PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 cJCIRC/DatoDue.p65-p.15 COMPUTATIONAL DESIGN OF MECHANICAL STRUCTURES IN ELASTICITY USING MULTI-RESOLUTION ANALYSIS By S udarsanam Che llappa A DISSERTATION Submitted to Michigan State University in partial fiilfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mechanical Engineering 2003 C OMPL’TATI A form elastic system various resol homogenizatit resolution an: elasticity OPE information mathces that models of la in‘v’OMng CO intensive CO, is applied u heterogmem mmPIES in ‘ presented. ABSTRACT COMPUTATIONAL DESIGN OF MECHANICAL STRUCTURES IN ELASTICITY USING MULTI-RESOLUTION ANALYSIS By Sudarswram Chellappa A formal methodology for reducing the size of models used in the analysis of elastic systems is presented. This involves an explicit representation of the model at various resolutions and is accomplished using a projection generated numerical homogenization procedure. The fiamework for this analysis is derived from the multi- resolution analysis associated with the construction of wavelet bases. This is applied to elasticity operators to average fine scale properties and behavior while limiting loss of information. In discretized form, the method produces equivalent smaller stiffness matrices that can be used as building blocks (super-elements) to construct reduced models of larger systems. The principal application envisioned is in design problems involving complex structural systems, such as in crash-worthiness design, where very intensive computations demand computational efficiency. This model reduction scheme is applied to the problem of layout optimization of structures involving multi-scale heterogeneities of sizes that may be comparable to the size of the structure. Numerical examples in which the heterogeneities are in the form of perforations of various sizes are presented. T 0 my parents iii l woult Your guidar Dr. Andre E time spent i like to thanl Lastly. motivation ' ACKNOWLEDGEMENTS I would like to express my Sincere gratitude to my advisor Dr. Alejandro Diaz. Your guidance and support has been invaluable. I thank the members of my committee: Dr. Andre Benard, Dr. Michael Frazier and Dr. Farhang Pourboghrat. Thank you for the time spent in reviewing this document and providing valuable suggestions. I would also like to thank Dr. Martin Bendsrae. Thank you for your insight and suggestions. Lastly, I would like to thank my family. Thank you for all your support and motivation without which this endeavor would not have been possible. iv LIST 1 LIST ( 1 Intro 2 Wave 3.4.: 3.4 3 35 Comp TABLE OF CONTENTS LIST OF TABLES viii LIST OF FIGURES ix 1 Introduction ........................................................................................................ l 2 Wavelets And Multi-Scalc Representation Of Functions ................................. 7 2.1 Introduction ................................................................................................. 7 2.2 Mum-Resolution Analysis ............................................................................ 8 2.3 Example Using the Haar Basis ...................................................................... 10 2.4 Generalized Orthogonal Scaling Functions and Wavelets .............................. 10 2.5 The Discrete Wavelet Transform .................................................................. 15 2.6 Constructing The Basic Scaling Function and Wavelet ................................. 19 2.7 Function Approximation Using Orthogonal Projection ................................. 23 2.8 Wavelets In Multiple Dimensions ................................................................. 25 3 Model Reduction In Elastostatics ...................................................................... 28 3.1 Scales And Material Properties In Elasticity ................................................. 29 3.2 The Fine-Scale Elasticity Problem ................................................................ 32 3.3 A Multi-Resolution Analysis Of Material Distribution ................................... 34 3 .4 A Multi-Resolution Analysis Of Displacement .............................................. 37 3.4.1 A Periodic Multi-Resolution Reduction Scheme ................................. 37 3.4.2 Transformation To Finite Element F orrn ............................................ 44 3.4.2.1 Method I ............................................................................... 45 3.4.2.2 Method H ............................................................................. 46 3.4.2.3 Method III ............................................................................ 48 3.4.3 Computational Aspects ...................................................................... 50 3.5 Comparison Of Model Reduction Schemes ................................................... 53 DJ a .3 . 3.5.1 Comparison Between MRA Of Material Distribution And MRA Of Displacements ................................................................................... 53 3.5.2 Comparison Between MRA Techniques And Classical Homogenization ................................................................................ 56 3 .6 Computing The Non-Periodic Reduced Stiffness Matrix Of A Substructure ................................................................................................ 62 3 .7 Computing Fine-Scale Stresses By Augmentation ........................................ 66 3.7 .1 Computing The Periodic Large-Scale Nodal Displacements ............... 68 3.7.2 Conversion To A Wavelet Basis ......................................................... 71 3.7.3 Periodic Multi-Resolution Refinement ................................................ 72 3.8 Numerical Examples .................................................................................... 74 3.8.1 Example 1 ......................................................................................... 74 3.8.2 Example 2 ......................................................................................... 81 3.8.3 Example 3 ......................................................................................... 87 4 Multi-Scale Layout Optimization Of Structures ............................................... 93 4.1 Background: Topology Optimization Of Structures ..................................... 94 4.2 Optimal Design Using Macro-Scale Heterogeneities .................................... 97 4.2.1 Building A Model Using Reduced Substructures ................................ 99 4.2.2 Constructing Libraries Of Perforated Substructures ........................... 101 4.2.3 Sensitivity Analysis ............................................................................ 103 4.3 Compliance Minimization Problems .............................................................. 105 4.4 Optimization Using Fixed Layouts Of Reduced Substructures ...................... 108 4.5 Optimization Using Variable Layouts Of Reduced Substructures .................. 110 4.5.1 Perforated Substructures And Layout Dependency ............................. 110 4.5.2 A Dividing Approach To Optimization With Variable Layouts ............ 113 4.5.3 A Merging Approach To Optimization With Variable Layouts ............ 115 4.6 Examples ..................................................................................................... 119 4.6.1 Example 1 .......................................................................................... 119 4.6.2 Example 2 .......................................................................................... 121 4.6.3 Example 3 .......................................................................................... 122 5 Conc L" n— .U‘ k) lil’t b.) BIBLIO' 4.6.4 Example 4 .......................................................................................... 125 4.6.5 Example 5 .......................................................................................... 132 5 Concluding Remarks .......................................................................................... 139 5.1 Summary ...................................................................................................... 139 5.2 Conclusions ................................................................................................. 140 5.3 Areas Of Future Work ................................................................................. 141 BIBLIOGRAPHY .................................................................................................. 142 ME Me A]... a). List Of Tables 3.1 Main Results For Example 1 .......................................................................... 79 3.2 Main Results For Example 2 .......................................................................... 86 3.3 Main Results For Example 3 .......................................................................... 92 viii 3_1 Ha: 22(3) Ori 22th) Apr 22(0) Apr 2.2(d) Der 2.3 Sin; 2.4 The 2.5 The 2.6 Sing 27 The 28 Son 29 Ortl 210 Am 3 1 A st 3-2 Sun. 33 Basi: 3'4 Com 35 Com 3‘6 Com 3'7 Choir 3'8 Boun using List Of Figures 2.1 Haar scaling firnction and wavelet ................................................................. 11 2.2(a) Original function ........................................................................................... 12 2.2(b) Approximation at scale 6 ............................................................................... 12 2.2(c) Approximation at scale 5 ............................................................................... 12 2.2(d) Detail at scale 5 ............................................................................................. 12 2.3 Single stage discrete wavelet decomposition .................................................. 17 2.4 The convolution operation ............................................................................. 17 2.5 The dyadic down sampling operation ............................................................. 18 2.6 Single stage discrete wavelet reconstruction .................................................. 18 2.7 The dyadic up sampling operation ................................................................. 19 2.8 Some commonly used orthogonal scaling functions and wavelets ................... 22 2.9 Orthogonal projections fiom wavelet to finite-element spaces ........................ 23 2.10 A multi-scale representation of two-dimensional functions ............................. 27 3.1 A structure built from substructures of difl‘erent scales ................................... 31 3.2 Structure of L J and HJ_k .......................................................................... 43 3.3 Basis functions in wavelet and finite-element spaces ...................................... 44 3 .4 Conversion from wavelet to finite-element spaces using method I .................. 46 3.5 Conversion from wavelet to finite-element spaces using method II ................. 47 3 .6 Conversion from wavelet to finite-element spaces using method [[1 ............... 49 3.7 Choices of the material distributions for the comparison ................................ 55 3 .8 Bounds on the maximum relative error between the strain energy computed using material MRA and displacement MRA ................................................. 55 3.9(a) Strain energy comparisons for material distribution (a) .................................. 60 3.9(b) Strain energy comparisons for material distribution (b) .................................. 60 3.9(c) Strain energy comparisons for material distribution (c) .................................. 61 3.9(d) Strain energy comparisons for material distribution ((1) .................................. 61 3.10 Substructure surrounded by a weak fictitious domain .................................... 63 3.11 Constant pre-strains applied to a patch of substructures ................................. 66 3.12 3.13 3.14 3.15 3.16 3.17 3.18 3.19 3.20 3.21 3.22 3.23 3.24 3.25 3.26 3.27 3.28 3.29 3.30 3.31 3.32 3.33 3.34 3.35 Decomposition of a general non-periodic fimction into a coarse, non-periodic firnction and a periodic fimction ..................................................................... 69 Geometry and boundary conditions for example 1 ......................................... 74 Von Mises stress distribution using the fine-scale model of the structure in example 1 .................................................................................................. 75 Assembly of substructures for example 1 ....................................................... 76 Von Mises stress distribution in reduced model using material MRA ............. 77 Von Mises stress distribution in reduced model using displacement MRA ...... 77 Von Mises stress in substructure (A) from the reduced model using displacement MRA ........................................................................................ 78 Von Mises stress in substructure (A) after augmentation ............................... 78 Detail of Von Mises stress near substructure (A) obtained fiom the fine-scale model ............................................................................................ 79 Geometry and boundary conditions for example 2 ......................................... 81 Von Mises stress distribution from the fine-scale model of the structure ........ 82 Arrangement of substructures for example 2 .................................................. 83 Von Mises stress distribution fi'om reduced model using material MRA ......... 84 Von Mises stress distribution from reduced model using displacement MRA .. 84 Von Mises stress in substructure (A) from reduced model using displacement MRA ........................................................................................ 85 Von Mises stress in substructure (A) after augmentation ............................... 8S Detail of the Von Mises stress in a region around substructure (A) fiom the fine-scale model ............................................................................................ 86 Geometry and boundary conditions for example 3 ......................................... 87 Von Mises stress distribution fiom fine-scale model ....................................... 88 Arrangement of substructures for example 3 .................................................. 88 Von Mises stress distribution fiom reduced model using material MRA ......... 89 Von Mises stress distribution from reduced model using displacement MRA . 89 Von Mises stress in substructure (A) from reduced model using displacement MRA ........................................................................................ 9O Von Mises stress in substructure (A) after augmentation ............................... 90 3.36 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 4.9 4.10 4.11 4.12 4.13 4.14 4.15 4.16 4.17 4.18 4.19 4 20 4.2 4 22 4.23 4.24 4.25 4 76 ‘5- 'C '1‘). '7‘ W. ’Y) Ombnnmmi. 3.36 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 4.9 4.10 4.11 4.12 4.13 4.14 4.15 4.16 4.17 4.18 4.19 4.20 4.21 4.22 4.23 4.24 4.25 4.26 Detail of the Von Mises stress in substructure (A) obtained from the fine-scale model ............................................................................................ 91 A structure with a composite microstructure ................................................. 96 A typical topology optimization problem and solution ................................... 97 A typical structure with macro-scale heterogeneities (perforations) ................ 98 A structure built from an assembly of substructures with circular perforations ................................................................................................... 99 Assembly of substructures ............................................................................. 100 A non-conforming assembly of discretized substructures ............................... 101 Substructures with circular perforations for various finest levels .................... 103 Substructures with rectangular and circular perforations ................................ 105 Optimization using fixed layout of substructures ............................................ 107 Optimization using variable layouts of substructures ...................................... 107 A design domain and a possible layout of reduced substructures .................... 109 Boundary traction test ................................................................................... 1 10 Comparison of strain energies in substructures with different number of perforations for various boundary tractions .................................................... l 1 1 Comparison of strain energies in substructures with different number of perforations for various constant pre-strains .................................................. 112 Illustration of possible evolution of layouts using the dividing approach ......... 113 Illustration of possible evolution of layouts using the merging approach ........ 116 Problem description for example 1 ................................................................. 120 Optimal layout for example 1 ........................................................................ 120 Problem description for example 2 ................................................................. 121 Optimal layout using multiple load cases for example 2 .................................. 