCEDURES<br />

2.2 Design variables<br />

The most important classification of design variables is into:<br />

• SIZE design variables<br />

• SHAPE design variables<br />

• TOPOLOGY design variables<br />

17


Size<br />

Shape<br />

Topology<br />

Examples<br />

18 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

stated here in order of difficulty to solve but also in order of increasing importance<br />

for the obtained objective value. It is therefore not surprising that recent research to<br />

some extent concentrates on topology design variables.<br />

The notion of size design variable, relates to the thickness of a beam, a plate or<br />

a shell (although this is often termed the shape of a beam, a plate or a shell). The<br />

area of a bar in a truss is also a size design variable, and the definition of size variable<br />

is related to the fact that the modelling domain is not changed. So, the line of the<br />

beam, rod or bar is unchanged, just like the reference surface of a plate or a shell is<br />

assumed unchanged when the concept of size design variable is used. In 3D-problems<br />

the mass density or a relative volume density is size. The orientations of non-isotropic<br />

material we also treat as size design variables.<br />

The notion of shape design variable, relates to the reference domain of the actual<br />

model. For beams, rods and bars we may treat the length as a design variable,<br />

which is then a shape design variable. Also the curvature of the reference line for<br />

these one-dimensional models is a shape design variable. For 2D-models likewise the<br />

boundary curve or the curvature of the reference surface are shape design variables.<br />

For 3D-models the boundary surface (including internal boundaries like holes) is a<br />

shape design variable. Stress concentration problems are often related to shapes of<br />

boundaries.<br />

Finally, the notion of topology design variable, relates to presence or absence of<br />

a certain design aspect. Should two joints in a truss be connected with a bar, - yes or<br />

no ?. Should a continuum like a plate have a hole, - yes or no ?. The complications<br />

in treating topology design variables are due to the fact that a change in topology<br />

results in a discontinuous change in the design response, while a continuous change<br />

in size or shape design variables normally results in continuous change in the design<br />

response.<br />

Let us exemplify the difference between size, shape and topology design variables.<br />

In a truss (2D as well as 3D), the bar areas (uniform or non-uniform) are the sizes,<br />

the positions of the joints determine the shape, and the chosen bars (among many<br />

possibilities) give the topology. In a shell the thickness and material density distributions<br />

are the sizes, the boundaries of the reference surface and its curvatures are the<br />

shapes, and the number of holes in the reference surface is the topology.<br />

2.2.1 Alternative classifications<br />

Many alternative names to classify design variables can be found in the literature, like<br />

cross-sectional, geometrical, configuration, layout etc. We try to avoid these names<br />

Other<br />

in order to avoid unnecessary confusion.<br />

classifications The design variables may also be classified from other points of view. Let us first<br />

discuss the distinction between continuous and discrete design variables. If only a


Names and concepts 19<br />

number of specific values for the design variable is acceptable, say when catalog values<br />

must be used, then the notion of discrete design variables is used, and procedures Continuous<br />

related to what is called Integer programming come into focus. This is not covered discrete<br />

in the present book that concentrates on continuous descriptions, but we include<br />

absolute limitations for the values of the design variables. Distributed<br />

Another meaning of the ”continuous” and ”discrete” relates to the modelling parameter<br />

of the design domain. A complete continuous description in space means design<br />

variables related to a point (like a design function) and not to a domain. Often this<br />

is termed distributed parameter description, in contrast to say a truss description<br />

where each bar is described as a unit. In a finite element modelling of a continuum,<br />

the element domains may be related to a number of design values, so in reality this<br />

is a discrete description. However, with the extensive number of elements and the<br />

fact that everything in a computer is discrete, the distinction between continuous and<br />

discrete related to the modelling of the design domain is of no practical importance. Parametrization<br />

For a successful optimization the choice of design parametrization is of vital importance,<br />

perhaps the most important decision to take. In the experience of the<br />

author it is wise to start with as few design variables as possible. A hierarchical description<br />

is suggested, and also it is important to make sure that the design variables<br />

serve different purposes. It is asking for practical problems, if the design variables are<br />

chosen such that different combinations of design variables can give the same design.<br />

The parametrization is also related to the chosen optimization procedure, so with<br />

an optimality criterion method large quantities, say 50.000 design variables, can be<br />

handled without problems.<br />

2.3 Design objective<br />

The design objective is a function or a functional that returns a single value from<br />

which different designs can be compared. The optimal design is then the design with<br />

a minimum (or maximum) value of the objective. In this book we often use the<br />

notation Φ to denote the objective. We shall not treat multi-objective formulations,<br />

which in most cases are reformulated into a single objective anyhow.<br />

Alternative names for the objective include criterion, cost, merit, goal as well as<br />

many others. The name ”criterion” is in this book used extensively in relation to<br />

optimality criterion formulations (see chapter 14), so we try only to use the name<br />

objective, although a name like cost may be more appealing. In fact, the objective<br />

Single value<br />

value is often a measure of the cost of the design. Minimum<br />

A minimum and maximum formulation may be interchanged by simply changing<br />

the sign of the objective. However, it is important to notice that many methods just<br />

locate a stationary value of the objective, which means that the convergence of the<br />

Maximum


Local<br />

Global<br />

Optimized<br />

design<br />

Existence<br />

Convergence<br />

tests<br />

Constraints<br />

20 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

procedure must be followed and the final design justified.<br />

A much more severe problem is related to the existence of local stationary solutions,<br />

and in reality very few (and often non-practical) methods are able to find a<br />

global optimal solution. Starting an optimal design procedure from different initial<br />

designs and always ending up in the same optimal design may be the most practical<br />

procedure for improving the probability of an obtained solution being the global optimal<br />

solution. It should be noticed that a number of optimal design formulations for<br />

idealized problems may include a proof of global optimal solution.<br />

However, for problems where a large number of practical constraints need to be<br />

taken into account, it is more safe to state that we have optimized the design as an<br />

alternative to obtaining the optimal design. Furthermore, it is not always easy to see<br />

from the formulation whether an optimal design exists. If an optimal design does not<br />

exist we talk about a not well formulated problem. Even so a procedure may return<br />

an optimized design, and the convergence often reveals the missing aspect(s) in the<br />

formulation.<br />

An important part of an optimization procedure is to decide when to stop. We<br />

talk about convergence tests. Two different aspects of convergence must be clarified,<br />

convergence of the design objective and convergence of the design variables. Often<br />

the rates of these two convergences are very different. Also the formulation of the<br />

specific stop condition can be mathematically formulated more or less complicated.<br />

The favourite formulation of the present author is as follows: When the design changes<br />

are somewhat smaller than the actual accuracy in the design production, then the<br />

design procedure should be stopped. At that instant the design objective is often<br />

converged at a much earlier step.<br />

2.4 Design constraints<br />

The design constraints serve many purposes and more or less reflect the goal of the<br />

design. In order to get some overview let us discuss them according to the following<br />

names:<br />

• LOAD conditions and equilibrium<br />

• BEHAVIOURAL constraints<br />

• SIDE constraints, absolute limits and move-limits<br />

• EQUALITY and INEQUALITY constraints<br />

• FULLY STRESSED design<br />

• SATISFIED and VIOLATED constraints


Names and concepts 21<br />

Often the loads on a structure are given and are design independent. However,<br />

loads like inertia forces and pressure are in most cases design dependent. Boundary<br />

conditions may also be given or may be part of the design variables. In the most<br />

simple cases as for the truss design in chapter 3, we in reality solve a problem of<br />

Loads and<br />

boundary<br />

conditions<br />

transfer of given loads to given possible supports. Load cases<br />

The main condition of all the design problems is that equilibrium must be satisfied<br />

for static as well as for dynamic problems. In most cases the analysis to obtain<br />

equilibrium is solved by the finite element method, a method that we assume the<br />

reader to be familiar with. Note that in most practical problems, we need to take a<br />

number of different load cases into account. Implicit<br />

A structure is subjected to a large number of behavioural constraints, that more or<br />

less are all calculated based on the results from the equilibrium. All these constraints<br />

depend in a very implicit manner on the design variables, and must be calculated for<br />

behaviour<br />

all load cases. Displacements<br />

and stiffness<br />

The stiffness or flexibility can be described locally by displacements or globally by<br />

total elastic energy. The displacements are restricted in magnitude, but very specific<br />

prescribed displacements may also be part of the design purpose. A large amount of Compliance<br />

optimal designs are based on minimum compliance, which is the same as minimum<br />

elastic energy (as defined more specifically in chapter 13). Stresses<br />

strength<br />

From displacements follow strains and then with a constitutive model the stresses.<br />

Constraints on stresses are related to the strength, as expressed by the von Mises stress<br />

or the strain/stress energy density. To account for these strength constraints we must<br />

find the most critical points in a structure, i.e. these constraints are formulated as<br />

point (element) constraints, where the dangerous domains must first be located (active<br />

set strategy). Eigenfrequencies<br />

For dynamic problems, the eigenvalue analysis needed to determine eigenfrequencies<br />

can be a finite element analysis, that returns sets of frequencies and corresponding<br />

vibration modes. The constraint can specify a minimum allowable frequency and/or<br />

a minimum gap between eigenfrequencies. The sensitivity analysis in chapter 18 gives<br />

the extended background for the iterative redesign of such design problems, and ex-<br />

tended experience is available. Stability<br />

Another eigenvalue problem is related to the analysis which determines the stabil- (buckling)<br />

ity of the structure against buckling. To a large extent these aspects of optimal design<br />

are similar to dealing with eigenfrequencies. More indirect behavioural constraints Fracture, creep<br />

are related to fracture, fatigue, creep and in reality to aspects that determine the life<br />

time of the structure. Absolute<br />

The notion of side constraints covers two different aspects. The first one we term<br />

absolute side limits and gives the designer the possibility of specifying a range of<br />

acceptable values for each design variable. The second aspect is directly related to<br />

side limits


Move-limits<br />

Active<br />

constraints<br />

or inactive ?<br />

Fully stressed<br />

Satisfied<br />

or violated ?<br />

22 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

the control of the iterative design process. The total optimal design is carried out<br />

in a number of steps, or we may say redesigns, i.e. the optimization is an iterative<br />

process.<br />

In each redesign it is often a good idea to describe so called move-limits, that<br />

control the change from one design to the next iterative design. By adjusting in<br />

an actual redesign the move-limits we can indirectly take the absolute limits into<br />

account, so that these are not involved in the formulation. When move-limits are<br />

given in terms of physical quantities the designer treats these as probably the most<br />

important parameters to control the design process.<br />

Constraints specified as equality constraints are conditions that must be satisfied,<br />

say the equilibrium constraints. Most behavioural constraints are specified as inequality<br />

constraints, or we may say bounds on the behaviour. Likewise the absolute limits<br />

and the move-limits, and these bound constraints may be active or inactive, and in<br />

general it is not known in advance what will be the case. Only when the optimal<br />

design is obtained, the final information about the active constraints is known.<br />

A fully stressed design reflects in reality a natural consequence, more than a<br />

described constraint. We will see that a number of idealized design problems lead to<br />

fully stressed designs, and if this can be assumed in advance, the design problem is<br />

much less complicated. Even with multiple load cases, the notion of fully stressed<br />

is meaningful; it describes that each structural member (or structural region) is fully<br />

stressed at least in relation to one of the load cases.<br />

The design problem is a very non-linear problem, but dividing the problem into<br />

an iterative process of optimal redesigns we linearize the description relative to the<br />

actual design, like a Taylor expansion of the behaviour. This means that errors in<br />

the analysis are involved, and we may have constraints that are violated. During the<br />

design process we have both satisfied and violated constraints. Ending the iterative<br />

process with small move-limits we can control these violations.


Names and concepts 23<br />

2.5 Design space<br />

The notion of design space gives a geometrical interpretation of our possibilities for<br />

an actual design. Dealing with many design parameters this design space is multidimensional,<br />

but anyhow, for most of us is an advantageously interpretation. By the<br />

following concepts we describe this design space in more detail:<br />

• FEASIBLE and NON-FEASIBLE design space<br />

• INTERIOR and EXTERIOR design space<br />

• CONSTRAINT BOUNDARIES<br />

• CONVEX design space<br />

• DIRECTION in design space<br />

• GRADIENTS for a design direction<br />

• STEP SIZE and LINE-SEARCH<br />

• JUMPS in design space<br />

Feasibility<br />

A number of these names are more or less self-explaining. The feasible design<br />

space is the collection of all design points (sets of design parameters) that satisfy all<br />

the design constraints. In the non-feasible design space at least one design constraint<br />

is violated. In the interior of the design space all neighbouring points are also feasible<br />

design points, while in the exterior design space all neighbouring points are also nonfeasible<br />

points. The separation between these two design spaces is the constraint<br />

boundary, where at least one constraint is active. Convex<br />

Changing the parametrization of the design description we change the design<br />

space, and if it is possible to enforce a parametrization where the design space is<br />

convex, then valuable conclusions can be drawn. The notion of convex means that<br />

any linear combination between two feasible design points is a feasible design point. Moving in<br />

design space<br />

A design change involves a movement in the design space. This can be specified<br />

by a direction and a step size. The change may be continuous or it may be a jump.<br />

If several step sizes are possible, then a search along a proposed direction may be<br />

actual.<br />

An advantageous move in the design space, corresponding to a redesign, is determined<br />

by the actual gradients corresponding to the different directions in the design<br />

space. The gradients relate to the objective as well as to all the constraints, so a lot of<br />

information is necessary and in most cases this information is only available in an im- Sensitivity<br />

plicit manner. Important results from sensitivity analysis may give this information,<br />

see chapters 14 and 18.<br />

analysis


24 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

2.6 Optimization formulations<br />

Choosing a formulation for an optimal design problem, we more or less restrict ourselves<br />

and we do so in order to obtain specific results. Modifications and extensions<br />

of a chosen formulation are in most cases necessary in the process of learning more<br />

about our specific problem. A number of names and concepts related to different<br />

formulations are:<br />

• DETERMINISTIC or PROBABILITY based optimal designs<br />

• ANALYTICAL or NUMERICAL formulation<br />

• LINEAR or NON-LINEAR behavioural constraints involved<br />

• STATIC or DYNAMIC constraints involved<br />

• NECESSARY and SUFFICIENT conditions for optimality<br />

• AUTOMATIC design, INVERSE problems and IDENTIFICATION<br />

• SIMULTANEOUSLY ANALYSIS and DESIGN<br />

Only<br />

deterministic A few comments to this list is given. In the present book we only deal with<br />

optimal designs based on deterministic behaviour.<br />

Mostly<br />

Only few analytically determined optimal designs are available, but because of<br />

numerical<br />

the insight they give, we describe some of these. However, the major part of the book<br />

focus on numerically obtained optimal designs. This does not imply that analytical<br />

results are not involved, because the basic theory behind a given numerical procedure<br />

Also<br />

is often based on analytically obtained knowledge, i.e., theorems and others.<br />

non-linear The importance of non-linearity in behavioural description along with our ability<br />

to day to analyse with the finite element method mean that to optimal designs, based<br />

on non-linear behaviour must be given proper attention. We strive to obtain this<br />

Multiple<br />

goal.<br />

eigenvalues Static and dynamic constraints are treated, and eigenvalue problems are put in<br />

focus. The specific problem with multiple eigenvalues forces us to go in more details<br />

Stationarity with these problems, that physically relate to both vibration and stability problems.<br />

extremum The energy theorems of mechanics are divided into stationarity principles and<br />

extremum principles. Likewise are the optimality conditions, and furthermore we<br />

need to distinguish between necessary and sufficient conditions for optimality. The<br />

close relation between the principles of mechanics and the principles of optimality is<br />

shown.<br />

Inverse<br />

In practice the restricted problem of finding just a feasible design may be of<br />

problems<br />

major importance. In the subject of optimal design this sub-problem is often termed<br />

automatic design. Also the very important problems related to formulation of inverse


Names and concepts 25<br />

problems and to problems of identification (estimation) are shown to involve the same<br />

tools and procedures as necessary for optimal design.<br />

2.7 Optimization methods and procedures<br />

Most optimal designs can only be obtained by an iterative solution of a sequence of<br />

sub-problems as shown in figure 2.1. Each of these sub-problems may be solved by<br />

a number of different methods and, furthermore, the selection of design parameters<br />

is by no means unique. Viewing a specific combination of methods of solution for<br />

these sub-problems as a ”new” optimal design method, we get too many methods<br />

and lose our overview of the subject. Therefore it is necessary to concentrate on the<br />

sub-problems, separately.<br />

From the author’s point of view, the splitting of our approaches into analytical<br />

methods versus numerical methods, or our subject into distributed parameter problems<br />

versus discrete parameter problems is also confusing. This splitting is more a<br />

matter of taste than necessity since almost all optimal design problems are solved<br />

numerically with a finite number of parameters.<br />

2.7.1 The sub-problems of optimal redesign<br />

Sub-problems<br />

The box of optimal redesign in figure 2.1 includes three sub-problems: <strong>Optimal</strong><br />

redesign<br />

• ANALYSIS by finite element method (FEM)<br />

• SENSITIVITY ANALYSIS<br />

• REDESIGN<br />

The term response analysis is a very wide one, which may include displacement-<br />

, stress-, vibration- and stability-analysis. It is a characteristic of optimal design<br />

that we are confronted with non-uniform thickness distribution, shapes which cannot<br />

be expressed by classical functions, and often with extreme behaviour like multiple<br />

eigenvalues. Thus, it is essential that we use a very flexible method and one in which<br />

we are experienced. It is advantageous if this method for response analysis can be<br />

used with different levels of accuracy.<br />

Many methods are available for the different response analyses, but none serves<br />

our purpose as well as the finite element method (FEM). It is therefore important to<br />

state: the unified approach to optimal design is based on the finite element method for<br />

response analysis. Advanced<br />

The process of optimal design should always terminate with a response analysis of<br />

the resulting optimal design. Stated in other terms: optimal design is not a separate<br />

subject, but an advanced extension of the response analysis.<br />

FEM-analysis


26 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

SELECTION OF<br />

DESIGN PARAMETERS<br />

TEMPORARY INITIAL<br />

DESIGN<br />

RESPONSE ANALYSIS<br />

OF THE DESIGN<br />

SENSITIVITY ANALYSIS<br />

OF THE DESIGN<br />

DECISION OF REDESIGN<br />

DESIGN<br />

YES CHANGED NO<br />

OPTIMIZED<br />

DESIGN<br />

Problem of optimal design:<br />

A highly non−linear problem<br />

Problem of optimal redesign:<br />

A linear programming problem<br />

DESIGN<br />

Figure 2.1: The approach to optimal design presented from the point of view of a<br />

number of sub-problems.


Names and concepts 27<br />

The advanced extension of response analysis has two aspects. Firstly, because<br />

of the iterative process of redesign, we have to do repeated response analysis. An<br />

efficient optimal design approach naturally takes this into account.<br />

Secondly and more important, for the decision of redesign we need to know how<br />

the response of a design changes with the design parameters. The analysis necessary<br />

for determining this is often named as sensitivity analysis. This analysis may be<br />

carried out numerically by differences, but often it is our goal to perform this analysis<br />

more analytically.<br />

As it is seen also from the present book, sensitivity analysis plays a leading role<br />

and many results are available. Entire books (Kleiber, Antunez, Hien and Kowalczyk<br />

1997) are written on the subject. Always<br />

sensitivities<br />

Some approaches to optimal design do not seem to involve the sensitivity analysis<br />

but a closer look shows that it is almost always present. As an example, the terms<br />

of an optimality criterion are gradients and must be obtained by a sensitivity analysis.<br />

It is therefore the opinion of the present author that: sensitivity analysis is the<br />

”cornerstone” of almost any approach to optimal design. Redesign<br />

When we focus on the third sub-problem, i.e. the problem of deciding the redesign,<br />

we find methods that are very much different in nature. There is no doubt that the<br />

decision of redesign is an optimization problem. Nevertheless, it is often attempted<br />

with ad hoc methods and without an objective, and it is difficult to assess the degree<br />

to which the results approximate the optimal redesign. Move-limits<br />

The problem of optimal design as whole is highly non-linear, and even the problem<br />

of optimal redesign is, in fact, non-linear. However, by move-limits we restrict the<br />

design changes so that linear expansions are acceptable. These move-limits are of<br />

major importance, as they are the key for the designer, to control the design process.<br />

In the primary redesigns we choose rather large move-limits, whereas in the final<br />

redesigns, small move-limits may be necessary. Classes of<br />

Still, if we choose to treat the decision of redesign as an optimization problem, we<br />

have many possible methods to choose among. For the optimal designs in the present<br />

book, the two major classes are:<br />

• OPTIMALITY CRITERIA methods<br />

• MATHEMATICAL PROGRAMMING methods<br />

and in the later class we shall primarily use<br />

• Linear Programming (LP), used sequentially (SLP)<br />

Further methods in the class of mathematical programming are:<br />

• Quadratic Programming (QP), used sequentially (SQP)<br />

• Dynamic Programming<br />

optimization<br />

procedures


Local/global<br />

parametrization<br />

28 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

• Geometric Programming<br />

• Method of Moving Asymptotes (MMA)<br />

• Interior or Exterior point methods<br />

• Stochastic methods<br />

• Generic Algorithms<br />

• Simulated Annealing<br />

• Neural Network<br />

• Evolutionary methods<br />

but the more detailed description of these procedures is not given in the present book.<br />

2.7.2 The overall strategy<br />

The selection of design parameters is by no means unique, and in relation to the<br />

optimization approach the manner in which the design is described is very important.<br />

This is especially true in relation to shape design.<br />

The one extreme is the local design variables, like thickness at a given point,<br />

position of a joint (node), or in relation to FEM, the element variables. The other<br />

extreme is the global design variables, which influence the design everywhere. The<br />

use of local design variables seems to dominate, although the use of global design<br />

variables has many advantages.<br />

One example of the use of global design variables is the description of a shape by a<br />

linear combination of a few known functions, in a way very parallel to mode expansion.<br />

The discrete design parameters are then the factors of the linear combination. Such<br />

an approach has the advantage of dealing only with a few design variables, and it is<br />

easy to ensure a smooth connection to regions of fixed design. The disadvantage is<br />

that the sensitivity analysis is enlarged, which counteracts the small number of design<br />

variables.<br />

With regard to the use of local design variables, the main advantage is that the<br />

gradients are often directly available, being, say specific energies that do not have to be<br />

computed separately. Naturally, the use of local design variables gives the possibility<br />

of a very detailed design but at the expenses of having to deal with many design<br />

variables. Furthermore, the use of local design variables gives the risk of obtaining<br />

degenerate designs and solutions that depend drastically on the discretization level.<br />

A final comment on the important question of selection of design variables is<br />

that many combinations of local and global design variables are possible, and that


Names and concepts 29<br />

the accuracy in description of the design need not be the same at the initial design<br />

iterations as at the final ones. This is easily obtained with local as well as with global<br />

design variables. The approach of design variable linking should also be kept in mind.<br />

The optimality condition for the optimal design in total (not for the individual<br />

redesigns) may also be viewed as a stopping condition. When to stop the sequence<br />

of redesigns ? As indicated in figure 2.1 there is a general practical answer to this<br />

question, which is: When the solution to the linear programming problem of redesign<br />

is to make no design changes, then the design is optimal. This is, of course, true<br />

in the local sense only, but we have to deal with the question of local versus global<br />

optimal solution anyhow, and base our beliefs on engineering judgment and optimal<br />

design from different initial designs.<br />

There is no contradiction between this stopping condition and other statements<br />

like stationarity or Kuhn-Tucker condition. The nice aspects of the simple condition<br />

are that the stopping condition is valid also for solutions with non-stationarity and<br />

the condition does not have to be expressed explicitly.<br />

Stopping<br />

condition


30 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...


Chapter 3<br />

Design of trusses<br />

Trusses are structures that are simple to analyse, because each point of the structure<br />

is only subjected to a unidirectional state of stress/strain. The basic element of a<br />

truss is a bar and the bars are connected through joints that only transfer a normal<br />

force (tensile or compression) to each bar. In this chapter we only deal with these<br />

simple models, but local stability of each bar is also considered. Simple<br />

In addition to being simple structures, trusses are also very efficient structures<br />

with the possibility of designing fully stressed structures, where each point in the<br />

structure is used to its limit. For a statically determined truss each bar can be designed<br />

independently and therefore in an optimal design must be fully stressed. The notion<br />

of fully stressed for multiple load cases means that each bar is fully stressed, at least<br />

in one of the load cases.<br />

With reference to chapters 16, 19, and 20 we shortly discuss the theoretical and<br />

numerical background and choose bar members with circular cross-section as our basic<br />

Bars and joints<br />

unidirectional<br />

element. The influence of hollow cross-sections in relation to design for local stability Compressive<br />

is studied.<br />

Through a number of specific examples (cantilever and bridge types) the importance<br />

of the difference between tensile bars and compression bars is put in focus. All<br />

the results for plane trusses as well as for space trusses are obtained by an interactive<br />

stability<br />

program that also includes the design of supports as an integrated part. Single<br />

Early results for cantilevers and bridges (Pedersen 1969) show how size and topol- load case<br />

ogy optimal design (bar areas and selection of bars) are combined with shape optimal<br />

design (joint positions), but these examples are restricted to single load case models.<br />

In the extensions to multiple load cases and/or including constraints on displacements<br />

(stiffness in addition to strength), the design parameters will only be size and<br />

shape, and we use a more traditional, iterative optimal design procedure with anal- Multi-purpose<br />

trusses<br />

31


32 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

ysis, sensitivity analysis and linear programming. Some attempts to include also<br />

topology is discussed, but in general these practical important design problems are<br />

still open for new research initiatives.<br />

3.1 Theoretical and numerical background<br />

If we assume that allowable stresses and joint positions are given, then the problem<br />

of finding a truss of minimum mass is a linear programming problem. Such optimal<br />

design problems can be solved using the simplex optimization procedure, and this was<br />

Statically<br />

done already by (Dorn, Gomory and Greenberg 1964) and by (Fleron 1964).<br />

determined With only a single load case, local stability is taken into account, and we can prove<br />

that a statically determined truss is a solution. This is proven in chapter 16 in a more<br />

general setting of minimum cost and starting also without a complete specification of<br />

supports.<br />

The modified simplex procedure, described in chapter 19, is used to obtain the<br />

solutions to a number of illustrative problems, which we discuss in sections 3.2 and<br />

Size, shape<br />

and topology<br />

3.3, but without going into detail about the optimization method itself. Also the size,<br />

shape, and topology combined solutions in section 3.4 are obtained with a similar<br />

procedure combined with a more traditional mathematical programming solution for<br />

the joint positions (shape variables).<br />

In the last sections 3.5 and 3.6, all practical constraints can be accounted for,<br />

but the truss topology is then assumed to be given. Sensitivity analysis plays an<br />

important role and the formulation as well as the solution procedure are described<br />

in chapter 20. In the present chapter we concentrate on the discussion related to the<br />

specific examples.<br />

3.1.1 Common parameters for the examples<br />

In addition to some specified joint positions and the actual loads to be transferred,<br />

only few parameters are needed for the optimization. In most of the examples we use<br />

the same values for these data and therefore list them here:<br />

General data<br />

for examples Modulus of elasticity E 2.0 · 10 11 Pa<br />

Stress limit of proportionality σP 1.6 · 10 8 Pa<br />

Allowable tensile stress σT 2.0 · 10 8 Pa<br />

Maximum allowable compressive stress σC 1.6 · 10 8 Pa<br />

Factor of safety for slender columns ξ 2.5<br />

Mass density ρ 7500 kg/m 3<br />

Cross-sectional parameter α := √ I/A see section 3.1.2


Design of trusses 33<br />

3.1.2 Truss members with circular cross-section<br />

We shall study the influence of local stability on the optimal topology by changing<br />

a cross-sectional parameter α. Also for visualization of the results, we need crosssectional<br />

outer dimensions. Therefore we specialize to truss members with circular<br />

cross-section (pipes) with outer diameter do and inner diameter di, specified by the<br />

ratio µ<br />

di = µdo with 0 ≤ µ


34 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

3.4 including optimal position of the free joints and with different number of free<br />

joints.<br />

3m<br />

3m<br />

80000 N 40000 N 20000 N 20000 N<br />

3m<br />

Figure 3.1: Load and geometry assumptions with the possible supports for the cantilever<br />

test example. In addition to the loaded and supported joints this example has<br />

three joints.<br />

The initial chosen topology with only size optimization (fully stressed) is shown<br />

in figure 3.2 together with the optimized topology. For this example the self-weight<br />

of the truss is not taken into account. The used colour code in all the examples are<br />

blue for tensile members, orange for compressive slender columns, red for compressive<br />

short columns, and green for zero strings. For definition of these different types, see<br />

chapter 16.<br />

Fully<br />

stressed As commented in general, both topology’s in figure 3.2 are fully stressed and statically<br />

determined. We note the repeated sub-structure in the optimal design, which without<br />

Self-weight<br />

the string support also has a column of double length. This problem is not actual in<br />

the optimal design shown in figure 3.3, where the self-weight is taken into account.<br />

included Note, that with self-weight we totally add loads in the same direction as the<br />

given external loads and still the optimal design has a smaller total mass (685 kg to<br />

compare with 707 kg). The explanation for this is that the vertical columns when<br />

self-weights are taken into account, then transfer smaller compressive forces, and in<br />

3m<br />

3m


Design of trusses 35<br />

Figure 3.2: Cantilever with given loads in gravity directions, but without self-weights.<br />

Initial size optimized design (mass = 902 kg) at the top and optimal topology design<br />

(mass = 707 kg) at the bottom. <strong>Solid</strong> cross-sections (α =0.3).<br />

this manner make a lighter structure possible. We notice that the topology is changed<br />

when self-weight is taken into account.<br />

Figure 3.3: Cantilever with given loads in gravity directions, and with self-weights.<br />

Initial size optimized design (mass = 922 kg) as in figure 3.2 and the optimal topology<br />

design here results in mass = 685 kg. <strong>Solid</strong> cross-sections (α =0.3).<br />

In a formulation without local stability constraints and with equal allowable tensile<br />

and compressive stress, the optimal design will be the same with opposite external<br />

Changing<br />

load direction


Changing<br />

type of<br />

cross-section<br />

36 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

loads. However, when compressive members are more expensive the optimal topology<br />

will normally be different. This is illustrated in figure 3.4. The self-weights now<br />

counteract the given external loads, and it is possible to obtain a lighter structure, for<br />

the initial topology as well as for the optimal topology. Comparing with the optimal<br />

topology in figure 3.3 we notice a completely different solution.<br />

Figure 3.4: Cantilever with given loads opposite gravity directions (opposite direction<br />

of forces in figure 3.1), and with self-weights. Initial size optimized design (mass =<br />

683 kg) at the top and optimal topology design (mass = 605 kg) at the bottom. <strong>Solid</strong><br />

cross-sections (α =0.3).<br />

The solutions presented in figures 3.2, 3.3 and 3.4 are based on an assumption of a<br />

solid cross-section (α =0.3), and thus compressive members dominate the total mass.<br />

With other cross-sectional assumptions we get different optimal designs, as illustrated<br />

in figures 3.5, 3.6, 3.7 and 3.8. Comparing the designs in figure 3.5 corresponding to<br />

α = 1 (a hollow circular cross-section with di =0.92d0) with the results in figure 3.2<br />

and 3.3, we first of all notice the much less necessary mass, but also the change in the<br />

optimal topology. Figure 3.5 shows that a number of the compressive members are<br />

now classified as short columns (in colour: red not orange). See chapter 16 for the<br />

classification.


Design of trusses 37<br />

Figure 3.5: Cantilever with given loads in gravity directions, and with self-weights.<br />

Initial size optimized design (mass = 336 kg) at the top and optimal topology design<br />

(mass = 280 kg) at the bottom. Cross-sections corresponding to α =1.<br />

Figure 3.6: Cantilever with given loads opposite gravity directions, and with selfweights.<br />

Initial design size optimized (mass = 282 kg) at the top and optimal topology<br />

design (mass = 264 kg) at the bottom. Cross-sections corresponding to α =1.


38 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

The results corresponding to extreme thin-walled cross-sections are shown in figure<br />

3.7 and 3.8. In these designs, all compressive members are short columns (red, not<br />

orange) and taking local stability into account is then not so important. However,<br />

with compressive allowable stress only 80% of allowable tensile stress, see section<br />

3.1.1, we still find different design when the directions of the given external loads are<br />

changed.<br />

Figure 3.7: Cantilever with given loads in gravity directions, and with self-weights.<br />

Initial size optimized design (mass = 223 kg) at the top and optimal topology design<br />

(mass = 188 kg) at the bottom. Thin walled cross-sections (α = 3).<br />

3.2.2 A cantilever truss with 17 joints<br />

The optimal cantilever designs presented depend on the basic 9 joints and their positions.<br />

Later we shall discuss the possibility of optimizing the positions of some of the<br />

joints, but let us primarily just double the number of joints in the cantilever length<br />

direction. The total given loads are not changed, with each load force just divided


Design of trusses 39<br />

Figure 3.8: Cantilever with given loads opposite gravity directions, and with selfweights.<br />

Initial size optimized design (mass = 216 kg) at the top and optimal topology<br />

design (mass = 180 kg) at the bottom. Thin walled cross-sections (α = 3).<br />

into two neighboring joints. Two cases are shown both with loads in the gravity direction,<br />

both with self-weights included, differing only in cross-sections corresponding<br />

to α =0.3 and α = 1, respectively.<br />

Comparing the results in figure 3.9 with the results in figure 3.3 certain similarities<br />

in topology are recognized and the resulting masses are lower with the many joints.<br />

This is due to the possibility of shorter columns, but in practice the higher cost of<br />

many connections may counteract this.<br />

Comparing the results in figure 3.10 with the result in figure 3.5 certain similarities<br />

in topology are again recognized and the resulting masses are again lower with the<br />

many joints, only 3 kg for the initial topology but 36 kg for the optimal topology. Repeated<br />

Notice again the repeated sub-structures.<br />

sub-structures<br />

3.2.3 A bridge truss with 16 joints<br />

This example is also an optimization taken from (Pedersen 1969), in order to illustrate<br />

a classical civil-engineering design problem in a simple model. Figure 3.11 shows the


40 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 3.9: Cantilever with given loads in gravity directions, and with self-weights.<br />

Initial size optimized design (mass = 836 kg) at the top and optimal topology design<br />

(mass = 548 kg) at the bottom. <strong>Solid</strong> cross-sections corresponding to α =0.3.<br />

Figure 3.10: Cantilever with given loads in gravity directions, and with self-weights.<br />

Initial size optimized design (mass = 333 kg) at the top and optimal topology design<br />

(mass = 244 kg) at the bottom. Cross-sections corresponding to α =1.


Design of trusses 41<br />

load and geometry assumptions.<br />

L/8<br />

7 forces, each 100000 N<br />

L =32m<br />

Figure 3.11: Load and geometry assumptions with the possible supports for the<br />

bridge test example. In addition to the loaded and supported joints this example has<br />

seven joints.<br />

Figure 3.12 shows the optimal topology based on joint positions with a parabolic<br />

reference and length/height = 5.5. The resulting total mass of the optimal bridge<br />

is 4157 kg. This bridge truss optimization is extended in section 3.4 to include the<br />

positions of the seven upper joints as design parameters. Furthermore, the solutions<br />

for different total length of the bridge are then presented. For all these problems the Same optimal<br />

optimal topology in figure 3.12 is unchanged.<br />

topology<br />

Figure 3.12: <strong>Optimal</strong> topology for the bridge problem defined in figure 3.11, with<br />

compressive bars along the parabola and tensile bars in the remaining part of the<br />

truss.<br />

3.2.4 A beam/bridge truss with 24 joints<br />

We design a beam/bridge that span 44 meter. At the lower joints the beam/bridge<br />

truss is loaded with a uniformly distributed load of 2500 N/m (totally 110000 N),<br />

plus the weight of the structure itself, which adds about 10%. The design depends<br />

on the possible height of the beam/bridge and especially on the possible supports.


42 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

With symmetry assumed we only model (and only show in the figures) half of the Half a model<br />

by symmetry<br />

structure, using for this 24 joints.<br />

In figure 3.13 the possible height is only 1 meter and fixed supports are assumed<br />

possible at the two left boundary joints in addition to the supports that model the<br />

symmetry line at the right boundary joints. The characteristic of the optimal design<br />

in figure 3.13 is a combination of a cantilever part that supports a mid-part of mostly<br />

tensile members.<br />

Figure 3.13: Half of a low beam with uniform loads at the lower joints, two left<br />

joints fixed and two right joints at symmetry line. Self-weights included. Initial size<br />

optimized design (mass = 845 kg) at the top and optimal topology design (mass =<br />

532 kg) at the bottom. <strong>Solid</strong> cross-sections corresponding to α =0.3.<br />

Cantilever +<br />

hanging part The relative length of the cantilever part and the tensile hanging part is depending<br />

upon the relative mass (cost) of tensile to compressive members (here the α parameter<br />

is equal to 0.3), and upon the possible height of the beam/bridge. Figure 3.14 shows<br />

the optimal design for a height of 3 meter, and for this case only the hanging part<br />

remains. Thus, we end up with a structure with only tensile members, a very mass<br />

effective structure.<br />

Importance<br />

of supports<br />

However, the total reaction at the left top support may be a limiting factor. We<br />

therefore try to optimize the design without this support possibility, and then obtain<br />

the result in figure 3.15. Now, also the optimal design is very much increased in mass.<br />

Note, that again in this optimal design we find repeated sub-structures, and in the<br />

mid-part of the upper structure short columns appear. The conclusion is that the<br />

cost and possibility of supports must be carefully evaluated, before the truss design<br />

is decided.<br />

It is concluded that parameter studies are necessary and parameter studies based<br />

on optimal designs are more conclusive than parameter studies based on non-optimal<br />

designs.


Design of trusses 43<br />

Figure 3.14: Half of a high beam with uniform loads at the lower joints, two left joints<br />

as possible supports and two right joints at symmetry line. Self-weights included.<br />

Initial size optimized design (mass = 897 kg) at the top and optimal topology design<br />

(mass = 195 kg) at the bottom. <strong>Solid</strong> cross-sections corresponding to α =0.3.<br />

Figure 3.15: Half of a bridge with uniform loads at the lower joints, only lower left<br />

joint fixed and two right joints at symmetry line. Self-weights included. Initial size<br />

optimized design (mass = 1320 kg) at the top and optimal topology design (mass =<br />

933 kg) at the bottom. <strong>Solid</strong> cross-sections corresponding to α =0.3.


Combined<br />

44 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

3.3 Size, supports and topology for<br />

optimal single load case space trusses<br />

The theoretical formulation in chapter 16 and the computer program based on this<br />

is for three dimensional trusses, and in section 3.2 just applied to plane trusses. The<br />

main challenge for space trusses is to present in graphics the results. For this reason<br />

only one example is presented here.<br />

3.3.1 A tower<br />

A tower space truss is shown in the perspective drawing in figure 3.16. The height of<br />

the tower is 16 meter and the assumed single load case consists of horizontal forces of<br />

magnitude 10000 N at each of the three top joints, all forces in the same direction. The<br />

further data are as for the plane trusses given in section 3.1.1 and the cross-sectional<br />

parameter α is set to α =1.<br />

This space truss has 15 joints and 39 bars which must be chosen out of 105 possible<br />

bars. The resulting optimal truss only gives 15 active bars and 24 ”zero strings”. This<br />

is often the case when only a few joints are subjected to external forces.<br />

As expected, this optimization results in two tensile bars connected directly to<br />

the foundation. The mass of the initial, size optimized topology is 373 kg, while<br />

the optimized topology returns a structure of 255 kg. The example is taken from<br />

(Pedersen 1993).<br />

3.4 Size, shape and topology for optimal 2D-trusses<br />

The size and topology optimized designs in the previous sections of this chapter are<br />

obtained by a fully stressed design, and topology optimization is obtainable due to<br />

the basic knowledge of a statically determined truss, proven in chapter 16.<br />

shape design This holds independent of the position of the actual joints and it is thus possible<br />

to combine the size and topology optimization with optimization of positions for joints<br />

that do not have a fixed position. This combination with shape optimization is applied<br />

in (Pedersen 1969) for cantilever trusses as well as for bridge trusses, corresponding<br />

to the cases in sections 3.2.1 and 3.2.3.<br />

The extended formulation where a number of joint positions are also treated as<br />

design variables thus includes shape optimization. The solutions are obtained by<br />

Two level<br />

iterative<br />

optimization<br />

iterative redesign where improved joint positions by analysis, sensitivity analysis and<br />

linear programming, are followed by a new size and topology optimization. <strong>Optimal</strong><br />

designs for the two specific cases in sections 3.2.1 and 3.2.3 are presented.


Design of trusses 45<br />

Figure 3.16: Perspective drawing of tower space truss. Left: initial topology resulting<br />

in 373 kg. Right: optimal topology resulting in 255 kg. Cross-sectional pipes<br />

corresponding to α =1.


Influence of<br />

number of<br />

joints<br />

Better<br />

46 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

3.4.1 The cantilever truss with 7 to 12 joints<br />

With reference to the optimal design in figure 3.3 that by size and topology optimization<br />

resulted in a total mass of 685 kg, we optimize also the positions of the three<br />

”free” joints and obtain now a total mass of 582 kg. This design is shown in figure<br />

3.17 together with solutions for alternative number of free joints.<br />

We note in figure 3.17 that the improvement from 5 to 6 free joints is larger<br />

than the improvements both from 3 to 4 free joints and from 4 to 5 free joints. The<br />

explanation is that with 6 free joints a third joint can be supported directly from<br />

the fixed upper supports, which with ”expensive” compressive bars is of dominating<br />

influence.<br />

The decision of the number of joints must be taken on the basic of cost of joints,<br />

and this is not included in the present formulation.<br />

3.4.2 The bridge truss with 16 joints<br />

Similar to the extension of the cantilever problem, the bridge problem described in<br />

figures 3.11 and 3.12 is also optimized with respect to joint positions, and figure 3.18<br />

shows the result. The general tendency in the shape optimization is that the free<br />

joints move closer to the supports and that the height of the structure increases. This<br />

increase in height is a balance between longer bars and the possibility to get the forces<br />

more vertically into the supports.<br />

force flow Figure 3.18 shows the free joint coordinates xi and the actual values can be found<br />

in figure 3.19. Relative values are given for different total length of the bridge, and<br />

the resulting masses as function of the total load and of the total length can be found<br />

in (Pedersen 1969).<br />

3.5 Size and shape for optimal multi-purpose plane<br />

trusses<br />

Also with shape optimization the optimal designs are fully stressed designs, and topology<br />

optimization is possible. The severe limitation behind these solutions is the<br />

assumption of only one load case, and the fact that only strength constraints are<br />

Not statically considered although also local stability of each bar is accounted for.<br />

determined With multiple load cases and/or with displacement constraints the knowledge of a<br />

statically determined optimal truss is lost. We are therefore forced to use the general<br />

procedure described in chapter 20, and the topology design problem can not be solved<br />

as done previously in the present chapter. In the following sections of this chapter<br />

we therefore restrict to size and shape optimization. On the other hand all kind of


Design of trusses 47<br />

1136 kg<br />

756 kg<br />

582 kg<br />

520 kg<br />

473 kg<br />

401 kg<br />

Figure 3.17: <strong>Optimal</strong> topology and shape design with different number of free joints.<br />

Same problem as in figure 3.3. Tensile members with thin lines, and compressive<br />

members with thicker lines.


Not fully<br />

stressed<br />

Results with<br />

different<br />

assumptions<br />

48 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

(x3,x4)<br />

(x7,x8)<br />

(x11,x12)<br />

(L/2,x16)<br />

Figure 3.18: <strong>Optimal</strong> shape design compared to the parabolic initial shape design.<br />

The optimal topology is unchanged, but the new sizes decrease the total mass from<br />

4157 kg to 3852 kg.<br />

practical constraints can then be involved, and in fact the more constraints, the more<br />

stable is the optimization process normally.<br />

As specific examples we again choose a bridge and a cantilever for the plane trusses<br />

and for space trusses a dome problem.<br />

Figure 3.20 shows a 12 joint bridge subjected to five equal forces, that act independently,<br />

say to model a moving load. This problem is solved in different versions, with<br />

and without local stability constraints, with and without displacement constraints.<br />

The optimal shape solutions are collected in figure 3.21.<br />

For comparison we list the results with only size optimization for the shape in<br />

figure 3.20. The following total masses are obtained (in the parentheses relative to<br />

the optimal design in figure 3.21):<br />

• without stability constraints and without displacement constraints 1876 kg<br />

(+13.3 %) for a fully stressed design and 1868 kg (+12.8 %) for an optimal<br />

size design.<br />

• without stability constraints but with maximum displacement constraint 3730<br />

kg (+28.1 %) for a proportional change of the optimal design without displacement<br />

constraint and 3526 kg (+21.1 %) for an optimal size design.<br />

• with stability constraints but without displacement constraints 3141 kg (+8.1<br />

%) for a fully stressed design equal to the optimal size design.<br />

• with stability constraints and with maximum displacement constraint 4030 kg<br />

(+21.6 %) for a proportional change of the optimal design without displacement


Design of trusses 49<br />

0.35<br />

0.30<br />

0.25<br />

0.20<br />

0.15<br />

0.10<br />

0.07<br />

xi/L (relative position)<br />

i = 11<br />

i = 16<br />

i = 12<br />

i = 7<br />

i = 8<br />

i = 4<br />

i = 3<br />

16 24 32 48 64<br />

L meter<br />

Figure 3.19: <strong>Optimal</strong> shape design for different total lengths of the bridge, given as<br />

relative positions of the free joints with numbering according to figure 3.18<br />

1st 2nd 3rd 4th 5th<br />

Figure 3.20: Initial parabolic shape design for a multi-purpose bridge truss with<br />

fixed topology. Five independent load cases.


50 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

constraint and 3668 kg (+10.6 %) for an optimal size design.<br />

3.6 Size and shape for optimal multi-purpose space<br />

trusses<br />

A dome space truss problem is solved in (Pedersen 1973) and we only comment the<br />

different solutions, which are shown in figures 3.22 and 3.23.<br />

Figure 3.22-left shows a paraboloid shape design for a dome with a diameter of 30<br />

meter. This dome is assumed to be loaded by four vertical load cases (forces toward<br />

the plane of the supports):<br />

1) a force at the central joint # 1 with the value 300000 N.<br />

2) at all the free joints, each force has the value 30000 N.<br />

3) a force at the central point # 1 with the value 150000 N, and forces at the joints<br />

# 4 and 5, each force has the value 100000 N.<br />

4) a force at the central point # 1 with the value 150000 N, and forces at the joints<br />

# 2, 3 and 4, each force has the value 70000 N.<br />

For fixed paraboloid shape, the optimal design of the bar areas results in a fully<br />

stressed design, i.e. at least one bar in each linked group is stressed to the allowable<br />

limit for at least one of the loading conditions. Such a design is easily obtainable<br />

by four to five fully stressed iterations, where the mutual connections are neglected.<br />

For heights of the paraboloid shape equal to 6, 9, 12, and 15 meter, the mass of<br />

the optimal size domes is 7247, 6679, 6929 and 7638 kg, respectively. The optimal<br />

paraboloid shape design is the one shown in figure 3.22-left with height 9.25 m and<br />

optimal mass 6675 kg.<br />

Figure 3.22-mid shows the optimal topology not restricted to the paraboloid<br />

shape. The total mass for this design is 5623 kg, i.e. the paraboloid shape implies<br />

18.7 % more mass. For this comparison, it must be noted that the load systems<br />

are assumed design independent, i.e. in the structure in figure 3.22-mid the forces<br />

acting on the joints # 2 to 13 are closer to the supports, but the more important<br />

forces acting on joint # 1 have fixed horizontal position.<br />

An objection to the design in figure 3.22-mid may well be that it is too high (11.34<br />

meter), and therefore the problem is optimized with a maximum height of the 9.25 m<br />

as the result in figure 3.22-left. We then get the design in figure 3.22-right with a total<br />

mass of 5733 kg, i.e. only 1.7 % more than the result without the height constraint.<br />

We also optimize the dome with a constraint on displacements. The maximum<br />

vertical displacement of joint # 1 for the designs in figure 3.22 are 0.026 m, 0.026 m,<br />

and 0.032 m, respectively. If we only allow this displacement to be 0.01 m we find<br />

the optimal design in figure 3.23-left. The total mass for this design is 5989 kg, and


Design of trusses 51<br />

.<br />

.<br />

.<br />

.<br />

(0, 7.53)<br />

1656 kg<br />

No stability and displacement constraints<br />

(0, 8.64)<br />

(5.53, 7.80)<br />

(6.33, 6.72)<br />

(11.53, 4.37)<br />

2911 kg<br />

No stability constraints, maximum displacement 0.01 m<br />

(10.65, 5.05)<br />

(0, 5.00) (6.29, 4.85)<br />

(11.38, 2.66)<br />

2905 kg<br />

Stability constraints, without displacement constraints<br />

(0, 7.39)<br />

(5.78, 6.53)<br />

(10.81, 4.37)<br />

3315 kg<br />

Stability constraints, maximum displacement 0.01 m<br />

Figure 3.21: <strong>Optimal</strong> size and shape design of a multi-purpose truss, subjected to<br />

five load cases. The topology is fixed and not optimized.


52 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 3.22: <strong>Optimal</strong> size and shape design of a multi-purpose dome space truss,<br />

subjected to four load cases. The topology is fixed and not optimized. Top: Enlarged<br />

picture to see the linking. Left: paraboloid shape resulting in 6675 kg. Mid: optimal<br />

shape without height constraint resulting in 5623 kg. Right: optimal shape with<br />

height constraint resulting in 5733 kg.


Design of trusses 53<br />

Figure 3.23: <strong>Optimal</strong> size and shape design of a multi-purpose dome space truss,<br />

subjected to four load cases. The topology is fixed and not optimized. Left: optimal<br />

shape with α =1.0 5989 kg. Mid: optimal shape with α =0.5 9832 kg. Right:<br />

optimal shape with α =1.5 4242 kg.


54 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

compared with the design in figure 3.22-mid we have reduced the actual displacement<br />

by a factor of 2.6 by an increase of only 6.5 % in the total mass.<br />

The optimal designs fulfill a balance between a favourable transmission of external<br />

forces to the supports (which normally give rise to rather high structures), and the<br />

influence of local stability which tends to shorten the bars. Therefore, the optimal<br />

design depends on the characteristics of the bar cross-section, here described by the<br />

parameter α. The dome results shown so far correspond to α =1.0. In figure 3.23-mid<br />

the optimal design is given for α =0.5 with increased total mass to 9832 kg, and in<br />

figure 3.23-right the optimal design is given for α =1.5 with decreased total mass to<br />

4242 kg.<br />

The optimal design for α =1.0 in figure 3.22-mid is in no way a design ”between”<br />

the optimal design for α =0.5 in figure 3.23-mid and the optimal design for α =1.5<br />

in figure 3.23-right. As we have no guaranty of global optima, a further ”test of<br />

the designs” was performed by optimizing based on α =1.0, starting from both the<br />

optimal designs corresponding to α =0.5 and α =1.5. In both cases we again ended<br />

up with the result of figure 3.22-mid.<br />

The same structural dome problem is in chapter 5 modelled as a frame structure<br />

with account for the much more complicated stress distributions.<br />

3.7 A large space truss with constraints on<br />

eigenfrequencies, displacements, stresses and buckling<br />

Section to be included, based on (Pedersen and Nielsen 2002)<br />

3.8 Approach to topology optimization for<br />

multi-purpose space trusses<br />

Section to be included, based on (da Silva Smith 1997)


Chapter 4<br />

Continua of<br />

uniform energy density<br />

4.1 General comments<br />

In this chapter we show a number of continua designed to obtain uniform (or almost<br />

uniform) energy density, restricted to only one load case. From chapter 14 we will<br />

know that these designs are optimal with respect to stiffness as well as with respect<br />

to strength, when the total volume is not changed.<br />

Stiffness is a global measure for the whole continuum and is here set equal to the<br />

total elastic energy, which in itself is equal to the work of the external loads. The<br />

objective in stiffness optimization is to minimize the stored elastic energy, equivalent<br />

Only single<br />

load case<br />

to minimization of the work of the external forces, which is often named compliance. Stiffness<br />

Strength is a local measure and thus focus is put on the most critical point(s) in a<br />

continuum. As a general measure for a critical quantity we choose the elastic energy<br />

density, and thus strength optimization is to minimize the maximum elastic energy<br />

density, located in the continuum. If uniform energy density is obtained, then the<br />

compliance is stationary, see chapter 12.<br />

4.1.1 Design variables and design constraints<br />

A continuum example is shown in figure 4.1, and with reference to this figure we<br />

define some quantities that are general for the different examples. The continuum is<br />

modelled by finite elements, and absolute quantities depend on the chosen model, thus<br />

we mainly present results by relative quantities and keep the finite element modelling<br />

55<br />

and strength


56 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 4.1: A cantilever shown as a continuum example, modelled by finite elements.<br />

Total volume ¯ V and total area Ā are kept constant during optimization. (Isolines for<br />

energy densities shown for a uniform thickness field with the indicated uniform line<br />

load).<br />

constant during optimization. The boundary conditions are indicated by green (gray)<br />

arrow heads, thus the cantilever in figure 4.1 is completely fixed at the left boundary.<br />

The actual load condition is indicated by black arrows, in figure 4.1 a uniform<br />

line-load on the top boundary. Blue (dark) contour line shows the non-deformed, and<br />

red (gray) contour line shows the deformed contour (magnified). Total volume ¯ V and<br />

total area Ā are kept constant, so, the mean thickness Given<br />

¯t is<br />

volume ¯t = ¯ V/ Ā and ¯ V = <br />

(4.1)<br />

Thicknesses<br />

or densities<br />

Side<br />

constraints<br />

where te is the thickness in element (domain) e and area ae is the corresponding area.<br />

Optimization is started from a uniform thickness field te = ¯t.<br />

The actual design variables are then te, giving the thickness field. We may as an<br />

alternative take the relative volume densities ρe as design variables, and this would be<br />

the appropriate quantities for 3D-problems. However, in this chapter we only show<br />

2D-problems. In addition to the global constraint on given volume, we have local side<br />

constraints<br />

αmin¯t = tmin ≤ te ≤ tmax = αmax¯t (4.2)<br />

with typical relative values being αmin =0.001 and αmax = 10. The lower limit has<br />

very little influence on the results and is seen as ”holes”, while the upper relative limit<br />

may be responsible for not reaching a complete uniform energy density. However, this<br />

limit is necessary because large gradients in thickness design put severe limitations<br />

to the accuracy of the finite element modelling, and often such a constraint is also<br />

important from a more practical point of view.<br />

e<br />

teae


Continua of uniform energy density 57<br />

4.1.2 Presentation of results<br />

In addition to graphical illustrations of thickness fields and fields of strain energy<br />

density, a few relative numbers are needed to clarify the obtained results. The starting<br />

analysis with a uniform thickness design gives an initial total elastic energy U0, from<br />

which an initial mean energy density umean0 = U0/ ¯ V is given. Also this starting<br />

analysis gives a maximum energy density umax0, and an energy concentration factor,<br />

defined as umax0/umean0.<br />

The total elastic energy of the optimized design is U and the goal of optimization<br />

is to obtain uniform energy density umax = umean = U/ ¯ V , from which follows that<br />

a gain factor for improvement in stiffness can be defined as umean0/umean = U0/U,<br />

and a gain factor for improvement in strength can be defined as umax0/umax. Reference<br />

Especially the upper bound tmax should also be seen in relation to the reliability<br />

of the finite element analysis. When the gradients of the thickness field get extended,<br />

the accuracy in modelling may be violated.<br />

The effect of the upper bound is that in most cases umax/umean > 1 results even<br />

for the optimized continuum that gives almost uniform energy density. A good measure<br />

for the influence from the upper bound tmax is given by this ratio umax/umean.<br />

The procedure for obtaining the optimized designs is rather simple and is described<br />

in section 4.11.<br />

values<br />

4.2 Bars, beams and beam-bars in 2D-formulation<br />

Axial load<br />

or bending<br />

Bars, beams and beam-bars are effectively analysed by one dimensional models when a<br />

number of conditions are satisfied. These conditions relate to slenderness, to simplicity<br />

in design, and to loads giving rise to a not too complicated displacement field.<br />

With two dimensional modelling we can analyse more complex bars, beams and<br />

beam-bars, but naturally also with limitations, before three dimensional modelling is<br />

necessary. The name bar merely reflects that the axial stress flow is of major concern.<br />

With 2D-modelling we can also analyse and design very short beams, where also the<br />

name of beam is questionable. Slender<br />

Figure 4.2 illustrates nine problems, for which we show thickness fields, that return<br />

almost uniform energy density. The pure beam problems and the pure bar problems<br />

might be solved with only half the models due to symmetry and skew-symmetry,<br />

but due to the bar-beam problems full models are used. The actual finite element<br />

models have 11042 degrees of freedom and 10800 design variables (thicknesses of 10800<br />

or short<br />

triangular elements). Finite element<br />

models<br />

Immediately, a suggested design for the bar problem would be a uniform bar<br />

only of the width corresponding to the area of uniform external load. However, such<br />

a design is questionable at the support where stress concentration can result. The


58 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

isolines in figure 4.2 for this case illustrate the complexity. Also the Saint-Venant<br />

principle is nicely illustrated in figure 4.2, and in the solutions to follow.<br />

Figure 4.2: The basic models for optimization of bars, beams, and beam-bars in<br />

three different lengths, i.e. in combination 9 cases. Isolines for energy densities are<br />

shown for the uniform thickness design with the loads shown at the free ends. The<br />

colour scale is red (gray) for low values, green for medium values and blue/violet<br />

(dark) for large values.


Continua of uniform energy density 59<br />

4.3 Bars: long, medium and short<br />

Figure 4.3 shows the results for a rectangular domain, which is fixed at the left end<br />

and loaded at the free right end, only at the mid third part and in the length direction,<br />

i.e. orthogonal to the supports.<br />

The optimized thickness design at the top of figure 4.3 is illustrated by isolines<br />

together with five cross-sectional ”cuts” that more clearly show the thickness distribution.<br />

The maximum thickness is set to ten times the average thickness (initial<br />

design), but is not reached.<br />

The mid figure illustrates directions of principal stresses for the optimized design,<br />

where isolines of energy density disappear due to the uniformity in design domain.<br />

The bottom figure shows the isolines for energy density distribution in the uniform<br />

Only axial<br />

forces<br />

initial design, added the principal stress directions. Theoretical<br />

For a design of uniform thickness (only in the part directly from the uniform<br />

external stress σ) a finite element model results in total elastic energy Uref = ¯ Vσ 2 /E<br />

with also uniform energy density uref = σ 2 /E, where E is the modulus of elasticity.<br />

Such a design can result in stress concentration at a support, which is wider than the<br />

width of a uniform thickness, rectangular domain, see the discussion in section 4.9.<br />

The obtained optimal designs give values which, within a few tenth of a percent, are<br />

equal to these reference values.<br />

In table 4.1 values for the initial design with uniform thickness over the whole<br />

rectangular domain, relative to the values for the optimal design are given. Compliance<br />

U0/U =1.18 is thus 18% more for the initial design and the stress concentration<br />

umax0/umean0 =11.8 is more than a factor of ten, with the actual value depending<br />

on the specific finite element model.<br />

For the optimized design with thickness design shown at the top in figure 4.3,<br />

the energy density field (shown in the mid of figure 4.3) is almost uniform and the<br />

relevant ratios are umax0/umax =13.8 and umax/umean =1.006. With the same<br />

finite element mesh the maximum energy density is thus decreased with a factor of<br />

14.9, corresponding to a stress decrease with a factor of 3.9 (= √ 14.9). A uniform<br />

stress distribution is obtained.<br />

Figure 4.4 shows the similar results for a shorter rectangular domain and for a<br />

squared domain, respectively. In table 4.1 we have collected the important ratios<br />

for energy (stiffness) and for energy density (strength). In general we note that the<br />

optimization has a stronger influence for the shorter domains.<br />

With uniform thickness over the full design domain, the compliances are 18%,<br />

27%, and 54% too high, and the energy concentrations are factors of 14.9, 14.0, and<br />

13.8 too high, respectively for the different lengths of the design domain. With optimal<br />

thickness design we obtain in all three cases uniform energy density within less than<br />

1% (umax/umean =1.006).<br />

reference


60 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 4.3: A bar with a 3 × 9 rectangular design domain. Top: resulting optimized<br />

thickness design. Mid: corresponding resulting field of energy densities (almost uniform).<br />

Bottom: the field of energy densities for a uniform thickness design. The black<br />

”hatch” is an attempt to show the principal stress directions.


Continua of uniform energy density 61<br />

Figure 4.4: Bars with a 3 × 3anda3× 6 rectangular design domains. Top: resulting<br />

optimized thickness designs. Mid: corresponding resulting fields of energy densities<br />

(almost uniform). Bottom: the fields of energy densities for uniform thickness design.<br />

The black ”hatch” is an attempt to show the principal stress directions.


Combined<br />

bending and<br />

axial forces<br />

62 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

<strong>Designs</strong> Uniform thickness Optimized design<br />

Resulting ratios U0/U umax0/umean0 umax0/umax umax/umean<br />

Design domain 9 × 3 1.18 11.8 14.9 1.006<br />

Design domain 6 × 3 1.27 11.1 14.0 1.006<br />

Design domain 3 × 3 1.54 9.8 13.8 1.006<br />

Table 4.1: Resulting ratios for energy and energy density, corresponding to three<br />

different lengths of rectangular domains, subjected to tension external load. For each<br />

case the results for two designs, which are uniform thickness design and optimized<br />

thickness design.<br />

4.4 Beam-bars: long, medium and short<br />

Figure 4.5 shows the results for a rectangular domain, which is fixed at the left end<br />

and loaded at the free short end, only at the mid third part. The load is in the length<br />

direction and in the transverse direction, i.e. relative to the previous case we have<br />

added bending, with equal magnitudes in the two directions.<br />

Bending has the dominating influence and for this case it is advantageous to<br />

use the full width of the design domain. We compare again the uniform thickness<br />

solution with the optimized thickness design. For shorter design domains the results<br />

are presented in figure 4.6, and all the numerical obtained ratios are contained in<br />

table 4.2. The illustrations in figures 4.5 and 4.6 are arranged in the same manner as<br />

used for the bars.<br />

<strong>Designs</strong> Uniform thickness Optimized design<br />

Resulting ratios U0/U umax0/umean0 umax0/umax umax/umean<br />

Design domain 9 × 3 2.15 10.8 7.1 3.3<br />

Design domain 6 × 3 1.87 11.2 5.9 5.9<br />

Design domain 3 × 3 1.60 10.4 3.5 4.8<br />

Table 4.2: Resulting ratios for energy and energy density, corresponding to three<br />

different lengths of rectangular domains, subjected to combined tension and bending<br />

external load, i.e. a non-symmetric case. For each case the results for two designs,<br />

which are uniform thickness design and optimized thickness design.<br />

We notice that the limitation of tmax =10tmean implies that even for the optimized<br />

designs we get an energy concentration (ratios 3.3, 5.9 and 4.8), and in general<br />

depending upon the actual finite element modelling, because of the high gradients of<br />

thicknesses.