121 Optimal layout using single load case for example 2 ....................................... 122 Problem description with uniform layout for example 3 ................................. 122 Problem description with multi-size layout for example 3 ............................... 123 Optimal layout using single-size substructures for example 3 ......................... 123 Layout using multi-size substructure for example 3 ........................................ 124 Problem description for example 4 ................................................................. 125 4428 4429 430 44 4- b.) D.) k) b) b) 440 (A (b Pn Ini lay Sec ()pi SITU Star corr Seqt ()pfii Sch subst LayOI 1.15an 4.27 4.28 4.29 4.30 4.31 4.32 4.33 4.34 4.35 4.36 4.37 4.38 4.39 4.40 Starting arrangement of substructures for optimization using the dividing approach for example 4 and the corresponding optimal layout ....................... 126 Sequence of steps in the dividing approach for example 4 .............................. 127 Optimal arrangement of substructures using the dividing approach for example 4 and the corresponding optimal layout ............................................ 128 Starting arrangement of substructures using the merging approach for example 4 and the corresponding optimal layout ............................................ 129 Sequence of steps in the merging approach for example 4 .............................. 130 Optimal arrangement of substructures using the merging approach and the corresponding optimal layout for example 4 .................................................. 131 Problem description for example 5 ................................................................. 132 Initial layout using the merging approach and optimal structure using this layout for example 5 ...................................................................................... 133 Sequence of layouts obtained using the merging approach for example 5 ....... 134 Optimal arrangement of substructures and the corresponding optimal structure obtained using the merging approach for example 5 ........................ 135 Starting layout of substructures for the dividing approach and the corresponding optimal structure for example 5 .............................................. 13 5 Sequence of layouts obtained using the dividing approach for example 5 ....... 136 Optimal arrangement of substructures and the corresponding optimal structure obtained using the dividing approach with prescribed number of substructures of size 2L for example 5 ......................................................... 137 Layout of substructures and corresponding optimal structure obtained using the dividing approach with a perimeter constraint for example 5 ........... 138 xii Cl lnt In the optimiz prevent an autt CompUll for exte cOfltplelt POSsible minimUm ln 44am is invoked Chapter 1 Introduction In the design of complex structural systems or structures with complex behavior to optimize certain properties such as crashworthinees, the complexity of the problem often prevents a detailed computational analysis of the structure. A typical computer model of an automotive structure involves around 106 degrees of freedom requiring days of computer time to run a single analysis. Under these conditions the current practice calls for extensive simplifications. The strategies for simplification of the optimal design of complex structures (or structures with complex behavior) can be divided into two possible approaches: one that uses a full-scale model but limits the evaluations to a minimum and the other that uses a reduced model. In the first approach, the optimization is applied on surrogate models (or response surfaces), constructed using a strategic (statistical) sampling of a lull-scale model, which is invoked only sparingly to save effort. The firll-scale model is still used as the principal source 0' evaluatior very effer [40]). As to build attractive ; Th full-scale r noted that inforrnatior rePlaces. 1 0f accuraq Same Ph)'si the results c It is one that a mode'15) eff. Satings that PIGSemS representatic Preliminary, i it ignOres fee NeVetheleSS source of information about the system and therefore the computational cost per evaluation is not changed. Instead, the number of evaluations is reduced. This strategy is very effective whenever the number of design variables involved is small (see Yang [40]). As the number of variables increases, the number of firnction evaluations required to build a reasonable response surface also increases, rendering the approach less attractive and not of much advantage compared to using the firll-scale model throughout. The main idea in the second strategy is to reduce computations by replacing the full-scale model by a reduced model, a model that is less expensive to evaluate. It is noted that the process of constructing the reduced model may result in the loss of critical information, causing the reduced model to be significantly less accurate than the model it replaces. In this strategy the computational cost per evaluation is reduced at the expense of accuracy. In addition, design variables in the reduced order model may not have the same physical meaning as the variables in the original problem, making interpretation of the results diflicult. It is likely that the most effective strategy for the design of complex structures is one that combines the two approaches (i.e. , response surface methods and reduced models) effectively. A response surface methodology would clearly benefit from the savings that result fi'om a carefirlly crafted, reduced model. With this in mind, this work presents a formal methodology for model reduction that involves an explicit representation of the model at various scales. The work presented here is still preliminary, in the sense that it addresses only the elastic behavior of a structure and thus it ignores features that are crucial to the full understanding of many complex problems. Nevertheless, this is a necessary first step and it provides an important understanding of how a structu T quite sr primaril reductic methods freedom degrees freedom componer the dynan cOITtponen 93d] inten PfOblem is PTOpenies r Coarse]- mor average the . The Cc from fine‘xa dlflEIEnt Seale‘ how a formal procedure can be derived to reduce the complexity of models used in structural design without losing information that can be relevant in the design problem. The problem of constructing reduced models of structures has been investigated for quite some time in the context of vibration and control of structural systems. These primarily comprise of methods such as pseudo static variants of the classical Guyan’s reduction method or component-mode based techniques. In the Guyan’s reduction type methods (see Guyan [22], Friswell et a1 [19], Wilson et al [39]), a set of degrees of fi'eedom of the system are chosen to be master and another set chosen as the slave degrees of freedom. The reduction process aims to eliminate the slave degrees of freedom and express the state equation in terms of the master degrees of freedom. In component mode based techniques (see Hurty [24], Craig and Bampton [8], Seshu [35]), the dynamics of a structure are described by selected sets of normal modes of individual components of the structure, plus a set of static vectors that account for the coupling at each interface where individual components are connected. In the present work, the problem is phrased in the language of homogenization and the computation of effective properties of composites: starting fi'om a structure that is modeled in fine detail, we seek coarser models of the same structure (i.e., “homogenized” or “effective” structures) that average the detail without losing the (relevant) fine scale information. The collection of mathematical methods for extracting the coarse-scale behavior from fine-scale models is termed homogenization. Typically, such problems are solved using asymptotic expansion techniques or weak-limits; see Bensoussan et a1 [5]. In these techniques, there is no accounting for structures with features involving distinctly different scales. In a recent paper by Pecullan et al [30] the subject of scale effects on the behavior tensors < volume r numerical appear in homogeni; resolution linear hom system of e original 53,-: these efl‘ect could deter Complete Cl: [30] applied eqmvalent tr method of 1 developed m more rObUSt situations tha techqu to . homogenizat, that it Com d l Dorabantu u‘ behavior of two-dimensional composites is discussed by comparing the apparent stiffness tensors of two-dimensional elastic composites for various sizes of the representative volume element. Also, there has been substantial effort to develop methods for numerical homogenization that facilitate the analysis of problems involving systems that appear in multiple scales. Brewster and Beylkin [7] outlined a procedure for numerical homogenization of a system of linear ordinary difl‘erential equations using the multi- resolution analysis (MRA) associated with the construction of a wavelet basis. Their linear homogenization procedure consists of an algorithm to produce an effective linear system of equations whose solutions are the coarse-scale projections of the solution of the original system of equations, called multi-resolution reduction, and one for augmenting these efi‘ective equations to produce the homogenized solution, called augmentation. One could determine the projection of the solution at any intermediate scale and obtain a complete description of the transition from fine to coarse scale representation. Gilbert [20] applied the same approach to a system of two ordinary differential equations that is equivalent to a one dimensional second order elliptic problem and compared the classical method of homogenization, i.e., the asymptotic expansion method, with this recently developed multi-resolution technique. It was noted that the MRA scheme is physically more robust than the classical theory, i.e. it could be applied to many more physical situations than the classical theory. Dorabantu and Engquist [15] applied the same MRA technique to a discrete elliptic second-order differential equation. They observed that this homogenization procedure produced an operator that preserved its divergence form and that it could be well approximated by a band diagonal matrix. The work by Gilbert and Dorabantu use the Haar basis (piecewise constants) for the discretization of the djfi‘eren difiereni In of elastic using fix: Glowinski work, the elasticity n could be it large stitfne generalized material in 2 this problem 10 the size 01 is Specifican} me Principa] ‘ To deve} 83.516015 finescar. differential equations. Beylkin and Coult [6] applied this MRA method to elliptic partial differential equations and studied the spectral characteristics of the reduced operators. In linear structural analysis, the associated differential equations are the equations of elasticity. Techniques for solving the elasticity problem defined on arbitrary domains using fixed-scale wavelet-Galerkin methods have been investigated for some time (see Glowinski et a1 [21], Wells et al [37], Diaz [14], DeRose and Diaz [11]). In the present work, the MRA based numerical homogenization scheme is applied to the equations of elasticity modeled using a wavelet-Galerkin technique. In this case, the model reduction could be thought of as a method to produce equivalent smaller stiffness matrices from large stiffness matrices. An application of this method is then presented in the context of generalized topology or layout optimization of structures, i.e., the optimal distribution of material in a given design space subject to prescribed loads. The existing methods for this problem do not account for the presence of finite scales that may even be comparable to the size of the design domain. Here, a method that uses a model reduction scheme that is Specifically tailored to the problem of layout optimization of structures such that finite scale heterogeneities can be accounted is presented. The principal goals of this dissertation are: 1. To develop a consistent scheme to compute equivalent reduced models of structural systems in linear elasticity at various coarse-scales that retain relevant features of fine-scale models. accc size The rema. introductic Chapter 3 framework Chapter 4 problem 0 heteIOgene possible dir 2. To develop a method for the layout optimization of structures in elasticity that accounts for multi-scale heterogeneities of sizes that may even be comparable to the size of the structure. The remainder of this dissertation is organized as follows. Chapter 2 gives a brief introduction to wavelets and the concepts of multi-resolution analysis of functions. Chapter 3 presents a model reduction scheme using the multi-resolution analysis framework. Numerical emples that illustrate the proposed scheme are provided. Chapter 4 discusses the application of the proposed model reduction technique to the problem of layout optimization of two-dimensional structures in elasticity, in which heterogeneities of finite scale can be accounted. Finally, some concluding remarks and possible directions for firture work are presented in Chapter 5. "H freq Stu} “av. hiera analc l Daub Chapter 2 Wavelets and Multi-Scale Representation of Functions 2.1 Introduction “Wavelets are mathematical fimctions that are used to cut up data into diflerent frequency components and then study each component with a resolution matched to its scale ”§. The concept of multi-resolution analysis is firndamental to the theory of wavelets. The main idea is the separation of the information to be analyzed hierarchically into principal and residual parts. In signal processing applications this is analogous to decomposing a signal into its low fi'equency and high fiequency components with the knowledge of when they occur. This has definite advantages over ’Daubechies [10] the standai information making “*8 Fourier trar been appliet solving ordi emphasis 0 chapter. F t refer to the references t1 2.2 Mull DCfiHIIIOn; filnctions ~ i. (i) {0; (ii) fl} (iii)g( (1") st.» (V) The the standard Fourier analysis, which identifies fi'equency information but no time information. A variety of efiicient algorithms using wavelets have been developed making wavelet transforms on par with computationally efficient methods such as fast Fourier transforms. Wavelets have become a very popular tool in engineering and have been applied to a wide range of problems in signal processing, image processing and in solving ordinary and partial differential equations. A brief introduction to wavelets with emphasis on applications in computational engineering analysis is presented in this chapter. For detailed information about the construction and applications of wavelets refer to the books by Daubechies [10], Frazier [18], Resnikofi' and Wells [31] and the references therein. 2.2 Multi-Resolution Analysis Definition: A multi-resolution analysis (MRA) of L2 (IR) - the space of square integrable firnctions — is a nested sequence of subspaces Vj such that (i) {0} c ---C V_1cV0 ch C---CL2(IR) (ii) njVj={o} and LIV—1:3 (1R) (iii) g(x)€ VI 4:) g(2x)€ Vj+1 (iv) g(x)€V0 4:) g(x—k)€ V0, k E Z (v) There exists a scaling function arJ_l CJ——-’ |——D 42h ——u 12 ._, cJ_1 Figure 2.3: Single-stage discrete wavelet decomposition In the figure, following the stande notation, 4: represents the convolution operator illustrated in figure 2.4 and i represents the down sampling operator defined as shown in figure 2.5; all the indices and arguments are evaluated modulo n. x(i) ——> *h .__., y(i)=Zh(k)x(i—k) Figure 2.4: The convolution operation 17 Similarly, Wavelet t as illustra and Here T d. denOte x(n) a} ytn)=x(2n) Figure 2.5: The dyadic down sampling operation Similarly, the discrete periodic form of equation (2.27) leads to the inverse discrete wavelet transform and is implemented by a sequence of convolutions and up samplings as illustrated in figure 2.6. The filters in this case are l T hzfilao 01 02 div-2 aN—l 0 01 (230) and 1 T g=:/—§[aN—I —aN_2 aN_3 01 —a0 0 "° 0] (2.31) dJ—1——h 12 ——> *g CJ CJ—1——N T2 *—> *h Figure 2.6: Single-stage discrete wavelet reconstruction Here T denotes the up sampling operator and can be defined as shown in figure 2.7 and 619 denotes the addition operator. l8 1n the p: hence all strictly convolut particula: input vet inverse t However influence “1115me Using Var 31C. 2.6 Co VNRIVnal The explic onh'the E accurate “(fights c x(n) ——> 12 L——->y(n): y(2n)=x(n) Figure 2.7: The dyadic up sampling operation In the preceding discussion it is assumed that the associated functions are periodic and hence all the indices and arguments are evaluated modulo n. It is noted that this is not strictly necessary and one could define the same transformations using linear convolutions rather than cyclic convolutions. In the periodic case, the formulation is particularly clean because the total number of terms in the transformed vectors and the input vector are always the same, the transformation matrix is square with a simple inverse that has an interesting structure and can be efficiently calculated using an FFT. However, there is the additional problem due to aliasing in the periodic case, this is the influence of the terms in one end of the function affecting the other end due to cyclic transformations. These problems are sometimes overcome by extrapolating the functions using various techniques such as zero padding, symmetric padding, reflective padding, etc. 2.6 Constructing the Basic Scaling Function and Wavelet The explicit use of the scaling firnction or wavelet is rare in most applications (one uses only the scaling and wavelet filter coefficients); however, they might be required for accurate function evaluations and visualization. In general, scaling functions and wavelets do not have a closed-form solution. Instead, they have to be computed 19 recursiw written t Evaluati: ;(ij)=< In matrix 01' In Order tC recursively from the dilation equation (equation 2.1). The dilation equation can be written explicitly as N—1. Thus, the only remaining equations are DD DD Di iiiib AAAAAAAAAAAAA 1.85 0.st '1 g firnction and wavelet '1 Daubechies-6 ' f v v v v r v v ‘ v w v v v r v ' 141444411111 4444111414144414411 1 A A A A A A A A A A A A A .- . A A . A . A . A . A v A . A . A . A r. 5 0 8 1 5 2 a g - u g 0 0 1 . _ . t .................. L T A a A v A j A n A A A 'J. Coiflet-Z scaling function and wavelet 4411141441441444414144 A A A A A A A A _ A A itblfihitiiifii’hiii’ A A l A A A Coiflet-8 scaling function and wavelet Figure 2.8: Some commonly used orthogonal scaling functions and wavelets '9 22 2.7. F 1 Projec In certair poly-nomiz multi—resc approximz illustrated analysis ar Let 1"” respeCIIVQI' v . FOr Element fur 2.7. Function Approximation Using Orthogonal Projections In certain applications, functions expressed in some commonly used basis (such as polynomials) need to be approximated in a suitable scaling function basis so that the multi-resolution analysis described earlier could be applied to them. Here, the approximation of functions in a scaling firnction basis using orthogonal projections is illustrated using a basis of bi-linear (hat) functions used commonly in finite element analysis and the Daubechies D6 scaling functions. Figure 2.9: Orthogonal projections from wavelet to finite-element spaces Let V w and V h be finite-dimensional spaces of periodic functions in 1.2 (1R) , spanned respectively by D6 scaling functions cpw and bi-linear (finite-element) shape functions h w w Ph w h - - - (,0 . For any f EV , let f EV be the prOJectron onto the space of finite element firnctions such that the L2 -norm of the error e, (e = f w - Ph (f w )) between the original firnction and the projected function is minimized, i.e., n—1 n—1 W” = Z fkw sot". P” (1"): Z fi'soi' (2.39) k=0 k=0 23 The Static In the vet Here C wavelet a integrals a the COmpL Problem; f ‘3‘ a! [27] transforms as Consider n' Find f)? 6 IR that n—1 n—1 h h 2 (2.40) min f 23 fkwwi" (y)—§: fk 90k (y) dy k=0 k=0 The stationary point of equation (2.40) is the solution to n—l h n-l h h h f 23/13"in 90j03’=f ka c.01.: ‘pjdy (2-41) k=0 1:20 In the vector form this can be written as Ct" = Nih (2.42) 4:) rh _—. N“Cr“’ (2.43) Here C and N are block-circulant matrices comprising of the inner products of the . w h h h ‘ wavelet and hat functions, f 80k cpjdy and f 90k cpl-dy , respectrvely. Values of these integrals are commonly referred to as connection coeflicients. Using the scaling relation, the computation of these quantities is usually reduced to the solution of an eigenvector problem; for details regarding the computation of these connection coefficients, see Latto et a1 [27], Dahmen and Michelli [9], Kunoth [26]. Thus a projection matrix that transforms vectors in the wavelet space to that in the finite element space can be written as P” = N‘ 1C (2.44) Consider now, the projection, Pw ( f h) of a given finite-element firnction f h E V b onto the wavelet space. As before, the statement of the projection problem can be written as n—1 n—1 Vf” = 2 that e V”. P” (f”)= Z) gin" (2.45) k=0 k=0 24 The statio (Note: (4‘) (Compare that transfr written as Findgi" 6 R that n—l h h n—l 2 (2.46) min] 2 ft 90k (”-2 giver? (y) dy The stationary point of equation (2.46) is the solution to n—l h h n—l f 21m spray= f Eire}! wydy=ff (2.47) k=0 k=0 (Note: (up, k e [0, n — 1]} forms an orthonormal basis for V‘” ). In vector form this is, CTrh = r‘” (2.48) (Compare the left hand sides of equations (2.41) and (2.47)). Thus a projection matrix that transforms vectors in the finite-element space to that in the wavelet space can be written as P” = CT (2.49) where, the matrix C is the same as the one defined earlier. 2.8 Wavelets In Multiple Dimensions In the preceding discussions it is assumed that the associated firnctions are one- dirnensional. In higher dimensions, the wavelet and scaling functions are defined as tensor products of the respective functions in 1-D. The space of two-dimensional functions at a scale J is given as VJ =VJ®VJ (2.50) where V J is the corresponding space of 1-D functions. For example, a two-dimensional scaling firnction is defined as follows: 25 All the functions fixed (fine It follows J—l can Thus. a mu notation as, “here the c' and the deta ‘PJ,k1 (X, y) = x1 x1 Figure 3.1: A structure built from substructures of different scales. Here 9: U06, where no is a substructure. Any point x in QC can be expressed as x = Tic + y , where Ye is the global coordinate of a reference point in the substructure and y is a coordinate local to the substructure. Then we can express the material distribution as p(x)=p(fc,y) (3.5) Also, it may be possible (and desirable) to express the stiffness matrix associated with the structure as an assembly of matrices corresponding to each substructure. Thus the problem of finding a reduced model of the structure becomes a problem of finding reduced stifliress matrices corresponding to each individual substructure. The concept of a structure being divided into substructures is similar to the notion of super-elements in stande finite-element analysis. 31 The homogeniz a mixturec with a fret infinitesima (weak limit scales in w the averag displacemer described ir 3.2. ml The Plane-s Such that Where 5 ( ll E is the e! elastic tense The above formulation of the problem has obvious similarities to that in periodic homogenization, where the goal is to find the effective material tensor that corresponds to a mixture of materials. The mixture is characterized by a cell that is repeated periodically with a frequency 1/ e, (e —i O) , i.e., material variation is assumed to take place at infinitesimal scales. Such problems are solved using asymptotic expansion techniques (weak limits) and the result is a homogenized material tensor. In the present problem the scales in which the material is distributed in Q are of finite dimensions and the result of the averaging process is an operator (a stifl’ness matrix) that relates loads and displacements in the reduced scale. The construction and reduction of the operators are described in the following sections. 3.2. The F ine-Scale Elasticity Problem The plane-stress elasticity problem on a prescribed domain 9 = Uflc seeks u E V (9) such that fEe(u)e(v)dQ= ftvdI‘ Vv€V°(Q) (3.6) Q 1‘t where 5(u) is the strain tensor associated with the displacement u; t is an applied traction on the boundary 1"; V is a space of kinematically admissible displacements and E is the elastic tensor defined on 9. Using the material model described earlier the elastic tensor within each substructure is expressed in the form E(y)=chO (3-7) 32 where EU a substruct problem (1 where the we assume with jump called the .s discretizati QC lS I’CSOl is the V'aIUe that resolv from subs: “ye discretiZed l discretizEltio e'g‘w 118ng f where E0 is a reference material tensor, pc characterizes the material distribution within a substructure and y E QC is a coordinate system local to the substructure. The elasticity problem defined on Q is now: ZfE(y)e(u)e(v)dy=ft.vdr Vv€V0(Q) (3.3) c Q Pt where the sum is interpreted in the sense of assembly. In order to facilitate computations we assume that pc is resolved with sufficient accuracy by a piecewise constant firnction with jump discontinuities at Cartesian grid lines spaced Sc = 2"J LC units apart (Sc is called the scale of the discretization), for some positive integer J called the level of the discretization. LC is the length of a side of the substructure. In practice, the geometry in QC is resolved by a digital image composed of 2J x 2‘] pixels of size Sc xSc , where pic], is the value of pc at the center of the pixel (LI). Sc and J are a measure of the scale that resolves pc. We denote this by writing pc (y) E p3 (y). Both Sc and J may vary from substructure to substructure. We refer to (3.8) as the fine-scale problem when all the substructures are discretized at their finest scale. In the typical problems of interest such a fine-scale discretization results in a system with too many degrees of freedom for emcient analysis, e.g., using finite elements. The large size of the fine-scale problem requires that the fine- scale stifliress matrix associated with each substructure be replaced by an equivalent stifiress matrix of a smaller dimension obtained through some consistent process of reduction. Here we propose two such reduction strategies: one based on a multi- 33 resolutic multi-ref inexpenS particula displacer resolutior rigorous relatively 3.3 t! Distril: All appro.‘ USlng a “1 15 resolved Where the 1 Pixel (k, I) and detail implemEme' Splits timer" 1 resolution analysis of the material distribution functions pc and the other based on a multi-resolution analysis of the displacements u. The first approach is computationally inexpensive but crude. It is used where there is no necessity to go from the solution at a particular discretization level to another, i.e., there is no consistent procedure linking the displacements at different levels of discretization. The second strategy, based on a multi- resolution analysis of displacements, is a more consistent procedure and provides a rigorous link between the displacements at various levels of discretization, but it is relatively a computationally expensive approach. 3.3 A Multi-Resolution Analysis of the Material Distribution An approximation scheme based on a representation of material distribution function pc using a wavelet expansion is presented. If the material distribution in a substructure QC is resolved by a 2‘] x 2‘, digital image, we can write N—l pc(y)=p5 (y)= E kaz 9011:! (y) (3.9) k,l=0 where the functions 90.1,“ are 2D Haar scaling fimctions, i.e., piecewise constant over the pixel (k,I) and N = 2J. A wavelet expansion of this function splits pc into coarse-scale and detail functions, ,5( y) and i)”( y) , respectively. The transformation is easily implemented using a 2-D discrete wavelet transform. More specifically, the transform splits functions pc 6 VJ , a space of dimension 22] = N 2 , into functions 5 6 VJ_1 and 34 fiefij_p dimension coarse Spa decomposr The coarse tensor prod Of pc OVe, filnction [7 lenSOr [)6 WJ_1, where WJ__1 is the orthogonal complement of VJ_1 in VJ. Space V J_1 is of dimension 220—1) 2 N 2 /4 , i.e., each application of the wavelet transform reduces the coarse space by a factor of four. After several applications of the wavelet transform the decomposition is of the form N —l y)= Z: pawn (y) k,1=0 (3.10) n—l J —l 2'" —l = Z EWJ-My yl+ )m2 2 Z (’W'ltmfluy k,l=0 -1 k, 1-0 r-l The coarse-scale function is of the form :20?“ also, “(y (3.11) and n=2f . Here the 2D scaling functions $131610) and the wavelets rpm (y) are tensor products of their corresponding 1D firnctions. Coefiicients 'p'k, are average values of pc over pixels of size A = 2‘] LC in an equally spaced grid of size nxn. The fimction 70'( y) is nothing but an arithmetic average of p] and in view of (3.7), the elastic tensor EA(y)=/7(y)E° (3.12) is an arithmetic average of E (y) over the substructure. A substructure made of material E A would over-estimate the stiflhess of the original substructure, as EA (y) is an upper bound for E (y). To compensate, we shall look for a harmonic average of E (y). For this purpose we define the firnction 35 whe coar by l CONE isal used a11d I] This E a fine N—l 950’): Z gilt VJ and “J,fJEVJ. Define W to be the transformation w : VJ _, VJ”l EBWJ_1 (3.26) that maps a function in the space VJ into functions in the space VJ—1 and its orthogonal complement WJ_1. The orthogonal transformation iJ—1 iiJ—r PJ—luJ WIIJ = QJ-1“J (3.27) represents the splitting of a vector “J into its coarse scale component i J_1 and the details frJ_1, where the operators PJ_1 :VJ —+VJ_1 and QJ_.1:VJ -> WJ"l are the coarse and detail projection operators respectively. The equation (3.23) can now be expanded as follows: LJ-r CJ—l E4 I'J—r 5.1—1 fiJ—r WLJwT(WuJ) = (3.28) T CJ—r AJ—r where, AJ—r == QJ—ILJQJ—l I WJ_1 —* WJ_1 CJ—l = PJ—rLJQJ—r 3WJ_1 —’ VJ—l (3-29) LJ—r = PJ—rLJPJ—l ZVJ—l —+ VJ_1 40 Here 6-1 and fJ_1 are the coarse-scale and detail components of the external force. The coarse scale component of the displacements iiJ_1 is found by block Gauss elimination, which yields the equation —1 T ._ - -—1 ~ (LJ—r —CJ-1AJ-1CJ—1)“J-1 = fJ—1_CJ—1AJ-lfJ—l (3.30) For the class of operators of interest here it can be shown that the operator AJ_1 is indeed invertible. This can be proved for our prototype problem in 2D elasticity as follows (The result can be extended to other cases easily). A 2D elastic stiffness matrix L J is positive semi-definite with two zero-eigenvalues that correspond to the two rigid body modes (translation in the x-direction and the translation in the y—direction). These two modes are constant functions and can be represented exactly at any scale, i.e., the detail components associated with these modes are always zero. This means that the operator Q J_1 is orthogonal to these rigid body modes. Let 8 denote the subspace spanned by the rigid body modes. From the positive-semi-definiteness of L J we know, that for all non-zero vectors x in the space V J that are orthogonal to the rigid body modes xTLJx >0, x:(x¢0,xEVJ\6} (3.31) Using the orthogonality of the wavelet transform operator, (3.31) can be written as (Wx)T (WLWT)(WX)>O, x:{x¢0, xEV‘I \9} (3.32) Using equation (3.28) this can be expressed as T LJ—r CJ—l Wx T CJ-r AJ—r Wx >0, x:{x=:0,xEVJ\9} (3.33) 41 Also, any vector in 6 has no detail components, i.e., VxEG, Wx= x (3 34 0 ‘ ) 0 Consider yEVJ such that Wy=[ l, where vEWJ-l. It can be seen fi'om equation v (3.34) (and using the linearity of the concerned operators) that y is orthogonal to any vector in the subspace 9. Now, for all such vectors y , yTLJy > o (3.35) Using the same approach as in equation (3.33), we have T LJ—l CJ—l 0 T CJ—r AJ—r 0 V > o, Vv 6 WJ‘1 (3.36) V Alter carrying out the multiplication we have vTAJ_1v > o, Vv 6 WJ—1 (3.37) Thus, fi'om equation (3.37) we can see that the matrix A J_1 is positive-definite and thus invertible. More generally, Engquist and Runborg [17] show that the operator AJ_1 is bijective for a class of elliptic operators L J obtained from bilinear forms. The matrix HJ—l i LJ—r — CJ—iAilrci—r (3 38) is the effective stifliress matrix at level J —1 associated with a periodic patch of identical substructures QC discretized at level J . Further reductions are possible by applying the reduction procedure recursively to obtain a sequence of effective stifl‘ness matrices HJ_2 , HJ_3, etc. It should be noted that the each reduction of the stiifiress matrix 42 results in a matrix that is relatively much denser than the original matrix. This is illustrated in figure 3.2 where the dark regions denote non-zero entries. LJ Figure 3.2: Structure of LJ and HJ_k If the structure considered involves forces that are slow-varying in nature, i.e., f = 0, then the above set of matrices (L J,{H J_k }) form a complete description of the substructure at various scales. The obtained efi‘ective stifliress matrices operate on vectors that result fiom a wavelet discretization of the associated functions. This creates some difficulties not only in the application of boundary conditions but also in the assembly of stifi‘ness matrices from several substructures. Moreover, the strain energy of an assembly of substructures is not the sum of strain energies of individual substructures since the associated basis functions in each substructure extend beyond the boundaries of the substructure. This is due to the fact that the support (=5) of the considered scaling fimctions (D6) is greater than 1. It might be advantageous to convert these wavelet efl‘ective stifi‘ness matrices into equivalent nodal stiffness matrices that operate on the nodal values of the associated forces and displacements defined on the substructure (similar to a bi-linear finite-element 43 stifiiress matrix). Also, these nodal-matrices would be more portable and can be incorporated into general-purpose widely available finite element codes. This conversion is described next. 3.4.2 Transformation to Finite-Element Form This section deals with finding an approximation to 111- that acts on finite-element instead of wavelet spaces. Introduce V” and V” as finite-dimensional spaces of periodic functions on DC spanned respectively by D6 scaling functions (cp) and bi-linear finite element shape functions (cph) shown in figure 3.3. '2 4 A D6 Scaling Function (2D) A 2D Bi-linear (hat) Function Figure 3.3: Basis functions in the wavelet and finite-element spaces respectively Representative functions in these spaces are of the form n-1 n-1 h h uw = Z ulcisokz and uh = Z um (3.39) k,l=0 k,l=0 where n=2j . Here it” ={ui’1} are wavelet coefficients and uh ={ullr'1} are displacements at nodes on an nxn uniformly spaced grid with spacing A: 2—j LC. Three approaches to convert a periodic stiffness matrix, H J- , into an equivalent nodal stiffness matrix in the (bi-linear) finite-element space are presented. 