Continua of uniform energy density 63<br />

Figure 4.5: A beam-bar with 3 × 9 rectangular design domain. Top: resulting<br />

optimized thickness design. Mid: corresponding resulting field of energy densities<br />

(almost uniform). Bottom: the field of energy densities for a uniform thickness design.<br />

The black ”hatch” is an attempt to show the principal stress directions.


64 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 4.6: Beam-bars with 3×3 and 3×6 rectangular design domain. Top: resulting<br />

optimized thickness designs. Mid: corresponding resulting fields of energy densities<br />

(almost uniform). Bottom: the fields of energy densities for uniform thickness design.<br />

The black ”hatch” is an attempt to show the principal stress directions.


Continua of uniform energy density 65<br />

4.5 Beams: long, medium and short<br />

Figure 4.7 shows the results for a rectangular domain, which is fixed at the left end<br />

and loaded at the free short end, only at the mid third part. The load is in the Only bending<br />

forces<br />

transverse direction as a bending load. As for the two previous cases the results<br />

for shorter design domains are also shown, here in figure 4.8. The illustrations in<br />

figures 4.7 and 4.8 are arranged in the same manner as used for the bars and for the<br />

beam-bars.<br />

Table 4.3 gives the ratios for the corresponding numerical results. The similarities<br />

in these numbers with the combined tension bending case are striking, although the<br />

optimized designs are quite different, being symmetric for the pure bending case.<br />

<strong>Designs</strong> Uniform thickness Optimized design<br />

Resulting ratios U0/U umax0/umean0 umax0/umax umax/umean<br />

Design domain 9 × 3 2.12 10.1 3.8 3.1<br />

Design domain 6 × 3 1.80 10.5 5.5 3.4<br />

Design domain 3 × 3 1.40 9.9 6.9 3.6<br />

Table 4.3: Resulting ratios for energy and energy density, corresponding to three<br />

different lengths of rectangular domains, subjected to bending external load. For each<br />

case the results for two designs, which are uniform thickness design and optimized<br />

thickness design.<br />

4.6 ”Bridge”s with different load distributions and<br />

supports<br />

Nine<br />

problems<br />

Figure 4.9 shows the results for a model of half a continuum bridge, i.e. a rectangular<br />

domain with one node supported at the left boundary and all the nodes at the right<br />

boundary supported to model the symmetry. The design depends on the position of<br />

the left supported node, in figure 4.9 at the top. The design also depends on the<br />

distribution of the load.<br />

We analyse three load cases, corresponding to: a) uniform loads at the full bottom<br />

boundary, b) uniform loads at only the mid half of the bottom boundary, and c)<br />

uniform loads at only the mid eights of the bottom boundary. We also analyse three<br />

different support cases at the left boundary, which are a single point support at the<br />

lower part, at the mid, and at the upper part, respectively. Totally nine cases, for<br />

which resulting numerical ratios are presented in table 4.4. Three<br />

figures


66 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 4.7: A beam with 3 × 9 rectangular design domain. Top: resulting optimized<br />

thickness design. Mid: corresponding resulting field of energy densities (almost uniform).<br />

Bottom: the field of energy densities for a uniform thickness design. The black<br />

”hatch” is an attempt to show the principal stress directions.


Continua of uniform energy density 67<br />

Figure 4.8: Beams with 3 × 3 and 3 × 6 rectangular design domain. Top: resulting<br />

optimized thickness designs. Mid: corresponding resulting fields of energy densities<br />

(almost uniform). Bottom: the fields of energy densities for uniform thickness design.<br />

The black ”hatch” is an attempt to show the principal stress directions.


68 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

The results of three designs are shown in the three figures 4.9, 4.10, and 4.11. In<br />

figure 4.9 corresponding to upper support and full load, in figure 4.10 mid support<br />

and load at only the half center of the bridge, and in figure 4.11 lower support and<br />

load at only an eights of the bottom boundary, concentrated on the center of the<br />

bridge.<br />

<strong>Designs</strong> Uniform thickness Optimized design<br />

Resulting ratios U0/U umax0/umean0 umax0/umax umax/umean<br />

Lower full load 2.12 10.1 6.9 3.1<br />

support half load 1.80 15.7 5.5 5.1<br />

eights load 1.40 29.8 3.8 10.9<br />

Mid full load 2.24 223.2 76.1 6.6<br />

support half load 2.19 120.0 62.1 4.2<br />

eights load 2.15 100.2 56.0 3.9<br />

Upper full load 2.23 224.0 75.5 6.6<br />

support half load 2.19 120.0 62.1 4.2<br />

eights load 2.05 23.0 12.0 3.9<br />

Table 4.4: Resulting ratios for energy and energy density, corresponding to nine<br />

different support and load cases of a bridge. For each case the results for two designs,<br />

which are uniform thickness design and optimized thickness design.<br />

Especially the energy concentration values in table 4.4 depend on the chosen finite<br />

element model, in particular due to the single point support, and also depend on the<br />

chosen tmax =10tmean. The modelling is the same for all nine cases, and thus the<br />

relative ratios give information.<br />

We notice that the improvement in compliance is almost a factor 2 (first number<br />

column), with less gained for the more concentrated load and lower support. Energy<br />

concentration is for the optimized designs brought down to values from 3 to 11 (last<br />

number column). The energy concentration factors for the initial uniform thickness<br />

design is described in second and third column, but as mentioned these values are<br />

related to the very detailed modelling (25000 elements) and the actual point supports.


Continua of uniform energy density 69<br />

Figure 4.9: A bridge model with upper left support and uniform loads. Top: resulting<br />

optimized thickness designs with red (dark) inserts to clarify the thickness design.<br />

Mid: corresponding resulting field of energy densities (almost uniform) with rough<br />

illustration of principle stress directions. Bottom: the field of energy densities for<br />

uniform thickness design.


70 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 4.10: A bridge model with mid left support and concentrated loads on mid<br />

half of the span. Top: resulting optimized thickness designs with red (dark) inserts<br />

to clarify the thickness design. Mid: corresponding resulting field of energy densities<br />

(almost uniform) with rough illustration of principle stress directions. Bottom: the<br />

field of energy densities for uniform thickness design.


Continua of uniform energy density 71<br />

Figure 4.11: A bridge model with lower left support and rather concentrated loads<br />

on mid eights of the span. Top: resulting optimized thickness designs with red (dark)<br />

inserts to clarify the thickness design. Mid: corresponding resulting field of energy<br />

densities (almost uniform) with rough illustration of principle stress directions. Bottom:<br />

the field of energy densities for uniform thickness design.


Approximately<br />

optimal hole<br />

72 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

4.7 A ”knee” domain<br />

Figure 4.12 shows as for the other examples at the top the optimal thickness design,<br />

at the mid the obtained almost uniform energy density field, and at the bottom the<br />

energy density field for the design with uniform thickness. In addition to the isolines<br />

we here and in the remaining examples add a full colour (gray scale) display.<br />

By the thickness optimization, we have improved the stiffness of the continuum<br />

as given by the ratio U0/U =2.52. Furthermore, the strength is much improved with<br />

specific ratios being umax0/umax =20.0 and umax/umean =2.15 as alternative to<br />

umax0/umean0 =9.33<br />

Completely uniform energy density is not obtained with an active design constraint<br />

of tmax/tmean = 10. This maximum thickness is reached along the inner<br />

curved boundary and except for the non-effective corner also at the outer straight<br />

sides.<br />

4.8 Biaxially loaded hole<br />

Figure 4.13 shows the results of optimizing the thickness field close to a hole that<br />

is biaxially stressed. A few initial comments are necessary before these results are<br />

discussed.<br />

The shape of the hole is described by an ellipse with ratio for the axes equal to the<br />

ratio of the given external stresses, in this case 2:1. As will be discussed in chapter 7,<br />

specifically devoted to shape design, this is the optimal shape for a very small hole (a<br />

hole in an infinite domain), but for a finite domain, as the present, this shape is only<br />

approximately optimal. Anyhow, the stress field, even for the uniform thickness, is<br />

expected to be rather uniform along the boundary of the hole. This is illustrated in<br />

bottom of figure 4.13.<br />

However, by thickness design we can obtain almost uniform energy density in the<br />

total domain, as illustrated in the mid figure. By the thickness optimization, we have<br />

improved the stiffness of the continuum only little (because of the almost optimal<br />

hole shape). The improvement is specified by the ratio U0/U =1.02. However, the<br />

strength is improved with specific ratios being umax0/umax =2.47 and umax/umean =<br />

1.06 as alternative to umax0/umean0 =2.26.<br />

Completely uniform energy density is not obtained with an active design constraint<br />

of tmax/tmean = 10. The optimal design in the figure top shows a clear<br />

reinforcement around the elliptic boundary shape with maximum thickness at the<br />

lower part, where the larger stresses must be released.


Continua of uniform energy density 73<br />

Figure 4.12: A knee model with bending external stress. Top: the optimal thickness<br />

design. Mid: resulting uniform energy density field. Bottom: Starting energy density<br />

field corresponding to uniform thickness design.


74 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 4.13: A hole model with biaxially external stress. Top: the optimal thickness<br />

design. Mid: resulting uniform energy density field. Bottom: Starting energy density<br />

field corresponding to uniform thickness design.


Continua of uniform energy density 75<br />

4.9 A foundation problem<br />

In order to relate to the bar problems studied initially, we here see how a uniform<br />

stress load can be distributed in a kind of foundation, modelled by only half due to<br />

symmetry. Figure 4.14 shows at the bottom the very non-uniform energy density<br />

field for a homogeneous foundation (uniform thickness design). The characteristic<br />

ratio is umax0/umean0 =26.4, which for the optimal thickness design is brought down<br />

to umax/umean =1.3<br />

The improvement of strength for the foundation can then be measured by umax0/umax =<br />

11.78<br />

Homogeneous<br />

nonhomogeneous<br />

4.10 Uniformly loaded, non-rectangular cantilever<br />

Non-rectangular<br />

cantilever<br />

The cantilever used as a generic problem in figure 4.1 is finally optimized, and within<br />

the design constraint of tmax/tmean = 10 we can obtain umax/umean =1.10, i.e. in<br />

reality uniform energy density.<br />

The results are presented in figure 4.15, where the top figure shows a kind of I<br />

beam design that gets smaller toward the free tip. The green inserts illustrate the<br />

thickness design. The mid figure shows the uniform energy density field, while the<br />

bottom figure shows the energy density field corresponding to the uniform thickness<br />

design.<br />

The improvement in stiffness is described by the obtained ratio U0/U =1.86, and<br />

the improvement in strength is characterized by umax0/umax =38.5.<br />

4.11 The numerical procedure<br />

As in general, optimal designs are obtained through a number of iterations. In each<br />

iteration a finite element analysis is performed and based on the information from this<br />

analysis (element energy densities) a better design is suggested. The simple recursive<br />

formula in agreement with optimality criteria methods is an update of element<br />

thicknesses by<br />

0.8 <br />

(ue)actual<br />

¯V<br />

<br />

(te)new =(te)actual<br />

(4.3)<br />

ūactual V<br />

with proper adjustment to the limits tmin and tmax and iteratively updating the<br />

actual volume V , so that the volume constraint V = ¯ V is satisfied in each redesign.<br />

Redesign<br />

iterations


76 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 4.14: Half of a foundation model with short uniform external line load (left<br />

lower corner). Top: the optimal thickness design. Mid: resulting uniform energy density<br />

field. Bottom: Starting energy density field corresponding to uniform thickness<br />

design.


Continua of uniform energy density 77<br />

Figure 4.15: A cantilever model with uniform external line load. Top: the optimal<br />

thickness design with green (gray) inserts to illustrate the thickness design. Mid:<br />

resulting uniform energy density field. Bottom: Starting energy density field corresponding<br />

to uniform thickness design


Modelling<br />

accuracy<br />

78 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

The relaxation power of 0.8 is used by most researchers but naturally it is not a<br />

physical constant. However without this relaxation, design cycling is often observed<br />

while with the relaxation power monotonic convergence is often obtained.<br />

To find the optimized designs shown we have only used about 10 design iteration<br />

for each, but this number is strongly related to the actual convergence test.<br />

4.12 Summing up<br />

We end this chapters with the major conclusions about the influence from maximum<br />

thickness constraint, the direct relation to free material design and design with alternative<br />

materials, and finally about the specific examples dealt with in this chapter.<br />

4.12.1 Influence from the maximum thickness constraint<br />

As mentioned before we might have used mass density or relative volume density ρe<br />

for each element as design variables, and then naturally ρmax = 1 would be the upper<br />

side constraint. With thicknesses te as design variables the upper side constraint must<br />

be set from practical point of view (say manufacturing possibilities) and/or estimated<br />

from the reliability of the finite element analysis of a continuum with large slopes in<br />

the thickness field.<br />

In some of the examples treated we end up with designs that call for a kind<br />

of stiffener enforcement. Then the results for the optimized designs depend on the<br />

chosen upper side constraint tmax, and a completely uniform energy density field is<br />

not obtained. The measure of the maximum energy density relative to the mean<br />

energy density umax/umean gives information about this situation.<br />

4.12.2 Relation to free material design<br />

All the examples are optimized for an isotropic material with zero Poisson’s ratio.<br />

This means that the stress field and the strain field are equal, except for the constant<br />

E, the modulus of elasticity (see chapter 11). The energy density ue in element e is<br />

therefore given simply by<br />

ue =(σ 2 1 + σ 2 2)e/E =(ɛ 2 1 + ɛ 2 2)eE (4.4)<br />

with the two principal stresses (σ1,σ2)e, and the two principal strains (ɛ1,ɛ2)e.<br />

Behind this choice of material description is the fact, that the ultimately optimal<br />

continua are then obtained by using for each element (in reality each point) the<br />

constitutive parameters<br />

C1111 = Eɛ 2 1,C2222 = Eɛ 2 2,C1122 = Eɛ1ɛ2, C1212 = 0 (4.5)


Continua of uniform energy density 79<br />

as described in detail in chapter 15.<br />

4.12.3 Relation to non-linear and non-isotropic material<br />

Although we have here focused on these extreme materials, the optimization procedure<br />

would be the same for non-zero Poisson’s ratio, isotropic material and the resulting<br />

optimized designs are not too different.<br />

For non-isotropic materials we may combine with optimal material orientation as<br />

described in chapter 17. Even for non-linear elastic materials, described with a power<br />

law, the procedure is valid as described in chapter 12. So, the examples do by no<br />

means indicate actual restrictions.<br />

4.12.4 The specific examples<br />

The examples in sections 4.3, 4.4, and 4.5 show the influence from rather concentrated<br />

external loads. A single force on a continuum results in a singularity and the finite<br />

element analysis of this depends on the specific modelling. We have chosen to apply<br />

the external load over a finite domain, but still only part of the actual external<br />

boundary. Away from loads a more homogeneous state is obtained and we notice this<br />

for the longer rectangular domains.<br />

This is reflected in the optimized designs for tension, for bending and for combined<br />

tension bending. For the squared domain the whole continuum is in a kind of<br />

transition zone. For the tension case we have compared with a narrowed domain that<br />

is externally loaded over the full boundary. This idealized situation is in reality questionable<br />

at the support, and the optimized designs based on the full domain agree<br />

in stiffness and in strength within a less than a percent. The cases which include<br />

bending result in designs comparable to I-beam and the maximum thickness then has<br />

a strong influence on the obtained results.<br />

In section 4.6 we optimized continua that transfer loads to two end supports,<br />

and in reality these also act like beams. We find again I-beam solutions with end<br />

modifications near the supports The examples include different load distributions<br />

and different positions of the end support. Again the maximum thickness constraint<br />

influences the obtained results.<br />

Bending is also active in the ”knee” problem where stress concentration also is<br />

active around a curved inner boundary. Large domains of minimum thickness result,<br />

indicating that shape design at inner holes as well as at outer boundaries would be<br />

valuable. This leads to the hole problem optimized in section 4.8, which for a biaxially<br />

loaded rectangular domain is based on an assumed almost optimal shape of the hole<br />

boundary. As the hole has finite size in the domain, the shape of the hole is not


80 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

completely optimal, and we find that the optimized thickness design gives different<br />

enforcement along the hole boundary.<br />

To illustrate the complexity in transferring loads to a foundation, we optimize<br />

a simple foundation problem in section 4.9. It is interesting to see the resulting<br />

combination of areas of enlarged stiffness (large thickness) with areas of enlarged<br />

flexibility (small thickness) which makes it possible to obtain almost uniform energy<br />

density.


Chapter 5<br />

Design of beams and frames<br />

Bars are one dimensional structural elements from which we can construct two- and<br />

three-dimensional trusses; beams are one-dimensional structural elements from which Structures<br />

we can construct two- and three-dimensional frames, including grids, curved beams<br />

and non-uniform beams.<br />

In bars the cross-sectional points are in a uniform state of stress, strain, and energy<br />

density, but this is not the case for beams. In beams and frames the aspect of bending<br />

is dominating, and point-wise uniform energy density can not be obtained. However,<br />

optimality criteria expressed in terms of finite design regions are still applicable,<br />

of 1D-elements<br />

and a number of important optimal design problems can be solved, if not explicit Finite design<br />

regions<br />

analytically, then implicit analytically.<br />

The early literature related to structural optimization contains a large number<br />

of specific results, which also include two-dimensional plate problems. For general<br />

reference, see the course notes by (Prager 1974) together with the books by (Banichuk<br />

1983) and by (Rozvany 1989). Extended lists of references are available in these Important<br />

books.<br />

In this chapter we, based on the theoretical results presented in chapter 14, show<br />

a very direct way to obtain explicit analytical optimal designs and then indicate the<br />

iterative solution to obtain implicit analytical optimal designs, but without repeating<br />

references<br />

the results found in the two books, just mentioned. Main contents<br />

of chapter<br />

We then concentrate on problems that are solved with application of methods<br />

from mathematical programming. At first a specific example of beam design with<br />

several eigenfrequency constraints is described in detail with focus on the difference<br />

between using simple beam theory and taking shear deformations into account.<br />

An example with a portal frame illustrates the inherent problem with unknown<br />

active constraints when multiple load cases are prescribed. This problem also illus-<br />

81


Focus on<br />

stress energy<br />

<strong>Optimal</strong>ity<br />

criterion b<br />

82 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

trates the influence on the resulting optimal designs from given boundary conditions.<br />

Shape optimization of frames involves positions of beam connections as design<br />

variables. For 3D-frames this important structural design problem is presented in the<br />

final section of this chapter. Analytical sensitivity analysis is still possible, but the<br />

”bookkeeping” is extensive. Two specific optimal designs are shown.<br />

5.1 Explicit analytical optimal designs<br />

For a number of problems explicit analytical optimal designs can be obtained. The<br />

derivation of these results is based on the optimality criterion for problems of minimum<br />

compliance with a given mass (volume), which is stated in chapter 14, section 14.3.1<br />

as: the ratio between sub-region energy and sub-region mass should be the same in all<br />

the design sub-regions.<br />

With a given ratio between strain energy, stress energy, and thus total elastic<br />

energy we may choose what is most convenient for the actual formulation. Here we<br />

choose stress energy. The optimality criterion also holds for non-linear elasticity,<br />

modelled by power law non-linearity.<br />

Taking a part of a beam, say from a to b with length (b − a), where the crosssectional<br />

area A is to be designed (uniform over the length, which may be infinitesimal).<br />

The sub-region stress energy Uσ for pure bending with a simple beam model<br />

and linear elasticity is known to be<br />

Uσ =<br />

b<br />

a<br />

M 2 b<br />

dx =<br />

2EI a<br />

M 2<br />

2Eα2 dx (5.1)<br />

A2 where M is the bending moment, E is the modulus of elasticity, and I is the crosssectional<br />

moment of inertia, for which we assume the simple relation I = α 2 A 2 to be<br />

valid with the same cross-sectional constant α as used in chapter 3.<br />

With ρ as a given mass density, the mass m for the design sub-region is m =<br />

b<br />

a ρAdx and the ratio to be constant according to the optimality criterion (dUσ/dA) =<br />

˜C(dm/dA), see chapter 14, is thus<br />

a<br />

A = C<br />

2 M<br />

d<br />

2Eα 2 A 2<br />

b<br />

a M 2 dx<br />

b − a<br />

<br />

/dA dx = ˜ C<br />

1/3<br />

b<br />

a<br />

(d(ρA)/dA) dx ⇒<br />

⇒ A = CM 2/3 for (b − a) → 0 (5.2)<br />

where ˜ C and C are constants when 2Eα 2 ρ is assumed constant too. The constant C


Design of beams and frames 83<br />

is determined by the given total volume ¯ V , i.e.<br />

¯V =<br />

L<br />

0<br />

Adx =<br />

L<br />

0<br />

CM 2/3 ¯V<br />

dx ⇒ C = L<br />

0 M 2/3dx (5.3)<br />

where L is the total length of the beam (or alternatively with the summation over<br />

design sub-regions).<br />

5.1.1 Statically determined elementary cases<br />

A number of elementary cases such as cantilever beams (or frames) and simply supported<br />

beams are statically determined, which means that the moment distribution<br />

is determined independently of the design, and then (5.3) directly gives explicit Known<br />

analytical optimal designs.<br />

26% energy<br />

39% energy<br />

65% energy<br />

100% energy<br />

Figure 5.1: Left: <strong>Optimal</strong> designs with obtained compliance. Right: the corresponding<br />

cantilever elementary load cases.<br />

Figure 5.1 shows the optimal designs for cantilever cases where the moment distribution<br />

is cubic dependent on x, quadratic dependent on x, linear dependent on x,<br />

and independent on x (constant), respectively. Characterizing these four cases by the<br />

power n, the optimal designs are<br />

x1<br />

A(x) =Cx 2n/3 with C determined by C<br />

L<br />

0<br />

x1<br />

x 2n/3 dx = ¯ V (5.4)<br />

where L is the length of the actual beam. As expected, the best improvement can be<br />

obtained for the triangular load distribution (almost a factor 4), and the pure moment<br />

moment<br />

distribution<br />

load for which the uniform beam is the optimal design (no improvements possible). Alternative<br />

The assumption of I = α 2 A 2 may not always be applicable, and alternatively we<br />

look at the case of a rectangular cross-section with cross-sectional moment of inertia<br />

design<br />

variables


Explicit<br />

optimal<br />

beam design<br />

<strong>Optimal</strong><br />

energy ratio<br />

84 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

I = bh 3 /12 and area A = bh, where the width b or the height h might be the actual<br />

design parameters. In terms of width b and height h, the sub-region energy and<br />

sub-region mass are<br />

Uσ =<br />

2 M<br />

dx =<br />

2EI<br />

2 M<br />

2Eh3 dx and m = hbdx (5.5)<br />

b/12<br />

with resulting optimality conditions for the two alternative optimality criteria<br />

M 2 /b 2 = constant for given h<br />

M 2 /h 4 = constant for given b (5.6)<br />

Thus, the only difference relative to the solution (5.4) is that the power 2n/3 must be<br />

changed either to 2n/2 =n or to 2n/4 =n/2, as seen by comparing the optimality<br />

criterion (5.3) with (5.6).<br />

In general for all the cases mentioned in this section, the optimal solution is of<br />

the type d(x) =Cx q , where d(x) is width, height, radius, or area at position x from<br />

the free tip of the cantilever. The power q depends on the load case as well as on the<br />

actual design parameter.<br />

The percentages in figure 5.1 are determined as follows directly from (5.1). For<br />

an uniform beam with moment distribution M(x) = ˜ Mx n and area Ā = ¯ V/L the<br />

total stress energy Ūσ is<br />

Ūσ =<br />

L<br />

0<br />

˜M 2 x 2n<br />

2Eα 2 ¯ V 2 /L 2 dx = ˜ M 2 L 2<br />

2Eα 2 ¯ V 2<br />

L<br />

0<br />

x 2n dx (5.7)<br />

For an optimal beam with the same moment distribution and area distribution A =<br />

¯V<br />

L<br />

0 x2n/3 dx x2n/3 the total stress energy Uσ is<br />

Uσ = ˜ M 2 ( L<br />

0 x2n/3dx) 2<br />

2Eα2 ¯ V 2<br />

L<br />

0<br />

which together with (5.7) gives the ratio<br />

Uσ<br />

Ūσ<br />

¯V<br />

=<br />

3<br />

C3L2 L<br />

0 x2ndx = (2n +1)¯ V 3<br />

C3L2n+3 x2n dx (5.8)<br />

x4n/3 (5.9)


Design of beams and frames 85<br />

5.2 Implicit ”analytical” optimal designs<br />

Two aspects prevent explicit analytically optimal designs, but implicit solutions are<br />

then obtained by a few iterations. The first aspect is related to size constraints, due to<br />

the missing knowledge about the regions where these size constraints are active. The<br />

second aspect concerns statically indetermined cases, where the moment distribution<br />

is depending on the actual design. Often both aspects must be taken into account, Simple<br />

which can be done simultaneously, but in the description here we separate them. iterations<br />

5.2.1 Side constraints<br />

We notice in the optimal solutions of figure 5.1, that when the moment in a beam crosssection<br />

is zero, the optimal cross-sectional area is also zero. To obtain more practical<br />

solutions we impose a minimum area constraint, say Amin =0.1Amean =0.1 ¯ V/L.<br />

The location of design regions where this constant is active is found by iteration<br />

and the optimal design in the remaining regions is described by (5.4), which we use<br />

as an example. In (5.4) the given volume ¯ V is substituted by the ”active” volume Active<br />

V = volume<br />

¯ V − Vconstraint of free design.<br />

5.2.2 Statically indetermined cases<br />

With statically indetermined boundary conditions, we can also use iterations to find<br />

the optimal design, very much similar to the procedure for obtaining the designs<br />

of uniform energy density in chapter 4, section ??. For multiple load cases and with<br />

different constraints we can not expect the optimal design to be statically determined.<br />

For a number of specific examples see (Rozvany 1989).<br />

5.3 Beam design with eigenfrequency constraints<br />

5.3.1 Sensitivity analysis for eigenfrequencies<br />

In chapter 18 the sensitivity analysis, for dynamic problems in general, is described,<br />

so here we concentrate on the actual beam problem, treating this as a finite element<br />

model with a limited number of beam elements. The finite element formulation for<br />

the structural eigenfrequencies is Problem of<br />

analysis<br />

[S]{D} = ω 2 [M]{D} (5.10)<br />

where ω 2 and {D} are, respectively, the squared eigenfrequency and the corresponding<br />

eigenvector. The total stiffness matrix [S] and the total mass matrix [M] are both


Sensitivity<br />

result<br />

Sensitivity for<br />

simple beams<br />

86 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

obtained by summation of the individual beam element matrices<br />

[S] = <br />

[Se] and [M] = <br />

[Me] (5.11)<br />

e<br />

This origin of the system matrices is essential.<br />

Inverse iteration as described in (Bathe 1982) (and later editions), including shift,<br />

sub-space iteration, and Sturm sequence check, provides a powerful analysis tool<br />

which allows the user to focus on an actual frequency domain and control the rate of<br />

convergence. Close or even equal eigenfrequencies are also taken care of.<br />

Assuming that the eigenpair ω 2 , {D} is known, the important point to note, as<br />

proven in chapter 18, is that this also contains the information necessary to determine<br />

the eigenfrequency sensitivities, dω 2 /dae, i.e. the change in eigenfrequency due to a<br />

change in the area of beam element e. The result, derived in chapter 18 and given in<br />

(18.42) is (with the normalization condition {D} T [M]{D} =1)<br />

dω2 = {D}<br />

dae<br />

T ( d[S] 2 d[M]<br />

− ω ){D} (5.12)<br />

dae dae<br />

With ae being a local design parameter, this result with the assumed normalization<br />

can be calculated on the element level with {De} contained in {D}<br />

dω2 = {De}<br />

dae<br />

T ( d[Se] 2 ∂[Me]<br />

− ω ){De} (5.13)<br />

∂ae ∂ae<br />

For simple beam theory we have d[Se]/dae =2[Se]/ae and d[Me]/dae =[Me]/ae, and<br />

therefore get a result which can be described in element elastic and element kinetic<br />

energies Ue,Te.<br />

dω2 dae<br />

= {De} T (2[Se] − ω 2 [Me]){De}/ae =(2Ue − ω 2 Te)/ae<br />

e<br />

(5.14)<br />

With other cross-sectional design parameters than area, the only difference, relative<br />

to (5.14), is another factor than 2 in front of Ue and a factor also in front of Te,<br />

see section 18.3.2.<br />

More complicated modifications of the sensitivity results are necessary when<br />

higher order beam theories are used. They are discussed in detail in section 5.3.5.


Design of beams and frames 87<br />

5.3.2 <strong>Optimal</strong> design with one eigenfrequency constraint<br />

Let us first consider a simply supported beam whose volume we want to minimize<br />

subject to the constraints that some of the eigenfrequencies must be greater than or<br />

equal to those of the corresponding uniform beam. We choose a slender beam, and<br />

thus the gradients (5.14) are valid. Putting constraint only on the first eigenfrequency<br />

we obtain the design in figure 5.2. This result agrees well with the early result of<br />

(Niordson 1965), although we do not reach zero area at the support when shear Classical<br />

deformations are taken into account.<br />

problem<br />

ω1 optimal = ω1 uniform<br />

ω2 optimal =0.928 ω2 uniform<br />

ω3 optimal =0.910 ω3 uniform<br />

ω4 optimal =0.902 ω4 uniform<br />

Voptimal =0.883 Vuniform<br />

Figure 5.2: <strong>Optimal</strong> design of slender, simply supported beam for minimum volume<br />

with the constraint put only on the first eigenfrequency.<br />

The volume is through the optimization minimized to 88.3% of that for the uniform<br />

beam, and as expected the second, third and fourth eigenfrequencies are lowered.<br />

5.3.3 <strong>Optimal</strong> design with several eigenfrequency constraints<br />

Extended<br />

Controlling also the second eigenfrequency, then the first three eigenfrequencies and problems<br />

finally all four eigenfrequencies, we obtain the optimal designs shown in figure 5.3.<br />

Solutions for other boundary conditions than simply supported, like clamped<br />

- simply supported or flexible clamped - simply supported, and also for different<br />

medium slender beams can be found in (Pedersen 1981).<br />

5.3.4 <strong>Optimal</strong>ity criterion with only a single constraint<br />

With only a single constraint the general optimality criterion of proportional gradients<br />

(chapter 14 (14.4)) gives for the actual problem<br />

dω2 dae<br />

= C dV<br />

= Cle<br />

dae<br />

(5.15)


88 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

ω1 optimal = ω1 uniform<br />

ω2 optimal = ω2 uniform<br />

ω3 optimal =0.959 ω3 uniform<br />

ω4 optimal =0.952 ω4 uniform<br />

ω1 optimal = ω1 uniform<br />

ω2 optimal = ω2 uniform<br />

ω3 optimal = ω3 uniform<br />

ω4 optimal =0.972 ω4 uniform<br />

ω1 optimal = ω1 uniform<br />

ω2 optimal = ω2 uniform<br />

ω3 optimal = ω3 uniform<br />

ω4 optimal = ω4 uniform<br />

Voptimal =0.934 Vuniform<br />

Voptimal =0.955 Vuniform<br />

Voptimal =0.967 Vuniform<br />

Figure 5.3: <strong>Optimal</strong> design of slender, simply supported beams for minimum volume<br />

with constraint on several eigenfrequencies. Top: constraints on the first two eigenfrequencies.<br />

Mid: constraints on the first three eigenfrequencies. Bottom: constraints<br />

on the first four eigenfrequencies.


Design of beams and frames 89<br />

where the constant C (Lagrange multiplier) is the same for all beam elements e. The<br />

element lengths le need not be equal. Substituting the simple gradient (5.14), we get<br />

2Ue − ω 2 Te = Caele ⇒ 2ue − ω 2 te = C (5.16)<br />

where ue,ω2te are the elastic energy density and the kinetic energy density, respectively.<br />

By summation (integration) we can then obtain the value of the constant C<br />

<br />

2Ue − <br />

ω 2 Te = C <br />

aele = CV (5.17)<br />

e<br />

e<br />

e<br />

and with normalization <br />

e Te =1→ <br />

e 2Ue =2ω 2 we have for the optimal design Value of<br />

multiplier<br />

C = ω 2 /V (5.18)<br />

For the results of this section we have used mathematical programming as the iteration<br />

tool, but with only a single constraint, optimality criterion iterations based on (5.16)<br />

with estimates of C from (5.18) would also be possible. Test of<br />

solutions<br />

Even though the optimality criterion is not used for finding the optimal design,<br />

it is valuable to test whether an optimal design fulfills this criterion, i.e. if it is valid.<br />

We do this in the examples to follow.<br />

5.3.5 Sensitivity analysis for Timoshenko beam theory<br />

The numerical analysis of beams and frames is often based on Timoshenko beam<br />

theory and although it is possible still to use the simplified results for sensitivity<br />

analysis (5.14), we present the extended result, as derived in (Pedersen 1982–83). Exact<br />

The result is<br />

gradients<br />

<br />

βe<br />

1+βe<br />

dω 2<br />

dae<br />

= 1<br />

ae ({De} T (2[Se] − ω 2 [Me]){De} +<br />

{De} T (2[Se] − ω 2 [Me]){De}− EIe<br />

le (θ1 − θ2) 2 e<br />

<br />

−<br />

ω 2 {De} T [ ˜ Me]){De}) (5.19)<br />

where θ1,θ2 are the two end rotations of the beam element. The element coefficient<br />

of shear βe is<br />

βe = 12µEIe<br />

Gael 2 e<br />

(5.20)<br />

where E,G are the material moduli and µ is the cross-sectional constant that follows<br />

from shear distribution. For slender beams we have βe = 0 and thus the same formula<br />

as (5.14). The matrix [ ˜ Me] is a modified mass matrix, that is listed in (Pedersen 1982–<br />

83), together with all the involved matrices.


Influence of<br />

gradients<br />

90 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

5.3.6 Optimization with different gradients<br />

The beam eigenfrequency analyses are in all the present cases based on Timoshenko<br />

beam theory, but it is possible to perform the optimization with the simple gradients<br />

(5.14), here termed approximate gradients or with the more complicated gradients<br />

(5.19), here termed exact gradients.<br />

To study the influence of slenderness and the difference between using exact gradients<br />

(5.19) and approximate gradients (5.14) in the case of a simply supported beam,<br />

we maximize the second eigenfrequency for a given volume, but without constraints on<br />

the first eigenfrequency. For Bernoulli-Euler beams this problem is treated in (Olhoff<br />

1976), and the optimality criterion (5.16) is known to be valid. Choosing a uniform<br />

beam as our initial design, figure 5.4 presents the results for a total slenderness ratio<br />

of 25. Based on these results we conclude:<br />

1. The design obtained by exact gradients is better with the eigenfrequency improved<br />

7.5% as compared to 6.9%, but for larger slenderness ratio this difference<br />

disappears as expected.<br />

2. All the designs agree with the optimality criterion, where (dω2 /dae)exact has<br />

the same value for all beam elements e. Thus, we have designs which differ<br />

only slightly, but which both are stationary.<br />

Figure 5.4: Two optimal designs of a simply supported beam for maximum second<br />

eigenfrequency with given volume. Thin line boundary: initial uniform design and<br />

result with approximate gradients. Heavy line boundary: Result with exact gradients.<br />

Taking the initial design to be uniform (symmetric design), we end up with symmetric<br />

optimal designs, which is to be expected since every re-design problem itself is


Design of beams and frames 91<br />

symmetric. However, taking an initial design with linear tapered thickness, we find<br />

the optimal design in figure 5.5, which are better than the designs in figure 5.4. Again<br />

the results fulfill optimality criterion (5.16). The improvement in eigenfrequency is<br />

now 22.8%, but the first eigenfrequency is then close to zero. Symmetry or<br />

non-symmetry<br />

Figure 5.5: <strong>Optimal</strong> beam design for maximizing the second eigenfrequency with<br />

given volume, and without constraint on symmetry. The first eigenfrequency is not<br />

constrained and exact gradients are used for redesigns.<br />

It is a general feature of optimal design that many local optima exist, and optimization<br />

should therefore be attempted from different initial designs. However, it<br />

should be noted that if we start with a good engineering design, and use the approach<br />

of mathematical programming, then we can not get a less favourable design.<br />

5.4 2D-Frame design with many load cases<br />

Most structures in practice must be designed for several load cases. In relation to the<br />

design problem, this means that a multiplicity of constraints have to be accounted<br />

for, although most of them turn out to be non-critical. Therefore, we cannot choose<br />

a design approach where the critical constraints have to be identified in advance.<br />

The combination of finite element analysis and sequential linear programming<br />

(SLP) constitutes an effective and reliable approach for these practical optimal designs.<br />

We show and discuss the resulting design of a portal frame for a crane, where<br />

the moving load is modelled by a number of single forces, acting independently, i.e.<br />

multiple load cases (actually 15 cases).<br />

Again sensitivity analysis plays a major role and analytical gradients are obtained.<br />

Figure 5.6 shows a model of the portal frame and the constraint of a straight upper<br />

Multiple<br />

load cases


Eccentricity<br />

Reference<br />

stress<br />

92 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

surface means that eccentricities need to be taken into account. (In frame analysis<br />

for ships this is also often the case.) Detail of the involved matrices can be found in<br />

(Pedersen and Jørgensen 1984).<br />

Figure 5.6: Model of a portal frame with one of the many load cases.<br />

5.4.1 Design problem, analysis and sensitivity analysis<br />

The objective of design is to minimize the total mass Φ<br />

Minimize Φ= <br />

cehe with given ce := ρelebe<br />

e<br />

(5.21)<br />

where the heights he are the design variables while lengths le, widths be and mass<br />

densities ρe are given.<br />

The constraints of the problem are specified as a limit on the squared von Mises<br />

stress (Fe)l in all elements e for all load cases l, i.e.<br />

(Fe)l = 1<br />

2 {σe} T l [Z]{σe}l ≤ Fmax<br />

(5.22)<br />

where {σe}l is the stress vector in element e for load case l and [Z] is a symmetric<br />

matrix of integer constants. The specified limit is Fmax. The stress vector we evaluate<br />

from<br />

{σe}l =[Qe]{De}l<br />

(5.23)


Design of beams and frames 93<br />

where [Qe] is the element stress-displacement matrix and with element displacements<br />

{De}l from the system equilibrium of load case l<br />

[S]{D}l = {A}l<br />

(5.24)<br />

where {A}l constitutes the load case l.<br />

The finite element analysis locates the critical (say within 5% from Fmax) stress<br />

elements and corresponding load cases. For these we perform sensitivity analysis by<br />

the chain rule of differentiation Chain rule of<br />

differentiation<br />

d(Fe)l/dhi =[Z]{dσe/dhi}l<br />

{dσe/dhi}l =[dQe/dhi]{De}l +[Qe]{dDe/dhi}l<br />

[S]{dD/dhi}l = {dA/dhi}l − [dS/dhi]{D}l<br />

(5.25)<br />

where [dQe/dhi] =0fore = i. Treating {dA/dhi}l − [dS/dhi]{D}l as a pseudo-load<br />

we get {dD/dhi}l by what is named the semi-analytical approach. Normally this does<br />

not give rise to numerical problems, but for possible errors see (Pedersen, Cheng and Semi-analytical<br />

Rasmussen 1989).<br />

The solution by SLP follows from the general description in chapter 19, and we<br />

can go directly to the results.<br />

5.4.2 Only stress constraints and influence from stiffness of<br />

columns<br />

The portal frame is modelled as a beam with boundary conditions given by the actual<br />

columns, that influence the solution. Three different columns are applied as examples.<br />

sensitivities<br />

Timoshenko beam theory with Poisson’s ratio ν =0.3 and the coefficient of shear Timoshenko<br />

distribution µ = (12 + 11ν)/(10 + 10ν) is the basis. A discretization in 30 uniform,<br />

eccentrically connected elements of equal length is used and symmetry is enforced to<br />

be able only to encounter 15 heights as design variables and 15 load cases, as figure<br />

5.7 shows.<br />

When the bending stiffnesses of the portal columns are large in relation to that of<br />

beam theory<br />

the portal beam, then the portal model corresponds to fixed-fixed beam, and when the Columns<br />

portal columns are very flexible, the portal model corresponds to a simply supported<br />

beam, i.e. a statically determinate structure. The optimal designs for three models<br />

with increasing stiffness of the columns (relative stiffness of 1, 8 and 64) are shown<br />

in figures 5.9, 5.10, and 5.11. In all cases, a uniform beam is taken as initial design,<br />

and the characteristics of optimal design (mass M and squared reference stress F )<br />

are given relative to those of the uniform beam. Only stress constraints are involved<br />

in these solutions, and the allowable Fmax is taken as the maximum value for the<br />

uniform beam.<br />

at beam<br />

boundaries


Initial<br />

designs<br />

94 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

1 2 3 4 5 6 7 8 9101112131415<br />

Figure 5.7: Design model of a portal frame with numbers on the 15 different load<br />

cases.<br />

Note that smooth stress results along the beam length are not to be expected,<br />

because the maximum stresses may refer to different loading cases. By the numbers<br />

(1 - 15) in the figures we have indicated the actual load case, also numbered in figure<br />

5.7.<br />

The optimal designs are not fully stressed, by which is meant that every beam<br />

element is stressed to the allowable level in at least one load case. However, the case<br />

of flexible columns is close to this condition and is discussed specifically.<br />

Figure 5.8: Initial design models used for optimization.<br />

The optimal designs are performed with the three different initial designs, shown<br />

in figure 5.8 and in all cases the same optimal designs are obtained. Other initial<br />

designs have also been tried, and everything points in the direction of the designs<br />

being global optimal solutions, but no proof is available.


Design of beams and frames 95<br />

Figure 5.9 shows the optimal design and the resulting maximum reference stresses<br />

with the flexible columns. A 20% decrease in mass is obtained and the portal beam<br />

is almost fully stressed. A fully stressed design can be found by simple fully stressed<br />

iterations and for this design we find only slightly more mass, with small design<br />

modifications only at the beam ends. Slender<br />

columns<br />

(Fe)l/Fmax<br />

<strong>Optimal</strong> design<br />

Moptimal/Muniform =0.805<br />

with slender columns<br />

Figure 5.9: Top: optimal design of portal beam on slender columns (hcolumns =<br />

huniform/2). Bottom: Relative maximum squared von Mises stress and with numbers<br />

of the corresponding load cases (see figure 5.7). Dotted line for the uniform beam.


Medium<br />

slender<br />

columns<br />

96 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 5.10 shows the optimal design and the resulting maximum reference stresses<br />

with the stiffness of the columns increased with a factor 8. A 14% decrease in mass<br />

is obtained and the portal beam is again almost fully stressed. If we iterate to find a<br />

fully stressed design, then as for the flexible columns it has only slightly more mass,<br />

with small design modifications only at the beam ends.<br />

(Fe)l/Fmax<br />

<strong>Optimal</strong> design<br />

Moptimal/Muniform =0.862<br />

with medium columns<br />

Figure 5.10: Top: optimal design of portal beam on medium columns (hcolumns =<br />

huniform). Bottom: Relative maximum squared von Mises stress and with numbers<br />

of corresponding load cases (see figure 5.7). Dotted line for the uniform beam.


Design of beams and frames 97<br />

Figure 5.11 shows the optimal design and the resulting maximum reference stresses<br />

with large stiffness of the columns. A 12% decrease in mass is obtained and the portal<br />

beam is again almost fully stressed. Non-slender<br />

columns<br />

(Fe)l/Fmax<br />

<strong>Optimal</strong> design<br />

Moptimal/Muniform =0.881<br />

with heavy columns<br />

Figure 5.11: Top: optimal design of portal beam on heavy columns (hcolumns =<br />

2huniform). Bottom: Relative maximum squared von Mises stress and with numbers<br />

of corresponding load cases (see figure 5.7). Dotted line for the uniform beam.


Further<br />

designs<br />

Size and shape<br />

with linking<br />

98 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

5.4.3 Including displacement constraints<br />

The presented optimal designs give a maximum displacement that is close to 20%<br />

larger than for the uniform beam, and severe design modifications are necessary if<br />

we add a constraint on the maximum displacement, equal to the displacement before<br />

optimization. Still a save in mass of 16% is possible also for the case in figure 5.9.<br />

For the case in figure 5.10 the save in total mass is 11% as compared to 14% without<br />

displacement constraints, and for the case in figure 5.11 the corresponding percentages<br />

are 9% instead of 12%. The changed optimal designs can be found in (Pedersen and<br />

Jørgensen 1984).<br />

A common conclusion from this portal frame problem is that displacement constraints<br />

change the optimal design, but the structure generally has the ability to<br />

adjust to the additional constraint with only a slight increase in mass.<br />

The optimization of the portal frame is based on Timoshenko beam theory and<br />

analytical sensitivity analysis is obtained. For slenderness ratio lower than about<br />

20 we see a clear difference between optimal designs based on Bernoulli-Euler beam<br />

theory and optimal designs based on Timoshenko beam theory.<br />

5.5 3D-Frame design with optimal joint positions<br />

This section mainly refers to the paper (Sergeyev and Pedersen 1996). The experience<br />

of optimal design shows that the shape of a construction is a fundamental factor for<br />

its mechanical characteristics. We thus here see the optimization of joint positions<br />

as our primary interest, but naturally the beam optimization plays the role as an<br />

important sub-problem, which with non-trivial beam cross-sections may be rather<br />

complicated.<br />

Linking design variables into groups we can take constructive requirements into<br />

consideration. Furthermore, we may then decrease the number of load cases and<br />

naturally we also decrease the total number of design variables. Constraints on stresses<br />

have to account for stresses resulting from tension, bending, shear and torsion with a<br />

definition of an effective stress measure.<br />

At first the sensitivity of this stress measure to joint positions may seem very<br />

complicated, but in fact the analytical sensitivity analysis is rather straight forward.<br />

We prefer this analytical sensitivity analysis not only because of its effectiveness with<br />

respect to computer time, but also for its robustness.<br />

Firstly, a problem which was earlier solved in a truss formulation (Pedersen 1973)<br />

is treated. This problem includes 52 beams and several load cases. Solutions with<br />

different assumptions are compared and the influence of including self-weights is illustrated.<br />

Then, a portal frame as defined in (Apostol, Santos and Goia 1995) is<br />

optimized, again studying the influence from the basic assumptions.


Design of beams and frames 99<br />

All solutions are found using a sequential quadratic programming algorithm, with<br />

a possibility for including move-limits.<br />

5.5.1 Design problem, analysis and sensitivity analysis<br />

From a formulation point of view the present problem is almost like the 2D-frame<br />

problem in section 5.4. The objective of design is to minimize the total mass Φ<br />

Minimize Φ= <br />

ρeaele = <br />

e<br />

k<br />

ρkak<br />

<br />

e<br />

le<br />

<br />

k<br />

(5.26)<br />

where the design variables are cross-sectional areas ae, and lengths le indirectly<br />

through the joint positions xj, while densities ρk are assumed given. By index k<br />

we denote a group of beams in order to get a formulation suited for linking.<br />

All our design variables are scalar quantities contained in a design vector {h} that<br />

includes cross-sectional parameters and joint coordinates. The influences of these two<br />

different groups of design parameters are very different, but for the formulation it<br />

is an advantage to treat them together. We may have side constraints on the size<br />

parameters (say thickness of beam cross-sections) as well as on the shape parameters<br />

(say a joint constrained to be in a certain domain). These side constraints are specified<br />

by Side<br />

{h}min ≤{h} ≤{h}max<br />

(5.27)<br />

For the iteration strategy it is often advantageous to specify move-limits<br />

{∆hp}min ≤{∆hp} ≤{∆hp}max<br />

(5.28)<br />

constraints<br />

indicating by index p that these may change from iteration step p to another iteration<br />

step. Note that the absolute side constraints (5.27) can be included in the move-limit<br />

constraints (5.28).<br />

The more physically based constraints relate to stresses and displacements, i.e. to<br />

strength and stiffness. As the formulation is based on multiple load cases we have to<br />

add an index l indicating the load case and for the state of strength write constraints<br />

like (5.22) Strength<br />

(Fe)l ≤ Fmax for all elements e and all load cases l (5.29) constraints<br />

The number of constraints is given by the ”nearby” critical cases. Thus, for every<br />

beam there may be several critical points and each point may in fact be critical in<br />

several load cases. Furthermore, during design iteration the critical cases may change<br />

and in most problems the number of constraints increases with the design evolution.


Stiffness<br />

100 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Another group of design constraints are related to the stiffness of the frame, as<br />

specified by the resulting displacements {D} of the joints<br />

{D}min ≤{D}l ≤{D}max for all load cases l (5.30)<br />

with interpretation again in terms of active set strategy.<br />

constraints The detail of analysis and of sensitivity analysis are given in (Sergeyev and Pedersen<br />

1996), and especially the sensitivities to change in joint positions are rather<br />

lengthy expressions, that should be tested against finite difference calculations.<br />

In the following two illustrative examples are presented. A number of data are<br />

common for all examples: Young’s modulus of elasticity E = 200 GPa, shear modulus<br />

of elasticity G =80GPa, allowable effective stress σeff = √ Fmax = 147 M Pa,<br />

density of mass ρ = 7799kg/m3 , and gravity acceleration in negative z-direction,<br />

which is orthogonal to the plane of supports (the xy plane).<br />

Specified<br />

linking<br />

Single<br />

load case<br />

5.5.2 <strong>Optimal</strong> design of a dome frame<br />

Figure 5.12 shows a dome with linking of joint positions as well as beam areas, which<br />

gives only 13 independent design variables. Furthermore, by this symmetry requirement,<br />

we can deal with fewer load cases, in the following illustrated by load cases<br />

restricted to a quarter part and a half part, respectively.<br />

000000000<br />

111111111<br />

000 11100<br />

11<br />

000 11100<br />

11<br />

000 11100<br />

11<br />

000 111 0000 1111 00 11<br />

0000 1111 00 11<br />

0000 1111 0000 1111<br />

0000 1111<br />

000<br />

000 111<br />

000 111<br />

000 111<br />

000 111<br />

111<br />

0000 1111<br />

0000 1111<br />

0000 1111<br />

0000 1111<br />

0000 1111<br />

000000<br />

0000 1111<br />

0000 1111<br />

0000 1111<br />

000 1110000 1111 00 11<br />

000 1110000<br />

1111 00 11 0000 1111<br />

000 11100<br />

11 0000 1111<br />

000000000<br />

111111111<br />

000 11100<br />

11 0000 1111 000000<br />

111111 01<br />

000000<br />

111111 01<br />

000000<br />

111111 01<br />

000000<br />

111111 01<br />

111111 01<br />

0000 1111 01<br />

0000 1111 01<br />

0000000<br />

1111111 0000 1111 01<br />

0000<br />

00 11<br />

00 11<br />

0000 11110000<br />

1111<br />

0000 11110000<br />

1111<br />

00 11<br />

0000 11110000<br />

1111<br />

0000000<br />

1111111<br />

01<br />

000 111<br />

0000 1111<br />

01<br />

0<br />

0000 11110000<br />

1111 0000000<br />

1111111 0000 1111 f 01<br />

1000<br />

111<br />

0000 1111<br />

11110000<br />

1111 0000 1111 01<br />

00000<br />

11111 01<br />

000 111<br />

0000 1111<br />

3<br />

7 g<br />

0000 1111 00 11 000 111 000000<br />

111111<br />

0000 1111 01<br />

00000<br />

11111 01<br />

000 111 0000 1111 b 00 11 000 111 000000<br />

111111<br />

d e 01<br />

h<br />

00000<br />

11111 01<br />

000 11100<br />

11 000 111 000000<br />

111111 01<br />

00 11<br />

00000<br />

11111<br />

000000000<br />

111111111<br />

00000<br />

11111 000 111 01<br />

01<br />

00 11 0000 1111 4 000000000<br />

111111111<br />

000 111 1 a 00 11 000 111 000000<br />

111111<br />

2 c 01<br />

6 00 11<br />

00 11<br />

00000<br />

11111 000 111 01<br />

01<br />

00 11 0000 1111<br />

01<br />

00 11<br />

00000<br />

11111 000 111 01<br />

00 11 0000 1111<br />

00000<br />

11111 000 111 01<br />

0000 1111 00 11 0000 1111<br />

00000<br />

11111 000 111 01<br />

0000 1111 5<br />

00 11<br />

01<br />

000 1110000<br />

1111 0000 1111 0000 1111<br />

00 1101<br />

000 1110000<br />

1111 0000 1111<br />

01<br />

00 110000000<br />

1111111<br />

01<br />

000 1110000<br />

1111<br />

01<br />

0000 1111<br />

0000 1111<br />

01<br />

Figure 5.12: Left: Illustrations of the dome topology. Right: Symmetry linking of<br />

joint positions and their numbers together with symmetry linking of beam members<br />

and letters for these.


Design of beams and frames 101<br />

The first problem for the dome is a single load case, with only a force at the<br />

central point # 1. The force is in the negative z-direction and has the value 632745<br />

N. The resulting optimal joint positions are illustrated in figure 5.13 with numerical<br />

values for beam thicknesses and joint positions in table 5.1.<br />

00 11<br />

00 11 00 11<br />

00 11<br />

Figure 5.13: Left: Horizontal view of the optimal dome with joint vertical positions.<br />

Right: Projection on the xy plane of the optimal dome.<br />

01<br />

01<br />

00 11<br />

The second problem formulation has four load cases:<br />

00 11<br />

00 11<br />

00 11<br />

00 11<br />

1. a force at the central joint # 1 in the negative z-direction with the value 300000<br />

N.<br />

2. at all the free joints, forces in the negative z-direction where each force has the<br />

value 30000 N.<br />

3. a force at the central point # 1 in the negative z-direction with the value 150000<br />

N, and forces at the joints # 4 and 5 in the negative z-direction where each<br />

force has the value 100000 N.<br />

4. a force at the central point # 1 in the negative z-direction with the value 150000<br />

N, and forces at the joints # 2, 3 and 4 in the negative z direction where each<br />

force has the value 70000 N.<br />

This problem is solved both without and with self-weight. The resulting optimal<br />

joint positions are illustrated in figure 5.14 with numerical values for beam thicknesses<br />

and joint positions again in table 5.1.<br />

Multiple<br />

load case<br />

With and<br />

without<br />

self-weights


102 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Design Minimum Maximum <strong>Optimal</strong> values in mm<br />

variables 1 load case 4 load cases 4 load cases<br />

in mm no self-weight self-weight<br />

ta 1.5 15.0 6.199 4.363 4.346<br />

tb 1.5 15.0 1.5 1.5 1.5<br />

tc 1.5 15.0 3.667 2.959 2.935<br />

td 1.5 15.0 1.677 2.401 2.476<br />

te 1.5 15.0 1.5 1.5 1.5<br />

tf 1.5 15.0 2.968 1.607 2.375<br />

tg 1.5 15.0 1.5 1.5 1.5<br />

th 1.5 15.0 1.5 1.5 1.5<br />

z1 7000. 11000. 11000. 11000. 11000.<br />

x2 18000. 22500. 18000. 19390. 20281.<br />

z2 6000. 10000. 8430. 8636. 8178.<br />

x6 23000. 27500. 23000. 23000. 24350<br />

z6 3000. 7000. 4305. 5583. 4613.<br />

Mass kg 1887. 1703. 1875.<br />

Table 5.1: Values for side constraints and resulting optimal dome designs for three<br />

problems, differing only in applied loads.


Design of beams and frames 103<br />

Including self-weight does not change for this case the optimal design much, but<br />

in figure 5.14 we see a clear tendency for the joints to move closer toward the supports<br />

(down and out).<br />

00 11<br />

00 11<br />

00 11<br />

00 11<br />

00 11<br />

00 11 0 0 1 1<br />

00 11<br />

Figure 5.14: Top-left: Horizontal view of the optimal dome with joint vertical positions.<br />

Top-right: Projection on the xy plane of the optimal dome, without selfweights.<br />

Bottom: the same but with self-weight of the optimal frame.<br />

01<br />

01<br />

01<br />

01<br />

01<br />

01<br />

00 11<br />

5.5.3 <strong>Optimal</strong> design of a mobile crane frame<br />

00 11<br />

00 11<br />

A model of a mobile crane structure is suggested by (Apostol et al. 1995), and figure<br />

5.15 shows the model with numbers and letters for the design variables. We optimize<br />

01<br />

01<br />

01<br />

01<br />

01<br />

01<br />

01<br />

01


Linking and<br />

no-linking<br />

104 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

this structure in two versions, first with linking given only 9 independent design<br />

variables and then without linking resulting in 22 independent design variables. Only<br />

the ”roof” is subjected to optimization.<br />

a<br />

c<br />

e<br />

Figure 5.15: Mobile crane frame, with initial joint positions.<br />

The loads are all acting in the negative z-direction at the six joints at the top.<br />

The actual value at these joints are 123607 N , i.e. total load of 741642 N. Figure<br />

5.16 shows the resulting optimal shapes of the mobile crane frame with actual design<br />

values given in table 5.2.<br />

Figure 5.16: <strong>Optimal</strong> shapes of mobile crane frame. Left: with linking to enforce a<br />

straight roof line. Right: without linking.<br />

From these 3D-frame design problems it is clear that by the shape design of space<br />

frames in addition to the beam cross-sectional design we have gained a lot in the<br />

possibilities for mass minimization.<br />

For frames the shape parameters are the positions of the joints where the involved<br />

beams meet. We have seen that the sensitivity analysis with respect to these joint<br />

positions are much more involved than the sensitivity analysis with respect to the<br />

b<br />

1<br />

d<br />

f<br />

2


Design of beams and frames 105<br />

Design Minimum Initial Maximum <strong>Optimal</strong> values in mm<br />

variables Without With<br />

in mm linking linking<br />

ta 0.8 8.0 15.0 5.54 5.73<br />

tb 0.8 8.0 15.0 4.41 3.62<br />

tc 0.8 8.0 15.0 1.32 1.36<br />

td 0.8 8.0 15.0 1.08 1.18<br />

te 0.8 8.0 15.0 1.70 1.16<br />

tf 0.8 8.0 15.0 0.94 0.98<br />

z1 4500. 5000. 6000. 5099. 5521.<br />

y1 300. 1000. 2200. 545. 548.<br />

x2 500. 1000. 4700. 500. 500.<br />

z2 4500. 5000. 6000. 5531. 5521.<br />

Mass kg 2457. 1679. 1684.<br />

Table 5.2: Values for side constraints, initial design and resulting optimal crane<br />

designs for two problems, differing only in linking.<br />

cross-sectional beam parameters. This sensitivity analysis is done analytically, and<br />

we obtain confidence and computational advantages at the same time. Design dependent<br />

loads, like temperature loads and inertia loads, are included in a more general<br />

formulation.


106 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...