3.4.2.1 Method I For a given periodic stifliress matrix H (corresponding to a periodically tiled domain) and a force f w in V w with mean value zero (i.e., orthogonal to the two rigid body modes) there exists a unique displacement u” E V w (and associated coefficients n”) such that Hu‘” = r” (3.40) Let f h be the orthogonal projection of f " onto Vb computed as r” = Phr’” (3.41) where Ph is an orthogonal projection matrix that depends only on the choice of the two basis fimctions (as described in chapter 2, section 2.7). We now look for a transformation matrix Q that transforms the vector of wavelet displacement coefiicients uw , into a vector of nodal displacements uh , in such a way that the work of the loads 1'" and fh is the same, i.e., u” = Qu‘” (3.42) and rWTu‘” = rhTu” (3.43) This is illustrated in figure 3.4. 45 Figure 3 .4: Conversion from wavelet to finite element spaces using method I The work due to the force and displacement in V h can be written as thuh = fWTPhTQuw (3.44) Combining equations (3.43) and (3.44) we see that —1 Q = [PM] (3.45) Now, we look for a matrix E that relates the forces and displacements in Vh as Ku” = r” (3.46) Using equations (3.42) and (3.45) in (3.40), we have -—1 Eu" = r" => HPhTuh = [Ph] r” => PhHPhTuh = r” (3.47) Thus the equivalent finite element matrix is 17: = PhHPhT (3.48) 3.4.2.2 Method II Here we start with a finite-element firnction (force) f h E V h and denote f w to be its orthogonal projection onto V” who’s coefiicients are computed as 46 r” = owh (3.49) where P” is an orthogonal projection matrix. The objective here is to find a transformation matrix R : V h —+ V w that transforms a finite-element displacement into an equivalent wavelet displacement and in turn the associated nodal stiffness matrix such that the work of the loads 1'” and f h is the same, i.e., u‘” = Ru” (3.50) and thuh = rWTu‘” (3.51) This process is illustrated in figure 3.5. Figure 3.5: Conversion from wavelet to finite element spaces using method 11 The work associated with u”, f w E V W can be expressed as {”11” =rhTPWTRu” (3.52) Combining equations (3.51) and (3.52) we see that the works are the same only if R =[PWTJ—1 (3.53) 47 As before, we look for a finite element stiffness matrix that relates uh and I” as in” = r” (3.54) Using equations (3.50) and (3.53) in (3.40), we have -1 Eu” = r” => 11hr”) uh = owh _1 -1 (3.55) 4(1)”) HlPWT] uh =r" The equivalent nodal stifliress matrix is then — w —1 WT —l K =[P ] H[P ] (3.56) 3.4.2.3 Method III Here we introduce a matrix B , defined such that {a ( .) 1") — y) e u e u dy (3.57) 0c The right hand side of equation (3.57) is the elasticity bilinear form with finite-element (hat) trial functions and wavelet weight functions (compare to 3.22). The elastic tensor E is an average of the material in the substructure and is assumed to be of the form E=p3(y)E° (3.58) where E0 is a reference elasticity tensor and p3 is a distribution of relative densities that is obtained by averaging the fine-scale distribution of relative densities to the coarse-scale using (3.18) (It should be noted that this is the only place where an average of the material is used in this method that is based on a multi-resolution analysis of displacements). The matrix B is one that transforms a finite-element displacement uh E Vh into a wavelet body force f W E V w, i.e., f w = Bub is the wavelet body force 48 tl {1'3 WC and that corresponds to a finite-element pre-strain 5(uh). If the basis firnctions used as the trial and weight fimctions are interchanged, matrix BT is the one that transforms a given (wavelet) displacement uw E V” into a finite-element body force f h E V h . Having defined this operator B , we look for an operator Q : V w —’ V k that transforms a given wavelet displacement u” E V" into a nodal displacement uh E V h such that the work due to the transformed forces and displacements is the same as the work of the corresponding firnctions in V” , i.e., uh = Quw (3.59) and thuh = rWTu‘” (3.60) This process is illustrated in figure 3.6. Figure 3.6: Conversion from wavelet to finite-element spaces using method H1 The work corresponding to the displacements and forces in V h can be written as f ”uh = uWTBQuw = f WTH-IBQuw (3.61) 49 From eql Using eQI Thus, the Tl equivaleni substrucn 3.4.3 Cc In this se matrix of 1- A: sul (di From equations (3.60) and (3.61) we see that Q = B-lflj (3.62) Using equation (3.62) in the equation I—(uh = f h we have, fin" = r” 4:) Ron" = 13%” e» I‘m—Inn” = BTu‘” (3.63) Thus, the equivalent nodal stiffness matrix is then given as R = 3711“]; (3.64) The matrix K obtained using either of these approaches is the “finite element equivalent” to Hj. It relates nodal degrees of fieedom to nodal forces in a periodic substructure reduced from level J to level j . 3.4.3 Computational Aspects In this section, the computational procedure involved in obtaining a reduced stifliress matrix of a substructure using the multi-resolution analysis of displacements is outlined. l. Assemble a fine-scale wavelet stimiess matrix of a substructure, L J , as NJ L J = Z p‘gl0 (sum interpreted in the sense of assembly) e=l where N J 2 2J x 2J is the number of pixels in the fine-scale discretization of the substructure, l0 is a pre-computed wavelet “element” stiffness matrix of a pixel (dimension 50 x 50 for D6 scaling function) with a reference material tensor and p! is the value of the relative density firnction at a pixel e. 50 For]:- .N 0 C1 DJ at Tl ha end [00p de an. Usi Var For j = J to J —k +1 (k is the number of reduction levels), do 2. Compute the wavelet decomposition of the stiffness matrix at level j N2- 1 2 W21) 2 H - 2N - 2 'CI-l 1“ T 3. Compute the Schur’s complement to obtain the efi‘ective reduced stiffiress matrix at level j -l -1 T Hj—l = Lj—l _Cj—IAj—le-l J . . . 3N3- 3N2- . . N}- . This mvolves the solution of a TX T system of equatrons With —2— right 2 2 N,- 3N]- 2 3N- _x_ J 2 N. 2 hand sides and the multiplication of a matrix with a matrix. end loop n 4. Assemble B= prgbo , where n: ZJ-k x2‘]—k , p3 is an averaged relative e=l density distribution at the reduced scale (obtained as shown in equation (3.18)) and b0 is a pre-computed “element” matrix of size 50 x8 that is constructed using D6 and Hat firnctions as the weight and trial fimctions respectively in the variational form of the elasticity equation. 5. Compute I—( = BTHjlkB 51 End. The mo This my as matrix stiffness 1 computat finest-sea] reference Steps 4 ar are much fine~SCale relatively , The matrix H J_k is positive semi-definite (it has two rigid body modes). It is inverted by adding a matrix 50 to remove the rigid body modes. Here, 5 is a 1 1 o 0 suitable scalar penalty and Q=pr, where p: 0 0 1 1 is a matrix of rigid body modes. End. The most computationally intensive step is computing the Schur’s complement (step 3). This involves the solution of a system of equations for multiple right hand sides as well as matrix multiplication. At the end of each reduction stage (steps 2 and 3) the size of the stiffness matrix (and hence the number of equations) reduces by a factor of 4 and thus the computations progressively reduce after each reduction step. The first step, involving the finest-scale stiffness matrix, can be implemented efficiently with a pre-computed reference “element” stiffness matrix and using sparse assembly. The computations in steps 4 and 5 are performed on matrices in the reduced-scale; the sizes of these matrices 22k are much smaller (factor of , where k is the number of reductions) compared to the fine-scale stiffness matrix. It should be noted that the reduced stifliress matrices are relatively much denser than the fine-scale matrices, as illustrated in Figure 3 .2. 52 3.5 C In this : respect 3.5.1 l displa C onsidc to an h the spa nxn‘ i matrix KE 2 t diagong basesf eigenve 3.5 Comparison of the Model Reduction Schemes In this section, comparisons of the proposed model reduction schemes are presented with respect to each other and with the classical homogenization scheme. 3.5.1 Comparison between MRA of material distribution and MRA of displacements Consider K : V —+ V and K : V —> V to be the reduced stiffiress matrices that correspond to an MRA of material distribution and an MRA of displacements, respectively, and V is the space of kinematically admissible functions. Let the dimensions of the matrices be nxn, where n is the number of degrees of freedom in the models. Define E to be a matrix of eigenvectors and Q to be a diagonal matrix of eigenvalues of K , i.e., KE = 9E. Similarly, define I}. to be the matrix of eigenvectors and fl to be the diagonal matrix of eigenvalues of K , i.e., Ki) = {If}. Since E and i) are orthonormal bases for V , any eigenvector e,- of K can be expressed as a linear combination of A eigenvectors (3;) of K , i.e., f The two stiffness matrices have two zero-energy (rigid body) modes that correspond to a translation in the x and y directions respectively. Let i) be the matrix of eigenvectors of K other than the rigid body modes, dim (E) = n xn —— 2 . This can be expressed as A F: = EB (3.66) f1 5? as DU] where the entries in matrix B are of the form Bi- =(a;,éj) (3.67) and E,- and E,- are colurrm vectors of I73 and B respectively. Any (displacement) vector u that is orthogonal to the rigid body modes can be written as, u = Ea (3.68) for some ai E R . Using the orthogonality of Ii) , we can write the strain energy associated with this displacement it using the material MRA model as uTKu = «TETKEa = (ITS-ill (3.69) where Q is a diagonal matrix of the non-zero eigenvalues of K. The displacement u can be expressed in terms of the eigenvectors of K as u = 13:13.: (3.70) where the entries of B are as defined in (3.67). Now, the strain energy associated with u computed using the MRA of displacements can be expressed as uTKu = «TBTETKEBa = 6713712136 (3.71) The relative error in strain energy obtained using the two models can then be expressed as IuTKu — uTKu laT [fl — BTSAIB] a — _ (3 .72) luTKu laTflal It can be shown using Rayleigh’s principle that the eigenvalues of the matrix in the numerator of the right hand side of equation (3.72) form an upper bound on the relative 54 error of rigid bo« Here, th five difi? Figure 3 obtained eight) (All = error of the strain energies, i.e., for a normalized displacement u that is orthogonal to rigid body modes, luTKu—uTKul __ l T . max T , *Smax eig(I—ST B QB) (3.73) MS; '11 Kul Ila: Here, the reduced stiffness matrices of a periodic tiling of a substructure for a choice of five difi‘erent material distributions (Figure 3 .7) are computed and compared. (a) (C) (6) Figure 3 .7: Choices of material distributions for the comparison Figure 3.8 shows the plots of the eigenvalues of the matrix (A = I —§-1BT§B) obtained for the various choices of the material distributions chosen. -A- (a) as . O V 0.25 -v- (d) ' 4 -o- M a 0.2L 711 a ' ‘- ‘ ...‘ ~ .- '5 ‘. 11 5: 0.15 - 0.1 0.05 l Figure 3 .8: Bounds on the maximum relative error between the strain energy computed using material MRA and displacement MRA 55 The r 6 spacing are 121 usually spatial error i1 combin onthel 3.5.2 1 Home Here, displace mfinites. 15 assun base cell Where represent: Wording,I The reduced stiffness matrices are at a discretization that corresponds to a uniform spacing of degrees of freedom in a 8 x 8 grid. The dimensions of these stiffiress matrices are 128x128. The figure 3.8 shows the first 64 eigenvalues of the matrix A, as we are usually not concerned with the higher energy modes that correspond to rapidly varying spatial displacements. The value of the plots at each ‘i’ corresponds to a bound on the error in the strain energy associated with a displacement that can be expressed as a linear combination of the first ‘i’ modes. It can be seen that for the first few modes, the bound on the error is small but it increases as more number of modes are included. 3.5.2 Comparison between MRA techniques and Classical Homogenization Here, the averaging schemes based on the MRA of material distributions and displacements are compared to the classical homogenization method. In classical homogenization, the material distribution is assumed to vary at infinitesimal scales and an asymptotic analysis is used to compute effective properties. It is assumed that the material distribution is characterized by the periodic repetition of a base cell (1’ ). The displacement field it is expanded in an asymptotic series u=uo(x,y)+€ul(x,y)+62u2(x,y)+~- (3.74) where y = 16/8 (3.75) represents the local (microscopic) coordinates and x is the global (macroscopic) coordinate. It can be shown (see Bensoussan et al [5]) that the effective material tensor is given as 56 where \ periodic 1 The straill Where E 50. Thu $1 a periodic Equation k1 1 3x EW— _ Y —f 5W —E,-qu-—’i- dY (3.76) Y ay‘l where xfg’ is the solution to the so-called cell-problem defined on the infinitesimal periodic characteristic cell, given as follows 8 faqu— -—X—” div—tor: fEiJ-Hade VvEV' 6y 6y 6y,- (377) V: {vzv rs Y-periodic}Y The strain energy form of (3.76) can be written as _H 50 80 _1 o * 0 * Y where E is the material tensor and e. is the resultant strain due to the applied pre-strain, 50. Thus, we can define the strain energy associated with a (finite size) substructure QC , corresponding to a periodic repetition of infinitesimal cells Y , as (PH g -}17£(Eiqu (€3- —e;-)(egq —e;,q))dY theas(Qc) (3.79) Similarly, the strain energy associated with applying a constant pre-strain (so) on a periodic tiling of a (finite-size) substructure QC can be defined as t t (I) Q f(E,-qu (83- ‘51)“qu -epq))d§2 (3.80) 9c Equation (3.80) can then be written as ‘1’: IE iqusije pq +fE queijem 2f}; 0109595 N (3°81) 57 The ela: substruc‘ Using (3 The first strain an prescribe second 1 displacen The strair Where K bang that The Strair Where R SCaJe SOlu The elasticity problem associated with the application of a constant pre-strain on a substructure no, is defined as: Find if, such that 1 mp.) .11: is. o-eu*r.1) so Using (3.82) in (3.81), the strain energy becomes =fEiqu 51]qu —fE,~J-pqe;je W (3.83) The first term on the right hand side of equation (3.83) does not involve the resultant strain and can be computed exactly without solving any elasticity problem (as 80 is a prescribed constant strain). The model reduction schemes are used to compute the second term using an approximation of either the material distribution or the displacements at a coarse scale, i.e., MRA of material or MRA of displacements. The strain energy using a multi-resolution analysis of material is then M *T " =fE ”megs pq —u Kcu (3.84) where KC is a reduced stiffness matrix obtained as shown in (3.19), the only difference being that the substructure involved is assumed to be periodic. The strain energy using a multi-resolution analysis of displacements is D *T- * ch =ch megs m —u Ku (3.85) where K is a reduced stiffness matrix obtained as shown in (3 .64) and u‘ is the coarse- scale solution to a constant pre-strain elasticity problem (It should be noted that the body 58 force (T.1 material T homogerj various 1 these me] to a pres a particu would be material . material directioni Strains 5‘ force (right hand side of 3.82) in this case is computed using an equivalent, reduced material tensor). The comparison between the multi-resolution reduction schemes and classical homogenization is carried out by applying constant pre-strains to substructures with various material distributions and comparing the resulting strain energies when either of these models is used to compute the strain energy. The stifl‘er material, when subjected to a prescribed constant pre-strain will result in higher strain energy. In order to say that a particular material is stifl‘er, this result has to hold for all possrble strains. However, it would be meaningful only to consider strains whose principal directions coincide with the material axes, as only then would the strain energy be maximmr, i.e., the best use of the material is only when it is oriented in such a way that it coincides with the principal directions of the applied strain (see Pedersen [31]). Thus, among all such (normalized) strains 80 = , the strain energy produced by the strain with ,8 = O is the highest. 77 Here, substructures of unit sizes are considered with five choices of material distributions and the strain energies as a result of applying constant pre-strains of the form 0 l 5‘: 0 1, where r) E [—1,1], are computed. 77 Figures 3.9 (a) to (e) show the plots of strain energy () versus the parameter 7) , computed using the various methods discussed here. 59 2.00 Ti, 1.80 « 1.60 « 1.40 . 1.20 - 1.00 0'80 - + Fine-Scale 05° " —I— Homogen-tion 0.40 + + Disp MRA j=3 0.20 ~— -x-— Mat MRA j=3 0.00 I r n -1.00 -0.50 0.00 0.50 1.00 Figure 3 .9(a): Strain energy comparisons for material distribution (a) 2.00 - (I) 1 8° - + Fine-Scale 1 60 -I— Homogenization ' + Disp MRAj=3 “'0 ‘ —x— Mat MRA j=3 1.20 d 1.00 . 0.80 - 0.60 0.40 0.20 now r I T 7 l n -1.00 -0.50 0.00 0.50 1.00 Figure 3.9(b): Strain energy comparisons for material distribution (b) 60 0.80 + Fine-Scale 0.60 - —I— Homogen-tion 0.40 q + Disp MRA i=3 0.20 q -—>(—— Mat MRA i=3 0.00 . 1 . . ’7 -1.00 -0.50 0.00 0.50 1.00 Figure 3 .9(c): Strain energy comparisons for material distribution (c) 2.00 “ Q 1.30 i + Fine-Scale 1.60 — —I— Homogenization 1_4o . + Disp MRA F3 1 20 J —)(— Mat MRA i=3 0.00 r 1 -1.00 -0.50 0.00 0.50 1.00 Figure 3 .9(d): Strain energy comparisons for material distribution (d) 61 hrthe at discretiz anmnnc energy'i xmemr uxd F propertic dumm underesfi complica loss of ; essentiall matrices the effect 3.6 ( Matri: The redl DOt 3’61 t build a r be exp! \ In the above plots, the reduction process involved starting fi'om a 64 x 64( J=6 ) element discretization and reducing it to a 8 x8 ( j=3 ) discretization. For simple material structures such as the perforated substructures in Figure 3.9 (a), (b) and (c), the strain energy in the reduced model is essentially indistinguishable from the energy in the filli- scale model, regardless of whether the material based or displacement based MRA is used. For comparison, the graphs also show the energy in a substructure whose elastic properties are those of a periodic mixture at infinitesimal scales, i.e., the result from classical homogenization methods. In all cases, using such efl‘ective properties will underestimate the strain energy (and hence the stiffness) of the substructures. For more complicated structures, such as in Figure 3.9 (d), the loss of resolution results in some loss of accuracy. However, even in this case the two reduction processes result in essentially identical approximations of the strain energy and the reduced stifi‘ness matrices still provide a more accurate of the strain energy in the firll-scale structure than the effective properties obtained from classical homogenization. 3.6 Computing the Non-Periodic Reduced Stiffness Matrix of a Substructure The reduced matrix I? , obtained using a multi-resolution analysis of the displacements is not yet ready to be used as a super-element and assembled with other such matrices to build a model of a complex structure. This is because, while the area of the structure can be expressed a union of smaller areas that correspond to smaller substructures, Q = UCQC , the total stiffness matrix of the structure cannot be expressed as an assembly 62 of pen} interpret that cha: For this different Ieclmiqu a single surround substruct layout is 3.10. Fic of periodic stiffness matrices of individual substructures, K(Q)¢Zl_(c(flc) (sum C interpreted in the sense of assembly). However, I_( can be used to construct a matrix K that characterizes a substructure as a single entity in a non-periodic arrangement. For this we turn to the well-known computational scheme used in the solution of partial differential equations on arbitrarily shaped domains called the fictitious domain technique, (e.g., see Bakhalov and Knyazev [1], Glowinski et al [21]). The properties of a single substructure that is not part of a periodic arrangement can be obtained by surrounding the substructure by weak material of a sumciently low strength so that the substructure is essentially unaffected by its surroundings. A periodic arrangement of this layout is characterized by a fictitious substructure QFc =Qc U000 as shown in figure 3.10. n U n n U u Fictitious Substructure QFc Periodic arrangement of fictitious substructures Figure 3.10. Substructure surrounded by a weak fictitious domain 63 For exar fine—soak the fictiti Thus. th substruct to level substruct Using thi Where B Recall rh the COTTe OV'er the SubStrUct 3 See Sec. For example, if the material distribution within 0, is resolved by 2’ x 2’ pixelss, the fine-scale problem in “Pa involves 2‘]+1 x 21+l pixels and the material distribution in the fictitious substructure is defined as follows U €<<1 iajEQOC Thus, the fine-scale problem on a substructure is now extended to one on the fictitious substructure and the reduction procedure discussed eariier is performed from level J +1 to level j +1 to compute a reduced stiffness matrix corresponding to this fictitious substructure, denoted as H i +1. Using the approach in section 3.5.2.3, we compute the nodal matrix as —1 —F F F F K 1+1 :3 1+1 [11 1+1] 131-+1 (3.87) where Bf“ is defined as uWTBfHuh = fEfH (y) €(uw)s(uh)dy (3.88) 0 Recall that when this method was illustrated with periodic stiffness matrices and vectors, the corresponding material tensor used in the definition of the matrix B was the average over the substructure (as given in (3.58)). In the case of a fictitious substructure (i.e., substructure surrounded by a weak fictitious material) the average material tensor at scale j+l is defined as § See section 3.2 This is c original resulting The pres analysis the actuz both the? 0f the fit is definet state of c Where ( dOma'm E 058 (y) The Den in fight". )= p§3(y)E0 if yEQc E . ( 1'1 y . (3.89) This is done so that the terms corresponding to the non-periodic stiffness matrix of the original substructure (without the fictitious domain) are the only non-zero terms in the resulting matrix obtained using (3.87) and thus can be directly obtained. The presence of the fictitious domain in the reduction process using the multi-resolution analysis of displacements causes the terms that are associated with the boundary between the actual substructure and the weak fictitious domain to have properties that are due to both these materials. These edge eflects need to be compensated using a judicious choice of the filnction p§3 , which is the only adjustable parameter in this process. Here p53 (y) is defined such that a patch of 3 x 3 isotropic, homogeneous substructures reproduces a state of constant strain exactly. We assume that Rig (y) is of the form PiB(y)=a§B(y)p§ (y) (390) where 053 (y) is a correction factor to compensate for the presence of the fictitious domain and p; (y) E] for a homogeneous substructure. The piecewise constant function 033 (y) is of the form n 0‘58 (y) = Z 01:190ka (y) (3.91) k,l=0 The patch is subjected to three prescribed constant strain fields 4,551,531 as illustrated infigure3.ll. 65 3....1-C-I1-I-I I I ---l _____ ____ ___.. ... ’2: ::flP--r----’ --1--1--1 I I I I I nun-uI-II-n' I I I I I In... ~ ‘- ‘~ s f. --- -————— ———— ———— II-J L... ———— ———— ————— Figure 3.11. Constant pre-strains applied to a patch of substructures The coefiicients a“ are chosen to minimize the function 2 s” rc+| ”1 —e({” I (3.92) I 11 where e , e 51” and are the strains resulting from the imposed pre-strains. K5511 is a matrix of size 2(4n2 x4112) for n = 2} , fi'om this we extract a sub-matrix K3- of dimensions 2((n+l)2 x (n+l)2), in the case of approach 1]], these are the only non- zero terms in K 1+1 This matrix characterizes the behavior of a substructure QC as a single entity in a non-periodic setting when reduced fi'om level J to level j < J . This can now be used as a building block in an assembly of other substructures to construct a complete reduced model of a structure. 3.7 Computing Fine-Scale Stresses By Augmentation The solution to the reduced system of equations obtained using either of the earlier described methods represents the coarse behavior of the original system. This is useful and may even be suficient for certain design problems that involve only large-scale 66 behavior. However, it may be necessary to compute some small-scale effects at certain locations (substructures) of interest, such as local stresses at locations of possible stress- concentrations such as sharp corners or abrupt changes in geometry or material properties. In this section, a method to compute the small-scale stresses (where necessary) given the coarse solution is presented. This problem can be compared to that of computing the small-scale stresses in classical periodic homogenization. In the case of classical homogenization, the large-scale (coarse) solution is used to compute a coarse-scale strain field. Since this method assumes that the small-scale is microscopic, the coarse-scale strain field at any point in the domain may be thought of as a constant strain over the (infinitesimal) region occupied by the microstructure. The small-scale displacements in the case of classical homogenization are computed as the solution to the cell problem (3.77) for a constant pre-strain, where the value of constant pre-strain is the value of the large-scale strain at a location of interest. The important difi‘erence in the present scheme is that the small- scale problem is solved on a domain of finite size rather than on an infinitesimal cell. Thus the small-scale displacements cannot be computed simply by applying a constant pre-strain over the substructure, as in classical homogenization. The large-scale strains in the substructure are in general not constant. In the case of the model reduction using a multi-resolution analysis of the displacements, it is possible to obtain the fine-scale displacements at any substructure using a consistent augmentation procedure. This is not possible in the case of the multi-resolution analysis of the material. Let c be a particular substructure of interest whose small-scale stresses need to be computed and let uj E V; be the large-scale nodal displacement in this substructure 67 obtained fiom the solution of the reduced system. The multi-resolution analysis discussed earlier assumes periodicity of the associated fimctions in their respective domains and also deals with functions in a wavelet basis (V w ). Thus, in order to use the multi-resolution procedure to compute the fine-scale displacements, we first need to find an equivalent, periodic large-scale nodal displacement function and convert it into an . . - W equlvalent filnctlon m V j . 3.7.1 Computing the periodic large-scale nodal dlsplacements It is assumed that the fine-scale strain in a substructure, refine , can be expressed as the sum of a periodic fine-scale strain in an equivalent periodic substructure and a large-scale non-periodic strain, i.e., Efine = eljjne + Ex/olgrrse (3.93) where the subscripts P and NP denote periodic and non-periodic, respectively. This assumption in fact means that the non-periodic component of the strain is large-scale in nature. In addition, it is assumed that the large-scale displacement in the substructure, u j , can also be expressed as a sum of a periodic component and a non-periodic component, i.e., P NP This assumption is illustrated in figure 3.12, where a typical non-periodic displacement filnction is shown to be composed of a non-periodic component and a periodic component. 68 A periodic function (uf) 21> 619 A non-periodic filnction (uj) A non-periodic fimction (14f? ) Figure 3.12. Decomposition of a general non-periodic fianction into a coarse, The coarse-scale strain energy in the substructure can be expressed as UC=%S{Eje(uj)s(uJ-)dy (3.95) Using (3.94) this can be expressed as U6 =51)ij e(uf)e(uf)dy+%({ F3]- e(aj.VP)e(a§/P)ay +1 516("i)€("5ypldy QC (3.96) It should be noted that the augmentation process does not add any strain energy into the system, it merely computes the fine-scale displacements associated with the same coarse- 69 scale strain energy. It can be shown that the strain energy associated with a periodic stifl’ness matrix at a level J is the same as the strain energy computed from a consistently reduced stifi‘ness matrix at level J-k, i.e., the multi-resolution reduction scheme conserves strain energy. Thus, it would be useful to express the strain energy purely as the sum of one due to periodic displacements and another due to non-periodic displacements. This would require the last term in the right hand side of (3.96) to be equal to zero, i.e., the periodic component of the displacement needs to be orthogonal to the non-periodic component with respect to the energy inner product, (0,0) E , defined as ME = f Ej5(°)5(')dy (3.97) QC such that the strain energy due to the coupled (periodic and non-periodic) terms is zero. Thus, we look to decompose the displacement 111- according to equation (3.94) such that P P (uj ,u}’ )E = o (3.98) Using equation (3.94), we can write this as E =0 (3.99) 1731- e(uf)e(uf)dy =17ij €(uj)e(uf)dy (3.100) [u]? T D -u - = I? 5(uj)€(uf)dy (3.101) In the discretized form, equation (3. 100) becomes 70 [uflT Kin? = [uflT Djuj (3.102) where K]- is the periodic nodal stifiress matrix of dimensions 2112 x 2n2. The matrix D] (of dimensions 2112 x2(n+ l)2) could be thought of as one that transforms a vector of non-periodic nodal displacements into a periodic nodal body force. A candidate a? is such that it satisfies the system of equations I—( -u’-’ = D -u- (3 103) J J .l J ' This ensures that equation (3.102) is satisfied and that the total strain energy can then be expressed as the sum of the strain energy due to periodic displacements and that due to the coarse, non-periodic displacements. The periodic displacement vector u? obtained by solving equation (3.103) is one that corresponds to the nodal values of the displacement filnction in a periodic substructure. In order to apply the multi-resolution analysis to this filnction it needs to be expressed in an equivalent wavelet basis, this conversion is described next. 3.7.2 Conversion to a Wavelet Basis The conversion of a periodic filnction in the finite-element (nodal) basis Vh into an equivalent filnction in a wavelet basis V‘” is proposed using the reverse of the process described in section 3.4.2.3. According to this scheme a periodic nodal displacement a; can be converted into the equivalent wavelet displacement filnction 171- E V jw as — —l P 71 where 11,- is a reduced stiffness matrix of the substructure in a wavelet basis and B j is a matrix that transforms a given periodic nodal displacement into an equivalent body force in a wavelet basis. Once this coarse-scale displacement filnction in the wavelet basis has been computed, it can then be refined using the reverse of the multi-resolution analysis discussed in section 3.4.1 to obtain the fine-scale displacement filnction. This process is described next. 3.7.3 Periodic Mum-Resolution Refinement Recall the multi-resolution reduction scheme discussed in section 3 .4. 1. As expressed in (3 .28) a single stage wavelet transform applied to a system represented by the equations L j+1ll j+l = j+l yields the partitioned system T Li Cf 6,- _ ’1 r ~. T". Cj Aj ll] f] Thus, given the coarse-scale displacement ii]- , the detail components III of the displacement at scale j can be obtained from {if =Aflij—A;‘C§fij (3.105) where A j and C J- are as defined in equation (3.29). Under the assumption that the force is slow-varying (i.e., f = 0), the detail component is given as (U = —A;1C§fij (3.106) The displacement at scale j +1 can then be obtained using the inverse wavelet transform, W—l, as follows 72 _ —1 if uj+1=W (3.107) fir This process (equations (3.106) and (3.107)) is recursively repeated several times to obtain the displacement at finest scale J . Once the fine-scale displacement has been computed, the total fine-scale strain in the substructure is then obtained by adding the fine-scale periodic strain, €(uJ), to the non- periodic coarse strain, €(llj-VP ), i.e., using a, =e(uJ)+e(u}”’) (3.108) The fine-scale stresses can be obtained using this strain and the material distribution in the given substructure ”J 00:15.! (y)€J (y) (3.109) where E J represents the elastic tensor. 73 3.8 Numerical Examples This section presents numerical examples that illustrate and compare the schemes presented in this chapter. (Images in this dissertation are presented in color.) 