Chapter 6<br />

Design of laminates and<br />

plates<br />

In the science of composites the problems of major importance are manufacturing<br />

and strength estimation. The natural wish to do things better, e.g. to optimize these<br />

materials and their use, has also created interest among researchers with a primary<br />

interest in optimal design, see (Gürdal, Haftka and Hajela 1999). Structure<br />

Detailed design of laminates, like a point-wise design of orientation and thickness,<br />

might be named material design. Therefore, there is no clear boundary between this<br />

chapter and chapter 8 on material optimal design, and overlapping also occurs in<br />

relation to chapter 10 on identification and inverse problems.<br />

Effective analysis and analytical sensitivity analysis give a good background for<br />

the performed optimization. We therefore first report some of these essential tools,<br />

and primarily concentrate on laminates. Laminate analysis is an important subject<br />

within the mechanics of composite materials. Among the many good textbooks on<br />

laminates we mention (Jones 1975), (Tsai and Hahn 1980), (Vinson and Sierakowski<br />

1989), (Whitney 1987).<br />

From the book (Tsai and Hahn 1980) we quote from page 210:<br />

”It is unfortunate that the use of composite materials is limited or penalized by the<br />

non-availability of analytical tools. It is important to understand how anisotropy<br />

and non-homogeneity arise in composite laminates and to what degree they can be<br />

manipulated to perform functions not possible with conventional materials”.<br />

or material<br />

There is still some truth in this statement. Plies - layers<br />

A laminate consists of two or more plies and acts as an integral plate. Plies are<br />

also named layers or laminas, and often they are available as prepregs. The plies<br />

may be made of different materials, although the same ply material is often used<br />

107<br />

- laminas<br />

- prepregs


Stress/strain<br />

vectors<br />

108 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

throughout the laminate. The main difference between the individual plies is thus<br />

their orientation in the laminate, and these orientations are very important because<br />

the plies are often strongly non-isotropic, with ratios of modulus of the order 5-40as<br />

illustrated in table 6.1 in section 6.1.3. The increasing possibilities in manufacturing<br />

with individual placement of fibers make optimization even more important.<br />

We first concentrate on analysis, and we use the √ 2-contracted notation that<br />

simplifies the rotational transformations, as argued in (Pedersen 1995).<br />

6.1 Laminate analysis and rotational transformations<br />

To a large extent the laminate analysis is based on ”book-keeping”, where the rotational<br />

transformations of stresses, strains, and constitutive relations play a major role.<br />

We list the necessary formulas by rather simple expressions, but without derivations,<br />

and we restrict the description to 2D.<br />

6.1.1 Rotational transformations of stress and strain vectors<br />

Orthogonal transformations of stresses and strains are, as argued in chapter 11, obtained<br />

simply with the vector definitions<br />

√<br />

2ɛ12} (6.1)<br />

y2<br />

{σ} T √<br />

:= {σ11 σ22 2σ12}, {ɛ} T := {ɛ11 ɛ22<br />

x2<br />

Figure 6.1: The two Cartesian coordinate systems x1,x2 and y1,y2 with definition<br />

of the relative angle θ.<br />

θ<br />

y1<br />

x1


Design of laminates and plates 109<br />

The orthogonal matrix [T ] that transforms second order tensors between the<br />

Cartesian coordinate systems x to y, as shown in figure 6.1, is Orthogonal<br />

[T ] matrix<br />

[T ]= 1<br />

⎡<br />

√<br />

1+c2 1 − c2 2s2<br />

⎣ 1 − c2 1+c2 −<br />

2<br />

√ 2s2<br />

− √ ⎤<br />

⎦<br />

√<br />

with [T ]<br />

2s2 2s2 2c2<br />

−1 =[T ] T<br />

(6.2)<br />

with the short notation for the trigonometric functions being<br />

c2 := cos(2θ), s2 := sin(2θ), c4 := cos(4θ), s4 := sin(4θ) (6.3)<br />

In general for stresses as well as for strains (and other similar vectors) we have<br />

(6.4)<br />

{σ}y =[T ]{σ}x, {σ}x =[T ] T {σ}y, {ɛ}y =[T ]{ɛ}x, {ɛ}x =[T ] T {ɛ} T y<br />

and conclude that only one matrix, the [T ] matrix, is necessary for these transformations.<br />

Note, that if the stress and strain vectors are defined differently (traditionally<br />

σ12 not √ 2σ12, and 2ɛ12 not √ 2ɛ12), then we need more than one matrix, which are<br />

not orthogonal.<br />

6.1.2 Rotational transformations of constitutive matrices<br />

A general non-isotropic constitutive matrix [C] for two dimensional problems is defined<br />

in relation to the vector definitions (6.1), and using the four index notation from<br />

tensor notation we have<br />

⎧ ⎫ ⎡<br />

⎨ σ11 ⎬<br />

= ⎣<br />

⎩<br />

σ22<br />

√ 2σ12<br />

⎭ x<br />

C1111<br />

C1122 √<br />

2C1112<br />

√<br />

C1122 2C1112 √<br />

C2222 2C2212<br />

√<br />

2C2212 2C1212<br />

⎤<br />

⎦<br />

x<br />

⎧<br />

⎨<br />

⎩<br />

ɛ11<br />

ɛ22<br />

√ 2ɛ12<br />

⎫<br />

⎬<br />

⎭ x<br />

(6.5)<br />

Constitutive<br />

matrix<br />

Constitutive<br />

vector<br />

In optimal material design it is very convenient to contract further the constitutive<br />

matrix [C] into a vector {C} with six components (see chapter 11 for further<br />

discussion),<br />

√<br />

{C} = {C1111 C2222 2C1212 2C1122 2C1112 2C2212} (6.6)<br />

and the orthogonal matrix [R] that transforms fourth order tensors between the Cartesian<br />

coordinate systems x to y is Orthogonal<br />

[R] = 1<br />

8 ·<br />

⎡<br />

√ √<br />

3+4c2 + c4 3 − 4c2 + c4 2 − 2c4 2 − 2c4 4s2 +2s4 4s2 − 2s4<br />

√ √<br />

⎢<br />

3 − 4c2 + c4 3+4c2 + c4 2 − 2c4 2 − 2c4 −4s2 +2s4 −4s2 − 2s4<br />

⎢ 2 − 2c4 2 − 2c4 4+4c4 −2<br />

⎢<br />

⎣<br />

√ 2+2 √ 2c4 −4s4 4s4<br />

√ √ √ √<br />

2 − 2c4 2 − 2c4 −2 √ 2+2 √ 2c4 6+2c4 −2 √ 2s4 2 √ 2s4<br />

−4s2 − 2s4 4s2 − 2s4 4s4 2 √ ⎤<br />

⎥<br />

2s4 4c2 +4c4 4c2 − 4c4<br />

⎦<br />

−4s2 +2s4 4s2 +2s4 −4s4 −2 √ 2s4 4c2 − 4c4 4c2 +4c4<br />

[R] matrix


Practical<br />

definitions<br />

110 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

with [R] −1 =[R] T (6.7)<br />

We can then by a simple matrix multiplication transform the constitutive parameters<br />

from one Cartesian coordinate system to another by<br />

{C}y =[R]{C}x, {C}x =[R] T {C}y<br />

(6.8)<br />

and similar transformations are valid for contracted compliance matrices [C] −1 as well<br />

as for the stress strength matrices [F ] and the strain strength matrices [G], that is<br />

used in relation to strength constraints.<br />

6.1.3 Alternative description by practical parameters<br />

With reference to the paper (Hammer, Bendsøe, Lipton and Pedersen 1997), we<br />

now show an alternative description of the constitutive matrix [C] in a coordinate<br />

system, rotated an angle θ counter-clockwise relative to a coordinate system where<br />

the following practical parameters are defined<br />

C1 := C1111 − C2 − C3<br />

C2 := (C1111 − C2222)/2<br />

C3 := (C1111 + C2222 − 2C1122 − 4C1212)/8<br />

C4 := C1122 + C3<br />

C5 := (C1 − C4)/2<br />

C6 := (C1112 + C2212)/2<br />

C7 := (C1112 − C2212)/2 (6.9)<br />

These parameters are used in the book by (Tsai and Hahn 1980), and are in the<br />

early work (Tsai and Pagano 1968) named invariants, which may be misleading. In a<br />

non-dimensional form αn = Cn/C1111 some actual parameters are listed in table 6.1<br />

for some orthotropic materials (C6 = C7 = 0).<br />

With definitions (6.9) the description in terms of the following five symmetric<br />

matrices, defined by<br />

⎡<br />

[Γ0] := ⎣<br />

C1 C4 0<br />

C4 C1 0<br />

0 0 2C5<br />

⎡<br />

C2 0<br />

[Γ1] := ⎣ 0 −C2<br />

√ √<br />

2C6 2C6<br />

⎤<br />

⎦ (6.10)<br />

√<br />

2C6 √<br />

2C6<br />

0<br />

⎤<br />

⎦ (6.11)


Design of laminates and plates 111<br />

Materials GPa (Giga Pascal) Non-dimensional parameters<br />

EL ET GLT νLT α0 α1 α2 α3 α4 α5<br />

181.0 10.30 7.17 0.28 0.9955 0.4200 0.4716 0.1084 0.1244 0.1478<br />

Graphite/Epoxy 138.0 8.96 7.10 0.30 0.9942 0.4298 0.4675 0.1027 0.1222 0.1538<br />

207.0 5.17 2.59 0.25 0.9964 0.3922 0.4875 0.1203 0.1266 0.1328<br />

204.0 18.50 5.59 0.23 0.9952 0.4278 0.4547 0.1175 0.1384 0.1447<br />

Boron/Epoxy 207.0 20.70 6.90 0.30 0.9910 0.4365 0.4500 0.1135 0.1435 0.1465<br />

213.7 23.44 5.17 0.28 0.9914 0.4358 0.4452 0.1190 0.1498 0.1430<br />

Aramid/Epoxy 76.0 5.50 2.30 0.34 0.9916 0.4233 0.4638 0.1129 0.1375 0.1429<br />

Glass/Epoxy 38.6 8.27 4.14 0.26 0.9855 0.5221 0.3929 0.0850 0.1407 0.1907<br />

53.8 17.90 8.96 0.25 0.9792 0.6021 0.3336 0.0643 0.1474 0.2273<br />

Isotropic E E E<br />

2(1+ν) ν 1 − ν2 1 0 0 ν 1−ν<br />

2<br />

Table 6.1: Actual parameters for orthogonal materials, calculated from (Tsai and<br />

Hahn 1980) using the definitions by αn := Cn/C1111. Note that α1 + α2 + α3 =1.<br />

is simply written<br />

⎡<br />

[Γ2] := ⎣<br />

⎡<br />

[Γ3] := ⎣<br />

[Γ4] :=<br />

C3<br />

−C3<br />

√ 2C7<br />

−C3 C3 − √ 2C7<br />

√ 2C7 − √ 2C7 −2C3<br />

⎤<br />

⎦ (6.12)<br />

2C6 0 −C2/ √ 0 −2C6<br />

2<br />

−C2/ √ −C2/<br />

2<br />

√ 2 −C2/ √ 2 0<br />

⎦ (6.13)<br />

⎡<br />

⎣<br />

C7 −C7 − √ −C7<br />

−<br />

C7<br />

2C3 √<br />

2C3<br />

√ 2C3<br />

√<br />

2C3 −2C7<br />

⎤<br />

⎦ (6.14)<br />

[C]x =[Γ0]+[Γ1] cos 2θ +[Γ2] cos 4θ +[Γ3] sin 2θ +[Γ4] sin 4θ (6.15)<br />

The parameters C1 − C7 can be interpreted as follows: If the material is orthotropic,<br />

we can choose a reference coordinate system where<br />

C6 = C7 =0 with reference to orthotropic directions (6.16)<br />

and such directions exist if the following condition can be satisfied in a coordinate Orthotropic<br />

system, where (6.16) is not satisfied. The condition is<br />

materials<br />

C7C 2 2 − 4C7C 2 6 − 4C6C3C2 = 0 (6.17)<br />


Orthotropic<br />

engineering<br />

4 parameters<br />

112 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

The parameters C6,C7 are therefore only necessary for non-orthotropic laminates<br />

(materials). For an orthotropic material we can express the parameters in what is<br />

normally named the engineering parameters EL,ET ,GLT ,νLT and get<br />

C1 = EL/α0 − C2 − C3<br />

C2 =(EL − ET )/(2α0)<br />

C3 =(EL +(1− 2νLT )ET )/(8α0) − GLT /2<br />

C4 = νLT ET /α0 + C3<br />

C5 =(C1 − C4)/2<br />

with the definition α0 := 1 − ν 2 LT ET /EL<br />

(6.18)<br />

The most important parameter C2 describes the level of non-isotropy, and the parameter<br />

C3 describes the relative shear stiffness, as discussed in detail in chapter 17.<br />

Figure 6.2 shows an overview of material classification, restricted to the two dimensional<br />

case.<br />

General<br />

Non-isotropic<br />

C7C 2 2 − 4C7C 2 6 − 4C6C3C2<br />

Orthotropic<br />

C2,C3<br />

zero<br />

Isotropic<br />

both zero<br />

not zero<br />

not both zero<br />

Non-orthotropic<br />

C3<br />

negative<br />

High shear modulus<br />

Low shear modulus<br />

positive<br />

Plane stress, isotropic<br />

Plane strain, isotropic<br />

Others<br />

Figure 6.2: Overview of material classification in 2D.


Design of laminates and plates 113<br />

If the laminate (the material) is isotropic, then in any coordinate system we have Isotropic<br />

2 parameters<br />

C2 = C3 = C6 = C7 =0 ⇒<br />

C1 = C1111, C4 = C1122, C5 =(C1111 − C1122)/2 (6.19)<br />

i.e. the only non-zero parameters C1, C4 and C5 are described by two parameters.<br />

6.1.4 Laminate stiffnesses and lamination parameters<br />

In the classical plate theory the global relation between the forces and moments per<br />

unit length {N}, {M} and the mid-plane strains {ɛ0 } and curvatures {κ} is<br />

<br />

0 {N} [A] [B] {ɛ }<br />

=<br />

(6.20)<br />

{M} [B] [D] {κ}<br />

In-plane and<br />

out-of-plane<br />

When the √ 2-notation is also used for {N}, {M}, {ɛ0 }, {κ},<br />

i.e. {N} T √<br />

:= {N11 N22 2N12} etc., then the laminate stiffness matrices [A], [B], [D]<br />

are in terms of the material parameters C1 − C7 defined in (6.9) and the twelve Stiffnesses:<br />

lamination parameters ξ membrane<br />

coupling<br />

bending<br />

A,B,D<br />

1,2,3,4 given as<br />

[A] = h [Γ0]+[Γ1]ξ A 1 +[Γ2]ξ A 2 +[Γ3]ξ A 3 +[Γ4]ξ A 4<br />

[B] = h 2 [Γ1]ξ B 1 +[Γ2]ξ B 2 +[Γ3]ξ B 3 +[Γ4]ξ B 4<br />

[D] = h 3<br />

<br />

1<br />

12 [Γ0]+[Γ1]ξ D 1 +[Γ2]ξ D 2 +[Γ3]ξ D 3 +[Γ4]ξ D <br />

4 (6.21)<br />

where the lamination parameters in a global coordinate system x are defined as the<br />

weighted trigonometric integrals over the thickness<br />

ξ A,B,D<br />

1,2,3,4 :=<br />

1/2<br />

z<br />

−1/2<br />

0,1,2 (cos(2θ(z)), cos(4θ(z)), sin(2θ(z)), sin(4θ(z))) dz (6.22)<br />

This compact notation implies, as an example, that ξD 3 is given by Lamination<br />

ξ D 1/2<br />

3 :=<br />

−1/2<br />

z 2 sin(2θ(z))dz (6.23)<br />

We use the lamination parameters as the design variables, mainly due to the fact<br />

that the laminate stiffnesses are linear functions of these variables, as seen from (6.21).<br />

The lamination parameters are not independent, as there exist trigonometric relations<br />

between them. Governing the pure membrane stiffnesses or the pure bending<br />

stiffnesses the constraints are<br />

parameters<br />

Linear<br />

functions


114 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 6.3: The design domains for the four lamination parameters with one parameter<br />

equal to zero in turn. From Hammer et al. (1997).


Design of laminates and plates 115<br />

2ξ 2 1(1 − ξ2)+2ξ 2 3(1 + ξ2)+ξ 2 2 + ξ 2 4 − 4ξ1ξ3ξ4 ≤ 1<br />

ξ 2 1 + ξ 2 3 ≤ 1<br />

−1 ≤ ξ2 ≤ 1 (6.24)<br />

as discussed further in (Hammer et al. 1997). By rotating the reference coordinate<br />

system for the lamination parameters, we can force one of these to be zero, and in this<br />

way obtain a graphical illustration of the convex feasible space for these parameters,<br />

as shown in figure 6.3. Mutual relations for the coupled problems with both A, B, D<br />

lamination parameters are not yet available.<br />

6.2 Elastic energy sensitivity analysis<br />

In this section our general design parameter is taken to be the material orientation θe<br />

in the region e, which could be the laminate as a whole, a specific ply in a laminate, or<br />

more detailed an element of a ply. With power law elasticity and design independent<br />

loads we have directly from (13.13) that<br />

dUɛ<br />

dθe<br />

= − Ve<br />

p<br />

<br />

∂(ūɛ)e<br />

∂θe<br />

fixed strains<br />

(6.25)<br />

where (ūɛ)e is the mean strain energy density in region e with volume Ve, and for<br />

linear elasticity we have p =1.<br />

For coupled plate/disc problems, the strain energy per plate area ūɛt, where ūɛ<br />

is the mean strain energy density through the thickness t of the plate, is for linear<br />

elasticity given by<br />

ūɛt = 1<br />

2 {ɛ0 } T {N} + 1<br />

2 {κ}T {M} (6.26)<br />

which with relations (6.20) is written Strain energy<br />

ūɛt = 1<br />

2 {ɛ0 } T [A]{ɛ 0 } + {ɛ 0 } T [B]{κ} + 1<br />

2 {κ}T [D]{κ} (6.27)<br />

Inserting this in (6.25) we get (linear elasticity)<br />

<br />

dUɛ 1<br />

= −ae<br />

dθe 2 {ɛ0 T ∂[Ae]<br />

e} {ɛ<br />

∂θe<br />

0 e} + {ɛ 0 T ∂[Be]<br />

e} {κe} +<br />

∂θe<br />

1<br />

<br />

T ∂[De]<br />

{κe} {κe}<br />

2 ∂θe<br />

(6.28)<br />

where ae is the area of the actual laminate region, and index e on the involved<br />

quantities relates to this actual region.<br />

per area


Only four<br />

stationary<br />

solutions<br />

Orthotropic<br />

analytical<br />

solutions<br />

116 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Even for the fully coupled problems this result can be written<br />

dUɛ<br />

= U1 sin(2θe)+U2 sin(4θe)+U3 cos(2θe)+U4 cos(4θe) (6.29)<br />

dθe<br />

which follows from (6.21) or directly from the rotational matrix [R] in (6.7). The<br />

energy coefficients U1, U2, U3, U4depend on the material as well as on the specific<br />

strain state.<br />

Before treating the specific and simplified problems, it should be appreciated that<br />

according to (6.29) we can, in the general case, find at most four different solutions<br />

to the stationarity condition dUɛ/dθe = 0. This follows from rewriting (6.29) as a<br />

fourth order polynomial. However, analytical expressions for this general case are too<br />

complicated to be shown here.<br />

For orthotropic materials and models where only the cosine terms are involved in<br />

(6.21), analytical solutions to dUɛ/dθe = 0 are not complicated. After differentiation<br />

we keep in (6.29) only the sine terms, and have for these (often named specially<br />

orthotropic or balanced) models<br />

<br />

dUɛ<br />

U1<br />

=2U2sin(2θe) + cos(2θe)<br />

(6.30)<br />

dθe<br />

2U2<br />

Stationarity is then obtained for<br />

θe =0, θe = π/2, θe = ± 1<br />

2 arccos<br />

<br />

− U1<br />

<br />

(6.31)<br />

2U2<br />

and, furthermore, supplementary angles return the same energy density uɛ(π − θ) =<br />

uɛ(θ) when only cos(2θ) and cos(4θ) appear in the stiffness expressions. Thus, for this<br />

case the orientational dependence is described completely by the interval 0 ≤ θ ≤ π/2.<br />

6.3 <strong>Optimal</strong> orientation of material in<br />

simply supported plates in bending<br />

Figure 6.4 illustrates a simply supported plate placed in a Cartesian coordinate system<br />

x1,x2. The displacement field for the transverse displacement w = w(x1,x2) of this<br />

plate in bending is given by<br />

w(x1,x2) =wmn sin(mπx1/a) sin(nπx2/b) (6.32)<br />

with number of half-waves m along the length a in the x1-direction and number of<br />

half-waves n along the length b in the x2-direction, in agreement with the actual load.


Design of laminates and plates 117<br />

x1<br />

b<br />

x3<br />

w(x1,x2)<br />

Figure 6.4: A rectangular plate with dimensions a, b, assumed simply supported<br />

along all four boundaries.<br />

a<br />

For rectangular, simply supported plates in pure bending we introduce a mode<br />

parameter η that describes the displacement pattern (for a square pattern we have<br />

η = 1). The mode parameter is then defined as the ratio between the two actual<br />

half-wave lengths of the deformation by Deformation<br />

ηmn := b/n mb<br />

=<br />

a/m na<br />

x2<br />

(6.33)<br />

A linear combination of the bending stiffnesses Φmn is for a specific deformation<br />

mode ηmn defined by<br />

mode<br />

parameter<br />

Φmn := 1 4<br />

ηmnD1111 + D2222 +2η<br />

8<br />

2 mn(D1122 +2D1212) <br />

(6.34)<br />

where D1111,D2222,D1122,D1212 are the four components of the orthotropic laminate Objective<br />

bending stiffness matrix, defined in (6.20). Φmn is our objective to extremize, because Φmn<br />

eigenfrequencies, buckling loads, and displacement amplitudes for the corresponding<br />

rectangular plate is then also extremized. The relations to these physical quantities<br />

can be found in the book (Jones 1975) and in other classical books, and they are for<br />

eigenfrequencies:<br />

ωmn = n2π2 b2 <br />

Φmn<br />

ρt<br />

(6.35)


Eigen<br />

frequencies<br />

Buckling load<br />

Maximum<br />

displacement<br />

Non-trivial<br />

optimal<br />

orientation<br />

118 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

where ρ is the mass density of the plate, t is the thickness of the plate, and m, n give<br />

the actual mode.<br />

For buckling load (per length) in the x1 − direction with most dangerous mode<br />

corresponding to m, n =1, 1 we have<br />

(Nx1 )buckling = π2<br />

η2 Φ11<br />

(6.36)<br />

b2 and finally corresponding to a distributed load p = p(x1,x2) we get the maximum<br />

displacement |w|max by<br />

p = pmn sin(mπx1/a) sin(nπx2/b) ⇒|w|max = pmn<br />

b 4<br />

1<br />

n4π4 Φmn<br />

(6.37)<br />

For laminates with orthotropic bending stiffness matrix , i.e. D1112 = D2212 =0<br />

with reference to the rectangular side directions, the orientations that give stationary<br />

objective Φ, (as defined in (6.34)), can be found analytically as by (6.31). The energies<br />

U1 and U2 in (6.31) depend on the specific material as characterized by C2,C3 defined<br />

in (6.9) and on the mode parameter ηmn. The derivation in (Pedersen 1987b) focus<br />

on the result<br />

(θe)optimal = 1<br />

2 arccos<br />

<br />

1 − η4 <br />

mn C2<br />

(6.38)<br />

4C3<br />

1 − 6η 2 mn + η 4 mn<br />

and figure 6.5 shows the results of optimizing the eigenfrequency, as also found in<br />

(Bert 1977).<br />

From an engineering point of view the main conclusions to be drawn from figure<br />

Practical<br />

6.5 are:<br />

conclusions • The optimal orientation depends mainly on the mode parameter ηmn. Thus<br />

if the deformation pattern is known, the optimal direction can be estimated<br />

directly.<br />

• Cases of inverse mode parameters ηa = η −1<br />

b have complementary solutions<br />

θb = π/2 − θa.<br />

• For ”extreme” values of η the optimal orientation is perpendicular to the long<br />

wavelength.<br />

• The change of optimal direction to a skew direction is very sensitive to the<br />

mode parameter.<br />

• The optimal orientation is rather insensitive to the material parameters.<br />

• The optimal orientation is independent of the position of the ply in the laminate,<br />

and thus the same for all plies.


Design of laminates and plates 119<br />

• It is not seen in the figure but local optima exist.<br />

By simple numerical analysis, problems with combined modes can also be optimized,<br />

(Pedersen 1986b) shows such solutions. Often a specific mode is dominating,<br />

and then solutions from figure 6.5 are a good starting design for more detailed iterations.<br />

Figure 6.5: <strong>Optimal</strong> orientation as a function of the mode parameter ηmn for plate<br />

bending. Note that α2/α3 = C2/C3, defined in (6.9).


Simple<br />

example<br />

Large<br />

improvements<br />

No energy<br />

concentration<br />

120 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

6.4 Laminate point-wise design<br />

The importance of laminates has put increasing demands on the possibility to manufacture<br />

graduated thickness as well as orientation, and even individual fiber laying<br />

is possible. Thus the point-wise design of orientation and thickness is a realistic<br />

optimization problem.<br />

Figure 6.6: A uniformly loaded cantilever example, for which the thickness distribution<br />

as well as the orientational field is optimized.<br />

To exemplify the different effects of thickness and orientational optimization the<br />

cantilever problem shown in figure 6.6 is solved. The objective is to minimize compliance<br />

(minimize stored elastic energy) and at the same time minimize concentration<br />

of energy density, but without involving this later aspect in the formulation. Simple<br />

heuristic iterations based on optimality criteria have been applied. For the thickness<br />

iterations this criterion is based on uniform energy density, and for the orientational<br />

iterations alignment with principle stress/strain directions is applied, see chapter 17.<br />

Figure 6.7 shows the results. In this simple figure a black and white visualization<br />

of the results is used. The design is characterized by thickness and orientation, which<br />

are shown by hatching the triangular finite elements in the direction of the larger<br />

modulus direction and with the hatch density proportional to the thickness. Dark<br />

areas are therefore areas with large thicknesses. The cantilever results shown in figure<br />

6.7 are based on a 720 triangular constant stress element model, assuming constant<br />

thickness and orientation in each element. For the uniform cantilever, the mean<br />

(normalized to 1.0) and the maximum relative values of energy density are 1.0, 41.8<br />

(relative measures of stiffness and stress concentration). Only thickness optimization<br />

gives 0.53, 0.57, which means stiffness improved by a factor 1.9 and almost no energy<br />

concentration. With simultaneous thickness and orientational optimization we obtain<br />

0.23, 0.25, i.e. the stiffness improved by a factor of 4.3, and again almost no energy<br />

concentration.


Design of laminates and plates 121<br />

Figure 6.7: Above: the optimal thickness distribution (dark equal to large thickness)<br />

for a uniformly loaded cantilever. Below: result of the combined thickness and<br />

orientational optimization.


Non-linear<br />

elasticity<br />

Design<br />

independence<br />

122 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

6.4.1 Thickness distribution only<br />

Design of laminate thickness distribution for maximum stiffness is characterized by<br />

resulting uniform energy density, and is with many examples illustrated in chapter<br />

4. With three examples taken from (Pedersen and Taylor 1993) we give results that<br />

show the solutions to be independent of a material non-linear elasticity, modelled by<br />

a power law, see chapter 12.<br />

Figure 6.8 gives the three specific examples and table 6.2 gives the relative resulting<br />

values. The main result is that the optimal thickness distribution is independent<br />

Figure 6.8: The three example cases: a) uniformly loaded cantilever of isotropic<br />

material, b) circular hole loaded biaxially 2:1, isotropic material, c) optimal designed<br />

shape of the hole, loaded biaxially 3:2, orthotropic material.<br />

of the power p. Thus the results for the problems of figure 6.8 mainly illustrate the<br />

influence from the power p on two given designs, i.e. the uniform thickness design<br />

and the optimal thickness design as obtained from linear elasticity.<br />

In terms of relative values table 6.2 gives the strain energy densities. Minimum,<br />

mean and maximum values are related to the elements of the finite element models.<br />

The table clearly shows the different results from uniform and optimal thickness distribution.<br />

It is well known from optimization based on linear elasticity that certain<br />

areas in a model cannot be fully stressed (too little energy density). Thus the minimum<br />

values are of minor interest, and the agreement of the maximum values with the<br />

mean values better shows the fulfillment of the optimality criterion of uniform energy<br />

density.<br />

The relative values of the objective function are given by the relative mean values<br />

in table 6.2. The factor between energy in uniform design and energy in optimal<br />

design is almost constant for the three cases, with a weak tendency to be more important<br />

with increasing non-linearity (decreasing p). The stronger effect for the cantilever<br />

problem reflects the initial less uniform/strain distribution. Also the actual<br />

stress/strain level (higher for the cantilever problem) has an influence, and thus the


Design of laminates and plates 123<br />

Uniform or Strain energy densities in % of reference energy density<br />

optimal cantilever, isotropic circular hole, isotropic opt. hole, orthotropic<br />

with p = min. mean max. min. mean max. min. mean max.<br />

Uni. 1.0 0.5 100 577 4 100 677 47 100 387<br />

Opt. 1.0 14 50 50 81 87 89 59 90 92<br />

Uni. 0.9 0.7 163 935 4 140 951 60 138 528<br />

Opt. 0.9 23 75 76 112 120 122 63 123 138<br />

Uni. 0.8 0.9 280 1599 3 202 1391 79 196 737<br />

Opt. 0.8 39 120 121 159 170 173 69 173 221<br />

Uni. 0.7 1 512 2916 2 304 2127 108 288 1060<br />

Opt. 0.7 68 201 202 236 251 256 80 253 369<br />

Uni. 0.6 2 1005 5730 0.3 482 3430 153 445 1581<br />

Opt. 0.6 126 357 359 365 387 394 90 385 641<br />

Uni. 0.5 2 2151 12310 0.1 808 5891 227 722 2464<br />

Opt. 0.5 251 683 686 595 630 640 107 618 1169<br />

Table 6.2: Table of relative results with uniform thickness (Uni.) and with optimal<br />

thickness distribution (Opt.) for linear elasticity (p = 1) and for five models of nonlinear<br />

elasticity modelled by the power p


Multiple<br />

124 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

For computational examples based on first locating the optimal lamination parameters<br />

and then identifying a laminate resulting in these parameters, see (Hammer<br />

et al. 1997). The laminate identification problem can be solved as described in chapter<br />

10, and a laminate with at most three plies can be constructed analytically, see<br />

(Lipton 1994).<br />

6.5 Strength optimization<br />

On a more or less heuristic basis, designs for optimal strength have been taken to be<br />

designs where principal stress direction is aligned with material orthotropy direction.<br />

With a non-symmetric strength criterion like the Tsai-Wu criterion, such solutions<br />

can not be the optimal ones, because optimal stiffness design (energy independent of<br />

sign) is different from optimal strength design.<br />

The strength problem is a local problem and therefore simple results from sensitivity<br />

analysis can not be expected. To solve the optimal design problem in a proper way,<br />

we must use mathematical programming. In this section we formulate the problem<br />

and show results from (Hammer 1994).<br />

load cases The optimization problem formulated in words is: maximize a common load factor,<br />

subject to given strength constraints. Multiple plies, multiple loads, and multiple<br />

strength constraints are included in the formulation. Let {Al} be the load distribution<br />

vector corresponding to the load case l, then from finite element analysis the resulting<br />

nodal displacement vector {Dl} is found by<br />

[S]{Dl} = λ{Al} for l =1, 2,...,L (6.39)<br />

where [S] is the stiffness matrix of the actual design (finally the optimal design) and<br />

Multiple<br />

λ is a load factor, common to all load cases L.<br />

plies From {Dl}, directly strains and stresses in every ply k of every element e follow<br />

Multiple<br />

strength<br />

constraints<br />

{Dl} ⇒ {ɛekl}, {σekl}<br />

for e =1, 2,...,E; k =1, 2,...,K; and l =1, 2,...,L (6.40)<br />

Now the load strength Fn corresponding to a given strength criterion n can be determined<br />

for each of these ekl and compared to the strength limit (F0)n. Formulating<br />

our problem in relation to first ply failure (FPF) we thus have the constraints<br />

(F/F0)ekln ≤ 1<br />

for e =1, 2,...,E; for k =1, 2,...,K;<br />

for l =1, 2,...,L; and for n =1, 2,...,N (6.41)


ective<br />

Design of laminates and plates 125<br />

i.e. strength limits should be satisfied in all K plies of all E elements, for all L load<br />

cases, and this should be the fulfilled for all N specified strength constraints.<br />

The objective of our optimization is by means of orientational design (and/or<br />

thickness distribution design) to<br />

Maximize λ subject to the strength constraints (6.42)<br />

i.e. constraints (6.41) must be fulfilled. Analytical sensitivity analysis, as derived<br />

in (Hammer 1994), can be carried through, and then the main practical problem is<br />

related to the existence of a large number of local optima. In a way this problem<br />

was expected. Different solutions to this problem include strategies for choosing the<br />

initial design before mathematical programming is applied. A valuable alternative<br />

is to describe the design with global design parameter. This approach is described<br />

before an example is shown.<br />

6.5.1 Global design parameters<br />

The idea of a global design description used frequently in shape optimization, see<br />

chapter 7, is extended from curve to surface parametrization. The design is given<br />

as a linear combination of orthogonal functions. Application is here to orientational<br />

design of laminates for strength optimization, but the technique is directly applicable<br />

to design also of thickness distribution. The advantages by this design description are<br />

many, including control on smoothness, control on slopes, and control on connections<br />

at the design boundaries. This simplified parametrization of the design makes it<br />

possible to work with only few design variables, say 25 for a whole design domain. 1D-design<br />

description<br />

Figure 6.9: Illustration of the mapping from reference line to the base curve.<br />

Let us first illustrate the design description from (Pedersen, Tobiesen and Jensen<br />

1992) in a form that makes our extension from curve to surface more clear. In figure


2D-design<br />

description<br />

126 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

6.9 we show the reference domain −1 ≤ ξ ≤ 1, and the base curve C0 with the<br />

mapping illustrated. The design curve C is then described by<br />

C = C0 +<br />

N<br />

znfn(ξ) (6.43)<br />

where fn(ξ) are given functions, and the linear combination factors zn are the design<br />

variables. Mutual orthogonal functions may be chosen as the first vibration modes<br />

to a clamped/clamped beam, if the positions and slopes at both ends of the curve C<br />

should be identical to those of C0. Naturally, this selection of functions has no relation<br />

to vibration at all - it is just a convenient way of obtaining well known orthogonal<br />

functions, and surely other possibilities exist. Note that other conditions at the ends<br />

of C0 can be obtained from other boundary conditions of the related beam problems.<br />

Now for our 2D-problem we show in figure 6.10 the reference domain −1 ≤ ξ1 ≤<br />

1, −1 ≤ ξ2 ≤ 1 and a finite element mesh put onto it. If a constant design (thickness<br />

and/or orientation) is wanted in each element, this may be controlled by the values<br />

in the center of the corresponding element in the reference domain. Mapping is<br />

from element in reference model to corresponding element in actual model, and from<br />

triangle center to corresponding triangle center.<br />

Figure 6.10: Illustration of the mapping from reference domain to the actual finite<br />

element model.<br />

n<br />

The material orientation θ in the reference domain is given by<br />

θ(ξ1,ξ2) = <br />

zmnfm(ξ1)fn(ξ2) form, n =0, 1, 2,...N (6.44)<br />

m<br />

n<br />

where the f functions are given mutual orthogonal functions, and thus the zmn in<br />

(6.44) are the design parameters. In this specific case we choose the f functions to be


Design of laminates and plates 127<br />

f0(ξ) =1; f1(ξ) =ξ;<br />

fm(ξ) = cosh(kmξ)+µm cos(km)ξ) for m =2, 4,...<br />

fm(ξ) = sinh(kmξ)+µm sin(km)ξ) for m =3, 5,... (6.45)<br />

where the constants km,µm are determined by<br />

tan(km)+(−1) mtanh(km) =0 for m =2, 3, 4 ...<br />

µm = cosh(km)/ cos(km) for m =2, 4,...<br />

µm = sinh(km)/ sin(km) for m =3, 5,... (6.46)<br />

Figure 6.11: Problem with an elliptic hole where symmetry is assumed, and the two<br />

uniform loads act independently.<br />

With this design description, i.e. using the zmn in (6.44) as design variables, an<br />

example from (Hammer 1994) illustrates the possibilities. The example is a two ply,<br />

two load cases problem with Tsai-Wu strength constraint as well as maximum strain<br />

constraint. The problem and the quarter model are shown in figure 6.11, and the<br />

optimal orientational designs with the optimized load factors for increasing number<br />

Design<br />

functions<br />

of design variables are presented in figure 6.12. Manufacturing<br />

The results in figure 6.12 show how the complexity of the optimized orientational<br />

field is controlled by the global design description. Thus, this parametrization is a<br />

tool that can take the possibilities for practical manufacturing into account. Note<br />

also that the gain in going from 9 design variables to 16 design variables is rather<br />

small (6.76 compared to 6.54).<br />

constraints


128 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 6.12: <strong>Optimal</strong> orientation with increasing number of design variables. Upper<br />

ply orientations to the left and lower ply orientations to the right. Density of hatch<br />

illustrates the resulting stress levels. From (Hammer 1994).


Chapter 7<br />

Shapes of minimum stress<br />

concentration<br />

In the design of structural connections as well as in the micro-mechanics design of<br />

materials, boundary shapes play an important role. The objective may be the stiffest<br />

design, the strongest design or just a design of uniform energy density along the<br />

shape. In an energy formulation in chapter 14 it is proven that these three objectives<br />

have the same solution, at least within the limits of geometrical constraints, including<br />

the parametrization. Without involving stress/strain fields, the proof holds for 3Dproblems,<br />

for power-law non-linear elasticity, and for non-isotropic elasticity. The<br />

proof is based on an assumption of a constant volume of material.<br />

Many results are available within optimal shape design, but still a number of important<br />

issues need to be addressed. The literature gives a picture of three different<br />

Structures<br />

or materials<br />

directions of research. In mechanical and civil engineering focus has for more than Different<br />

25 years been on design for minimum stress concentration, see (Ding 1986). In the<br />

material and more mathematical oriented research the focus has been on design for<br />

minimum compliance, see (Vigdergauz 1997). Some heuristic approaches have focused<br />

on design for uniform stress, see (Xie and Steven 1997). Not too many mutual references<br />

are given among these three research directions. A goal of the present chapter,<br />

together with chapter 14, is to show that a unification is possible, because the three<br />

different objectives give the same solution under certain conditions.<br />

Studying the results of a variety of rather different shape design problems, it is<br />

objectives<br />

seen that very general knowledge can be obtained. The striking generality is the General<br />

property of uniform energy density along the designed boundaries for all the different<br />

formulations mentioned.<br />

Mainly from formulations with minimum stress concentration we list here a num-<br />

129<br />

knowledge


Neighbouring<br />

redesign<br />

Modelling<br />

issues<br />

Minimum<br />

of maximum<br />

Only<br />

minimum<br />

130 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

ber of analytically and/or numerically obtained results and the theoretical background<br />

for this is presented in chapter 14.<br />

7.1 Statement of the actual problems<br />

Like most optimal design problems, as discussed in section 2.7, the problem of shape<br />

design for minimum stress concentration is highly non-linear, and must be solved<br />

iteratively. Thus, the problem is converted to a sequence of problems of optimal<br />

redesigns; where a given shape is changed into a better neighbouring shape. The<br />

solution to this problem involves three steps: finite element analysis for a given shape,<br />

sensitivity analysis with respect to the parameters describing the design, and optimal<br />

decision for redesign.<br />

Furthermore, it should be kept in mind that the quantitative description of stress<br />

concentration is very much depending on the finite element modelling, and very small<br />

changes in shape can drastically influence the actual values. Modelling is therefore<br />

related to analytically smooth shape modelling, and the element sizes in the critical<br />

domains are to a large extent kept very small and with only small changes from one<br />

iteration to the next iteration.<br />

In mathematical terms, the objective of this chapter of shape design is to<br />

Minimize Maximum σ 2 eff<br />

(over feasible shapes) (over the structural space x and load cases) (7.1)<br />

where the squared effective stress σ 2 eff<br />

could also be the strain energy density uɛ.<br />

Converting (7.1) to a pure minimization problem on redesign by increments of design<br />

parameters ∆hi, and including for problems of hole design a volume constraint V = ¯ V ,<br />

the actual redesign formulation is<br />

Minimize σ2 max<br />

(within move−limits on ∆hi)<br />

subject to the constraints :<br />

σ 2 eff (x)+<br />

i<br />

∂σ 2<br />

eff (x)<br />

∂hi<br />

∆hi − σ 2 max ≤ 0<br />

V + <br />

i ∂V<br />

∂hi ∆hi − ¯ V = 0 (7.2)<br />

where σ 2 max is an unknown in addition to the unknown increments ∆hi. This general<br />

technique of changing from a minimum-maximum formulation to a pure minimum<br />

formulation is often used in optimal design, and there named the bound formulation.<br />

The statement (7.2) is a problem within linear programming as described in general<br />

in chapter 20.


Shapes of minimum stress concentration 131<br />

Specific problems, with and without the volume constraint, are in the following<br />

sections presented with focus on the obtained results.<br />

7.2 The 2D-fillet in tension<br />

The 2D-fillet-problem, defined with the goal of minimizing stress concentration was<br />

solved by (Tvergaard 1973) using a finite difference method for stress analysis. In<br />

(Kristensen and Madsen 1974) and in (Francavilla, Ramakrishnan and Zienkiewicz<br />

1975) the same problem was solved using the finite element method for stress analysis.<br />

Within the limits of the imposed geometrical constraints (length of the fillets and the<br />

Early<br />

references<br />

parametrization) the results of optimization give constant tangential stress along the Constant<br />

designed boundary. At this boundary we have an unidirectional state of stress and<br />

therefore constant energy density as well as constant von Mises stress. Thus these<br />

early papers point toward the importance of constant energy density.<br />

In relation to the fillet model in figure 7.1 we illustrate the importance of design<br />

to obtain remarkably better stress fields. Primarily we quantify the very negative<br />

effect of small circular connection zones.<br />

l1<br />

l2<br />

lf<br />

l3<br />

l4<br />

Figure 7.1: Fillet model for the 2D-fillet in tension. All calculations and illustrations<br />

to follow are related only to the lower left quarter part, due to symmetries.<br />

energy<br />

density


Overall<br />

parameters<br />

FE model<br />

Severe<br />

concentration<br />

Influence of<br />

stiffness<br />

132 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

In the examples to follow we use the following dimensions and parameters:<br />

• length l1 =30mm, length l2 =60mm, length l3 =90mm<br />

• length l4 =40→ 60 mm , length of fillet lf =10→ 60 mm<br />

• isotropic material and Poisson’s ratio ν =0.3<br />

• resulting maximum stress levels relative to the applied uniform stresses<br />

7.2.1 Circular connection fillet<br />

Figure 7.2 shows the results for 6 different models, all with circular shapes, but with<br />

increasing radius and thus increasing fillet length and corresponding increasing larger<br />

dimension l4, as defined in figure 7.1. Figure 7.2 shows by isolines the resulting fields<br />

of energy densities. In this figure it is, by the green (gray) areas added to the boundary<br />

of the hole, illustrated how the energy density varies along the boundary of the hole<br />

(the same technique is applied in other figures of the chapter).<br />

We conclude that a larger circular radius is advantageous, but even the largest<br />

one shows a significant stress concentration of more than 50 %. In relation to specific<br />

numbers it should be noted that the finite element model used, is in the range of<br />

20.000 constant stress elements and 20.000 degrees of freedom, and the element sizes<br />

close to the stress concentration areas have length of about 0.1 mm. We have listed<br />

the results with four digits to show the relative tendencies. Note however, this is not<br />

an indication of the actual accuracy for the model.<br />

In the examples in figure 7.2 we note that the largest possible circular fillets are<br />

used (radius = l4 − l1). The stress concentration is even larger, if smaller circular<br />

fillets are applied to larger change in dimensions. To illustrate this we show in<br />

figure 7.3 the circular fillet of 10 mm applied also to the case of l4 = 60 mm. The<br />

stress concentration factor is then increased from 1.868 to 2.111. The explanation<br />

for this is that larger dimension of l4 increases the stiffness close to the area of stress<br />

concentration and thus has a negative influence.


Shapes of minimum stress concentration 133<br />

Figure 7.2: Stress distributions for circular shapes with increasing fillet length, equal<br />

to 10, 14, 18, 22, 26, 30 mm, respectively. The length of the dimension l4 in the<br />

model figure is thus 40, 44, 48, 52, 56, 60 mm, respectively. The corresponding stress<br />

concentration factors are 1.868, 1.774, 1.695, 1.627, 1.568, 1.516, respectively.


134 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 7.3: Stress distributions for a circular shape with circular radius equal to 10<br />

mm, although the change in dimensions is 30 mm. The corresponding stress concentration<br />

factor is 2.111, to be compared to the value of 1.868 for the corresponding<br />

case with change in dimensions equal to the fillet radius, i.e. 10 mm.


Shapes of minimum stress concentration 135<br />

7.2.2 Super-circular connection fillet<br />

A family of shapes is defined by<br />

x η<br />

1 + xη2<br />

= r (7.3)<br />

including the circle for η = 2 and in the limit the square for η →∞. For η>2we<br />

name this super-circles, as described in section 14.5. Discontinuous<br />

The circular connection to a straight domain is not the best solution, because a curvature<br />

discontinuity in curvature takes place, with large jumps especially for small circular<br />

radii. We have seen that large stress concentration factors result for the defined model.<br />

The super-circular shapes have continuous change in curvature from the straight line,<br />

and is therefore expected to perform better. Figure 7.4 shows the results with different<br />

powers η for the case of a fillet length equal to the change in dimensions, here 30<br />

mm. For this case we conclude that η =2.275 is to be preferred, and the stress<br />

concentration is now minimized from 1.516 to 1.496, which is however, only a minor<br />

improvement.<br />

In relation to smaller fillets relatively more can be obtained by the super-circular<br />

design, but in general the very important design change is to change from supercircular<br />

design to super-elliptic design, which we illustrate in the next subsection.<br />

7.2.3 Extended length of the fillet<br />

Of major importance for minimization of stress concentration, is the possible length<br />

of the fillet. With unlimited length we can totally avoid stress concentration, so an<br />

optimization must be based on a given length of the fillet. Figure 7.5 illustrates this<br />

major influence, here related to super-elliptic designs with η =2.25. We notice that<br />

almost uniform stress (energy density) along the shape is obtained for the long elliptic<br />

shapes, and thus according to chapter 14 confirms the versatility of the simple design<br />

Major<br />

importance<br />

parametrization. Down to 15 %<br />

concentration<br />

Still, the best stress concentration is 15 % and especially for fillets in tension,<br />

it is very difficult to avoid stress concentration. By adding more design variables<br />

we are able to minimize the stress concentration further, as it is verified in the next<br />

subsection.<br />

7.2.4 Multi-parameter optimal shape design<br />

With more or less single parameter design studies we have seen that the fillet design<br />

is very sensitive to many aspects of design possibility. Based on this we now conclude<br />

that the length of the fillet should be as long as possible, thus we assume this<br />

parameter to be given. Discontinuity in change of curvature of the shape should be


136 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 7.4: Stress distributions for super-circular shapes with increasing power of<br />

the super-circle (η =2.0 for circle), equal to 2.0, 2.1, 2.2, 2.275, 2.3, 2.4, respectively.<br />

The corresponding stress concentration factors are 1.516, 1.503, 1.497, 1.496, 1.496,<br />

1.498, respectively.


Shapes of minimum stress concentration 137<br />

Figure 7.5: Stress distributions for super-elliptic shapes with increasing relative<br />

length of the super-ellipse (η =2.25). The lengths relative to the change in dimensions<br />

are equal to 2/3, 2.5/3, 1, 4/3, 5/3, 6/3, respectively. The corresponding stress<br />

concentration factors are 3.124, 1.625, 1.496, 1.327, 1.222, 1.153, respectively.


Down to 8 %<br />

concentration<br />

Axisymmetric<br />

modelling<br />

138 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

avoided, thus we use super-elliptic power greater than two. The major optimization<br />

parameter is the super-elliptic axis orthogonal to the length direction of the fillet,<br />

taken relative to the change in dimensions. In the chosen specific case for which we<br />

minimize the stress concentration we take, with reference to the model in figure 7.1,<br />

l4 =2l1 =60mm and length of the fillet of the same size. Figure 7.6 shows the stress<br />

field with the optimized shape, and a stress concentration of only 8 % is obtained.<br />

Figure 7.6: Stress distributions for a three-parameter optimized shape with given<br />

length of the fillet. Stress concentration factor for this design is 1.085<br />

7.3 The 3D-fillet in tension, bending and torsion<br />

The 3D-fillet of an axisymmetric solid is optimized in (Pedersen and Laursen 1982–83).<br />

Tension, bending and torsion are all treated as separated load cases and the length of<br />

the fillet is imposed as a geometrical constraint. The detailed stress distributions in<br />

the paper (Pedersen and Laursen 1982–83) all show constant von Mises stress, until a<br />

point of the designed boundary where the geometrical constraint imposes a decrease<br />

in this stress. Note, that von Mises stress and energy density are only proportional<br />

for isotropic, non-compressive materials, see (Pedersen 1998).<br />

For the axisymmetric model shown in figure 7.7, the results are presented in figure<br />

7.8, 7.9, 7.10, 7.11, 7.12, and 7.13. All these results are from (Pedersen and Laursen<br />

1982–83).


Shapes of minimum stress concentration 139<br />

Figure 7.7: Shaft with a shoulder fillet loaded in bending, tension, or torsion.<br />

Stress concentration factor : 1.025 1.014 1.007 1.000<br />

1.146 for circular design<br />

Length of the fillet : 0.9 1.0 1.1 1.3 ×∆R<br />

∆R = R2 − R1<br />

R1<br />

BENDING load<br />

R2/R1 =11/6<br />

Figure 7.8: <strong>Optimal</strong> shapes with minimum stress concentration for bending load.<br />

Solutions for different lengths of the transition zone.<br />

R2


140 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 7.9: Stress levels corresponding to a bending load, for a circular design at the<br />

top and for an optimal design at the bottom.


Shapes of minimum stress concentration 141<br />

For the case of bending load, the optimal shapes as a function of fillet length<br />

are shown in figure 7.8, indicating that with the actual finite element solution (1980)<br />

no stress concentration is found for fillet length greater than 1.3 times the change in<br />

radial dimension. We especially note that optimal fillets are obtained by removing<br />

material relative to the circular design.<br />

Stress levels for a specific fillet length are presented in figure 7.9, comparing the<br />

stress field for a circular design with the field for the optimal design.<br />

Stress concentration factor :1.120 1.099 1.080 1.056 1.012<br />

1.288 for circular design<br />

Length of the fillet : 0.9 1.0 1.1 1.3 2.0 ×∆R<br />

∆R = R2 − R1<br />

R1<br />

T ENSION load<br />

R2/R1 =11/6<br />

Figure 7.10: <strong>Optimal</strong> shapes with minimum stress concentration for tension load.<br />

Solutions for different lengths of the transition zone.<br />

For the case of tension load, the optimal shapes as a function of fillet length are<br />

shown in figure 7.10, indicating that stress concentration is still found for fillet length<br />

greater than 2 times the change in radial dimension. Thus as found in section 7.2<br />

with a two dimensional model, stress concentration with tension load is very difficult<br />

to minimize. Stress levels for a specific fillet length are presented in figure 7.11,<br />

comparing the stress field for a circular design with the field for the optimal design.<br />

For the case of torsion load, the optimal shapes as a function of fillet length<br />

are shown in figure 7.12, indicating that stress concentration is still found for fillet<br />

length greater than 1.3 times the change in radial dimension, but not so severe as<br />

for the tension loading case. The optimal shapes for the three different loading cases<br />

are compared in figure 7.13. Further detail can be found in (Pedersen and Laursen<br />

1982–83).<br />

R2


142 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 7.11: Stress levels corresponding to a tension load, for a circular design at<br />

the top and for an optimal design at the bottom.


Shapes of minimum stress concentration 143<br />

Stress concentration factor : 1.030 1.021 1.016 1.009<br />

1.106 for circular design<br />

Length of the fillet : 0.9 1.0 1.1 1.3 ×∆R<br />

∆R = R2 − R1<br />

R1<br />

T ORSION load<br />

R2/R1 =11/6<br />

Figure 7.12: <strong>Optimal</strong> shapes with minimum stress concentration for torsion load.<br />

Solutions for different lengths of the transition zone.<br />

R1<br />

Length of the fillet : 1.3 × ∆R<br />

∆R = R2 − R1<br />

TENSION<br />

BENDING<br />

TORSION<br />

R2/R1 =11/6<br />

Figure 7.13: Comparison of optimal shapes for different loads.<br />

R2<br />

R2


Load<br />

direction<br />

144 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

7.4 The 2D-hole subjected to biaxial stress<br />

2b<br />

x2<br />

2a<br />

x1<br />

σ2<br />

σ1<br />

Numerical model<br />

Figure 7.14: Left: plane problem for the study of stress/strain concentration around<br />

an elliptical hole in a finite domain, with ellipse axes a, b and external stresses σ1,σ2<br />

in directions x1,x2, respectively. Right: the actual quarter model based on symmetry.<br />

The 2D-hole-problem is illustrated in figure 7.14 and is shown for a biaxial load<br />

case. Numerical results for this case are contained in (Kristensen and Madsen 1976),<br />

and these are compared to what is mentioned as the ”analytical” results. Analytical<br />

studies of the problem are reported by (Banichuk 1977) with reference to even<br />

earlier results by (Cherepanov 1974). These analytical studies prove that a constant<br />

tangential stress is obtained with an elliptical shape design where the ratio of the<br />

two half axis a, b equals the ratio of the stresses, assuming equal signs for σ1 and σ2.<br />

More recent studies by (Cherkaev, Grabovsky, Movchan and Serkov 1998) deal with<br />

non-equal signs for σ1 and σ2. In all these cases constant energy density is obtained<br />

but it should be remembered that the hole is assumed infinitesimally small.<br />

The influence of a limited domain, i.e. holes of finite size, is of major importance,<br />

and we discuss this aspect in more detail. First we present the analytical results.<br />

7.4.1 The classic solution for stress concentration around small<br />

elliptical holes<br />

Figure 7.15 shows an elliptical hole with half-axes a, b in the x1,x2 directions, respectively.<br />

Two angles are involved in the problem, the angle φ from the x1-direction to<br />

the direction of the load. For the load a uniaxial stress state is assumed and the<br />

stress level far away from the hole is σ∞. The other angle θ specifies the point of the<br />

elliptical boundary at which we determine the actual tangential stress σθ, which is a<br />

principal stress because the stress normal to and at the elliptical boundary is zero.


Shapes of minimum stress concentration 145<br />

x2<br />

σθ(φ)<br />

θ<br />

a<br />

x1<br />

b<br />

Figure 7.15: Elliptical hole in an infinite plane domain, subjected to stress far away<br />

from the hole.<br />

The coordinates for the point at which the tangential stress is given are<br />

x1 = a cos t, x2 = b sin t (7.4)<br />

with the parameter 0 ≤ t


Concentration<br />

factor of 2<br />

Minimum<br />

stress<br />

concentration<br />

Essential<br />

assumptions<br />

Influence from<br />

finite size<br />

Elliptical are<br />

not optimal<br />

146 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

and we may talk about a stress concentration factor of 2 for this very simple case. In<br />

the more general case of (7.6) the maximum tangential stress at the circular boundary<br />

is<br />

3σ2 − σ1 ≤ σθ ≤ 3σ1 − σ2 for σ1 >σ2<br />

3σ1 − σ2 ≤ σθ ≤ 3σ2 − σ1 for σ2 >σ1<br />

(7.9)<br />

Let us go back to the elliptical hole and for the result (7.6) find the condition for<br />

a constant tangential stress, i.e. dσθ/dθ = 0 for all θ. We find after some algebra<br />

a/b = σ1/σ2 ⇒ σθ constant (7.10)<br />

the result frequently used as a test example in optimal shape design for minimum<br />

stress concentration. For this case the constant stress is<br />

σθ(a/b = σ1/σ2) =σ1 + σ2<br />

(7.11)<br />

Now all these classical results are based on assumptions of plane, linear elasticity,<br />

infinite domains, and isotropic material. We use some numerical examples to study<br />

the influence of these assumptions.<br />

7.4.2 Numerical solutions for finite domains<br />

The essential influence from a finite width of a strip with a circular hole can be found<br />

in (Savin 1961) based on early results using series expansions, i.e. before the finite<br />

element method was developed. Today such results can be obtained with almost any<br />

finite element program, and graphical illustrations of stress state, strain state, energy<br />

density state give direct access to the result. With the reference model shown in<br />

figure 7.14 we discuss such results for finite domains. The sizes of the ellipse axes are<br />

specified as a percent of the corresponding lengths.<br />

The results in 7.16 illustrate the influence of the ratio of the hole size to the finite<br />

domain size. With isotropy and linear elasticity and relative stresses σ1 =3,σ2 =2<br />

together with a/b =3/2, for infinite domain from (7.11) we have σθ = 5 and constant<br />

along the hole boundary.<br />

Figure 7.16 shows four solutions corresponding to relative hole sizes of 10%, 25%,<br />

33%,and 50%, respectively. For the smallest hole model we get (σθ)max =5.18, and<br />

for the largest hole model we get (σθ)max =9.4.<br />

The general conclusion from the results in figure 7.16 is in good agreement with<br />

the results in (Savin 1961). It tells us that the analytical results for infinite domains<br />

should be used with great caution, because the ratio of hole size to domain size has<br />

an influence which should not be neglected. For the largest hole the tangential stress<br />

along the elliptical boundary varies from 4.1 to 9.4 , while for the smallest hole the<br />

variation is only from 5.01 to 5.18.


Shapes of minimum stress concentration 147<br />

50% size :4.1


Strong<br />

influence from<br />

orthotropy<br />

148 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

7.5 The 2D-hole in an orthotropic material<br />

With the increasing use of non-isotropic material, like laminated composites, there is<br />

a need for shape design also for non-isotropic material. The orthotropic case is studied<br />

in (Pedersen et al. 1992). As expected the optimal shape design is very dependent on<br />

the level of non-isotropy.<br />

Results for the case illustrated in figure 7.14 (biaxial relative external stresses<br />

σ1/σ2 =3/2), and with material data as listed in table 7.1 are determined.<br />

Material EL/10 11 Pa ET /10 11 Pa νLT GLT /10 11 Pa EL/ET<br />

Isotropic 1.0 1.0 0.3 0.3846 1.0<br />

5% fiber 3.450 1.052 0.3 0.4044 3.281<br />

10% fiber 5.900 1.109 0.3 0.4264 5.322<br />

20% fiber 10.80 1.244 0.3 0.4784 8.683<br />

30% fiber 15.70 1.416 0.3 0.5448 11.08<br />

Strong orthotropic 60.91 1.450 0.3 0.1378 42.00<br />

Table 7.1: Applied material data in shape optimal design.<br />

The optimal boundary shapes for the six materials are shown in figure 7.17, and<br />

the concentration of energy density relative to the original elliptical boundary are for<br />

the four fiber materials<br />

(uɛ)max elliptical shape<br />

(7.12)<br />

(uɛ)max optimal shape =1.8 5%fiber, 1.6 10%fiber, 1.5 20%fiber, 1.2 30%fiber<br />

We conclude that as expected the level of orthotropy has a strong influence on<br />

design. The important conclusion from this study is that we obtain constant strain<br />

energy density along the designed boundary.<br />

Another important conclusion from this study, discussed in (Pedersen et al. 1992),<br />

is that the optimal design is not very influenced by the finite element modelling. This<br />

holds for a change in element type as well as for a change in the number of elements<br />

(modelling refinement). Naturally, the stress/strain/energy values are sensitive to the<br />

accuracy of the finite element modelling, but the designed shape is not.<br />

7.6 The 2D-hole in a non-linear elastic material<br />

Here we discuss the 2D-hole with non-linear material. In a thesis by (Stokholm 1998)<br />

we can find optimal shape designs based on a constitutive behaviour that follows from


Shapes of minimum stress concentration 149<br />

<br />

Figure 7.17: <strong>Optimal</strong> boundary shapes for different ”degrees” of orthotropy.<br />

the stress energy density uɛ given by uɛ = σ n+1<br />

eff /((n +1)En ) where n ≥ 1 and σeff is<br />

the effective stress, as defined in chapter 12. The result with this material constitutive<br />

model is again uniform strain energy density along the designed boundary shape for<br />

the problem in figure 7.14. In reality the optimal shapes differ so little from the shapes<br />

obtained with linear elasticity that a more detailed numerical study is necessary to<br />

conclude that there is a difference.<br />

For the problem with the smallest hole (10 % size in figure 7.16), we study the<br />

influence from possible material non-linearity, using a power law model as described Stress<br />

in chapter 12. Concentrating on the field close to the hole boundary figure 7.18 shows<br />

the stress results for four different values of the power p =1.0, 0.75, 0.50 and 0.25<br />

(p =1/n).<br />

The main effect of the non-linearity on the stress field is a releasing effect. The<br />

maximum principal stress for the four cases is σmax =5.14, 4.69, 4.32, 3.94, respectively,<br />

and the fields also clearly show more uniformity for increasing non-linearity<br />

(power p decreasing).<br />

release<br />

The strain fields and the energy density fields show the opposite effect, which Strain<br />

is increasing localization with increasing non-linearity. Figure 7.19 shows principal<br />

strain fields with hatching proportional to strain intensity. The maximum principal<br />

strain for the four cases is ɛmax =0.0026, 0.0028, 0.0033, 0.0051, respectively.<br />

The variation of stresses and strains along the hole boundary is very limited for<br />

concentration


150 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

<br />

<br />

<br />

<br />

<br />

<br />

Figure 7.18: Stress field for solutions with non-linear elasticity.<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

Figure 7.19: Strain fields for solutions with non-linear elasticity.


Shapes of minimum stress concentration 151<br />

all the cases, but this is naturally due to the initial choice of a/b = σ1/σ2. However,<br />

it tells that the ”optimal” hole shape depends only a little on the power of the nonlinearity.<br />

A more detailed discussions of this is contained in chapter 8, section 8.4.4.<br />

7.7 The 3D-cavity<br />

For the cavity problem in 3-D the shape of an inclusion with minimum stress concentration<br />

has also been studied analytically as well as numerically. As expected<br />

in an infinite model, the optimal shape is an ellipsoid, which follows from the work<br />

of (Eshelby 1957). The von Mises reference stress is constant on the surface of the<br />

cavity. In (Dybbro and Holm 1986) the optimal ellipsoidal half axes a, b and c are<br />

given implicitly in terms of the three principal stresses, and the ellipsoidal axes are<br />

aligned with the principal stress directions. A numerical study is also performed in<br />

(Dybbro and Holm 1986) and within the accuracy of the applied finite element model<br />

the uniform energy density is found on the boundary of the shape.<br />

Also for composite materials we find three dimensional studies, as by (Vigdergauz<br />

1994). However, the number of studies are limited, and it is not expected that the<br />

analytical results for 2D-models (based on analytically determined stress fields) can<br />

be extended to 3D-models.<br />

7.8 Shape design in materials for maximum bulk<br />

modulus<br />

Chapter 8 is specifically devoted to material design and therefore is not contained<br />

in the present chapter. For more extended explanation than given in chapter 8, see<br />

(Pedersen 2001a).<br />

7.9 Stress release at a crack tip<br />

The method of hole drilling near the crack tip is often used in fatigue damage repair.<br />

In the survey by (Shin, Wang and Song 1996) comparisons with alternative methods<br />

are presented, and it is concluded that hole drilling is the most effective method. A<br />

number of experimental results are presented.<br />

From shape optimization we know that the circular shape is by no means optimal,<br />

as seen in sections 7.2 and 7.3. It is therefore important to find the shape of a hole,<br />

which in a very effective way reduces the stress concentration. This shape optimization<br />

problem is in this section restricted to holes directly at the crack tip. <strong>Optimal</strong><br />

stress release


152 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

2h<br />

000000<br />

111111<br />

000000<br />

111111<br />

000000<br />

111111<br />

000000<br />

111111<br />

000000<br />

111111<br />

000000<br />

111111<br />

00 11<br />

2a<br />

2w<br />

x2<br />

θ<br />

x1<br />

crack tip<br />

Figure 7.20: Left: the analyzed elementary case (uniform external stresses or forced<br />

displacements) with holes at the crack tips indicating the area of analysis (hatched)<br />

and the area of the graphs (cross-hatched) with hole size 1 mm. Right: the<br />

Cartesian coordinate system with origo at the crack-tip.<br />

r


Shapes of minimum stress concentration 153<br />

In detail we analyze the elementary case shown in figure 7.20. The simple<br />

parametrization, used in earlier shape optimization, is applied and it is shown that<br />

an almost uniform stress state can be obtained along the boundary of the hole. The<br />

shape of the hole is parameterized by the super-elliptic equation<br />

(x1/a) η +(x2/b) η = 1 (7.13)<br />

where the x1-direction is the direction of the crack, as shown in figure 7.20. If we use<br />

a quadratic design domain a = b (super-circle), then the only design parameter is η<br />

(assuming that the area of the hole is kept constant). Simple<br />

To make this shape optimization problem more clear, the results of a parameter<br />

study are presented in figure 7.21, which also gives the possibility for explaining the<br />

graphical display of the results from the study related to shape. We analyzed three<br />

designs corresponding to η = 2 (circular), η =2.5 (optimized) and η = 6. Values for<br />

relative maximum strain energy densities for the three designs are 1.00, 0.67 and 1.13<br />

corresponding to relative maximum tangential stresses 1.00, 0.82 and 1.06. We note<br />

that considerably better distributions of stresses are obtained with a non-circular<br />

design. (Even more uniform fields along the hole boundary can be obtained when<br />

more design parameters are included.) Figure 7.21 shows by isolines the resulting<br />

fields of energy densities. In this figure it is, by the red (dark) areas added to the<br />

boundary of the hole, illustrated how the energy density varies along the boundary<br />

of the hole (the same technique is applied in other figures of the chapter). Note that<br />

the illustrations are rotated 90 degrees relative to figure 7.20, i.e. the crack direction<br />

parametrization<br />

is now vertically downward. Illustrative<br />

example<br />

The objective of the optimal shape design in relation to cracks is not completely<br />

clear. At first we may argue that the objective should be to minimize the stress<br />

intensity factor as is done for drilled holes away from the crack tip. However, for<br />

non-sharp crack tips the interpretation of the stress intensity factor is not clear. Thus<br />

for this case we choose to minimize the maximum tangential stress at the boundary<br />

of the drilled hole. In the case of plane stress this also corresponds to minimizing Objective<br />

the maximum von Mises stress or the energy density. For the cases of non-isotropic<br />

materials it seems very relevant to minimize the maximum energy density.<br />

The optimization is performed by a simple parameter study, of the shape design<br />

for non-circular holes at the crack tip. Influences of the size of the hole, of the relative<br />

ellipse axes a/b, of the external load (stresses or forced displacements), of material<br />

non-linearity, and of material non-isotropy are discussed and a number of specific<br />

solutions are presented. The central aspect of the present study is to show that<br />

circular holes by no means are optimal, as already illustrated in the introduction.<br />

Restricting this study to holes at the crack tip we with a few (one or two) parameters<br />

optimize the shape of the hole boundary.<br />

The example shown in figure 7.20 is based on a number of assumptions and a


154 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 7.21: The energy distribution around half the holes, corresponding to η =<br />

2.0, 2.5, and 6.0. The red (dark) areas added to the boundary of the hole, illustrate<br />

how the energy density varies along the boundary of the hole.<br />

study of the sensitivity to these assumptions is needed. We discuss the influence from<br />

the external load and size of the hole, from material power law non-linearity, from<br />

material non-isotropy, and from the allowable design domain (square or rectangular)<br />

for the hole.<br />

7.9.1 Influence from the external load and size of the hole<br />

Different external loads can be examined, either given stresses (forces) or given forced<br />

displacements. As expected only little influence is seen as long as the crack is loaded<br />

mostly in mode I. This follows from the fact that the near crack tip field, as a function<br />

of the external load, only changes with a common factor (the stress intensity factor).<br />

Optimization for cracks in mode II, mode III and combined modes needs further<br />

studies.<br />

In table 7.2 we show the relative concentration of the energy density, for different<br />

sizes of the holes and for two alternative load cases. In all cases the best of the<br />

analyzed designs corresponds to a super-elliptic power of η =2.5.<br />

Relative to the size of the hole (0.5 , 1.0 , 1.5 and 2.0 mm), with the 1 mm size<br />

as reference, the relative values of maximum energy density are 1.92, 1.00, 0.68, 0.52,<br />

respectively. Thus, as expected, the larger holes give a more efficient stress release.<br />

The size of the hole is assumed to be determined by alternative considerations, and<br />

is not part of the optimization.


Shapes of minimum stress concentration 155<br />

Shape parameter η 2.0 2.25 2.5 3.0 3.5 4.0 4.5<br />

size stress load 1.0 0.857 0.840 0.876 0.934 0.994 1.05<br />

0.5 mm displ. load 1.0 0.854 0.835 0.863 0.916 0.973 1.03<br />

size stress load 1.0 0.857 0.843 0.894 0.970 1.05 1.13<br />

1.0 mm displ. load 1.0 0.852 0.833 0.875 0.944 1.02 1.09<br />

size stress load 1.0 0.856 0.842 0.895 0.974 1.06 1.14<br />

1.5 mm displ. load 1.0 0.850 0.830 0.872 0.943 1.02 1.10<br />

size stress load 1.0 0.854 0.841 0.894 0.973 1.06 1.15<br />

2.0 mm displ. load 1.0 0.847 0.826 0.866 0.937 1.01 1.09<br />

Table 7.2: Relative values of maximum energy density (for circle, η = 2 , the value is<br />

set to 1.0). Corresponding values for stress are equal to the square root of the shown<br />

values. Optimized values are shown in bold.<br />

7.9.2 Influence from material non-isotropy<br />

<strong>Optimal</strong> shape<br />

independent of<br />

load and size<br />

It is expected that anisotropic material behaviour influences the optimal shape to a<br />

large extent, see section 7.5. When the material is stiffer in the crack direction we<br />

see little influence on the optimal shape, but when it gets more flexible in the crack<br />

direction the influence is pronounced. We illustrate this in figure 7.22, where the ratio<br />

of the two moduli is set to EC/ET =0.25.<br />

We found that an energy concentration always appears for simple super-elliptic<br />

designs. With one modification function to the shape of the hole, as described in<br />

details in (Pedersen et al. 1992) and in chapter 6, we obtain designs with almost<br />

uniform energy density along the hole boundary. A look at the stress distribution in<br />

figure 7.23 for this anisotropic case, may give an understanding for the need of more<br />

advanced designs for these cases. From the results, a two parameter description of<br />

the boundary shape seems sufficient. Non-isotropy<br />

influences<br />

7.9.3 Influence from material power law non-linearity<br />

Figure 7.24 shows results based on analysis with material non-linearity. As expected<br />

from earlier results ((Pedersen 2001a)) the optimal shape of the hole is rather insensitive<br />

to the power p (p ≤ 1) of the non-linearity.<br />

We still obtain almost uniform energy density (here von Mises stress) along the<br />

boundary of the hole. The isolines show equal levels of reduced stiffness, described


156 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 7.22: Levels of energy density for EC =0.25ET corresponding to designs<br />

with η =2.5, 4.0, and 4.0. The design to the right is a two parameter design.<br />

Larger principal stresses<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 20 40 60 80 100 120 140 160 180<br />

Angle from crack direction<br />

Isotropic<br />

E(crack) = 0.25E(transverse)<br />

Figure 7.23: Larger principal stresses as function of the angle relative to the crack<br />

direction to illustrate the more complicated dependence when the crack direction is<br />

more flexible.


Shapes of minimum stress concentration 157<br />

by the factor (ɛe/ɛ0) p , where ɛe is effective strain and ɛ0 is the corresponding value<br />

that gives the transition from linearity to non-linearity. The assumptions behind the<br />

calculations leading to the results analyzed for η =2, 2.5, and 6 (as for the linear<br />

material) correspond to deformation theory with a power law of p =0.1 Relative<br />

values for these results are for the squared maximum von Mises stress 1.0, 0.98, 1.11<br />

and for the minimum stiffness reduction factors 0.305, 0.336, 0.280. Non-linear<br />

less influence<br />

Figure 7.24: In the interior: isolines of stiffness reduction based on material power<br />

law non-linearity. Along the boundary of the shapes: the variation of this stiffness<br />

reduction. The three designs correspond to η =2.0, 2.5, and 6.0.<br />

7.9.4 Influence from the allowable domain of the hole<br />

By including the elliptic half-axes a, b as design parameters we may further improve<br />

the levels of squared von Mises stress. The added condition of ab = constant practically<br />

fixes the area of the hole (the parameter η has only slight influence).<br />

In all cases we obtain almost uniform distribution along the highly stressed boundary.<br />

Figure 7.25 shows the results corresponding to a/b =1.0, 0.9, 0.8, 0.7, 0.6, and<br />

0.5, and give the resulting relative maximum values 1.0, 0.92, 0.84, 0.75, 0.66, and 0.59<br />

for the resulting maximum energy densities, and thus the super-ellipse has distinct<br />

advantages over the super-circle. The optimal super-elliptic power η changes with<br />

the ratio a/b and for the solutions referred, we got η =2.5, 2.4, 2.3, 2.15, 2.10, 1.95,<br />

respectively. Elongated<br />

The same six designs were analyzed based on a strong material non-linearity<br />

(σe/σ0) 10 and again almost constant von Mises stress is obtained along the boundary,<br />

although now decreased with almost a factor of four, relative to the results given for<br />

the linear solution. The values with the non-linear solution were 0.24, 0.23, 0.22, 0.21,<br />

0.20, 0.18, respectively. The strong non-linearity levels out the difference, but still the<br />

better designs


158 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

possibility for smaller ratios a/b gives a better solution.<br />

With illustrative examples we have shown that the stress field, at the boundary<br />

of a stress releasing hole at the crack tip, can be significantly improved. To a large<br />

extent the one parameter super-elliptic shape is able to return a field of constant<br />

tangential stress along the boundary. A shape like this diminish the possibility for<br />

further fatigue crack initiation, relative to a circular hole.<br />

Figure 7.25: Levels of squared von Mises stress for optimized super-elliptic power,<br />

when the ratio a/b are prescribed to 1.0, 0.9, 0.8 / 0.7, 0.6, and 0.5, respectively.


Chapter 8<br />

Material design<br />

Advanced materials are now used frequently in engineering design and have opened<br />

the possibility of material design. A general characteristic of these materials is that<br />

they are non-isotropic, which puts new demands on the analysis capabilities and on<br />

the optimization methods. Also for materials we can classify the design parameters<br />

in size, shape and topology. Size, shape<br />

and topology<br />

Let us imagine that a material point is described by a finite domain and then treat<br />

this domain as a continuum or structure with periodic boundary conditions. This is<br />

named a cell model. In a cell model of a material point, the size variables relate to<br />

the density distribution, and also the orientation field within the material cell model,<br />

belongs to this class.<br />

For a material cell model with a hole, the boundary of the hole is classified as a<br />

shape design parameter. The number of holes is a topology parameter, and so are the<br />

selection of actual bar elements in a truss model, or the actual beam elements in a<br />

frame model. Basically, the principal differences between designing a structure and<br />

designing a material cell model are small.<br />

Design of a material and of a structure are in principle identical problems. How- Material as<br />

a structure<br />

ever, for structures, we usually assume the external loads to be given, while in design<br />

for material moduli the loads are evaluated from forced boundary displacements and<br />

the loads are therefore design-dependent.<br />

In the present study we also focus on the influence of these different load assumptions.<br />

The influence of Poisson’s ratio and the influence of material non-linearity are<br />

shown by examples, and more general insight is obtained by sensitivity analysis.<br />

The objective can be the stiffest design, the strongest design or just a design of Different<br />

uniform energy density, say along a boundary. In an energy formulation it is proven<br />

in chapter 7 that these three objectives have the same solution, at least within the<br />

159<br />

objectives,<br />

but the<br />

same design


Theory in<br />

chapter 15<br />

<strong>Optimal</strong><br />

3D-modulus<br />

matrix<br />

160 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

limits of geometrical constraints, including the parametrization. Without involving<br />

stress/strain fields, this proof of the same solution holds for 3D-problems, for power<br />

law non-linear elasticity and for non-isotropic elasticity. With this background knowledge,<br />

the results of a number of specific influence studies are presented.<br />

The chapter is divided into five sections, the first three of which are related to<br />

size optimization. Firstly, the material design with the individual components of the<br />

constitutive matrix as design parameters is presented, resulting in analytical solutions<br />

for three–dimensional cases as well as for two-dimensional cases.<br />

Secondly, the total local relative density is treated as the only design parameter.<br />

For two-dimensional problems this is identical to the thickness optimization, treated<br />

in chapter 4. The relation to material topology optimization is exemplified.<br />

Thirdly and mainly for illustration, an analytical solution for a two parameter,<br />

orthotropic case is derived, based on general knowledge from orientational design in<br />

chapter 17 and a necessary optimality criterion, derived in chapter 14.<br />

Section 8.4 is concentrated on shape optimal design as applied to design of materials<br />

with a single hole in each cell of a periodic material model. Finally, in section<br />

8.5 comments on the active research area of material design with multi-physics constraints<br />

in addition to an objective are given. As an example such a problem could<br />

be to optimize the bulk modulus with a lower bound on the shear modulus or bounds<br />

on the thermal-elastic properties.<br />

8.1 Free material design<br />

In chapter 15 with reference to the papers (Bendsøe, Guedes, Haber, Pedersen and<br />

Taylor 1994) and (Pedersen 1998), we have obtained analytical results for the individual<br />

design of material constitutive components with the objective of maximizing<br />

the stiffness of a material (minimize the stored elastic energy). The single constraint<br />

for the optimization problem is a given total size of the constitutive matrix, measured<br />

by the Frobenius norm or by the trace norm, see also (Taylor 2000).<br />

For 3D-problems the non-dimensional part of the optimal constitutive matrix is<br />

presented in (15.10) and this result also holds for non-linear elasticity modelled by<br />

power law non-linearity. We give the results here with a dimensional factor L<br />

[L]optimal =<br />

⎡<br />

L<br />

(ɛ1 + ɛ2 + ɛ3) 2<br />

⎢<br />

⎣<br />

ɛ 2 1 ɛ1ɛ2 ɛ1ɛ3 0 0 0<br />

ɛ1ɛ2 ɛ 2 2 ɛ2ɛ3 0 0 0<br />

ɛ1ɛ3 ɛ2ɛ3 ɛ 2 3 0 0 0<br />

0 0 0 0 0 0<br />

0 0 0 0 0 0<br />

0 0 0 0 0 0<br />

⎤<br />

⎥<br />

⎦<br />

(8.1)


Material design 161<br />

For 2D-problems the corresponding optimal constitutive matrix is<br />

C<br />

[C]optimal =<br />

(ɛ1 + ɛ2) 2<br />

⎡<br />

⎣ ɛ21 ɛ1ɛ2<br />

ɛ1ɛ2<br />

ɛ<br />

0<br />

2 0<br />

2<br />

0<br />

⎤<br />

0 ⎦<br />

0<br />

(8.2)<br />

The matrix [L] presented in (8.1) and the matrix [C] presented in (8.2) both have<br />

only one non-zero eigenvalue, and thus only stiffness in relation to the specific strain<br />

condition for which they are designed. We can obtain the same effective strain and<br />

strain energy density with an isotropic, zero Poisson’s ratio material, but then the<br />

corresponding material cost is six times greater for the 3D-problem with [L] =L[I],<br />

and three times greater for the 2D-problem with [C] =C[I]. As shown in (Bendsøe<br />

et al. 1994), the zero Poisson’s ratio material is valuable in numerical calculation,<br />

<strong>Optimal</strong><br />

2D-modulus<br />

matrix<br />

because of the degeneracy of the ultimate optimal material. Examples in<br />

chapter 4<br />

In chapter 4 a number of designs are based on zero Poisson’s ratio, and thus may<br />

serve as illustrations of ultimate optimal fields for specific load conditions.<br />

8.2 Design of density distribution<br />

Design of density distribution in a material cell with the objective of minimizing<br />

the stored elastic energy, subjected only to the single constraint of given amount<br />

of cell material can be solved with the use of an optimality criterion method, and<br />

then the number of design variables is not a practical problem for the solution. Nonisotropic<br />

and non-linear elasticity do not add further to the complexity of the solution<br />

procedure, because the optimality criterion is in all cases uniform energy density, as Theory in<br />

proven in chapter 12. To be specific our objective is to<br />

chapter 12<br />

Minimize Uɛ = <br />

(uɛ)eρe<br />

(8.3)<br />

where Uɛ is the total strain energy, (uɛ)e is the specific strain energy that multiplied<br />

with the density ρe gives the strain energy (Uɛ)e in design domain e. The total amount<br />

of density (resource) ¯ R is given, i.e. the optimization constraint is<br />

R := <br />

ρe = ¯ R (8.4)<br />

e<br />

We notice directly, that the optimality criterion of constant specific strain energy<br />

(uɛ)e =ūɛ with the results of chapter 12 imply that the necessary condition of sta-<br />

e<br />

tionarity dUɛ = 0 is fulfilled with the constraint dR = 0. To actually find a density Procedure<br />

distribution that satisfies the optimality criterion (uɛ)e =ūɛ we can use the heuristic<br />

and examples<br />

in chapter 4


Material,<br />

yes or no ?<br />

Modified<br />

optimality<br />

criterion<br />

List of<br />

practical<br />

problems<br />

162 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

procedure, described in section 4.11 and used in chapter 4, again illustrating that design<br />

of structural thickness distribution and design of material cell density distribution<br />

are almost identical problems.<br />

8.2.1 ”Black and white” designs<br />

This book does not cover in detail the very active research area of the topology<br />

optimization method, where topology design parameters are converted into size design<br />

parameters; reference is given to the book of (Bendsøe and Sigmund 2003).<br />

However, a comment in relation to the present description is given. To obtain a<br />

”black and white” design, the intermediate densities must be penalized. A very used<br />

technique to obtain this is to modify the objective (8.3) to<br />

Minimize Uɛ = <br />

(uɛ)e(ρe) q<br />

(8.5)<br />

where q is some power (often q = 3 is chosen). The sensitivity of this objective with<br />

the results from chapter 12 is<br />

dUɛ<br />

= −<br />

dρe<br />

1<br />

<br />

dUɛ<br />

= −<br />

p dρe<br />

1<br />

q−1<br />

(uɛ)eq(ρe) (8.6)<br />

p<br />

e<br />

fixed strains<br />

(p = 1 for linear elasticity, but the final result is not influenced by p). We find that<br />

if (uɛ)e(ρe) q−1 is constant, then the stationarity condition is fulfilled. This follows<br />

directly from the condition of proportional gradients (proven in chapter 14), because<br />

from the unchanged resource condition (8.4) follows dR/dρe =1.<br />

The literature with methods to solve this modified problem is very rich, both with<br />

respect to necessary specific methods and obtained designs. In fact for solution, the<br />

methods of mathematical programming are used more than the optimality criteria<br />

methods. A number of practical problems occur in relation to the search for these<br />

”black and white” designs, and in fact some of these are not only related to design<br />

for ”black and white”. We mention:<br />

• non-unique solutions and local optimal solutions<br />

• solutions with checkerboard patterns in the finite element models<br />

• solutions depending on mesh<br />

• solutions that violate limits of the actual model<br />

• slow convergence in objective and/or design parameters<br />

• solutions depending on stopping criterion for design iterations


Material design 163<br />

In addition to these difficulties in obtaining a specific solution, it should be noted<br />

that an optimized design may be strongly depending on the following aspects:<br />

• total relative volume density<br />

• modelling of loads and boundary conditions, especially with relation to rather<br />

concentrated loads and reactions<br />

• finite element modelling, i.e. element type chosen, mesh quality, and number<br />

of degrees of freedom<br />

• penalizing power and possible continuation approach, i.e. the strategy of penalizing<br />

• approach for smoothing (filtering) of design and/or sensitivities<br />

• approach for allowable design changes, such as move-limits<br />

• approach for redesign, such as optimality criterion or mathematical programming<br />

with specific choice of linear programming (LP) or method of moving<br />

asymptotes (MMA) by (Svanberg 1987)<br />

• initial design<br />

• neglected modelling<br />

• lower bound on density<br />

In relation to all these aspects, see (Bendsøe and Sigmund 2003) and the references<br />

in there as well as the two theses (Bendsøe 1995) and (Sigmund 2001).<br />

8.2.2 An illustrative example<br />

Without going into the discussion of all the mentioned aspects of the topology optimization<br />

methods, we solve in a very simple manner with an optimality criterion, a<br />

specific example and discuss the resulting designs with and without penalized intermediate<br />

densities.<br />

Figures 8.1, 8.2, 8.3, and 8.4 and table 8.1 show the result of the specific Quarter cell<br />

example<br />

problem, solved with a simple heuristic approach, based on the optimality criterion<br />

(uɛ)e(ρe) q−1 = constant, comparing solutions with q = 3 and q = 1, i.e. with and<br />

without penalizing. To be specific we use in all cases 16 iterations with q = 1, followed<br />

by 16 iterations with q = 3, and the solutions are rather stable after these 16<br />

iterations.<br />

We find a strong influence of the penalization, and the ”black and white” designs<br />

are obtained on the cost of lower stiffness and higher stress concentration. It must be


Penalized and<br />

not-penalized<br />

164 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

noted that the total volume is kept the same in all the examples , and furthermore<br />

the stiffness analyses are based on the actual densities, not the penalized densities<br />

which are only used for the redesign procedure. (This is not the case in most topology<br />

optimization programs). Also it should be mentioned that densities are not scaled to<br />

the range of zero to one, but a common scaling factor does not change the penalization.<br />

Figure 8.1: Results for 80% material usage. Left: square root of energy densities. Right:<br />

<strong>Optimal</strong> designs. Upper not penalized with 0.001 ≤ ρe ≤ 1.25 (corresponding to 80% with<br />

1.25·0.8 = 1), and lower penalized to ”black and white”. The colour code is red for minimum,<br />

green for medium, blue/violet for maximum, and white indicates ”no” material.


Material design 165<br />

Figure 8.2: Results for 50% material usage. Left: square root of energy densities. Right:<br />

<strong>Optimal</strong> designs. Upper not penalized with 0.001 ≤ ρe ≤ 2 (50%) , and lower penalized to<br />

”black and white”. The colour code is red for minimum, green for medium, blue/violet for<br />

maximum, and white indicates ”no” material.


166 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 8.3: Results for 25% material usage. Left: square root of energy densities. Right:<br />

<strong>Optimal</strong> designs. Upper not penalized with 0.001 ≤ ρe ≤ 4 (25%) , and lower penalized to<br />

”black and white”. The colour code is red for minimum, green for medium, blue/violet for<br />

maximum, and white indicates ”no” material.


Material design 167<br />

Figure 8.4: Results for 12.5% material usage. Left: square root of energy densities. Right:<br />

<strong>Optimal</strong> designs. Upper not penalized with 0.001 ≤ ρe ≤ 8 (12.5%) , and lower penalized to<br />

”black and white”. The colour code is red for minimum, green for medium, blue/violet for<br />

maximum, and white indicates ”no” material.


168 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Constraints Design umean umax Figures Note<br />

Uniform 1.00 22.7 all upper left a)<br />

0.001 ≤ ρe ≤ 1.25 Optimized 0.86 14.5 8.1 upper right b)<br />

q =1 not penalized<br />

0.001 ≤ ρe ≤ 1.25 Optimized 0.90 14.5 8.1 lower c)<br />

q =3and 80% penalized<br />

0.001 ≤ ρe ≤ 2.00 Optimized 0.71 5.43 8.2 upper right d)<br />

q =1 not penalized<br />

0.001 ≤ ρe ≤ 2.00 Optimized 0.84 12.3 8.2 lower e)<br />

q =3and 50% penalized<br />

0.001 ≤ ρe ≤ 4.00 Optimized 0.65 1.28 8.3 upper right f)<br />

q =1 not penalized<br />

0.001 ≤ ρe ≤ 4.00 Optimized 1.02 36.2 8.3 lower g)<br />

q =3and 25% penalized<br />

0.001 ≤ ρe ≤ 8.00 Optimized 0.65 0.68 8.4 upper right h)<br />

q =1 not penalized<br />

0.001 ≤ ρe ≤ 8.00 Optimized 2.44 129.3 8.4 lower i)<br />

q =3and 12.5% penalized<br />

Table 8.1: Table of relative values for the solutions shown in figures 8.1, 8.2, 8.3, and<br />

8.4.


Material design 169<br />

Figures 8.1, 8.2, 8.3, and 8.4 show the results of optimal design with given boundary<br />

stresses, and in each figure a chosen relative maximum density 1.25, 2, 4, 8,<br />

respectively, corresponding for the penalized cases to total relative density of 80%,<br />

50%, 25%, and 12.5%. Table 8.1 gives the relative values for compliance (measured<br />

by umean) and for stress concentration (measured by umax). We comment the results,<br />

the figures and the table by the following notes:<br />

The finite element model has 19600 constant stress triangular elements and totally<br />

19882 degrees of freedom with a bandwidth of 284. Discussion<br />

a) The uniform design is used as reference for all the cases, and for this we find a<br />

stress concentration factor equal to 22.7, active in the loaded corner as seen in the upper<br />

left picture of all the figures 8.1, 8.2, 8.3, and 8.4, with a colour scale proportional<br />

to the square root of element energy density, (i.e. proportional to stresses).<br />

b) For maximum density only 25% higher than the mean densitiy not much can<br />

be obtained by optimal design. Figure 8.1 shows this. Accepting intermediate densities<br />

(”gray” design) we improve the compliance from 1.00 to 0.86 and the stress<br />

concentration from 22.7 to 14.5.<br />

c) The same case as in b) but now penalized to obtain ”black and white” design.<br />

The lower part of figure 8.1 shows this design to the right, with its resulting distribution<br />

of energy density to the left. The numbers are almost as for the ”gray” design<br />

with compliance improved to 0.90 and stress concentration equal to 14.5.<br />

d) For maximum density being the double of the mean density the optimal designs<br />

are presented in figure 8.2. Accepting intermediate densities (”gray” design) we<br />

improve the compliance from 1.00 to 0.71 and especially the stress concentration is<br />

improved from 22.7 to 5.43.<br />

e) The same case as in d) but now penalized to obtain ”black and white” design.<br />

The lower part of figure 8.2 shows this design to the right, with its resulting distribution<br />

of energy density to the left. The results are now different from those for the<br />

”gray” design with compliance only improved to 0.84 and stress concentration still<br />

rather high and equal to 12.3.<br />

f) For maximum density equal to four times the mean density the optimal designs<br />

are presented in figure 8.3. Accepting intermediate densities (”gray” design)<br />

we improve the compliance almost to the limit, from 1.00 to 0.65 and the stress<br />

concentration is then controlled from 22.7 to 1.28, but still active.<br />

g) The same case as in f) but now penalized to obtain ”black and white” design.<br />

The lower part of figure 8.3 shows this design to the right, with its resulting<br />

distribution of energy density to the left. The compliance is no longer improved, but<br />

almost unchanged from 1.00 to 1.02, and the stress concentration is now increased<br />

to 36.2. However, this is partly due to the finite element modelling of the resulting<br />

complicated design and boundary smoothing could be very effective. So the design in<br />

8.3 lower right must be post-processed.<br />

of results


170 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

h) For maximum density equal to eight times the mean density the optimal designs<br />

are presented in figure 8.4. Accepting intermediate densities (”gray” design)<br />

we improve the compliance almost to the limit, from 1.00 to 0.65 and the stress concentration<br />

is then almost completely eliminated from 22.7 to 0.68. We know from<br />

chapter 14 that without size constraints the optimal design has uniform energy density,<br />

and see that this is almost obtained. The practical question then is related to<br />

the modelling accuracy with high density gradients at the loaded corner, as seen in<br />

figure 8.4 upper right.<br />

i) The same case as in h) but now penalized to obtain ”black and white” design.<br />

The lower part of figure 8.4 shows this design to the right, with its resulting distribution<br />

of energy density to the left. The compliance is now changed from 1.00 to 2.44,<br />

and the calculated stress concentration is 129.3. However, this is partly due to the<br />

finite element modelling of the resulting complicated design.<br />

Thus, for the examples with ”black and white” solutions we need post-processing<br />

with e.g. shape design variables. We return to this problem in section 8.4.<br />

8.3 Design of a two parameter cell model<br />

The design of material with total freedom in section 8.1 and the design with only<br />

a single parameter (the density) in section 8.2 may be seen as extreme cases. For<br />

illustration we go through the two-parameter case, for which an analytical solution is<br />

Orthotropic possible to obtain, but is much more difficult to derive.<br />

material Let us focus only on the objective of maximized stiffness (minimized strain energy)<br />

based on orthotropic materials classified as low shear modulus materials. In reality<br />

we go directly to the material model from (Bendsøe 1991), as illustrated in figure 8.5.<br />

The total, relative volume densities (0


Material design 171<br />

Figure 8.5: A material cell model with two directional densities.<br />

From (Bendsøe 1991) we take the orthotropic constitutive matrix, expressed in<br />

known modulus E and Poisson’s ratio ν and in the directional parameters µ, ξ as<br />

[C] = Eξ<br />

α−β2 <br />

1<br />

β<br />

β<br />

α<br />

<br />

⇒ [C] −1 = 1<br />

<br />

Eξ<br />

α<br />

− β<br />

−β<br />

1<br />

<br />

(8.9)<br />

with the definitions α := ξ2 + ξ(1 − ξ)/µ and β := ξν.<br />

The shear modulus is not involved, and we assume the major principal axis chosen<br />

by<br />

Constitutive<br />

matrix<br />

αξ/(1 + ξ)<br />

The conditions of positive definite and limited [C] are satisfied for<br />

(8.10)<br />

0


172 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

This is the result from the orientational optimization. Next, in the same way as<br />

in the thickness optimization, we have uniform energy density on the global level to<br />

determine the global ρ field. Also on the local level we must have the same energy<br />

density in the µ and ξ ”fibers”. The statement of this local optimization problem for<br />

given total density ρ is:<br />

Determine the directional densities µ, ξ that<br />

Minimize Uɛ = uɛ ˜ Vρ or<br />

Minimize Uσ = uσ ˜ Vρ<br />

subject to the constraint : ρ = µ + ξ − µξ (8.14)<br />

The condition of proportional gradients from chapter 14 gives (after extended calculations<br />

for the uɛ formulation) the analytical result<br />

µ =<br />

ρ<br />

1+|ζ|<br />

ξ =<br />

|ζ|ρ<br />

1+|ζ|−ρ<br />

in terms of the total relative density ρ and stress ratio ζ := σ2/σ1.<br />

8.4 Design of material with a single hole<br />

(8.15)<br />

The problem of designing the shape of a single hole in order to maximize the homog-<br />

Optimize<br />

enized bulk modulus is given special attention and a very simple parametrization<br />

bulk modulus is chosen. The same parametrization is used in chapter 7. Agreement with earlier<br />

results of (Vigdergauz 1986) and with Hashin-Shtrikman bounds, see (Grabovsky and<br />

Kohn 1995), is obtained. Independence of Poisson’s ratio on the optimal shapes is<br />

numerically found for these square-symmetric problems. As is shown in section 8.4.3,<br />

this independence is generally not the case.<br />

A very simple optimality criterion, derived in section 14.5, gives the background<br />

for an efficient optimization procedure. Using this tool the goal of the present section<br />

Influence<br />

is to study the influence of the following aspects:<br />

studies • Influence from volume constraint.<br />

• Influence from boundary conditions.<br />

• Influence from Poisson’s ratio.<br />

• Influence from material non-linearity.<br />

Without involving optimization it is useful first to study solutions where the shape<br />

of the model in figure 8.6 is determined only by the parameter η (power of the superelliptic<br />

function). We assume a symmetric model A = B, which with a constant solid<br />

area gives the values α = β for a chosen parameter η.


Material design 173<br />

Figure 8.6: The simplified shape parametrization used also for material design.<br />

In summary, the main conclusions of the parameter study in this section are:<br />

• The compliance of a material cell is rather insensitive to a variation<br />

of the shape.<br />

• In contrast the energy density concentration is very sensitive to a variation of<br />

the shape.<br />

• If a shape of constant energy density along the shape boundary is found, then<br />

the compliance is stationary. Results and<br />

conclusions<br />

• A design for stationary compliance may return high energy density concentration<br />

if geometrical constraints (parametrization) are a controlling parameter.<br />

• In a finite domain problem, load applied by stress and load applied by forced<br />

displacements give different optimal shapes.<br />

• Especially for dense structures with given forced displacements, different Poisson’s<br />

ratios of an isotropic material give in genera different optimal shapes.<br />

However, for given applied stresses the optimal shape is proven to be independent.<br />

• Also the non-linearity of a material is reflected in the optimal design, but the<br />

influence is not very strong.<br />

8.4.1 Influence from the volume constraint<br />

Before the detailed results from the study of influence of boundary conditions, Poisson’s<br />

ratio and material non-linearity are presented, we show the influence from the


Hole designs<br />

with resulting<br />

bulk modulus<br />

<strong>Designs</strong><br />

independent of<br />

Poisson’s ratio<br />

174 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

volume constraint, i.e. the total relative density. With reference to figure 8.6 we<br />

define the relative density as the ratio of the area of the solid part (hatched) and the<br />

total area. The asymptotic case of ρ → 1 is a small hole for which classical analytical<br />

design results are available. The other asymptotic case of ρ → 0 gives a frame<br />

structure and we concentrate on the intermediate cases of say 0.2


Material design 175<br />

which with a constant solid area gives the values α = β for a chosen parameter η.<br />

Three different load cases are analyzed:<br />

• forced boundary displacements at x1 = A, x2 = B, corresponding to pure mean<br />

dilatation ¯ɛx2 =¯ɛx1 =0.001 Different<br />

load cases<br />

• forced boundary stresses at x1 = A, x2 = B, corresponding to pure mean hydrostatic<br />

pressure ¯σx2 =¯σx1 =0.0006 (modulus of elasticity C1111 =1)<br />

• a mixed problem with forced boundary displacements at x1 = A, corresponding<br />

to ¯ɛx1 =0.001, and forced boundary stresses at x2 = B, corresponding ¯σx2 =<br />

0.0006<br />

Note that for an infinitesimally small hole, these three problems are identical. Equal cases for<br />

small holes<br />

With zero Poisson’s ratio the Hashin-Shtrikman bound for the bulk modulus with<br />

relative solid area ρ =0.75 is 0.3 and thus ¯σ/¯ɛ = σx1 /2ɛx1 =0.3 gives ¯σx1 =0.6ɛx1 ,<br />

see (Pedersen 2001a).<br />

The results of the analyses are shown in figure 8.8 in terms of the mean strain<br />

energy densities in the solid areas (macroscopic bulk modulus = (1/2)(uɛ)mean · ρ =<br />

(3/8)(uɛ)mean). Note, how insensitive the compliance (proportional to (uɛ)mean) is<br />

in this rather large design domain 1.7


176 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 8.8: Results of one-parameter studies, with mean strain energy density<br />

umean = umean(η) as a function of the shape parameter η. The value from the<br />

Hashin-Shtrikman bound is umean =0.8.


Material design 177<br />

Figure 8.9: Starting designs in upper row with η = 2 and optimal solutions in<br />

lower row with the left column: showing forced displacements with η =2.26, middle<br />

column: showing forced stresses with η =1.84, right column: showing mixed loads<br />

with η =1.96 and α =0.92β. Energy densities along the boundaries in black and<br />

isolines for the larger principal stresses in the interiors.


Influence<br />

on designs<br />

Cases of no<br />

influence<br />

Sensitivity<br />

analysis<br />

178 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

8.4.3 Isotropic cases with influence from Poisson’s ratio<br />

The studies of section 8.4.2 were based on isotropic, zero Poisson’s ratio linear elasticity.<br />

This was done to put the focus on the different boundary conditions. In<br />

earlier optimization it was found that the optimal shape depends strongly on the<br />

non-isotropy, see chapter 7.<br />

However, weak dependence on Poisson’s ratio and on the non-linearity was found.<br />

In the present section we go deeper into the influence of Poisson’s ratio and focus<br />

on the influence on the optimal shape design, rather than on the resulting stresses,<br />

strains and energies. It is seen that for non-symmetric problems ¯ɛx1<br />

= ¯ɛx2 the<br />

influence cannot always be neglected.<br />

In the three papers (Thorpe and Jasiuk 1992), (Cherkaev, Lurie and Milton 1992)<br />

and (Christensen 1993) the homogenized modulus of elasticity is proven to be in-<br />

dependent of the Poisson’s ration ν for the basic material of an isotropic composite,<br />

with critical comments found in (Zheng and Hwang 1997). In papers by Vigdergauz,<br />

as referred in (Grabovsky and Kohn 1995), the shape for maximum bulk modulus<br />

is also found independent of ν although the bulk modulus itself, as shown in figure<br />

8.7, depends strongly on ν, which means that we have to be careful with the term<br />

independent.<br />

For more general problems with given forced displacements the objective and also<br />

the optimal shape depend on ν, and the intention of this section is to show this<br />

by an example. Before doing this we through sensitivity analysis get more detailed<br />

information and separate this influence of ν into the local influence when the stress or<br />

strain fields are kept unchanged and the more global influence from a redistribution<br />

of these fields. From this it follows that the three sources of sensitivity are:<br />

• a general scaling factor<br />

• a further local change for fixed strains or fixed stresses<br />

• a global redistribution of displacements, strains and stresses<br />

For an isotropic material the energy densities can be evaluated from the principal<br />

stresses σ1,σ2 and/or the principal strains ɛ1,ɛ2.<br />

constitutive relations<br />

Thus, we only need part of the<br />

<br />

σ1<br />

<br />

C1111<br />

=<br />

C1122<br />

<br />

ɛ1<br />

<br />

(8.16)<br />

σ2<br />

C1122 C1111<br />

With the same C1111 all over the cell model this is only a scaling factor and the<br />

Material<br />

ratio<br />

parameter µ := C1122/C1111 with − 1


Material design 179<br />

is thus the important parameter. With a plane stress assumption we have µ = ν<br />

(Poisson’s ratio) and with a plane strain assumption we have µ = ν/(1 − ν). Also the<br />

ratio of principal stresses and the ratio of principal strains are important parameters<br />

and we define as done in relation to (8.12) and (8.13) Stress/strain<br />

parameters<br />

γ := ɛ2/ɛ1 with − 1 ≤ γ ≤ 1, and ζ := σ2/σ1 with − 1 ≤ ζ ≤ 1 (8.18)<br />

thus assuming both |ɛ1| ≥|ɛ2| and |σ1| ≥|σ2| (ordering). The constitutive behaviour<br />

(14.3) is rewritten to<br />

<br />

1<br />

1 µ 1<br />

σ1 = C1111ɛ1<br />

(8.19)<br />

ζ<br />

µ 1 γ<br />

It follows from (8.19) that we can express ζ in µ and λ or γ in µ and ζ<br />

ζ =(γ + µ)/(1 + µγ), γ =(ζ − µ)/(1 − µζ) (8.20)<br />

which is always valid as µγ < 1 and µζ < 1. The conditions in (8.18) are satisfied,<br />

i.e. the numerically larger ɛ1 and σ1 are aligned. Note that γ = 0 implies ζ = µ and<br />

ζ = 0 implies γ = −µ. Aligned larger<br />

stress/strain<br />

For linear elasticity the strain energy density uɛ and the stress energy density uσ<br />

have the same value, which is determined by<br />

uɛ = uσ = 1<br />

2 (σ1ɛ1 + σ2ɛ2) = 1<br />

2 σ1ɛ1(1 + γζ) (8.21)<br />

Using relations (8.19) and (8.20) we can express this as<br />

and as<br />

uσ =<br />

uɛ = C1111ɛ 2 1(1 + 2µγ + γ 2 ) (8.22)<br />

σ 2 1<br />

2C1111(1 − µ 2 ) (1 − 2µζ + ζ2 ) (8.23)<br />

We have the global scaling factors C1111 and (1 − µ 2 ). In fact, for plane stress<br />

models we have C1111(1 − µ 2 )=(E/(1 − ν 2 ))(1 − ν 2 )=E and therefore no scaling.<br />

The main conclusion from (8.22) and (8.23) is that even in a fixed strain field or a<br />

fixed stress field the energy density depends, in addition to the global scaling factor,<br />

on the ratio µ = C1122/C1111. We have no influence with unidirectional strain, γ =0<br />

or with unidirectional stress, ζ =0.<br />

At the free boundary to be designed we have unidirectional stress. Thus if we<br />

can prove the stress field to be unchanged, then the optimal shape is independent of<br />

µ, i.e. independent of Poisson’s ratio. With given boundary stresses in equilibrium Unchanged<br />

with<br />

stress loads


Sensitivity<br />

analysis<br />

Changed<br />

designs<br />

with forced<br />

displacements<br />

180 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

this is true for 2D linear elastic problems, see (Muskhelishvili 1953), pp. 160-161.<br />

Numerical results for our actual problems confirm this.<br />

However, with given boundary displacements the optimal design depends on µ. To<br />

get some information also for this case we set up a global sensitivity analysis, primarily<br />

with respect to the nodal displacements in a finite element model. In addition to the<br />

ratio µ = C1122/C1111 we have for isotropic models C1212/C1111 =(1− µ)/2. With<br />

this being the case for the total model, the total stiffness matrix is linear in µ and we<br />

have<br />

[S]{D} =([S0]+µ[S1]){D} = {A} (8.24)<br />

where {D} are the nodal displacements and {A} the nodal loads obtained by either<br />

prescribed boundary stresses or by prescribed boundary displacements. For {A} not<br />

independent on µ we must solve<br />

[S] ∂{D}<br />

∂µ = −[S1]{D} (8.25)<br />

and for {A} = −([S0]+µ[S1]){Df }, where {Df } is the vector of forced displacements,<br />

we must solve<br />

[S] ∂{D}<br />

∂µ = −[S1]({D} + {Df }) (8.26)<br />

to get ∂{D}/∂µ and from this the sensitivities of the strain and the stress fields.<br />

For the symmetric problems in section 8.4.2 we have seen independence on µ,<br />

but for non-symmetric problems this is not the case. Figure 8.10 shows rather strong<br />

influence for an actual problem with only weak non-symmetry described by ¯ɛ1 =1.2¯ɛ2.<br />

8.4.4 Influence from non-linear elasticity<br />

In the sensitivity analysis relative to Poisson’s ratio we have identified three sources<br />

of possible influence from the ratio µ := C1122/C1111. On the local level the scaling<br />

influence through C1111 and (1 − µ 2 ) and the further influence depending on the ratio<br />

of principal stresses or strains according to (8.22) and (8.23). On the global level we<br />

have the influence through a redistribution of the stress/strain fields. We perform a<br />

somewhat similar analysis for non-linear elasticity described by σeff = Eɛ p<br />

, where<br />

σeff , ɛeff are the effective stress, effective strain defined relative to energy densities,<br />

see chapter 12. In general we have also for non-linear elasticity<br />

eff<br />

uɛ = uσ =(σ1ɛ1 + σ2ɛ2) (8.27)<br />

and with the relation uσ = puɛ, proven in chapter 12 for power law non-linearity, we<br />

get the generalization of (8.22) and (8.23) to be<br />

uɛ = 1<br />

1+p C1111ɛ 2 1(1 + 2µγ + γ 2 ) (8.28)


Material design 181<br />

Figure 8.10: Influence of Poisson’s ratio for isotropic material, non-symmetric displaced<br />

¯ɛ1 =1.2¯ɛ2. Strain energy densities at boundaries in black and isolines for the<br />

larger principal stresses in the interiors.


Sensitivity<br />

analysis<br />

Unchanged with<br />

size design<br />

Changed with<br />

shape design<br />

182 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

and<br />

σ 2 1<br />

uσ = p<br />

1+p 2C1111(1 − µ 2 ) (1 − 2µζ + ζ2 ) (8.29)<br />

On the local level we have directly from (8.28) and (8.29) that in a fixed strain field<br />

or in a fixed stress field we get proportional changes with respect to the non-linearity<br />

parameter p<br />

∂uɛ<br />

= −<br />

∂p fixed strains<br />

1 ∂uσ<br />

uɛ, =<br />

1+p ∂p fixed stresses<br />

1<br />

p(1 + p) uσ<br />

(8.30)<br />

Thus, this local influence from the non-linearity parameter p do not change a characteristic<br />

of constant energy density and the optimality criterion is still satisfied.<br />

We then perform a sensitivity analysis to get information about the global change<br />

of the displacement field d{D}/dp, strain field and stress field. The finite element<br />

formulation for this is<br />

[S] d{D}<br />

dµ<br />

= d{A}<br />

dµ<br />

d[S]<br />

− {D} (8.31)<br />

dµ<br />

Let us assume a constant stress, strain, energy density element and that the strain<br />

state in element e is of magnitude ɛeff that relates to the actual strain energy density<br />

uɛ by<br />

uɛ = 1<br />

1+p Eɛ1+p<br />

eff<br />

(8.32)<br />

where E is a fixed modulus of elasticity, see chapter 12 for discussion of the definition<br />

of ɛeff . The element stiffness matrix [Se] can be written with relative strain ˜ɛ :=<br />

ɛeff /ɛref, where ɛref is a pre-defined reference strain<br />

[Se] =˜ɛ p−1 [ ˜ Se] (8.33)<br />

with [ ˜ Se] being independent of p, and then we get<br />

d[Se]<br />

dp = ln(˜ɛ)˜ɛp−1 [ ˜ Se] =ln(˜ɛ)[Se] (8.34)<br />

Thus, for elements of equal energy density (equal effective strains ɛeff )weget<br />

proportional change of the stiffness matrices. From this follows the result, seen in<br />

(Pedersen and Taylor 1993), that in thickness design the optimal design does not<br />

depend on the non-linearity p.<br />

However, for shape design the effective strain ɛeff is only constant along the<br />

boundary and thus a redistribution of displacements, strains and stresses takes place.<br />

As we see in the example in figure 8.11, the optimal shape only changes slightly,<br />

probably due to the fact that the change in ɛeff is only in the direction orthogonal<br />

to the shape.


Material design 183<br />

Figure 8.11: Change in optimal design for increasing non-linearity p =<br />

0.75, 0.50, 0.25. Energy densities at boundaries in black and isolines for the larger<br />

principal stresses in the interiors.


184 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

8.5 Design with multi-physics constraints<br />

See (Bendsøe and Sigmund 2003) for interesting examples and references.


Chapter 9<br />

Bone mechanics and damage<br />

evolution<br />

9.1 Introduction<br />

The structure of bone and of the bone material itself is complicated. Furthermore,<br />

the bone material evolves and adapts to the actual load conditions. Much insight may<br />

be gained by the mathematical modelling of bone mechanics even though it includes<br />

rough approximations. In this chapter, results related to the research on optimal<br />

structural and material design are communicated, although like in general for this<br />

book, mostly limited to an ”in house” review.<br />

9.1.1 Points of view<br />

The complexity is very well described in (Huiskes and Kaastad 2000), a paper which<br />

describes formulations, computations, and experiments together with expressing a<br />

number of points of view. Also the present chapter starts with some points of view<br />

of the author:<br />

• Bone-mechanics is a very important subject which must be subjected to intensive<br />

research for many years to come. We should attack the problems with all<br />

the up-to-date tools available, i.e. extensive FEA (finite element analysis), advanced<br />

experimental techniques at the micro-level, optimization formulations<br />

and tools, identification techniques, and long term observations at the macrolevel.<br />

There is no meaning in abandoning some of these tools.<br />

185<br />

Material and<br />

structure


3D analysis<br />

Identification<br />

Formulation<br />

186 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

• The modelling for FEA must be three dimensional, due to the fact that two<br />

dimensional models are too restrictive. In the light of improved computer<br />

facilities extensive 3D modelling is possible.<br />

• Static FEA are necessary to obtain homogenized material quantities and study<br />

the influence of micro-structures.<br />

• Long term simulations by FEA are necessary to model the re-modelling of bone,<br />

to be tested against experimental observations. With this tool and improved<br />

modelling, the response to change in loads or reactions to implants might then<br />

be estimated.<br />

• In relation to these simulations and observations, interesting identification (estimation)<br />

of parameters can be solved using optimization procedures. Find the<br />

parameters (the model) that minimize the difference between observations and<br />

simulations.<br />

in time • Like 3D is a must, so is a formulation in time. Loads and bone are time<br />

dependent, and it make little sense to deal with combinations of loads. At any<br />

time there is one and only one load situation. The re-modelling at any time<br />

may however also depend on the history (recent history) of loads. In a similar<br />

Energy density<br />

Phenomenological<br />

manner the sensing may also be non-local. However, models for time history<br />

influence as well as space non-local influence can be formulated.<br />

• With non-isotropy, a local measure of energy density must be more reasonable<br />

than a measure of strain or stress. Note, that stiffness and strength are independent<br />

quantities, although sometimes good stiffness also give good strength.<br />

However, in certain cases good strength can be obtained by actually removing<br />

material and decreasing the stiffness.<br />

• Optimization formulations in bone-mechanics has a broad range of applications,<br />

in particular in relation to re-modelling. So at any time step in a simulation, an<br />

optimization model is formulated and solved. Note, that advanced optimization<br />

models must include a number of constraints, and learning by doing the models<br />

must be improved.<br />

re-modelling • These phenomenological models, that are tested against observations, will also<br />

be influenced by the increasing understandings obtained on the micro levels,<br />

Shape and size<br />

from FEA analysis and experiments.<br />

optimization • <strong>Optimal</strong> re-modelling can be formulated as shape optimization (surface) as well<br />

as size optimization (density), and combination of these. In addition the orientational<br />

re-modelling is important, either in a macro sense controlled by<br />

principal stresses or in a micro sense by changes in micro-structure, which give<br />

changes in homogenization results.


Bone mechanics and damage evolution 187<br />

9.1.2 Contents of the chapter<br />

Analysis, evolution and layout in bone mechanics are described as seen from an optimal<br />

design and solid mechanics perspective. Non-isotropic behaviour and three<br />

dimensional modelling are necessary and focus is therefore put on the aspects of dealing<br />

with a large number of material parameters. The tools of homogenization, inverse<br />

homogenization and some basic results from optimal structural design are described. Non-isotropic<br />

Bone structures/materials are not isotropic, and thus the non-isotropic nature<br />

should be properly modelled. This holds for the bone stiffness as well as for the<br />

bone strength and the bone evolution. There is also evidence that three dimensional<br />

modelling is necessary, so just for stiffness we need 21 material parameters. The<br />

homogenization method is a method to obtain these parameters, the method is based<br />

on a detailed description of a substructure or a material point model.<br />

Normally, an actual bone shows a very specific layout in terms of distribution<br />

of densities and material orientations. Can we postulate an objective that leads to<br />

these layouts, such as the postulate of maximum stiffness for minimum material? In<br />

general the answer is no, but we can learn from such a postulate. The example of<br />

a femur bone structure from the thesis by (Bagge 1999) and the paper by (Bagge A thesis<br />

2000) is discussed. It is found that the uncertainty in knowledge about actual loads<br />

(force transfer) is also a critical issue in bone mechanics.<br />

The load conditions on a bone change with time due to the different activities<br />

of the total body. The bone adapts to these changes, and we need a model for this Adaptation<br />

adaptation. Several models are suggested, and we will her describe a model based<br />

on stimulus from an energy density situation. The energy density situation is time<br />

dependent and the modelling may include some form of memory.<br />

With reliable tools for the simulation of bone evolution, we can then consider<br />

the important problems related to design of the implants that replace bones, e.g. Implants<br />

a hip prosthesis. Traditionally, these implants are made of homogeneous, isotropic<br />

materials, in large contrast to the bone structure which they substitute. Mathematical<br />

simulations can show the advantages of alternative implant design, although from a<br />

practical point of view, it may be difficult to introduce such implants. A more feasible<br />

design may be obtained by pure shape optimization of the implants already in use.<br />

9.2 Bone layouts from an optimization process<br />

In a discretized model of a bone, we assume that each element in a finite element<br />

model has stiffness and strength given by a restricted number of parameters, e.g. the<br />

density µe and the orientation(s). The principal (extreme) material directions align Aligned<br />

directly with the principal stress directions, and we can then obtain the stiffness as a<br />

function (interpolation) of the density µe. When this information is available, we can<br />

orientations


188 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

iteratively determine the material distribution that will maximize the total stiffness<br />

(minimize the stored elastic energy) for a given total amount of material. For such a<br />

procedure to be effective, the experience obtained through the extensive research in<br />

Proximal<br />

the area of topology optimization is needed. See (Bendsøe and Sigmund 2003).<br />

femur For a proximal femur, it is shown in (Bagge 1999) and (Bagge 2000) that realistic<br />

layouts in agreement with observed layouts are obtained. Much attention has been<br />

paid to the loads on the femoral head due to compressive body weight forces and<br />

to tension forces from the muscles. From this research it has been concluded that<br />

3D-modelling is necessary.<br />

The extension of homogenization based on frame models to continuum models<br />

leads to plate-like structures with closed walls. It is concluded in (Sigmund 1999) that<br />

to obtain the open wall micro-structures seen in reality, the optimization objective<br />

of maximum stiffness must be modified, or appropriate constraints must be added to<br />

the formulation.<br />

Let us now assume an initial bone structure corresponding to a certain load level<br />

is given, and then we change the load level, the load distribution or the bone structure.<br />

The evolution of µe as a function of time can then be based on the tendency to reach<br />

Memory<br />

function<br />

a new optimal solution, but a maximum rate of change dµe/dt must be included. In<br />

(Bagge 1999) a memory function is also included in the formulation, and different<br />

simulations over a range of 10,000 days are performed and shown. This procedure<br />

is also used for modelling the influence of implants, and it may be a tool for design<br />

of implants or training procedures. However, a much closer cooperation with actual<br />

observations is needed before the role of these simulations can be clarified.<br />

The modelling behind (Bagge 1999) and (Bagge 2000) is restricted to cubic symmetry<br />

for the bone material model, and thus a one parameter description µe is possible.<br />

In (Bendsøe et al. 1994) a procedure for free material evolution is described.<br />

9.3 Finite element analysis and load modelling<br />

The essential tool for mechanical analysis of bone is the finite element method with<br />

nodal displacements {D} obtained from solving the linear system of force equilibrium<br />

[S]{D} = {A}, where [S] is the total stiffness matrix and {A} the total load vector.<br />

The total stiffness matrix, often of the order 10,000 to 100,000, is obtained by accumulating<br />

element stiffness matrices [Se] of the order 10. The constitutive matrix [Le],<br />

that for linear elasticity {σe} =[Le]{ɛe} relates element stresses {σe} and element<br />

strains {ɛe}, is the essential basis for determining the element stiffness matrix [Se].<br />

With reference to the thesis (Bagge 1999) we know that it is necessary to model<br />

three dimensional and non-isotropic behaviour of bone. This means that the constitutive<br />

matrix [Le] is determined from up to 21 material parameters. This does not


Bone mechanics and damage evolution 189<br />

complicate the finite element analysis itself, but it is almost impossible to measure the<br />

many necessary parameters and even more difficult to model the evolution of these.<br />

A major part of this chapter is related to this problem. Load estimation<br />

A further complication is to obtain knowledge about the actual load (in reality<br />

load cases), as here described by the nodal loads {A}. A literature review and measures<br />

of the hip joint forces, force directions and moments during walking and running<br />

can be found in (Bergmann, Graichen and Rohlmann 1993). Muscle force magnitudes<br />

during gait are reported in (Pedersen, Brand and Davy 1997). For more extensive<br />

reference in relation to the proximal femur, see (Bagge 2000). Also it should be kept<br />

in mind that the loads are by no means stationary.<br />

The discretization into elements (or regions, or cells) should be commented. Although<br />

stresses, strains and energy densities are defined with reference to a point, we<br />

here assume constant stress, strain and energy density in an element e. Dealing with<br />

models containing 10,000 to 100,000 elements, this will not restrict the generality.<br />

The finite element analysis will not be further commented, but a comment on a<br />

non-traditional formula for the elastic energy density is needed. We write the elastic<br />

energy density ue as a scalar product { ˆɛe} T {Le} by defining the vector of constitutive<br />

components {Le} and the vector of second order strain components { ˆɛe} Constitutive<br />

vector<br />

and<br />

{Le} T := {{L1111,L2222,L3333}, √ 2{L1122,L1133,L2233},<br />

2{L1112,L1113,L1123,L2212,L2213,L2223,L3312,L3313,L3323},<br />

2{L1212,L1313,L2323}, 2 √ 2{L1213,L1223,L1323}} (9.1)<br />

{ ˆɛe} T := {{ɛ 2 11,ɛ 2 22,ɛ 2 33}, √ 2{ɛ11ɛ22,ɛ11ɛ33,ɛ22ɛ33},<br />

2{ɛ11ɛ12,ɛ11ɛ13,ɛ11ɛ23,ɛ22ɛ12,ɛ22ɛ13,ɛ22ɛ23,ɛ33ɛ12,ɛ33ɛ13,ɛ33ɛ23},<br />

{ɛ 2 12,ɛ 2 13,ɛ 2 23}, 2 √ 2{ɛ12ɛ13,ɛ12ɛ23,ɛ13ɛ23}} (9.2)<br />

where we have omitted the index e on the individual components. This writing Energy density<br />

hopefully makes the following sections more clear. It is directly verified that the<br />

result ue = { ˆɛe} T {Le} of these definitions agrees with alternative forms like ue =<br />

{ɛe} T [Le]{ɛe}. In reality {Le} is just the vector notation for the fourth order elasticity<br />

tensor defined with invariant length, see (Pedersen 1995). Note, that the energy<br />

density in this chapter is the sum of the strain energy density and the stress energy<br />

density.


190 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

9.4 Homogenization and inverse homogenization<br />

For calculating effective material properties we use the homogenization method and<br />

assume the existence of a base cell that in a periodic distribution represents the bone<br />

material. Subjecting this base cell to any test strain field that satisfies the periodicity,<br />

we demand the elastic energy of the homogenized base field to equal the elastic energy<br />

Homogenization of the detailed base cell (the micro-structure).<br />

analysis For a homogeneous micro-structure with volume V and material parameters {LH }<br />

the elastic energy U is directly given by U = V {ˆɛ} T {LH } when the strain field<br />

{ˆɛ} is constant. With only six strain states we can relate the 21 material parameters<br />

in {LH } to combinations of elastic energy densities (assuming linear elasticity).<br />

For the first state we choose ɛ11 = 1, and others ɛij = 0; then similar for<br />

ɛ22,ɛ33, √ 2ɛ12, √ 2ɛ13, √ 2ɛ23, i.e. only one strain component not equal to zero in each<br />

state. The resulting formulas are listed in (Pedersen and Tortorelli 1998), see also<br />

chapter 11.<br />

For the inhomogeneous micro-structure, the bone micro-structural model is subjected<br />

to boundary displacements corresponding to the six strain states and we perform<br />

a finite element analysis. Except for the periodicity of the boundary conditions<br />

this analysis is exactly the same as for any macro-structure with loads only from<br />

forced displacements. With elastic energy from the six different load cases we have<br />

the homogenized constitutive parameters.<br />

Frame model<br />

with cubic<br />

symmetry<br />

Micro-structure<br />

optimization<br />

In the thesis by (Bagge 1999) a micro-structural three dimensional frame model<br />

was applied as idealized in (Guedes and Kikuchi 1990). This chosen micro-structure<br />

gives cubic symmetric material properties, and calculations for discrete values of the<br />

relative volume fraction µ are found using the program PREMAT3D by (Guedes<br />

1995). Finally the homogenized material parameters are fitted to polynomials, thus<br />

making derivatives available.<br />

As an alternative to postulate a micro-structural model and then by homogenization<br />

find the constitutive parameters as a function of e.g. a density parameter, we<br />

may find the micro-structure as a result of an optimization procedure. This local<br />

optimization problem is similar to the global problem presented in the next section<br />

and results obtained with the techniques of topology optimization are given in (Sigmund<br />

1999). This paper discusses the relation between closed and open wall cells and<br />

conclude that closed wall cells have optimal stiffness properties whereas open wall<br />

cells are non-optimal. Specific stress states are considered with focus on optimized<br />

bulk modulus.<br />

In a more abstract setting the paper by (Bendsøe et al. 1994) treats the ”free<br />

material” with direct design of the individual constitutive parameters. Analytical<br />

results are obtained and ultimate bounds are thereby available. Alternative formulations<br />

are found in (Rodrigues, Jacobs, Guedes and Bendsøe 1999) and in (Jacobs,


Bone mechanics and damage evolution 191<br />

Simo, Beaupre and Carter 1997).<br />

The problem of micro-mechanical design is understood as the problem of finding a<br />

distribution in a base cell of homogenization, in order to obtain prescribed constitutive<br />

parameters or to satisfy constraints on these. For obvious reasons the problem Inverse<br />

is also termed the inverse homogenization problem or identification of the micromechanical<br />

structure. These problems can as most optimization problems be solved<br />

by a number of different iterative methods, where each iteration includes analysis of<br />

a micro-mechanical structure, sensitivity analysis and decision of an improved design.<br />

For the sensitivity analysis we need the derivatives of the constitutive parameters,<br />

which can be found from derivatives of elastic energies.<br />

9.5 Bone layout based on<br />

one parameter optimization<br />

homogenization<br />

If we for a frame-like bone structure postulate that the density distribution and the<br />

field of orientations in a bone structure are controlled by an extremum principle, like<br />

maximum stiffness for a given amount of material, then we should be apple to find the<br />

bone layout by solving an optimization problem. Maximum stiffness corresponds to<br />

minimum of elastic energy U, and the available amount of material ¯ V we accumulate<br />

over the elements, each with relative volume fraction µe, i.e. the constraint of the<br />

problem is µeVe − ¯ V = 0. Omitting at first the orientationel problem, then the<br />

optimization problem is stated Optimization<br />

formulation<br />

F ind µe which Minimize U subject to g = µeVe − ¯ V = 0 (9.3)<br />

The elastic energy is accumulated over element elastic energies Ue<br />

U = Ue = Veue = Ve{ ˆɛe} T {Le} (9.4)<br />

Using the results from (Pedersen 1998), see chapter 13, the derivative of U with respect<br />

to the local constitutive parameters (without influence on the loads) simplifies to Sensitivity<br />

dU ∂U<br />

= −(<br />

d{Le} ∂{Le} )¯ɛ = −( ∂Ue<br />

∂{Le} )¯ɛ = −Ve( ∂ue<br />

∂{Le} )¯ɛ = −Ve{ ˆɛe} T<br />

(9.5)<br />

where sub-index ¯ɛ indicates the fixed strain field.<br />

For an optimization problem with only one constraint the optimality criterion is<br />

that the ratio between the derivative of the objective (U) and the derivative of the<br />

constraint (g = 0) should be the same for all optimization variables, see chapter 14.<br />

of energy


<strong>Optimal</strong>ity<br />

criterion<br />

Special case<br />

Incremental<br />

formulation<br />

<strong>Optimal</strong><br />

re-modelling<br />

192 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

These derivatives are with the result (9.5)<br />

dU<br />

dµe<br />

=<br />

dU d{Le}<br />

T d{Le}<br />

= −Ve{ ˆɛe}<br />

d{Le} dµe<br />

dµe<br />

(9.6)<br />

dg/dµe = Ve (9.7)<br />

and thus the actual numerical problem is to find a distribution µe that satisfies the<br />

optimality criterion<br />

{ ˆɛe} T (d{Le}/dµe) =constant (same for all e) (9.8)<br />

As mentioned earlier the relation between the vector of constitutive parameters {Le}<br />

and the relative volume fraction µe follows from the basic homogenization of the<br />

material model. It might be advantageous to fit the results of the homogenization<br />

with functions that then gives analytical derivatives d{Le}/dµe. The special case of<br />

a homogeneous relation {Le} = µ q e ˜ {Le}, where {Le} ˜ do not depend on µe, results in<br />

ue/µe =a constant, i.e. constant elastic energy density per optimization variable.<br />

Different methods are available for the numerical solution, see (Bagge 1999). For<br />

a proximal femur, it is shown in (Bagge 2000) and (Bagge 1999) that realistic layouts<br />

in agreement with observed layouts are obtained. For these solutions the principal<br />

material directions are aligned directly with the principle stress directions, as by the<br />

law of Wolff and in general proven for low shear stiffness materials in (Pedersen 1989),<br />

see chapter 17.<br />

With time varying loads the optimization problem is formulated as finding the<br />

optimal redesign within a time-step where the load case is fixed. Constraints based on<br />

maximum rate of change of the relative volume fractions as well as the maximum reorientations<br />

are taken into account. The specific numerical procedures are described<br />

in (Bagge 2000).<br />

9.6 Multi-parameter evolution<br />

Based on the conclusion that it will not be possible to measure individually the<br />

simultaneously evolution of 21 material parameters, a procedure using an optimization<br />

principle was suggested in (Hammer and Pedersen 1999) for laminated materials and<br />

in (Tortorelli and Pedersen 1999) for bone materials.<br />

Although experimental verification is still missing the procedure will be shortly<br />

described here. We postulate that in a time-step of a simulation the change in {Le},i.e.<br />

{∆Le}, will be such that the change in total elastic energy ∆U is extremized. From<br />

(9.5) we get<br />

∆U = <br />

−Ve{ ˆɛe} T {∆Le} (9.9)<br />

e


Bone mechanics and damage evolution 193<br />

We see directly that extreme ∆U is obtained when {∆Le} and { ˆɛe} are aligned, i.e.<br />

{∆Le} = Ce{ ˆɛe} and only the constant Ce needs to be determined.<br />

The vector {Le} contains the 21 independent material parameters for an element<br />

constitutive behaviour. The length Fe of this vector (F 2 e := {Le} T {Le}) is known as<br />

the Frobenius norm of the constitutive matrix [Le], and we assume that the magnitude<br />

of evolution of stiffness is described by the rate of change ˙<br />

Fe = dFe/dt. Figure 9.1 Rate of<br />

shows such phenomenological(postulated) bone evolution. Several different domains<br />

are identified: degradation due to low energy density, quiescence due to ”normal<br />

energy density”, growth due to high energy density, and damage due to extensive<br />

energy density.<br />

f<br />

ud<br />

(urd,frd)<br />

ug<br />

(urg,frg)<br />

(uf ,ff )<br />

0 2 4 6 8 10<br />

u<br />

degradation quiescence growth failure<br />

Figure 9.1: Possible evolution function for the velocity factor f in an element as a<br />

function of the actual element energy density u. The postulated function is specified<br />

by the 8 parameters shown in the figure with an addition of three power parameters,<br />

for this figure: nd = −1.5,ng =2.0, and nf =4.0.<br />

The actual function ˙ F = ˙ F (u) is only illustrated by a non-dimensional factor<br />

f and must actually be identified from observations. The present description by<br />

constitutive<br />

norm


Example of<br />

rate function<br />

Constitutive<br />

increments<br />

Constitutive<br />

constraints<br />

Non-bone<br />

materials<br />

194 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

functions is<br />

f = fd<br />

nd ud−u<br />

ud<br />

0 ≤ u ≤ ud<br />

f = 0 ud ≤ u ≤ ug<br />

f = frg<br />

u−ug<br />

f = ff +(frg − ff )<br />

ng<br />

urg−ug<br />

nf u−uf<br />

urg−uf<br />

ug ≤ u ≤ urg<br />

ug ≤ u (9.10)<br />

but alternatively a pure numerical tabulation may be chosen.<br />

By differentiation of F 2 e we get 2Fe ˙ Fe =2{Le} T { ˙ Le} and with { ˙ Le}∆t = {∆Le} =<br />

Ce{ ˆɛe} we obtain<br />

Fe ˙ Fe∆t = {Le} T Ce{ ˆɛe} (9.11)<br />

In (9.11) we know Fe, Fe, ˙ ∆t, {Le}, { ˆɛe} , then the unknown constant Ce is determined<br />

by<br />

Ce =(Fe ˙ Fe∆t)/({Le} T { ˆɛe}) (9.12)<br />

and with ue = {Le} T { ˆɛe} we get the result<br />

FeFe∆t ˙<br />

{∆Le} = { ˆɛe} (9.13)<br />

ue<br />

During a simulation we can also put constraints on { ˙<br />

Le} = {∆Le}/∆t and control<br />

that the matrix [Le] behind the definition of the vector {Le} stays positive definite.<br />

For details and example of evolution of a beam problem, see (Tortorelli and Pedersen<br />

1999). For laminate problems a slightly modified procedure is described in (Hammer<br />

and Pedersen 1999).<br />

For metal a damage limit σd may define the initiation of the stiffness reduction<br />

with increasing rate for increasing effective stress σeff . Figure 9.2 shows such a<br />

phenomenological(postulated) curve for damage. This might seem as a rather unusual<br />

alternative to keeping track of strength surfaces and their evolution. In (Hammer and<br />

Pedersen 1999) another approach to evolution of strength surfaces is presented and<br />

applied to laminate damage.