3.8.1 Example 1 The first example considered is a square plate with circular perforations of various sizes subject to a uniform compressive load at the tip as shown in figure 3.13. Here the solid material is assumed to have a Young’s modulus of 0.91 and the weak (void) material with a Young’s modulus of 0.045 with a Poisson’s ratio of 0.3 in both. / E = 0.91 E = 0.045 A V I: Figure 3.13. Geometry and boundary conditions for example 1 This structure is modeled at a fine-scale that resolves all the features of the material distribution using a commercial finite-element software. The total number of degrees of freedom in the model is 74,498. The compliance of the structure is 0.671. The maximum displacement is at the center of the right edge and is of magnitude 1.44. Figure 3.14 74 shows the distribution of Von Mises stress from this fine-scale model. The maximum Von Mises stress in the structure is 2.69 near the center perforation. it Figure 3.14. Von Mises stress distribution using the fine-scale model of the structure in example 1. (Number of degrees of fieedom = 74,498) Figure 3.15 shows the assembly of 33 substructures used in constructing the reduced model of the structure. The substructure in the center is modeled at a fine-scale corresponding to a 64x64 pixel discretization and reduced to a scale corresponding to a 16x16 discretization. All the other substructures are also modeled at a fine-scale that 75 corresponds to a 64x64 discretization but arereduced to onethatcorresponds toan 8x8 discretization. It is noted that if all the substructures were to be modeled in the given fine scales then the displacement across the boundary between the central substructure and those surrounding it would not be continuous, i.e., additional constraints would need to be applied in order to enforce the continuity. However, the reduction is done is such a way that the displacement across substructure boundaries in the reduced models are continuous. The total number of degrees of freedom in the reduced model is 4802, i.e., it is approximately %6 of the fine-scale model. Reduction: 64x64—+16xl6 Reduction: 64x64 —» 8x8 Figure 3.15. Assembly of substructures for example 1 76 Figure 3.17. Von Mises stress distribution in reduced model using displacement MRA 77 ' 0.8 0.6 0.4 m Figure 3.18. Von Mises stress in substructure (A) using the reduced model .25 Figure 3.19. Von Mises stress in substructure (A) after augmentation l' I. Mn = 1.053-01 Figure 3.20: Detail of the Von Mises stress near substructure (A) obtained using the fine-scale model Error Max. om Error 2.69 Table 3.1: Main results for example 1 The reduced model using an MRA of material yields a compliance of 0.667 and a maximum displacement of 1.43. The reduced model using the MRA of displacements yields a compliance of 0.675 and a maximum displacement of 1.44. These values are very close (less than 1% difference) to that obtained using the fine-scale model. Figures 79 3.16 and 3.17 show the distribution of the Von Mises stress using reduced models constructed using the material MRA and the displacement MRA respectively. The maximum Von Mises obtained from the reduced model using material MRA is 2.21 and that from the reduced model using displacement MRA is 2.25. These values are almost 17% less than that obtained using the fine-scale model. Figure 3.18 shows a more detailed look at the stresses in the central substructure (A). The augmentation procedure is performed on the solution in this substructure to compute the fine-scale stresses. The augmented stress distribution is shown in figure 3.19. The maximum stress obtained afier the augmentation is 2.57, only 4% ofi‘ from the fine-scale. It is noted that the coarse-scale results obtained from the two reduction procedures are in general not too difi‘erent. However, in the case of displacement MRA, it was possible to carry out an augmentation procedure and compute the stresses with a greater accuracy at the desired location. 80 3.8.2 Example 2 The next example considered is the structure shown in figure 3.21. As before, the solid material has Young’s modulus of 0.91 and the weak material (void) has Young’s modulus of 0.045. Poisson’s ratio is 0.3 in both cases. f=2 E=0m E = 0.045 V A V7 Figure 3.21. Geometry and boundary conditions for example 2 The fine-scale model of the structure consists of 163,840 degrees of freedom and is computed using a commercial finite element sofiware. The compliance of the structure is 6.96 and the maximum displacement is 7.66 and is observed at the center of the top edge. Figure 3.22 shows the distribution of Von Mises stress in the structure. The maximum stress (as expected) is observed near the sharp comers and the magnitude of the maximum stress obtained is 4.72. 81 > 4.04a+00 < 411494-00 < 3.37e+00 < 2.7De+00 < 2,029+!!!) < 1,353+00 < 83740-01 < 4280-05 Max = 4.729+00 Min = 4280-05 Figure 3.22: Von Mises stress distribution fi'om fine-scale model of the structure with 163,840 degrees of freedom Figure 3.23 shows the arrangement of 50 substructures used in constructing the reduced model of the given structure. The substructures in the outer periphery of the structure with large perforations are modeled at a fine-scale that corresponds to a 64x64 uniform discretization and reduced to an equivalent of a 16x16 discretization. The other smaller substructures are also modeled using the same fine-scale discretization but are reduced to an equivalent of an 8x8 discretization. The total number of degrees of fieedom in the reduced model is 10,240. 82 Reduction: Figure 3.23. Arrangement of substructures for example 2 The reduced model using material MRA predicts a compliance of 6.7 and a maximum displacement of 7.35. The compliance and maximum displacement obtained from the reduced model using displacement MRA are 6.7 and 7.4 respectively. These values are very close to those obtained fiom the fine-scale model. Figures 3.24 and 3.25 show the Von Mises distribution obtained from reduced models build using the material MRA and displacement MRA respectively. The maximum Von Mises stress from the material MRA reduced model is 3.19 and that from the displacement MRA reduced model is 3.22. These are approximately 30% less than the maximum stress obtained from the fine-scale model. Figure 3.26 shows the coarse-scale stress in substructure (A). The fine-scale stresses in (A) are computed by the augmentation procedure and are shown in figure 3.27. The maximum stress after augmentation is 4.2. This is still ofi~ fiom the fine-scale result by about 11%. 83 Figure 3.24. Von Mises stress distribution from a reduced model using material MRA Figure 3.25. Von Mises stress distribution from a reduced model using displacement MRA Figure 3.26: Von Mises stress in substructure (A) from the reduced model using displacement MRA F45 4 3.5 Figure 3.27: Von Mises stress in substructure (A) alter augmentation < 4280—05 Max 2 ave-+00 Min = 4280-05 Figure 3.28: Detail of the Von Mises stress in a region around substructure (A) from the finescale model Max 7.66 7.35 7.4 Table 3.2: Main results for example 2 In this case, the maximum stress even afier augmentation has greater than 10% error as compared to the fine-scale. This is due to the fact that sharp corners are locations of high stress gradients and the discretization level at the coarse-scale is not enough to capture 86 these gradients. Locations of high stress gradients usually need to be modeled at a much finer discretization than the other regions 3.8.3 Example 3 In the last example the material distribution is as shown in figure 3.29. The bottom edge of the frame is clamped and the top edge is subject to a uniform unit load. Figure 3.29. Geometry and boundary conditions for example 3 The compliance of the structure obtained from a fine-scale model using a commercial finite element software is 0.765. The maximum displacement is 1.86. Figure 3.30 shows the distribution of the Von Mises stresses obtained fiom a fine-scale model of the structure. The maximum Von Mises stress is 0.83. The assembly of substructures used in the construction of a reduced model of this structure is illustrated in figure 3.31. The solid rim is not reduced and is modeled in accordance with the continuity requirement across the substructure boundaries. The interior of the structure is modeled using 12 substructures modeled at a fine scale corresponding to a 64x64 mesh and reduced to a 8 x 8 mesh. 87 < 3.110-03 Max = 8309—111 Mr) == 3.110-03 Figure 3.30. Von Mises stress distribution from a fine-scale model Rim Not Reduced Figure 3.31. Arrangement of substructures for example 3 88 Figure 3.32: Von Mises stress distribution fi'om a reduced model using material MRA Figure 3.33: Von Mises stress distribution fiom a reduced model using displacement MRA F025 “0.2 '0.15 Figure 3.34: Von Mises stress in substructure (A) from the reduced model using displacement MRA H0. 8 ' 0.7 ‘0.6 ‘0.5 : 30.4 0.3 Figure 3.35. Von Mises stress in substructure (A) alter augmentation 90 Figure 3.36: Detail of the Von Mises stress in substructure (A) obtained from the fine-scale model The reduced model using material MRA yields a compliance of 0.724 and a maximum displacement of 1.71. The corresponding results fi'om the reduced model using displacement MRA are: compliance of 0.756 and a maximum displacement of 1.90. Figures 3.32 and 3.33 show the distribution of Von Mises stress obtained fiom the material MRA model and the displacement MRA model respectively. The maximum Von Mises stress from the material MRA reduced model is 0.61 and that from the displacement MRA model is 0.69. These values of maximum stresses are observed at the bottom comers. The maximum stresses from the reduced model are approximately 20- 30% less than that obtained from the fine-scale model. The augmentation procedure is then carried out on substructure (A) and the resulting mardmum stress computed. Figures 91 3 .34 and 3.34 show the coarse-scale stress in substructure (A) and the augmented stress respectively. The maximum Von Mises stress afier augmentation is 0.817, only 2% off from the fine-scale result. Max Disp % Error Compliance % Error Max. om % Error Fine-Scale 1.86 - 0.77 - 0.83 - Material 1.71 8% 0.72 < 7% 0.61 27% MRA Displacement 1.9 2% 0.75 < 3% 0.69 17% MRA ' Augmented - - - - 0.82 < 2% Table 3.3: Main results for example 3 It is noted that the maximum stress fi'om the reduced models are observed at the bottom comers of the outer rim. However, in the fine-scale, the bottom corner stress though considerably large, is not the maximum. The maximum stress in the fine—scale model is observed in the interior substructures (A). In the reduced models, the stresses in the substructure (A) are much less than that in the fine-scale model (compare figures (3.34) and (3.36)). After augmentation the maximum stresses are only 2% less than the fine- scale result. But, the augmentation over estimates the locations of the maximum stresses, i.e., the locations of maximum stresses afier augmentation are spread out over a larger area than as computed from the fine-scale result. Nevertheless, as seen by comparing figures (3.34) and (3.35), there is a tremendous increase in the accuracy of the magnitude of maximum stress computed in substructure (A) using the reduced model after augmentation. 92 Chapter 4 Multi-Scale Layout Optimization Of Structures In this chapter, some strategies for the optimal layout design of structural systems using materials with finite-scale heterogeneities are presented. The standard methods that are commonly used to solve these problems are either homogenization-based techniques that involve materials with infinitesimal heterogeneities (microstructures) (refer Bendsoe and Kikuchi [2], Diaz and Bendsoe [12]) or methods that use fictitious material models (see Bendsoe [3], Rozvany et al [33,34]). Structural systems involving many finite-scale heterogeneities yield very large systems of equations that are not suitable for the kind of iterative solution schemes that are required in the case of optimization problems. The main goal in this chapter is to propose strategies that incorporate the model reduction techniques discussed earlier into a problem of layout optimization of structural systems where finite-scale features can be accounted. 93 This chapter is arranged as follows: section 4.1 gives a brief introduction to the problem of layout optimization of structures and the standard techniques used to solve these problems. The next section, section 4.2, introduces the problem of optimizing the layout with finite-scale materials. There, a simplified version of this problem using perforations as a prototype for heterogeneities is presented and the associated sensitivity analysis is outlined. Section 4.3 discusses the formulation of compliance minimization problems using perforated substructures. Section 4.4 illustrates the solution of compliance minimization problems using fixed layouts of substructures. The next section, section 4.5, deals with the dependence of the optimal layout on the arrangement of various sizes of substructures. Here, the formulation and solution of compliance minimization problems with varying layouts of substructures are discussed. Finally, numerical results are presented in section 4.6. 4. 1. Background: Topology Optimization of Structures Topology optimization problems in general seek to find the optimal layout of a structure in a prescribed design space using a given amount of material, subject to constraints on the response of the structure under prescribed loading conditions. Typical objective firnctions involve: mean compliance, total mass, eigenvalues. Typical constraints include volume and stress (e.g. see Suzuki and Kikuchi [36], Diaz and Kikuchi [13], Duysinx and Bendsoe [l6], Haber et al [23]). A typical topology optimization problem can be expressed as a problem that seeks an optimal distribution of the elasticity tensor Eijkl (x) over the design domain, 0 , by writing 94 where Em is a reference tensor and x (x) is an indicator filnction for the part 9'" of O that is occupied by material, i.e., 1 if xefl’" x(x)= 14.2) 0 if x€Q\Q’" An approach to solving such optimization problems using finite elements results in each point, x , in the domain having the discrete choice of having material or no material, i.e., this distributed parameter optimization problem is formulated using a discrete valued parameter firnction. The solution of this type of problems requires the use of discrete optimization algorithms. However, such an approach would be unstable with respect to the choice of elements and the discretization mesh, as the distributed problem, in general, does not have a solution unless composite materials are introduced (see Kohn and Strang [25], Murat and Tartar [29]). An alternative solution to this problem was introduced by Bendsoe and Kikuchi [2]. According to this scheme, rather than determining the mixture of two materials at the macroscopic level, the mixture is allowed to occur at an infinitesimal scale. This leads to a problem formulation using material with a microstructure, i.e., a material with microscopic perforations of different sizes controlled by the introduction of a parameter called effective density (volume occupied by material in a characteristic unit-cell), which may vary continuously fi'om 0 to l, the two limiting cases being the void and a solid material and the intermediate densities correspond to a composite material. The relation between the effective density and the material tensor, Erjkl (x), is determined through the use of a homogenization method, 95 where the material distribution at the microscopic level is used to detemrine the effective properties of the material at the macroscopic level. Figure 4.1 shows a structure that is composed of a periodic composite microstructure. \ microstructure characteristic \cill Figure 4.1: A structure with a composite microstructure Another approach to solve the problem is by introducing an artificial density function 11(x), x69, 0l and defining the elasticitytensoras Ey'kl (x) = [#(x)lp 51ij (4-3) This model is known as the Solid Isotropic Material with Penalization (SIMP) model (see Rozvany et a1 [33, 34]) and yields results with fictitious materials of low stifliless for intermediate densities for sufiiciently large value of the penalty factor, i.e., it forces the material at each point to have an efi‘ective density that is either close to being solid or void. Figure 4.2 illustrates a typical topology optimization problem and its solution. 96 e—m—s :3 Figure 4.2: A typical topology optimization problem and solution 4. 2. Optimal Design Using Macro-Scale Heterogeneities The work presented here is concerned with obtaining optimal mixtures of two materials (solid and void) at finite scales (macro-scale), i.e., scales that may even be comparable to the dimensions of the structure. Furthermore, it is assumed that the boundary of the desired structure is known a-priori. This is ditferent from the case of standard topology optimization problems where finding the boundaries is part of the problem. The problem is now reduced to one of finding the optimal arrangement of finite-size heterogeneities on a prescribed domain. The heterogeneities may be of any type; however, to illustrate the idea, this discussion deals with perforations as the macro-scale heterogeneities. A typical structure with finite-scale perforations is illustrated in Figure 4.3. The Optimization problem seeks to find the locations and the sizes of various perforations in the given domain. The implementation of this kind of an optimization problem using standard techniques such as finite element methods requires the domain to be suitably discretized and this usually involves adapting the geometry of the structure to a conforming mesh. Since the layout of the material keeps changing during the course of the optimization process, repeated re-meshing of the structure would be required at each 97 stage of the optimization problem, making this process not only computation-intensive but also difficult to automate. Figure 4.3: A typical structure with macro-scale heterogeneities (perforations) One possible approach to solving such problems is proposed here, using the sub- structuring idea described in the previous chapter. The proposed method seeks to build structures in a given design domain 0, that can be expressed as the union of regular substructures, DC, in such a way that all the perforations lie in the interior of the substructures, i.e., n = U12, and 611, map = e (4.4) C where 0” represents the perforated portions of the domain. The substructures can be of different sizes and in general each substructure can include more than one perforation. One could construct libraries (databases) of such substructures (stifi‘ness matrices) of various sizes and types of perforations. The optimization problem would then be approximated into one that seeks to build a structure as an assembly of an optimal 98 selection of substructures chosen from a library of perforated substructures of various sizes. As a result of the assumption on the design domain (4.4), the optimal structure obtained as an assembly of selected substructures cannot have overlapping perforations and the geometry of the perforations are limited to those in the pre-computed library of perforated substructures. An example of such a structure is illustrated in Figure 4.4. O Figure 4.4: A structure built from an assembly of substructures with circular perforations 4.2.1 Building a model using reduced substructures Here, some relevant features of the construction of a model (stifliless matrix of a structure) using reduced stiffiless matrices of substructures are presented. Consider a structure assembled from square substructures of various sizes and with various sizes of circular perforations as shown in Figure 4.5. The structure is represented using five difi‘erent substructures of three difl‘erent sizes, L , 2L and 4L. Let Jc be the level of the discretization required to resolve the material distribution in a substructure c, such that 99 . . . L the discretization corresponds to a umforrn mesh wrth spacrng equal to Sc = Tc , where 2 C LC is the dimension of the substructure. 10L 1 4L 1 ‘2L. 0 O O 0 if? J=7 ' " ( O O 00 1:5 c=s Q Q J=5 8L 0 Q 0 021 F) F, O O V A J 5 0:4 00 :5 4H— 00 ’ #35 c=5 c = 2 Figure 4.5: Assembly of substructures Substructures of the same size (e.g., l, 2 and 3, 4) may require a different resolution scale, depending upon the size of the perforations in them. In the figure, it is assumed that substructures 2, 4 and 5 are discretized at the same level Jc = 5. However, the spacing of the degrees of freedom in each of these substructures is: S2 = -‘%= g, 2 2L L L . . . S4=——=— and S5=—=— While substructures 1 and 3 are discretized at 25 16 25 3 different levels, J]: 7 and J3 = 6, the spacing in these substructures is the same, S1=%=% and S3 -£=%. Notice that in the fine scale problem when Sc is 2 26 difi‘erent in two adjacent substructures, a one pixel — one element finite element discretization is in general not confirming and would require additional constraints in order to enforce continuity across substructure boundaries. An example of such a non- conforming mesh is shown in Figure 4.6. Indicates non- conforming boundary Figure 4.6: A non-conforming assembly of discretized substructures The reduction of each substructure is performed such that the displacement across boundaries of adjacent reduced substructures is continuous, thus avoiding additional constraints to enforce inter-substructure compatibility. This requirement leads to the following condition on the reduced discretization level of a substructure, A, =Lc/zjc =A (4.5) for c=l,2,---,Nc. For the substructures shown in Figure 4.5, this implies that jl = j; = j5 +2 and j3 = j4 = j5 +1. It is emphasized at this point that these relations do not determine the finest scale at which a substructure is to be modeled. 4.2.2 Constructing libraries of perforated substructures Here, the construction of libraries (databases) of stiffness matrices and material distributions associated with perforated substructures is discussed. A particular type of (pararneterizable) perforation is chosen and a set of material distribution filnctions 101 (k) . . . . {p5} associated with a discrete set of desrgn vanables in a substnrcture of srze L are constructed at a suitably chosen fine scale J . If a reduction based on the MRA of material distributions is to be used, then the library consists of reduced material . . . c (k) . . . . . . drstnbutlons {pf} at various levels 1, obtalned usrng the process described in section 3.2. If an MRA of displacement is used, the library consists of reduced stifi‘ness matrices c (k) , , . . {Kj} (sections 3.3, 3.4), at levels j=J,J—l,---,3 (1:3 13 the smallest level possible when using D6 wavelets). This library is comparable to the library of effective material tensors used in some topology optimization problems (see Suzuki and Kikuchi [36]) and needs to be constructed only once and can be re-used in other problems with little difi'rculty. The resolution and the number of different perforation diameters in the library depend on the choice of the finest discretization level, J. An interpolation scheme is then used to approximate K3- as a continuous function of the design variables in the specified range of allowable values the design variables can take. To illustrate, Figure 4.7 shows the largest, intermediate and smallest perforation that can be modeled at resolutions corresponding to scales J = 4 , J = 5 and J = 6. The spacing between nodal degrees of fieedom would be the same after reduction, regardless of the starting level. Thus the choice of J only affects the offline computations and not the computations within the optimization process. 102 a = 0.96875 a = 0.5 a = 0.03125 Figure 4.7: Substructures with circular perforations for various finest levels 4.2.3 Sensitivity Analysls The sensitivity analysis using substructures with centered circular perforations is illustrated here. The analysis can be extended to substructures with other kinds of perforations easily. In the case of substructures with circular perforations, the design variable is the diameter of the perforation in a substructure given by the following relation dc = aLc, 0 S a 3 am (4.6) where LC is the dimension of the substructure and ozmax is a prescribed bound on the size of the perforations. When the reduction of the substructures is based on an MRA of material distribution, entries in the library are pixel values of pi, obtained using a 103 (J — j )-level reduction of the fine-scale material distribution firnction p J (a(k)). The pixel values (1);)” are obtained fiom an interpolation of the entries in the library. The effective stiffness matrix [(3- (a) is then computed using these interpolated reduced material distribution functions. The gradients of the stifl‘ness matrix with respect to the design variables is computed as BKC- 1 6K0- 1 n J _ J _ Z 0 6 LC 0 c k,l=1 ’ a C. p] is computed using the interpolation filnction and k0 is the element where p' = stifl‘hess matrix of a solid element (reference). When the reduction is based on an MRA of displacements, the effective stifi'ness 1k) matrix K;- (o) is computed by interpolation in a of the entries in the library {K5} . This interpolated firnction is similarly used in the computation of the gradients of the stifi‘ness matrix with respect to the design variables, Bdc LC Ba ' 6K6- 1 where the derivative with respect to a , — , is computed using the interpolation a firnction. 104 4.3. Compliance Minimization Problems While the proposed procedure can be applied using any geometry of (parameterizable) perforations, here the formulation of the compliance mininrization problem is illustrated for two particular cases: substructures with centered rectangular perforations and substructures with circular perforations. It is a common practice in standard topology optimization problems to consider characteristic cells with rectangular voids (see Suzuki and Kikuchi [36]). This is usually done so that firlly void regions can be modeled by letting the perforation extend to the whole cell (void cells), see Figure 4.2. In the present case, the exterior boundary of the design domain is known a-priori and we are only interested in the distribution of the perforations inside this boundary. The choice of the shape of the perforation in this case may be dictated by other considerations such as case of implementation, etc. Figure 4.8 shows two typical perforated substructures and the associated parameters. Figure 4.8: Substructures with rectangular and circular perforations A compliance-minimization problem for a library of substructures of centered rectangular perforations can be written as follows: For each designable substructure c, find comma, that 105 subject to 21L: — acbc) gymeas (Q) C aIIIinLc S ac COS 66: b6 COS 6c 5 LC (49) aminLc S ac sin 90, be sin BC 3 LC 0 3 BC < 1r/2 Ku = f The design variables in this case are the dimensions of the perforations and the orientation of the perforations with respect to the substructures. Similarly, the optimization problem for a library of substructures of centered circular perforations can be expressed as follows: for each designable substructure c , find dc that minimize CéfTu subject to E C OSchamec Ku=f 2 d Lc—7r—46— _<_7meas(fl) (4.10) The design variable in this problem is the diameter dc of the perforation in each substructure. In these problems, f is a prescribed load vector, 7 is a prescribed volume fraction, 0 < 7 <1, and the bounds am“ and amax on the size of the perforations are given data. Compliance minimization problems (or any other problems) solved using the proposed scheme can be divided into two broad categories: fixed and variable discretization of the domain. In the problems of the first kind, the discretization of the design domain (2 into substructures is prescribed a-priori and remains unchanged throughout the optimization, i.e., only the size of the perforation in each substructure 106 needs to be determined. Figure 4.9 illustrates a possible starting and final layout of perforations in an optimization problem using a fixed layout of substructures. Figure 4.9: Optimization using fixed layout of substructures In problems of the second kind, the discretization of the design domain into subsnuctures varies at each stage of the optimization, i.e., at each step the size of each substructure and size of the perforation are determined. Figure 4.10 illustrates the optimization using variable layouts of substructures. However, in both types of problems, since (consistently) reduced substructures are used the number of degrees of freedom in the reduced model of the structure is always the same regardless of the number of substructures used or their sizes. .. 1' {-11 Initial layout Figure 4.10: Optimization using variable layouts of substructures 107 4.4 Optimization Using Fixed Layouts 0! Reduced Substructures In this type of problems the size LC of each substructure is prescribed a-priori, i.e., the optimization problem seeks to find only the sizes of the perforations within a prescribed assembly of substructures. Furthermore, from (4. 5), in order to guarantee compatibility across substructures, LC is of the form, Lc=2m0L (4.11) where me is a non-negative integer and L is a prescribed dimension in the problem, see Figure 4.6. The exponent mc may vary fi'om substructure to substructure. For simplicity, it is assumed that the geometry in all the perforated substructures is resolved by a discretization of the same level, J . Thus, the fine scale at any substructure is Sc = LC /2J (4.12) i.e., the material in each substructure is modeled using 2'] x2] (pixels) elements. In the reduced system, the degrees of fi'eedom are spaced such that displacements are continuous across boundaries. This determines the reduction level jc for each substructure (see equation 4.5). If the nodal spacing after reduction is A = L / 2”'0 (4.13) for some positive integer m0 , then a perforated substructure is reduced from level J to level jc =m0+mc (4.14 Clearly, m0 must be such that for all substructures jc 3 J . 108 A possible (fixed) layout of substructures is shown in Figure 4.11. Three sizes of substructures are used in this layout, L , 2L and 4L. In order to satisfy the continuity requirements across substructure boundaries the reduction levels in the substructures are relating according to the following rule: if the substructures of size 4L are reduced to a level j, then the substructures of sizes 2L and L are reduced to levels j —l and j—2 respectively. The lowest allowable level for any substructure is j = 3 when D6 scaling functions are used. Thus, the substructures of size 4L can be reduced at the most to a level j = 5 in order to satisfy the previous criterion. 4 16L —__§ l 16L —>l|<——>I|<— 2L L k—H 4L Figure 4.11: A design domain and a possible layout of reduced substructures 109 4.5 Optimization Using Variable Layouts Of Reduced Substructures 4.5.1 Perforated Substructures and Layout Dependency In problems with variable layouts of substructures, one needs suitable criteria to choose between assemblies of smaller substructures and large substructures. This necessitates a comparison between the stiffness properties of a single perforated substructure with an assembly of smaller substructures keeping the total amount of material in both equal. Here, an assembly of four substructures, each of size LXL, is compared with a single large substructure of size 2L x 2L . Two tests are performed: a prescribed traction on the boundary test and a prescribed constant pre-strain test. The first test involves computing the resulting strain energies for various arbitrary tractions applied on three edges and constraining all the degrees of freedom on one edge, as shown in Figure 4.12. 11“ HF Q J; U9: g 1/2‘ \ x a 311‘ Pgan Up“ I‘gBQ Figure 4.12: Boundary traction test In a structure made up of several substructures subject to some prescribed loading, this test seeks to answer the question: whether removing a large substructure with a single perforation and replacing it with four small substructures (of half the size and with the amount material being the same in both) would make the structure stiffer or alternatively, 110 removing a patch of four small substructures and replacing them with a single large substructure makes the structure stifi‘er. For the same tractions, the stiffer substructure would result in a lesser defamation and hence the resulting complementary strain energy of a stifl‘er structure would be lesser. The random tractions are chosen arbitrarily fi'om a uniform random distribution in lRlo’l]. The resulting strain energies in each case are shown in Figure 4.13. The results show that a configuration of four substructures, of size L each, is stifi‘er (lesser strain energy) that a substructure of size 2L with a single perforation and having the same amount of material. 33 5 G: l-I-l E l: (I) 8 E Q E O 0 who Random Boundary Tractions Figure 4.13: Comparison of strain energies in substructures with different number of perforations for various boundary tractions The second test involves the application of constant pre-strains of the form, 0 0 . . . . e = ,nE[—l,l], (see sectron 3.x). For the same prescribed strain a strfi‘er 0 111 structure would result in higher strain energy. As before, the resulting strain energy due to these pre-strains as obtained from a single substructure of size 2L and a patch of four substructures of size L are shown in Figure 4.14. It can be seen fiom the figure that the strain energy in the patch of four smaller substructures is always greater than that in the large substructure and thus conforming that the patch of four substrucmres of size L each is indeed stifi‘er than a single substructure of size 2L for the same amount of material in both. Strain Energy Figure 4.14: Comparison of strain energies in substructures with difl‘erent number of perforations for various constant pre-strains It should be noted that these results are local to a particular substructure, i.e., they assume that the change in the layout of a substructure does not alter the overall stress distribution in the entire structure. However, this may not be true in general. Nevertheless, this result may be used as a criterion in an updating scheme to determine 112 the optimal arrangement of substructures that result in the stifi‘est structure. This needs to be done iteratively and is discussed in the following sections. 4.5.2 A Dlvldlng Approach to Optlmlzatlon with Variable Layouts In this approach the structure is initially built entirely using as many large substructures as possible and each update of the layout corresponds to a subdivision of a large substructure into four smaller ones. This process is illustrated in Figure 4.15. Figure 4.15: Illustration of the possible evolution of the layouts using the dividing approach Here, additional constraints such as a bound on the total perimeter of the perforations (PM) or a limit on the number of substructures of a particular size are introduced in order to have a suitable stopping criterion for the layout updating. This process may be summarized as follows. (Here, the method is illustrated using a bound on the perimeter of the perforations as the stopping criterion.) (0). Start with an initial layout of substructures {LC } , c = lto N0 , with perforations of suitable sizes such that the volume constraint is satisfied, e.g., dc =2Lc 1:1 7r 113 where dc is the diameter of a perforation, 7 is a prescribed volume fraction and LC is the size of a substructure. (l). Solve a fixed-layout optimization problem (4.10), i.e., for c =1toN0 , find optimal dc (2). Sort the substructures according to decreasing order of magnitude of strain energy in each, such that ungiui Zunguj, if i aminL (if (P+7rD)SPmax dc = dN+1 = dN+2 = dN+3 = Nlhw|b Lc = LN+1 = LN+2 = LN+3 = N=N+3 P=P+7rD Kendif kendif where amin is a suitably chosen minimum (relative) size of a perforation (e.g., 0.05 ). Else, go to step 4 (break loop). (4). Solve a fixed-layout optimization problem (4.10) to determine the optimal sizes of perforations, i.e., d , 0 =1 to N (where, N > No) The resulting layout may be used again as a starting layout and the entire process is repeated recursively till the stopping criterion is met, i.e., any further division results in a violation of the perimeter constraint. A method that directly follows from the reverse idea of this approach is discussed next. 4.5.3 A Merging Approach to Optimization with Variable Layouts In this approach, the starting layout consists of purely small substructures and each update of the layout corresponds to merging four small substructures to create a bigger substructure. This is illustrated in Figure 4.16. 