Bone mechanics and damage evolution 195<br />

f<br />

σd<br />

0 2 4 6 8 10<br />

no damage damage<br />

(σrd,frd)<br />

σeff<br />

Figure 9.2: Possible evolution function for the velocity factor f in an element as a<br />

function of the actual element effective stress σeff . The postulated damage function<br />

is specified by the 3 parameters shown in the figure with an addition of the power<br />

n =1.5, for the damage part.


Note<br />

196 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

9.7 Conclusion<br />

The present chapter informs about methods and formulations that hopefully can be<br />

helpful in further understanding the complicated behaviour of bone. Simultaneously<br />

analysis on the micro-structural level as well as on the macro-structural level is now<br />

possible. Large computer simulations in 3D-formulations with non-isotropic behaviour<br />

are possible, but further efforts on clarifying the experimental possibilities are needed.<br />

Close relation between research in bone mechanics and in optimal material design<br />

has been the focus of international meetings, like an informal workshop on ”Comparing<br />

structural optimization and tissue adaptation models”, arranged by B. Kohn and<br />

S. Cowin in New York January 1997, an the IUTAM symposium in May 1998 with<br />

proceedings in (Pedersen and Bendsøe 1999), and an AIMETA mini-symposium in<br />

September 2001 with a journal special issue (Prendergast and Contro 2002). It seems<br />

clear that many years of research is still needed within this interesting and important<br />

area of science.<br />

It must be stressed that the suggested evolution models illustrated in figures 9.1<br />

and 9.2 are only an idea, and further research with experimental backup is very much<br />

needed.


Chapter 10<br />

Identification (estimation)<br />

and inverse problems<br />

This chapter describes solutions to two problems related closely to laminate analysis<br />

and laminate design, but not with the objective of optimal design as in chapter Not optimal<br />

design<br />

6. The first part could be termed estimation of material parameters for composite<br />

laminates and is closely related to the note (Pedersen 1999a) and the references in<br />

this note, especially the papers by P. S. Frederiksen. The solution approach belongs<br />

to the experimental-numerical methods for which a much wider area of application<br />

is possible, as seen in estimation of stiffness models in (Pedersen 1986a), estimation<br />

of applied loads, estimation from a hardness test (Pedersen, Mortensen and Larsen<br />

1998), and estimation from a deep drawing process (a project in progress).<br />

The second part could be termed inverse problems for laminates and materials, Laminate<br />

where the goal is to find a design that returns specified specifications, say the bending<br />

stiffnesses of a laminate or the constitutive parameters of a material. A recursive<br />

iteration technique based on an optimality criterion is described, and may serve as<br />

an illustration for the many other problems within the class of inverse problems in<br />

general.<br />

10.1 Estimation of orthotropic material parameters<br />

inverse<br />

problems<br />

A combined experimental-numerical method is presented with the goal of obtaining Experimental<br />

the material stiffness for composite materials. The identification is based on eigenfre- - numerical<br />

quencies for a free rectangular plate, because excellent agreement between measured<br />

and calculated eigenfrequencies can be obtained. The numerical identification prob-<br />

197


Complicated<br />

laboratory<br />

tests<br />

Simple<br />

laboratory<br />

tests<br />

198 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

lem is formulated as an optimization problem, and in one experiment we obtain all<br />

the involved moduli.<br />

The total approach consists of experiment, numerical , eigenvalue sensitivity analysis,<br />

optimization, and error estimation. Results from these different parts of the<br />

approach are commented and a general conclusion is that the method has several<br />

significant advantages compared to traditional determination of material moduli. An<br />

overview of the state of the art is intended.<br />

Traditional laboratory tests are designed to give the desired quantities in a direct<br />

manner. As an example, determination of material parameters of constitutive models<br />

leads to special design of test samples and choices of applied load in an attempt to<br />

obtain homogeneous stress-strain fields. These idealized tests are not easy to perform,<br />

and especially with composite systems problems often arise. Furthermore, the tests<br />

have a local nature, e.g. the point of the strain gauge. So for more general material<br />

information a series of tests is required.<br />

The identification approach is an experimental strategy from an opposite point<br />

of view. The experiment is made as simple as possible to give reliable results.<br />

On the other hand, this often leads to complex inhomogeneous stress-strain fields<br />

where the desired quantities are not among the directly measured quantities. This<br />

causes a complicated interpretation, but nowadays this is comfortably taken care of<br />

by computer calculations. Although our aim is to find static material data, we decide<br />

to measure eigenfrequencies of a free rectangular plate because excellent agreement<br />

between measured and calculated eigenfrequencies can be obtained. A structural<br />

Research<br />

groups<br />

eigenfrequency is an integrated quantity and we thus obtain material quantities that<br />

are valid in the mean for the entire laminate.<br />

In a pilot project by (Markworth and Petersen 1987) this technique was found<br />

most promising (some of the results can be found in (Pedersen and Frederiksen 1992).<br />

The same conclusion was reported in the doctoral thesis by (Sol 1986), and the research<br />

groups in Brussels, Belgium and in Eindhoven, The Netherlands have continuously<br />

refined the method and obtained impressive results. The proceedings from an<br />

Euromech colloquium, (Sol and Oomens 1997), give a picture of the state of the art.<br />

A doctoral thesis by (Kuttenkeuler 1998) shows the importance of these techniques<br />

for composite design.<br />

In the author’s home department the extended work of P. S. Frederiksen reports<br />

to a great depth about the method. Especially the work related to thick plate formulations,<br />

(Frederiksen 1995), (Frederiksen 1997b), should be noted and also the work<br />

related to parameter uncertainty (Frederiksen 1998) deserves special attention. The<br />

research group in Portugal has concentrated on the combination with the finite element<br />

method and has also applied thick plate formulation, see (Soares, de Freitas,<br />

Araujo and Pedersen 1993) and (Araujo, Soares, Freitas, Pedersen and Herskovits<br />

2000).


Identification (estimation) and inverse problems 199<br />

Here we first formulate the identification problem from an optimization point of<br />

view. Then the experimental set-up is described, followed by numerical eigenvalue<br />

analysis and sensitivity analysis. The optimization algorithms are shortly commented,<br />

and then results from the literature are commented. Among these are the study of<br />

material moduli in different environments.<br />

10.1.1 Identification formulation<br />

A structure can be described in many ways, directly or indirectly. The direct approach<br />

can be by parameters, like length, width, height, weight, topology and material parameters.<br />

Alternatively, we can describe it by its response like static displacements<br />

or dynamic eigenfrequencies. Combination approaches are also possible. A laminated<br />

plate is a structure where the ply stacking is an important aspect. The thickness<br />

of the plies and their positions in addition to ply orientations are the parameters<br />

that describe this stacking. Also boundary conditions are of vital importance for the<br />

response of the plate.<br />

The laminate membrane stiffnesses [A], as well as the laminate bending stiffnesses<br />

[D] are linearly depending on the material stiffness [C] for each ply, i.e. upon the<br />

constitutive matrix. The chosen identification problem we describe as a problem where<br />

we do not have information enough about [C] , but we have information about the<br />

structural response, say in terms of measured eigenfrequencies ωi. Let us be specific<br />

and formulate the identification problem as an optimization problem. A number of<br />

parameters pk are unknown for the determination of the material stiffnesses<br />

[C] =[C(pk)] (10.1)<br />

and therefore a calculated eigenfrequency response also depends on these parameters<br />

ωi = ωi(pk) (10.2)<br />

The functional relations (10.1) and (10.2) can be found in textbooks on laminates,<br />

say in (Jones 1975) and in (Pedersen 1987b).<br />

The optimization problem is to find best possible agreement between measured<br />

eigenfrequencies ¯ωi and calculated eigenfrequencies ωi. An objective could be to Objective<br />

Minimize Φ= (ωi − ¯ωi) 2<br />

i<br />

¯ω 2 i<br />

(10.3)<br />

i.e., a weighted least-squares estimator, but naturally other choices are possible. For<br />

example, more weight could be given to more accurate measurements, and a priori


Move-limits<br />

Non-dimensional<br />

parameters<br />

Constraint on<br />

definiteness<br />

200 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

expectation on parameter values can be included. The solution must be obtained in<br />

an iterative process, where the parameters are simultaneously improved<br />

(pk)n+1 =(pk)n +(∆pk)n for n =1, 2,... (10.4)<br />

and constraints can be put on the absolute values of the parameters (side constraints)<br />

as well as on the changes of the parameters (move-limits)<br />

(pk)min ≤ (pk) ≤ (pk)max and (∆pk)min ≤ (∆pk) ≤ (∆pk)max (10.5)<br />

With this kind of constraints we do not need in advance to distinguish very clearly<br />

between known and unknown parameters. For orthotropic two-dimensional models,<br />

two parameters are especially important. The level of non-isotropy described by the<br />

parameter α2<br />

α2 := (1 − ET /EL)/2 (10.6)<br />

where ET ,EL are the engineering moduli in transverse and fiber direction, respectively.<br />

The relative shear stiffness is described by the parameter α3<br />

α3 := (1 + (1 − 2νLT )ET /EL − 4α0GLT /EL)/8 (10.7)<br />

where νLT is the major Poisson’s ratio and GLT is the shear modulus. The nonimportant<br />

parameter α0 is given by α0 = 1 − ν 2 LT ET /EL, and a further not so<br />

important parameter is<br />

α4 = α3 + νLT (1 − 2α2) (10.8)<br />

i.e. only new information from the Poisson’s ratio. Solving the problem (10.3) in the<br />

formulation<br />

Minimize Φ=Φ(α2,α3,α4) ≥ 0 (10.9)<br />

subject to the constraint of a positive definite constitutive matrix. Written in the<br />

non-dimensional parameters we find<br />

1 − 2α2 > 0; 1 − α2 − 3α3 − α4 > 0; 1 − 2α2 − (α4 − α3) 2 > 0 (10.10)<br />

When α2,α3,α4 are determined we can invert the relations (10.6), (10.7) and (10.8)<br />

and identify EL, ET , GLT and νLT .<br />

10.1.2 The experimental setup<br />

A detailed description of the experimental procedure and the critical aspects of this<br />

can be found in (Frederiksen 1997a). Here we describe only the main aspects. As<br />

shown in figure 10.1 the setup consists of six main components:<br />

1) the test specimen, i.e. the laminated plate


Identification (estimation) and inverse problems 201<br />

2) the excitation source, i.e. an impact hammer or a loudspeaker<br />

3) the response detector, i.e. an accelerometer, a microphone or<br />

a laser vibrometer<br />

4) a power amplifier<br />

5) a transient recorder and frequency analyzer<br />

6) a personal computer, which includes the identification program(s)<br />

Figure 10.1: Scheme of the experimental setup.<br />

The test specimens in the works of Frederiksen were chosen as rectangular, orthotropic<br />

and symmetric laminates. In the work by (Kuttenkeuler 1998) also non-<br />

rectangular test specimens were used, but then the numerical analysis needs to be Test<br />

based on finite element modelling. In the section on uncertainties we discuss the other<br />

physical aspects for the test specimens.<br />

specimens<br />

As excitation source we mostly use a small impact hammer, which imparts a force Excitation<br />

with a broad frequency range to the plate. This simultaneously excites all the modes<br />

of vibration. The pulse frequency range can be altered by adjusting the tip-hardness<br />

and the hammer weight. In the identification of complex moduli by (de Vissher, Sol,<br />

de Wiede and Vantomme 1997) the excitation by a loudspeaker or a disconnected<br />

shaker is also described. The joke is: ”ask the plate for its moduli”, but do it loudly<br />

(in any language).<br />

In the early research a light weight accelerometer with a mass of 2.2 gram was<br />

source<br />

used as response detector, and the concentrated mass was taken into account in the Response<br />

eigenfrequency calculations. In more recent research, especially for the study of small<br />

and light plates, a microphone is used as response detector. In (de Vissher et al. 1997)<br />

detector


Further<br />

tools<br />

Rayleigh-Ritz<br />

or FEM<br />

Higher order<br />

plate theory<br />

Subspace<br />

iterations<br />

202 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

a laser vibrometer is shown as response detector. In general we can state that good<br />

experimental repeat-ability is obtained and reported by the different authors.<br />

The power amplifier, the transient recorder and the PC equipment are standard<br />

experimental tools. The software separation between the transient recorder and the<br />

PC is not very specific. Thus, the fast Fourier transformation could be contained in<br />

either of them.<br />

10.1.3 Plate eigenfrequency analysis and sensitivity analysis<br />

Plate eigenfrequencies can be calculated with different methods. For the numerical<br />

modelling as well as for the numerical procedure we have many possibilities and in<br />

this section we comment on the ones used for the actual identification problem. For<br />

detail see (Frederiksen 1995) and (Frederiksen 1997b).<br />

The numerical modelling is performed either by the Rayleigh-Ritz method with<br />

global expansion functions or by the finite element method with local expansion functions.<br />

Both methods are applicable to rectangular plates, but for more complicated<br />

boundary shapes only the finite element method is practical. The main issue of the numerical<br />

modelling is to choose the relevant plate theory, and it is generally agreed that<br />

higher order plate theories are often required. In the studies of (Frederiksen 1997b)<br />

an estimation of the errors that result from the use of the classical plate theory is<br />

included. The higher order shear deformation theory by (Reddy 1984), (Levinson<br />

1981) is a good compromise between accuracy and complexity.<br />

When the numerical modelling is chosen we end up with a generalized, linear<br />

eigenvalue problem<br />

[S]{Di} = ω 2 i [M]{Di} (10.11)<br />

where [S] is a ”stiffness” matrix, [M] a ”mass” matrix, and ω 2 i , {Di} a squared natural<br />

frequency and its corresponding eigenvector. In a Rayleigh-Ritz model the order of<br />

the matrices is say 500 , while in a finite element model the order may be say 20.000 .<br />

(Comparisons of results from the two different methods can be found in (Frederiksen<br />

1997b)). The assumption of linearity is justified by very small strains during the<br />

vibrations (1-10 micro-strains).<br />

An effective numerical procedure to solve the matrix eigenvalue problem (10.11)<br />

is the subspace iteration method, see (Bathe 1982). By iteration this method de-<br />

termines the eigenvectors and then evaluates the eigenvalues in a selected domain of<br />

the spectrum, say the lowest eigenfrequencies. The method is capable of dealing with<br />

multiple eigenfrequencies, and in general computer time is no problem.<br />

For the identification we also need the partial derivatives of the eigenfrequencies<br />

with respect to the material moduli, here symbolized by a parameter pk. The general


Identification (estimation) and inverse problems 203<br />

result for distinct eigenfrequencies is<br />

when the normalization<br />

dω 2 i<br />

dpk<br />

= {Di} T (−ω 2 i<br />

d[M]<br />

dpk<br />

+ d[S]<br />

){Di} (10.12)<br />

dpk<br />

{Di} T [M]{Di} = 1 (10.13)<br />

is assumed. Detailed description of this sensitivity analysis is presented in chapter<br />

18.<br />

For the actual problem where d[M]/dpk = [0], because the material moduli have Eigenvalue<br />

no influence on the mass matrix, we get<br />

sensitivities<br />

dω 2 i<br />

T d[S]<br />

= {Di} {Di} (10.14)<br />

dpk dpk<br />

For multiple eigenfrequencies a special calculation is necessary to obtain the eigenvectors<br />

that uncouple the sensitivities. With {D}i, {D}j being two eigenvectors corresponding<br />

to the same eigenfrequency ωi, we determine these vectors uniquely by the<br />

two orthogonality conditions:<br />

where δij is Kronecker’s delta.<br />

{D} T i [S]{D}j = δijω 2 i and {D} T i<br />

10.1.4 Example results<br />

d[S]<br />

dpk<br />

dω<br />

{D}j = δij<br />

2<br />

dpk<br />

(10.15)<br />

In the different studies of identification techniques, different objectives and different<br />

optimization algorithms have been applied. Different algorithms for the different steps<br />

have been necessary to remove the necessity of good initial guesses. As local minima<br />

are an actual problem, (Frederiksen 1992) suggested initial steps based on rather<br />

rough estimation to get good initial guesses before the more advanced optimization<br />

algorithms are applied. The thesis (Frederiksen 1992) should be consulted for more<br />

detailed advice on the optimization algorithms.<br />

Let us comment on results for weakly, moderately and strongly non-isotropic Weakly<br />

materials. A rolled aluminum plate is identified in the early work by (Markworth and<br />

Petersen 1987) which based on ten frequencies found EL =68.9GP a, ET =66.0GP a,<br />

GLT =24.8GP a, νLT =0.343 and fairly good agreement with traditional uniaxial<br />

tests. (Frederiksen 1997a) for another aluminum plate identified EL =71.1GP a,<br />

ET =70.7GP a, GLT =25.9GP a, νLT =0.336, GLZ =27.9GP a, GTZ =28.0GP a,<br />

thus showing the possibility for identifying the out-of-plane shear moduli GLZ,GTZ.<br />

non-isotropic


Strongly<br />

non-isotropic<br />

Citation from<br />

Frederiksen<br />

204 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

For glass-epoxy also the early studies report good agreement with static strain Moderately<br />

gauge test results in the identification of the in-plane moduli. Based on 14 frequencies,<br />

(Frederiksen 1997a) identified the following parameters EL = 42.4GP a,<br />

ET =11.6GP a, GLT =4.68GP a, νLT =0.305, GLZ =4.55GP a, GTZ =4.07GP a,<br />

with details on iteration history and the extremely small residuals from experimental<br />

and calculated frequencies.<br />

Carbon-epoxy identification is properly of most interest and extended results are<br />

available, in early as well as recent studies. Let us here again comment the results of<br />

(Frederiksen 1997a), who reports EL = 113.0GP a, ET =8.50GP a, GLT =4.45GP a,<br />

νLT = 0.323, GLZ = 4.43GP a, GTZ = 2.97GP a. We find almost transversely<br />

isotropic stiffness for this unidirectional laminate and the ratio GTZ/GLZ =0.666<br />

is in other studies reported between 0.548 and 0.624. Note that the determination<br />

of out-of-plane shear moduli is very difficult to perform by other experimental techniques.<br />

Finally, we show in 10.2 the results from identification of the temperature dependence<br />

for glass-epoxy moduli; mainly to illustrate the possibilities when knowledge<br />

on environmental influence is important.<br />

10.1.5 Uncertainties and optimal experiments<br />

In the early thesis of (Sol 1986) and also in the recent thesis of (Kuttenkeuler 1998)<br />

assessment of uncertainties was given priority. In the extended study of (Frederiksen<br />

1998) these questions are analysed based on statistics, and from the abstract in this<br />

paper we cite ”This paper investigates an inverse technique for the identification of<br />

orthotropic elastic constants from measured plate natural frequencies. In general, the<br />

accuracy of the identified parameters depends on the method of estimation, modelling<br />

errors and measurement errors. The paper addresses the parameter uncertainty due<br />

to errors in the measurements. Based on assumptions of the measurement errors,<br />

second-order statistics of parameters are approximated by linearization schemes. The<br />

main focus is on the possibility of designing the experiment to minimize the uncertainty<br />

of the estimated parameters. The uncertainty of each estimate as function of<br />

the experimental design variables is investigated. Also the overall optimality of the<br />

experimental design defined as the hyper-volume of the confidence region is considered.<br />

The results show that not all parameters are estimated with a sufficient precision in<br />

the general case, but by carefully designing the experiment, the parameter uncertainties<br />

can be greatly reduced. Both thin and thick plates are considered with focus on<br />

single-layer plates, but the results for laminates plates are also discussed.”<br />

The individual errors can be grouped into<br />

• Experimental errors<br />

• Physical modelling errors<br />

non-isotropic


Identification (estimation) and inverse problems 205<br />

Figure 10.2: Temperature dependence for glass-epoxy ply obtained by identification.


206 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

• Numerical modelling errors<br />

• Plate defects<br />

Experimental The experimental errors are reported as limitations of the instruments, and are<br />

errors generally low. (Frederiksen 1998) states the standard deviations of the frequency<br />

values to be less than 0.2% . Measuring the thickness of a laminated plate may be<br />

Modelling<br />

errors<br />

Numerical<br />

errors<br />

Plate<br />

errors<br />

Alternative<br />

materials<br />

non-unique and a method based on volume measure in water is therefore suggested.<br />

The physical modelling errors include the assumptions related to linearity, damping,<br />

surrounding air, gravity, suspension and frequency dependence. Most critical of<br />

these sources might be the suspension, but with a free plate only hanged in rubberbands<br />

the boundary conditions are well established. Other boundary conditions normally<br />

cause severe errors. Lowest and highest frequency in the identification often<br />

deviate by one order, and if the material moduli are frequency dependent we should<br />

conclude that averaged properties in some sense are found.<br />

The numerical modelling errors have been investigated in (Frederiksen 1995).<br />

For the first 15 frequencies, the maximum error is found to be 0.4% for a moderately<br />

thick orthotropic plate. For most of the frequencies the error is considerably smaller.<br />

Through many experiments the most essential problems were related to plate defects,<br />

such as non-rectangular shape, non-uniform thickness, non-zero plate curvature,<br />

and most important a non-homogeneity on the macro level. This last aspect has different<br />

influence on the individual eigenmodes, eigenvalues and therefore often makes<br />

identification impossible, which in a way is better than an artificial adjustment of the<br />

moduli.<br />

10.1.6 Comments<br />

The identification technique is a fascinating approach involving very close cooperation<br />

between experimental and numerical work. The studies have shown that reliable<br />

values of the six most important moduli of orthotropic laminates can be determined<br />

from one experiment, measuring plate frequencies. Because of the simplicity of the<br />

free isolated specimens, environmental influence on material moduli can easily be<br />

determined. Results from temperature dependence are shown to illustrate these potentials.<br />

Naturally also other materials than composite laminates can be studied with<br />

the method. With thick plate modelling these potentials are greatly enlarged. As ex-<br />

ample we mention bone material (see early study by (Thomsen 1990)), ceramics, and<br />

sandwich combinations.<br />

The design of optimal experiments for identification of specific parameters is an<br />

important new aspect that follows from the statistical analysis for determination of<br />

parameter uncertainties. Bounds related to the identified parameters on the laminate<br />

level as well as on the ply (lamina) level are important additional information to<br />

obtain.


Identification (estimation) and inverse problems 207<br />

10.2 Laminate inverse problems<br />

From searching the very detailed information in constitutive parameters, we here<br />

focus on problems at another level. Given the constitutive description of the plies<br />

(layers, laminas) in a laminate, how to choose the stacking (thicknesses, orientations,<br />

positions) to obtain a laminate with specified stiffnesses (if possible) ? This laminate<br />

inverse problem is not simple to solve without the tools of optimization. Laminate<br />

In two-dimensional formulation of laminates we deal with a number of 3 × 3<br />

matrices: the constitutive (material stiffness) matrix [C], the laminate membrane<br />

stiffness matrix [A], the laminate bending stiffness matrix [D], the laminate coupling<br />

stiffness matrix [B], the laminate stress strength matrix [F ], and the laminate strain<br />

strength matrix [G]. In relation to the present inverse problem they are all treated<br />

analogously and as example we formulate the problem of finding a laminate that<br />

results in a given bending matrix [ ¯ D]. This example is described more detailed in<br />

(Pedersen 1999b).<br />

10.2.1 Laminate with given bending stiffness<br />

design<br />

With the four index notation from the tensor formulations the matrix [D] of laminate Bending<br />

bending stiffnesses is defined as<br />

⎡<br />

[D] = ⎣<br />

D1111<br />

D1122 √<br />

2D1112<br />

√<br />

D1122 2D1112 √<br />

D2222 2D2212<br />

√<br />

2D2212 2D1212<br />

or with matrix [D] in contracted vector notation {D}<br />

{D} T = {D1111 D2222 2D1212<br />

⎤<br />

⎦ with [D] positive definite (10.16)<br />

√ 2D1122 2D1112 2D2212} (10.17)<br />

The factors √ 2 and 2 in (10.17 ) are related to a moment vector definition {M} T √ =<br />

2κ12} that<br />

stiffnesses<br />

√<br />

{M11 M22 2M12} and to a curvature vector definition {κ} T = {κ11 κ22<br />

imply orthogonal rotation transformations and then, as an example, invariant trace<br />

of [D] equal to (D1111 + D2222 +2D1212).<br />

The matrix [D] = [D(pk)] depends on the design parameters pk (thicknesses,<br />

orientations, positions), and our goal is to find parameters that give [D] =[ ¯ D]orat<br />

least obtain as close as possible agreement. This closeness is measured by the residual<br />

matrix ([D] − [ ¯ D]). To get an objective that is invariant we define an error function Objective<br />

Φ as the squared Frobenius norm of this matrix, i.e.<br />

Φ=({D} T −{ ¯ D} T )({D}−{ ¯ D}) (10.18)


Optimization<br />

problem<br />

Iteration<br />

strategy<br />

208 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

It follows that Φ ≥ 0, and if Φ = 0, we have located an exact solution [D] =[ ¯ D]. A<br />

solution approach is then by iterations to minimize Φ and in such an approach we<br />

need the gradients<br />

dΦ<br />

=2({D}<br />

dpk<br />

T −{ ¯ D} T ) d{D}<br />

(10.19)<br />

dpk<br />

where formulas for d{D}/dpk can be found in (Pedersen 1987b).<br />

10.2.2 A redesign procedure<br />

A redesign described by<br />

then gives<br />

(pk)next = pk +∆pk<br />

(Φ)next =Φ+ <br />

(dΦ/dpk)∆pk<br />

i<br />

(10.20)<br />

(10.21)<br />

and since (dΦ/dpk) is only correct for small ∆pk a possible approach would be in each<br />

iteration to<br />

Minimize (Φ)next subjected to g = <br />

(∆pk) 2 − R = 0 (10.22)<br />

where R is a redesign limit. Problem (10.22) can be solved analytically by proportionality<br />

between the gradient of the constraint and the gradient of the objective (see<br />

chapter 14) and we get<br />

∆pk = a dΦ<br />

= a2({D}<br />

dpk<br />

T −{ ¯ D} T ) d{D}<br />

dpk<br />

where the constant a is determined by the condition in (10.22)<br />

a<br />

2 <br />

i<br />

i<br />

(10.23)<br />

(dΦ/dpk) 2 = R (10.24)<br />

A practical approach is to start the redesign iterations with rather large resource for<br />

change R, and ending up with a redesign based on smaller resources. Thus, a strategy<br />

for R as a function of the iteration history is needed, very much like a strategy for<br />

move-limits.


Identification (estimation) and inverse problems 209<br />

10.3 Material identification and inverse homogenization<br />

We now turn our attention to design of micro-structures, using optimization techniques,<br />

not necessary to get optimal material but just to get a material that satisfies Inverse<br />

certain specifications.<br />

In chapter 8 we deal more specific with optimal material design, that for the<br />

completely free material design is also covered in chapter 15. In the present chapter<br />

only inverse problems related to material design are included.<br />

10.3.1 Homogenization<br />

material<br />

problems<br />

For calculating effective material properties we use homogenization and assume the Base cell<br />

homogenization<br />

existence of a base cell that, in a periodic distribution, represents the material. Subjecting<br />

this base cell to any test strain field that satisfies the periodicity, we demand<br />

that the strain energy of the homogenized base field be equal to the strain energy of<br />

the detailed base cell (the micro-structure).<br />

For a homogeneous micro-structure no calculations are needed and we can simply<br />

set up the relations between total strain energies U for the cell subjected to elementary<br />

strain states and the homogenized constitutive parameters C H ijkl in 2D or LH ijkl<br />

in 3D<br />

as specified in detail in chapter 11.<br />

For the inhomogeneous micro-structure we have to perform a detailed analysis<br />

(often a finite element analysis) to obtain the strain energy. In 2D three different<br />

boundary conditions are prescribed (in 3D six), corresponding to the corner condition Elementary<br />

for the homogeneous micro-structure when evaluating the elementary strain states.<br />

Note that we assumed periodicity (shape not known) in the boundary displacements,<br />

which, for the homogeneous micro-structure, are straight lines.<br />

Except for the forced periodicity of the boundary conditions, the analysis of the<br />

inhomogeneous micro-structure is exactly the same as for any macro-structure with<br />

load only from forced displacements. From a finite element analysis with the microstructural<br />

stiffness matrix, we get the displacement, which gives the strain energy.<br />

We have established the direct relation between the strain energies as obtained in<br />

a traditional finite element analysis of the inhomogeneous micro-structure and the<br />

homogenized constitutive parameters, listed in chapter 11.<br />

10.3.2 Inverse homogenization<br />

The problem of micro-mechanical design is here understood as the problem of finding<br />

a distribution of a given material (materials) in a base cell of homogenization in order<br />

to obtain prescribed material parameters. For obvious reasons, the problem is also<br />

strain cases


Multiple<br />

solutions<br />

References<br />

to results<br />

210 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

termed the inverse homogenization problem or identification of the micro-mechanical<br />

structure. As in the case of most identification problems, we cannot expect a unique<br />

solution, but, in fact, this is an advantage and we can introduce penalties and/or<br />

priorities as exemplified by a minimum mass micro-mechanical structure. Even so<br />

there may be several possible solutions.<br />

Like most optimal design problems, the inverse homogenization problem can be<br />

solved by a number of different methods. Most of them are based on an iterative<br />

scheme in which each iteration includes analysis of a micro-mechanical structure,<br />

sensitivity analysis and determination of an improved design. For the sensitivity<br />

analysis we need, for each design variable pe, (related to element e of the micromechanical<br />

structure), the change in the homogenized constitutive parameter, i.e. in<br />

3D dL H ijkl /dpe. Using the results of chapter 12 we can determine this in a localized<br />

calculation.<br />

In the Ph.D.-thesis by (Sigmund 1994a) and in the papers (Sigmund 1994b),<br />

(Sigmund 1995), as well as in the dr.techn.-thesis (Sigmund 2001), we find 2D and 3Dsolutions<br />

to inverse homogenization problems, often with minimum micro-structural<br />

volume as a solution priority. Ultimate optimal materials as described in chapter 15<br />

are also obtained, as well as materials with negative Poisson’s ratio. In papers by<br />

(Sigmund and Torquato 1996), (Sigmund and Torquato 1997) the inverse homogenization<br />

problem is formulated and solved for the thermoelastic problem. Properties<br />

of thermal expansion cannot be designed independently of the constitutive elastic behaviour.<br />

Very interesting results for the bounds that limit our possibilities are given<br />

in these papers. Many examples of interesting base cells are shown in (Sigmund and<br />

Torquato 1997), where one of the more interesting situations is a homogenized material<br />

with negative thermal expansion, obtained by mixing two phases both of positive<br />

thermal expansion together with void.


Chapter 11<br />

Alternative Descriptions of<br />

Constitutive Parameter<br />

11.1 Stress/strain relations<br />

The constitutive parameters relate stresses to strains, and we describe them in many<br />

different forms, say as fourth order tensors, as squared matrices, as vectors, or by Many<br />

invariants of these quantities. In a three-dimensional formulation the parameters<br />

are pure material parameters, but in two-dimensional formulations the parameters<br />

incorporate the modelling from three to two dimensions.<br />

Even for one-dimensional formulation with the modulus of elasticity E as constitutive<br />

parameter, giving normal stress σ from normal strain ɛ by σ = Eɛ, it is not<br />

always simple. For non-isotropic materials, E depends on the direction we choose. Reference<br />

Therefore, to be more specific, the constitutive parameters might be added an index<br />

that specifies the actual reference coordinate system.<br />

11.2 Two-dimensional formulations<br />

possibilities<br />

coordinates<br />

In tensor formulation the constitutive relation is Tensors<br />

σij = Cijklɛkl<br />

(11.1)<br />

and due to symmetry of the stress tensor σji = σij and of the strain tensor ɛlk = ɛkl<br />

we may contract these tensors to vectors {σ} and {ɛ} with each having only three Vectors<br />

components. The constitutive parameters are then described by the matrix [C] giving<br />

Matrices<br />

211


212 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

{σ} =[C]{ɛ} and {ɛ} =[C] −1 {σ} (11.2)<br />

The square matrix [C] of order three is symmetric and positive definite. To obtain<br />

vectors of invariant length (the same in mutually rotated coordinate systems), we<br />

define the row vectors {σ} T and {ɛ} T by<br />

{σ} T √<br />

:= {σ11 σ22 2σ12} and {ɛ} T √<br />

:= {ɛ11 ɛ22 2ɛ12} (11.3)<br />

knowing that σ2 11 +σ2 22 +2σ2 12 and ɛ2 11 +ɛ2 22 +2ɛ2 √Definition of<br />

12 are not changed by coordinate rota-<br />

2 contraction tion. (This follows from the invariant trace (σ11 + σ22) and the invariant determinant<br />

Constitutive<br />

trace<br />

Constitutive<br />

Frobenius<br />

(σ11σ22 −σ 2 12) of the stress matrix, so (σ11 +σ22) 2 −2(σ11σ22 −σ 2 12) =σ 2 11 +σ 2 22 +2σ 2 12<br />

is an invariant).<br />

In order to be specific we use the four-index notation from the tensor formulation<br />

and then define the matrix [C] by<br />

⎧<br />

⎨<br />

⎩<br />

σ11<br />

σ22<br />

√ 2σ12<br />

⎫<br />

⎬<br />

⎭ =<br />

⎡<br />

⎣<br />

C1111<br />

C1122 √<br />

2C1112<br />

Among the invariants of [C] we have the trace<br />

√<br />

C1122 2C1112 √<br />

C2222 2C2212<br />

√<br />

2C2212 2C1212<br />

Tr([C]) = C1111 + C2222 +2C1212<br />

⎤ ⎧<br />

⎨<br />

⎦<br />

⎩<br />

ɛ11<br />

ɛ22<br />

√ 2ɛ12<br />

⎫<br />

⎬<br />

⎭<br />

(11.4)<br />

(11.5)<br />

The second order norm, named the Frobenius norm F ([C]), is the ”length of the<br />

matrix”, i.e. F 2 is equal to the sum of all the squared matrix components<br />

(length) F 2 = C 2 1111 + C 2 2222 +4C 2 1212 +2C 2 1122 +4C 2 1112 +4C 2 2212 (11.6)<br />

Orthonormal<br />

rotational<br />

Although rather unusual, in relation to optimal material design, it is very convenient<br />

also to contract further the constitutive matrix [C] into a vector {C} with six<br />

component<br />

√<br />

2C1122 2C1112 2C2212} (11.7)<br />

{C} = {C1111 C2222 2C1212<br />

that has invariant length, equal to the Frobenius norm of the matrix [C]<br />

F 2 = {C} T {C} (11.8)<br />

as seen by direct comparison with (11.6).<br />

Furthermore, the vector description of not only stress {σ} and strain {ɛ} but also<br />

of the constitutive parameters {C} makes rotational transformations from a coordinate<br />

system y to another rotated coordinate system x rather simple<br />

transformations {σ}x =[T ]{σ}y , {ɛ}x =[T ]{ɛ}y , {C}x =[R]{C}y (11.9)


Alternative Descriptions of Constitutive Parameter 213<br />

where the components of the matrix [T ] of order three and the components of the<br />

matrix [R] of order six are given by trigonometric combinations, for details see (Pedersen<br />

1995). The matrices [T ] and [R] are both orthonormal matrices so the inverse<br />

transformations to (11.9) are easily found<br />

{σ}y =[T ] T {σ}x , {ɛ}y =[T ] T {ɛ}x , {C}y =[R] T {C}x<br />

(11.10)<br />

Like for other squared matrices, the eigenvalues of [C] give important information.<br />

With [C] defined as in (11.4) these positive invariants can easily be determined<br />

numerically, but the analytical expressions for the general case are rather lengthy. So,<br />

we only give the results for the orthotropic case, i.e. for C1112 = C2212 =0 Constitutive<br />

λ = 1<br />

2 (C1111<br />

<br />

+ C2222 ± (C1111 − C2222) 2 +4C2 1122 ) ,λ=2C1212 (11.11)<br />

Note that the sum of the three eigenvalues is equal to the trace of [C]. The<br />

minimum eigenvalue is the smallest energy density u that any normed state of strain<br />

(ɛ 2 11 + ɛ 2 22 +2ɛ 2 12 = 1) can give, and the maximum eigenvalue gives the largest energy<br />

density.<br />

The total energy density u defined here as the sum of strain energy density uɛ<br />

and stress energy density uσ (also named complimentary energy density) is<br />

which we may write in a linear description<br />

if the vector of second order strains {ˆɛ} is defined by<br />

{ˆɛ} T := {ɛ 2 11 ɛ 2 22<br />

The description (11.14) that for linear elasticity gives<br />

u = uɛ + uσ := {ɛ} T [C]{ɛ} (11.12)<br />

u = {C} T {ˆɛ} (11.13)<br />

√ 2ɛ11ɛ22 2ɛ11ɛ12 2ɛ22ɛ12 2ɛ 2 12} (11.14)<br />

uɛ = uσ = 1<br />

2 {C}T {ˆɛ} (11.15)<br />

is very useful in optimal material design and in modelling evolution of bone materials.<br />

This is due to the fact that sensitivities can often be determined in a fixed strain field<br />

where the change of {ˆɛ} is not involved.<br />

eigenvalues


Analog<br />

in 3D<br />

Eigenvector<br />

description<br />

Elementary<br />

strain cases<br />

214 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

11.3 Three-dimensional formulations<br />

The analog and complete description of the three-dimensional case is not given in this<br />

book, but can be found in (Pedersen and Tortorelli 1998). However, we add some<br />

further aspects that are also of interest for the two-dimensional case.<br />

For 3D-problems the constitutive matrix [L] is a square, symmetric, and semi-<br />

positive definite matrix of order 6. The corresponding contracted vector description<br />

{L} has 21 components. The strain vector {ɛ} has 6 components and the vector of<br />

second order strains {ˆɛ} has 21 components. The total energy density u expressed in<br />

these quantities is<br />

u = {ɛ} T [L]{ɛ} = {L} T {ˆɛ} (11.16)<br />

11.3.1 Spectral decomposition<br />

The constitutive matrix [L] has six non-negative eigenvalues λi and corresponding<br />

mutual orthogonal eigenvectors {Λi} for i =1, 2,...,6. We may choose to order and<br />

normalize them by<br />

0 ≤ λ1 ≤ λ2 ≤ λ3 ≤ λ4 ≤ λ5 ≤ λ6<br />

{Λi} T {Λj} = δij, {Λi} T [L]{Λj} = δijλi (11.17)<br />

where δij is the Kronecker delta. Again with general results from linear algebra we<br />

can describe the constitutive matrix (vector) by the spectral decomposition<br />

[L] =<br />

6<br />

λi{Λi}{Λi} T , {L} =<br />

i=1<br />

6<br />

λi{ ˆ Λi} (11.18)<br />

with {Λi}{Λi} T being the dyadic product of the eigenvector {Λi}. From these dyadic<br />

products we may construct vectors { ˆ Λi} of order 21 in analog to the definition of the<br />

vector {ˆɛ}.<br />

11.3.2 Constitutive parameters by energy densities<br />

Finally we show how constitutive parameters can be described in terms of energy<br />

densities, very much in parallel to the tools used in homogenization theory, and for<br />

the two-dimensional case listed in (Pedersen 1997).<br />

We define only six cases of elementary strain states (here specified by the nonzero<br />

components only) and the corresponding resulting mean strain energy densities<br />

uɛ =ūstrain case<br />

uɛ11 =ū(ɛ11 =1), uɛ22 =ū(ɛ22 =1), uɛ33 =ū(ɛ33 =1)<br />

uɛ12 =ū( √ 2ɛ12 =1), uɛ13 =ū( √ 2ɛ13 =1), uɛ23 =ū( √ 2ɛ23 = 1) (11.19)<br />

i=1


Alternative Descriptions of Constitutive Parameter 215<br />

For linear elastic material the homogenized constitutive parameters are given by<br />

L1111 = 2uɛ11, L2222 =2uɛ22, L3333 =2uɛ33<br />

L1122 = uɛ11+ɛ22 − uɛ11 − uɛ22, L1133 = uɛ11+ɛ33 − uɛ11 − uɛ33<br />

L2233 = uɛ22+ɛ33 − uɛ22 − uɛ33 (11.20)<br />

for the relations between normal components,<br />

2L1212 = 2uɛ12, 2L1313 =2uɛ13, 2L2323 =2uɛ23<br />

2L1213 = uɛ12+ɛ13 − uɛ12 − uɛ13, 2L1223 = uɛ12+ɛ23 − uɛ12 − uɛ23<br />

2L1323 = uɛ13+ɛ23 − uɛ13 − uɛ23 (11.21)<br />

for the relations between shear components, and finally<br />

√ 2L1112 = uɛ11+ɛ12 − uɛ11 − uɛ12, √ 2L1113 = uɛ11+ɛ13 − uɛ11 − uɛ13<br />

√ 2L1123 = uɛ11+ɛ23 − uɛ11 − uɛ23<br />

√ 2L2212 = uɛ22+ɛ12 − uɛ22 − uɛ12, √ 2L2213 = uɛ22+ɛ13 − uɛ22 − uɛ13<br />

√ 2L2223 = uɛ22+ɛ23 − uɛ22 − uɛ23<br />

√ 2L3312 = uɛ33+ɛ12 − uɛ33 − uɛ12, √ 2L2213 = uɛ33+ɛ13 − uɛ33 − uɛ13<br />

√ 2L2223 = uɛ33+ɛ23 − uɛ33 − uɛ23 (11.22)<br />

for the coupling relations between normal and shear components.<br />

Note, that due to the superpositions in, say uɛ11+ɛ22, the relations (11.20), (11.21),<br />

(11.22) are only valid for linear elasticity. Only linear<br />

elasticity<br />

For the material design in chapter 8 we need sensitivities of individual material<br />

constitutive components. Using the relations (11.20), (11.21), (11.22) we can find<br />

these sensitivities from sensitivities of elastic energy, using directly the results in<br />

chapter 12 for the linear case.


3D-bulk<br />

modulus<br />

2D-bulk<br />

modulus<br />

216 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

11.4 Bulk modulus and alternative isotropic parameters<br />

In this section, concentrated on isotropic materials, we primarily derive the relations<br />

between strain energy density, bulk modulus and eigenvalues for constitutive matrices.<br />

11.4.1 Three-dimensional case<br />

In the three-dimensional case, only for the principal components, we have for {σ} =<br />

[L]{ɛ} the 3 × 3 constitutive matrix [L] defined by<br />

[L] =<br />

⎡<br />

E(1 − ν)<br />

⎣<br />

(1 + ν)(1 − 2ν)<br />

1 ˜ν ˜ν<br />

˜ν 1 ˜ν<br />

˜ν ˜ν 1<br />

⎤<br />

⎦ , with ˜ν := ν<br />

1 − ν<br />

(11.23)<br />

and for a pure hydrostatic stress state and a pure dilatational strain state we have<br />

the stress/strain vectors<br />

{σ} T = {¯σ ¯σ ¯σ}, {ɛ} T = {¯ɛ ¯ɛ ¯ɛ}/3 (11.24)<br />

because dilatation ∆V/V is determined by (ɛ11 + ɛ22 + ɛ33) =¯ɛ.<br />

The bulk modulus K is physically defined as the ratio ¯σ/¯ɛ, i.e. from (11.23) and<br />

(11.24)<br />

E(1 − ν)<br />

E<br />

¯σ =<br />

(1 + 2˜ν)1 ¯ɛ ⇒ K =<br />

(11.25)<br />

(1 + ν)(1 − 2ν) 3 3(1 − 2ν)<br />

The strain energy density (uɛ)1 corresponding to the normalized strain state gives<br />

(uɛ)1 = 1<br />

2 {ɛ}T E<br />

[L]{ɛ} =<br />

2(1 − 2ν)<br />

(11.26)<br />

The largest eigenvalue λmax of the constitutive matrix [L] is3K, and thus for the<br />

3D-case we have<br />

K = 2<br />

3 (uɛ)1 = 1<br />

3 λmax<br />

E<br />

=<br />

3(1 − 2ν)<br />

(11.27)<br />

(The eigenvector corresponding to λmax is actually pure dilatation {ɛ} T = {1 11}/ √ 3).<br />

11.4.2 Two-dimensional case<br />

The ”bulk” modulus K2 in the two-dimensional case is depending on the modelling<br />

from 3D to 2D and thus has a less physical definition. With a plane stress assumption


Alternative Descriptions of Constitutive Parameter 217<br />

we have for {σ} =[C]{ɛ} the constitutive matrix [C] and the stress/strain vectors<br />

defined by<br />

E<br />

[C] =<br />

(1 − ν2 <br />

1 ν<br />

) ν 1<br />

and thus with the definition K2 =¯σ/¯ɛ we have<br />

gives<br />

{σ} T = {¯σ ¯σ}, {ɛ} T = {¯ɛ ¯ɛ}/2 (11.28)<br />

E<br />

K2 =<br />

(11.29)<br />

2(1 − ν)<br />

The strain energy density (uɛ)1 for the normalized strain state {ɛ} T = {1 1}/ √ 2<br />

(uɛ)1 = 1<br />

2 {ɛ}T E<br />

[C]{ɛ} =<br />

2(1 − ν)<br />

and the largest eigenvalue λmax (with the eigenvector {ɛ} T = {1 1}/ √ 2) is<br />

λmax =<br />

Thus, the relations for the 2D-case are<br />

E<br />

(1 − ν)<br />

K2 =(uɛ)1 = 1<br />

2 λmax =<br />

E<br />

2(1 − ν)<br />

(11.30)<br />

(11.31)<br />

(11.32)<br />

11.4.3 Relations between other alternative moduli<br />

For isotropic materials, two parameters are enough to describe the constitutive behaviour,<br />

even for the 3D-case. Different combinations of these two parameters are<br />

used for different purposes, and we here relate them to the minimum and the maximum<br />

eigenvalues of the 6 × 6 constitutive matrix [L].<br />

The minimum eigenvalue λmin is twice the shear modulus G that can also be 3D-isotropic<br />

expressed by the longitudinal modulus E and the Poisson’s ratio ν<br />

modulus<br />

relations<br />

λmin =2G = E<br />

=2µ (11.33)<br />

1+ν<br />

where the Lame’ constant µ (describing distortion) is equal to G.<br />

The maximum eigenvalue λmax is (as derived in section 11.4.1) three times the<br />

bulk modulus K that can also be expressed by the longitudinal modulus E and the<br />

Poisson’s ratio ν, or in the Lame’ constant λ and the Poisson’s ratio ν<br />

λmax =3K = E ν<br />

= λ (11.34)<br />

1 − 2ν 1+ν


Collected<br />

results<br />

218 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

The Lame’ constant λ describes the dilatation.<br />

Alternative descriptions for the non-dimensional Poisson’s ratio ν are<br />

ν =<br />

λ<br />

2(λ + µ)<br />

and an alternative description for the Lame’ constant λ is<br />

ν<br />

λ = E<br />

(1 + ν)(1 − 2ν)<br />

1 E E<br />

= (1 − )= − 1 (11.35)<br />

2 3K 2G<br />

(11.36)<br />

as seen from (11.34). In most textbooks on solid mechanics we also find the longitudinal<br />

modulus E expressed in the Lame’ constants<br />

11.5 Summing up<br />

3λ +2µ<br />

E = µ<br />

λ + µ<br />

In this chapter the important results to focus on are:<br />

(11.37)<br />

• Many different descriptions of constitutive relations are possible, and we may<br />

choose what is most simple for specific purposes<br />

• Energy density described as a scalar product (11.13) and (11.16) is important<br />

in sensitivity analysis that can be performed with fixed strain fields<br />

• The formulation with square root definitions as in (11.4) is important for obtaining<br />

invariant lengths of strain vector, stress vectors, and constitutive matrices<br />

or vectors<br />

• The bulk modulus is derived for the 3D-case as well as for the 2D-case, because<br />

this information is needed for the material optimization


Chapter 12<br />

Effective stress/strain and<br />

energy densities<br />

The goal of the present chapter is to put forward some results that can be used more General<br />

generally. We deal with non-linear problems but with the simplest possible extension results<br />

from linear elasticity, which is the power law non-linear elasticity.<br />

Because this class of problems can also be used to describe plasticity theory in the<br />

deformation plasticity formulation and stationary creep, this simple extension covers<br />

a broad range of practical important problems. Non-isotropic description is included Non-linear,<br />

although a number of specific results are given only for the important orthotropic<br />

case.<br />

12.1 Analysis by secant formulation<br />

non-isotropic<br />

The analysis of non-isotropic, non-linear elastic structures/continua are presented in<br />

the secant formulation, as described in (Pedersen and Taylor 1993) and in (Pedersen<br />

1995). We firstly concentrate on the compliance matrix and the definition of effective<br />

stress. The preferred effective stress is defined in an energy-consistent way and the<br />

relation to the von Mises stress is pointed out.<br />

Then an important relation between strain and stress energy densities is established.<br />

This relation is not well known, but was already mentioned by (Hill 1956). We<br />

use the following notation: for the stress/strain vectors {σ} and {ɛ} (for 3D-problems, Notation<br />

each having 6 elements); for the scalar effective stress/strain the notation σeff and<br />

ɛeff ; and for the constant reference modulus of elasticity E. The non-linearity is<br />

described by the powers n or p , where n =1/p.<br />

219


Compliance<br />

power law<br />

description<br />

Sub-vectors,<br />

sub-matrices<br />

General<br />

non-isotropic<br />

Orthotropic<br />

220 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

12.1.1 Non-dimensional compliance matrix<br />

The one-dimensional power law model is often written ɛ =(σ/E) n but with proper account<br />

for signs it should be written ɛ =(|σ| n−1 E −n )σ. In two- and three-dimensional<br />

models we take |σ| = σeff , where σeff ≥ 0 is a scalar measure. Based on this we<br />

take the following compliance description (see end of this chapter for an alternative<br />

derivation of this postulate)<br />

{ɛ} = σ n−1<br />

eff E−n [β]{σ} and σ 2 eff := {σ} T [β]{σ} (12.1)<br />

the non-dimensional matrix [β] describes the non-isotropy, and the only restriction<br />

on this matrix is that it must be symmetric and positive semi-definite, i.e.<br />

[β] T =[β] and σ 2 eff ≥ 0 (12.2)<br />

Now, separating the stress vector into normal terms with index N and shear<br />

terms with index S, we have<br />

√<br />

2σ23}} (12.3)<br />

{σ} T = {{σ} T N {σ} T S } = {{σ11 σ22 σ33}{ √ √<br />

2σ12 2σ13<br />

and, accordingly, we have the following separated [β] matrix<br />

<br />

[β]NN<br />

[β] =<br />

[β]<br />

[β]NS<br />

T NS [β]SS<br />

<br />

with the sub-matrices containing at most 21 different parameters<br />

[β]NN =<br />

⎡<br />

β1111<br />

⎣ β1122<br />

β1122<br />

β2222<br />

β1133<br />

β2233<br />

⎤<br />

⎦<br />

[β]NS = √ 2<br />

[β]SS<br />

=2<br />

⎡<br />

⎣<br />

⎡<br />

⎣<br />

β1133 β2233 β3333<br />

β1112 β1113 β1123<br />

β2212 β2213 β2223<br />

β3312 β3313 β3323<br />

β1212 β1213 β1223<br />

β1213 β1313 β1323<br />

β1223 β1323 β2323<br />

⎤<br />

⎦<br />

(12.4)<br />

⎤<br />

⎦ (12.5)<br />

In the orthotropic directions, the orthotropic case is characterized by the simpler form<br />

of two of the matrices in (12.5)<br />

⎡<br />

⎤<br />

[β]SS =2⎣<br />

⎦ , [β]NS = [0] (12.6)<br />

β1212 0 0<br />

0 β1313 0<br />

0 0 β2323<br />

and is thus described by at most only 9 different parameters.


Effective stress/strain and energy densities 221<br />

The isotropic case is described by only 2 parameters with Isotropic<br />

⎡<br />

⎤<br />

[β]NN =<br />

⎦ , [β]NS = [0]<br />

[β]SS =<br />

⎣ β1111 β1122 β1122<br />

β1122 β1111 β1122<br />

β1122 β1122 β1111<br />

⎡<br />

⎣ β1111 − β1122<br />

0<br />

0<br />

β1111 − β1122<br />

0<br />

0<br />

⎤<br />

⎦ (12.7)<br />

0 0 β1111 − β1122<br />

For the following analysis we list the conditions of incompressibility of an orthotropic<br />

description for any stress state, which is General<br />

incompressible<br />

β1111 + β1122 + β1133 = 0<br />

β1122 + β2222 + β2233 = 0<br />

β1133 + β2233 + β3333 = 0 (12.8)<br />

while incompressibility in relation to hydrostatic pressure is obtained by the single<br />

condition Specific<br />

β1111 + β2222 + β3333 +2(β1122 + β1133 + β2233) = 0 (12.9) incompressible<br />

12.1.2 The von Mises effective stress<br />

In traditional plasticity theory the effective stress is not defined as shown in (12.1)<br />

but instead by means of the von Mises stress σMises defined by Deviatoric<br />

σ 2 Mises := 3<br />

2 {s}T {s} (12.10)<br />

stresses<br />

where {s} is the vector of deviatoric stresses, i.e. the hydrostatic pressure is eliminated.<br />

In (Pedersen 1987a) it is shown that this deviatoric stress vector can be<br />

obtained by a projection with the projection matrix [P ] (i.e. [P ] T =[P ],<br />

[P ], [P ] singular), and we have<br />

[P ][P ]=<br />

Projection<br />

{s} =[P ]{σ} with [P ]:= 1<br />

3<br />

⎡<br />

⎢<br />

⎣<br />

2 −1 −1 0 0 0<br />

− 1 2 −1 0 0 0<br />

− 1 −1 2 0 0 0<br />

0 0 0 3 0 0<br />

0 0 0 0 3 0<br />

0 0 0 0 0 3<br />

⎤<br />

⎥<br />

⎦<br />

(12.11)<br />

matrix


Matrix of<br />

von Mises<br />

Condition for<br />

proportionality<br />

Matrix of<br />

Hill<br />

222 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Inserting (12.11) in (12.10), we get the von Mises stress in terms of the total<br />

stresses<br />

σ 2 Mises = 3<br />

2 {σ}T [P ]{σ} (12.12)<br />

and comparing this to the definition of σ2 eff in (12.1), we find that σMises is proportional<br />

to σeff only for the specific compliance matrix that corresponds to an isotropic<br />

and incompressible material<br />

[β]isotropic and incompressible proportional to [P ] (12.13)<br />

For other materials there is in general a difference between the von Mises stress σMises<br />

and the effective stress σeff based on total energy density. In optimal designs which<br />

take strength constraints into account the solution naturally depends on the chosen<br />

effective measure. Therefore, with σMises not proportional to σeff , we get solutions<br />

related to the specific choice. Note, that the more general results in sensitivity analysis<br />

are based on the energy related definition σeff .<br />

12.1.3 The Hill strength measure<br />

The Hill (Hill 1950) strength reference σHill, for non-isotropic materials is<br />

or in matrix notation<br />

σ 2 Hill = F (σ22 − σ33) 2 + G(σ33 − σ11) 2 + H(σ11 − σ22) 2 +<br />

2Lσ 2 23 +2Mσ 2 13 +2Nσ 2 12<br />

σ 2 Hill = {σ} T<br />

⎡<br />

⎢<br />

⎣<br />

G + H −H −G 0 0 0<br />

− H F + H −F 0 0 0<br />

− G −F F + G 0 0 0<br />

0 0 0 N 0 0<br />

0 0 0 0 M 0<br />

0 0 0 0 0 L<br />

(12.14)<br />

⎤<br />

⎥ {σ} (12.15)<br />

⎥<br />

⎦<br />

We see that for hydrostatic stress {σ} T =¯σ{1 11000} we obtain σHill = 0 so this<br />

measure is insensitive to hydrostatic pressure. Note also that the Hill measure has<br />

only 6 parameters (plus one constraint for hydrostatic pressure), while the orthotropic<br />

[β] matrix has 9 parameters.


Effective stress/strain and energy densities 223<br />

12.2 Strain energy density and<br />

stress energy density<br />

As used in (Pedersen and Taylor 1993), the secant description (12.1) can also be Modulus<br />

stated,<br />

{σ} = ɛ p−1<br />

eff E[α]{ɛ} and ɛ2 eff := {ɛ} T [α]{ɛ} (12.16)<br />

where the non-dimensional constitutive matrix [α] is just the inverse of the compliance<br />

matrix [β] (here assumed positive definite)<br />

[α] =[β] −1<br />

with sub-matrix definitions in complete analogy to (12.5), (12.6) and (12.7).<br />

The integrated strain energy density uɛ based on (12.16) is<br />

(12.17)<br />

power law<br />

description<br />

uɛ = E 1<br />

p +1 ɛp+1<br />

eff<br />

(12.18)<br />

and the stress energy density uσ based on (12.1) is Energy<br />

uσ = 1<br />

En 1<br />

n +1 σn+1<br />

eff<br />

(12.19)<br />

densities<br />

where n =1/p. It follows (most easily from uɛ + uσ = σeff ɛeff ) that we have the<br />

following important relation Energy<br />

uσ = puɛ<br />

(12.20)<br />

and this relation gives rise to simplified sensitivity analysis as well as rather general<br />

optimality criteria, see chapters 13, 14 and 15.<br />

An alternative to start with (12.1) and then derive (12.19) is to start with (12.19)<br />

and then derive (12.1). From<br />

{ɛ} = duσ 1<br />

=<br />

d{σ} T En and from the definition of σ 2 eff<br />

1<br />

n +1 (n +1)σn dσeff<br />

eff<br />

d{σ} T<br />

(12.21)<br />

density<br />

RELATION<br />

in (12.1) Derivation<br />

dσeff<br />

2σeff<br />

=2[β]{σ} (12.22)<br />

d{σ} T<br />

we get directly (12.1) as an alternative to starting with this as a postulate.<br />

from elastic<br />

potential


224 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

12.3 Summing up<br />

In this chapter the important results to focus on are: Collected<br />

results<br />

• Secant relations (12.1) and (12.16) give the energy densities (12.18) and (12.19).<br />

Alternatively, the energy densities give the secant relations.<br />

• Sub-matrix descriptions are valuable to clarify the difference between different<br />

definitions of effective stress/strain.<br />

• For power law, non-linear and non-isotropic elastic materials, the proportionality<br />

(12.20) between strain energy density and stress energy density holds<br />

independently of the level of energy density.


Chapter 13<br />

Elastic potentials, relations<br />

and sensitivities<br />

13.1 Elastic potentials<br />

The general equation of energy equilibrium is Energy<br />

equilibrium<br />

Uɛ + Uσ + Uext = 0 (13.1)<br />

with elastic strain energy Uɛ and elastic stress energy Uσ (elastic complementary en-<br />

ergy) from the corresponding densities uɛ, uσ integrated over the structure/continuum<br />

volume V<br />

<br />

<br />

Uɛ = uɛdV and Uσ = uσdV (13.2)<br />

and the external potential Uext defined by Potential<br />

<br />

Uext := −( TividA + pividV ) (13.3)<br />

with surface tractions Ti, volume forces pi , corresponding displacements vi, and area<br />

A surrounding the volume V .<br />

definitions<br />

For materials resulting in the density relation uσ = puɛ, as proven by (12.20) Basic<br />

in chapter 12, to hold everywhere in the continuum or structure, we have in total<br />

Uσ = pUɛ and the equilibrium (13.1) is simplified to<br />

(1 + p)Uɛ = −Uext<br />

with 0


Total<br />

potential<br />

Potential<br />

relations<br />

Extremum<br />

relations<br />

Using<br />

virtual work<br />

principle<br />

Non-changed<br />

loads<br />

Important<br />

RESULT<br />

226 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Defining the total potential Π by<br />

we have the relations<br />

Π:=Uɛ + Uext<br />

(13.5)<br />

Uɛ = Uσ/p = −Π/p = −Uext/(1 + p) (13.6)<br />

and by p>0 and Uɛ > 0 we have Uσ > 0, Π < 0 and Uext < 0. From this follows<br />

that design for a number of differently stated extremum problems are equivalent and<br />

that their values at the solutions are related as<br />

min Uɛ = min Uσ/p = −max Π/p = −max Uext/(1 + p) (13.7)<br />

with p = 1 for the specific case of linear elasticity.<br />

13.2 Derivatives of elastic potentials<br />

The derivative of the total potential Π with respect to an arbitrary parameter, say a<br />

design parameter h, is<br />

dΠ<br />

dh =<br />

<br />

∂Π<br />

+<br />

∂h fixed strains<br />

∂Π dɛ<br />

∂ɛ dh =<br />

<br />

∂Π<br />

(13.8)<br />

∂h fixed strains<br />

because of stationary total potential ∂Π/∂ɛ = 0 (virtual work principle) with respect<br />

to kinematically admissible strain variations.<br />

For design-independent external loads, (∂Uext/∂h)fixed strains = 0, the definition<br />

(13.5) gives <br />

<br />

∂Π<br />

∂Uɛ<br />

=<br />

(13.9)<br />

∂h fixed strains ∂h fixed strains<br />

and then from (13.6), (13.8) and (13.9) we get the result that is frequently used<br />

<br />

dUɛ ∂Uɛ<br />

= −1<br />

(13.10)<br />

dh p ∂h<br />

fixed strains<br />

For a local design parameter he that only changes the design in the domain e of<br />

the structure/continuum this gives the possibility of a localized determination of the<br />

sensitivity for the total elastic strain energy<br />

dUɛ<br />

dhe<br />

= − 1<br />

p<br />

<br />

∂((ūɛ)eVe)<br />

∂he<br />

fixed strains<br />

(13.11)


Elastic potentials, relations and sensitivities 227<br />

where (ūɛ)e is the mean strain energy density in the domain of he and where Ve is the<br />

corresponding volume. We note that the only difference between linear (p = 1) and<br />

non-linear material is the factor 1/p, and for a condition on stationarity dUɛ/dhe =0,<br />

p has no influence.<br />

Note, that the sensitivity is not physically localized, but still we can without approximation<br />

determine the sensitivity localized. Two specific cases are exemplified in<br />

more detail. Local<br />

13.2.1 Constant constitutive matrix, a specific case<br />

For the case of a constant constitutive matrix (non-changed material) in a fixed strain<br />

field, the result (13.11) simplifies to<br />

dUɛ<br />

= −<br />

dhe<br />

(ūɛ)e ∂Ve<br />

(13.12)<br />

p ∂he<br />

and often ∂Ve/∂he is simply equal to Ve/he (or proportional to), e.g. when he is a<br />

thickness, a relative volume density, or an area.<br />

13.2.2 Constant volume, a specific case<br />

For the case of constant volume, the result (13.11) simplifies to<br />

dUɛ<br />

= −<br />

dhe<br />

Ve<br />

<br />

∂(ūɛ)e<br />

p ∂he<br />

fixed strains<br />

(13.13)<br />

Two such even more specific cases should be mentioned: firstly the case of sensitivity<br />

to material orientation θe, for which ∂(ūɛ)e/∂θe must be evaluated, and secondly<br />

sensitivity to a specific constitutive parameter Lijkl, for which ∂(ūɛ)e/∂Lijkl must be<br />

evaluated. Due to the fact that these sensitivities are determined in a fixed strain<br />

field it is possible to obtain analytical expressions, as shown in chapters 15, 17 and<br />

used in chapters 8, 9, 10.<br />

design<br />

parameter


Collected<br />

results<br />

228 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

13.3 Summing up<br />

In this chapter the important results to focus on are:<br />

• Due to the simple relations (13.6) between potentials, alternative statements of<br />

optimal design problems are possible.<br />

• In an energy formulation, and thus with great generality, the possibility of<br />

determining sensitivities in a fixed strain field (13.10) is proven.<br />

• Local design parameters often give further simplifications by, more or less,<br />

bringing calculation from system (total structure/continuum) level to element<br />

(domain) level.


Chapter 14<br />

Some necessary conditions for<br />

optimality<br />

In this chapter we primarily present two necessary conditions for optimal solutions in<br />

general, i.e. not directly related to optimal design. The first condition is only valid General<br />

for non-constrained problems and the second condition is valid for problems with only<br />

a single constraint in addition to the objective.<br />

After the two first sections we then determine, expressed in physical terms, the<br />

condition for solution of the most simple optimal design problems, for which sizes (or<br />

field of size) optimize stiffness as well as strength. The optimal designs to these two<br />

different problems are shown to be the same. Design<br />

In close relation to the analysis for size optimization, we treat optimization of<br />

shape, again to optimize stiffness as well as strength. The shape solution to these two<br />

different problems is shown in many cases also to be the same, very much in parallel<br />

optimization<br />

optimization<br />

to the result for size optimization. Size and<br />

shape<br />

A simple parametrization for shape optimization is finally described, this approach<br />

is used in the examples presented in chapter 7. The goal of the present chapter is to<br />

obtain basic understanding of very simple optimal design problems, without involving<br />

extended numerical calculations.<br />

14.1 Non-constrained problems<br />

A non-constrained optimization problem may be defined as<br />

Extremize Φ=Φ(he) with variables he non − constrained (14.1)<br />

229


Stationary<br />

objective<br />

Converted to<br />

230 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

and a necessary condition for this is that the objective Φ is stationary with respect<br />

to all the independent variables he<br />

dΦ/dhe =0for all e (14.2)<br />

The optimization of material orientation is an important example of such a nonconstrained<br />

problem. Unfortunately, this problem is not simple to solve, because<br />

many local optima exist. Each variable (orientational angle θe in a domain e) has<br />

several solutions to the stationarity condition (14.2), and this inherent problem is<br />

described in detail in chapter 17.<br />

14.2 Problems with a single constraint<br />

An optimization problem with only a single constraint may be defined as<br />

Extremize Φ = Φ(he) with variables he<br />

constrained by g = g(he) = 0 (14.3)<br />

To obtain an optimality condition we convert this problem to a non-constrained<br />

non-constrained problem, using a Lagrangian function L =Φ− λg to be made stationary for arbitrary<br />

value of λ (the Lagrangian multiplier). It follows from (14.2) that a necessary<br />

optimality condition is<br />

Proportional<br />

gradients<br />

dL/dhe = 0 for alle ⇒<br />

dΦ/dhe = λdg/dhe (14.4)<br />

This general result we read as proportionality between the gradient of the objective<br />

and the gradient of the single constraint. The factor of proportionality, the Lagrangian<br />

multiplier λ, is determined by the constraint condition g(he) = 0, that follows from<br />

dL/dλ = −g =0.<br />

The use of the optimality condition (14.4) for obtaining more general information<br />

about optimal design is very important. It should be noted that behind this result is<br />

the assumption that the constraint is active. Extensions to two and more constraints<br />

are possible, but the uncertainty about the active constraints is then often a limiting<br />

factor on the usefulness.<br />

An alternative look at the problem is to postulate (14.4) and then see that dg =0<br />

implies dΦ = 0, i.e.<br />

dΦ = ∂Φ<br />

∆he = λ<br />

∂he<br />

∂g<br />

∆he = λdg = 0 (14.5)<br />

∂he<br />

e<br />

e


Some necessary conditions for optimality 231<br />

14.3 Size optimization for stiffness and strength<br />

The theoretical results for size optimization are more developed than those for shape<br />

optimization. Let us therefore start with some basis knowledge from size optimization,<br />

as it can be found in (Pedersen 1998) for non-linear elasticity or in (Wasiutynski 1960)<br />

for linear elasticity.<br />

14.3.1 Size design with optimal stiffness<br />

If the objective is to minimize compliance (minimize elastic energy) for given total<br />

mass then we have (for optimal stiffness design with homogeneous assumptions and Homogeneous<br />

design independent loads): the ratio between sub-domain energy and sub-domain mass mass (volume)<br />

should be the same in all the design sub-domains.<br />

dependence<br />

Let the design parameters be he, then homogeneous mass relations are obtained<br />

with M = <br />

e Me = <br />

e hme ¯ Me, where M is the total mass, Me is the mass in domain<br />

e, m is a given positive value, and ¯ Me is independent of the design parameters. The<br />

homogeneous energy relations are obtained with Uɛ = <br />

Uɛe e = e hne Ūɛe , where Uɛ<br />

is the total strain energy, Uɛe is the strain energy in domain e, n is a given positive<br />

value, and Ūɛe is explicitly independent of the design parameters.<br />

Restricted to problems with constant mass density we get, in all design domains,<br />

the same mean strain energy density. Furthermore, if the model has constant energy<br />

density within a design domain, then the result for the optimal design is uniform Stiffest<br />

strain energy density u design<br />

∗ ɛ , i.e.<br />

u ∗ ɛe =ūɛ for all free design domains (14.6)<br />

where lower and upper size constraints are not reached. The symbolism here is a<br />

super-index ∗ related to the optimal design, and a overhead bar ¯indicating a constant<br />

value for each domain e (mean value).<br />

Assume now that the necessary condition (14.6) give a global minimum solution,<br />

then for any other design the total strain energy Uɛ is larger (or equal to)<br />

Uɛ = <br />

uɛeVe ≥ U ∗ ɛ = <br />

e<br />

e<br />

u ∗ ɛ V ∗<br />

e =ūɛ<br />

<br />

e<br />

V ∗<br />

e =ūɛ<br />

<br />

Ve = <br />

e<br />

e<br />

u ∗ ɛe Ve<br />

(14.7)<br />

where V ∗<br />

e is the optimal volume of the design domain e. For an alternative design<br />

with design volumes Ve we have the same total volume V V = <br />

e Ve = ∗<br />

e Ve . From<br />

(14.7) we get<br />

<br />

(uɛe − u∗ɛe )Ve ≥ 0 (14.8)<br />

e


Also best<br />

232 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

14.3.2 Size design with optimal strength<br />

strength With positive volumes Ve we read from (14.8), that at least one uɛe is not less than<br />

u∗ ɛe . Thus if the strongest design is defined by minimum of maximum uɛe , then the<br />

stiffest design characterized by the optimality condition (14.6) is also the strongest<br />

design.<br />

We note that the strength may also be defined in relation to the von Mises stress<br />

or an alternative effective stress, and these measures are not always proportional to<br />

the energy density. For a detailed discussion of these aspects see (Pedersen 1998).<br />

General<br />

knowledge<br />

Localized<br />

volume change<br />

14.4 Shape optimization for stiffness and strength<br />

In the following we use the same kind of reasoning to draw conclusions about shape<br />

optimization, without involving a solution to the actual stress problem. Thus we gain<br />

general knowledge, valuable for 3D and 2D-problems, for non-linear elastic as well as<br />

for linear problems, for non-isotropic or isotropic problems, for any external, design<br />

independent load. Also valid for non-homogeneous problems and independent of the<br />

solution procedure.<br />

In order to simplify the mathematics the design parametrization is chosen as<br />

illustrated in figure 14.1. An alternative parametrization with expansion in terms of<br />

shape design functions is formulated in (Dems and Mroz 1978), a paper closely related<br />

to this presentation.<br />

We assume a homogeneous state for the strain energy density uɛe<br />

within the<br />

volume Ve related to the shape parameter he, say a constant stress finite element.<br />

Let us now subject the shape to variation using only two parameters hi and hj.<br />

Furthermore, let the total volume V of the structure (continuum) be fixed, then<br />

∆V = dV<br />

∆hi +<br />

dhi<br />

dV<br />

dhj<br />

∆hj = dVi<br />

dhi<br />

∆hi + dVj<br />

∆hj = 0 (14.9)<br />

dhj<br />

because we also assume the domain volumes to be depending only on one design<br />

parameter and with a positive gradient (to be used later)<br />

Ve = Ve(he) and dVe/dhe > 0 (14.10)<br />

14.4.1 Shape design with optimal stiffness<br />

In shape optimization for extremum elastic strain energy the increment of the objective<br />

corresponding to increments ∆hi, ∆hj is<br />

∆Uɛ = dUɛ<br />

dhi<br />

∆hi + dUɛ<br />

∆hj<br />

dhj<br />

(14.11)


Some necessary conditions for optimality 233<br />

Figure 14.1: Discretized design parametrization, showing two design domains i and<br />

j.<br />

which for power law non-linear elasticity σ = Eɛp can be written as Design<br />

∆Uɛ = − 1<br />

<br />

∂Uɛ<br />

∆hi +<br />

p ∂hi<br />

∂Uɛ<br />

<br />

∆hj<br />

∂hj fixed strains<br />

(14.12)<br />

This is proven in (Pedersen 1998) for design independent loads, and follows from<br />

(13.10) in chapter 13. Therefore only the local energies Uɛi = uɛiVi and Uɛj = uɛj Vj<br />

are involved and the variations in the strain energy densities need not be determined,<br />

because the constitutive relations are unchanged. We have Localized<br />

∆Uɛ = − 1<br />

p (uɛi<br />

dVi dVj<br />

∆hi + uɛj ∆hj)<br />

dhi dhj<br />

and inserting (14.9) in (14.13) we obtain<br />

(14.13)<br />

∆Uɛ = − 1<br />

(uɛi − uɛj )dVi ∆hi<br />

p dhi<br />

(14.14)<br />

A necessary condition for optimality ∆Uɛ = 0 with dVi/dhi > 0 is therefore<br />

uɛi = uɛj .<br />

With all design parameters, eq. (14.9) and (14.13) are written<br />

∆V = dVe<br />

∆he<br />

dhe<br />

e<br />

independent<br />

loads<br />

energy change


Constant<br />

energy density<br />

Basic<br />

assumption<br />

Detail of<br />

proof<br />

234 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

∆Uɛ = − 1<br />

p<br />

<br />

e<br />

uɛe<br />

dVe<br />

∆he<br />

dhe<br />

(14.15)<br />

and we conclude that a necessary condition for optimality ∆U = 0 with constraint<br />

∆V = 0 is constant strain energy density uɛe . Thus for the stiffest design the energy<br />

density along the shape(s) to be designed, here denoted uɛs , must be constant<br />

14.4.2 Shape design with optimal strength<br />

uɛs =ūɛ (14.16)<br />

We now relate the stiffest design (minimum compliance) to the strongest design (minimum<br />

maximum strain energy density). Let us assume that the highest strain energy<br />

density is at the shape to be designed. With index s referring to shape design domains<br />

and index n referring to domains not subjected to design changes, this means that<br />

for the stiffest design we assume<br />

uɛs =ūɛ >uɛn (14.17)<br />

A design domain that depends on design parameter is given index s (hs) and a design<br />

domain which is not subjected to design change is given index n (hn). For the total<br />

design domain we use index S and for the total domain not subjected to design, index<br />

N. The total elastic strain energy Uɛ is obtained from<br />

<br />

Uɛ = UɛS + UɛN = Uɛs + Uɛn = uɛsVs + <br />

uɛnVn, i.e.<br />

<br />

Vs + <br />

s<br />

Uɛ = ūɛ<br />

uɛn<br />

s n<br />

Vn<br />

With unchanged domain N and for the stiffest design Uɛ >U∗ ɛ we obtain<br />

as <br />

s<br />

<br />

uɛs<br />

s<br />

Vs + <br />

uɛn<br />

n<br />

Vn > <br />

ūɛV<br />

s<br />

∗<br />

s + <br />

<br />

(uɛs<br />

n<br />

− ūɛ)Vs > <br />

(u ∗ (14.19)<br />

ɛn − uɛn )Vn<br />

s<br />

n<br />

n<br />

s<br />

n<br />

u ∗ ɛn V ∗ n , i.e.<br />

(14.18)<br />

ūɛV ∗<br />

s = <br />

s ūɛVs due to given total volume, and furthermore individual un-<br />

changed in the non-design domains V ∗ n = Vn.<br />

The right hand side might be negative, so we can not directly draw conclusions<br />

as from (14.8). However, in a complementary formulation with stress energies we can<br />

prove that the right hand side is non-negative and then the proof holds.<br />

The proof of increasing energy in the shape domain is as follows. We write the


Some necessary conditions for optimality 235<br />

total stress energy Uσ as the sum of stress energy in the shape domain UσS and stress<br />

energy in the non-shape domain UσN and obtain<br />

Uσ = UσS<br />

+ UσN ⇒ dUσ<br />

dh<br />

= dUσS<br />

dh<br />

From the principle of complementary virtual work follows<br />

dUσ/dh =(∂Uσ/∂h)fixed stress field and we get<br />

dUσ<br />

dh =<br />

∂UσS<br />

∂h<br />

+ ∂UσN<br />

∂h<br />

<br />

+ dUσN<br />

dh<br />

fixed stress field<br />

(14.20)<br />

(14.21)<br />

where the last term is zero when h has no direct influence on the non-shape domain.<br />

Finally for the stiffest design we have dUσ/dh > 0 and from this we conclude<br />

∂UσS<br />

=<br />

∂h fixed stress field<br />

dUσ<br />

dh<br />

1 dUɛS<br />

=<br />

p dh<br />

> 0 (14.22)<br />

Summarizing the theoretical results of this section; we have for the general threedimensional<br />

case with non-isotropic, power law non-linear elastic material in an nonhomogeneous<br />

structure, and for any design independent single load case that:<br />

The minimum compliance shape design (stiffest shape design) has uniform energy<br />

density along the designed shape, as far as the geometrical constraints make this<br />

possible.<br />

If we furthermore assume that the highest energy densities are found at the designed<br />

shape, then the stiffest design is also the strongest design, as defined by a design<br />

which minimizes the maximum energy density. Design for<br />

Note that these results are obtained without calculating the stress/strain fields<br />

and without specifying the constitutive behaviour. This behaviour need not be homogeneous<br />

and thus we can also include the multi-material case. The many different<br />

problems in chapter 7 support these results.<br />

14.5 Conditions with a<br />

simple shape parametrization<br />

stiffness<br />

and strength<br />

In the final conclusions in section 14.4.2, we have added the note ”as far as the<br />

geometrical constraints make this possible”. Also it was commented that normally<br />

the shape parametrization implies such a geometrical constraint. In this section we<br />

use a simple shape parametrization that makes a rather simple optimality condition<br />

possible. The limitations of using this simple parametrization can be evaluated by<br />

the possibility to obtain almost uniform energy density distribution along the shape Good<br />

experience


Two or<br />

only one<br />

parameter(s)<br />

236 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 14.2: A three parameter (α, β, η) description of an internal hole in a rectangular<br />

domain, specified by A, B.<br />

to be designed. The examples in chapters 7 and 8 illustrate that the parametrization<br />

is in fact able to describe optimal shapes in many cases.<br />

Figure 14.2 shows a single inclusion hole, where the shape of the boundary is modelled<br />

as a super-elliptic shape, described by only three non-dimensional parameters,<br />

relative axes α, β and power η<br />

<br />

x<br />

<br />

η<br />

η<br />

y<br />

+ = 1 (14.23)<br />

αA βB<br />

With known area of the hole we only have two parameters and if furthermore symmetry<br />

is enforced, say αA = βB, we only have one free parameter, which might be<br />

the power η. Figure 14.3 shows the great flexibility even for this one parameter<br />

description. This parametrization naturally has its limitation, but several examples<br />

show its usefulness, and furthermore it can easily be extended to 3D-problems by<br />

<br />

x<br />

<br />

η<br />

η η y z<br />

+ + = 1 (14.24)<br />

αA βB γC<br />

In the 2D-model (14.23) the area of the hole is<br />

αA <br />

4 βB 1 − (<br />

0<br />

x<br />

αA )η 1/η ) dx =2αβABg(η) (14.25)<br />

with the function g = g(η) defined by<br />

<br />

1 η +1 2<br />

g(η) :=Γ Γ /Γ<br />

(14.26)<br />

η η η


Some necessary conditions for optimality 237<br />

Figure 14.3: Shapes giving equal area of the hole, with powers of the super-elliptic<br />

shape being η = 0.75, 1.25. 1.75 and 3.00, respectively.<br />

where Γ is the Gamma-function. With the rectangular area being 4AB the relative<br />

area of the hole φ (relative to the area 4AB) and the relative area of the solid (relative Relative<br />

density) ρ are<br />

φ = 1<br />

αβg(η) =1− ρ (14.27)<br />

2<br />

An optimal design problem is formulated in order to extremize the elastic energy<br />

U for constant relative area<br />

hole area<br />

or density<br />

Extremize U subject to φ(α, β, η) = ¯ φ (14.28)<br />

Within the possibilities of the three parameters α, β, η this also minimizes energy Design<br />

concentration and returns constant energy density along the boundary of the hole,<br />

as discussed in section 14.4.2. Using the result (14.12) from sensitivity analysis we<br />

determine the differential of the elastic energy (p = 1 for linear elasticity)<br />

dU = − 1<br />

<br />

∂U ∂U ∂U<br />

dα + dβ +<br />

p ∂α ∂β ∂η dη<br />

<br />

(14.29)<br />

fixed strains<br />

and the differential of the constraint follows from (14.27) (using a formula manipulation<br />

program to differentiate the Gamma-functions)<br />

<br />

dα dβ p(η)dη<br />

dφ = φ + +<br />

α β η2 <br />

(14.30)<br />

problem


Available<br />

functions<br />

<strong>Optimal</strong>ity<br />

condition<br />

For model<br />

by the FEM<br />

Strength or<br />

stiffness<br />

238 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

with the function p = p(η) defined by<br />

<br />

2 1 η +1<br />

p(η) :=Ψ − Ψ − Ψ<br />

(14.31)<br />

η η η<br />

where Ψ is the Psi-function. To illustrate that the functions g(η) and p(η) are wellbehaved<br />

functions we show in figure 14.4 these functions and there derivatives. These<br />

functions are available in many libraries of computer routines.<br />

Figure 14.4: Left g-function and right the p-function with their derivatives as a<br />

function of the shape power η.<br />

The condition of dU = 0 when dφ = 0 is a necessary condition for optimality and<br />

thus (as in general with only a single constraint) we from (14.29) and (14.30) get the<br />

optimality condition by proportional gradients (14.4), i.e.<br />

<br />

α ∂U<br />

<br />

∂U η2 ∂U<br />

= β = (14.32)<br />

∂α ∂β p(η) ∂η fixed strains<br />

In a fixed strain field the energy densities u are constant and only the volumes of<br />

domains (elements) connected to the hole boundary change. Thus in a finite element<br />

formulation the optimality condition (14.32) is written<br />

α ∂Vs<br />

∂Vs η2 ∂Vs<br />

us = β us = us<br />

(14.33)<br />

∂α ∂β p(η) ∂η<br />

s<br />

s<br />

s<br />

where index s refers to an element connected to the hole boundary. The only information<br />

needed in addition to the results from analysis is ∂Vs/∂α, ∂Vs/∂β, ∂Vs/∂η,<br />

i.e. only information from geometry. We note, in agreement with section 14.4.2, that<br />

if us is constant along the hole boundary then <br />

s ∂Vs/∂α = ∂V/∂α = φ/α etc., and<br />

the optimality criterion (14.33) is satisfied by usφ = usφ = usφ. Thus a constant<br />

energy density along the boundary of the hole implies stationary total elastic energy.<br />

However, we can have stationary energy without constant energy density, if the<br />

possible designs are restricted. This is illustrated by the examples in chapter 7.


Some necessary conditions for optimality 239<br />

14.5.1 Possible iterative procedure<br />

The problem is how to find a boundary shape that satisfies (14.32) or in finite element<br />

formulation (14.33). The heuristic approach of successive iterations could be to<br />

estimate the Lagrange multiplier λ by the mean value<br />

λestimated = 1<br />

3<br />

<br />

α ∂U ∂U<br />

+ β<br />

∂α ∂β<br />

<br />

η2 ∂U<br />

+<br />

p(η) ∂η<br />

(14.34)<br />

and then redefine α, β, η by Estimated<br />

αnew = λ/( ∂U<br />

∂α )old, βnew = λ/( ∂U<br />

∂β )old, ( η2<br />

p(η) )new = λ/( ∂U<br />

∂η )old<br />

with iterations on λ to satisfy the constraint of (14.28)<br />

14.5.2 Extended design space<br />

(14.35)<br />

φnew = 1<br />

2 αnew βnew g(ηnew) = ¯ φ (14.36)<br />

Let us assume that the three-parameter boundary shape does not in a satisfactory way<br />

give constant energy density along the boundary shape, i.e. the energy concentration<br />

must be made smaller. We can then add modification functions fk = fk(s) where s is<br />

the natural coordinate along the super-elliptic shape. The functions fk (k =1, 2, ...)<br />

are chosen as described in (Pedersen 1988) and used in (Pedersen et al. 1992). Figure<br />

multiplier<br />

6.9 in section 6.5.1 illustrates in detail this approach. The added design parameters Modification<br />

are the amplitudes ck of the modification functions and gradients ∂φ/∂ck and ∂Vs/∂ck<br />

must be evaluated, numerically or eventually analytically. Better solutions to problem<br />

(14.28) can then be obtained. Alternatively we may formulate the problem directly<br />

as minimum energy concentration problem<br />

Minimize by α, β, η, ck<br />

over all domains e<br />

subject to φ(α, β, η, ck) = ¯ φ (14.37)<br />

maximum uɛe<br />

Solutions can be obtained with the integrated FEM-SLP approach, see (Pedersen<br />

1981) and (Pedersen et al. 1992). With the modification functions we also have the<br />

possibility to introduce discontinuities in the slope of the boundary shape.<br />

functions


Collected<br />

results<br />

240 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

14.6 Summing up<br />

In this chapter the important results to focus on are:<br />

• For non-constrained problems the necessary optimality condition is stationarity<br />

of the objective with respect to all the independent variables.<br />

• For problems with a single constraint the necessary optimality condition is<br />

proportionality between the gradient of the objective and the gradient of the<br />

single constraint.<br />

• For size optimization of stiffness and strength the stiffest design is characterized<br />

by the optimality condition of uniform energy density and this design is also<br />

the strongest design.<br />

• For shape optimization of stiffness, the minimum compliance shape design<br />

(stiffest shape design) has uniform energy density along the designed shape,<br />

as far as the geometrical constraints make this possible.<br />

• For shape optimization of strength, if we assume that the highest energy densities<br />

are found at the designed shape, then the strongest design, as defined by a<br />

design which minimizes the maximum energy density, is also the stiffest design.<br />

• For shape optimization a simple super-elliptic description makes it possible to<br />

design a wide spectrum of shapes, and the analytical treatment of this case is<br />

almost as simple as the classic elliptic case.


Chapter 15<br />

The ultimate optimal material<br />

15.1 The individual constitutive parameters<br />

In ultimate optimal material design, also named free material design, we represent the<br />

material properties in the most general form possible for an elastic continuum, namely Free<br />

the unrestricted set of components in positive semi-definite constitutive matrices.<br />

For a given material (given constitutive relations), we normally measure cost by<br />

the amount of material, say by thickness or density. With the free material we need<br />

a measure of the ”amount of a matrix”, and cost is then measured on the basis of<br />

invariants of these matrices.<br />

With reference to the paper by (Bendsøe et al. 1994), we extend the results obtained<br />

in that paper to be valid also for power law non-linear elasticity, as done in<br />

(Pedersen 1998). If we choose as cost constraint the Frobenius norm (length of a ma-<br />

material<br />

trix) of the constitutive matrix, then the analytical proof of the optimal constitutive Frobenius<br />

matrix, even for 3D-problems, is rather direct.<br />

norm<br />

15.2 Sensitivity analysis<br />

With localized sensitivity analysis as shown in chapter 13, and also given specifically<br />

by (13.13), we have<br />

<br />

<br />

dUɛ ∂Uɛ<br />

1 ∂uɛ<br />

= −1<br />

= −Ve<br />

(15.1)<br />

dh p ∂h fixed strains p ∂he fixed strains<br />

where Ve is the volume of the domain of the localized design variable he, (here a Localized<br />

component of the constitutive matrix). Thus minimum total strain energy Uɛ implies<br />

241<br />

determined<br />

sensitivities


Design<br />

problem<br />

Principal<br />

strains only<br />

Direct<br />

conclusions<br />

Positive<br />

definite<br />

242 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

maximum strain energy density uɛ in a fixed strain field. (In domains of non-constant<br />

strain energy density, the notion of mean value ūɛ should be used). The strain energy<br />

density depends homogeneously on the squared effective strain ɛeff , see (12.18) in<br />

chapter 12. The problem formulation can therefore be stated as<br />

Maximize ɛ 2 eff := {ɛ} T [α]{ɛ} subject to F robenius([α]) = 1 (15.2)<br />

where the matrix [α] describes the non-dimensional part of the constitutive matrix in<br />

the secant formulation (12.16).<br />

In the invariant formulation for [α] we can choose the coordinate system of principal<br />

strains<br />

{ɛ} T = {{ɛ1 ɛ2 ɛ3} {000}} (15.3)<br />

and obtain<br />

ɛ 2 eff = {ɛ} T [α]{ɛ} = {ɛ1 ɛ2 ɛ3}<br />

⎡<br />

⎣ α1111 α1122 α1133<br />

α1122 α2222 α2233<br />

α1133 α2233 α3333<br />

⎤ ⎧<br />

⎨<br />

⎦<br />

⎩<br />

ɛ1<br />

ɛ2<br />

ɛ3<br />

⎫<br />

⎬<br />

⎭<br />

(15.4)<br />

Now, the Frobenius norm of a matrix is defined as the square root of the sum<br />

of the squares of all the elements of the matrix (equal to the squared length of the<br />

contracted vector). It thus follows directly that for optimality, the matrix elements<br />

not involved in (15.4) must be zero. This means directly that also for the non-linear,<br />

power law materials we have:<br />

• the optimal material is orthotropic<br />

• principal directions of material, strain and stress are aligned<br />

• there is no shear stiffness<br />

This result for linear elastic material is proven in (Bendsøe et al. 1994), based<br />

also on a constraint on the trace of the constitutive matrix. Here, the extension to<br />

non-linear elastic material follows directly from the localized sensitivity result (15.1).<br />

For simplicity of proof we have chosen the Frobenius norm as the constraint.<br />

15.3 Final optimization<br />

The further analysis relates only to the sub-matrix in (15.4). To fulfill the condition<br />

of being positive definite, we have as necessary conditions<br />

α1111 > 0, α2222 > 0, α3333 > 0<br />

α1111α2222 >α 2 1122, α1111α3333 > α 2 1133, α2222α3333 >α 2 2233 (15.5)


The ultimate optimal material 243<br />

The problem formulation (15.1) can now be written as<br />

Maximize ɛ 2 eff = α1111ɛ 2 1 + α2222ɛ 2 2 + α3333ɛ 2 3 +<br />

2α1122ɛ1ɛ2 +2α1133ɛ1ɛ3 +2α2233ɛ2ɛ3<br />

(15.6)<br />

constrained by (15.5) and by given Frobenius norm<br />

F<br />

New design<br />

problem<br />

2 − 1=α 2 1111 + α 2 2222 + α 2 3333 +2α 2 1122 +2α 2 1133 +2α 2 2233 − 1 = 0 (15.7)<br />

The general necessary condition for optimality is proportional gradients<br />

chapter 14), i.e. for this specific case<br />

(see <strong>Optimal</strong>ity<br />

condition<br />

which gives the result<br />

ɛ 2 1<br />

α1111<br />

d(ɛ 2 eff )/dαiijj = λd(F 2 )/dαiijj<br />

= ɛ22 =<br />

α2222<br />

ɛ23 =<br />

α3333<br />

ɛ1ɛ2<br />

=<br />

α1122<br />

ɛ1ɛ3<br />

=<br />

α1133<br />

ɛ2ɛ3<br />

α2233<br />

(15.8)<br />

(15.9)<br />

and we can finally write the resulting constitutive matrix in the directions of principal<br />

strains/stresses (evaluating λ to satisfy (15.7): <strong>Optimal</strong><br />

modulus<br />

matrix<br />

1<br />

[α]optimal =<br />

(ɛ1 + ɛ2 + ɛ3) 2<br />

⎡<br />

⎤<br />

⎢<br />

⎥<br />

⎢<br />

⎥<br />

⎢<br />

⎥<br />

⎢<br />

⎥<br />

⎢<br />

⎥ (15.10)<br />

⎢<br />

⎥<br />

⎣<br />

⎦<br />

ɛ 2 1 ɛ1ɛ2 ɛ1ɛ3 0 0 0<br />

ɛ1ɛ2 ɛ 2 2 ɛ2ɛ3 0 0 0<br />

ɛ1ɛ3 ɛ2ɛ3 ɛ 2 3 0 0 0<br />

0 0 0 0 0 0<br />

0 0 0 0 0 0<br />

0 0 0 0 0 0<br />

15.4 Numerical aspects and comparison with isotropic<br />

material<br />

The result (15.10) is valid also for power law, non-linear elastic materials. We note<br />

that the matrix in (15.10) has only one non-zero eigenvalue and that the material<br />

therefore only has stiffness in relation to the specified strain condition. For the ultimate<br />

optimal material, the effective strain ɛeff , the strain energy density uɛ, and the<br />

Frobenius norm F are<br />

ɛ 2 eff = ɛ 2 1 + ɛ 2 2 + ɛ 2 3<br />

uɛ = E 1<br />

p +1 (ɛ2 1 + ɛ 2 2 + ɛ 2 3) (p+1)/2<br />

F = 1 (15.11)


Collected<br />

results<br />

244 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

We can obtain the same effective strain and strain energy density with an isotropic,<br />

zero Poisson’s ratio material [α] =[I], but then the corresponding Frobenius norm is<br />

F = 6, i.e. the material cost is six times greater. As shown in (Bendsøe et al. 1994),<br />

the zero Poisson’s ratio material may be valuable in numerical calculation, because<br />

of the degeneracy of the ultimate optimal material.<br />

All the examples in chapter 4 are calculated with zero Poisson’s ratio material, and<br />

therefore may be interpreted as examples of ultimate optimal material distributions.<br />

15.5 Summing up<br />

In this chapter the important results to focus on are:<br />

• The ultimate optimal material is very degenerate and is only stable in relation<br />

to the specific strain state for which it is designed.<br />

• The obtained solution is also valid for power law non-linear elastic materials,<br />

and simple arguments lead to the obtained analytical solution.<br />

• The direct comparison with isotropic, zero Poisson’s ratio material is most interesting,<br />

and can be used for obtaining numerical solutions to specific problems.<br />

For specific problems this is demonstrated in chapter 4.


Chapter 16<br />

Conditions for statically<br />

determinate trusses<br />

In the early papers on optimal design of truss topology by (Dorn et al. 1964) and<br />

by (Fleron 1964), the optimization of truss topology was formulated as a linear programming<br />

problem with bar forces as unknowns. From this follows that there exist<br />

statically determinate optimal trusses. These formulations are based on design in- Early<br />

dependent allowable stresses, possibly different for each bar and possibly different for<br />

bars in tension and bars in compression, but given in advance.<br />

Extension to include compressive allowable stresses that account for local stability<br />

references<br />

of the bars is presented in (Pedersen 1969) and (Pedersen 1970), and still statically Extensions:<br />

determinate trusses resulted. In (Pedersen 1992) and (Pedersen 1993) unknown supports<br />

with individual costs were included in the formulation, and the presentation<br />

below very much follows these 92-93 proceedings.<br />

In a statically determinate truss the force in each bar is determined only by<br />

stability,<br />

supports<br />

equilibrium and is not dependent on the size of the bars. Therefore, the size of each Individual<br />

bar can be designed individually and for a minimum cost solution each bar must be<br />

designed to its minimum size, i.e. it must be fully stressed.<br />

In the formulation below we assume that each bar carries two forces, a tensile<br />

force and a compressive force which are both non-negative. In the minimum cost<br />

truss either the tensile or the compressive force is zero, but in the formulation the<br />

account for two independent forces is important. In general the presentation in this<br />

chapter is closely related to the theory of linear programming, and therefore may<br />

seem complicated if this theory is unknown to the reader. However, the chapter is<br />

written to be be self-contained.<br />

245<br />

bar design


Assumptions<br />

Statically<br />

determinate<br />

solution<br />

3D<br />

Equilibrium<br />

at joints<br />

Unknown<br />

support<br />

reactions<br />

246 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

16.1 Theorem for the single load case<br />

If each bar n of the truss is designed with a minimum cost φn (fully<br />

stressed), which as a function of the actual non-negative bar force Pn,<br />

satisfies<br />

dφn<br />

dPn<br />

then there exists at least one statically determinate truss that minimizes<br />

the cost of the total truss.<br />

≥ 0 and dφn<br />

non − increasing with Pn, i.e.<br />

dPn<br />

d2φn dP 2 ≤ 0 (16.1)<br />

n<br />

16.2 Proof with supports as further unknowns<br />

The proof follows from equilibrium, change in equilibrium by introducing an additional<br />

bar and the corresponding change in the total cost. We directly deal with threedimensional<br />

trusses, and thus a two-dimensional truss is just a special case.<br />

16.2.1 Force equilibrium<br />

At each joint j three force equilibriums must be satisfied, in all<br />

[Rm]{Pm} = {L} (16.2)<br />

where [Rm] is a matrix of direction cosine’s for the members (bars) in the truss (matrix<br />

order 3J × 2N with totally J joints and totally 2N = J(J − 1) possible bar forces<br />

(N bars)). The vector {L} is a vector of 3J external forces including the reactions.<br />

Let us express the forces {L} in known external forces {B} added a number K of<br />

possible support reactions {Bk} for k =1, 2,...,K (K


Conditions for statically determinate trusses 247<br />

and the total vector (order 2N + K) of unknown member forces {Pm} plus reactions<br />

{Ps} at the possible supports<br />

{P } T := {{Pm} {Ps}} (16.5)<br />

then with a completely known right hand side {B} the total equilibrium is written<br />

[R]{P } = {B} (16.6)<br />

To state it differently, we have treated the possible reactions as possible truss members<br />

connected only to one joint. Note that it is possible to describe any support direction Supports<br />

at a joint and that a joint with totally fixed displacements requires three columns in<br />

the matrix [Rs].<br />

In sub-matrix form we write the force equilibriums Basis<br />

<br />

{PI}<br />

[[RI] [RO]]<br />

{PO}<br />

<br />

= {B} (16.7)<br />

where the squared matrix [RI] of order equal to the rank of the total matrix [[R] {B}]<br />

is non-singular. There may be several possibilities for choosing [RI], but at least one<br />

exists.<br />

We solve system (16.7) in terms of {PI} and get<br />

{PI} = {<br />

Basis<br />

solution<br />

˜ PI}−[Z]{PO} = { ˜ PI}− <br />

{Zo}Po<br />

(16.8)<br />

with the definitions<br />

o<br />

[Z] := [RI] −1 [RO] (16.9)<br />

{ ˜ PI} := [RI] −1 {B} (16.10)<br />

and where {Zo} is column o of the matrix [Z]. The force Po is force o of the set {PO},<br />

and later we use forces Pi that belong to the other set of forces {PI}.<br />

16.2.2 The total cost with all forces<br />

as members<br />

separation<br />

When designed to its minimum allowable value, the cost to of carrying force Pn in a Cost<br />

certain bar is given by a function φn<br />

functions<br />

φn = φn(Pn) (16.11)<br />

satisfying the conditions stated in the theorem. The total cost Φ for the truss corresponding<br />

to all the forces is then<br />

Φ= <br />

φi(Pi)+ <br />

φo(Po) (16.12)<br />

i<br />

o


Derivative<br />

of cost<br />

Cost nonincreasing<br />

248 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

which, with the solution (16.8), we write as<br />

Φ= <br />

16.2.3 The change in total cost<br />

i<br />

φi<br />

<br />

˜Pi − <br />

<br />

(Zi)oPo + <br />

φo(Po) (16.13)<br />

o<br />

Assuming now that all the Po forces are zero, i.e. {PO} = {0}, a total cost ˜ ΦI is<br />

defined<br />

From (16.13) we also calculate dΦ/dPo to<br />

i<br />

o<br />

˜ΦI = <br />

φi( ˜ Pi) (16.14)<br />

i<br />

dΦ<br />

=<br />

dPo<br />

<br />

<br />

dφi(Pi)<br />

(−(Zi)o) +<br />

dPi<br />

dφo(Po)<br />

dPo<br />

16.2.4 Monotonous behaviour and new basis solution<br />

(16.15)<br />

We can prove dΦ/dPo to be a non-increasing function of Po: For the second term<br />

in (16.15) that is directly the assumption of our theorem. In the summation term<br />

of (16.15), then for (−(Zi)o) > 0 we have increasing Pi as seen from (16.8), thus<br />

non-increasing dφi(Pi)/dPi and therefore also (dφi(Pi)/dPi)(−(Zi)o) non-increasing.<br />

Finally, terms with (−(Zi)o) < 0 give decreasing Pi, thusdφi(Pi)/dPi non-decreasing<br />

with Po and thus also for these terms (dφi(Pi)/dPi)(−(Zi)o) non-increasing.<br />

All terms of dΦ/dPo are thus non-increasing with Po, and the change in total<br />

cost is monotone. In order to minimize Φ we should therefore choose Po as large as<br />

possible, remembering that all forces by definition are positive.<br />

The maximum ∆Po is determined by the condition that all the values of Pi should<br />

stay non-negative. This condition is by (16.8) stated<br />

Pi = ˜ Pi − <br />

(Zi)o∆Po ≥ 0 for alli (16.16)<br />

o<br />

which means that a new basis is obtained. The bar i that gives rise to the smallest ∆Po<br />

in (16.16) leave basis and the new basis with Pi = 0 and ∆Po introduced corresponds<br />

to a new statically determinate truss.


Conditions for statically determinate trusses 249<br />

16.3 <strong>Optimal</strong>ity condition<br />

In other words, our optimality condition is that all neighbouring basis solutions give<br />

larger cost ∆Φ > 0 than the present basis solution. In mathematical terms with ∆Po<br />

determined by (16.16), the optimality condition is Finite cost<br />

<br />

increment<br />

∆Φ = φo(∆Po)+ <br />

i<br />

φi<br />

˜Pi − (Zi)o∆Po<br />

− ˜ ΦI > 0 for all o (16.17)<br />

i.e. for all forces not in basis, tensile as well as compressive forces. The special case<br />

of ∆Po = 0 is controlled by the gradient condition from (16.15), i.e.<br />

dΦ/dPo > 0 for ∆Po = 0 (16.18)<br />

16.4 Alternative proofs and solution procedures<br />

In the literature on mathematical programming, as in (Gass 1964) and in (Hadley Convex<br />

1964) we can find geometrical interpretations showing directly that the optimal solution<br />

is at a vertex of the convex feasible space described by the linear constraints<br />

(16.6). When the objective is a concave function and not just a linear function the<br />

solution is still at a vertex, but local optima may exist if the concave functions are too<br />

curved relative to the curvature of the feasible space. In the actually solved problems Concave<br />

of topology design of trusses such local optima have not been seen.<br />

cost<br />

For the pure linear programming problem, the simplex procedure described in<br />

most literature on mathematical programming and in (Press, Teukolsky, Vetterling<br />

and Flannery 1992) can be applied. For the actual problem with a concave objective<br />

function, the modified simplex procedure described in (Pedersen 1994) can be applied.<br />

This modified procedure also illustrates how the doubling of unknowns (both tensile<br />

and compressive forces in each bar) can be avoided and directly build into the solution<br />

procedure.<br />

16.5 A truss member model<br />

The size of a truss member is determined directly by the force P transmitted through<br />

the member, and the size is assumed to be unchanged through the length l of the<br />

member. The member cross-sectional area A, moment of inertia I and, finally, the<br />

mass/volume/cost φ are thus only dependent on the force P in addition to given<br />

parameters.<br />

The dependence on the force, however, is not very simple because practical design<br />

codes, which account for local stability, divide the force domain into four cases:<br />

feasible space


Member<br />

classes<br />

Slender<br />

or short<br />

250 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

• Tensile bars, i.e. P>0<br />

• Zero strings, i.e. P =0 +<br />

• Slender columns, i.e. PL


Conditions for statically determinate trusses 251<br />

16.5.3 Short columns, i.e. P 1 (16.26)<br />

we write the linear cost function<br />

<br />

φ = ηCT (PL − P )+CC |PL| ⇒ dφ/d|P | = ηCT > 0 ⇒<br />

16.5.4 Zero strings, i.e. P =0 +<br />

d 2 φ/d|P | 2 = 0 (16.27)<br />

In trusses with only a single load case, the optimal topology is often be a mechanism.<br />

To deal with this problem, we include tensile bars of vanishing area and cost No cost<br />

assumed<br />

A = 0 (16.28)<br />

φ =0 ⇒ dφ/dP = CT > 0 ⇒ d 2 φ/dP 2 = 0 (16.29)<br />

According to (16.24) the derivative corresponding to 0 − is infinite; thus, by definition,<br />

the zero strings belong to the tensile domain.<br />

16.5.5 Concave cost functions<br />

We note that all the cost functions (16.22 ), (16.24), (16.27), (16.29) satisfy<br />

φ(0)=0and dφ/dP > 0 and d 2 φ/dP 2 ≤ 0 (16.30)<br />

and the only remaining condition for concave cost functions is then related to the<br />

transition from a slender column to a short column. From (16.24) and (16.27) we get<br />

the condition of non-increasing derivative for the transition to be Concave<br />

CC<br />

2 |PL| >ηCT<br />

(16.31)<br />

transition


Assumption<br />

for cost of<br />

supports<br />

Collected<br />

results<br />

252 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Let us from (16.23 ,16.24) assume CC = l/(πα E/ξ) which with (16.19) gives<br />

CC<br />

2 √ |PL| = |PL|ξ/σL. Then, the left hand side of (16.31) is ξ/(2σL). With η := σT /σ0<br />

and from (16.21, 16.22) CT =1/σT we can finally write the condition (16.31)<br />

ξ ≥ 2σL/σ0<br />

(16.32)<br />

When the factor of safety has the value 2σL/σ0, the slopes have a continuous transition.<br />

Totally, with (16.31) or specifically (16.32), we conclude that the cost functions<br />

are all concave functions.<br />

φ = φ(|P |) (16.33)<br />

16.5.6 Cost functions for the supports<br />

We assume that the support costs are also described by concave functions, with possibility<br />

of a different function when the support reaction is either positive or negative.<br />

This assumption is as an example fulfilled by simple linear functions.<br />

Totally, we conclude that the cost functions for bar members as well as for supports<br />

are all concave functions.<br />

16.6 Summing up<br />

In this chapter the important results to focus on are:<br />

φ = φ(|P |) (16.34)<br />

• The single load case assumption behind the results in this chapter is most<br />

essential.<br />

• Local stability of each bar in the truss is accounted for, treating both slender<br />

columns and short columns.<br />

• Support conditions need not be specifically given and need not be statically<br />

determinate. Support reactions are found as part of the optimization.<br />

• Specific optimal truss examples are presented in chapter 3, and the necessary<br />

numerical procedure for the optimization is described in chapter 19.


Chapter 17<br />

Orientation of orthotropic<br />

material<br />

A problem of major practical importance is related to the optimization procedure for<br />

locating optimal orientation, i.e. how to find an optimal design, which satisfy the Many local<br />

extremum<br />

optimality criterion, stated in chapter 14 by (14.2). Procedures like steepest decent<br />

are not reliable due to the inherent problem of many local extremum.<br />

Recursive procedures are used extensively, but also here several possibilities exist.<br />

We may redefine according to actual larger principal strain direction, according to<br />

actual larger principal stress direction or according to some combination of these<br />

directions. The relative effectiveness of these procedures is problem dependent, i.e. Recursive<br />

dependent on the degree of non-isotropy and not at least on the degree of statically<br />

indeterminacy of the problem.<br />

In the present analysis a stress characteristic of the optimal solution is proven.<br />

This states a condition for the principal stress ratio σ2/σ1 (where |σ1| ≥|σ2|), which<br />

must be satisfied for the design with optimal orientation. To a large extent the<br />

presentation in this chapter follows (Pedersen and Bendsøe 1995).<br />

17.1 Inherent practical problems<br />

In order to use a non-isotropic material effectively it should be oriented optimally<br />

with respect to the actual stress strain condition. We are not here concerned with the<br />

design of material as in chapter 8, but merely with the use of a given non-isotropic<br />

material. Complete analytical results are obtained for orthotropic materials in two<br />

procedures<br />

dimensions, showing that point-wise local as well as global maxima and minima Gradient<br />

result. Thus a pure gradient information is not sufficient and we need to obtain a<br />

deeper understanding.<br />

253<br />

information<br />

non-sufficient


Alignments<br />

The match<br />

problem<br />

254 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

In optimal orientation for maximum total stiffness (minimum compliance) it is<br />

shown that for orthotropic materials we get alignment between principal strain directions,<br />

principal stress directions and principal material directions, at least in most<br />

cases, see (Seregin and Troitskii 1982), (Banichuk 1983), (Fedorov and Cherkaev<br />

1983), (Pedersen 1989), (Rovati and Taliercio 1991), (Pedersen and Taylor 1993).<br />

Materials with relative high shear stiffness (to be defined later) are not aligned<br />

with the principal strains and principal stresses, but also for these materials are the<br />

principal strain directions and the principal stress directions aligned when the criterion<br />

for optimal orientation is satisfied, (Pedersen 1990).<br />

However, alignment is not satisfactory information, because even for 2D-problems<br />

we have two principal strains, two principal stresses and two orthotropic material<br />

directions. Thus a match problem is identified. If the criterion for optimal orientation<br />

is satisfied, then the numerically larger principal strain and the numerically larger<br />

principal stress together with the numerically larger material modulus are all aligned.<br />

For all relative low shear stiffness materials we obtain this result, but not in general<br />

for all materials.<br />

17.2 From global to local optimality criterion<br />

The actual optimization problem is stated as orientational design for minimum compliance<br />

(equal to the negative external potential Uext), for minimum elastic strain energy<br />

Uɛ, for maximum stiffness (-Uɛ), or for maximum potential energy Π := Uɛ + Uext.<br />

Different<br />

These different statements are all the same with a constant ratio between elastic<br />

objectives stress energy Uσ and elastic strain energy Uɛ. For power law elasticity (including<br />

linear elasticity) this ratio is p (see chapter 12) with 0


Orientation of orthotropic material 255<br />

independent of the power p of the non-linear elasticity model. In a point-wise (continuum)<br />

formulation the mean strain energy density (ūɛ)e is substituted by the strain<br />

energy density uɛ itself, and the angle θe of element e with the angle θ at the actual<br />

point. In the following, for simplicity, we use this notation without index e and<br />

without the mean value bar super-index¯.<br />

17.3 Multiplicity of extremum<br />

We restrict the discussion here and in the following to the case of linear, two-dimensional<br />

elasticity. The conditions for optimal orientation for an orthotropic material are presented<br />

here, following the developments in (Pedersen 1989), and (Cheng and Pedersen<br />

1997). First consider the maximization or minimization of the point-wise strain energy<br />

density for an orthotropic material. Using the axes of principal strains (ɛ1,ɛ2)<br />

as the frame of reference, the strain energy density can be written as The θ function<br />

for strain<br />

uɛ = C2(ɛ energy density<br />

2 1 − ɛ 2 2) cos 2θ + C3(ɛ1 − ɛ2) 2 cos 4θ (17.3)<br />

with the definitions (used in the theory of laminates, see also chapter 6)<br />

2C2 := C1111 − C2222, 8C3 := C1111 + C2222 − 2C1122 − 4C1212<br />

(17.4)<br />

where θ is the angle between the principal strain frame and the axes of orthotropy,<br />

and where ⎡<br />

⎤<br />

Non-isotropic<br />

parameters<br />

[C] = ⎣<br />

⎦ (17.5)<br />

C1111 C1122 0<br />

C1122 C2222 0<br />

0 0 2C1212<br />

is the constitutive matrix in the frame of orthotropy. Ordering so that |ɛ1| ≥|ɛ2| and<br />

C1111 ≥ C2222, we see that all coefficients to the trigonometric functions in (17.3), Solution<br />

except C3, are positive. The condition of stationarity (17.2) becomes<br />

equation<br />

sin 2θ(C2(ɛ 2 1 − ɛ 2 2)+4C3(ɛ1 − ɛ2) 2 cos 2θ) = 0 (17.6)<br />

and from second order conditions or direct evaluation at the stationary points, we<br />

obtain, with the definition of the strain related parameter γ for C3(ɛ1 − ɛ2) = 0,<br />

γ := C2 ɛ1 + ɛ2<br />

4C3 ɛ1 − ɛ2<br />

(17.7)


Results in<br />

strains<br />

Relative<br />

low or high<br />

shear stiffness<br />

The θ function<br />

for stress<br />

energy density<br />

Results in<br />

stresses<br />

256 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

C3 > 0: uɛ is max for θ = 0<br />

uɛ is min for θ = ±(1/2) arccos (−γ) if |γ| ≤1<br />

uɛ is min for θ = ±(π/2) if |γ| > 1 (17.8)<br />

C3 < 0:uɛ is max for θ = ±(1/2) arccos (−γ) if |γ| < 1<br />

uɛ is max for θ = 0 if |γ| > 1<br />

uɛ is min for θ = ±(π/2) (17.9)<br />

and in all cases (seen by inspection) are the axes of principal strains and principal<br />

stresses (at the optimum) aligned. In line with (Pedersen 1989), we say that a material<br />

with C3 > 0 has relative low shear stiffness, and relative high shear stiffness if C3 <<br />

0. The match between larger principal stress directions and larger principal strain<br />

directions is discussed in section 17.4.<br />

We note here that the results above can be used directly to obtain similar results<br />

for the dual problem of minimization or maximization of the stress energy density uσ<br />

with the definitions<br />

uσ = H2(σ 2 1 − σ 2 2) cos 2θ + H3(σ1 − σ2) 2 cos 4θ (17.10)<br />

2H2 := H1111 − H2222, 8H3 := H1111 + H2222 − 2H1122 − 4H1212<br />

Here [H] =[C] −1 in the frame of orthotropy is given as<br />

[H] =<br />

1<br />

C1111C2222 − C 2 1122<br />

⎡<br />

⎣<br />

C2222 −C1122 0<br />

− C1122 C1111 0<br />

0 0<br />

C1111C2222−C 2<br />

1122<br />

2C1212<br />

(17.11)<br />

⎤<br />

⎦ (17.12)<br />

We get in analogy to (17.8, 17.9), with the definition of the stress related parameter<br />

ζ for H3(σ1 − σ2) = 0,<br />

ζ := H2 σ1 + σ2<br />

(17.13)<br />

4H3 σ1 − σ2<br />

H3 > 0: uσ is max for θ = 0<br />

uσ is min for θ = ±(1/2) arccos (−ζ) if |ζ| ≤1<br />

uσ is min for θ = ±(π/2) if |ζ| ≥1 (17.14)


Orientation of orthotropic material 257<br />

H3 < 0:uσ is max for θ = ±(1/2) arccos (−ζ) if |ζ| < 1<br />

uσ is max for θ = 0 if |ζ| > 1<br />

uσ is min for θ = ±(π/2) (17.15)<br />

It can be shown, see (Cheng and Pedersen 1997), that C3 < 0 leads to H3 > 0, but if<br />

C3 > 0 the sign of H3 can be positive as well as negative. Maximization of stiffness<br />

corresponds to maximization of uɛ in a fixed strain field or minimization of uσ in a<br />

fixed stress field (or vice versa for minimization of stiffness).<br />

Taking for example a case with C3 < 0 and H3 > 0, we see from (17.9), (17.14) Example<br />

that the strain and stress fields at the optimum are related by the condition γ = ζ.<br />

Also, note that if a given (non-optimal) set of stress and strain fields do not satisfy<br />

this relation, the optimum rotation angles read off from (17.8), (17.9), (17.14), (17.15)<br />

may be quite different. The complete analysis for the various combinations of signs<br />

of C3 and H3 can be found in (Cheng and Pedersen 1997). In the following we<br />

concentrate on finding characteristic bounds for optimal stress fields, and consider<br />

the effects these bounds have on computations.<br />

17.4 The match problem<br />

Even for 2D-problems we have two principal strain directions, two principal stress Principal<br />

directions and two orthotropic material directions. Thus knowledge on alignment of<br />

principal directions is not enough information, and in this section this problem is<br />

analysed in more detail to end up with a specific strategy.<br />

Firstly, we consider principal strains ɛ1,ɛ2 =1,η aligned with principal stresses<br />

σa,σb and related by the constitutive equation<br />

<br />

σa C1111 C1122 1<br />

=<br />

(17.16)<br />

σb C1122 C2222 η<br />

with constitutive parameters C1111,C2222,C1122. For well-ordered strains, i.e. ɛ2 = Relative<br />

directions<br />

ηɛ1 with −1 ≤ η ≤ 1, we now derive the size-ordering of the stresses, given in terms strains<br />

of the constitutive parameters and the relative strain value η. First, note that the<br />

condition for a positive definite constitutive matrix implies that Positive<br />

definite<br />

C1111C2222 − C 2 1122 > 0, C1111 > 0, C2222 > 0, (17.17)<br />

With the stresses given by (17.16), we can obtain the necessary<br />

investigation of the sign of a second order polynomial in η<br />

information by an 2nd order<br />

polynomial<br />

σ 2 a − σ 2 b =(C 2 1122 − C 2 2222)η 2 +2C1122(C1111 − C2222)η +(C 2 1111 − C 2 1122) (17.18)


Stress<br />

ordering<br />

from strains<br />

258 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

over the interval −1 ≤ η ≤ 1. This function defines a parabola parameterized in the<br />

variable η which at the interval end-points takes the values<br />

Moreover, the two roots σ 2 a − σ 2 b<br />

(C1111 − C2222)(C1111 + C2222 ± 2C1122) (17.19)<br />

= 0 are found to be at<br />

η± = ∓ C1111 ± C2222<br />

C2222 ± C1122<br />

which lie in the interval −1 ≤ η ≤ 1if(C2222 ± C1122) 2 > (C1111 ± C1122) 2 , i.e. if<br />

(17.20)<br />

−(C1111 − C2222)(C1111 + C2222 ± 2C1122) ≥ 0 (17.21)<br />

The factor (C1111 + C2222 ± 2C1122) common to (17.19) and (17.21) is positive as we<br />

have C1111C2222 − C2 1122 > 0. Thus if (C1111 − C2222) > 0, the end values by (17.19)<br />

are positive and no roots fall in the interval −1 ≤ η ≤ 1 indicating that σ2 a − σ2 b > 0<br />

for all −1 ≤ η ≤ 1. On the other hand, if (C1111 − C2222) < 0 the end values by<br />

(17.19) are negative and the roots fall in the interval −1 ≤ η ≤ 1. Then σ2 a − σ2 b > 0<br />

holds between the roots (17.20). In conclusion we have :<br />

C1111 >C2222 : |σa| > |σb| for any − 1 ≤ η ≤ 1<br />

C1111 |σb| for − C1111 + C1122<br />

C2222 + C1122<br />


Orientation of orthotropic material 259<br />

The analysis (17.16) to (17.22) may be started from the assumption of stress<br />

ordering |σ1| ≥|σ2| ending up with conditions like (17.22) for |ɛa| ≥|ɛb|. We here<br />

only list this specific result that is needed for the further discussion<br />

H1111 |ɛb| for − H1111 + H1122<br />


Strain or<br />

stress<br />

directions<br />

Collected<br />

results<br />

260 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

optimization is given in (Pedersen 1991), and here we comment on the two alternatives<br />

for orientational optimization, i.e. redefinition according to actual larger principal<br />

strain direction or according to larger principal stress direction.<br />

The experience from many solved problems shows that redefinition according<br />

to larger principal stress gives the most stable optimization procedure. This can be<br />

understood from the fact that the stress distribution is not strongly effected by the<br />

design changes. In contrast the redefinition according to larger principal strain often<br />

gives fluctuating results, and when this is the case we see that the characteristic<br />

(17.24) is not satisfied. Thus, when (17.24) is violated the principal stress and strain<br />

direction may be close to alignment, but the larger principal stress does not match<br />

the larger principal strain, which then is very sensitive to design changes.<br />

Should we redefine material orientation according to the principal strain field or<br />

according to the principal stress field ? Normally, principal stresses is to be preferred<br />

but it depends on the specific problem. For an almost statically determinate problem<br />

the stress field changes very little, and if it does not change at all then only one<br />

iteration is necessary. On the other hand in maximizing flexibility we may have<br />

rather unchanged strain field and then a different strategy is needed.<br />

17.6 Summing up<br />

In this chapter the important results to focus on are:<br />

• The elastic energy as a function of material orientation has multiple stationary<br />

solutions, in most cases corresponding to alignment of principle stress, principle<br />

strain and orthotropic material directions.<br />

• The case of high relative shear stiffness material, should be treated with special<br />

care. See (Cheng and Pedersen 1997) and (Pedersen 1989) for more detail.<br />

• The optimal match of the specific principle directions for stress, strain and<br />

material is to a large extent known for two-dimensional problems, and recursive<br />

iterations work in most cases. The three-dimensional problems still need some<br />

clarification.<br />

• Especially for strongly non-isotropic materials and strongly statically indeterminate<br />

cases, severe restrictions are put on the optimal stress state, as stated<br />

by (17.24). Thus the existence of solutions that satisfy the optimality criterion<br />

is not always guaranteed.<br />

• Gradient based methods are not reliable, due to the mentioned multiplicity of<br />

stationary solutions.