115 Figure 4.16 Illustration of the possible evolution of the layouts using the merging approach This approach is more flexible in terms of feasible starting layouts. In the previous approach, there is only one way to divide a substructure into four equal smaller substructures. However, in this approach, each substructure may have up to four ways of merging with neighboring substructures to make up a large substructure. Thus, the number of possible layouts of substructures in this approach is a lot more than that using the earlier approach. This approach may be summarized as follows. (As before, the method is illustrated using a bound on the perimeter as the stopping criterion.) (0). Start with a layout of small substructures {LC}, c =1toNo , with perforations such that the volume constraint is satisfied, e.g., dc =2Lc 1.17. 71' where dc is the diameter of a perforation, 7 is a prescribed volume fraction and LC is the size of a substructure. ( l). Solve a fixed-layout optimization problem (4.10), to find the optimal dc for C=1tON0 116 (2). Group neighboring substructures (of the same size)§ as possible candidates to be merged. A substructure can be a part of up to four groups depending on whether it is located in an edge or in the interior of the domain. Create a table G , of dimensions NG x 4 , that consists of all possible groups of four neighboring substructures, where NO is the number of groups. (3). Sort the groups according to increasing magnitude of total strain energy of the substructures in the groups Set P=Z7rdc and N=No C (4).For each group in the sorted table, g =1 to N6 , if the total perimeter exceeds the allowable limit and if the maximum deviation of the perforation size fiom the mean is within a prescribed bound (and the perforations are of significant size), merge the four substructures in the group, i.e., [if P> P11m {d}g ={d,-,dJ-,dk,d,} (if mean({d}g)>aminLg and ::(({;};))—1 <6 D=fl£+fi+£+fi) P=P—tr((d,- +dj +dk +d,)—D) N=N—3 Kendif \endif ’ relevant for recursive restarts 117 where, {i, j, k,l} is any group in G , am“ is a prescribed bound on the minimum relative size of perforations, L8 is the size of the substructures in the group, 6 is a prescribed bound on the maximum deviation of the perforation sizes allowable in a group to be merged, e. g., 30% and D is the new size of the merged perforation. Each such merging reduces the number of substructures by 3 and the reduction in the perimeter depends on the size of the perforations merged. (5). Solve a fixed-layout optimization problem (4.10), to find the optimal dc , for c=1toN. Here, N Figure 4.33: Problem description for example 5 As before, this problem seeks to build the optimal structure in the given design space using perforated substructures of two sizes L and 2L. Here, to illustrate other possible stopping criteria for the layout updating, a constraint on the number of substructures of size 2L is prescribed as 11. This is used as the stopping criterion in the merging and the dividing approaches to determine the optimal arrangement of substructures. Merging Aggmach The initial layout of substructures for the merging approach and the optimal structure using this layout are shown in Figure 4.34. The compliance of the structure is 25.6. The total perimeter of the perforations is 325. 132 III III Figure 4.34: Initial layout for the merging approach and optimal structure using this layout for example 5 A sequence of layouts obtained using the merging approach is shown in Figure 4.35. The arrangement of substructures afier successive merging and the corresponding optimal structure using that layout is shown in Figure 4.36. The compliance of that structure is 25.4. The total perimeter of the perforations is 250. Although this perimeter is not used at this point to determine the termination of the merging process, it is used later to compare the optimal structures obtained using the dividing approach with two possible termination criteria, i.e., number of substructures and total perimeter. 133 Figure 4.35: Sequence of layouts obtained using the merging approach for example 5 134 Figure 4.36: Optimal arrangement of substructures and the corresponding optimal structure obtained using the merging approach for example 5 Dividing Approach The starting layout for the dividing approach and the optimal structure for this layout are shown in Figure 4.37. The compliance of this structure is 29.9 and the total perimeter of the perforations is 177. Figure 4.37: Starting layout of substructures for the dividing approach and the corresponding optimal structure for example 5 135 000- “““ 0000 ...... O. r- O. OOOoo- one... 9. '- O. ......ooo eeeeeee 00.9.09... 0... OOOO OOOOOOOOOOO.... --tosooooo.oo . .QOOOOO.....O 00...... 0:00.00. 000000;; Figure 4.38: Sequence of layouts obtained using the dividing approach for example 5 136 A sequence of steps using the dividing approach is shown in Figure 4.38. The dividing process continues till the number of substructures of size 2L becomes equal to the prescribed value of 11. The final layout in Figure 4.39 is then used as the starting guess for a fixed-layout optimization problem. The layout of substructures and the corresponding optimal structure are shown in Figure 4.40. The compliance of this structure is 25.3. 0- - 000 . COO. .0 - Figure 4.39: Optimal arrangement of substructures and the corresponding optimal structure obtained using the dividing approach with prescribed number of substructures of size 2L for example 5 It can be seen that the optimal layout of substructures and hence the optimal structure obtained using the dividing approach when the number of substructures of size 2L are prescribed is the same as that obtained using the merging approach. However, if instead of the constraint on the number of substructures, a constraint on the perimeter of the perforations (Pmax = 250) is used as the termination criterion in the dividing approach, the dividing process terminates at step 24 shown in Figure 4.39. (The perimeter constraint prescribed is total perimeter of the optimal structure obtained using 137 the merging approach.) The number of substructures of size 2L in this layout is 17. This layout is then used in a fixed-layout optimization problem to determine the optimal structure for this arrangement of substructures. The arrangement of substructures and the corresponding optimal structure is shown in Figure 4.40. The compliance of this structure is 25.7. This is higher than the previously obtained optimal layout using the constraint on the number of substructures of size 2L . 0 ° ' “.OOOOOOOO . O a e . Figure 4.40: Layout of substructures and corresponding optimal structure obtained using the dividing approach with a perimeter constraint for example 5 138 .J Chapter 5 Concluding Remarks 5.1 Summary An approach for the analysis of structural systems in linear elasticity using an assembly of consistently reduced substructures (smaller components) was presented. Two strategies for constructing reduced models were introduced: one based on a multi— resolution analysis of material distributions and the other based on a multi-resolution analysis of displacements. For relatively simple geometries, it was seen that the two methods produce results that are not too different in the coarse-scale. However, using an augmentation procedure it was shown that some parameters such as stresses could be computed more accurately at certain desired locations in the case of the approach using a multi-resolution analysis of displacements. 139 The concept of structural analysis using reduced models was then applied to the problem of layout optimization of structures in which finite size heterogeneities of multiple scales could be accounted. This involved constructing libraries of reduced substructures with suitable heterogeneities (illustrated using perforations here) of various sizes. The dependency of the optimal layout on the arrangement of substructures was discussed. Two approaches for optimization of layouts with finite size heterogeneities were indicated: one using fixed discretization of the design domain into substructures and the second in which the arrangement of the substructures was part of the problem. 5.2 Conclusions 0 The concept of analysis of large structures using assemblies of reduced substructures proved very convenient and successful in problems such as layout optimization of structures. 0 A particular advantage is derived from the fact that the reduction of a given substructure needs to be performed only once but the results are reusable. This facilitates the construction of libraries of pre-computed reduced stifliress matrices to be used as needed. 0 The use of reduced substructures of difl'erent sizes and reduction levels accomplishes the goal of accounting for heterogeneities of different sizes and scales. 0 As the large-scale parameters, such as compliance, are almost the same when computed using either of the two reduction schemes the computationally 140 expensive MRA of displacements need only be used in cases where certain small- scale parameters such as local stresses are required. Finally, It has to be acknowledged that many serious issues need to be addressed before this strategy can be usefirl in problems such as those arising in crashworthiness design, such as: how to deal with dynamic behavior? — splitting of ii into coarse-scale and details — how to accommodate nonlinear material behavior and non-linearity associated with large deformations? 5.3 Areas of Future Work An immediate extension of the model reduction scheme presented is to incorporate harmonic loading. This involves the computation of a reduced mass matrix as well as a reduced stiffness matrix. This is of interest in the design of wave-guides and in acoustics. Similarly, a direct extension of the presented layout optimization scheme is the problem of optimal arrangement of beadings or corrugations on plates for improved dynamics properties such as natural frequencies. An approach to suitably combine the dividing and merging approaches in the optimization using variable layouts of substructures. Extension of the proposed schemes for three-dimensional problems in elasticity. Incorporation of transient dynamics, non-linearity arising from material properties as well as fiom large displacements. 141 BIBLIOGRAPHY 142 Bibliography [1] NS. Bakhalov and AV. Knyazev, Fictitious domain methods and computations of homogenized properties of composites with a periodic structure of essentially different components, in G. I. Marchuk (editor), Numerical Methods and Applications, CRC Press, Boca Raton, FL, (1994), 221-266 [2] MP. Bendsoe and N. Kikuchi, Generating optimal topologies in structural design using a homogenization method, Computer Methods in Applied Mechanics and Engineering, 71 (1988) 197-224. [3] MP. Bendsoe, Optimal shape design as a material distribution problem, Structural Optimization, I (1989), 193-202. [4] MP. Bendsoe, A.R. Diaz and N. Kikuchi, Topology and generalized layout optimization of elastic structures, in Topology design of structures, edited by MP. Bendsoe and C .A Mota Scares, Kluwer Academic Publishers, (1993), 159-205. [5] A. Bensoussan, J .L. Lions and G. Papanicolaou, Asymptotic analysis for periodic structures, North-Holland, (1978). [6] G. Beylkin and N. Coult, A multi-resolution strategy for reduction of elliptic PDEs and eigenvalue problems, Applied and Computational Harmonic Analysis, 5, (1998), 129-155. [7] M. Brewster and G. Beylkin, A multi-resolution strategy for numerical homogenization, Applied and Computational Harmonic Analysis, 2 (1995), 327-349 [8] RR. Craig and M.C.C. Bampton, Coupling of substructures for dyncunic analyses, AIAA Journal, 6 (7), (1968), 1313-1319. [9] W. Dahmen and CA. Micchelli, Using the refinement equation for evaluating integrals of wavelets, SIAM Journal of Numerical Analysis, 30, (1993), 507-537 [1011. Daubechies, Ten Lectures 0n Wavelets, CBMS-NSF Series in Applied Mathematics, 61, SIAM, Philadelphia, 1992. [I 1] G. C - A. DeRose, Jr. and AR Diaz, Single scale wavelet approximations in layout optimization, Structural Optimization, 18 (1999), 1-1 I. 143 [12] AR. Diaz and MP. Bendsoe, Shme optimization of structures for multiple loading conditions using a homogenization method, Structural Optimization, 4 (1992), 17-22 [13] AR Diaz and N. Kikuchi, Solutions to shape and topology eigenvalue optimization problems using a homogenization method, International Journal of Numerical Methods in Engineering, 35, (1992), 1487-1502 [14] AR. Diaz, A wavelet-Galerkin scheme for analysis of large scale problems on simple domains, International Journal of Numerical Methods in Engineering, 44, (1999), 1599-1616 [1 5] M. Dorabantu and B. Engquist, Wavelet-based numerical homogenization, SIAM Journal of Numerical Analysis, 35 (1998), 540-559 [16] P. Duysinx and MP Bendsoe, Topology optimization of continuum structures with local stress constraints, International Journal of Numerical Methods in Engineering, 43, (1998), 1453-1478 [17] B. Engquist and O. Runborg, Wavelet-based numerical homogenization with applications, In T.J. Barth, T.F. Chan and R Haimes, editors, Multi-scale and Multi- resolution Methods: Theory and Applications, 97-148, Vol. 20, Lecture Notes in Computational Science and Engineering, Heidelberg 2001 , Springer-Verlag. [18] M. Frazier, An introduction to wavelets through linear algebra, Springer-Veriag, New York, (1999). [19]M.I. Friswell, S.D. Garvey and J.E.T. Penny, Model reduction using dynamic and iterated IRS techniques, Journal of Sound and Vibration, 186(2), (1995), 311-323 [20] AC. Gilbert, A comparison of multi-resolution and classical one-dimensional homogenization schemes, Applied and Computational Harmonic Analysis. 5 (1998), 1-3 5 [21]R. Glowinski, T.W. Pan, R.O. Wells and X. Zhou, Wavelet and finite element solutions for the Neumann problem using fictitious domains, J oumal of Computational Physics, 126, (1996), 40-51 [22] RJ. Guyan, Reduction of stiffitess and mass matrices, AIAA Journal, 3 (2), (1965), 380. [23]R.B. Haber, C.S. Jog and MP. Bendsae, A new approach to variable topology shape design using a constraint on perimeter, Structural Optimization, 11, (1996), 1-12 [24] W.C. Hurty, Dynamic analysis of structural systems using component modes, AIAA Journal, 3 (4), (1965), 678-685. 144 [25] RV. Kohn and G. Strang, Optimal design and relaxation of variational problems, Communications in Pure and Applied Mathematics, 39, (1986), 113-137 (part I), 139-182 (part H), 353-377 (part 11]) [26] A Kunoth, Computing refinable integrals, Technical report ISC-95-02-MATH (1995), Institute for Scientific Computation, Texas A&M University, College Station, Texas, USA. [27] A Latto, H.L. Resnikofl‘ and E. Tenebaum, The evaluation of connection coefiicients of compactly supported wavelets, in Y. Maday (editor), Proceedings of F tench-USA Workshop on Wavelets and Turbulence, Princeton University, Springer, New York, (1992). [28] 8G. Mallat, A theory for multi-resolution signal decomposition: fire wavelet representation, Communications in Pure and Applied Mathematics, 41 (7), 674-693 (1988) [29] F. Murat and L. Tartar, Optimality conditions and homogenization, in A Marino et al (editors), Nonlinear Variational Problems, Pittman Advanced Publishing Program, Boston, (1995), 1-8 [30] S. Pecullan, L.V. Gibiansky and S. Torquato, Scale Eflects on the elastic behavior of periodic cmd hierarchical two-dimensional composites, Journal of the Mechanics and Physics of Solids, 47 (1999), 1509-1542. [31]P. Pedersen, 0n optimal orientation of orthotropic materials, Structural Optimization, 1, (1989), 101-106 [3 2] H.L. Resnikoff and RD. Wells, Wavelet Analysis: The scalable structure of information, Springer-Veflag, New York, (1999). [33] G.I.N. Rozvany, M. Zhou and T. Birker, Generalized shape optimization without homogenization, Structural Optimization, 4, (1992), 250-252 [34] G.I.N. Rozvany, M. Zhou, T. Birker and O. Sigmund, Topology optimization using iterative continuum-We optimality criteria (C 0C) methods for discretized systems, in Topology design of structures, edited by MP Bendsoe and CA Mota Soares, Kluwer Academic Publishers, (1993), 273-286 [3 5] P. Seshu, Substructuring and component mode synthesis, Shock and Vibration, 4 (3), (1997), 199-210. [3 6] K. Suzuki and N. Kikuchi, A homogenization method for shwe and topology optimization, Computer Methods in Applied Mechanics and Engineering, 93 (1991) 291-318 [3 7] R0. Wells and X. Zhou, Wavelet solutions for the Dirichlet problem, Numerische Mathematik, 70, (1995), 379-396 145 [3 8] EL. Wilson, The static condensation algorithm, International Journal of Numerical Methods in Engineering, 8 (1974), 199-203 [39]E.L. Wilson, M-W. Yuan and J .M. Dickens, Dynamic analysis by direct superposition of Ritz vectors, Earthquake Engineering and Structural Dynamics, 10, (1982), 813-821 [40] R-J. Yang, Metarnodeling development for vehicle fiontal impact simulation, (DETC2001/DAC-21012), Proceedings of the 27th Annual ASME Design Automation Conference. Pittsburgh, PA, Sept 9-12, (2001). 146 un]rurrrrnmrgrrrjru