Chapter 18<br />

Sensitivity analysis for<br />

dynamic problems<br />

18.1 Including non-conservative problems<br />

The dynamic behaviour of systems subjected to non-conservative loads is often against<br />

our physical intuition and can only be understood with a deep insight. As exam- Non-intuitive<br />

ples: vibration frequency may increase with increasing load and damping may act behaviour<br />

destabilizing. The non-conservative forces make it necessary to deal with complex<br />

functionals in the description, and the intuitive understanding of the physical results<br />

is not easy because the indirect effects may be the determining ones. However, in this<br />

chapter high priority is given to the physical interpretation.<br />

The usual (familiar) analysis for the displacements etc., could be based either<br />

on the equations of equilibrium, set up directly, or on the stationarity of an energy<br />

functional. For the non-selfadjoint problems such functionals are now also available in<br />

the literature and we term them here the mutual energy functionals - mutual in the Mutual<br />

sense that displacements of the physical as well as of the adjoint problem are involved. potentials<br />

The sensitivity analysis may also be based directly on the equations of equilibrium,<br />

but the simplicity is striking when we base it on the mutual energy, because then the<br />

variations of displacements do not have to be considered.<br />

The response analysis and sensitivity analysis can be formulated in either a continuous<br />

or in a discrete form. As almost all problems are solved numerically with Finite<br />

only finite degrees of freedom, the sensitivity analysis is presented in a matrix for- degrees of<br />

mulation, which has the advantage that boundary conditions are an integrated part. freedom<br />

Furthermore, this formulation relates directly to the practical methods of analysis,<br />

such as the finite element method and the Galerkin method.<br />

261


Linear<br />

systems<br />

Stationary<br />

systems<br />

Normalized<br />

mass matrix<br />

Conservative<br />

systems<br />

262 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

18.2 Classification of dynamic systems and their behaviour<br />

Restricting ourselves to linear, elastic system, we can in matrix notation write the<br />

dynamic equilibrium by<br />

[M]{ ¨ D} +[C]{ ˙ D} +[S]{D} = {A} (18.1)<br />

where [M] is termed the mass matrix, being the coefficient matrix to the vector<br />

{ ¨ D} of accelerations. Thus derivative with respect to time t is denoted by the ˙ (dot)<br />

superscript. The matrix [C] is termed the damping matrix, being the coefficient<br />

matrix to the vector { ˙ D} of velocities. The stiffness matrix [S] is the coefficient<br />

matrix to the displacements {D}, and the external forces {A} then make the system<br />

inhomogeneous.<br />

18.2.1 System classification<br />

The system classification follows the classic book by (Ziegler 1968), and we immediately<br />

identify the class termed instationary systems. If only one element in one of the<br />

matrices [M], [C], [S], {A} is explicitly a function of time t, then the system is termed<br />

instationary. Sensitivity analysis for these systems is described in (Pedersen 2001b).<br />

With stationary external forces {A} , we can concentrate on the homogeneous system<br />

that describes the dynamics relative to the static equilibrium [S]{D} = {A}. Thus<br />

we continue our classification based on system description by<br />

[M]{ ¨ D} +[C]{ ˙ D} +[S]{D} = {0} (18.2)<br />

and furthermore we assume this to be normalized in such a way that [M] is symmetric<br />

and positive definite. Now, the remaining classification relates directly to the contents<br />

of the matrices [C] and [S]. Writing these matrices as a sum of their symmetric and<br />

skew-symmetric parts<br />

[C] = 1<br />

2 ([C]+[C]T )+ 1<br />

2 ([C] − [C]T ), [S] = 1<br />

2 ([S]+[S]T )+ 1<br />

2 ([S] − [S]T ) (18.3)<br />

we can directly define our remaining four classes of systems.<br />

A conservative, non-gyroscopic system has no velocity depending terms and a<br />

symmetric stiffness matrix, which in physics may describe simple, free, undamped<br />

vibrations. The characteristics of the possible solutions are well known, and these<br />

solutions often serve as expansions for more complicated systems. A conservative,<br />

gyroscopic system may in addition include gyroscopic forces which are reflected in<br />

the skew-symmetric part of the damping matrix which in physics often is a possible


Sensitivity analysis for dynamic problems 263<br />

description for a rotor system. The non-conservative systems are systems with dissipative<br />

forces and/or with circulatory forces. The dissipative forces are described by<br />

the symmetric part of the damping matrix, and the circulatory forces are described<br />

by the skew-symmetric part of the stiffness matrix. Thus for a dissipative system Non-conservative<br />

we have ([C] +[C] T ) = [0] without further condition on our general system. For a<br />

circulatory system we have ([S] − [S] T ) = [0] which means that non-potential forces<br />

are included in the system.<br />

18.2.2 Classification of the behaviour<br />

We are concerned about the stability of the solution {D} = {0} of (18.2) to small<br />

disturbances. Now to classify as far as possible the behaviour of the above classes of<br />

systems, we assume the time separation by<br />

{D} = {Φ} exp ((α + ıω)t) ={Φ} exp (αt)(cos ωt + ı sin ωt) (18.4)<br />

and thus in this separation from the initial formulation uses a complex exponential<br />

factor λ = α + ıω as time coefficient. With this formulation we can avoid the physical<br />

systems<br />

nonsense of complex frequency. Two distinct classes of instability follow from this Flutter<br />

separation. The first one is flutter instability (also termed dynamic instability), it<br />

corresponds to solutions α>0,ω = 0. The second one is divergent instability (also<br />

- dynamic<br />

termed static instability), it corresponds to solutions α>0,ω = 0. Stability is Divergence<br />

defined by solutions having only negative α (real part), and a critical solution is the<br />

terminology for the case of a solution with at least one α = 0 and negative α values<br />

for the remaining solutions. For ω = 0 this corresponds to a harmonic vibration part<br />

of the solution. To be more specific about stability limits we should include gradient<br />

information. With p as a load parameter we define a flutter load pF by the conditions<br />

that at least for one solution we have<br />

ωF = 0, αF =0, (∂α/∂p)F > 0 (18.5)<br />

and non-positive values for the remaining α solutions. In parallel a divergence load<br />

pD is obtained for<br />

ωD =0, αD =0, (∂α/∂p)D > 0, (∂ω/∂p)D = 0 (18.6)<br />

Now, one of the most important theorems in stability tells that a conservative, nongyroscopic<br />

system can only lose its stability by a divergent instability. This means<br />

that the definiteness of the stiffness matrix gives complete information about stability,<br />

but naturally not about the dynamic behaviour in general. Thus, such a system is<br />

stable if and only if [S] is positive definite.<br />

- quasi-static


Different<br />

sensitivities<br />

System<br />

matrix<br />

Adjoint<br />

problem<br />

264 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

18.3 Sensitivity analysis<br />

The notion of sensitivity analysis is a very wide one, even when we restrict it to<br />

response sensitivity for dynamic systems. It has been argued that sensitivity is just a<br />

new word for partial derivative, and to some extent this is correct. When our response<br />

is measured by a scalar like an eigenvalue and we ask for the sensitivity with respect<br />

to another scalar then it is a partial derivative. However, we may also ask for the<br />

sensitivity with respect to a function, and then the notion of gradient function is often<br />

used. Furthermore, the response may be a function like an eigenfunction or another<br />

displacement description, and then we need a notion for the change of a function with<br />

respect to another function. So sensitivity analysis is a notion that covers the work<br />

of determining all this important information. The description in this chapter follows<br />

closely to (Pedersen 1984).<br />

18.3.1 General variational analysis<br />

We are interested in studying the dynamic behaviour, and with the separation (18.4)<br />

the spatial problem is described by the homogeneous matrix equation<br />

[L]{Φ} = {0} with [L] =[L(α + ıω,p,h)] (18.7)<br />

where the system matrix [L] for the present analysis depends on the complex eigenvalue<br />

λ with α as a stability measure and ω as frequency according to the separation<br />

by {D} = {Φ} exp (α + ıωt). Furthermore [L] depends on the load level described<br />

by the real parameter p and on design, load distribution, damping etc., all of which<br />

we symbolize by the real quantity h, which - in individual cases - may be a scalar<br />

parameter or a spatial parameter function. In addition to the physical system we also<br />

analyze the adjoint problem<br />

[L] T {Ψ} = {0} or {Ψ} T [L] ={0} T<br />

(18.8)<br />

i.e. we need both right {Φ} and left {Ψ} eigenvectors, for the same eigenvalue. We<br />

also remember that with a symmetric system matrix (self-adjoint problem) we have<br />

{Φ} = {Ψ}. The results we obtain in this section are expressed as ratios between<br />

complex functionals, and in order to make it less abstract, an interpretation in terms<br />

of specific, mutual energies is presented. The term specific means that the energy may<br />

be per area or per volume and a (for the sensitivity analysis) non-important factor<br />

may be omitted. The term mutual relates to the fact that two different eigenvectors<br />

determine the energy. The functional, termed the total, specific, mutual energy Π is<br />

defined by<br />

Π:={Ψ} T [L]{Φ} = 0 (18.9)


Sensitivity analysis for dynamic problems 265<br />

where the zero follows from (18.7) as well as from (18.8). Taking general variations<br />

of Π, which is also zero with the variations, we get from δΠ =0 Total mutual<br />

potential<br />

{δΨ} T [L]{Φ} + {Ψ} T [δL]{Φ} + {Ψ} T [L]{δΦ} = 0 (18.10)<br />

We find that the variations of the eigenvectors {Φ}, {Ψ} disappear, which is naturally<br />

the reason for introducing the adjoint eigenvector and for taking variations of the<br />

mutual energy, δΠ = 0, instead of variations relative to the equilibrium (18.7).<br />

In relation to the sensitivity analysis where [δL] = [0] we write (18.10) more<br />

specifically in the variations of the involved independent parameters δp and δh. To<br />

have a shorter notation we define the important functionals (normally complex) Complex<br />

functionals<br />

A = {Ψ} T (∂[L]/∂λ){Φ}, B = {Ψ} T (∂[L]/∂p){Φ}, C = {Ψ} T (∂[L]/∂h){Φ} (18.11)<br />

and then get<br />

A(δα + ıδω)+Bδp + Cδh = 0 (18.12)<br />

Direct interpretation of this equation (18.12) gives the results needed.<br />

Sensitivity with respect to load level. With a fixed ”design” δh = 0 the<br />

sensitivity from (18.12) makes possible a more rigorous definition of terms like critical<br />

load and flutter load. Assuming A = 0 we read from (18.12) with δh =0<br />

∂α/∂p = −Re(B/A), ∂ω/∂p = −Im(B/A) (18.13)<br />

where the notation for real part is Re() and the notation for imaginary part is Im().<br />

To clarify the term instability initiation, we must in addition to α = 0 require ∂α/∂p ><br />

0. By (18.13) this instability condition is Well defined<br />

instability<br />

α =0 and Re(B/A) < 0 (18.14) criterion<br />

In the case of ω = 0 condition (18.14) corresponds to initiation of flutter and in the<br />

case of ω = 0 together with Im(B/A) = 0 condition (18.14) corresponds to initiation<br />

of divergence.<br />

Sensitivity with respect to design. With a fixed load level δp = 0 we read<br />

from (18.12), again assuming A = 0<br />

∂α/∂h = −Re(C/A), ∂ω/∂h = −Im(C/A) (18.15)<br />

We are often interested in knowing whether a certain parameter h acts in a stabilizing Stabilizing or<br />

or destabilizing manner, which, from (18.15), we find to<br />

destabilizing<br />

Re(C/A) > 0 ⇒ h stabilizing, Re(C/A) < 0 ⇒ h destabilizing (18.16)


Flutter load<br />

sensitivity<br />

Flutter<br />

frequency<br />

sensitivity<br />

Kinetic,<br />

dissipative,<br />

elastic and<br />

external<br />

potentials<br />

Sensitivity<br />

functionals<br />

266 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Flutter load sensitivity to design. A main question is related to the change<br />

in instability level as a function of design, i.e. all variations in (18.12) are involved.<br />

Since we focus on a load level of instability initialization, the variations are restricted<br />

by<br />

α =0 and δα = 0 (18.17)<br />

Again for A = 0 we from (18.12) divided by A with (18.17) and using that ıδω is<br />

purely imaginary, get<br />

∂pF /∂h = −Re(C/A)/Re(B/A) (18.18)<br />

Note that by (18.18) the sign of ∂pF /∂h is equal to the sign of Re(C/A). This simply<br />

states the natural fact from (18.12) that if h is stabilizing, then pF increases with h.<br />

If ω = δω = 0 the load pF , here termed flutter load, is as a special case a divergence<br />

load. No specific analysis for this is therefore needed. The change in flutter frequency<br />

is in a similar way obtained with δα = 0 and we get<br />

∂ωF /∂h = −Im(C/B)/Re(A/B) (18.19)<br />

because δpF<br />

from (18.14).<br />

is a real quantity. In (18.19) we have used that B = 0 which follows<br />

18.3.2 Results with finite element formulation<br />

We now restrict the description to system matrices [L] given by<br />

[L] =λ 2 [M]+λ[C]+[S]+p[K] (18.20)<br />

with λ as the only complex quantity. In order to get convenient short notation we<br />

define specific, mutual energies: kinetic T , dissipative D, elastic U and external W<br />

by<br />

T := {Ψ} T [M]{Φ}, D := {Ψ} T [C]{Φ},<br />

U := {Ψ} T [S]{Φ}, W := −{Ψ} T [K]{Φ} (18.21)<br />

and with (18.20) we may express W by<br />

W =(λ 2 T + λD + U)/p (18.22)<br />

The important functionals A and B<br />

energies<br />

can then directly be expressed in the mutual<br />

A =2λT + D, B = −W = −(λ 2 T + λD + U)/p (18.23)<br />

and thus the results for pure load change (18.13) are directly available<br />

∂λ/∂p = ∂α/∂p + ı∂ω/∂p =(λ 2 T + λD + U)/(p(2λT + D)) (18.24)


Sensitivity analysis for dynamic problems 267<br />

To obtain the functional C we need more specific information, and even for the case<br />

of [K] independent of h , the evaluation of<br />

C = {Ψ} T (λ 2 ∂[M]/∂h + λ∂[C]/∂h + ∂[S]/∂h){Φ} (18.25)<br />

seems rather complicated. For a finite element model we have<br />

[M] = [Me], [C] = [Ce], [S] = [Se] (18.26)<br />

This has to be read symbolically because [M], [C] and [S] are matrices of order say<br />

10 3 while [Me], [Ce] and [Se] are element matrices of order 10. Often a parameter he<br />

only influences a specific element and then (18.25) is brought down to<br />

C = {Ψe} T (λ 2 ∂[Me]/∂he + λ∂[Ce]/∂he + ∂[Se]/∂he){Φe} (18.27)<br />

where the components of the lower order vectors {Ψe}, {Φe} are contained in the<br />

higher order vectors {Ψ}, {Φ}. Thus the sensitivity analysis is brought down to the<br />

element level. Furthermore, the he dependence is often homogeneous, such that Element<br />

level FEM<br />

∂[Me]/∂he = m[Me]/he, ∂[Ce]/∂he = l[Ce]/he, ∂[Se]/∂he = k[Se]/he (18.28)<br />

where m, l, k are given factors. Then (18.27) can be written directly in element mutual<br />

energies defined in parallel to (18.21)<br />

C =(mλ 2 Te + lλDe + kUe)/he<br />

(18.29)<br />

With the assumption behind (18.29), localized sensitivities are available , and the<br />

general results (18.15), (18.18), and (18.19) simplify to<br />

and<br />

∂ωF<br />

∂pF<br />

∂he<br />

= Re((mλ2 Te + lλDe + kUe)/(2λT + D))<br />

heRe((λ 2 T + λD + U)/(2λT + D))<br />

= −Im((mλ2 Te + lλDe + kUe)/(λ 2 T + λD + U))<br />

heRe((2λT + D)/(λ 2 T + λD + U))<br />

∂he<br />

These formulas are directly suited for evaluation or programming.<br />

(18.30)<br />

(18.31)<br />

18.3.3 Results with Galerkin modelling (global expansion)<br />

An alternative to the finite element method, which we can interpret as a Galerkin Expansion<br />

method with local expansion functions, we have methods with global expansion func- functions<br />

tion often named Galerkin, Ritz and weighted residual. The advantage of these analysis<br />

methods is the moderate number of degrees of freedom, and the disadvantage is


Galerkin<br />

discretization<br />

Gradient<br />

function<br />

268 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

that expansion function valid over the whole structural domain V must be obtainable.<br />

With x as the space parameter in V , we assume given expansion functions uj = uj(x)<br />

for the physical eigenfunction and vj = vj(x) for the adjoint eigenfunction. Let us<br />

rewrite (18.27) without matrix notation<br />

C = (λ 2 ∂mij/∂h + λ∂cij/∂h + ∂sij/∂h)ΨiΦj<br />

(18.32)<br />

where mij,cij,sij are matrix elements of [M], [C], [S], respectively. These elements<br />

are for the global Galerkin discretization determined by<br />

<br />

<br />

<br />

mij = mH(ui,vj)dx, cij = cG(ui,vj)dx, sij = sF (ui,vj)dx, (18.33)<br />

with F, G, H as given expressions in the functions u, v and their spatial derivatives.<br />

The factors m = m(x),c = c(x),s = s(x), depend on the design function h = h(x).<br />

By the partial derivative with respect to h(x) we mean change of h at position x in<br />

space. Assuming only local influence we get<br />

∂mij<br />

∂h(x)<br />

∂m(x)<br />

∂cij<br />

= H(ui,vj),<br />

∂h(x) ∂h(x)<br />

Inserting this in (18.32) we have<br />

∂c(x)<br />

∂sij<br />

= G(ui,vj),<br />

∂h(x) ∂h(x)<br />

∂s(x)<br />

= F (ui,vj)<br />

∂h(x)<br />

(18.34)<br />

C(x) = 2 ∂m<br />

∂c<br />

∂s<br />

(λ H(ui,vj)+λ G(ui,vj)+<br />

∂h(x) ∂h(x) ∂h(x) F (ui,vj))ΨiΦj (18.35)<br />

The sensitivities are then a function of space and we term them gradient functions.<br />

For example in relation to the most important sensitivity (18.18) we have<br />

g(x) :=∂pF /∂h(x) =−Re(C(x)/A)/Re(B/A) (18.36)<br />

and the resulting flutter load variation is obtained by integration<br />

<br />

δpF = g(x)δh(x)dx (18.37)<br />

The following example has results evaluated by these formulas.<br />

18.3.4 Example of a cantilever beam<br />

In figure 18.1 (left) we have shown a cantilever beam of non-uniform mass m(x), bending<br />

stiffness s(x), Kelvin-Voigt internal damping with coefficient γ, external damping<br />

with coefficient β, and loaded according to the distribution pq(x) as follower forces.


Sensitivity analysis for dynamic problems 269<br />

Figure 18.1: Left: extended Leipholz and Hauger columns including external and<br />

internal damping. Right: characteristic curves for the Becks column for a uniform<br />

well as for a linearly tapered column.<br />

With eigenvalue λ and eigenfunction u(x) , the continuous eigenvalue problem is<br />

((su ′′ ) ′′ + pQu ′′ )+λ(βu + γ(su ′′ ) ′′ )+λ 2 <br />

mu ≡ 0, Q := pq(ξ)dξ, (18.38)<br />

with the boundary conditions Continuous<br />

problem<br />

u(0) = u ′ (0)=0, (1 + λγ)(su ′′ )x=1 =(1+λγ)(su ′′ ) ′ x=1 = 0 (18.39)<br />

Special cases are known as Leipholz’s column and Hauger’s column, and with an<br />

end load Beck’s column. Discretizing this problem by Galerkin expansions as given<br />

in details in Pedersen and Seyranian (1983) we obtain the analysis results shown in


270 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

figure 18.1 (right) for the Beck column and with very similar results for the Leipholz<br />

and Hauger columns. The sensitivity analysis follows the theory described and the<br />

specific functional for this problem is<br />

C(x) = ((1 + λγ)2mv ′′<br />

i u ′′<br />

j + λ 2 viuj)ΨiΦj<br />

(18.40)<br />

and figure 18.2 shows results from sensitivity analysis. Note, that at positions (horizontal<br />

axes) where the curves are below the zero level we have negative flutter load<br />

gradient, and at these beam positions it is advantageous to decrease the beam area.<br />

18.3.5 Multiple eigenvalues<br />

The case of multiple eigenvalues is often excluded in sensitivity formulations, but in<br />

reality it is a physical important example that must be properly dealt with. The most<br />

simple case of a conservative, non-gyroscopic system for which<br />

[L] =−ω 2 [M]+[S] (18.41)<br />

With [M] and [S] symmetric we always have a complete set of eigenvectors, even<br />

when some of them have the same eigenvalue. All eigenvalues and eigenvectors are<br />

real with left and right eigenvectors being identical (self-adjoint). A straight forward<br />

sensitivity analysis as derived from (18.25) gives for the specific solution (ω2 Distinct<br />

, {Φ})<br />

eigenvalues ∂ω2 ∂h = {Φ}T 2 ∂[M] ∂[S]<br />

(−ω + ){Φ} (18.42)<br />

∂h ∂h<br />

when the normalization<br />

{Φ} T [M]{Φ} = T = 1 (18.43)<br />

Eigenmodes<br />

identification<br />

is assumed. This result can be found in (Wittrick 1962) and in many later references.<br />

The problem is that when the eigenvector {Φ} is not unique, the sensitivity is only<br />

defined as a directional derivative (h positive scalar coefficient to a specified design<br />

direction), and furthermore {Φ} can be any linear combination of the set of eigenvectors<br />

all with the same eigenvalue ω 2 . Only for some specific { ˆ Φ} the sensitivity<br />

(18.42) has a physical meaning. To find the specific { ˆ Φ} vectors let us deal only with<br />

the bimodal case, where with {Φ1}, {Φ2} being two arbitrary, but mutual orthogonal<br />

eigenvectors. We have<br />

{Φ} = c1{Φ1} + c2{Φ2}<br />

{Φ1}[M]{Φ2} = 0, c 2 1 + c 2 2 =1 ⇒ {Φ} T [M]{Φ} = 1 (18.44)<br />

Inserting this expansion for {Φ} in (18.42) we get<br />

∂ω 2 /∂h = c 2 1g11 + c 2 2g22 +2c1c2g12 with<br />

gmn := {Φm} T (−ω 2 ∂[M]/∂h + ∂[S]/∂h){Φn} (18.45)


Sensitivity analysis for dynamic problems 271<br />

and thus depending on c1,c2. Extreme values of ∂ω2 /∂h can be found by solving the<br />

eigenvalue problem constituted by these sensitivities Additional<br />

<br />

eigenvalue<br />

<br />

<br />

problem<br />

g11<br />

<br />

− g g12 <br />

<br />

g22 − g = 0 (18.46)<br />

g12<br />

The eigenvalues of this matrix are the true sensitivities, with corresponding eigenvectors<br />

{ ˆ Φ}. When the eigenvalues ga,gb to (18.46) are found, we also have the<br />

corresponding eigenvectors {Φa}, {Φb} that return gab = gba = 0, which means that<br />

the mutual sensitivities are zero.<br />

Note, that this case learns us the primary importance of eigenmodes as compared<br />

to eigenvalues. A recent, detailed discussion of the multiple modal problem is given by<br />

(Seyranian, Lund and Olhoff 1994). This reference includes several specific examples.<br />

18.3.6 Double eigenvalue with only a single eigenvector<br />

For non-selfadjoint problems the sensitivity analysis may be further complicated by<br />

the fact that we do not have a complete set of eigenvectors. Thus the case of a double<br />

eigenvalue with the same eigenvector has to be dealt with. Often this is necessary<br />

to describe important physical problems like flutter initialization and interaction of<br />

eigenvalue branches. The paper (Seyranian and Pedersen 1995) treats this in detail.<br />

18.4 Summing up<br />

In this chapter the important results to focus on are: Collected<br />

results<br />

• For non-conservative problems we often find non-intuitive response. This implies<br />

that the necessary sensitivity analysis requires special attention.<br />

• However, with definition of mutual potentials (complex potentials) a major part<br />

of the simplicities are recovered. Especially the possibility to determine eigenvalue<br />

sensitivities without involving eigenvector change is of great importance.<br />

• As for static problems, localized design parameters lead to localized determination<br />

of sensitivities, although physically not localized. Especially for homogeneous<br />

design dependence of mass and stiffness, the resulting expressions are<br />

simple and may be expressed directly in localized potentials.<br />

• Special attention to multiple eigenvalues must be given, and very special attention<br />

to the cases with non-complete eigenvector spaces. For details see (Pedersen<br />

2001b) with an extended list of the references.


272 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Figure 18.2: Gradient functions g(x) for the Hauger columns. Left: Uniform beam.<br />

Right: Linearly tapered beam. 1st row: both external and internal damping. 2nd<br />

row: only internal damping. 3rd row: only external damping. 4th row: no damping.<br />

For more detail see (Pedersen and Seyranian 1983).


Chapter 19<br />

Simplex and a modified<br />

simplex procedure<br />

A modified simplex procedure is developed for solution of truss topology optimization.<br />

The procedure is presented here from a more general point of view, and in (Pedersen<br />

1994) it is documented by the involved Fortran subroutines.<br />

A simple, yet storage-effective ’linear’ programming code is available, and the<br />

assumption of non-negative variables is bypassed without increasing the size of the<br />

problem. Furthermore, the objective is allowed to be summed over not just linear,<br />

but also concave functions.<br />

The procedure differs in three ways from the classic simplex procedure: LP<br />

modifications<br />

• reformulation to non-negative variables is not necessary<br />

• the objective is not restricted to the linear case, but extended to concave functions<br />

• full pivoting and change of basis is obtained without reordering the constraint<br />

coefficient matrix.<br />

Some or all of these aspects can probably be found in other linear programming<br />

procedures. However, linear programming is to a large extent used as a ’black box’,<br />

and the main idea of the present chapter is to show the simplicity of the necessary<br />

programs and thereby make it easier to perform individual modifications. We start<br />

by describing the standard linear programming problem.<br />

273


Separated<br />

objective<br />

Separated<br />

constraints<br />

Regular<br />

sub-matrix<br />

Basis<br />

solution<br />

Feasible<br />

solution<br />

Global<br />

optimal<br />

274 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

19.1 Standard linear programming form<br />

A linear programming problem in a standard form is defined by a linear objective Φ,<br />

here assumed to be minimized<br />

Minimize Φ={C} T {X} = {CI} T {XI} + {CO} T {XO} (19.1)<br />

where the vector {C} contains given cost coefficients and the vector {X} ≥ {0}<br />

contains the unknown non-negative quantities to be found. For later use we have<br />

immediately divided these unknowns into two groups with index I referring to In and<br />

index O referring to Out. This separation is not known in advance, and in essence<br />

constitutes the solution to be found.<br />

The constraints in addition to the constraint of a non-negative solution are described<br />

by the linear set of equations<br />

[A]{X} = {B} or [AI]{XI} +[AO]{XO} = {B} (19.2)<br />

where the coefficient matrices are of order M ×N for matrix [A], order M ×M for the<br />

sub-matrix matrix [AI] and order M × (N − M) for the sub-matrix [AO]. We assume<br />

N>M and [AI] regular, i.e. the inverse matrix [AI] −1 exists. Normally with more<br />

unknowns N than the number of constraints M we have many feasible solutions, and<br />

we must find the one that minimizes the objective Φ.<br />

In the literature on mathematical programming, such as (Gass 1964) and (Hadley<br />

1964), it is proven that the solution is a basis feasible solution (if any), i.e. a solution<br />

with {XO} = {0} from which follows the index notation of O and o for out of basis.<br />

An actual feasible basis solution is found by solving (19.2) in terms of { ˜ XI}, with a<br />

super-index˜ to remember the assumption of {XO} =0,<br />

{ ˜ XI} =[AI] −1 {B} ≥{0} (19.3)<br />

We assume that a set of M unknowns (the selected {XI}) makes a solution like<br />

(19.3) possible, and our goal is then to determine the set which minimizes the objective<br />

(19.1).<br />

19.2 Effect of changing the basis set<br />

Based on the theoretical knowledge that the optimal solution is a global optimal<br />

solution, we in one iteration, search for a better set among the better neighbour solutions.<br />

When this is not possible the limited number of linear programming iterations<br />

are ended and we have found a global optimal (basis feasible) solution.<br />

If several neighbour solutions are better, we have to decide the change, i.e. what<br />

unknown Xo from the set {XO} to be introduced. Among the possible procedures


Simplex and a modified simplex procedure 275<br />

for this we have: here listed with increasing complexity of calculation but also with<br />

increasing effect: Deciding<br />

• The first found unknown with the effect of dΦ/dXo < 0<br />

• The unknown with minimum gradient, i.e. minimum of dΦ/dXo < 0<br />

• The unknown with the best effect, i.e. minimum of ∆Φ<br />

For the criteria based on dΦ/dXo, we determine this gradient, and for this we write<br />

the equilibrium (19.2), with only one variable from the set {XO} not zero, say Xo<br />

[AI]{XI} + {Ao}Xo = {B} or<br />

{XI} = { ˜ XI}−{Zo}Xo with {Zo} := [AI] −1 {Ao} (19.4)<br />

The objective for this solution as seen from (19.1) with (19.4) is<br />

neighbour<br />

solution<br />

Change in<br />

equilibrium<br />

Φ={CI} T ({ ˜ XI}−{Zo}Xo)+CoXo<br />

(19.5)<br />

with a resulting gradient Objective<br />

dΦ/dXo = −{CI} T {Zo} + Co<br />

(19.6)<br />

derivative<br />

This is information enough for the first two procedures.<br />

However, to carry through the change of basis we also need to know the unknown<br />

actually in basis, Xi from the set {XI}, that should be eliminated, i.e. transferred to<br />

the set out of basis {XO}. This is determined by the condition of only non-negative<br />

solution. From (19.4) this condition reads Eliminated<br />

19.3 Non-linear objective<br />

˜Xi − (Zi)oXo > 0 for all i giving<br />

Xo = Minimum ˜ Xi/(Zi)o for (Zi)o > 0<br />

and this also locates the actual i (19.7)<br />

variable<br />

from basis<br />

With a non-linear objective we may still be interested in locating a solution at a<br />

vertex, i.e. a set {XO} = {0}, for which all neighbour vertex solutions are not better.<br />

If the objective function is concave, then this locates an optimal solution, possible<br />

local but if the non-linearity is not very strong, then even a global optimal solution,<br />

see (Hadley 1964) and chapter 16 of this book.<br />

The procedure for obtaining such a solution is very closely related to a normal Best vertex<br />

solution


Feasible<br />

basis<br />

solution<br />

276 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

simplex procedure, except that ∆Φ is used for deciding change of basis and not just<br />

dΦ/dXo. Thus for all the unknowns at the present not in basis we first determine Xo<br />

from (19.7). Then the changed solution follows from (19.4) and the changed objective<br />

follows from ∆Φ = ˜ Φ − Φ with ˜ Φ as well as Φ non-linear depending on {XI} and Xo.<br />

It is thus possible to locate the unknown Xo with the best effect, giving minimum<br />

of ∆Φ. If this is negative we change, eliminating the unknown Xi determined from<br />

(19.7). If minimum ∆Φ is positive a vertex solution which is better than all neighbour<br />

vertices solutions is found. This solution is obtained in a finite number of iteration<br />

steps.<br />

19.4 Positive and negative variables<br />

When the unknowns are not necessarily non-negative, a common trick is to use twice<br />

the number of variables {X} = {X + }−{X − } with {X + }≥{0} and {X − }≥{0},<br />

and we then reformulate the constraints (19.2) to<br />

[A]{X + }−[A]{X − } = {B}<br />

[AI]({X +<br />

I }−{X−<br />

+<br />

I })+[AO]({X O }−{X− O })={B} (19.8)<br />

The total coefficient matrix [[A] (−[A])] is now twice the size, which may be critical<br />

for solving a large problem.<br />

However, it should be clear that we do not need to store the information in [A]<br />

twice. But how should we modify the simplex procedure to account for this when<br />

only [A] is stored ? We have already separate the unknowns into the basis variables<br />

being part of the two M × 1 vectors {X +<br />

I }, {X− I } and the out of basis unknowns in<br />

}, {X−}.<br />

A basis solution is obtained by setting<br />

the two (N − M) × 1 vectors {X +<br />

O<br />

} = {X−}<br />

= {0} , and we have<br />

{X +<br />

O<br />

O<br />

O<br />

{ ˜ X +<br />

I }−{˜ X −<br />

I } =[AI] −1 {B} = { ˜ B} (19.9)<br />

In fact we do not evaluate [AI] −1 , so here we just assume that it exists, i.e. that the<br />

M × M matrix [AI] is regular.<br />

In terms of the individual components the interpretation of the solution (19.9) is<br />

then<br />

˜Bi ≥ 0 ⇒ X +<br />

i = ˜ Bi, X −<br />

i =0<br />

˜Bi < 0 ⇒ X +<br />

i =0,X− i = − ˜ Bi (19.10)<br />

i.e. the signs of the right-hand-side elements after elimination directly focus on half<br />

of the unknowns in { ˜ X +<br />

}, and we have as usual only M basis unknowns.<br />

I }, { ˜ X −<br />

I


Simplex and a modified simplex procedure 277<br />

With the linear combinations [Z] :=[AI] −1 [AO], as in the classic simplex procedure,<br />

the constraint (19.8) is written as<br />

{ ˜ X +<br />

I }−{˜ X −<br />

I } = { ˜ B}−[Z]{X +<br />

O<br />

} +[Z]{X− O } (19.11)<br />

and with only one out-of-basis variable Xo different from zero, equation (19.8) gives<br />

{ ˜ X +<br />

I }−{˜ X −<br />

I } = { ˜ B}−{Zo}X + o + {Zo}X − o (19.12)<br />

The limiting numerical value of X + o or of X− o is obtained by the conditions that<br />

the basis variables do not change sign (in classic simplex do not become negative).<br />

Two cases, each with two sub-cases, are considered here.<br />

Case 1: Assume X + o to be introduced, i.e. X− o = 0, then the row i of (19.12) is<br />

˜X +<br />

i − ˜ X −<br />

i = ˜ Bi − (Zi)oX + o (19.13)<br />

According to (19.10) we have two sub-cases:<br />

Case 1a:<br />

˜Bi ≥ 0 ⇒ X +<br />

i = ˜ Bi − (Zi)oX + o<br />

and X +<br />

i only keeps its sign for ˜ Bi − (Zi)oX + o ≥ 0 (19.14)<br />

Case 1b: Value of<br />

˜Bi < 0 ⇒ −X −<br />

i = ˜ Bi − (Zi)oX + o<br />

and X −<br />

i only keeps its sign for ˜ Bi − (Zi)oX + o ≤ 0 (19.15)<br />

Case 2: Assume X− o to be introduced, i.e. X + o = 0, then the row i of (19.12) is<br />

˜X +<br />

i − ˜ X −<br />

i = ˜ Bi +(Zi)oX − o (19.16)<br />

with the two sub-cases:<br />

Case 2a:<br />

˜Bi ≥ 0 ⇒ X +<br />

i = ˜ Bi +(Zi)oX − o<br />

and X +<br />

i only keeps its sign for ˜ Bi +(Zi)oX − o ≥ 0 (19.17)<br />

Case 2b: Two<br />

˜Bi < 0 ⇒ −X −<br />

i = ˜ Bi +(Zi)oX − o<br />

and X −<br />

i only keeps its sign for ˜ Bi +(Zi)oX − o ≤ 0 (19.18)<br />

The condition (19.14) on X + o is only active for (Zi)o > 0, and in (19.15) only for<br />

(Zi)o < 0. Together these two conditions are written ˜ Bi(Zi)o > 0 for case 1.<br />

variable<br />

introduced<br />

sub-cases


Collected<br />

results<br />

278 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

In (19.17) and (19.18) we have opposite signs, i.e. totally ˜ Bi(Zi)o < 0 for case 2.<br />

With these conditions programmed we can determine the value of X + o or of X − o and<br />

the corresponding new values of the old basis unknowns. This information is enough<br />

to test the optimality criterion and if not optimal then decide the simplex iteration,<br />

i.e. change of basis.<br />

19.5 Summing up<br />

In this chapter the important results to focus on are:<br />

• A very compact description of standard simplex procedure is presented, (some<br />

basis knowledge of linear programming is assumed).<br />

• Three possibilities for deciding the change of basis are described.<br />

• Still having a convex feasible solution space, we locate the best vertex solution,<br />

also with a non-linear objective.<br />

• From a procedure point of view, the doubling of the number of unknowns is<br />

not necessary to obtain a formulation with non-negative unknowns.


Chapter 20<br />

Linear programming<br />

reformulations with SLP<br />

20.1 Introduction<br />

In a standard formulation, the linear programming problem is in matrix notation<br />

stated Standard<br />

LP<br />

Determine non − negative variables {X} ≥{0} that<br />

Minimize the scalar product {C} T {X}<br />

subjected to the constraints that {X} satisfies the equation<br />

[A]{X} = {B} (20.1)<br />

where the vector {C} of cost coefficients and the matrix [A] as well as the vector {B}<br />

are also assumed to be given.<br />

Most formulations from initial an setup are not presented directly in this standard<br />

form. The unknowns may not be non-negative, the objective and the constraints<br />

may not be linear, the constraint equation may be a combination of equalities and<br />

inequalities, maximizing may be the objective and not minimizing.<br />

By proper reformulations we can obtain the standard formulation (20.1), for nonlinear<br />

problems at least for an iterative step. Using linear programming iteratively<br />

we name it sequential linear programming (SLP). Sequential<br />

The goal of the present chapter is to show these reformulations. However, the<br />

chapter also gives a more general, but very short, introduction to the aspects of linear<br />

programming (LP), presented in matrix notation. The linear programming literature<br />

279<br />

LP, i.e. SLP


IMSL routine<br />

DDLPRS<br />

Maximize<br />

by minimize<br />

280 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

is very rich and the present author is most familiar with the earlier books of (Gass<br />

1964) and (Hadley 1964).<br />

Using available software like the IMSL routine DDLPRS, some of the reformulations<br />

are taken care of by the routine. The final part of the chapter is written to show<br />

how input to this routine can be reformulated directly into a starting basis solution,<br />

using notation related to this IMSL routine.<br />

20.2 Actual optimization problem after sign reformulations<br />

To get some uniformity we treat a problem where an objective Φ should be minimized<br />

subjected to less than or equal to inequalities. Thus maximization and greater than<br />

inequalities must be reformulated, which is simply done by proper change of sign<br />

Maximize Φ ⇒ Minimize − Φ and<br />

gj > ¯gj ⇒ −gj < −¯gj (20.2)<br />

Greater than<br />

by less than Without loss of generality we can therefore state the optimal design problem as<br />

Taylor<br />

expansion<br />

Determine design variables hi for i =1, 2,...,I that<br />

Minimize Φ as objective subjected to the constraints<br />

gj ≤ ¯gj for j =1, 2,...,J orinvector notation<br />

{g} ≤{¯g} (20.3)<br />

20.3 From non-linear to sequential linear problem<br />

Most practical optimal design problems are not linear and in general both the objective<br />

Φ and the constraint vector {g} may be non-linear dependent on the vector of<br />

design variables {h}.<br />

A Taylor expansion is assumed to be valid, at least close to an actual design<br />

specified at design iteration n by {h}n. The expansions in terms of sensitivity vector<br />

{dΦ/dh}n for the objective and sensitivity matrix [d{g}/d{h}]n for the constraints<br />

are<br />

Φn+1 = Φn + {dΦ/dh} T n {∆h}n<br />

{g}n+1 = {g}n +[d{g}/d{h}]n{∆h}n ≤{¯g} (20.4)


Linear programming reformulations with SLP 281<br />

Then with Φn a given objective value corresponding to the design {h}n the optimal<br />

redesign problem is stated<br />

Determine redesign variables {∆h}n that<br />

Minimize {dΦ/dh} T n {∆h}n as objective subjected to the constraints<br />

[d{g}/d{h}]n{∆h}n ≤{¯g}−{g}n<br />

(20.5)<br />

After solving this problem the design is updated to {h}n+1 = {h}n + {∆h}n and the Redesign<br />

errors of the Taylor expansion are taken care of in following iterations, assuming that<br />

smaller and smaller move-limits are put on the changes {∆h}.<br />

20.4 From inequalities to equalities<br />

A standard technique to transform inequalities to equalities is to introduce additional<br />

variables, named slack variables. With less than or equal to constraints these slack Slack<br />

variables are by definition non-negative variables. We indicate them by a super-posed<br />

˜, and show them here by the vector {˜g}<br />

[d{g}/d{h}]n{∆h}n + {˜g}n = {¯g}−{g}n<br />

(20.6)<br />

The slack variables have no influence on the objective, normally specified by the<br />

fact that their cost coefficients are zero. If the constraint without slack variables<br />

must be equality, we obtain that by forcing the corresponding slack variable to be<br />

zero. This is done by a relative large cost coefficient, as explained in more detail in<br />

section 20.9.<br />

20.5 Obtaining non-negative variables<br />

formulation<br />

variables<br />

When a minimum allowable redesign vector {∆h}min is specified ({∆h}n ≥{∆h}min),<br />

as is natural with Taylor expansions, then the transformation to non-negative redesign<br />

variables {∆h} + can be obtained by the change of variables Shift of<br />

unknowns<br />

{∆h} + = {∆h}n −{∆h}min ≥{0} or<br />

{∆h}n = {∆h} + + {∆h}min (20.7)<br />

and this transformation must then be introduced into the objective as well as into the<br />

constraints. Having solved the redesign problem in terms of {∆h} + we back-substitute<br />

to obtain the ”physical” redesign {∆h}. Detail is given in section 20.8.<br />

In addition to the minimum redesign vector {∆h}min we also specify a maximum<br />

redesign vector {∆h}max, and therefore have the following bounds on the redesign<br />

variables Bounds on<br />

redesign


Non-negative by<br />

move-limits<br />

Equality by<br />

slack variables<br />

Constraint<br />

types<br />

282 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

{∆h}min ≤{∆h} ≤{∆h}max or<br />

{0} ≤{∆h} + ≤{∆h}max −{∆h}min<br />

(20.8)<br />

In order to get more compact notation and without direct reference to our design<br />

variables, we in the following sections shift to the nomenclature used in the description<br />

of the IMSL routine, see (Press et al. 1992). The description begins with the movelimits<br />

as exemplified in (20.9).<br />

20.6 Move-limits and non-negative variables<br />

This section more or less repeats section 20.5 with alternative notation, the notation<br />

of the mentioned IMSL routine. The solution vector is {X} and the move-limits {XL}<br />

and {XU } give the constraints<br />

{XL} ≤{X} ≤{XU } (20.9)<br />

The number of variables is N. To convert to non-negative variables {X + } we define<br />

this vector as<br />

{X + } = {X}−{XL} ⇒ {X} = {X + } + {XL} (20.10)<br />

and the move-limits are then written as the constraints<br />

{0} ≤{X + }≤{XU }−{XL} (20.11)<br />

To convert the right most inequality to equality we introduce the non-negative slack<br />

variables { ˜ X}≥{0} and finally write the move-limit constraints<br />

20.7 Problem constraints<br />

{ ˜ X} + {X + } = {XU }−{XL} ≥{0} (20.12)<br />

The right hand side of a specific constraint Ri is defined by<br />

Ri = {Ai} T {X} = <br />

A(i, j)X(j) (20.13)<br />

For Ri a constraint type is specified by Ii =0, 1, 2, or3, and the constraints are for<br />

the individual types:<br />

j


Linear programming reformulations with SLP 283<br />

Ii = 0 ⇒ Ri = BiL = BiU (20.14)<br />

Ii = 1 ⇒ Ri ≤ BiU (20.15)<br />

Ii = 2 ⇒ Ri ≥ BiL ⇒ −Ri≤−BiL (20.16)<br />

Ii = 3 ⇒ BiL ≤ Ri ≤ BiU, i.e. (20.17)<br />

Ii = 3 ⇒ Ri ≤ BiU and − Ri ≤−BiL (20.18)<br />

where the BiL and the BiU are user specified limits. Double<br />

The constraints of type Ii = 3 we treat as two separated constraints (type Ii =<br />

1 and type Ii = 2). For this we define an extended matrix [ Ā] where the rows<br />

corresponding to the actual Ii = 3 are repeated and an extended vector { ¯ B} with<br />

elements BiL or BiU in agreement with the type Ii<br />

dimension of the matrix [<br />

as specified in (20.18). The<br />

Ā] and the vector { ¯ B} is K<br />

K = M + number of type Ii = 3 (20.19)<br />

where M is the number of specified problem constraints.<br />

The problem constraints can now all together be written as<br />

{ ˜ Y } +[ Ā]{X} = { ¯ B} (20.20)<br />

where we have introduced K further non-negative slack variables by the vector { ˜ Y }≥<br />

{0}. Note that the slack variables corresponding to problem constraints of type Ii =0<br />

constraints<br />

separated<br />

in the final solution must be zero for the solution to be feasible. Further<br />

Changing the constraints (20.20) by introducing the non-negative variables {X + }<br />

from (20.10) we get<br />

{ ˜ Y } +[ Ā]{X+ } = { ¯ B}−[ Ā]{XL} (20.21)<br />

If the right hand side of (20.21) is non-negative, i.e. { ¯ B}−[ Ā]{XL} ≥{0} then we<br />

have a starting solution. However, for some elements in this vector this may not be<br />

the case. Let there be L such element and we then introduce L artificial variables<br />

collected in the vector {Z}. For the corresponding L equations in (20.20) we change<br />

sign and write these L constraints<br />

{Z}−{ ˜ Y − }−[ Ā− ]{X + } = −{ ¯ B − } +[ Ā− ]{XL} (20.22)<br />

slack variables<br />

now with non-negative right hand side. The remaining K − L constraints of (20.21)<br />

is kept unchanged (except for the fewer number of constraints) Artificial<br />

variables<br />

{ ˜ Y + } +[ Ā+ ]{X + } = { ¯ B + }−[ Ā+ ]{XL} (20.23)


Collected<br />

constraints<br />

First basis<br />

solution<br />

284 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

20.8 Total LP coefficient matrix for the constraints<br />

Collecting all the unknown vectors { ˜ X}, {Z}, { ˜ Y + }, { ˜ Y − }, {X + } of dimensions N, L, K−<br />

L, L, N, respectively, we write the 3 constraint matrix equations (20.12), (20.22),<br />

(20.23) in one total matrix equation<br />

⎡<br />

⎣<br />

[I] [0] [0] [0] [I]<br />

[0] [I] [0] −[I] −[ Ā− ]<br />

[0] [0] [I] [0] [ Ā+ ]<br />

⎧<br />

⎤<br />

⎪⎨<br />

⎦<br />

⎪⎩<br />

{ ˜ X}<br />

{Z}<br />

{ ˜ Y + }<br />

{ ˜ Y − }<br />

{X + }<br />

⎫<br />

⎧<br />

⎪⎬ ⎨<br />

=<br />

⎩<br />

⎪⎭<br />

{XU }−{XL}<br />

−{ ¯ B − } +[ Ā− ]{XL}<br />

{ ¯ B + }−[ Ā+ ]{XL}<br />

⎫<br />

⎬<br />

⎭ (20.24)<br />

The three rows of equations have the dimensions N, L, K − L, so totally the order of<br />

the total coefficient matrix is (N + K) × (2N + K + L). Note that in a programming<br />

procedure, the original sequence of the slack variables { ˜ Y } need not be resequenced<br />

as shown here by { ˜ Y + }, { ˜ Y − }; the change of signs are just done appropriately.<br />

20.9 The vector of cost coefficients<br />

We see directly from (20.24) that a possible solution to this system of equations is<br />

⎧ ⎫ ⎧<br />

⎫<br />

⎪⎨<br />

⎪⎩<br />

{ ˜ X}<br />

{Z}<br />

{ ˜ Y + }<br />

{ ˜ Y − }<br />

{X + }<br />

⎪⎬ ⎪⎨<br />

=<br />

⎪⎭<br />

⎪⎩<br />

{XU }−{XL}<br />

−{ ¯ B − } +[ Ā− ]{XL}<br />

{ ¯ B + }−[ Ā+ ]{XL}<br />

{0}<br />

{0}<br />

⎪⎬<br />

⎪⎭<br />

(20.25)<br />

Although all elements of the right hand side are non-negative, this is not necessary a<br />

feasible solution (only a basic solution), because the artificial variables in {Z} are not<br />

necessarily all zero and also the slack variables in { ˜ Y + } and/or { ˜ Y − } corresponding<br />

to type Ii = 0 must be zero for the solution to be feasible.<br />

To obtain this as part of the iterative improvement of the solution, the cost factors<br />

for these variables are set to a large value, here indicated by ∞. The objective Φ to<br />

be minimized is then given by<br />

Φ={0} T { ˜ X}+{∞} T {Z}+{∞ or 0} T { ˜ Y + }+{∞ or 0} T { ˜ Y − }+{C} T {X + } (20.26)<br />

where the vector {C} contains the coefficient of the original objective<br />

Φ={C} T {X} = {C} T {X + } + {C} T {XL} (20.27)


Linear programming reformulations with SLP 285<br />

with the last term {C} T {XL} without influence on the optimization, because {XL}<br />

is a user given vector.<br />

20.10 Sequential linear programming<br />

In the presentation of {XL} and {XU } as move-limits it is assumed that the linear LP in SLP<br />

programming solution is used in a sequence of LP to solve a non-linear optimization<br />

problem. The total procedure is then named sequential linear programming (SLP).<br />

For a non-linear problem absolute limits on the resulting final values of the Absolute<br />

unknowns may also be given. Adjusting the move-limits in each linear programming<br />

setup, we can account also for these constraints, without introducing further variables<br />

and constraints.<br />

20.11 Summing up<br />

limits<br />

In this chapter the important results to focus on are: Collected<br />

results<br />

• The initial sections directly focus on optimal design as the actual optimization<br />

problem. Especially the formulation of sequential linear programming to solve<br />

non-linear design problems is important, see also the illustrative figure 2.1 in<br />

chapter 2.<br />

• The chapter assumes some basic knowledge on linear programming from the<br />

reader, and in the description from section 20.6 it is directly related to the often<br />

used IMSL routine DDLPRS.<br />

• The use of move-limits convert to non-negative variables is essential. Furthermore,<br />

move-limits are very useful for user control of more difficult design<br />

optimization.<br />

• The use of slack variables is the tool to convert inequalities to equalities, and<br />

this as a result gives information about non-used bounds.<br />

• The use of artificial variables is the tool to obtain a first basis solution. Furthermore,<br />

if the artificial variables are not zero after the final iteration, then<br />

no solution exists. Thus the artificial-variables are also the tool that give information<br />

about existence of solution.<br />

• The sub-matrix presentation (20.24) is a valuable tool to get an overview of the<br />

problem formulation.


286 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...


Bibliography<br />

Apostol, V., Santos, J. L. T. and Goia, I. (1995), Design sensitivity analysis and optimization<br />

of truss/beam structures with arbitrary cross-sections, in N. Olhoff and G. I. N.<br />

Rozvany, eds, ‘WCSMO1 First World Congress on Structural and Multi-disciplinary<br />

Optimization’, ISSMO.<br />

Araujo, A. L., Soares, C. M. M., Freitas, M. J. M., Pedersen, P. and Herskovits, J. (2000),<br />

‘Combined numerical-experimental model for the identification of mechanical properties<br />

of laminated structures’, Composite Structures 50, 363–372.<br />

Arora, J. S. (1989), Introduction to optimum design, McGraw-Hill. 623 p.<br />

Bagge, M. (1999), Remodeling of Bone Structures, DCAMM S84, <strong>Solid</strong> <strong>Mechanics</strong>, DTU -<br />

thesis for the Ph.D.<br />

Bagge, M. (2000), ‘A model of bone adaption as an optimization process’, J. of Biomechanics<br />

33, 1349–1357.<br />

Banichuk, N. V. (1977), ‘<strong>Optimal</strong>ity conditions in the problem of seeking the hole shapes in<br />

elastic bodies’, PMM 41(5), 946–951.<br />

Banichuk, N. V. (1983), Problems and Methods of <strong>Optimal</strong> Structural Design, Plenum Press,<br />

New York.<br />

Bathe, K. J. (1982), Finite Element Procedures in Engineering Analysis, Prentice-Hall.<br />

Bendsøe, M. P. (1991), <strong>Optimal</strong> topology design and homogenization, in Blanc and R. ans<br />

Suquet, eds, ‘<strong>Mechanics</strong>, Numerical Modelling and Dynamics of Materials’, CNRS.<br />

Bendsøe, M. P. (1995), Optimization of structural topology, shape and material, Springer.<br />

Bendsøe, M. P. and Sigmund, O. (1999), ‘Material interpolation schemes in topology optimization’,<br />

Arch. Appl. Mech.<br />

Bendsøe, M. P. and Sigmund, O. (2003), Topology Optimization - Theory, Methods and<br />

Applications, Springer. 370 pages.<br />

Bendsøe, M. P., Guedes, J. M., Haber, R. B., Pedersen, P. and Taylor, J. E. (1994), ‘An analytical<br />

model to predict optimal material properties in the context of optimal structural<br />

design’, J. Applied <strong>Mechanics</strong> 61, 930–937.<br />

287


288 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Bergmann, G., Graichen, F. and Rohlmann, A. (1993), ‘Hip joint loading during walking<br />

and running, measured in two patients’, J. of Biomechanics 26(8), 969–990.<br />

Bert, C. W. (1977), ‘<strong>Optimal</strong> design of a composite-material plate to maximize its fundamental<br />

frequency’, J. Sound and Vibration 50, 229–239.<br />

Borkowski, A., Jendo, S. and Reitman, M. I. (1990), Structural optimization, mathematical<br />

programming, Vol. 2, Plenum Press.<br />

Brandt, A. M., ed. (1989), Foundations of optimum design in civil engineering, Nijhoff,<br />

Dordrecht, The Netherlands.<br />

Cheng, G. and Pedersen, P. (1997), ‘On sufficiency conditions for optimal design based on<br />

extremum principles of mechanics’, J. <strong>Mechanics</strong> of Physics and <strong>Solid</strong>s 45(1), 135–150.<br />

Cherepanov, G. P. (1974), ‘Inverse problems of a plane theory of elasticity’, PMM.<br />

Cherkaev, A., Lurie, K. and Milton, G. W. (1992), ‘Invariant properties of the stress in plane<br />

elasticity and equivalence classes of composites’, Proc. R. Soc. Lond. A 438, 519–529.<br />

Cherkaev, A. V., Grabovsky, Y., Movchan, A. B. and Serkov, S. K. (1998), ‘The cavity of<br />

the optimal shape under shear stresses’, Int. J. <strong>Solid</strong>s and Structures.<br />

Christensen, R. M. (1993), ‘Effective properties of composite materials containing voids’,<br />

Proc. R. Soc. Lond. A 440, 461–473.<br />

da Silva Smith, O. (1997), ‘Topology optimization of trusses with local stability constraints<br />

and multiple loading conditions - a heuristic approach’, Structural Optimization<br />

13, 155–166.<br />

de Vissher, J., Sol, H., de Wiede, W. P. and Vantomme, J. (1997), Identification of the complex<br />

moduli of thin fibre reinforced polymer plates using measured modal parameters,<br />

in H. Sol and C. W. J. Oomens, eds, ‘Proc. Euromech 357’, Kluwer, Keikrade, The<br />

Netherlands, pp. 1–9.<br />

Dems, K. and Mroz, Z. (1978), ‘Multiparameter structural shape optimization by finite<br />

element method’, Int. J. Num. Meth. Eng. 13, 247–263.<br />

Ding, Y. (1986), ‘Shape optimization of structures: A literature survey’, Computers and<br />

Structures 24(6), 985–1004.<br />

Dorn, W. S., Gomory, R. E. and Greenberg, H. J. (1964), ‘Automatic design of optimal<br />

structures’, J. de Mechanique 3, 25–52.<br />

Dybbro, J. D. and Holm, N. C. (1986), ‘On minimization of stress concentration for threedimen-<br />

sional models’, Computers and Structures 4, 637–643.<br />

Eschenauer, H. A. and Olhoff, N. (2001), ‘Topology optimization of continuum structures -<br />

a review’, Applied <strong>Mechanics</strong> Reviews 54, 331–390. 425 references.<br />

Eshelby, J. D. (1957), ‘The determination of the elastic field of an ellepsoidal inclusion, and<br />

related problems’, Proc. R. Soc. Ser. A.<br />

Fedorov, A. V. and Cherkaev, A. V. (1983), ‘Choice of optimal orientation of axes of elastic<br />

symmetry for an orthotropic plate’, MTT 18, 135–142.


Bibliography 289<br />

Fleron, P. (1964), ‘The minimum weight of trusses’, Bygningsstatiske Meddelelser 35, 81–96.<br />

Francavilla, A., Ramakrishnan, C. V. and Zienkiewicz, O. C. (1975), ‘Optimization of shape<br />

to minimise stress concentration’, J. Strain Analysis 10, 63–70.<br />

Frederiksen, P. S. (1992), Identification of material parameters in anisotropic plates - a<br />

combined numerical/experimental method, DCAMM Sxx - thesis for the Ph.D.<br />

Frederiksen, P. S. (1995), ‘Single-layer plate theories applied to the flexural vibration of<br />

completely free thick laminates’, J. of Sound and Vibration 186, 743–759.<br />

Frederiksen, P. S. (1997a), ‘Experimental procedure and results for the identification of<br />

elastic constants of thick orthotropic plates’, J. Composite Materials 31, 360–382.<br />

Frederiksen, P. S. (1997b), ‘Numerical studies for the identification of orthotropic elastic<br />

constants of thick plates’, Eur. J. Mech., A/<strong>Solid</strong>s 16, 117–140.<br />

Frederiksen, P. S. (1998), ‘Parameter uncertainty and design of optimal experiments for the<br />

estimation of elastic constants’, Int. J. <strong>Solid</strong>s Structures 35, 1241–1260.<br />

Gass, S. I. (1964), Linear Programming, second edn, McGraw-Hill. 280 pages.<br />

Grabovsky, Y. and Kohn, R. V. (1995), ‘Microstructures minimizing the energy of a two<br />

phase elastic composite in two space dimensions, ii: The vigdergauz microstructure’, J.<br />

Mech. Phys. <strong>Solid</strong>s 43(6), 949–972.<br />

Guedes, J. M. (1995), Users Manual for the program PREMAT3D.<br />

Guedes, J. M. and Kikuchi, N. (1990), ‘Preprocessing and postprocessing for materials based<br />

on the homogenization method with adaptive finite element methods’, Computer Metrhods<br />

in Applied <strong>Mechanics</strong> and Engineering 83, 143–198.<br />

Gürdal, Z., Haftka, R. T. and Hajela, P. (1999), Design and Optimization of Laminated<br />

Composite Materials, Wiley. 337 pages.<br />

Hadley, G. (1964), Nonlinear and Dynamic Programming, Addison-Wesley. 484 pages.<br />

Haftka, R. T., Gurdal, Z. and Kamat, M. P. (1990), Elements of Structural Optimization,<br />

Kluwer.<br />

Hammer, V. B. (1994), Design of composite laminates with optimized stiffness, strength and<br />

damage properties, DCAMM S72, <strong>Solid</strong> <strong>Mechanics</strong>, DTU - thesis for the Ph.D.<br />

Hammer, V. B. and Pedersen, P. (1999), ‘On an orthotropic model for progressive degradation’,<br />

Composite Structures 46, 217–228.<br />

Hammer, V. B., Bendsøe, M. P., Lipton, R. and Pedersen, P. (1997), ‘Parametrization in<br />

laminate design for optimal compliance’, Int. J. of <strong>Solid</strong>s and Structures 34(4), 415–434.<br />

Hill, R. (1950), The Mathematical Theory of Plasticity, Oxford University Press, London.<br />

Hill, R. (1956), ‘New horizons in the mechanics of solids’, J. Mech. Phys. <strong>Solid</strong>s 5, 66–74.<br />

Huiskes, R. and Kaastad, T. S. (2000), ‘Biomechanics, bone quality and strength’, pp. 54–71.<br />

Jacobs, C., Simo, C., Beaupre, G. and Carter, D. (1997), ‘Adaptive bone remodelling incorporating<br />

simultaneous density and anisotropy considerations’, J. of Biomechanics<br />

30, 603–613.


290 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Jones, R. M. (1975), <strong>Mechanics</strong> of Composite Materials, McGraw-Hill.<br />

Kirch, U. (1993), Structural optimization: fundamentals and applications, Springer. 302<br />

pages.<br />

Kleiber, M., Antunez, H., Hien, T. D. and Kowalczyk, P. (1997), Parameter Sensitivity in<br />

Nonlinear <strong>Mechanics</strong>, Wiley. 406 pages.<br />

Kristensen, E. S. and Madsen, N. F. (1974), Optimering af kærv i plan spændingstilstand<br />

(in Danish), M.Sc. Thesis, <strong>Solid</strong> <strong>Mechanics</strong>, DTU.<br />

Kristensen, E. S. and Madsen, N. F. (1976), ‘On the optimum shape of fillets in plates<br />

subjected to multiple in-plane loading cases’, Int. J. Numer. Meth, Eng. 10, 1007–<br />

1019.<br />

Kuttenkeuler, J. (1998), Aircraft composites and aeroelastic tailoring, Thesis for the Ph.D. -<br />

Royal Institute of Technology, Sweden.<br />

Levinson, M. (1981), ‘An accurate, simple theory of the statics and dynamics of elastic<br />

plates’, Mech. Res. Commun 7, 81–87.<br />

Lipton, R. (1994), ‘On optimal reinforcement of plates and choice of design parameters’,<br />

Control Cypernetics 23, 481–493.<br />

Markworth, N. J. and Petersen, C. (1987), Identifikation af Materialeparametre for Fiberbaserede<br />

Laminatplader (McS thesis in Danish), <strong>Solid</strong> <strong>Mechanics</strong>, DTU.<br />

Muskhelishvili, N. I. (1934), A new General Method of Solution of the Fundamental Boundary<br />

Value Problems in Plane Theory of Elasticity, Dokl. Akad. Nauk SSSR.<br />

Muskhelishvili, N. I. (1953), Some Basic Problems of the Mathematical Theory of Elasticity,<br />

Noordhoff.<br />

Niordson, F. and Pedersen, P. (1973-74), A review of optimal structural design, in ‘Thirteenth<br />

Int. Congress of Theoretical and Applied <strong>Mechanics</strong>’, IUTAM, Moscow, Russia, pp. 264–<br />

278. 148 references.<br />

Niordson, F. I. (1965), ‘On the optimal design of a vibrating beam’, Quarterly of Applied<br />

Mathematics 23, 47–53.<br />

Olhoff, N. (1976), ‘Optimization of vibrating beams with respect to higher order natural<br />

frequencies’, J. Structural <strong>Mechanics</strong> 4, 87–122.<br />

Olhoff, N. (1980), <strong>Optimal</strong> design with respect to structural eigenvalues, in F. P. J. Rimrott<br />

and B. Tabarrot, eds, ‘IUTAM Congress’, North-Holland, Canada, pp. 133 –149.<br />

Olhoff, N. and Taylor, J. E. (1983), ‘On structural optimization’, Journal of Applied <strong>Mechanics</strong><br />

50, 1139–1151. 58 references.<br />

Pedersen, D. R., Brand, R. A. and Davy, D. T. (1997), ‘Pelvic muscle and acetabular contact<br />

forces during gait’, J. of Biomechanics 30, 959–965.<br />

Pedersen, N. L. and Nielsen, A. K. (2002), ‘Optimization of practical trusses with constraints<br />

on eigenfrequencies, displacements, stresses and buckling’, Structural Optimization.


Bibliography 291<br />

Pedersen, P. (1969), On the minimum mass layout of trusses, in ‘AGARD Symposium on<br />

Structural Optimization’, Vol. 36, NATO, Istanbul, Turkey, pp. 11–1–17.<br />

Pedersen, P. (1970), <strong>Optimal</strong> Layout of Trusses, DCAMM S1, <strong>Solid</strong> <strong>Mechanics</strong>, DTU - thesis<br />

for the Ph.D.<br />

Pedersen, P. (1973), ‘<strong>Optimal</strong> joint positions for space trusses’, J. of the Structural Division,<br />

ASCE 99(ST 12), 2459–2476.<br />

Pedersen, P. (1981), The integrated approach of fem-slp for solving problems of optimal<br />

design, in E. Haug and J. Cea, eds, ‘Optimization Distributed Parameter Structures’,<br />

Vol. 49 of NATO ASI series, Sijthoff and Noordhoff, pp. 757–780.<br />

Pedersen, P. (1982–83), ‘Design with several eigenvalue constraints by finite elements and<br />

linear programming’, J. Structural <strong>Mechanics</strong> 10(3), 243–271.<br />

Pedersen, P. (1984), Sensitivity analysis for non-selfadjoint problems, in V. Komkov, ed.,<br />

‘Sensitivity of Functionals with Applications to Engineering Sciences’, Vol. 1086 of<br />

Lecture Notes in Mathematics, American Mathematical Society, Springer, New York,<br />

USA, pp. 119–130.<br />

Pedersen, P. (1986a), Identification of models for dynamic systems, in G. Bianchi and<br />

W. Schielen, eds, ‘Dynamics of Multi-body Systems’, Springer, Udine, Italy, pp. 165–<br />

175.<br />

Pedersen, P. (1986b), Minimum flexibility of non-harmonic loaded laminated plates, in<br />

R. Pyrz, ed., ‘Mechanical Characterization of Fiber Composite Materials’, Aalborg,<br />

Denmark, pp. 182–196.<br />

Pedersen, P. (1987a), ‘A note on plasticity theory in matrix notation’, Comm. in Applied<br />

Numerical Methods 3, 541–546.<br />

Pedersen, P. (1987b), ‘On sensitivity analysis of optimal design of specially orthotropic laminates’,<br />

Engineering Optimization 11, 305–316.<br />

Pedersen, P. (1988), Design for minimum stress concentration - some practical aspects.,<br />

in G. Rozvany and B. Karihaloo, eds, ‘Structural optimization’, Kluwer, Melbourne,<br />

Australia, pp. 225–232.<br />

Pedersen, P. (1989), ‘On optimal orientation of orthotropic materials’, Structural Optimization<br />

1, 101–106.<br />

Pedersen, P. (1990), ‘Bounds on elastic energy in solids of orthotropic materials’, Structural<br />

Optimization 2, 55–63.<br />

Pedersen, P. (1991), ‘On thickness and orientational design with orthotropic materials’,<br />

Structural Optimization 3, 69–78.<br />

Pedersen, P. (1992), Topology optimization of three-dimensional trusses, in M. Bendsøe and<br />

C. A. Mota-Soares, eds, ‘Topology Design of Structures’, Vol. 227 of NATO ASI series,<br />

Kluwer, Sesimbra, Portugal, pp. 19–30.<br />

Pedersen, P. (1993), Topology optimization of 3d trusses with cost of supports, in J. Herskovits,<br />

ed., ‘Structural Optimization 93’, Vol. I, pp. 11–20.


292 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Pedersen, P. (1994), ‘Modified simplex optimization program’, Comm. in Numerical Methods<br />

in Engineering 10, 303–312.<br />

Pedersen, P. (1995), ‘Simple transformation by proper contracted forms’, Comm. in Numerical<br />

Methods in Engineering 11, 821–829.<br />

Pedersen, P. (1997), Recent developments of optimization formulations for microstructural<br />

based material systems, in ‘4th National Congress on Theoretical and Applied <strong>Mechanics</strong>’,<br />

Leuven, Belgium, pp. 1–6.<br />

Pedersen, P. (1998), ‘Some general optimal design results using anisotropic power law nonlinear<br />

elasticity’, Structural Optimization 15, 73–80.<br />

Pedersen, P. (1999a), Identification techniques in composite laminates, in C. A. Mota Soares,<br />

C. M. Mota Soares and M. J. M. Freitas, eds, ‘<strong>Mechanics</strong> of Composite Materials and<br />

Structures’, NATO ASI series, Kluwer, Troia, Portugal, pp. 443–452.<br />

Pedersen, P. (1999b), Sensitivity analysis and inverse problems for laminates and materials,<br />

in C. A. Mota Soares, C. M. Mota Soares and M. J. M. Freitas, eds, ‘<strong>Mechanics</strong><br />

of Composite Materials and Structures’, NATO ASI series, Kluwer, Troia, Portugal,<br />

pp. 453–463.<br />

Pedersen, P. (2001a), ‘On the influence of boundary conditions, Poisson’s ratio and material<br />

non-linearity on the optimal shape’, Int. J. of <strong>Solid</strong>s and Structures 38(3), 465–477.<br />

Pedersen, P. (2001b), Six lectures on sensitivity analysis for dynamic stability problems, in<br />

‘Stability of Structures: Modern problems and Unconventional Solutions - Udine, Italy’,<br />

CISM. 57 p.<br />

Pedersen, P. and Bendsøe, M. P. (1995), On strain-stress fields resulting from optimal orientation,<br />

in N. Olhoff and G. I. N. Rozvany, eds, ‘WCSMO1 First World Congress on<br />

Structural and Multi-disciplinary Optimization’, ISSMO, Pergamon, pp. 243–249.<br />

Pedersen, P. and Bendsøe, M. P., eds (1999), Synthesis in Bio <strong>Solid</strong> <strong>Mechanics</strong>, Vol. 69 of<br />

<strong>Solid</strong> <strong>Mechanics</strong> and its Applications, IUTAM, Kluwer, Lyngby, Denmark.<br />

Pedersen, P. and Frederiksen, P. S. (1992), ‘Identification of orthotropic material moduli by<br />

a combined experimental/numerical methods’, Measurement 10(3), 113–118.<br />

Pedersen, P. and Jørgensen, L. (1984), ‘Minimum mass design of elastic frames subjected to<br />

multiple load cases’, Computers and Structures 18, 147–157.<br />

Pedersen, P. and Laursen, C. L. (1982–83), ‘Design for minimum stress concentration by<br />

finite elements and linear programming’, J. Structural <strong>Mechanics</strong> 10(4), 375–391.<br />

Pedersen, P. and Taylor, J. E. (1993), <strong>Optimal</strong> design based on power-law non-linear elasticity,<br />

in P. Pedersen, ed., ‘<strong>Optimal</strong> Design with Advanced Materials,’, Elsevier, pp. 51–66.<br />

Pedersen, P. and Tortorelli, D. A. (1998), ‘Constitutive parameters and their evolution’, J.<br />

of Control and Cybernetics 27(2), 295–310.<br />

Pedersen, P., Cheng, G. and Rasmussen, J. (1989), ‘On accuracy problems for semi-analytical<br />

sensitivity analysis’, <strong>Mechanics</strong> of Structures and Materials 17(3), 373–384.


Bibliography 293<br />

Pedersen, P., Mortensen, S. A. and Larsen, D. G. (1998), A procedure for contact problems<br />

used in hardness test analysis, in A. Eriksson and C. Pacoste, eds, ‘NSCM11: Nordic<br />

Seminar on Computational <strong>Mechanics</strong>’, pp. 31–34.<br />

Pedersen, P., Tobiesen, L. and Jensen, S. H. (1992), ‘Shapes of orthotropic plates for minimum<br />

energy concentration’, <strong>Mechanics</strong> of Structures and Machines 20(4), 499–514.<br />

Prager, W. (1974), Introduction to structural optimization, Springer, CISM, Udine, Italy. 80<br />

p.<br />

Prendergast, P. J. and Contro, R., eds (2002), <strong>Mechanics</strong> of tissues and tissue implants,<br />

Vol. 37 of Meccanica, AIMETA, Kluwer.<br />

Press, W. H., Teukolsky, S. A., Vetterling, W. T. and Flannery, B. P. (1992), Numerical<br />

Recipes in Fortran 77: The Art of Scientific Computing, 2 edn, Cambridge University<br />

Press.<br />

Reddy, J. N. (1984), ‘A simple higher order theory for laminated composite plates’, J. Appl.<br />

Mech. 51, 745–752.<br />

Rodrigues, H., Jacobs, C., Guedes, J. M. and Bendsøe, M. P. (1999), Global and local<br />

material optimization models applied to anisotropic adaptation, in P. Pedersen and<br />

M. Bendsøe, eds, ‘Synthesis in Bio <strong>Solid</strong> <strong>Mechanics</strong>’, <strong>Solid</strong> <strong>Mechanics</strong> and its Applications,<br />

IUTAM, Kluwer, Lyngby, Denmark, pp. 209–220.<br />

Rovati, M. and Taliercio, A. (1991), <strong>Optimal</strong> orientation of the symmetry axes of orthotropic<br />

3-d materials, in H. A. Eshenauer, C. Mattheck and N. Olhoff, eds, ‘Engineering Optimization<br />

in Design Processes’, Lecture Notes in Engineering, Springer, Karlsruhe,<br />

Germany, pp. 127–134.<br />

Rozvany, G. N. I. (1989), Structural Design via <strong>Optimal</strong>ity Criteria, Kluwer, Dordrecht, The<br />

Netherlands.<br />

Save, M., Prager, W. and Sacchi, G. (1985), Structural optimization, optimality criteria,<br />

Vol. 1, Plenum Press.<br />

Savin, G. N. (1961), Stress Concentration around Holes, Pergamon Press. 430 p.<br />

Seregin, G. A. and Troitskii, V. A. (1982), ‘On the best position of elastic symmetry planes<br />

in an orthotropic body’, PMM 45, 139–142.<br />

Sergeyev, O. and Pedersen, P. (1996), ‘On design of joint positions for minimum mass 3d<br />

frames’, Structural Optimization 11, 95–101.<br />

Seyranian, A. P. and Pedersen, P. (1995), On two effects in fluid/structure interaction theory,<br />

in P. W. Bearman, ed., ‘6th Int. Conference on Fluid-Induced Vibration’, Balkema,<br />

London, UK, pp. 565–576.<br />

Seyranian, A. P., Lund, E. and Olhoff, N. (1994), ‘Multiple eigenvalues in structural optimization<br />

problems’, Structural Optimization 8(4), 207–227.<br />

Sheu, C. Y. and Prager, W. (1968), ‘Recent developments in optimal structural design’,<br />

Applied <strong>Mechanics</strong> Reviews 21(10), 985–992. 146 references.


294 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...<br />

Shin, C. S., Wang, C. M. and Song, P. S. (1996), ‘Fatigue damage repair: A comparison of<br />

some possible methods’, Int. J. of Fatigue 18, 535–546.<br />

Sigmund, O. (1994a), Design of Material Structures using Topology Optimization, DCAMM<br />

S69, <strong>Solid</strong> <strong>Mechanics</strong>, DTU - thesis for the Ph.D.<br />

Sigmund, O. (1994b), ‘Materials with prescribed constitutive parameters: An inverse homogenization<br />

problem’, Int. J. <strong>Solid</strong>s and Structures 31(17), 2313–2329.<br />

Sigmund, O. (1995), ‘Tayloring materials with prescribed elastic properties’, Mech. of Materials<br />

20(4), 351–368.<br />

Sigmund, O. (1999), On the optimality of bone microstructures, in P. Pedersen and<br />

M. Bendsøe, eds, ‘Synthesis in Bio <strong>Solid</strong> <strong>Mechanics</strong>’, <strong>Solid</strong> <strong>Mechanics</strong> and its Applications,<br />

IUTAM, Kluwer, Lyngby, Denmark, pp. 221–234.<br />

Sigmund, O. (2001), Topology Optimization Methods with Applications in Mechanism,<br />

MEMS and Material Design, <strong>Solid</strong> <strong>Mechanics</strong>, DTU - thesis for the dr. techn..<br />

Sigmund, O. and Torquato, S. (1996), ‘Composites with extremal thermal expansion coefficients’,<br />

Applied Physics Letters 69(21), 3203–3205.<br />

Sigmund, O. and Torquato, S. (1997), ‘Design of materials with extreme thermal expansion<br />

using a three-phase topology optimization method’, J. of the <strong>Mechanics</strong> and Physics of<br />

<strong>Solid</strong>s 45(6), 1037–1067.<br />

Soares, C. M. M., de Freitas, M. J. M., Araujo, A. L. and Pedersen, P. (1993), ‘Identification<br />

of material properties of composite plate specimens’, Composite Structures 25, 277–285.<br />

Sol, H. (1986), Identification of anisotropic plate rigidities using free vibration data, Thesis<br />

- Free University of Brussels, Belgium.<br />

Sol, H. and Oomens, C. W. J., eds (1997), Euromech 357 - Material Identicication using<br />

mixed numerical experimental methods, Euromech, Kluwer, Keikrade, The Netherlands.<br />

Stokholm, K. A. (1998), Formoptimering med ikke-lineær elasticitet (in Danish), M.Sc. Thesis,<br />

<strong>Solid</strong> <strong>Mechanics</strong>, DTU.<br />

Sundstrøm, B., ed. (1998), Handbok och Formelsamling i H˚allfasthetslaera (in Swedish),<br />

KTH, Stockholm. 398 pages.<br />

Svanberg, K. (1987), ‘The method of moving asymptotes - a new method for structural<br />

optimization’, Int. J. for Numerical Methods in Engineering 24, 359–373.<br />

Taylor, J. E. (2000), A formulation for optimal structural design with optimal materials, in<br />

G. I. N. Rozvany and N. Olhoff, eds, ‘Topology optimization of structures and composite<br />

continua’, NATO Science Series, Kluwer, Budapest, Hungary, pp. 49–59.<br />

Thomsen, J. J. (1990), ‘Modelling human tibia structural vibrations’, J. Biomechanics<br />

23(3), 215–228.<br />

Thorpe, M. F. and Jasiuk, I. (1992), ‘New results in the theory of elasticity for twodimensional<br />

composites’, Proc. R. Soc. Lond. A 438, 531–544.


Bibliography 295<br />

Tortorelli, D. A. and Pedersen, P. (1999), Similarities in bone re-modelling and damage<br />

evolution, in P. Pedersen and M. P. Bendsøe, eds, ‘Synthesis in Bio <strong>Solid</strong> <strong>Mechanics</strong>’,<br />

<strong>Solid</strong> <strong>Mechanics</strong> and its Applications, IUTAM, Kluwer, Lyngby, Denmark, pp. 387–400.<br />

Tsai, S. W. and Hahn, H. T. (1980), Introduction to Composite Materials, Technomic. 457<br />

pages.<br />

Tsai, S. W. and Pagano, N. J. (1968), Invariant properties of composite materials, in Tsai,<br />

Halpin and Pagano, eds, ‘Composite materials workshop’, Technomic, Westport Com.,<br />

pp. 233–253.<br />

Tvergaard, V. (1973), On the optimum shape of a fillet in a flat bar with restrictions.,<br />

in Sawczyk and Z. Mroz, eds, ‘IUTAM Symp. on Optimization in Structural Design’,<br />

Springer, Warsaw, Poland, pp. 181–195.<br />

Vigdergauz, S. B. (1986), ‘Effective elastic parameters of a plate with a regular system of<br />

equal-strength holes’, MTT 21(2), 165–169.<br />

Vigdergauz, S. B. (1994), ‘Three-dimensional grained composites of extreme thermal properties’,<br />

J. Mech. Phys. <strong>Solid</strong>s 42(5), 729–740.<br />

Vigdergauz, S. B. (1997), ‘Two-dimensional grained composites of minimum stress concentration’,<br />

Int. J. <strong>Solid</strong>s and Structures 34(6), 661–672.<br />

Vinson, J. R. and Sierakowski, R. L. (1989), The Behaviour of Structures Composed of<br />

Composite Materials, Nijhoff. 323 pages.<br />

Wasiutynski, Z. (1960), ‘On the congruency of the forming according to the minimum potential<br />

energy with that according to equal strength’, Bull. de l’Academie Polonaise des<br />

Sciences, Serie des Sciences Techniques 8(6), 259–268.<br />

Wasiutynski, Z. and Brandt, A. (1963), ‘The present state of knowledge in the field of optimum<br />

design of structures’, Applied <strong>Mechanics</strong> Reviews 16(5), 341–350. 234 references.<br />

Whitney, J. M. (1987), Structural Analysis of Laminated Anisotropic Plates, Technomic. 342<br />

pages.<br />

Wittrick, W. H. (1962), ‘Rates of change of eigenvalues, with reference to buckling and<br />

vibration problems’, J. Royal Aeronautical Soc. 66, 590–591.<br />

Xie, Y. M. and Steven, G. P. (1997), Evolutionary Structural Optimization, Springer.<br />

Zheng, Q. S. and Hwang, K. C. (1997), ‘Two-dimensional elastic compliances of materials<br />

with holes and microcracks’, Proc. R. Soc. Lond. A 453, 353–364.<br />

Ziegler, H. (1968), Principles of Structural Stability, Blaisdell. 150 pages.


296 Copyright c○Pauli Pedersen: <strong>Optimal</strong> designs ...


Index<br />

1D-design description, 125<br />

2D-bulk modulus, 216<br />

2D-design description, 126<br />

2nd order polynomial, 257<br />

3D, 246<br />

3D analysis, 186<br />

3D-bulk modulus, 216<br />

3D-isotropic modulus relations, 217<br />

A thesis, 187<br />

Absolute limits, 285<br />

Absolute side limits, 21<br />

Active constraints or inactive ?, 22<br />

active set strategy, 100<br />

Active volume, 85<br />

Adaptation, 187<br />

Additional eigenvalue problem, 271<br />

Adjoint problem, 264<br />

Advanced FEM-analysis, 25<br />

Aligned larger stress/strain, 179<br />

Aligned orientations, 187<br />

Aligned stresses and strains, 171<br />

alignment<br />

principal material direction, 254<br />

principal strain direction, 254<br />

principal stress direction, 254<br />

Alignments, 254<br />

Also best strength, 232<br />

Also non-linear, 24<br />

Alternative design variables, 83<br />

Alternative materials, 206<br />

Always sensitivities, 27<br />

Analog in 3D, 214<br />

analytical optimal design<br />

explicit, 81<br />

implicit, 81<br />

analytical sensitivity analysis, 125<br />

Approximately optimal hole, 72<br />

artificial<br />

297<br />

cost, 284<br />

variable, 283<br />

Artificial variables, 283<br />

Assumption for cost of supports, 252<br />

Assumptions, 246<br />

Available functions, 238<br />

Axial load or bending, 57<br />

axial stress flow, 57<br />

Axisymmetric modelling, 138<br />

bar<br />

force, 245<br />

long, medium and short, 59<br />

short column, 250<br />

slender column, 250<br />

tensile, 249<br />

zero string, 250<br />

Bars and joints, 31<br />

Base cell homogenization, 209<br />

Basic assumption, 225, 234<br />

basis<br />

change, 275<br />

Basis separation, 247<br />

Basis solution, 247, 274<br />

beam<br />

active size constraint, 85<br />

approximate gradient, 90<br />

Bernoulli-Euler theory, 90<br />

eigenfrequency, 85<br />

exact gradient, 90<br />

fixed-fixed, 93<br />

long, medium and short, 65<br />

minimum compliance, 82<br />

second eigenfrequency, 90<br />

simple beam theory, 86<br />

simply supported, 93<br />

slenderness, 90<br />

statically indetermined, 85<br />

stress constraint, 93


298 Index<br />

stress energy, 82<br />

Timoshenko theory, 89<br />

beam-bar<br />

long, medium and short, 62<br />

beam/bridge truss<br />

24 joints, 41<br />

cantilever part, 42<br />

hanging part, 42<br />

parameter studies, 42<br />

possible height, 41<br />

possible support, 41<br />

Beams and frames, 15<br />

Beck’s column, 270<br />

bending stiffness<br />

contracted vector, 207<br />

Bending stiffnesses, 207<br />

Bernoulli-Euler beam theory, 98<br />

Best vertex solution, 275<br />

Better force flow, 46<br />

black and white<br />

design, 162<br />

penalize, 162<br />

bone<br />

non-isotropic, 187<br />

a thesis, 187<br />

adaptation, 187<br />

aligned orientations, 187<br />

analysis, 186<br />

constitutive constraints, 194<br />

constitutive increments, 194<br />

constitutive vector, 189<br />

energy density, 186, 189<br />

formulation in time, 186<br />

frame model, 190<br />

homogenization, 190<br />

identification, 186<br />

implants, 187<br />

incremental formulation, 192<br />

inverse homogenization, 191<br />

load estimation, 189<br />

material, 185<br />

memory function, 188<br />

micro-structure optimization, 190<br />

note, 196<br />

optimal re-modelling, 192<br />

optimality criterion, 192<br />

optimization formulation, 191<br />

phenomenological re-modelling, 186<br />

proximal femur, 188<br />

rate function example, 194<br />

rate of constitutive norm, 193<br />

sensitivity of energy, 191<br />

shape and size, 186<br />

special case, 192<br />

structure, 185<br />

bound formulation, 130<br />

boundary<br />

displacement, 159<br />

load, 159<br />

post-process smoothing, 169<br />

boundary condition, 261<br />

boundary shapes, 129<br />

Bounds on redesign, 281<br />

bridge<br />

nine problems, 65<br />

bridge truss<br />

16 joints, 39, 46<br />

positions for joints, 46<br />

Buckling load, 118<br />

bulk modulus<br />

Hashin-Shtrikman bound, 175<br />

macroscopic, 175<br />

cantilever<br />

beam, 83<br />

elementary load cases, 83<br />

non-rectangular, 75<br />

statically determined, 83<br />

Cantilever + hanging part, 42<br />

cantilever truss<br />

17 joints, 38<br />

7 to 12 joints, 46<br />

9 joints, 33<br />

Cases of no influence, 178<br />

cavity<br />

3D inclusion, 151<br />

ellipsoid, 151<br />

Chain rule of differentiation, 93<br />

change basis, 275<br />

Change in equilibrium, 275<br />

Changed designs with forced displacements, 180<br />

Changed with shape design, 182<br />

Changing load direction, 35<br />

Changing type of cross-section, 36<br />

checkerboard, 162<br />

circular cross-section, 31<br />

Circular hole, 145<br />

circulatory<br />

force, 263<br />

Citation from Frederiksen, 204


Index 299<br />

Classes of optimization procedures, 27<br />

Classical problem, 87<br />

classification<br />

behaviour, 263<br />

dynamic system, 262<br />

material, 112<br />

names and concepts, 17<br />

coefficient<br />

cost, 279<br />

Collected constraints, 284<br />

Collected results, 218, 224, 228, 240, 244, 252,<br />

260, 271, 278, 285<br />

Columns at beam boundaries, 93<br />

Combined bending and axial forces, 62<br />

Combined shape design, 44<br />

complementary energy, 225<br />

complex<br />

eigenvalue, 264<br />

functional, 261<br />

Complex functionals, 265<br />

Compliance, 21<br />

compliance<br />

matrix, 219<br />

isotropic, 221<br />

non-dimensional, 220<br />

non-orthotropic, 220<br />

orthotropic, 220<br />

measured by umean, 169<br />

Compliance power law description, 220<br />

Complicated laboratory tests, 198<br />

composite, 107<br />

compression bar, 31<br />

Compressive - stability, 31<br />

Concave cost, 249<br />

concave function, 273<br />

Concave transition, 251<br />

Concentration factor of 2, 146<br />

condition<br />

optimality, 230, 249<br />

proportional gradient, 172<br />

stationarity, 255<br />

Condition for proportionality, 222<br />

Conservative systems, 262<br />

Constant energy density, 131, 234<br />

constitutive<br />

coupling component, 215<br />

eigenvalues, 213<br />

individual parameter, 241<br />

matrix, 211, 212, 223<br />

non-dimensional, 242<br />

normal component, 215<br />

parameter, 257<br />

alternative description, 211<br />

positive semi-definite matrix, 241<br />

rotational transformation, 212<br />

shear component, 215<br />

tensor, 211<br />

vector, 212<br />

Constitutive constraints, 194<br />

Constitutive eigenvalues, 213<br />

Constitutive Frobenius (length), 212<br />

Constitutive increments, 194<br />

Constitutive matrix, 109, 171<br />

constitutive matrix<br />

practical parameters, 110<br />

constitutive parameter<br />

transformation, 110<br />

Constitutive trace, 212<br />

Constitutive vector, 109, 189<br />

constraint<br />

displacement, 98<br />

eigenfrequency, 87<br />

equation, 274<br />

geometrical, 159<br />

manufacturing, 127<br />

multi-physic, 160, 184<br />

parametrization, 159<br />

several eigenfrequencies, 87<br />

side, 200<br />

strength, 222<br />

type, 282<br />

vector, 280<br />

Constraint on definiteness, 200<br />

Constraint types, 282<br />

Constraints, 20<br />

Continuous - discrete, 19<br />

Continuous problem, 269<br />

contracted<br />

constitutive matrix, 214<br />

second order strains, 214<br />

strain matrix, 214<br />

stress matrix, 214<br />

control of design process, 22<br />

convergence<br />

design variable, 259<br />

objective, 259<br />

Convergence tests, 20<br />

Converted to non-constrained, 230<br />

Convex, 23<br />

Convex feasible space, 249


300 Index<br />

corresponding displacement, 225<br />

cost<br />

change in total, 248<br />

coefficient, 279<br />

coefficient vector, 274<br />

large coefficient, 281<br />

support, 252<br />

total, 247<br />

Cost functions, 247<br />

Cost non-increasing, 248<br />

crack tip<br />

allowable domain, 157<br />

external load, 154<br />

hole size, 154<br />

material non-isotropy, 155<br />

power law non-linearity, 155<br />

stress release, 151<br />

cross-section<br />

constant, 82<br />

hollow circular, 36<br />

moment of inertia, 82<br />

rectangular, 83<br />

solid, 36<br />

cross-sectional<br />

parameter, 250<br />

curvature<br />

discontinuous, 135<br />

Cylindrical pipe data, 33<br />

damage<br />

evolution, 194<br />

non-bone materials, 194<br />

damping<br />

destabilizing, 261<br />

DDLPRS<br />

IMSL routine, 280<br />

Deciding neighbour solution, 275<br />

Definition of √ 2 contraction, 212<br />

Deformation mode parameter, 117<br />

degrees of freedom, 57<br />

Derivation from elastic potential, 223<br />

Derivative of cost, 248<br />

design<br />

advanced material, 159<br />

code, 251<br />

global parametrization, 123<br />

iteration, 280<br />

local parameter, 226<br />

optimal strength, 124<br />

thickness and orientation, 123<br />

thickness distribution only, 122<br />

design constraint<br />

implicit, 21<br />

life time, 21<br />

list, 20<br />

satisfied, 22<br />

stability, 21<br />

stiffness, 21<br />

strength, 21<br />

vibration, 21<br />

violated, 22<br />

design cycling, 78<br />

design domain<br />

graphical illustration, 115<br />

lamination parameters, 113<br />

Design for stiffness and strength, 235<br />

Design functions, 127<br />

Design independence, 122<br />

Design independent loads, 233<br />

design influence<br />

finite element modelling, 163<br />

initial design, 163<br />

load modelling, 163<br />

lower bound, 163<br />

move-limit, 163<br />

neglected modelling, 163<br />

optimization procedure, 163<br />

penalization approach, 163<br />

smoothing (filtering), 163<br />

total density, 163<br />

design limit<br />

absolute side, 21<br />

move-limit, 22<br />

design objective<br />

convergence, 20<br />

cost, merit, goal, 19<br />

existence of solution, 20<br />

local and global, 20<br />

minimum and maximum, 19<br />

optimized not optimal, 20<br />

Design optimization, 229<br />

Design problem, 33, 237, 242<br />

design space<br />

boundary, 23<br />

convex, 23<br />

direction, 23<br />

exterior, 23<br />

geometrical interpretation, 23<br />

interior, 23<br />

list of concepts, 23


Index 301<br />

step size, 23<br />

design variable<br />

continuous, 19<br />

convergence, 20<br />

discrete, 19<br />

distributed, 19<br />

hierarchical, 19<br />

linking, 29<br />

local, global, 28<br />

parametrization, 19<br />

shape, 18<br />

size, 18<br />

topology, 18<br />

vector, 280<br />

design-independent<br />

external load, 226<br />

<strong>Designs</strong> independent of Poisson’s ratio, 174<br />

Detail of proof, 234<br />

Deviatoric stresses, 221<br />

Different assumptions, 33<br />

Different load cases, 175<br />

Different objectives, 129, 254<br />

Different objectives, but the same design, 159<br />

Different sensitivities, 264<br />

Direct conclusions, 242<br />

direction<br />

cosine, 246<br />

principal strain/stress, 243<br />

Discontinuous curvature, 135<br />

Discussion of results, 169<br />

displacement constraints, 46<br />

displacement pattern, 117<br />

Displacements and stiffness, 21<br />

dissipative<br />

force, 263<br />

mutual energy, 266<br />

Distinct eigenvalues, 270<br />

Distributed parameter, 19<br />

Divergence - quasi-static, 263<br />

domain<br />

non-shape, 235<br />

shape, 235<br />

dome frame, 100<br />

Double constraints separated, 283<br />

Down to 15 % concentration, 135<br />

Down to 8 % concentration, 138<br />

dyadic product, 214<br />

dynamic behaviour, 261<br />

Early references, 131, 245<br />

Eccentricity, 92<br />

effective<br />

strain, 219, 242<br />

stress, 219<br />

Eigenfrequencies, 21, 118<br />

eigenfrequency<br />

calculated, 197, 199<br />

measured, 197<br />

thick plate formulation, 198<br />

eigenfunction, 269<br />

Eigenmodes identification, 270<br />

eigenvalue<br />

Sturm sequence check, 86<br />

bimodal case<br />

complex, 264<br />

constitutive, 213<br />

continuous problem, 269<br />

double with one eigenvector, 271<br />

inverse iteration, 86<br />

multiple, 270<br />

only one non-zero, 161<br />

shift, 86<br />

sub-space iteration, 86<br />

Eigenvalue sensitivities, 203<br />

eigenvector<br />

left, adjoint, 264<br />

normalization, 203<br />

right, physical, 264<br />

variation, 265<br />

Eigenvector description, 214<br />

eigenvectors<br />

mutual orthogonal, 270<br />

elastic energy<br />

sensitivity analysis, 115<br />

total, 55<br />

elastic mutual energy, 266<br />

elasticity<br />

non-isotropic, 160, 219<br />

non-linear, 219<br />

powers n or p, 219<br />

orthotropic, 219<br />

power law non-linear, 160, 233, 241<br />

Element energies, 254<br />

Element level FEM, 267<br />

Elementary strain cases, 209, 214<br />

elementary strain state, 214<br />

Eliminated variable from basis, 275<br />

ellipsoidal shape, 151<br />

Elliptical are not optimal, 146<br />

Elongated better designs, 157


302 Index<br />

Email, 11<br />

energy concentration factor, 57<br />

Energy densities, 223<br />

Energy density, 186, 189<br />

energy density<br />

elastic, 55<br />

initial mean, 57<br />

maximum, 57<br />

mean, 57<br />

strain, 179, 182, 213, 223<br />

stress, 179, 213, 223<br />

total, 213<br />

uniform, 55<br />

Energy density RELATION, 223<br />

Energy equilibrium, 225<br />

Energy principles, 15<br />

energy theorems of mechanics, 24<br />

engineering parameters, 112<br />

enlarged<br />

flexibility, 80<br />

stiffness, 80<br />

Equal cases for small holes, 175<br />

Equality by slack variables, 282<br />

equilibrium<br />

energy, 225<br />

Equilibrium at joints, 246<br />

Essential assumptions, 146<br />

Estimated multiplier, 239<br />

estimation, 197<br />

applied load, 197<br />

deep drawing, 197<br />

hardness test, 197<br />

material parameter, 197<br />

orthotropic material, 197<br />

stiffness model, 197<br />

estimator<br />

weighted least-square, 199<br />

Exact gradients, 89<br />

Example, 257<br />

example<br />

12.5% total relative density, 164<br />

25% total relative density, 164<br />

50% total relative density, 164<br />

80% total relative density, 164<br />

analytical solution, 170<br />

Beck’s column, 269<br />

boundary displacement, 175<br />

boundary stress, 175<br />

cantilever beam - Galerkin, 268<br />

Hauger’s column, 269<br />

identification - estimation<br />

carbon-epoxy, 204<br />

glass-epoxy, 203<br />

orthotropic aluminum, 203<br />

temperature dependence, 204<br />

Leipholz’s column, 269<br />

material with a single hole, 172<br />

mixed problem, 175<br />

optimal constitutive matrix, 160<br />

quarter cell, 163<br />

shape influence from boundary condition,<br />

174<br />

shape influence from material non-linearity,<br />

180<br />

shape influence from Poisson’s ratio, 178<br />

shape influence from volume constraint, 173<br />

two parameter cell model, 170<br />

ultimate optimal field, 161<br />

with penalizing, 163<br />

without penalizing, 163<br />

Example of rate function, 194<br />

Example values, 259<br />

Examples, 18<br />

Examples in chapter 4, 161<br />

Excitation source, 201<br />

Existence, 20<br />

Existence of solution, 259<br />

expansion<br />

adjoint eigenfunction, 268<br />

physical eigenfunction, 268<br />

expansion function<br />

global, 267<br />

local, 267<br />

Expansion functions, 267<br />

Experimental - numerical, 197<br />

Experimental errors, 206<br />

experimental setup<br />

excitation source, 200<br />

frequency analyzer, 200<br />

identification program, 200<br />

power amplifier, 200<br />

response detector<br />

accelerometer, 200<br />

laser vibrometer, 200<br />

microphone, 200<br />

test specimen, 200<br />

transient recorder, 200<br />

Explicit optimal beam design, 84<br />

Extended problems, 87<br />

Extensions: stability, supports, 245


Index 303<br />

external damping, 268<br />

external load<br />

design-independent, 226<br />

work, 55<br />

external mutual energy , 266<br />

external potential, 225<br />

extremum<br />

local, 253<br />

multiplicity, 255<br />

Extremum relations, 226<br />

fatigue damage repair, 151<br />

FE model, 132<br />

Feasibility, 23<br />

Feasible basis solution, 276<br />

Feasible solution, 274<br />

feasible space<br />

convex, 249<br />

fillet<br />

2D in tension, 131<br />

3D bending, 138<br />

3D tension, 138, 141<br />

3D torsion, 138, 141<br />

axisymmetric, 138<br />

bending stress fields, 141<br />

circular connection, 132<br />

geometrical constraint, 131<br />

length, 135<br />

multi-parameter, 135<br />

super-circular connection, 135<br />

tension stress fields, 141<br />

Finite cost increment, 249<br />

Finite degrees of freedom, 261<br />

Finite design regions, 81<br />

finite element, 146, 238, 266<br />

finite element modelling, 55<br />

Finite element models, 57<br />

First basis solution, 284<br />

first ply failure, 124<br />

fixed strain, 226, 254<br />

fixed stress field, 235<br />

flutter<br />

frequency, 266<br />

load, 263<br />

Flutter - dynamic, 263<br />

Flutter frequency sensitivity, 266<br />

Flutter load sensitivity, 266<br />

Focus on stress energy, 82<br />

follower force, 268<br />

For model by the FEM, 238<br />

force<br />

bar, 245<br />

circulatory, 263<br />

compressive, 245<br />

dissipative, 263<br />

equilibrium, 246<br />

follower, 268<br />

gyroscopic, 262<br />

non-negative, 245<br />

non-potential, 263<br />

tensile, 245<br />

Formulation in time, 186<br />

Fortran subroutine, 273<br />

foundation problem, 75<br />

Fracture and creep, 21<br />

frame<br />

3D, 98<br />

dome, 100<br />

linking, 98<br />

mobile crane, 103<br />

optimal joint position, 98<br />

portal, 81, 92<br />

self-weight, 98<br />

shape optimization, 82<br />

side constraint, 99<br />

total mass, 99<br />

frame from beams, 81<br />

Frame model with cubic symmetry, 190<br />

Free material, 241<br />

free material design, 160<br />

frequency measure, 264<br />

Frobenius norm, 207, 241<br />

full pivoting, 273<br />

Fully stressed, 22, 34<br />

fully stressed, 31, 245<br />

fully stressed design, 22<br />

functional<br />

complex, 261<br />

mutual energy, 261<br />

Further designs, 98<br />

Further slack variables, 283<br />

Further tools, 202<br />

gain factor, 57<br />

Galerkin<br />

method, 267<br />

Galerkin discretization, 268<br />

gamma-function, 237<br />

General data for examples, 32<br />

General incompressible, 221


304 Index<br />

General knowledge, 14, 129, 232<br />

General non-isotropic, 220<br />

General optimization, 229<br />

General results, 219<br />

geometrical constraint, 235<br />

Given volume, 56<br />

global<br />

maximum, 253<br />

minimum, 253<br />

global design parameters, 125<br />

curve, 125<br />

orthogonal functions, 126<br />

surface, 125<br />

Global optimal, 274<br />

Glossary, 14, 17<br />

Good experience, 235<br />

Gradient function, 268<br />

gradient function, 264<br />

Gradient information non-sufficient, 253<br />

gray design<br />

intermediate density, 169<br />

Greater than by less than, 280<br />

gyroscopic<br />

force, 262<br />

Half a model by symmetry, 42<br />

Hashin-Shtrikman bound<br />

bulk modulus, 175<br />

Hauger’s column, 270<br />

heuristic approach<br />

successive iteration, 239<br />

Heuristic methods, 259<br />

heuristic procedure, 161<br />

Higher order plate theory, 202<br />

hole<br />

approximately optimal, 72<br />

biaxial load, 144<br />

biaxial load , 72<br />

classic solution, 144<br />

limited domain, 144<br />

orthotropic material, 148<br />

relative hole size, 146<br />

strain localization, 149<br />

stress releasing, 149<br />

Hole designs with resulting bulk modulus, 174<br />

hollow circular cross-section, 36<br />

homogeneous<br />

energy relation, 231<br />

mass relation, 231<br />

system, 262<br />

Homogeneous mass (volume) dependence, 231<br />

Homogeneous non-homogeneous, 75<br />

homogenization<br />

constitutive parameter, 209<br />

inverse, 209<br />

material, 209<br />

prescribed material parameter, 209<br />

Homogenization analysis, 190<br />

hydrostatic pressure<br />

deviatoric stress, 221<br />

incompressibility, 221<br />

Identification, 13, 186<br />

identification, 197<br />

formulation, 199<br />

micro-mechanical structure, 210<br />

identification - estimation<br />

bone material, 206<br />

ceramic, 206<br />

environmental influence, 206<br />

errors<br />

experimental, 204<br />

numerical modelling, 204<br />

physical modelling, 204<br />

plate defects, 204<br />

sandwich combination, 206<br />

identification approach<br />

eigenvalue analysis, 198<br />

error estimation, 198<br />

experiment, 198<br />

optimization, 198<br />

sensitivity analysis, 198<br />

Illustrative example, 153<br />

imaginary part, 265<br />

Implants, 187<br />

Implicit behaviour, 21<br />

Importance of supports, 42<br />

Important references, 81<br />

Important RESULT, 226<br />

IMSL<br />

move-limit, 282<br />

routine, 282<br />

IMSL routine DDLPRS, 280<br />

In-plane and out-of-plane, 113<br />

incompressibility<br />

hydrostatic pressure, 221<br />

orthotropic, 221<br />

Incremental formulation, 192<br />

Individual bar design, 245<br />

inequality


Index 305<br />

greater than, 280<br />

less than, 280<br />

inequality to equality, 281<br />

influence<br />

material non-linearity, 159<br />

material Poisson’s ratio, 159<br />

maximum thickness, 78<br />

non-isotropic, 79<br />

non-linear, 79<br />

Poisson’s ratio, 78<br />

Influence from finite size, 146<br />

Influence of gradients, 90<br />

Influence of number of joints, 46<br />

Influence of stiffness, 132<br />

Influence on designs, 178<br />

Influence studies, 172<br />

Initial designs, 94<br />

instability<br />

divergent, 263<br />

dynamic, 263<br />

flutter, 263<br />

initiation, 265<br />

static, 263<br />

intermediate density<br />

gray design, 169<br />

internal damping<br />

Kelvin-Voigt, 268<br />

invariant<br />

determinant, 212<br />

matrix, 241<br />

matrix length, 241<br />

trace, 212<br />

vector length, 212<br />

inverse<br />

homogenization, 209<br />

Inverse homogenization, 13, 191<br />

inverse iteration, 86<br />

Inverse material problems, 209<br />

inverse problem, 197<br />

bending stiffness, 207<br />

laminate, 197, 207<br />

material, 197<br />

Inverse problems, 13, 24<br />

inverse transformation<br />

orthonormal, 213<br />

Isotropic, 221<br />

Isotropic 2 parameters, 113<br />

iteration<br />

monotonic convergence, 78<br />

recursive formula, 75<br />

relaxation, 78<br />

Iteration strategy, 208<br />

iterative improvement, 284<br />

joints in trusses, 31<br />

joints positions<br />

optimization, 44<br />

Kelvin-Voigt internal damping, 268<br />

kinematically admissible, 226<br />

kinetic mutual energy, 266<br />

Kinetic, dissipative, elastic and<br />

external potentials, 266<br />

knee problem, 72<br />

Known moment distribution, 83<br />

Kronecker delta, 214<br />

laboratory tests<br />

eigenfrequency, 198<br />

strain gauge, 198<br />

Lagrange<br />

function, 230<br />

multiplier, 89, 230, 239<br />

lamina, 107<br />

laminate<br />

analysis, 107<br />

balanced, 116<br />

bending stiffness, 199<br />

identification, 124<br />

material stiffness, 199<br />

membrane stiffness, 199<br />

ply layer lamina, 107<br />

rotational transformation, 108<br />

specially orthotropic, 116<br />

stiffness, 113<br />

Laminate design, 207<br />

Laminate inverse problems, 197<br />

Laminate parametrization for design, 123<br />

Lamination parameters, 113<br />

Large improvements, 120<br />

layer, 107<br />

Leipholz’s column, 270<br />

level of non-isotropy, 112<br />

limit<br />

lower bound, 56<br />

upper bound, 57<br />

Linear dependence, 250, 251<br />

Linear functions, 113<br />

linear programming, 245<br />

sequential, i.e. SLP, 279


306 Index<br />

standard form, 274, 279<br />

linear programming problem, 32<br />

Linear systems, 262<br />

Linking and no-linking, 104<br />

linking design variables, 98<br />

linking of design variables, 29<br />

List of practical problems, 162<br />

load<br />

condition, 56<br />

critical, 265<br />

design dependent, 21<br />

design independent, 21<br />

distribution, 264<br />

flutter, 265<br />

level, 264<br />

one case, 55<br />

Load cases, 21<br />

Load direction, 144<br />

Load estimation, 189<br />

Loads and boundary conditions, 21<br />

local<br />

design parameter, 226<br />

determination of sensitivity, 226<br />

Local - Global, 20<br />

Local design parameter, 227<br />

Local/global parametrization, 28<br />

Localized determined sensitivities, 241<br />

Localized energy change, 233<br />

Localized volume change, 232<br />

LP in SLP, 285<br />

LP modifications, 273<br />

Main contents of chapter, 81<br />

Major importance, 135<br />

Manufacturing constraints, 127<br />

Many local extremum, 253<br />

Many possibilities, 211<br />

mass density, 82<br />

material<br />

base cell, 209<br />

constitutive component, 160<br />

effective property, 209<br />

homogenization, 209<br />

identification, 209<br />

micro-structure, 209<br />

non-isotropic, 253<br />

orientational result, 171<br />

orthotropic, 253<br />

Material and structure, 185<br />

Material as a structure, 159<br />

material design, 107<br />

free material, 241<br />

negative Poisson’s ratio, 210<br />

solution priority, 210<br />

thermoelastic, 210<br />

ultimate design, 241<br />

Material models, 13<br />

Material parameter, 178<br />

Material, yes or no ?, 162<br />

mathematical programming, 81, 124, 274<br />

Matrices, 211<br />

matrix<br />

coefficient, 274<br />

compliance, 219<br />

inverse, 223<br />

constitutive, 211<br />

non-dimensional, 223<br />

damping, 262<br />

direction cosine’s, 246<br />

element level, 267<br />

Frobenius norm, 212<br />

Hill strength, 222<br />

invariant, 241<br />

inverse, 274<br />

length, 212<br />

mass, 262<br />

orthotropic constitutive, 170<br />

positive definite, 212<br />

projection, 221<br />

regular, 274<br />

sensitivity, 280<br />

skew-symmetric, 262<br />

stiffness, 262<br />

symmetric and positive definite, 262<br />

system, 264<br />

system level, 267<br />

total coefficient, 276<br />

total equation, 284<br />

total stiffness, 180<br />

trace norm, 212<br />

von Mises strength, 221<br />

Matrix of Hill, 222<br />

Matrix of von Mises, 222<br />

Maximize by minimize, 280<br />

maximum<br />

stiffness, 253<br />

strain energy density, 241<br />

Maximum displacement, 118<br />

mechanism in truss, 251<br />

Medium slender columns, 96


Index 307<br />

Member classes, 250<br />

Memory function, 188<br />

method<br />

Galerkin, 267<br />

heuristic, 259<br />

Ritz, 267<br />

subspace iteration, 202<br />

weighted residual, 267<br />

micro-structure<br />

finite element analysis, 209<br />

homogeneous, 209<br />

inhomogeneous, 209<br />

Micro-structure optimization, 190<br />

minimize the total mass, 33<br />

minimum<br />

compliance, 234, 253<br />

cost, 245<br />

energy concentration, 239<br />

gradient, 275<br />

maximum strain energy density, 234<br />

size, 245<br />

total strain energy, 241<br />

Minimum - Maximum, 19<br />

Minimum of maximum, 130<br />

Minimum stress concentration, 146<br />

minimum-maximum formulation, 130<br />

mode parameter, 117, 118<br />

model/plane strain, 179<br />

model/plane stress, 179<br />

Modelling accuracy, 78<br />

Modelling errors, 206<br />

Modelling issues, 130<br />

Moderately non-isotropic, 204<br />

Modification functions, 239<br />

Modified optimality criterion, 162<br />

Modulus power law description, 223<br />

monotonous behaviour, 248<br />

Mostly numerical, 24<br />

move-limit, 98<br />

side constraint, 99<br />

smaller and smaller, 281<br />

Move-limits, 22, 27, 200<br />

Moving in design space, 23<br />

moving load, 48<br />

multi-purpose plane truss, 46<br />

multi-purpose space truss, 50<br />

Multi-purpose trusses, 31<br />

multiple<br />

loads, 124<br />

plies, 124<br />

strength constraints, 124<br />

multiple eigenvalue<br />

sensitivity analysis, 270<br />

Multiple eigenvalues, 24<br />

Multiple load case, 101<br />

Multiple load cases, 91, 124<br />

multiple load cases, 46<br />

Multiple plies, 124<br />

Multiple solutions, 210<br />

Multiple strength constraints, 124<br />

mutual energy<br />

dissipative, 266<br />

elastic, 266<br />

external, 266<br />

kinetic, 266<br />

specific, 264<br />

variation, 265<br />

Mutual potentials, 261<br />

necessary condition, 229<br />

a single constraint, 230<br />

non-constrained, 229<br />

positive definite, 242<br />

proportional gradient, 243<br />

proportionality, 230<br />

necessary, sufficient, 24<br />

Neighbouring redesign, 130<br />

New design problem, 243<br />

Nine problems, 65<br />

No cost assumed, 251<br />

No energy concentration, 120<br />

nodal<br />

displacement, 180<br />

load, 180<br />

Non-bone materials, 194<br />

Non-changed loads, 226<br />

non-conservative<br />

load, 261<br />

Non-conservative systems, 263<br />

Non-dimensional parameters, 200<br />

non-increasing function, 248<br />

Non-intuitive behaviour, 261<br />

Non-isotropic, 187<br />

non-isotropic<br />

material, 253<br />

Non-isotropic parameters, 255<br />

non-isotropy<br />

degree, 253<br />

level, 200<br />

relative shear stiffness, 200


308 Index<br />

Non-isotropy influences, 155<br />

Non-linear - less influence, 157<br />

Non-linear elasticity, 122<br />

Non-linear, non-isotropic, 219<br />

Non-negative by move-limits, 282<br />

non-negative unknown<br />

in basis, 274<br />

out of basis, 274<br />

non-potential<br />

force, 263<br />

Non-rectangular cantilever, 75<br />

non-selfadjoint<br />

problem, 261<br />

Non-slender columns, 97<br />

Non-trivial optimal orientation, 118<br />

norm<br />

Frobenius, 160, 242<br />

trace, 160<br />

Normalized mass matrix, 262<br />

Not fully stressed, 48<br />

Not optimal design, 197<br />

Not proper references, 14<br />

Not statically determined, 46<br />

Notation, 219<br />

Note, 196<br />

Numerical errors, 206<br />

numerical procedure, 75<br />

Objective, 125, 153, 199, 207<br />

objective, 161, 273, 280<br />

concave, 275<br />

concave function, 249<br />

maximum bulk modulus, 174<br />

non-linear, 275<br />

stiffest design, 159<br />

strongest design, 159<br />

Objective Φmn, 117<br />

Objective derivative, 275<br />

Only axial forces, 59<br />

Only bending forces, 65<br />

Only deterministic, 24<br />

Only four stationary solutions, 116<br />

Only linear elasticity, 215<br />

Only minimum, 130<br />

Only single load case, 55<br />

optimal<br />

orientation, 253<br />

orientation in plate, 118<br />

<strong>Optimal</strong> 2D-modulus matrix, 161<br />

<strong>Optimal</strong> 3D-modulus matrix, 160<br />

optimal design<br />

books, 13<br />

<strong>Optimal</strong> designs, 13<br />

<strong>Optimal</strong> energy ratio, 84<br />

optimal material<br />

aligned, 242<br />

degenerate, 244<br />

no shear stiffness, 242<br />

non-zero eigenvalue, 243<br />

orthotropic, 242<br />

<strong>Optimal</strong> modulus matrix, 243<br />

<strong>Optimal</strong> re-modelling, 192<br />

<strong>Optimal</strong> redesign, 25<br />

optimal redesign<br />

analysis, 25<br />

sensitivity analysis, 25<br />

<strong>Optimal</strong> shape independent of load and size, 155<br />

<strong>Optimal</strong> stress release, 151<br />

optimal topology, 33<br />

optimal truss<br />

statically determinate, 245<br />

<strong>Optimal</strong>ity condition, 238, 243<br />

optimality condition, 230, 249<br />

<strong>Optimal</strong>ity criterion, 82, 192<br />

optimality criterion, 254<br />

density distribution, 161<br />

uniform energy density, 161<br />

optimality criterion test, 89<br />

optimization<br />

material orientation, 230<br />

non-linear by LP, 285<br />

Optimization formulation, 191<br />

optimization methods<br />

list of formulations, 24<br />

list of procedures, 27<br />

Optimization problem, 208<br />

optimize<br />

stiffness, 229<br />

strength, 229<br />

Optimize bulk modulus, 172<br />

Optimized design, 20<br />

orientation<br />

in strain parameter, 256<br />

in stress parameter, 256<br />

optimal, 253<br />

orientational field, 254<br />

Orthogonal [R] matrix, 109<br />

Orthogonal [T ] matrix, 109<br />

orthogonal matrix<br />

fourth order tensor, 109


Index 309<br />

second order tensor, 109<br />

Orthonormal rotational transformations, 212<br />

Orthotropic, 220<br />

orthotropic<br />

condition, 111<br />

directions, 111<br />

material, 253<br />

Orthotropic analytical solutions, 116<br />

Orthotropic engineering 4 parameters, 112<br />

Orthotropic material, 170<br />

Orthotropic materials, 111<br />

Other classifications, 18<br />

Overall parameters, 132<br />

Parametrization, 19<br />

parametrization<br />

boundary shape, 235<br />

super-ellipse, 235<br />

penalize<br />

stationarity condition, 162<br />

Penalized and not-penalized, 164<br />

Phenomenological re-modelling, 186<br />

plate<br />

buckling load, 118<br />

eigenfrequencies, 117<br />

eigenfrequency analysis, 202<br />

higher order shear deformation, 202<br />

mode parameter, 117<br />

Rayleigh-Ritz method, 202<br />

sensitivity analysis, 202<br />

simply supported, 116<br />

Plate errors, 206<br />

Plies - layers - laminas - prepregs, 107<br />

ply, 107<br />

Points or domains, 13<br />

Poisson’s ratio<br />

global redistribution, 178<br />

independence, 174<br />

local change, 178<br />

major, 200<br />

scaling factor, 178<br />

zero, 161, 244<br />

portal frame, 92<br />

Positive definite, 242, 257<br />

positive definite<br />

matrix, 212<br />

post-process<br />

boundary smoothing, 169<br />

potential<br />

derivative, 226<br />

elastic complementary energy, 225<br />

elastic strain energy, 225<br />

elastic stress energy, 225<br />

external, 225<br />

stationary total, 226<br />

total, 226<br />

Potential definitions, 225<br />

Potential relations, 226<br />

Practical conclusions, 118<br />

Practical definitions, 110<br />

practical problem<br />

checkerboard, 162<br />

local optimal, 162<br />

mesh dependence, 162<br />

non-unique solution, 162<br />

slow convergence, 162<br />

stopping criterion dependence, 162<br />

violate model, 162<br />

prepreg, 107<br />

Principal directions, 257<br />

Principal strains only, 242<br />

Problem of analysis, 85<br />

procedure<br />

simplex, 273<br />

Procedure and examples in chapter 4, 161<br />

Projection matrix, 221<br />

proportional change, 182<br />

proportional gradient, 162<br />

Proportional gradients, 230<br />

proportionality relation<br />

elastic energy, 225<br />

energy density, 223<br />

Proximal femur, 188<br />

pseudo-load, 93<br />

psi-function, 238<br />

Quarter cell example, 163<br />

Rate of constitutive norm, 193<br />

Rayleigh-Ritz or FEM, 202<br />

reaction<br />

support, 246<br />

truss member, 247<br />

recursive formula, 75<br />

recursive iteration<br />

optimality criterion, 197<br />

Recursive procedures, 253<br />

Redesign, 27<br />

redesign<br />

procedure, 208


310 Index<br />

limit, 208<br />

strategy, 208<br />

Redesign formulation, 281<br />

Redesign iterations, 75<br />

redistribution<br />

displacements, strains and stresses, 178<br />

Reference coordinates, 211<br />

Reference stress, 92<br />

Reference values, 57<br />

References to results, 210<br />

reformulation<br />

LP, 279<br />

Regular sub-matrix, 274<br />

relative angle, 108<br />

Relative hole area or density, 237<br />

Relative low or high shear stiffness, 256<br />

relative shear stiffness, 112<br />

Relative strains, 257<br />

Repeated sub-structures, 39<br />

Research groups, 198<br />

response analysis<br />

continuous, 261<br />

discrete, 261<br />

Response detector, 201<br />

result<br />

high energy density, 173<br />

insensitive compliance, 173<br />

load dependence, 173<br />

non-linearity dependence, 173<br />

Poisson’s ratio dependence, 173<br />

sensitive concentration, 173<br />

stationary compliance, 173<br />

Results and conclusions, 173<br />

Results in strains, 256<br />

Results in stresses, 256<br />

Results with different assumptions, 48<br />

Robustness, convergence, 259<br />

rotational transformation<br />

constitutive relation, 109, 212<br />

orthonormal, 212<br />

strain, 108, 212<br />

stress, 108, 212<br />

Same optimal topology, 41<br />

Satisfied or violated ?, 22<br />

secant formulation, 219, 242<br />

self-weight, 33<br />

Self-weight included, 34<br />

semi-analytical, 93<br />

Semi-analytical sensitivities, 93<br />

sensitivity<br />

constant constitutive matrix, 227<br />

constant volume, 227<br />

distinct eigenfrequency, 202<br />

flutter load, 266<br />

local determination, 226<br />

matrix, 280<br />

multiple eigenfrequency, 203<br />

to constitutive parameter, 227<br />

to material orientation, 227<br />

vector, 280<br />

with respect to design, 265<br />

with respect to load level, 265<br />

Sensitivity analysis, 23, 178, 180, 182<br />

sensitivity analysis, 93, 261, 264<br />

pseudo-load, 93<br />

effectiveness, 98<br />

localized, 241<br />

multiple eigenvalue, 270<br />

robustness, 98<br />

semi-analytical, 93<br />

Sensitivity for simple beams, 86<br />

Sensitivity functionals, 266<br />

Sensitivity of energy, 191<br />

Sensitivity result, 86<br />

Separated constraints, 274<br />

Separated objective, 274<br />

separation<br />

spatial problem, 264<br />

sequential linear programming, 91<br />

sequential linear programming (SLP), 285<br />

Sequential LP, i.e. SLP, 279<br />

sequential quadratic programming, 98<br />

Severe concentration, 132<br />

shaft<br />

shoulder fillet, 138<br />

Shape, 18<br />

shape<br />

boundary curve, 18<br />

boundary surface, 18<br />

crack tip, 151<br />

curvature, 18<br />

effective stress, 130<br />

energy density, 130<br />

heuristic approach, 129<br />

influence from boundary condition, 172<br />

influence from material non-linearity, 172<br />

influence from Poisson’s ratio, 172<br />

influence from volume constraint, 172<br />

joint positions, 18


Index 311<br />

length, 18<br />

material boundary, 159<br />

maximum bulk modulus, 151<br />

minimum compliance, 129<br />

minimum stress concentration, 129<br />

neighbouring shape, 130<br />

redesign, 130<br />

single hole boundary, 172<br />

smooth modelling, 130<br />

stiffest design, 129<br />

strongest design, 129<br />

Shape and size optimization, 186<br />

shape optimization<br />

stiffness, 232<br />

shear modulus, 200<br />

shear stiffness<br />

high, 112<br />

low, 112<br />

shift of eigenvalues, 86<br />

Shift of unknowns, 281<br />

short column, 251<br />

side constraint<br />

move-limit, 99<br />

Side constraints, 56, 99<br />

Simple example, 120<br />

Simple iterations, 85<br />

Simple laboratory tests, 198<br />

Simple parametrization, 153<br />

Simple unidirectional, 31<br />

simplex<br />

modified procedure, 249, 276<br />

procedure, 249, 273<br />

Single load case, 31, 100<br />

Single value, 19<br />

Size, 18<br />

size<br />

area of bar, 18<br />

mass density, 18<br />

material density, 159<br />

material orientation, 159<br />

orientation, 18<br />

thickness of beam, plate, shell , 18<br />

volume density, 18<br />

Size and shape, 229<br />

Size and shape with linking, 98<br />

size optimization<br />

stiffness, 231<br />

strength, 232<br />

Size, shape and topology, 32, 159<br />

Slack variables, 281<br />

slender column, 250<br />

Slender columns, 95<br />

Slender or short, 57, 250<br />

solid cross-section, 36<br />

solution<br />

basis feasible, 274<br />

existence, 259<br />

neighbour feasible, 274<br />

optimal, 274<br />

starting basis, 280<br />

vertex, 275<br />

Solution equation, 255<br />

space truss, 44<br />

Special case, 192<br />

Specific incompressible, 221<br />

Specified linking, 100<br />

Square root dependence, 250<br />

stability<br />

bar in a truss, 31<br />

Stability (buckling), 21<br />

stability measure, 264<br />

Stabilizing or destabilizing, 265<br />

Standard LP, 279<br />

statically determinate<br />

optimal truss, 245<br />

Statically determinate solution, 246<br />

Statically determined, 32<br />

statically determined, 31<br />

stationarity<br />

condition, 255<br />

energy functional, 261<br />

Stationarity - extremum, 24<br />

Stationary energy density, 254<br />

Stationary objective, 230<br />

Stationary systems, 262<br />

statistics<br />

uncertainties, 204<br />

steepest decent<br />

orientation, 253<br />

Stiffest design, 231<br />

stiffest design<br />

shape, 234<br />

size, 231<br />

Stiffness and strength, 55<br />

Stiffness constraints, 100<br />

stiffness global measure, 55<br />

Stiffnesses: membrane coupling bending, 113<br />

Stopping condition, 29<br />

strain<br />

direction of larger principal, 253


312 Index<br />

effective, 219, 242<br />

elementary state, 214<br />

mean energy density, 254<br />

principal, 242<br />

ratio, 171<br />

related parameter, 255<br />

rotational transformation, 212<br />

tensor, 211<br />

vector, 212<br />

Strain concentration, 149<br />

strain energy, 225<br />

per plate area, 115<br />

uniform density, 231<br />

strain energy density, 213<br />

θ function, 255<br />

integrated, 223<br />

Strain energy per area, 115<br />

Strain or stress directions, 260<br />

Strain ordering from stresses, 259<br />

strategy<br />

optimization, 259<br />

strength, 31<br />

constraint, 222<br />

Hill, 222<br />

von Mises, 222, 232<br />

Strength constraints, 99<br />

strength local measure, 55<br />

Strength or stiffness, 238<br />

stress<br />

allowable in compression, 245<br />

allowable in tension, 245<br />

bound for optimal ratio, 257<br />

characteristic, 253, 259<br />

deviatoric, 221<br />

direction of larger principal, 253<br />

effective, 219<br />

energy-consistent, 219<br />

Hill, 222<br />

von Mises, 219<br />

limit of proportionality, 250<br />

normal term, 220<br />

principal ratio, 253<br />

ratio, 171<br />

rotational transformation, 212<br />

shear term, 220<br />

tension, bending, shear, torsion, 98<br />

tensor, 211<br />

unidirectional, 179<br />

vector, 212<br />

von Mises, 92<br />

stress concentration<br />

analytical, 145<br />

factor, 169<br />

measured by umax, 169<br />

stress energy, 225<br />

complementary formulation, 234<br />

stress energy density, 213<br />

θ function, 256<br />

integrated, 223<br />

Stress ordering from strains, 258<br />

stress ratio<br />

value, 259<br />

Stress release, 149<br />

Stress/strain parameters, 179<br />

Stress/strain vectors, 108<br />

Stresses and strength, 21<br />

Strong influence from orthotropy, 148<br />

strongest design<br />

shape, 235<br />

size, 232<br />

Strongly non-isotropic, 204<br />

Structural models, 13<br />

Structure or material, 107<br />

Structures of 1D-elements, 81<br />

Structures or materials, 129<br />

Sturm sequence check, 86<br />

Sub-problems, 25<br />

sub-space iteration, 86<br />

Sub-vectors, sub-matrices, 220<br />

Subspace iterations, 202<br />

successive iteration<br />

heuristic approach, 239<br />

sufficient, necessary, 24<br />

super-ellipse<br />

power, 172<br />

super-elliptic shape, 236<br />

Supports as members, 247<br />

surface traction, 225<br />

Symmetry or non-symmetry, 91<br />

system<br />

conservative, gyroscopic, 262<br />

conservative, non-gyroscopic, 262<br />

homogeneous, 262<br />

inhomogeneous, 262<br />

instationary, 262<br />

non-conservative, circulatory, 263<br />

non-conservative, dissipative, 263<br />

System matrix, 264<br />

Table comments, 123


Index 313<br />

tangential stress, 144<br />

constant, 145<br />

Taylor expansion, 280<br />

tensile bar, 31, 250<br />

tensor<br />

constitutive, 211<br />

strain, 211<br />

stress, 211<br />

Tensors, 211<br />

Test of solutions, 89<br />

Test specimens, 201<br />

The θ function for strain energy density, 255<br />

The θ function for stress energy density, 256<br />

The match problem, 254<br />

theorem<br />

stiffest shape design, 235<br />

strongest shape design, 235<br />

truss, 246<br />

Theoretical reference, 59<br />

Theory and tools, 13<br />

Theory in chapter 12, 161<br />

Theory in chapter 15, 160<br />

thermal expansion<br />

negative, 210<br />

Thicknesses or densities, 56<br />

Three figures, 65<br />

Timoshenko beam theory, 93, 98<br />

Topology, 18<br />

topology<br />

a bar or no bar, 18<br />

a beam or no beam, 18<br />

a hole or no hole, 18<br />

material, 159<br />

truss, 245<br />

topology optimization, 162<br />

Total mutual potential, 265<br />

Total potential, 226<br />

tower space truss, 44<br />

trace<br />

invariant, 212<br />

transition<br />

slender - short, 251<br />

transmission of external forces, 54<br />

truss<br />

positions for joints, 44<br />

topology, 245<br />

truss from bars, 81<br />

Trusses, 14<br />

Two level iterative optimization, 44<br />

Two or only one parameter(s), 236<br />

Two sub-cases, 277<br />

ultimately optimal continua, 78<br />

uncertainties<br />

statistics, 204<br />

Unchanged with size design, 182<br />

Unchanged with stress loads, 179<br />

Uniaxial stress state, 145<br />

unidirectional state, 31<br />

Uniform energy density, 14<br />

uniform energy density, 55<br />

uniform strain energy density, 231<br />

Unknown support reactions, 246<br />

Using virtual work principle, 226<br />

Value of multiplier, 89<br />

Value of variable introduced, 277<br />

variable<br />

artificial, 283<br />

change, 281<br />

non-negative, 273, 281<br />

positive and negative, 276<br />

slack, 281<br />

variational analysis<br />

general, 264<br />

vector<br />

acceleration, 262<br />

constraint, 280<br />

cost coefficient, 274, 284<br />

design variable, 280<br />

displacement, 262<br />

element level, 267<br />

force, 262<br />

invariant length, 212<br />

maximum redesign, 281<br />

minimum redesign, 281<br />

non-negative unknown, 274<br />

redesign bound, 281<br />

redesign variable, 281<br />

sensitivity, 280<br />

strain, 212<br />

stress, 212<br />

system level, 267<br />

velocity, 262<br />

Vectors, 211<br />

vertex<br />

solution, 275<br />

vibration<br />

frequency, 261<br />

simple, free, undamped, 262


314 Index<br />

Vibrations, 15<br />

virtual work principle, 226<br />

volume<br />

relative density, 56<br />

volume force, 225<br />

von Mises stress, 92<br />

Weakly non-isotropic, 203<br />

weighted trigonometric integral, 113<br />

Well defined instability criterion, 265<br />

well-ordered strains, 257<br />

With and without self-weights, 101<br />

work, 55<br />

zero string, 251

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!